You are on page 1of 15

Applied Surface Science 536 (2021) 147326

Contents lists available at ScienceDirect

Applied Surface Science


journal homepage: www.elsevier.com/locate/apsusc

High-index crystal plane of ZnO nanopyramidal structures: Stabilization, T


growth, and improved photocatalytic performance
Taehoon Lima, Pegah S. Mirabedinib, Kichang Junga,c, P. Alex Greaneyb,d,
Alfredo A. Martinez-Moralesa,b,c,

a
College of Engineering - Center for Environmental Research and Technology, University of California - Riverside, Riverside, CA 92507, United States
b
Materials Science and Engineering Program, University of California - Riverside, Riverside, CA 92521, United States
c
Department of Chemical and Environmental Engineering, University of California - Riverside, Riverside, CA 92521, United States
d
Department of Mechanical Engineering, University of California - Riverside, Riverside, CA 92521, United States

ARTICLE INFO ABSTRACT

Keywords: While the low-index planes of Wurtzite ZnO, such as {0001} and {101̄0} are well-understood, the high-index
Zinc oxide crystal surfaces have not yet been thoroughly researched despite possessing structural characteristics that make
Pyramidal structure them suitable for many important surface chemistry processes. The high surface energy of high-index ZnO
High-index facet crystal surfaces makes synthesis challenging to achieve due to their instability during crystal growth. In this
CVD synthesis
work, we present a combined experimental and theoretical analysis of growth and photocatalytic activity of ZnO
Oxygen evolution reaction
high-index crystal facets. Density functional theory calculations are performed to determine the thermodynamic
conditions necessary to stabilize the high-energy semi-polar {112̄2} facets of pyramidal ZnO nanostructures
grown via chemical vapor deposition (CVD). The photocatalytic properties of as-synthesized nanopyramidal
structures for water splitting applications showed a 73% improved photocatalytic performance compared to
CVD-grown hexagonal ZnO nanorods with dominating low-index {101̄0} facets.

1. Introduction Improvements in photocatalytic performance have been achieved via


interface engineering [14], surface decoration [15], morphology en­
ZnO is a versatile wide bandgap semiconductor used in a diverse gineering [16,17], hybridization [18,19], and doping [20]. However,
array of applications due to its piezoelectric [1], photocatalytic [2], and the hexagonal prismatic structure is not an optimal morphology for
intrinsic n-type semiconducting [3] properties. The naturally occurring surface catalysis applications involving water. The non-polar {101̄0}
Wurtzite crystal structure favors a hexagonal prismatic crystal shape surfaces interact weakly with water molecules, while the polar {0001}
bounded by low-index crystal facets, with non-polar {101̄0} surfaces on surfaces have difficulty with desorption [21–24] due to the stronger
the sides and polar {0001} facets terminating the ends. The intrinsic binding forces on these surfaces.
properties of low-dimensional crystalline materials can be tuned via While the {101̄0} and {0001} crystal planes have been well re­
quantum confinement, or by controlling the surface-to-volume ratio. searched, both theoretically and experimentally [25], the high-index
This makes the ZnO nanostructures suitable for many optical, elec­ facets such as {112̄2} have not been fully investigated due to the un­
tronic, and optoelectronic applications. As a result, ZnO nanostructures stable nature of the surface. Diagonal surface facets of ZnO such as
are commonly used in anti-reflective coatings [4], nanowire-based {101̄1} , {101̄2} and {112̄2} planes can form a pyramidal morphology.
transistors [5], LEDs [6], lasers [7], and photodetector devices [8]. These semi-polar surfaces make the pyramidal ZnO nanocrystals a
According to a previous study, quasi-1-dimensional ZnO shows im­ strong candidate for a photocatalytic material due to their high surface
proved photocatalytic activity in comparison to the bulk state due to its electrochemical activity and better reversibility of the ad­
larger surface-to-volume ratio [9]. Morphologies such as nanowire and sorption–desorption process [26,27].
nanorod arrays have been extensively researched in photocatalytic Synthesis of pyramidal ZnO crystals in colloidal form via hydro­
applications including photoelectrochemical water splitting [10], self- thermal reaction has been previously reported [28,29]. However, for
cleaning surface [11], and air/water purification [12,13]. many applications, it is desirable to grow crystalline pyramidal ZnO


Corresponding author at: College of Engineering - Center for Environmental Research and Technology, University of California - Riverside, 1084 Columbia Ave,
Riverside, CA 92507, United States.
E-mail address: alfmart@cert.ucr.edu (A.A. Martinez-Morales).

https://doi.org/10.1016/j.apsusc.2020.147326
Received 18 February 2020; Received in revised form 20 July 2020; Accepted 20 July 2020
Available online 24 July 2020
0169-4332/ © 2020 Elsevier B.V. All rights reserved.
T. Lim, et al. Applied Surface Science 536 (2021) 147326

(a) γ(1010)-S (b) γ(1010)-Zn (d) γ(1122)-S-Zn (e) γ(1122)-O

(c) (f)

γ(1010)-S γ(1010)-O γ(1122)-S-O γ(1122)-Zn


Fig. 1. Profile view of ZnO surface facets after relaxation. (a-c) Slabs of the (101̄0) surfaces viewed along the c-axis, and (d-f) slabs of the (112̄2) surfaces viewed from a
direction perpendicular to the c-axis. (a) Shows the low energy stoichiometric surfaces, (b) and (c) show the same surface depleted of O and Zn, to generate the Zn-
rich and O-rich surfaces, respectively. (d) Shows the Zn and O terminated stoichiometric planes, with (e) showing the O-rich surface, and (f) the Zn-rich surface.

arrays directly on a conductive substrate to minimize interfacial elec­ generating a small polarization perpendicular to the surface. Non-stoi­
trical resistance, while making the material more robust and easier to chiometric {101̄0} surfaces that are deficient in either Zn or O can also
handle. be generated by removal of Zn or O atoms from the surface layer. The
In this paper, we report the chemical vapor deposition (CVD) (101̄0) facet, therefore, has three distinct surface energies, (101̄0) S ,
growth of crystalline ZnO pyramidal arrays with surface {112̄2} facets on (101̄0) Zn , and (101̄0) O for the stoichiometric, Zn-rich, and O-rich struc­
a transparent conductive substrate. Despite the unstable nature of high- tures, respectively.
index crystal surfaces, synthesis was achieved by stabilizing the surface The surface energies of stoichiometric surfaces, (hkwl) S , were cal­
during crystal growth. Experimental conditions were predicted and culated by computing the facet energy of relaxed slabs, each containing
guided by density functional theory (DFT) calculations. Crystal growth a single pair of parallel stoichiometric facets, using the following
conditions were controlled via a modified CVD ZnO synthesis process to equation [31]:
prevent the premature oxidation of Zn precursor in the low-temperature Eslab NO Ebulk
process [30]. The high-energy surface was stabilized by controlling the =
(hkwl) S
2A (1)
chemical potential of oxygen. As predicted by the DFT calculations and
confirmed by the hydrogen and oxygen evolution reactions from water where Eslab is the total energy of the slab, NO is the number of oxygen
splitting experiments, the as-synthesized ZnO nanopyramidal structures atoms in the slab, Ebulk is the energy of one formula unit of bulk ZnO,
show bi-functionality of p-type and n-type semiconducting character­ and A is the area of the surface. The factor of 2 in the denominator
istics. To understand the origin of the improved photocatalytic prop­ indicates the presence of a top and bottom surface in the slab model.
erties of pyramidal ZnO, the surface geometry, electrochemical prop­ The surface energy of the non-stoichiometric facets depends on the
erties, and electrostatic potential of different surfaces were chemical potential of the reservoir from which the excess surface spe­
investigated. cies are drawn. The energies of these surfaces were obtained from
calculations of slabs with one stoichiometric surface and a non-stoi­
chiometric opposing surface. The surface energy of the non-stoichio­
2. Calculation and experimental section metric facet is given by:
1
2.1. Surface energy calculations = [Eslab Eref (NO , NZn, µO )]
(hkwl) X
A (hkwl) S (2)

Electronic structure calculations were performed to compute the with the reference energy
surface energy of the non-polar (101̄0) and semi-polar (112̄2) facets of Eref (NO , NZn, µO ) = NZn (Ebulk µO EO2 ) + NO ( µO EO2 ) (3)
ZnO and to determine how the electronic structure of these surfaces
affects the crystal growth process and photocatalytic activity. DFT was where NO and NZn are the number of O and Zn atoms in the system,
used to compute the total energy of infinite slabs of ZnO with exposed respectively. EO2 is the cohesive energy per atom of oxygen in the O2
parallel (101̄0) or (112̄2) facets, as shown in Fig. S1. The slabs for DFT dimer. The oxygen chemical potential, µO , is defined as the difference
calculations were cut from bulk ZnO crystal and relaxed. Depending on between chemical potential of oxygen in the reactor and half of the O]
where the cut-plane intersects the unit cell, several surface structures O bond energy, and it is plotted over the range for which ZnO is stable.
are possible, enabling the formation of stoichiometric surfaces and This gives the low µO limit as the chemical potential at which ZnO is
surfaces that are O-rich or Zn-rich. In the [101̄0] direction, the packing reduced to Zn metal and O2 gas,
of atomic planes follows an AaBbAa sequence. The (101̄0) facet can be (Ebulk EZn HCP EO2) µO 0 (4)
cut between the narrowly spaced Aa planes or the wider aB planes. The
surface energy of both possible cuts was computed. The resulting sur­ where EZn HCP is the cohesive energy per atom of HCP Zn metal. The
faces cut between the wider aB planes were found to be significantly relaxed low energy geometries of the stoichiometric, Zn-rich, and O-
lower in energy as they are more compact. These lower energy surfaces rich (101̄0) facets are shown in Fig. 1(a)–(c), respectively.
are used for the DFT calculation in this work. The geometric permutations of the semi-polar {112̄2} surfaces are
The {101̄0} surfaces lie parallel to the polar c-axis of the crystal. Each more complex with four distinct surface terminations: O- and Zn-ter­
layer of atoms in the AaBb packing sequence is stoichiometric, making minated stoichiometric surfaces (opposite faces of a stoichiometric slab)
the facets nominally non-polar. However, under relaxation, the stoi­ and O-rich and Zn-rich non-stoichiometric surfaces. Although four
chiometric (101̄0) surface undergoes a small buckling distortion, different permutations of the top and bottom surface structures can be

2
T. Lim, et al. Applied Surface Science 536 (2021) 147326

created, from the ZnO crystal structure, it is only possible to determine done to control the chamber pressure and distance between precursor
three independent surface energy metrics. While the distinct surface and substrate. In short, to prevent the surface oxidation of Zn powder
energies for the (101̄0) facets can be determined, for the (112̄2) surface, (Sigma-Aldrich) precursor, a quartz tube (3 mm inner diameter, 12 cm
the energy of the stoichiometric surface, (112̄2) S , is taken as half of the length) with one end closed was used as a precursor boat. 0.2 g of Zn
cleavage energy, (112̄2) S = 2 ( (112̄2) S Zn + (112̄2) S O ) . This is used as
1 powder was loaded in the precursor boat and placed in the middle of the
the reference surface in Eq. (2) to define the energies of the Zn- and O- right zone in the tube furnace. The FTO substrate (Sigma-Aldrich) was
rich surfaces, (112̄2) -Zn and (112̄2) -O. These relaxed structures of the placed downstream next to the open end of the precursor boat. During
(112̄2) facets are shown in Fig. 1(d)-(f). synthesis, all three zones of the furnace (900 mm total heating zone)
The energy calculations were performed using supercells with were heated to 450 °C for the nanopyramidal and 500 °C for the nanorods
(2 × 1) and (2 × 2) periodicity in the plane of the slab for {101̄0} and at a ramping rate of 15 °C/min, under the flow of 100 SCCM nitrogen and
{112̄2} structures to allow the system freedom to undergo surface re­ 4 SCCM oxygen. The furnace temperature was maintained constant for
constructions that break the crystal symmetry. To obtain a consistent 30 min to during the crystal growth. The measured pressure during the
set of fully converged surface energies, calculations were performed for synthesis process was 12 torr. The synthesized ZnO was analyzed by
a series of slabs by systematically varying the slab thickness and va­ scanning electron microscopy (SEM), x-ray diffraction (XRD) analysis,
cuum separation. The computational details are given in the supple­ and Raman spectroscopy to characterize the morphology, crystal struc­
mentary material along with the detailed geometry of the computed ture, and crystal defects, respectively. J-V characteristics and electro­
systems. The surface energy of the (101̄0) terminated slabs was com­ chemical impedance spectroscopy (EIS) under dark and illuminated
puted to an accuracy of 0.01 J/m2. Convergence was achieved with conditions were measured to analyze photocatalytic properties and are
slabs having a thickness of 11 or more atomic layers ( 31 Å) and a discussed later in the manuscript. The photocatalytic performance of
vacuum space of 20 Å. In contrast, the convergence of the semi-polar synthesized ZnO nanopyramids and nanorods was tested under 365 nm
(112̄2) surface was harder to achieve due to the existence of an internal UV illumination in an 0.5 M Na2SO4 aqueous solution. Measurements
field formed in the polar slabs. In this case, surface energies were were performed in a three-electrode photoelectrochemical cell with Ag/
computed following the convergence method by Meyer et al. [32], AgCl as the reference electrode and Pt wire as the counter electrode.
where the energies of slabs with a series of different thicknesses, t, were
computed and then the 1/ t dependence of the slab energy was extra­
3. Results and discussion
polated to remove the effect of the internal field. This method resulted
in the surface energy of the (112̄2) terminated slabs being computed to
3.1. Prediction of crystal growth morphology
an accuracy of only 0.1 J/m2.
All DFT calculations were performed using the Vienna ab-initio
A growing crystal arrives at its final morphology either as a result of
Simulation Package (VASP) [33–36] within the Perdew-Berke-Ern­
anisotropic attachment kinetics producing a shape that becomes kine­
zerhof (PBE) general gradient approximation (GGA) ex­
tically frozen in place, or as a result of the growing crystal relaxing
change–correlation functional [37,38]. The calculations were spin-po­
towards its thermodynamically preferred shape. The latter is the shape
larized with a plane wave cutoff energy of 400 eV. Core electrons were
that minimizes the total surface energy of the structure and is described
modeled implicitly using the projector-augmented wave method (PAW)
by the Wulff construction. Here we consider only the thermo­
pseudopotentials [39,40]. Structures were relaxed using the conjugate
dynamically preferred shape and further restrict consideration to the
gradient algorithm [41] to minimize the forces for ionic relaxation to
Wulff construction of the crystal containing only {101̄0} and {112̄2} facets
lower than 1 × 10-3 eV/Å. A great deal of care was taken to obtain a
to identify conditions under which pyramidal crystals will be thermo­
fully converged and relaxed bulk ZnO reference structure, as errors in
dynamically preferred over nanorods.
the energy prediction of this reference state are the largest source of
The energy of the different variants of the {101̄0} and {112̄2} facets
uncertainty in the surface energy predictions. The cohesive energy of
are plotted as a function of oxygen chemical potential in Fig. 2. For the
the bulk crystal structure was found to be well converged using a
stoichiometric surfaces, the (101̄0) facet is more compact and has a
5 × 5 × 5 k-points grid for the Brillouin zone integration. Using this
approach, the modeled bulk ZnO had equilibrium lattice parameters of
a = 3.19 Å and c = 5.30 Å. For the slab calculation, a 5 × 5 × 1 k- (a) Oxygen chemical potential ΔμO (eV)
point mesh was used for Brillouin zone sampling with one k-point along -2 -1 0
the non-periodic direction perpendicular to the slab. 6
Surface energy γ (J/m )

{1122}
2

The surface energies computed using the described procedure above


{1010}
were used in the Wulff construction to predict the equilibrium mor­
phology of ZnO crystallite growing on a substrate with its c-axis per­ 4 O-rich
pendicular to the substrate. The energy of the ZnO/substrate interface is
assumed to be zero. The equilibrium morphologies were used for the Zn-rich
theoretical predictions of the height-to-width ratio and the {112̄2} to 2
{101̄0} surface area ratio of ZnO pyramidal nanocrystals grown under Stoichiometric
different oxygen chemical potential conditions. For growth on most
substrates, it is likely that the effective surface energy of the ZnO/ (b) 04
substrate is negative due to the reduction in energy from covering the
Ratio

exposed substrate and ZnO-(0001) surfaces by bringing ZnO into contact 2 {1122}/{1010}
with the substrate. If the effective ZnO-(0001) /substrate interface en­ Height/Width
ergy is negative, the width-to-height ratio and {112̄2} to {101̄0} surface 0
area ratio of ZnO pyramidal nanocrystals will be underestimated. (c)
2.2. Experimental procedures

Nanopyramid and nanorod ZnO structures were synthesized based on Fig. 2. (a) Plot of the surface energy of the {112̄2} (black) and {101̄0} (red) facets
our modified CVD method in a three-zone tube furnace (300 mm length/ as a function of oxygen chemical potential, (b) corresponding surface area
zone, OTF-1200X-III, MTI Corporation) [42]. Further optimization was (green) and aspect ratio (blue), and (c) corresponding Wulff construction.

3
T. Lim, et al. Applied Surface Science 536 (2021) 147326

lower surface energy of 0.9 J/m2, while the surface energy of the respectively) and the nanorods (Fig. 3(d) and 3(e)). In Fig. 3(c), the φ
stoichiometric (112̄2) surface is 1.8 J/m2. However, for the non-stoi­ (angle of the plane from the c-axis) observed on the {112̄2} crystals
chiometric surfaces, both the O-rich and Zn-rich (112̄2) facets have planes is calculated as 32°, which corresponds to the measured angle at
lower surface energy than their corresponding (101̄0) facet. The surface the apex of the pyramidal structure 2φ = 64° when viewed along
energy versus the chemical potential of oxygen plot in the system 101̄0 . The nanopyramids in projection from the side will appear to
(Fig. 2(a)) shows that there is a small range of O-starved growth con­ have an apex angle range between 64° and 72° (when viewed along
ditions under which the Zn-rich {112̄2} facets become lower in energy 21̄1̄0 ). The pyramidal peak angles measured in Fig. 3(b), provide
than even the stoichiometric{101̄0} surface. However, for the {112̄2} compelling evidence that although the {112̄2} crystal facets are not fa­
pyramidal facets, to completely geometrically exclude the {101̄0} facets vored under regular conditions due to their high energy, they have
from the Wulff construction, (112̄2) must be considerably lower than nonetheless been successfully synthesized through our modified CVD
process. The crystal structure of both ZnO nanostructures are char­
(101̄0) , satisfying the inequality,
(101̄0)
(112̄2)
2
3
1+ ()
a 2
c
(≈1.36 for ZnO).
acterized by XRD analysis, as shown in Fig. 4(a). The XRD results in­
Although this condition is not met in the plot in Fig. 2, it is not ne­ clude the peaks from the FTO substrate and the ZnO products. The ZnO
cessary for the {101̄0} facets to be completely excluded for the crystals to XRD pattern confirms that the product is Wurtzite ZnO. The ZnO peaks
grow in a pyramidal morphology. Fig. 2(b) shows the height/width and at 34.59°, 36.37°, 47.70°, 63.02°, and 68.04°, refer to the (0002) , (101̄1) ,
the {112̄2} /{101̄0} area ratio of the Wulff construction (Fig. 2(c)) is (101̄2) , (101̄3) , and (112̄2) crystal planes, respectively. While the peak
plotted on the same x-axis. Under the oxygen-starved growth condi­ ratio of (0002) is higher for the ZnO nanorods, the peak ratio for the
tions, the equilibrium morphology of ZnO crystals transitions from rod- diagonal planes (101̄1), (101̄2) , (101̄3) , and (112̄2) are higher for the ZnO
like to compact pyramidal structures is marked by the drastic increase nanopyramids. The XRD peak ratios for the nanopyramids and na­
in the fraction of semi-polar {112̄2} surfaces. norods crystal planes are listed in Table 1. The XRD results confirm that
The DFT calculations anticipate that ZnO with high-index {112̄2} the nanopyramid and nanorod structures have different crystal or­
surface planes can be synthesized when the chemical potential of oxygen ientations. The morphology and crystal structure of pyramidal ZnO
is low. This prediction is consistent with the experimental results. ZnO provide important clues to the growth mechanism. In the top-view SEM
pyramidal structures are successfully synthesized when the Zn vapor imaging, the incomplete ZnO nanopyramids have an irregular surface
partial pressure is significantly higher than that of oxygen near the pattern. The voids parallel to the c-axis indicate that the pyramids were
substrate. This extremely biased precursor ratio environment is achieved formed through side-by-side lateral growth of ZnO rather than layer-by-
by modifying the design of the CVD reactor. The prevention of the pre­ layer vertical growth, as it occurs in the growth of ZnO nanorods [43].
mature oxidation of Zn precursor plays an important role in growing the The crystal growth mechanism is further supported by Raman
ZnO crystal with {112̄2} surfaces by creating an environment with highly spectroscopy as shown in Fig. 4(b). The two most significant peaks are
concentrated Zn vapor. When the crystals are removed from the reactor observed in both structures at 100 and 440 cm−1, assigned as E2low and
and exposed to air, the surfaces quickly oxidize and become stoichio­ E2high, respectively. The observed differences in the Raman spectra are
metric, but the crystal morphology remains in the pyramidal shape. attributed to the nanorods with peaks at 121, 382, and 483 cm−1. The
peak at 382 cm−1 is assigned as A1(TO), a phonon vibration mode
parallel to the c-axis, forbidden for the c-plane oriented Wurtzite
3.2. Morphology and crystal structure analysis structure. The forbidden A1(TO) vibration is observed in the ZnO na­
norods due to its angled crystal orientation. The ZnO nanopyramids
Fig. 3 shows the top-view and cross-sectional SEM images of the show no A1(TO) vibration because the pyramidal structure is better
synthesized ZnO nanopyramidal structures (Fig. 3(a) and 3(b),

Fig. 3. (a) Top-view and (b) cross-sectional SEM images of synthesized pyramidal ZnO, (c) schematic drawing of (112̄2) plane for geometric analysis. (d) Top-view
and (e) cross-sectional SEM images of synthesized nanorod ZnO, (f) schematic drawing of (101̄0) plane.

4
T. Lim, et al. Applied Surface Science 536 (2021) 147326

(a) (0002)
(b)
Nanopyramids E2low Nanopyramids
Nanorods Nanorods

Intensity (Arb. Unit)


Intensity (Arb. Unit) FTO substrate
E2high

ZB TA(X) A1(TO) ZB TO(X)


(1013)
(1012)

(1011) (1122)
30 40 50 60 70 100 200 300 400 500
-1
600 700
2θ (degrees) Raman shift (cm )
Fig. 4. (a) XRD and (b) Raman spectra results of synthesized nanopyramids and nanorods.

Table 1
XRD peak ratios of ZnO nanopyramids and nanorods from diverse planes. (a) 0.5
Nanopyramid dark
0.4

Current density (mA/cm )


Nanopyramid light

2
Orientation Rectangular Diagonal Nanorod dark
0.3 Nanorod light
Miller index (0002) (101̄1) (101̄2) (101̄3) (112̄2)
Nanopyramid 1 1.77 1.69 1.65 1.35
Nanorod 1.13 1 1 1 1
0.2 OER
0.1

aligned perpendicularly to the substrate. The two peaks at 121 and 0.0
483 cm−1 observed in the nanorods are assigned as TA(X) and TO(X),
-0.1
respectively, which are from the planar crystal defect forming a zinc HER
blende isomorph layer [44]. This planar defect originates from the -0.2
stacking fault in the layer-by-layer crystal growth mechanism in ZnO -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6
nanorods. As supported by SEM, these peaks are not observed in the Potential (V vs Ag/AgCl)
nanopyramids due to their side-by-side crystal growth mechanism.
(b) Anodic current (c) Cathodic current
3.3. Photoelectrochemical water splitting
H+ /H2 H+ /H2
Fig. 5(a) shows the J-V characteristics of the water splitting op­
eration of the nanopyramids and nanorods arrays under illumination
and in dark conditions. The figure shows the average plot for 7 samples hν H2O/O2 H2O/O2
of each nanostructure to confirm the reproducibility and reliability.
While the statistical data (Fig. S2) shows that both nanostructures are
reproducible, further controllability is needed to reduce the built-in
experimental variation in the CVD reactor setup, as described in the
supplementary material. When illuminated, both ZnO morphologies
mediate the oxygen evolution reaction (OER), as observed from the FTO ZnO electrolyte FTO ZnO electrolyte
positive current. The onset potential (determined by measuring the x-
intercept of tangent at maximum slope [45]) of ZnO nanopyramids is Fig. 5. Water splitting operation of ZnO nanopyramids and nanorods; (a) J-V
characteristics under dark and 365 nm illumination, (b) energy diagram under
−0.31 V, which is negatively shifted by 45 mV from that of ZnO na­
illumination, and (c) dark condition.
norods. The cathodic onset potential shift in the nanopyramidal ZnO
suggests that the energy barrier for the hole injection to the electrolyte
is reduced [46]. The Tafel slope of the ZnO nanopyramids (43.2 mV/ in the nanopyramid structures. The photoelectrochemical performance
dec) is lower than the nanorods (73.1 mV/dec), indicating a faster of ZnO nanopyramids and nanorods are listed in Table 2 and compared
catalytic reaction and higher OER activity [47]. The nanopyramidal with other results in the literature in Table S1.
ZnO (red line) generates a current density of 0.45 mA/cm2, re­ The increased photocatalytic activity of ZnO nanopyramids is at­
presenting a 73% improvement over the 0.26 mA/cm2 produced by the tributed to the increased surface energy, polarity, enlarged active sur­
ZnO nanorod at 0.6 V vs Ag/AgCl reference electrode. The steeper slope face area, reduced interfacial charge transfer resistance, and increased
in the linear region before the saturation area of the nanopyramidal electron lifetime, as supported by EIS, transient voltage decay analysis,
ZnO indicates that the photogenerated charge carriers are more effec­ and electron lifetime measurement (Fig. 6 and S3 (a) – (b), respec­
tively separated in the nanopyramidal ZnO nanostructures [48]. The tively). The origin of the improved photocatalytic performance is ana­
steady-state illumination induces non-equilibrium electron and hole lyzed by EIS at open circuit potential, as shown in the Nyquist plot
concentrations, resulting in the splitting of the quasi-Fermi level and (Fig. 6(a)) and Bode plot (Fig. 6(b)). The circuit diagram in Fig. 6(c)
band bending, as schematically illustrated in Fig. 5(b) [49]. The pho­ represents the equivalent circuit model used to analyze results from the
tocatalytic operation for OER is due to the innate n-type semi­ Nyquist and Bode plots (listed in Table 3). The EIS results are separated
conducting nature of ZnO and is further improved by the crystal surface into two semi-circles in the Nyquist plot. RS, R1, and CPE1 are internal

5
T. Lim, et al. Applied Surface Science 536 (2021) 147326

Table 2 Table 3
Electrochemical performance of ZnO nanopyramids and nanorods. Equivalent circuit element values.
Property Nanopyramids Nanorods Circuit element RS (Ω) R1 (Ω) CPE1 (μF) RCT (kΩ) CPEDL (μF)

Onset Potential −0.31 V −0.26 V Nanopyramids dark 37.0 1860 7.61 90.1 5.65
Tafel Slope 43.2 mV/dec 73.1 mV/dec Nanopyramids light 33.1 225 32.9 22.5 153
Slope at Linear Region 0.90 mA/Vcm2 0.48 mA/Vcm2 Nanorods dark 49.0 342 22.3 1300 5.30
Current Density at 0.6 V 0.45 mA/cm2 0.26 mA/cm2 Nanorods light 49.2 462 29.4 48.7 13.5
Current Density at −0.6 V −0.19 mA/cm2 −0.05 mA/cm2

nanopyramidal ZnO is derived from the ambipolar nature of the (112̄2)


resistance, bulk resistance, and capacitance, respectively. The lower crystal facet observed in the DFT study. The electron-withdrawing and
frequency semi-circle is assigned to the surface charge transfer, where donating operations are both possible on the surface by having both
RCT is the charge transfer resistance, and CPEDL is the double-layer electro-positive and electro-negative functionalities on the same plane.
capacitance due to the slower charge transfer between the electrode The electrostatic potential of the DFT computed stoichiometric
and ions over the bulk electronic process [50]. ZnO nanopyramids (101̄0) and (112̄2) surfaces was analyzed to determine which structure
under illumination show the lowest RCT indicating that the charge has the highest polarity and greater photoelectrochemical activity. Only
transfer is easier in the pyramidal structure, and the CPEDL is generally stoichiometric planes are considered because although we predict that
proportional to the active surface area [51]. Therefore, a higher CPEDL the pyramidal morphology is stabilized during growth in oxygen-defi­
in pyramidal ZnO represents that the active surface area of pyramidal cient conditions, these surfaces will quickly become stoichiometric
ZnO is larger than that of the nanorod structure. The lower charge under ambient conditions. Fig. 7 shows the atomic structure of each
transfer resistance and larger surface area are the main reasons for the surface with the left-hand side of each plot showing the surfaces’ local
improved ZnO nanopyramids’ performance. The ZnO nanopyramids are electrostatic potentials plotted on the surface of constant charge den­
bounded by semi-polar {112̄2} crystal planes with higher surface energy sity. The charge density contour value is the same for all plots. The
and polarity. The high energy surface tends to lower its surface energy color scale indicating the local voltage is the same for all surfaces. It is
by having adsorbents on the surface [52], and the high electrostatic observed that the two (112̄2) surfaces present a wider variation in the
potential surface attracts the ionic molecules such as H+ and OH– [53]. local electrostatic potential than the (101̄0) surface. The compactly co­
Another key difference between the ZnO nanopyramid and nanorod ordinated and charge-balanced structure of the (101̄0) surface offers
structures is that the nanopyramids have electrocatalytic properties for smaller variation in the surface potential presented to water molecules
the hydrogen evolution reaction (HER) irrespective of whether they are than both the Zn and O terminated stoichiometric (112̄2) facets which
illuminated or not. In contrast to OER, HER is generally mediated by p- are polarized. The results are supportive of the observations that both
type semiconductors and metals with a relatively low Fermi level or Zn- and O-terminated (112̄2) surfaces are more reactive than the (101̄0)
work function. The proposed operation mechanism for electrochemical non-polar facet. Therefore, the (112̄2) facets are more likely to react
HER is driven by the applied bias-induced band bending, as illustrated with water molecules and are more suitable for water splitting appli­
in Fig. 5(c) [54]. This bi-functionality in OER and HER catalysis of cations.

(a) -2500 (b)


0 Nanopyramids dark
θ (degree)

Nanopyramids light
-20 Nanorods dark
Nanorods light
-2000 -40
-60
-80
-1500
105 104 103 102 101 100 10-1
Zim (Ω)

Frequency (Hz)
-1000
105 Nanopyramids dark
Nanopyramids light
|Z| (Ω)

4 Nanorods dark
Nanopyramids dark 10 Nanorods light
-500 Nanopyramids light 103
Nanorods dark
102
Nanorods light
0
0 500 1000 1500 2000 2500 105 104 103 102 101 100 10-1
Zre (Ω) Frequency (Hz)
(c) RS R1 RCT

CPE1 CPEDL

Fig. 6. (a) Nyquist plot of ZnO nanopyramids and nanorods under dark and illumination; (b) Bode phase (top) and amplitude (bottom) plots; and (c) equivalent
circuit model.

6
T. Lim, et al. Applied Surface Science 536 (2021) 147326

Fig. 7. Electrostatic potential on the stoichiometric


ZnO facets (left side of the surface) and atomic lo­
cations with red atoms indicating O and silver for Zn
(right side) on the surface of (a) (101̄0) , (b) Zn-ter­
minated (112̄2) , and (c) O-terminated (112̄2) .

4. Conclusion References

Although the pyramidal structures with {112̄2} facets are generally [1] J. Briscoe, S. Dunn, Piezoelectric nanogenerators – a review of nanostructured
not stable due to their relatively high surface energy, our DFT calcu­ piezoelectric energy harvesters, Nano Energy 14 (2015) 15–29, https://doi.org/10.
1016/j.nanoen.2014.11.059.
lations show that these morphologies are stabilized under a Zn-rich [2] C.B. Ong, L.Y. Ng, A.W. Mohammad, A review of ZnO nanoparticles as solar pho­
growth condition. The Wulff crystal structures show a transition to tocatalysts: Synthesis, mechanisms and applications, Renew. Sustain. Energy Rev.
eliminate {101̄0} nonpolar facets and form pure pyramidal structures 81 (2018) 536–551, https://doi.org/10.1016/j.rser.2017.08.020.
[3] Ü. Özgür, D. Hofstetter, H. Morkoç, ZnO Devices and Applications: A Review of
with {112̄2} semi-polar surfaces as the environment becomes oxygen- Current Status and Future Prospects, Proc. IEEE. 98 (2010) 1255–1268, https://doi.
deficient. This suggests a correlation between the growth condition and org/10.1109/JPROC.2010.2044550.
the final morphology of materials. Using this correlation, we tailored [4] M. Alimanesh, J. Rouhi, Z. Hassan, Broadband anti-reflective properties of grown
ZnO nanopyramidal structure on Si substrate via low-temperature electrochemical
our CVD process to obtain materials that are normally unstable and
deposition, Ceram. Int. 42 (2016) 5136–5140, https://doi.org/10.1016/j.ceramint.
difficult to grow. The nanopyramidal ZnO structures with semi-polar 2015.12.033.
{112̄2} facets were synthesized for improved photocatalytic perfor­ [5] G.C. Gazquez, S. Lei, A. George, H. Gullapalli, B.A. Boukamp, P.M. Ajayan, J.E. ten
Elshof, Low-Cost, Large-Area, Facile, and Rapid Fabrication of Aligned ZnO
mance. The synthesis process for the pyramidal ZnO was accomplished
Nanowire Device Arrays, ACS Appl. Mater. Interfaces 8 (2016) 13466–13471,
via the prevention of the premature oxidation of Zn vapor to create an https://doi.org/10.1021/acsami.6b01594.
oxygen-deficient condition, allowing for the crystal growth environ­ [6] C. Pan, L. Dong, G. Zhu, S. Niu, R. Yu, Q. Yang, Y. Liu, Z.L. Wang, High-resolution
ment necessary to achieve the ZnO nanopyramidal morphology. The electroluminescent imaging of pressure distribution using a piezoelectric nanowire
LED array, Nat. Photonics. 7 (2013) 752–758, https://doi.org/10.1038/nphoton.
pyramidal ZnO shows higher photocurrent density under the UV illu­ 2013.191.
mination in comparison to the ZnO nanorod structure. The improved [7] M.H. Huang, S. Mao, H. Feick, H. Yan, Y. Wu, H. Kind, E. Weber, R. Russo, P. Yang,
photocatalytic performance of the pyramidal ZnO is a result of the Room-Temperature Ultraviolet Nanowire Nanolasers, Science. 292 (2001)
1897–1899, https://doi.org/10.1126/science.1060367.
lower interfacial resistance at the surface of ZnO due to its higher [8] Y. Ning, Z. Zhang, F. Teng, X. Fang, Novel Transparent and Self-Powered UV
surface energy, electrostatic potential, and polarity. The technique for Photodetector Based on Crossed ZnO Nanofiber Array Homojunction, Small 14
the controlled crystal growth of high-index facets demonstrated in this (2018) 1703754, https://doi.org/10.1002/smll.201703754.
[9] Y.-C. Chang, P.-S. Lin, F.-K. Liu, J.-Y. Guo, C.-M. Chen, One-step and single source
work opens up opportunities for the application of ZnO pyramidal na­ synthesis of Cu-doped ZnO nanowires on flexible brass foil for highly efficient field
nostructures to a wide range of applications in surface physics and emission and photocatalytic applications, J. Alloys Compd. 688 (2016) 242–251,
chemistry. https://doi.org/10.1016/j.jallcom.2016.07.193.
[10] X. Li, S. Liu, K. Fan, Z. Liu, B. Song, J. Yu, MOF-Based Transparent Passivation Layer
Modified ZnO Nanorod Arrays for Enhanced Photo-Electrochemical Water Splitting,
CRediT authorship contribution statement Adv. Energy Mater. 8 (2018) 1800101, https://doi.org/10.1002/aenm.201800101.
[11] G. Kenanakis, D. Vernardou, N. Katsarakis, Light-induced self-cleaning properties of
ZnO nanowires grown at low temperatures, Appl. Catal. A Gen. 411–412 (2012)
Taehoon Lim: Conceptualization, Methodology, Investigation, 7–14, https://doi.org/10.1016/j.apcata.2011.09.041.
Validation, Formal analysis, Writing - original draft. Pegah S. [12] J.-C. Grivel, Photocatalytic Water Purification with ZnO Thin Films in
Mirabedini: Formal analysis, Methodology, Software, Investigation, Demineralized as Well as Natural Waters, Phys. Status Solidi A 216 (2019)
1800880, https://doi.org/10.1002/pssa.201800880.
Visualization, Data curation, Writing - original draft. Kichang Jung: [13] A. Fathy, M. Le Pivert, Y.J. Kim, M.O. Ba, M. Erfan, Y.M. Sabry, D. Khalil,
Investigation, Validation. P. Alex Greaney: Formal analysis, Y. Leprince-Wang, T. Bourouina, M. Gnambodoe-Capochichi, Continuous
Methodology, Software, Resources, Writing - review & editing. Alfredo Monitoring of Air Purification: A Study on Volatile Organic Compounds in a Gas
Cell, Sensors 20 (2020) 934, https://doi.org/10.3390/s20030934.
A. Martinez-Morales: Resources, Supervision, Funding acquisition, [14] Z. Kang, H. Si, S. Zhang, J. Wu, Y. Sun, Q. Liao, Z. Zhang, Y. Zhang, Interface
Project administration, Writing - review & editing. Engineering for Modulation of Charge Carrier Behavior in ZnO
Photoelectrochemical Water Splitting, Adv. Funct. Mater. 29 (2019) 1808032,
https://doi.org/10.1002/adfm.201808032.
Declaration of Competing Interest [15] R. Bin Wei, P.Y. Kuang, H. Cheng, Y.B. Chen, J.Y. Long, M.Y. Zhang, Z.Q. Liu,
Plasmon-Enhanced Photoelectrochemical Water Splitting on Gold Nanoparticle
Decorated ZnO/CdS Nanotube Arrays, ACS Sustain, Chem. Eng. 5 (2017)
The authors declare that they have no known competing financial 4249–4257, https://doi.org/10.1021/acssuschemeng.7b00242.
interests or personal relationships that could have appeared to influ­ [16] C. Chen, W. Mei, C. Wang, Z. Yang, X. Chen, X. Chen, T. Liu, Synthesis of a flower-
ence the work reported in this paper. like SnO/ZnO nanostructure with high catalytic activity and stability under natural
sunlight, J. Alloys Compd. 826 (2020) 154122, , https://doi.org/10.1016/j.jallcom.
2020.154122.
Acknowledgements [17] R. Taourati, M. Khaddor, A. El Kasmi, Stable ZnO nanocatalysts with high photo­
catalytic activity for textile dye treatment, Nano-Struct. Nano-Objects 18 (2019)
100303, , https://doi.org/10.1016/j.nanoso.2019.100303.
Electron microscopy was performed using the Nova Nano-SEM 450 [18] X. Liu, C. Chen, Mxene enhanced the photocatalytic activity of ZnO nanorods under
in CFAMM at UC Riverside. This research was partially supported by a visible light, Mater. Lett. 261 (2020) 127127, , https://doi.org/10.1016/j.matlet.
2019.127127.
grant from UC Solar MR-15-328386.
[19] C. Han, Z. Chen, N. Zhang, J.C. Colmenares, Y.-J. Xu, Hierarchically CdS Decorated
1D ZnO Nanorods-2D Graphene Hybrids: Low Temperature Synthesis and Enhanced
Appendix A. Supplementary data Photocatalytic Performance, Adv. Funct. Mater. 25 (2015) 221–229, https://doi.
org/10.1002/adfm.201402443.
[20] N.A. Putri, V. Fauzia, S. Iwan, L. Roza, A.A. Umar, S. Budi, Mn-doping-induced
Supplementary data (computational details, photocatalytic statis­ photocatalytic activity enhancement of ZnO nanorods prepared on glass substrates,
tical data and electron lifetime measurement) to this article can be Appl. Surf. Sci. 439 (2018) 285–297, https://doi.org/10.1016/j.apsusc.2017.12.
246.
found online at https://doi.org/10.1016/j.apsusc.2020.147326.

7
T. Lim, et al. Applied Surface Science 536 (2021) 147326

[21] H. Zhang, J. Sun, C. Liu, Y. Wang, Distinct water activation on polar/non-polar https://doi.org/10.1103/PhysRevLett.78.1396.
facets of ZnO nanoparticles, J. Catal. 331 (2015) 57–62, https://doi.org/10.1016/j. [39] G. Kresse, D. Joubert, From ultrasoft pseudopotentials to the projector augmented-
jcat.2015.08.016. wave method, Phys. Rev. B 59 (1999) 1758–1775, https://doi.org/10.1103/
[22] Y.A. Du, Y.-W. Chen, J.-L. Kuo, First principles studies on the redox ability of PhysRevB.59.1758.
(Ga1−xZnx)N1−xOx solid solutions and thermal reactions for H2 and O2 production [40] P.E. Blöchl, Projector augmented-wave method, Phys. Rev. B 50 (1994)
on their surfaces, Phys. Chem. Chem. Phys. 15 (2013) 19807, https://doi.org/10. 17953–17979, https://doi.org/10.1103/PhysRevB.50.17953.
1039/c3cp53091d. [41] W.H. Press, S.A. Teukolsky, W.T. Vetterling, B.P. Flannery, Numerical Recipes 3rd
[23] T. Morimoto, M. Nagao, Adsorption anomaly in the system zinc oxide-water, J. Edition: The Art of Scientific Computing, 2007.
Phys. Chem. 78 (1974) 1116–1120, https://doi.org/10.1021/j100604a014. [42] T. Lim, G. Ico, K. Jung, K.N. Bozhilov, J. Nam, A.A. Martinez-Morales, Crystal
[24] S. Akhter, W.H. Cheng, K. Lui, H.H. Kung, Decomposition of methanol, for­ growth and piezoelectric characterization of mechanically stable ZnO nanostructure
maldehyde, and formic acid on nonpolar (100), stepped (501), and (0001) surfaces arrays, CrystEngComm 20 (2018) 5688–5694, https://doi.org/10.1039/
of ZnO by temperature-programmed decomposition, J. Catal. 85 (1984) 437–456, C8CE00799C.
https://doi.org/10.1016/0021-9517(84)90233-1. [43] P.X. Gao, Z.L. Wang, Substrate Atomic-Termination-Induced Anisotropic Growth of
[25] S.B.A. Hamid, S.J. Teh, C.W. Lai, Photocatalytic Water Oxidation on ZnO: A Review, ZnO Nanowires/Nanorods by the VLS Process, J. Phys. Chem. B 108 (2004)
Catalysts 7 (2017) 93, https://doi.org/10.3390/catal7030093. 7534–7537, https://doi.org/10.1021/jp049657n.
[26] D. Tang, L.F. Allard, A. Boley, D.J. Smith, J. Liu, Structure and morphology of polar [44] Y. Yu, J. Zhou, H. Han, C. Zhang, T. Cai, C. Song, T. Gao, Ab initio study of
and semi-polar pyramidal surfaces coating wurtzite ZnO micro-wires, J. Mater. Sci. structural, dielectric, and dynamical properties of zinc-blende ZnX (X=O, S, Se,
48 (2013) 3857–3862, https://doi.org/10.1007/s10853-013-7187-y. Te), J. Alloys Compd. 471 (2009) 492–497, https://doi.org/10.1016/j.jallcom.
[27] E. Debroye, J. Van Loon, H. Yuan, K.P.F. Janssen, Z. Lou, S. Kim, T. Majima, 2008.04.039.
M.B.J. Roeffaers, Facet-Dependent Photoreduction on Single ZnO Crystals, J. Phys. [45] D. Cao, W. Luo, J. Feng, X. Zhao, Z. Li, Z. Zou, Cathodic shift of onset potential for
Chem. Lett. 8 (2017) 340–346, https://doi.org/10.1021/acs.jpclett.6b02577. water oxidation on a Ti4+ doped Fe2O3 photoanode by suppressing the back re­
[28] N.S. Bell, D.R. Tallant, R. Raymond, T.J. Boyle, Synthesis and self-assembly of zinc action, Energy Environ. Sci. 7 (2014) 752–759, https://doi.org/10.1039/
oxide nanoparticles with septahedral morphology, J. Mater. Res. 23 (2008) C3EE42722F.
529–535, https://doi.org/10.1557/JMR.2008.0053. [46] P. Zhao, C.X. Kronawitter, X. Yang, J. Fu, B.E. Koel, WO3–α-Fe2O3 composite
[29] M.Z. Ahmad, J. Chang, M.S. Ahmad, E.R. Waclawik, W. Wlodarski, Non-aqueous photoelectrodes with low onset potential for solar water oxidation, Phys. Chem.
synthesis of hexagonal ZnO nanopyramids: Gas sensing properties, Sens. Actuators Chem. Phys. 16 (2014) 1327–1332, https://doi.org/10.1039/C3CP53324G.
B Chem. 177 (2013) 286–294, https://doi.org/10.1016/j.snb.2012.11.013. [47] M. Arif, G. Yasin, M. Shakeel, M.A. Mushtaq, W. Ye, X. Fang, S. Ji, D. Yan,
[30] T. Lim, A.A. Martinez-Morales, Template-free and low temperature CVD synthesis Hierarchical CoFe-layered double hydroxide and g-C3N4 heterostructures with en­
of vertically aligned 1-D ZnO nanostructures for photovoltaic devices by precursor hanced bifunctional photo/electrocatalytic activity towards overall water splitting,
oxidation protection, in, MRS Proc. (2015) 35–40, https://doi.org/10.1557/opl. Mater. Chem. Front. 3 (2019) 520–531, https://doi.org/10.1039/C8QM00677F.
2015.734. [48] J.-Y. Jung, M.J. Choi, K. Zhou, X. Li, S.-W. Jee, H.-D. Um, M.-J. Park, K.-T. Park,
[31] J.G. Lee, Computational Materials Science : an Introduction, Second Edition, Taylor J.H. Bang, J.-H. Lee, Photoelectrochemical water splitting employing a tapered si­
& Francis, Boca Raton, FL, 2016. licon nanohole array, J. Mater. Chem. A 2 (2014) 833–842, https://doi.org/10.
[32] B. Meyer, D. Marx, Density-functional study of the structure and stability of ZnO 1039/C3TA14439A.
surfaces, Phys. Rev. B 67 (2003) 035403, , https://doi.org/10.1103/PhysRevB.67. [49] M.G. Walter, E.L. Warren, J.R. McKone, S.W. Boettcher, Q. Mi, E.A. Santori,
035403. N.S. Lewis, Solar Water Splitting Cells, Chem. Rev. 110 (2010) 6446–6473, https://
[33] G. Kresse, J. Furthmüller, Efficiency of ab-initio total energy calculations for metals doi.org/10.1021/cr1002326.
and semiconductors using a plane-wave basis set, Comput. Mater. Sci. 6 (1996) [50] F. Le Formal, N. Tétreault, M. Cornuz, T. Moehl, M. Grätzel, K. Sivula, Passivating
15–50, https://doi.org/10.1016/0927-0256(96)00008-0. surface states on water splitting hematite photoanodes with alumina overlayers,
[34] G. Kresse, J. Furthmüller, Efficient iterative schemes for ab initio total-energy cal­ Chem. Sci. 2 (2011) 737–743, https://doi.org/10.1039/C0SC00578A.
culations using a plane-wave basis set, Phys. Rev. B 54 (1996) 11169–11186, [51] S.T. Mayer, R.W. Pekala, J.L. Kaschmitter, The Aerocapacitor: An Electrochemical
https://doi.org/10.1103/PhysRevB.54.11169. Double-Layer Energy-Storage Device, J. Electrochem. Soc. 140 (1993) 446, https://
[35] G. Kresse, J. Hafner, Ab initio molecular dynamics for liquid metals, Phys. Rev. B 47 doi.org/10.1149/1.2221066.
(1993) 558–561, https://doi.org/10.1103/PhysRevB.47.558. [52] D.H. Bangham, The Gibbs adsorption equation and adsorption on solids, Trans.
[36] G. Kresse, J. Hafner, Ab initio molecular-dynamics simulation of the liquid-meta­ Faraday Soc. 33 (1937) 805, https://doi.org/10.1039/tf9373300805.
l–amorphous-semiconductor transition in germanium, Phys. Rev. B 49 (1994) [53] H.H. Kristoffersen, T. Vegge, H.A. Hansen, OH formation and H2 adsorption at the
14251–14269, https://doi.org/10.1103/PhysRevB.49.14251. liquid water–Pt(111) interface, Chem. Sci. 9 (2018) 6912–6921, https://doi.org/10.
[37] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized Gradient Approximation Made 1039/C8SC02495B.
Simple, Phys. Rev. Lett. 77 (1996) 3865–3868, https://doi.org/10.1103/ [54] C. Jiang, S.J.A. Moniz, A. Wang, T. Zhang, J. Tang, Photoelectrochemical devices
PhysRevLett.77.3865. for solar water splitting – materials and challenges, Chem. Soc. Rev. 46 (2017)
[38] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized Gradient Approximation Made 4645–4660, https://doi.org/10.1039/C6CS00306K.
Simple [Phys. Rev. Lett. 77, 3865 (1996)], Phys. Rev. Lett. 78 (1997) 1396–1396.

8
Supplementary Material
High-Index Crystal Plane of ZnO Nanopyramidal Structures: Stabilization,
Growth and Improved Photocatalytic Performance
Taehoon Lima, Pegah S. Mirabedinib, Kichang Junga,c, P. Alex Greaneyb,d, Alfredo A. Martinez-Moralesa,b,c,*
a
College of Engineering - Center for Environmental Research and Technology, University of California - Riverside, Riverside, CA 92507, United States
b
Materials Science and Engineering Program, University of California - Riverside, Riverside, California 92521, United States
c
Department of Chemical and Environmental Engineering, University of California - Riverside, Riverside, California 92521, United States
d
Department of Mechanical Engineering, University of California - Riverside, Riverside, California 92521, United States

1. Computational details

The {101̅0} and {112̅2} slabs are cut out of the bulk crystal as shown in Fig. S1. Surface structures with (2×1) and
(2×2) periodicity for {101̅0} and {112̅2} are generated, resulting in a surface area of 34.87 and 71.05 Å for the
{101̅0} and {112̅2} slabs, respectively. A surface area larger than the smallest possible unit is chosen to allow for
symmetry breaking reconstruction.
A consistent set of fully converged surface energies is obtained by performing a series of energy calculations for slabs
with systematically varying slab thickness and vacuum separation. To converge the surface energy of the (101̅0) facet
with respect to slab thickness, t, slabs with 5 to 16 atomic layers (thickness of 14 to 45 Å ) are generated and fully
relaxed following the computational method described in the main article. For the (112̅2) surface energy slabs are
generated with 5 to 10 double layers (thickness of 15 to 30 Å ). To find the optimum vacuum distance and avoid
interaction between layers, slabs with a vacuum space, d, of 15 to 35 Å are tested. All atoms are allowed to fully relax
in the x- y- and z- directions. For the (101̅0) surface, the surface energy is converged for slabs with 11 or more layers
(𝑡 ≥ 31 Å ) and computed to an accuracy of 0.01 Jm-2. Convergence of the semi-polar (112̅2) surface is harder to
achieve due to the existence of an internal field set up in the polar slabs. The converged surface energy is obtained
through extrapolating the 1/𝑡 dependence of the slab energy on thickness to remove the effect of the internal field,
following the convergence method by Meyer et al [1]. As a result of this more complex procedure, the surface energy
of the (112̅2) facets is computed to a lower accuracy of 0.1 Jm-2. A vacuum spacing of 20 Å is found to be sufficient
to avoid interactions between layers. The surface energies of stoichiometric surfaces were calculated by computing
the facet energy of relaxed slabs, each containing a single pair of parallel facets by using equation 1 in the main
document [2].
To study the change in surface energy with respect to the chemical potential of oxygen, nonstoichiometric slabs, with
an excess or deficiency of oxygen, are generated by cutting the top or bottom surface from a different position in the
c-axis direction. The energy of these surfaces were computed using the energy of the stoichiometric slab using
equation 2 in the main document.
The oxygen chemical potential, 𝛥𝜇𝑂 , defines the potential of oxygen relative to half of the cohesive energy of the
1
oxygen dimer, 𝐸𝑂𝑡𝑜𝑡
2
, and is plotted over the range for which ZnO is stable. This gives the low 𝛥𝜇𝑂 limit as the
2

chemical potential at which ZnO is reduced to Zn metal and O2 gas (equation 4 in the main document.)

1
Fig. S1. Computational cell geometries of the generated {101̅0} and {112̅2} slabs, demonstrating surface areas, slabs
thickness (t), vacuum space (d), and the cell vectors in terms of unit cell lattice parameters a, b, and c (where x is an
integer number in range 5-11 for the {101̅0} surface and 5-8 for the {112̅2} slabs).

2
2. Photocatalytic statistical data

Fig. S2 shows the statistical data of the onset potential, slope and current density at 0.6 V for the
photoelectrochemical performance measurements of synthesized ZnO nanopyramids (NPy) versus nanorods (NR).
Although the results show some deviation within the same structure type, there is a noticeable difference between the
performance of the NPy and NR nanostructures. The electrochemical performance results are related to the density
and morphology of nanostructures, effective surface area and surface state [3–5]. These properties are derived from
the experimental conditions and parameters during crystal growth and are inherently affected by built-in experimental
variation in the CVD equipment setup and reaction process. Experimentally, factors such as the distance between the
source and substrate, temperature profile along the CVD reactor, mass of Zn power precursor, and O and Zn vapor
partial pressures affect the density and surface state of as-synthesized nanostructures [6–9]. The measured performance
is compared with other works reported in the literature focused on the morphology engineering of ZnO nanostructures,
as listed in Table S1.

Fig. S2. Statistical data of photoelectrochemical performance of ZnO nanopyramid (black) and nanorod (red)
structures; (a) Onset potential, (b) Slope, (c) Current density at 0.6 V, and (d) Box chart legend.

3
Table S1. Comparison of photoelectrochemical performance results in this work versus results in the literature.

Onset Slope Current at 0.6


Morphology Electrolyte Light power Reference
Potential (V) (S) V (mA/cm2)
Our nanopyramid 0.5 M Na2SO4 1.0 mW/cm2 -0.40 0.97 0.47 -
2
Our nanorod 0.5 M Na2SO4 1.0 mW/cm -0.33 0.44 0.28 -
2
Nanowire 0.01 M Na2SO4 0.4 mW/cm -0.24 0.68 0.32 [10]
Urchin-like 0.01 M Na2SO4 0.3 mW/cm2 -0.22 0.79 0.42 [11]
2
Nanoparticle 0.5 M NaClO4 100 mW/cm 0.37 0.23 0.06 [12]
2
Nanoforest 0.5 M Na2SO4 100 mW/cm -0.19 1.06 0.68 [13]
2
Tetrapod 0.5 M Na2SO4 100 mW/cm -0.22 0.53 0.34 [14]
Nanorod-nanosheet 0.5 M Na2SO4 100 mW/cm2 -0.27 1.15 0.65 [15]
2
Patterned nanorod 1.0 M Na2SO4 100 mW/cm -0.09 1.14 0.61 [16]
2
Thinfilm (002) 0.5 M Na2SO4 72 mW/cm -0.2 0.55 0.36 [17]
2
Nanotree 0.5 M Na2SO4 100 mW/cm 0 0.88 0.47 [18]
Nanotube 1.0 M KOH 100 mW/cm2 -0.36 1.48 0.91 [19]

3. Electron lifetime measurement

The transient photovoltage decay (Fig. S3(a)) is measured to calculate the electron lifetime (Fig. S3(b)) of the photo-
excited electron in the ZnO nanopyramids and nanorods using the following equation:

𝑘𝐵 𝑇 𝑑𝑉𝑂𝐶 −1
τ= ( )
𝑒 𝑑𝑡

The pyramidal ZnO shows higher stability of photo-generated carriers, resulting in a prolonged electron lifetime as
calculated from the voltage drop result. The photocatalytic performance of pyramidal ZnO is enhanced by the longer
photo-excited carrier lifetime and lower interfacial resistance (Fig. 6 in the main manuscript).

Fig. S3. (a) Open circuit voltage drop measurement, and (b) calculated electron lifetime.

4
References

[1] B. Meyer, D. Marx, Density-functional study of the structure and stability of ZnO
surfaces, Phys. Rev. B 67 (2003) 035403. https://doi.org/10.1103/PhysRevB.67.035403.

[2] J.G. Lee, Computational Materials Science : an Introduction, Second Edition, Taylor &
Francis, Boca Raton, FL, 2016.

[3] J. Oh, T.G. Deutsch, H.-C. Yuan, H.M. Branz, Nanoporous black silicon photocathode for
H2 production by photoelectrochemical water splitting, Energy Environ. Sci. 4 (2011)
1690. https://doi.org/10.1039/c1ee01124c.

[4] C.-F. Du, K.N. Dinh, Q. Liang, Y. Zheng, Y. Luo, J. Zhang, Q. Yan, Self-Assemble and
In Situ Formation of Ni1− xFexPS3 Nanomosaic-Decorated MXene Hybrids for Overall
Water Splitting, Adv. Energy Mater. 8 (2018) 1801127.
https://doi.org/10.1002/aenm.201801127.

[5] M. Li, J. Deng, A. Pu, P. Zhang, H. Zhang, J. Gao, Y. Hao, J. Zhong, X. Sun, Hydrogen-
treated hematite nanostructures with low onset potential for highly efficient solar water
oxidation, J. Mater. Chem. A 2 (2014) 6727. https://doi.org/10.1039/c4ta00729h.

[6] M. Ali, M. Winterer, Influence of Nucleation Rate on the Yield of ZnO Nanocrystals
Prepared by Chemical Vapor Synthesis, J. Phys. Chem. C 114 (2010) 5721–5726.
https://doi.org/10.1021/jp907544g.

[7] O. Lupan, V.V. Ursaki, G. Chai, L. Chow, G.A. Emelchenko, I.M. Tiginyanu, A.N.
Gruzintsev, A.N. Redkin, Selective hydrogen gas nanosensor using individual ZnO
nanowire with fast response at room temperature, Sens. Actuators B Chem. 144 (2010)
56–66. https://doi.org/10.1016/j.snb.2009.10.038.

[8] A. Menzel, K. Subannajui, R. Bakhda, Y. Wang, R. Thomann, M. Zacharias, Tuning the


Growth Mechanism of ZnO Nanowires by Controlled Carrier and Reaction Gas
Modulation in Thermal CVD, J. Phys. Chem. Lett. 3 (2012) 2815–2821.
https://doi.org/10.1021/jz301103s.

[9] P.-C. Chang, Z. Fan, D. Wang, W.-Y. Tseng, W.-A. Chiou, J. Hong, J.G. Lu, ZnO
Nanowires Synthesized by Vapor Trapping CVD Method, Chem. Mater. 16 (2004) 5133–
5137. https://doi.org/10.1021/cm049182c.

5
[10] E.S. Babu, S.-K. Hong, T.S. Vo, J.-R. Jeong, H.K. Cho, Photoelectrochemical water
splitting properties of hydrothermally-grown ZnO nanorods with controlled diameters,
Electron. Mater. Lett. 11 (2015) 65–72. https://doi.org/10.1007/s13391-014-4227-y.

[11] H.N. Hieu, N.Q. Dung, J. Kim, D. Kim, Urchin-like nanowire array: a strategy for high-
performance ZnO-based electrode utilized in photoelectrochemistry, Nanoscale 5 (2013)
5530. https://doi.org/10.1039/c3nr00889d.

[12] A. Wolcott, W.A. Smith, T.R. Kuykendall, Y. Zhao, J.Z. Zhang, Photoelectrochemical
Study of Nanostructured ZnO Thin Films for Hydrogen Generation from Water Splitting,
Adv. Funct. Mater. 19 (2009) 1849–1856. https://doi.org/10.1002/adfm.200801363.

[13] X. Sun, Q. Li, J. Jiang, Y. Mao, Morphology-tunable synthesis of ZnO nanoforest and its
photoelectrochemical performance, Nanoscale 6 (2014) 8769–8780.
https://doi.org/10.1039/C4NR01146E.

[14] N.K. Hassan, M.R. Hashim, N.K. Allam, ZnO nano-tetrapod photoanodes for enhanced
solar-driven water splitting, Chem. Phys. Lett. 549 (2012) 62–66.
https://doi.org/10.1016/j.cplett.2012.08.038.

[15] T.-F. Hou, R. Boppella, A. Shanmugasundaram, D.H. Kim, D.-W. Lee, Hierarchically
self-assembled ZnO architectures: Establishing light trapping networks for effective
photoelectrochemical water splitting, Int. J. Hydrog. Energy 42 (2017) 15126–15139.
https://doi.org/10.1016/j.ijhydene.2017.04.121.

[16] Y. Hu, X. Yan, Y. Gu, X. Chen, Z. Bai, Z. Kang, F. Long, Y. Zhang, Large-scale
patterned ZnO nanorod arrays for efficient photoelectrochemical water splitting, Appl.
Surf. Sci. 339 (2015) 122–127. https://doi.org/10.1016/j.apsusc.2015.02.074.

[17] A.U. Pawar, C.W. Kim, M.J. Kang, Y.S. Kang, Crystal facet engineering of ZnO
photoanode for the higher water splitting efficiency with proton transferable nafion film,
Nano Energy 20 (2016) 156–167. https://doi.org/10.1016/j.nanoen.2015.11.035.

[18] X. Ren, A. Sangle, S. Zhang, S. Yuan, Y. Zhao, L. Shi, R.L.Z. Hoye, S. Cho, D. Li, J.L.
MacManus-Driscoll, Photoelectrochemical water splitting strongly enhanced in fast-
grown ZnO nanotree and nanocluster structures, J. Mater. Chem. A 4 (2016) 10203–
10211. https://doi.org/10.1039/C6TA02788A.

6
[19] W.C. Lee, Y. Fang, R. Kler, G.E. Canciani, T.C. Draper, Z.T.Y. Al-Abdullah, S.M.
Alfadul, C.C. Perry, H. He, Q. Chen, Marangoni ring-templated vertically aligned ZnO
nanotube arrays with enhanced photocatalytic hydrogen production, Mater. Chem. Phys.
149–150 (2015) 12–16. https://doi.org/10.1016/j.matchemphys.2014.10.046.

You might also like