effects of surface condition on the work function and valence ......tively radio-frequency sputtered...

9
Vol.:(0123456789) 1 3 Appl. Phys. A (2017) 123:735 DOI 10.1007/s00339-017-1341-3 Effects of surface condition on the work function and valence- band position of ZnSnN 2 Amanda M. Shing 1  · Yulia Tolstova 1  · Nathan S. Lewis 2  · Harry A. Atwater 3  Received: 17 May 2017 / Accepted: 27 October 2017 / Published online: 7 November 2017 © Springer-Verlag GmbH Germany 2017 1 Introduction ZnSnN 2 is an earth-abundant semiconductor that has a direct band gap of ~ 2 eV [13]. For crystals with orthorhombic symmetry, ZnSnN 2 is predicted to have a low effective mass (< 0.16 m e [4, 5]) for electrons in the conduction band. The promise of a high-mobility semiconductor with a suitable band gap for use in photonic devices such as solar cells, LEDs, and optical sensors has driven interest in the devel- opment of ZnSnN 2 devices [59]. Alloying ZnSnN 2 with ZnGeN 2 produces ZnSn x Ge 1−x N 2 , a non-phase segregat- ing, wide-band gap alloy series for 0 ≤ x ≤ 1, that can pro- vide band gap-tunable semiconductors with applications similar to ZnSnN 2 [8]. Early fabrications of ZnSnN 2 and ZnSn x Ge 1−x N 2 were reported in 2012 [9], with subsequent improvements in synthetic methods accompanying charac- terization of the material properties of this system [13, 10, 11]. ZnSnN 2 has been synthesized by methods such as vapor–liquid–solid (VLS) plasma-assisted growth [2], reac- tive radio-frequency (RF) sputtering, and molecular-beam epitaxy (MBE) [3, 10, 11]. VLS growth yields discontinuous islands of ZnSnN 2 on top of Zn–Sn from a melt, whereas sputtering provides more homogeneous films of ZnSnN 2 . Both VLS and sputtering yield crystallites of sizes on the order of 100 nm or less. Orthorhombic crystal structures have been observed for ZnSnN 2 samples grown using sput- tering methods on various substrates and using MBE on lith- ium gallate [1, 12, 13]. Monoclinic crystal structures have also been observed for ZnSnN 2 thin films grown by MBE on yttrium-doped zirconia substrates, although droplets of metal residue on the film surfaces accompany the formation of sin- gle-crystalline ZnSnN 2 [11]. The ZnSnN 2 samples prepared to date exhibit poor semiconductor properties, with major- ity-carrier concentrations at or above degenerate doping Abstract ZnSnN 2 is an emerging wide band gap earth- abundant semiconductor with potential applications in pho- tonic devices such as solar cells, LEDs, and optical sensors. We report the characterization by ultraviolet photoelectron spectroscopy and X-ray photoelectron spectroscopy of reac- tively radio-frequency sputtered II–IV-nitride ZnSnN 2 thin films. For samples transferred in high vacuum, the ZnSnN 2 surface work function was 4.0 ± 0.1 eV below the vacuum level, with a valence-band onset of 1.2 ± 0.1 eV below the Fermi level. The resulting band diagram indicates that the degenerate bulk Fermi level position in ZnSnN 2 shifts to mid-gap at the surface due to band bending that results from equilibration with delocalized surface states within the gap. Brief (< 10 s) exposures to air, a nitrogen-plasma treatment, or argon-ion sputtering caused significant chemical changes at the surface, both in surface composition and interfacial energetics. The relative band positioning of the n-type semi- conductor against standard redox potentials indicated that ZnSnN 2 has an appropriate energy band alignment for use as a photoanode to effect the oxygen-evolution reaction. Electronic supplementary material The online version of this article (https://doi.org/10.1007/s00339-017-1341-3) contains supplementary material, which is available to authorized users. * Amanda M. Shing [email protected] 1 Department of Materials Science, California Institute of Technology, Pasadena, CA 91125, USA 2 Division of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, CA 91125, USA 3 Department of Applied Physics, California Institute of Technology, Pasadena, CA 91125, USA

Upload: others

Post on 09-Mar-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Effects of surface condition on the work function and valence ......tively radio-frequency sputtered II–IV-nitride ZnSnN 2 thin films. For samples transferred in high vacuum, the

Vol.:(0123456789)1 3

Appl. Phys. A (2017) 123:735 DOI 10.1007/s00339-017-1341-3

Effects of surface condition on the work function and valence-band position of  ZnSnN2

Amanda M. Shing1 · Yulia Tolstova1 · Nathan S. Lewis2 · Harry A. Atwater3 

Received: 17 May 2017 / Accepted: 27 October 2017 / Published online: 7 November 2017 © Springer-Verlag GmbH Germany 2017

1 Introduction

ZnSnN2 is an earth-abundant semiconductor that has a direct band gap of ~ 2 eV [1–3]. For crystals with orthorhombic symmetry, ZnSnN2 is predicted to have a low effective mass (< 0.16 me [4, 5]) for electrons in the conduction band. The promise of a high-mobility semiconductor with a suitable band gap for use in photonic devices such as solar cells, LEDs, and optical sensors has driven interest in the devel-opment of ZnSnN2 devices [5–9]. Alloying ZnSnN2 with ZnGeN2 produces ZnSnxGe1−xN2, a non-phase segregat-ing, wide-band gap alloy series for 0 ≤ x ≤ 1, that can pro-vide band gap-tunable semiconductors with applications similar to ZnSnN2 [8]. Early fabrications of ZnSnN2 and ZnSnxGe1−xN2 were reported in 2012 [9], with subsequent improvements in synthetic methods accompanying charac-terization of the material properties of this system [1–3, 10, 11].

ZnSnN2 has been synthesized by methods such as vapor–liquid–solid (VLS) plasma-assisted growth [2], reac-tive radio-frequency (RF) sputtering, and molecular-beam epitaxy (MBE) [3, 10, 11]. VLS growth yields discontinuous islands of ZnSnN2 on top of Zn–Sn from a melt, whereas sputtering provides more homogeneous films of ZnSnN2. Both VLS and sputtering yield crystallites of sizes on the order of 100 nm or less. Orthorhombic crystal structures have been observed for ZnSnN2 samples grown using sput-tering methods on various substrates and using MBE on lith-ium gallate [1, 12, 13]. Monoclinic crystal structures have also been observed for ZnSnN2 thin films grown by MBE on yttrium-doped zirconia substrates, although droplets of metal residue on the film surfaces accompany the formation of sin-gle-crystalline ZnSnN2 [11]. The ZnSnN2 samples prepared to date exhibit poor semiconductor properties, with major-ity-carrier concentrations at or above degenerate doping

Abstract ZnSnN2 is an emerging wide band gap earth-abundant semiconductor with potential applications in pho-tonic devices such as solar cells, LEDs, and optical sensors. We report the characterization by ultraviolet photoelectron spectroscopy and X-ray photoelectron spectroscopy of reac-tively radio-frequency sputtered II–IV-nitride ZnSnN2 thin films. For samples transferred in high vacuum, the ZnSnN2 surface work function was 4.0 ± 0.1 eV below the vacuum level, with a valence-band onset of 1.2 ± 0.1 eV below the Fermi level. The resulting band diagram indicates that the degenerate bulk Fermi level position in ZnSnN2 shifts to mid-gap at the surface due to band bending that results from equilibration with delocalized surface states within the gap. Brief (< 10 s) exposures to air, a nitrogen-plasma treatment, or argon-ion sputtering caused significant chemical changes at the surface, both in surface composition and interfacial energetics. The relative band positioning of the n-type semi-conductor against standard redox potentials indicated that ZnSnN2 has an appropriate energy band alignment for use as a photoanode to effect the oxygen-evolution reaction.

Electronic supplementary material The online version of this article (https://doi.org/10.1007/s00339-017-1341-3) contains supplementary material, which is available to authorized users.

* Amanda M. Shing [email protected]

1 Department of Materials Science, California Institute of Technology, Pasadena, CA 91125, USA

2 Division of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, CA 91125, USA

3 Department of Applied Physics, California Institute of Technology, Pasadena, CA 91125, USA

Page 2: Effects of surface condition on the work function and valence ......tively radio-frequency sputtered II–IV-nitride ZnSnN 2 thin films. For samples transferred in high vacuum, the

A. M. Shing et al.

1 3

735 Page 2 of 9

levels [1, 3], and with low mobilities (< 10 cm2 V−1 s−1) [1]. Only recent work using molecular-beam epitaxy has yielded ZnSnN2 with higher mobilities. Explorations in the off-stoichiometric compound Zn1+xSn1−xN2 have also shown to influence the band gap [10] and lower doping levels to < 1018 carriers per cm3 [14]. More development of growth techniques that improve the quality of ZnSnN2 samples will provide greater insight into the potential of this material for device applications.

The work function is a critical property for designing and incorporating a material into optoelectronic devices. Knowl-edge of the work function at the surface of a semiconductor, in combination with knowledge of the band gap and position of the valence band relative to the Fermi level, allows deter-mination of the absolute energetic positions of the Fermi level as well as of the electronic bands at the surface. This information allows evaluation of the likely suitability of possible solid-state or liquid contacts based on the relative disposition of their Fermi levels, energy bands, or electro-chemical potentials of the contacting phases. Knowledge of the work function also assists in predicting the electronic behavior—equilibration of the electrochemical potentials and band bending—at the interface with a second material, by the Anderson technique for semiconductor–semiconduc-tor interfaces or by the Schottky–Mott technique for inter-faces between the semiconductor and metals or redox-active liquids [15, 16]. Work function values also provide informa-tion about the appropriate band alignment for the construc-tion of homojunctions, heterojunctions, or photoelectrodes [15–17]. These alignments, combined with band gaps, car-rier densities, and other material descriptors, are used to predict whether materials will form Type I, Type II, or Type III junctions and to estimate the carrier dynamics at junc-tions. Knowledge of the relative positions between the Fermi levels and band edges of a semiconductor and a redox couple in solution also provides information about the electron- and hole-transfer possibilities across a liquid junction, and thus about the suitability of a material for use as a photoanode or photocathode for photoelectrochemistry [17–20].

The work function can be measured using photoelectron spectroscopy (PES) or Kelvin Probe Force Microscopy (KPFM). In PES, an X-ray or ultraviolet (UV) radiation source excites electrons in the material, and the energies of electrons ejected from the material via the photoelectric effect are determined using a spectrometer. The work func-tion of the material is determined from the threshold energy at which electrons are barely emitted from the material, in conjunction with knowledge of the initial excitation energy [21].

Conversely, KPFM measures a contact potential differ-ence (CPD) between the probe tip and the material surface. The work function of the material can then be calculated rel-ative to the work function of the probe tip. Both techniques

require a clean surface to accurately determine the work function of the material, because any contamination, oxida-tion, or adsorbed molecules will affect the surface electro-chemical potential [22–24]. Thus, experiments performed in ultra-high vacuum (UHV) are preferred, in addition to sample preparation that minimizes surface contamination. In situ cleaning, such as sputter ablation or annealing, can remove contaminants. Crystal cleavage in UHV, or material fabrication in a cluster tool attached to the analysis instru-ment, is also an option for creating clean surfaces, although cluster tools are costly and scarce. Furthermore, the work function changes with crystal orientation, so inhomogene-ous, polycrystalline, or defective surfaces generally exhibit mutually different work functions even for the same bulk material [24, 25].

We report herein the work function and electronic struc-ture of thin-film ZnSnN2 as determined by PES in UHV. We have separately determined the ZnSnN2 work function using both UV and X-ray excitation energies, and compare measurements for vacuum-transferred samples to measure-ments of samples exposed to air, in situ nitrogen plasma treatment, or argon-ion sputtering, respectively. The spectra also have allowed determination of the energy position of the valence-band maximum (VBM) which, in combination with the measured band gap, enables formulation of a band diagram for ZnSnN2 and its interface to vacuum.

2 Methods

Thin-film samples were fabricated by reactive RF sputter-ing from separate Zn and Sn metal targets. Each target was 99.99+% pure, with RF power of the Zn target held at 44 W and RF power of the Sn target held at 74 W during sput-tering. Before deposition, epitaxial-ready c-plane sapphire and GaN template substrates were rinsed with isopropanol, dried with a stream of N2(g), and heated to 175 °C in the vacuum sputter chamber at a base pressure of 5 × 10−8 Torr. Before film deposition, polished, degenerately doped (111) n-Si substrates were etched with HF(aq) to remove surface oxides. The Si substrates were then dried with a stream of N2(g) and heated to 175 °C. During sputtering, the plasma in the chamber was maintained by a 1:3 Ar:N gas flow ratio at a pressure of 3 mTorr. The sputtered films were between 300 and 900 nm thick.

Except as noted, samples were transferred from the sput-tering chamber to a Kratos Axis Ultra spectrometer using a vacuum suitcase (< 3 × 10−7 Torr). The nitrogen plasma treatment, argon-ion sputter etch, and vacuum annealing were also performed in the Kratos spectrometer.

X-ray photoelectron spectroscopy (XPS) was performed with Al Kα X-rays (hυ = 1486.7 eV) in an analysis cham-ber that had a base pressure < 3 × 10−9 Torr. Ultraviolet

Page 3: Effects of surface condition on the work function and valence ......tively radio-frequency sputtered II–IV-nitride ZnSnN 2 thin films. For samples transferred in high vacuum, the

Effects of surface condition on the work function and valence-band position of  ZnSnN2

1 3

Page 3 of 9 735

photoelectron spectroscopy (UPS) was performed with He I excitation obtained from a UV lamp that was mounted on the analysis chamber and directed at the sample. A strong peach-color light was attained for He I energy (hυ = 21.21 eV). Spectra of the valence-band region to the secondary-electron cut-off region were acquired at a 25 meV step size and a 5 eV pass energy at 300 μm × 700 μm, 110 μm diameter, and 55 μm diameter aperture sizes. XPS secondary-electron cut-off regions were acquired at a 25 meV step size and a 5 eV pass energy, with the X-ray gun held at 0 mA and 15 keV, to not overload the analyzer. Work functions (WF) were calculated by taking the secondary-electron cut-off energies (SEC) and subtracting them from the excitation energies (hυ) (Eq. 1). SECs shifted as expected with changes in the applied external bias, indicating that the SECs were representative of the behavior of sample as opposed to being experimental artifacts.

XPS survey scans were taken with a 1 eV step size, 100 ms collection time, 40 eV pass energy, and using a ‘slot’ aperture. High-resolution scans of the valence-band region and of the carbon, oxygen, zinc, tin, germanium, and nitro-gen peaks were acquired with a 25 meV step size, 10 eV pass energy, and using a 300 μm × 700 μm aperture. Peaks were analyzed with Casa XPS software using Gaussian–Lor-entzian shapes and Shirley baselines [SI]. Stoichiometries were calculated from the Zn 2p, Sn 3d, N 1s, C 1s, and O 1s peaks.

After PES analysis, samples were removed from vac-uum and subjected to analysis by X-ray diffraction (XRD). The diffraction patterns (Fig. S1) from the samples were collected and analyzed using a P’Analytical brand high-resolution X-ray diffractometer employing Cu Kα radia-tion (λ = 1.5418 Å). The stoichiometry of the samples was measured using a scanning-electron microscope (SEM) equipped with an energy-dispersive spectrometer (EDS). The dielectric functions were determined by spectroscopic ellipsometry, with the data modeled using a wavelength-by-wavelength layer on top of a substrate layer. Hall measure-ments were performed to obtain carrier concentrations.

3 Results

Figure 1a shows a UPS spectrum for vacuum-transferred ZnSnN2 samples, as compared with UPS spectra for ZnSnN2 samples exposed to air for < 10 s (short expo-sure), to air for < 1 min (long exposure), to a nitrogen plasma treatment, or to argon-ion sputtering, respectively. For vacuum-transferred samples, the sharp increase in the peak intensity at a binding energy of ~ 17 eV is the secondary-electron cut-off (SEC). The fine structure at

(1)WF = h� − SEC.

10.1 eV represents the Zn 3d states [26], and the increase in counts at ~ 1.2–3 eV is the valence-band onset. The prominence of the Zn 3d feature in the UPS attests to the cleanliness of the surface of the vacuum-transferred sam-ple. The reduced fine structure in the air-exposed sample (Fig. 1a) is consistent with the presence of contamination overlayers, while the shift in the SEC and VB onset rep-resents a change in the work function and in the valence-band position.

The work function measured for vacuum-transferred ZnSnN2 samples was 4.04 ± 0.09 eV (Table 1), with 11 sam-ples measured. Spectra were collected at several locations on the surface of each sample, and were consistent to within a standard deviation of 2 meV (Figure S2). All of the work function measurements for vacuum-transferred ZnSnN2 samples fell within a range of 0.3 eV. With the exception of one sample, all of the vacuum-transferred samples exhib-ited a preferential crystalline orientation toward the (002) ZnSnN2 direction. The one sample having multiple orienta-tions in the (002), (201), (211), and (021) directions exhib-ited a work function of 3.85 ± 0.01 eV, which was the lowest work function measured of any sample.

Exposure of the ZnSnN2 samples to air resulted in an increase of > 1.2 eV in the binding energy associated with the valence-band maxima (VBM). This result is consistent with surface oxidation of the nitride, forming various oxides or oxynitrides, because the N 2p contribution to the valence-band states penetrates more deeply into the band gap than the O 1s contribution [27, 28].

Exposure to air also resulted in a shift to lower binding energies in the position of the SEC, indicating an increase in the work function [average 0.3 eV increase (Table 1)]. As with surfaces exposed to air, relative to vacuum-transferred samples, the samples cleaned with a nitrogen plasma exhib-ited a shift in the SEC to lower binding energies. However, samples cleaned with a nitrogen plasma showed a shift of ~ 2 eV in the SEC, resulting in an increase in the work func-tion to > 6 eV. These values are significantly higher than those observed for air-exposed samples, and a 1 eV increase in the valence-band onset was also observed.

Vacuum-transferred samples (Fig. 1b) showed continu-ous UPS electron counts between the valence-band onset and 0 eV binding energy. In addition to increases in the binding energy of the valence-band onset (VBO), air expo-sure eliminated the continuous electron emission from the valence-band onset to the Fermi level at 0 eV binding energy. Although in the air-exposed spectra emission still exists above the valence-band onset, it is difficult to discern whether these electron counts originate from gap states as opposed to the surface composition representing a com-bination of ZnSnN2 (VBO ~ 1.2 eV), ZnO (VBO ~ 2.2 eV [29]), SnO2 (VBO ~ 2.8 eV – 4 eV [30]), and/or various oxynitrides.

Page 4: Effects of surface condition on the work function and valence ......tively radio-frequency sputtered II–IV-nitride ZnSnN 2 thin films. For samples transferred in high vacuum, the

A. M. Shing et al.

1 3

735 Page 4 of 9

Figure 2 shows high-resolution XPS data for the N 1s, Zn 2p, Sn 3d, and valence-band regions for vacuum-transferred ZnSnN2 samples as well as for ZnSnN2 samples that had been cleaned with a nitrogen plasma or cleaned by Ar-ion sputtering for at least 1 min. The XP spectra in the N 1s region for vacuum-transferred samples exhibited a single peak at 397 eV, which shifted to lower binding energies (396.5 eV) for the plasma-cleaned or ion-sputtered samples. A second peak was observed in the N 1s spectra for either plasma-cleaned samples (402 eV) or Ar-ion-sputtered sam-ples (404.5 eV). For the nitrogen-plasma-cleaned samples, the peak in the N 1s region (402 eV) was characteristic of

nitrogen donating electrons to oxygen (N–O bonding) [28]. For argon-ion-sputtered samples, the peak in the N 1s region (404.5 eV) was indicative of N=O bonding [31]. XP spectra of ZnSnN2 samples exposed to air did not exhibit a peak at 402 eV.

The Zn 2p spectra show the Zn 2p 1/2 (1044 eV) and Zn 2p 3/2 (1022 eV) spin–orbit split states, with an Sn Auger peak [32, 33] at ≥ 1050 eV. Similarly, the Sn 3d 3/2 (495 eV) and 3d 5/2 (485 eV) peaks are present for the vacuum-trans-ferred samples, with a Zn LMM Auger shoulder at 497 eV.

Both Zn 2p and Sn 3d peaks of the plasma-cleaned and Ar-ion-sputtered samples shifted relative to the spectra of

Fig. 1 Comparison of UPS spectra for vacuum-transferred ZnSnN2 samples with spectra obtained for ZnSnN2 samples after a air exposure, c argon-ion sputter treatment, and d nitrogen plasma treatment. Short air exposure was for < 10 s in atmosphere, and long air exposure was for < 1 min. The spectra were obtained using a He I source (hυ = 21.21 eV). b UPS spectrum of the ZnSnN2 valence-band region. Significant electron emission was observed between the VBM and the Fermi level, indicating occupied surface states in the gap region for vacuum-transferred samples. Air-exposed surfaces did not retain continuous emission extending to the Fermi level at 0 eV, indicating that oxygen passivates the lower binding energy gap states in form-ing various metal oxides or oxynitrides

SnO2 (VBO ~2.8 eV – 4 eV [30]), and/or various oxynitrides.

a b

c d

Table 1 Average values for work functions (WF) and valence-band (VB) onsets obtained from UPS and XPS, and stoichiometry at the ZnSnN2 surface, including adventitious carbon and oxygen surface coverage

Error values shown are one standard deviationa Tailing states from gap to extrapolated VB max

UPS WF(eV) XPS WF(eV) % O % C XPS VB onset (eV) UPS VB onset (eV)

Vacuum transferred 4.04 ± 0.09 4.03 ± 0.09 4 ± 1 2 ± 1 1.20 ± 0.14 1.22 ± 0.09Air exposure 4.3 ± 0.2 4.0 ± 0.16 25 ± 10 27 ± 5 1.37 ± 0.09 3.44 ± 0.39Plasma cleaned 6.40 ± 0.04 N/A 36 ± 3 0 1.23 ± 0.08 2.43 ± 0.02a

Page 5: Effects of surface condition on the work function and valence ......tively radio-frequency sputtered II–IV-nitride ZnSnN 2 thin films. For samples transferred in high vacuum, the

Effects of surface condition on the work function and valence-band position of  ZnSnN2

1 3

Page 5 of 9 735

vacuum-transferred samples. The shift in the peaks in the N 1s, Zn 2p, and Sn 3d regions, as well as the appearance of new peaks at high binding energies in the N 1s region, are consistent with chemical changes ascribable to surface oxidation [34]. ZnSnN2 is a ternary compound in which ide-ally Sn is bonded to both N and Zn atoms. However, defects including vacancies, anti-site substitutions, and impurities may be involved in bonding, which will affect the ideal sys-tematic core-level shifts [34].

Sputtering resulted in a loss of nitrogen from the sur-face, as evidenced by a reduction in the integrated area of the nitrogen peak (Fig. 2; Table 2). A loss of nitrogen from Ar-ion-sputtered samples was also apparent in the Sn 3d

doublet, where each peak could be decomposed into signals arising from multiple oxidation states, with increasing pref-erence toward neutral tin, indicating that a change in bond-ing had occurred between the tin cations and the nitrogen anions. Samples that had been sputtered for 1 min showed a shift in the valence-band onset to 0 eV and a continuous electron population to the Fermi level at 0 eV (Fig. 1c), indi-cating that the surfaces of sputtered samples were metallic in nature. Argon-ion sputter cleaning performed for shorter times (< 10 s) showed mitigated effects; however, changes in surface stoichiometry were still apparent and the surface was not of the same quality as the vacuum-transferred ZnSnN2 samples (Table 2).

Fig. 2 High-resolution XPS data for the N 1s, Zn 2p, Sn 3d, and valence-band regions of freshly vacuum-transferred samples and Ar+ sputter-treated samples

Table 2 Surface composition and UPS values for argon-ion sputter-cleaned surfaces from one representative sample. Values in parenthesis are the raw integrated peak areas [SI]. Due to the decrease in proportion of nitrogen, other elements show increased contribution in the composition

% N % Zn % Sn % O % C UPS WF (eV) UPS VB onset (eV)

Vacuum transferred 51 (14,988) 22 (76,832) 20 (100,004) 3 (1692) 2 (305) 4.08 1.17< 10 s 46 (8374) 25 (53,769) 22 (69,009) 4 (1309) 3 (258) 4.15 0.58~ 1 min 40 (7766) 27 (61,494) 26 (83,132) 5 (1280) 1 (118) 4.26 0

Page 6: Effects of surface condition on the work function and valence ......tively radio-frequency sputtered II–IV-nitride ZnSnN 2 thin films. For samples transferred in high vacuum, the

A. M. Shing et al.

1 3

735 Page 6 of 9

Sputtering can change the chemical composition of the material and the band gap. When surface nitrogen is lost, the surface valence band terminates at the Fermi level (no band gap). The sputter spot could readily be visually observed, and the thin film in that region had a reflective, metallic tint, in accord with expectations for a change in the band gap after sputtering. The values of the band gap measured herein were obtained far from the sputtered region, and are similar to ellipsometric methods used in Ref. [1], and bulk meth-ods used in Refs. [2, 3, 10] for near-stoichiometric ZnSnN2 Eg ~ 1.8–2.2 eV.

Table 1 summarizes the work functions and valence-band onsets measured using UPS and XPS, as well as the stoichi-ometry of the ZnSnN2 surface including oxygen and carbon contamination as determined from XPS, for vacuum-trans-ferred, air-exposed, plasma-cleaned, and Ar-ion-sputtered samples, respectively. XPS penetrates deeper into the mate-rial bulk (~ 10 nm for Al Kα (1486.7 eV excitation)) than UPS (~ 1–2 nm), so XPS measurements, therefore, exhibit less convolution with surface states and satellite peaks than UPS measurements. Hence, measurement of the VBM by XPS is more representative than measurement by UPS [35]. The UPS measurements of the VBM demonstrate the surface effects caused by the different treatments, which are less apparent in the XPS data due to the increase in penetration depth.

Veal et al. have performed hard X-ray photoemission spectroscopy using 4000  eV excitation energy on their molecular-beam-epitaxy-grown ZnSnN2 thin films [10] for greater penetration depth, achieving an average valence band onset of ~ 2 eV from two samples. Although their values are within 1 eV of the derived values in this work, it is unknown if incorporation of oxidation, crystal quality, or orientation had influenced the valence band onset obtained in Ref. [10].

The average surface contamination for the vacuum-trans-ferred ZnSnN2 samples was 3% C and 4% O, as determined by XPS. Exposure to air for more than a day increased the surface contamination to 25% O and 27% C. Nitrogen-plasma-cleaned samples showed 0% C on the surface, but showed an increase in O to 36%. Argon-ion-sputtered sam-ples showed only a slight increase (to 5%) in O contamina-tion, and showed 1% C; however, these samples showed a significant decrease in N content (from 51 to 40%).

Figure 3 shows the work function progression as well as the C and O coverage of a sample that had undergone suc-cessive air exposures. This increasing trend correlates with the surface oxygen concentration, and agrees with greater percentages of detected Zn–O and Sn–O bonding. ZnO has a work function range of 4.3–5 eV [36–38] and SnO2 has a work function range of 4.7–5.7 eV [39, 40], depend-ing on surface preparation. Thus, the average work func-tion after each air exposure is expected to be higher when additional metal oxide is incorporated in the measured area.

This agreement is shown over several samples in Figure S3, in which the work functions of freshly vacuum-transferred samples are compared to the values observed after air expo-sure. The coverage of % O and % C increased from ~ 4% O to > 20% O, and the work functions increased to an average of 4.3 ± 0.2 eV (Table 1).

The optical band gap of the ZnSnN2 samples was 1.9 ± 0.2 eV, as determined from spectroscopic ellipsom-etry [SI]. Hall-effect measurements yielded carrier con-centrations of (3.6 ± 3.2) × 1020 cm−3, and mobilities of 2.4 ± 1.2 cm2 V−1 s−1. Error estimates reflect one standard deviation of the total 11-sample population.

4 Discussion

The influences of air exposure and in situ cleaning tech-niques were examined carefully to ensure that the work functions and valence-band maxima were reflective of clean ZnSnN2 surfaces rather than of an oxide or other contami-nated forms of ZnSnN2. Even brief (< 10 s) exposure to air resulted in oxidation of the ZnSnN2 surface. Surface oxida-tion was accompanied by an increase in the VBM and by a shift of the SEC to lower binding energy, thus increasing the work function. Attempts to eliminate surface contamination with in situ cleaning techniques, specifically nitrogen-plasma cleaning and argon-ion sputtering, caused chemical changes to the surface, as evidenced the UPS and XPS spectra. The nitrogen-plasma cleaning was performed in the load-lock chamber of the photoelectron spectrometer, and due to residual oxygen in the load-lock chamber, oxygen could not be eliminated from the ZnSnN2 surface using this method. Although argon-ion sputtering was used at the lowest energy (500 eV) available for our instrument, the treatment resulted in a loss of nitrogen from the surface and produced sam-ples that had increased metallic character. Not only did the work functions and valence-band onsets obtained for freshly

Fig. 3 Progression of the work function and of total C and O cover-age of a ZnSnN2 sample following successive exposures to air and for a nitrogen-plasma-cleaned sample

Page 7: Effects of surface condition on the work function and valence ......tively radio-frequency sputtered II–IV-nitride ZnSnN 2 thin films. For samples transferred in high vacuum, the

Effects of surface condition on the work function and valence-band position of  ZnSnN2

1 3

Page 7 of 9 735

vacuum-transferred samples best represent the ZnSnN2 surface, but exposures to air or in situ cleaning techniques resulted in significant changes, with air exposure increas-ing the valence-band maximum by 170 meV relative to the average value for vacuum-transferred samples (1.20 eV) by XPS, and plasma treatment increasing the UPS-derived work function by 2.36 eV relative to the average value for vacuum-transferred samples (4.04 eV) (Table 1).

For freshly vacuum-transferred samples, the valence-band maximum was 1.20 eV from the Fermi energy at 0 eV bind-ing energy. Combining the VBM with the measured band gap of ~ 1.7–2.1 eV places the Fermi level in the gap ~ 70% of the way from the valence band to the conduction band, indicating that the majority carriers are electrons. This Fermi level position agrees with the carrier type determined by Hall or hot-probe measurements. However, Hall measure-ments indicated carrier densities on the order of 1020 cm−3, implying that the Fermi energy should be in the degener-ate regime, close to the conduction band if the films were purely n-type, or in the conduction band if films exhibited the Burstein–Moss effect [1, 10]. The apparent discrepancy between the position of the Fermi level within the band gap as determined by PES and the degeneracy observed in the Hall measurements can consistently be ascribed to the influence of delocalized, mid-gap surface states that are observed preferentially using the surface-sensitive PES tech-niques, while Hall measurements are more reflective of bulk properties.

The ZnSnN2 UPS spectra showed continuous electron counts at energies between the valence-band onset and past the Fermi level at 0 eV binding energy. Although instrumen-tal broadening may account for small errors in the valence-band onset, the instrumental broadening Gaussian contribu-tion should be negligible in the region close the Fermi level [SI], and emission is representative of electron states. Metals also show counts to 0 eV binding energy in their valence-band spectra, albeit at much higher intensities than seen for ZnSnN2, because metal valence electron states are filled to

the Femi level. For ZnSnN2, the electron counts seen in the region between the valence-band onset and 0 eV arise from electrons with energies inside the band gap, and are indica-tive of the existence of mid-gap surface states [41–45].

Figure 4 presents a band diagram for the ZnSnN2/vacuum interface, including the effect of mid-gap surface states. Prior to equilibration, the Fermi level is found in a degener-ate position, but equilibration with mid-gap surface states results in band bending and a shift in the Fermi level toward a mid-gap position near the surface. After equilibration, the bulk material stays degenerate, with its Fermi level close to the conduction band, but equilibration with surface states causes the bands to bend such that the equilibrated Fermi position at the surface is closer to the center of the gap [25, 46–48].

The UPS spectra (Fig. 1b) indicated that oxygen pas-sivated the lower binding energy surface states, resulting in reduced electron-count intensity extending to the Fermi level after air exposure [46, 47]. Additionally, the distance between the Fermi level and the valence-band onset wid-ened upon air exposure, with the average XPS valence-band onset increasing by 170 meV (Table 1). This result is con-sistent with surface oxidation of the nitride, because the N 2p contribution to the valence-band states penetrates more deeply into the band gap than the O 1s contribution [27, 28]. Notably, the oxidized UPS valence-band onset changed to greater than 3 eV, surpassing the 2 eV ZnSnN2 band gap, but is only representative of the top ~ 2 nm of oxidized ZnSnN2. The density of surface states and the band bending could be determined from the bulk density of states in conjunc-tion with the carrier density, but the bulk DOS for thin-film ZnSnN2 has not yet been obtained. A crude approximation may be formed from the calculated theoretical DOS for sin-gle crystals [4, 5].

Although the density of surface states, their energies, and the amount band bending in the ZnSnN2 are unknown, Fig. 5 displays the current ZnSnN2 work function measure-ment on the vacuum energy and normal hydrogen electrode

Fig. 4 Energy band diagram for the ZnSnN2/vacuum interface prior to equilibration with mid-gap surface states (left) and after equilibration (right). Electrons from the bulk fill the empty surface states, caus-ing the bands to bend upward, reducing the valence band onset at the surface from a degenerate position to 1.2 eV

Page 8: Effects of surface condition on the work function and valence ......tively radio-frequency sputtered II–IV-nitride ZnSnN 2 thin films. For samples transferred in high vacuum, the

A. M. Shing et al.

1 3

735 Page 8 of 9

(NHE) scales. The figure also compares ZnSnN2 with other common semiconductors as well as with the electrochemical potentials for the water-splitting half-reactions. The ZnSnN2 valence band and conduction band are approximately posi-tioned to associate with the material’s measured 2 eV band gap and 1020 cm−3 n-type carrier concentration, placing the ZnSnN2 Fermi level close to the conduction band. This arrangement shows that the valence band of ZnSnN2 may lie in an energetically favorable position for a photoanode to effect the oxygen-evolution reaction, which has an electro-chemical potential of ~ 5.7 eV below vacuum.

5 Conclusion

Using UPS, the work function of reactive RF sputtered ZnSnN2 thin films was determined to be 4.04 ± 0.09 eV versus vacuum, with a VBM at 1.22 ± 0.09 eV below the Fermi level, which agrees with XPS-derived measurements. Although minor surface contamination was observed (< 4% O, C), the data for freshly vacuum-transferred samples most accurately represented the work function of the ZnSnN2 surface. More surface contamination yielded higher work functions (4.3 ± 0.2 eV for air-exposed and 6.40 ± 0.04 eV for plasma-cleaned samples) and increased the valence-band onset with respect to the Fermi energy (by 170 meV for air-exposed samples), as expected for greater percentages of SnO2 and ZnO at the sample surface. In situ cleaning tech-niques created chemically altered surfaces, as manifested in XPS spectra by either loss of nitrogen or nitrogen–oxygen bonding, and in the UPS spectra by shifts in the SECs as well as in the valence-band onsets.

The work function and valence-band maxima, combined with band gap measurements, yielded the band positions for the ZnSnN2 surface. Surface states cause band bending that positions the ZnSnN2 Fermi level towards mid-gap at the surface, instead of the bulk degenerate position, typical of n-type semiconductor surfaces. The relative band alignment of ZnSnN2 versus vacuum is valuable as it allows predic-tions of the ZnSnN2 interactions with the electrochemical potentials of other semiconductors, metals, and redox cou-ples. The ZnSnN2 PES data additionally provide guidance for future surface preparation studies to create quality inter-faces for devices.

Acknowledgements We gratefully acknowledge support from the Dow Chemical Company under the earth-abundant semiconductor pro-ject. We also acknowledge the Joint Center for Artificial Photosynthesis and the Molecular Materials Research Center of the Beckman Institute at Caltech for instrument access. The authors thank Bruce Brunschwig and Kimberly Papadantonakis for guidance.

Compliance with ethical standards

Funding sources Dow Chemical Company, Caltech Molecular Mate-rials Research Center.

References

1. L. Lahourcade, N. Coronel, K. Delaney, S.K. Shukla, N.A. Spal-din, H.A. Atwater, Adv. Mater. 25, 2562 (2013)

2. P. Quayle, K. He, J. Shan, K. Kash, MRS Commun. 3, 135 (2013) 3. N. Feldberg, J.D. Aldous, W.M. Linhart, L.J. Phillips, K. Durose,

P.A. Stampe, R.J. Kennedy, D.O. Scanlon, G. Vardar, R.L. Field

Fig. 5 ZnSnN2 Fermi level alignment with various common semiconductors and redox cou-ples on the vacuum energy scale and normal hydrogen electrode (NHE) scale. In vacuum, the ZnSnN2 valence band at the surface would be 1.2 eV below the Fermi level; however, the bulk ZnSnN2 valence band is expected to be further below the Fermi energy. The valence-band position for bulk ZnSnN2 aligns appropriately for the oxygen-evolution reaction at 1.23 V versus the hydrogen reduction potential. Adapted with permis-sion from Ref. [49]

Page 9: Effects of surface condition on the work function and valence ......tively radio-frequency sputtered II–IV-nitride ZnSnN 2 thin films. For samples transferred in high vacuum, the

Effects of surface condition on the work function and valence-band position of  ZnSnN2

1 3

Page 9 of 9 735

III, T.Y. Jen, R.S. Goldman, T.D. Veal, S.M. Durbin, Appl. Phys. Lett. 103, 042109 (2013)

4. K.T. Delaney, S.K. Shukla, N.A. Spaldin, in First-Principles Theoretical Assessment of Earth-Abundant Nitrides for Photo-voltaic and Optoelectronic Applications using Hybrid Density Functionals

5. A. Punya, W.R.L. Lambrecht, Phys. Rev. B 84, 165204 (2011) 6. A. Punya, T.R. Paudel, W.R.L. Lambrecht, Phys. Status Solid C

8, 2492 (2011) 7. A. Punya, W.R.L. Lambrecht, Phys. Rev. B 88, 075302 (2013) 8. A.M. Shing, N.C. Coronel, N.S. Lewis, H.A. Atwater, APL Mater.

3, 076104 (2015) 9. N.C. Coronel, L. Lahourcade, K.T. Delaney, A.M. Shing, H.A.

Atwater, Proc. 38th IEEE PVSC, p. 003204 (2012) 10. T.D. Veal, N. Feldberg, N.F. Quackenbush, W.M. Linhart,

D.O. Scanlon, L.F.J. Piper, S.M. Durbin, Adv. Energy Mater. 5, 1501462 (2015)

11. N. Feldberg, J.D. Aldous, P.A. Stampe, R.J. Kennedy, T.D. Veal, S.M. Durbin, J. Electron. Mater. 43(4), 884 (2014)

12. N. Senabulya, N. Feldberg, R.A. Makin, Y. Yang, G. Shi, C.M. Jones, E. Kioupakis, J. Mathis, R. Clarke, S.M. Durbin, AIP Adv. 6, 075019 (2016)

13. C.H. Kuo, K.S. Chang, Cryst. Growth Des. 17(9), 4696–4702 (2017)

14. A.N. Fioretti, A. Stokes, M.R. Young, B. Groman, E.S. Toberer, A.C. Tamboli, A. Zakutayev, Adv. Electron. Mater. 3, 1600544 (2015)

15. R.L. Anderson, Solid State Electron. 5, 341 (1962) 16. A.G. Milnes, D.L. Feucht, Heterojunctions and Metal Semicon-

ductor Junctions (Academic Press, New York, 1972) 17. M.J. Adams, A. Nussbaum, Solid State Electron. 22, 783–791

(1979) 18. A. Kudo, Y. Miseki, Chem. Soc. Rev. 38, 253–278 (2009) 19. L. Yang, H. Zhou, T. Fan, D. Zhang, Phys. Chem. Chem. Phys.

16, 6810–6826 (2014) 20. M.G. Walter, E.L. Warren, J.R. McKone, S.W. Boettcher, Q.X.

Mi, E.A. Santori, N.S. Lewis, Chem. Rev. 110, 6446–6473 (2010) 21. R. Schlaf, H. Murata, Z.H. Kafafi, J. Electron Spectrosc. Relat.

Phenom. 120, 149–154 (2001) 22. M. Schulz, E. Klausmann, A. Hurrle, CRC Crit. Rev. Solid State

Sci. 5(3), 319–325 (1975) 23. C.T. Au, W. Hirsch, W. Hirschwald, Surf. Sci. 197(3), 391–401

(1988) 24. F. Streicher, S. Sadewasser, M.C. Lux-Steiner, Rev. Sci. Instrum.

80, 013907 (2009) 25. G.V. Hansson, R.I.G. Uhrberg, Surf. Sci. Rep. 9, 197–292 (1988) 26. P.T. Andrews, L.A. Hisscott, J. Phys. F Met. Phys. 5, 1568–1572

(1975)

27. J. Kubota, K. Domen, ECS Interface 24(2), 57–62 (2013) 28. T. Sahm, A. Gurlo, N. Barsan, U. Weimar, Sens. Actuators B 118,

78–83 (2006) 29. C.A. Dearden, M. Walker, N. Beaumont, I. Hancox, N.K. Uns-

worth, P. Sullivan, C.F. McConville, T.S. Jones, Phys. Chem. Chem. Phys. 16, 18926–18932 (2014)

30. A. Fuchs, H.J. Schimper, A. Klein, W. Jaegermann, Energy Proc. 10, 149–154 (2011)

31. K. Burger, F. Tschismarov, H. Ebel, J. Electron Spectrosc. Relat. Phenom. 10, 461 (1977)

32. Positions of photoelectron and auger lines on the binding energy scale (Al X-rays) (XPS International LLC, 2017). http://www.xpsdata.com/XI_BE_table.htm. Accessed 06 Dec 1999

33. C.D. Wagner, W.M. Riggs, L.E. Davis, J.F. Moulder, G.E. Muilen-berg, Line positions from Al X-rays in numerical order, in Hand-book of X-ray Photoelectron Spectroscopy, 1st edn, (Perkin-Elmer Corporation Physical Electronics, 1979), p. 187 (Table 4)

34. X. Bai, W. Jie, G. Zha, W. Zhang, P. Li, H. Hua, L. Fu, Appl. Surf. Sci. 255(18), 7966 (2009)

35. B. Conings, L. Baeten, C.D. Dobbelaere, J. D’Haen, J. Manca, H.G. Boyen, Adv. Mater. 26(13), 2041–2046 (2014)

36. F. Vaz, N. Martin, M. Fenker, Metallic Oxynitride Thin Films by Reactive Sputtering and Related Deposition Methods (Bentham Science Publishers, Sharjah, 2013)

37. H. Moormann, D. Kohl, G. Heiland, Surf. Sci. 80, 261–264 (1979) 38. K. Jacobi, G. Zwicker, A. Gutmann, Surf. Sci. 141(1), 109–125

(1984) 39. M. Batzill, U. Diebold, Prog. Surf. Sci. 79, 47–154 (2005) 40. V.I. Nefedov, I.A. Zakharova, I.I. Moiseev, M.A. Porai-koshits,

M.N. Vargoftik, A.P. Belov, Zh. Neorg. Khimil 18, 3264 (1973) 41. D.E. Eastman, J.K. Cashion, Phys. Rev. Lett. 24(7), 310–313

(1970) 42. C.M. Eggleston, J.J. Ehrhardt, W. Stumm, Am. Miner. 81, 1036–

1056 (1996) 43. J.L. Freeouf, D.E. Eastman, CRC Crit. Rev. Solid State Sci. 5(3),

245–258 (1975) 44. K. Nomura, T. Kamiya, E. Ikenaga, H. Yanagi, K. Kobayashi, H.

Hosono, J. Appl. Phys. 109, 073726 (2011) 45. T.E. Fischer, Surf. Sci. 13(1), 30–51 (1969) 46. L.F. Wagner, W.E. Spicer, Phys. Rev. Lett. 28(21), 1381–1384

(1972) 47. J.N. Miller, I. Lindau, W.E. Spicer, Philos. Mag. B 43(2), 273–282

(1981) 48. D.R. Palmer, S.R. Morrison, C.E. Dauenbaugh, Phys. Rev. 129(2),

608–613 (1963) 49. A.J. Nozik, R. Memming, J. Phys. Chem. 100(31), 13061–13078

(1996)