Skip to main content
  • Research article
  • Open access
  • Published:

Electron-transport properties of degenerate ZnSnN2 doped with oxygen

Abstract

In this study, analysis of the electron mobility in ZnSnN2 epilayers that were unintentionally doped with oxygen (ZnSnN2−xOx) was performed to elucidate the reason for the low mobilities of ~ 20 cm2 V−1 s−1. While roughly 30% of the incorporated oxygen donated electrons, the rest existed as neutral impurities. Seebeck-effect measurements revealed that scattering by neutral impurities governed the electron transport. The theoretical mobility calculated taking into account the scattering by neutral impurities and ionized impurities reproduced the experimental Hall mobility. We concluded that the low electron mobility is attributed to the presence of the neutral oxygen impurities in high concentration.

Introduction

Zn–IV–N2 (IV = Si, Ge, Sn) compounds, which are derived from wurtzite-type group III nitride by replacing the group III elements with an equal number of Zn and group IV elements, can be regarded as a pseudo group III nitride [1, 2]. It has been demonstrated that similar to the well-known InxGaxN system, the bandgap (Eg) of alloyed ZnSnxGe1−xN2 was tunable from 3.1 to 2.0 eV by varying the Sn content (x) from 0 to 1 [3]. This makes Zn–IV–N2 an intriguing semiconductor system, and studies on ZnSnxGe1−xN2 as a counterpart to the InxGa1−xN system have been carried out in recent years [3, 4].

Zinc tin nitride (ZnSnN2) is the least studied Zn–IV–N2 compound. For example, the synthesis of ZnSnN2 thin films [5, 6] and powders [7, 8] has been reported in only the last 5 years. Since theoretical studies showed that the Eg value of 1.0–2.0 eV for ZnSnN2 is appropriate for a photovoltaic absorber [1, 2, 9,10,11], experimental investigations have been carried out on this material [12,13,14,15]. Indeed, ZnSnN2 has a large optical absorption coefficient (~ 105 cm−1) in most of the solar spectrum [13]. In addition, p-Si/n-ZnSnN2 and p-SnO/n-ZnSnN2pn junctions for photovoltaic application were successfully fabricated recently [15, 16].

ZnSnN2 has two phases that depend on the ordering of the cation sublattice: the ordered phase, which is derived by alternately replacing the cation sublattice with Zn and Sn in the wurtzite structure; the disordered-phase, in which Zn and Sn randomly occupy the cation sublattice. Theoretical studies showed that Eg of ZnSnN2 is dependent on the cation-sublattice ordering: Eg = 2 eV for the ordered phase and Eg = 1 eV for the disordered phase [10, 17]. Optical studies on ordered- and disordered-ZnSnN2 thin films obtained Eg values close to the above-mentioned theoretical values [10, 12, 14, 17, 18]. Owing to intensive study from theoretical and experimental points of view, the intrinsic Eg value of ZnSnN2 has gradually become clear.

However, even though control of the transport properties is also important for photovoltaic applications, such properties of ZnSnN2 are not well understood. ZnSnN2 thin films usually show degenerate n-type conductivity owing to unintentional oxygen doping [12, 18]. The oxygen impurities occupy the nitrogen sites and behave as electron donors [11]. As a result, ZnSnN2 films have electron densities of the order of 1019–1021 cm−3 [9, 10, 12, 14, 17, 18]. The electron mobility of ZnSnN2 thin films are usually 10 cm2 V−1 s−1 or lower [9, 10, 12, 14, 17, 18] even though the films are single crystalline. To establish ZnSnN2 as a photovoltaic absorber, the major challenges include suppression of the conduction electron density and enhancement of the electron mobility.

For the suppression of the conduction electron density, Floretti et al. proposed a method in which hydrogen doping takes place during film growth and subsequent annealing [19]. They successfully obtained nondegenerate Zn1+δSn1−δN2 films with a carrier density of 1016 cm−3. On the other hand, no method for improving the electron mobility has been reported so far, and the low electron mobility remains one of the open questions of the electron-transport properties of ZnSnN2. In the case of heavily-doped GaN epilayers with an electron density of ~ 1020 cm−3 [20], an electron mobility above 100 cm2 V−1 s−1 has been obtained. In comparison, the electron mobility of ZnSnN2 appears to be unusually low, but an understanding of the limiting factor of the electron mobility in ZnSnN2 is lacking.

In this study, detailed analysis of the electron-transport properties of epitaxially grown ZnSnN2 thin films was carried out to elucidate the reason for the low electron mobility. Herein, we present the analysis results and discuss in detail the limiting factor of the mobility.

Results and discussion

Cation-disordered ZnSnN2 films were grown on the (111) plane of yttria-stabilized zirconia (YSZ) single-crystalline substrates with a temperature of 300 °C; the films were grown by reactive radio-frequency magnetron sputtering using a Zn–Sn alloy target with Zn concentration of 50 at.%. In general, high growth temperature is preferable for the growth of thin films composed of larger crystalline grains with higher crystallinity. In the case of ZnSnN2, however, the highest growth temperature is limited to ~ 300 °C, because the decomposition temperature of this compound was estimated to be ~ 350 °C [8]. Indeed, we previously confirmed that no film was grown on YSZ (111) at 350 °C [18]. Accordingly, 300 °C is the highest temperature for the growth of ZnSnN2 thin films.

Out-of-plane and in-plane X-ray diffraction (XRD) measurements confirmed that the ZnSnN2 films were epitaxially grown on YSZ(111) with the epitaxial relationship of \(\left( {000{1}} \right)[{{2\bar{1}\bar{1}}}0]_{{{\text{ZnSnN}}_{{2}} }}\) || (111)[\({\bar{{1}}}\)10]YSZ (Additional file 1: Fig. S1). The Zn/(Zn + Sn) atomic ratio in the films was determined by X-ray photoelectron spectrometry (XPS) to be in the range of 0.52–0.55. That is, slightly Zn-rich off-stoichiometric films were obtained in this study. The XPS measurements also confirmed that all the films were unintentionally doped with oxygen, forming ZnSnN2−xOx. The oxygen content x could be controlled by varying the nitrogen partial pressure (\(P_{{{\text{N}}_{{2}} }}\)) during the growth [18]. When \(P_{{{\text{N}}_{{2}} }}\) was decreased from 2.0 to 1.2 Pa, x in ZnSnN2−xOx increased from 0.056 to 0.079 (the x values were semi-quantitatively determined by using the integrated intensity of the O 1s core spectra). The XPS spectra and elemental composition data are given in Additional files 2 and 3, respectively. Impurity substitution usually expand or shrink a unit cell, leading to a small shift of XRD peak. In the case of the ZnSnN2−xOx films with x < 0.08, however, significant peak shift by the incorporation of oxygen was not observed (Additional file 4: Fig. S3a), indicating that the lattice constant did not change (Additional file 4: Fig. S3b). The difference of the ionic radii of N3− (146 pm) and O2− (138 pm) is so small that the oxygen substitution hardly caused the change of the lattice constant.

Hall-effect measurements were carried out to determine the carrier density (ne) and Hall mobility (μH) in the films. As shown in Fig. 1a, ne increased with increasing x, indicating that the unintentionally incorporated oxygen (ON) served as an electron donor. Temperature (T)-independent behavior of ne can be seen in Fig. 1b, which suggests that the ZnSnN2−xOx films were degenerate semiconductors. The curve in Fig. 1a represents ne calculated assuming that 30% of ON was ionized (the ionization rate of η = 30%). The curve reasonably approximated the experimental ne values, indicating that ON was highly compensated at room temperature (RT). The reason for such high compensation is probably related to the Zn-rich composition of the films. The Zn excess introduces ZnSn acceptor-like defects in ZnSnN2 [19]. Recent theoretical calculation suggests that electrically neutral ZnSn–2ON complexes forms when ZnSn and ON coexist in ZnSnN2 [21]. The formation of the ZnSn–2ON complexes implies the compensation of the ON donors. Hence, many ZnSn–2ON complexes are likely to present in the epilayers.

Fig. 1
figure 1

a Electron density (ne) and c Hall mobility (μH) at room temperature as functions of x in ZnSnN2−xOx (closed circles). The solid line in a represents the carrier density calculated based on the assumption that the ionization rate of the incorporated oxygen was 30%. The cross marks in c represent the optically derived mobility (μopt). Temperature dependence of bne and dμH; the circles, triangles, inverted triangles, squares, and diamonds represent the values for films with x = 0.055, 0.069, 0.071 0.074, and 0.079, respectively. All the data are included in Additional file 5

Figure 1c shows the x dependence of μH at RT. The unintentional oxygen doping caused the decrease in μH, suggesting that ON acted as an impurity scattering center. Indeed, the μH values were independent of T (Fig. 1c), representing typical behavior of mobility limited by ionized- or neutral-impurity scattering in degenerate semiconductors. Furthermore, the T-independent behavior of μH clearly indicates that the contribution of T-dependent phonon scattering to electron transport was negligible.

Recently, Hamilton et al. reported that grain-boundary scattering is one of mobility-limiting factors in polycrystalline films in ZnSnN2−xOx [19]. Even in epitaxial films, grain-boundary scattering sometimes cannot be ignored, because epitaxial films frequently have a biaxially-oriented grain structure [22,23,24]. As we previously reported, ZnSnN2 epitaxial films sputtered on YSZ (111) have a compact grain structure with lateral grain diameters of ~ 30 nm [18]. In other words, sputtered ZnSnN2−xOx epilayers are not single-crystalline films but rather biaxially-oriented polycrystalline films. Accordingly, grain-boundary scattering should be considered, when we analyze the electron transport properties of the sputtered ZnSnN2−xOx epilayers. The optically derived mobility (μopt) using the Drude model corresponded to the intra-grain mobility without the grain boundary contribution, whereas DC-measured μH included the contribution of the grain-boundary scattering. To obtain the μopt values, we measured the infrared reflectance spectra of the ZnSnN2−xOx films at near-normal incidence (~ 5°) in the wavelength range of 1–5 μm, and then fitting analysis was carried out using the Drude model. In our previous study, we showed that the frequency (ω)-dependent dielectric function of the ZnSnN2−xOx (ε(ω)) can be modeled by combining the Drude function (εD(ω)) with double Tauc–Lorentz (TL) functions [εTL1(ω) and εTL2(ω)]: i.e., ε(ω) = εTL1(ω) + εTL2(ω) + εD(ω). The Drude function is given by

$$\varepsilon_{{\text{D}}} \left( \omega \right) = - \frac{{\omega_{{\text{p}}} }}{{\omega^{2} - i\varGamma_{{\text{D}}} \omega }},$$
(1)
$$\omega_{{\text{p}}}^{2} = \frac{{en_{{\text{e}}} }}{{\varepsilon_{0} m^{*} }} ,$$
(2)

where ωp is the plasma frequency, ΓD is the scattering rate, ε0 is the static dielectric constants of free space, and m* is the effective mass. The TL function is given in Ref. [25]. The theoretical reflectance spectra calculated through the Fresnel formula combined with the modeled dielectric function were fitted to the experimental spectra (ωp and ΓD were used as the fitting parameters). In the fitting analysis, the parameters in the TL functions were fixed for simplicity at constant values that were the same as reported values in the literature [13]. We obtained good fits, as shown in Fig. 2a–d. The best-fit parameters are listed in Additional file 7: Table S2 (Additional file 7). Using the best-fit parameters, the μopt values were calculated using the relationship

$$\mu_{{{\text{opt}}}} = \frac{1}{{en_{{\text{e}}} }}\frac{{\varepsilon_{0} \omega_{{\text{p}}}^{2} }}{{\varGamma_{{\text{D}}} }}{ }.$$
(3)

The values of μopt and μH were almost identical, as shown in Fig. 1c, indicating that the contribution of grain-boundary scattering to the electron transport in the ZnSnN2−xOx epilayers was negligible. In other words, μH corresponded to the intra-grain mobility, implying that the electron transport properties of the ZnSnN2−xOx epilayers were close to those of single crystalline films.

Fig. 2
figure 2

Infrared reflectance spectra of ZnSnN2−xOx epilayers with x = a 0.055, b 0.069, c 0.074, and d 0.079. The results of least-squares fitting using the modeled dielectric function are represented by solid lines. All the data are included in Additional file 6

Here, we compare μH values of the ZnSnN2−xOx epilayers with those of InN epilayers. Figure 3 shows ne versus μH plot for the ZnSnN2−xOx and InN epilayers [26, 27]. The μH values of the present ZnSnN2−xOx films are in comparable level with those of the degenerate InN epilayers reported in the 1990s. Nevertheless, the μH values in the present study are substantially lower than those of the InN layers fabricated in 2010s. We believe that there is still room for the improvement of electron mobility, if the mobility-limiting factors are fully understood.

Fig. 3
figure 3

Comparison of Hall mobility (μH) as a function of electron density (ne) for ZnSnN2−xOx and degenerate InN epilayers [26, 27]. All the data are included in Additional file 8

To obtain insight into the intra-grain electron transport in ZnSnN2−xOx, the μH values were compared with theoretical models. We first considered ionized-impurity scattering, which generally governs the electron transport in heavily-doped degenerate semiconductors. The theoretical mobilities limited by the ionized-impurity scattering (μI) were calculated by using the Brooks–Herring model [28]. We used an effective mass of 0.37m0 [18] and a static dielectric constant (εs) of 11[1] for the calculation. The calculated μI is plotted as a function of ne in Fig. 4a. The μI curve gives the upper limit of the electron mobility in ZnSnN2−xOx; namely, the electron mobility of ZnSnN2−xOx potentially reaches ~ 100 cm2 V−1 s−1. Nevertheless, the experimental μH values had a maximum of 20 cm2 V−1 s−1, implying that an additional scattering source must be taken into account.

Fig. 4
figure 4

a Dependence of Hall mobility on carried density. The circles represent data for the same films shown in Figs. 1 and 2. The triangles represent data for films grown under conditions similar to those employed for the grown of films shown in Figs. 1 and 2. The lines labeled μI and μN are theoretical curves of mobility limited by ionized- and neutral-impurity scatterings, respectively. The curve labeled μtotal represents the total mobility calculated from μtotal = (1/μI + 1/μN)−1. b Seebeck coefficient as a function of (carrier density)−2/3; theoretical lines for r = 3/2 (dashed line) and 0 (solid line) are also included. All the data are included in Additional file 9

Thermopower measurements were carried out for further investigation. For highly degenerate n-type semiconductors with a parabolic conduction band, Seebeck coefficient (S) is given by [28]

$$S = - \frac{{8\pi^{2} }}{3}\left( {r + \frac{3}{2}} \right)\frac{{k_{B}^{2} m^{*} T}}{{eh^{2} }}n_{e}^{{ - \frac{2}{3}}} ,$$
(4)

where r represents the scattering constant (e.g., r = 3/2 for ionized-impurity scattering, and r = 0 for neutral-impurity scattering), kB is the Boltzmann constant, and h is the Planck constant. A plot of S versus ne−2/3 allows us to gain insight into the electron-transport mechanisms from the viewpoint of r values [29]. As shown in Fig. 4b, the experimental data lie close to the theoretical curve for r = 0 (solid line), whereas the experimental S values are located away from the theoretical S curve (dashed line) for the ionized-impurity scattering (r = 3/2). These trends suggest that neutral-impurity scattering was the dominant scattering mechanism. The temperature-independent behavior of μH is consistent with this interpretation because neutral-impurity scattering does not depend on temperature [30, 31]. Since ~ 70% of ON was not ionized (Fig. 1a), it is reasonable to suppose that the compensated oxygen that is likely to form the ZnSn–2ON complex acted as the neutral-impurity scattering centers. In fact, Hamilton et al. recently pointed out the possibility that the ZnSn–2ON complexes behave as the neutral-impurity scattering centers [21].

We further analyzed the neμH data by taking account of the neutral-impurity scattering. We hypothesized that all the compensated oxygen served as neutral-impurity scattering centers, namely, the density of the neutral impurity (nN) was defined as nN = (ne/η)(1 − η). The mobility limited by the neutral-impurity scattering (μN) was calculated on the basis of the model developed by Mayer et al. [31]:

$$\mu_{{\text{N}}} = \frac{{e^{3} m^{*} }}{{4\pi \varepsilon_{{\text{s}}} \varepsilon_{0} A\left( w \right)\hbar^{3} }}\frac{1}{{n_{{\text{N}}} }}{ },$$
(5)

where w = (kFaB)2,

$$A\left( w \right) = \frac{35.2}{{w^{\frac{1}{2}} }}\frac{{\left( {1 + e^{ - 50w} } \right)\left( {1 + 80.6w + 23.7w^{2} } \right)}}{{1 + 41.3w + 133w^{2} }}\left[ {\frac{1}{w}{\ln}\left( {1 + w} \right) - \frac{{1 + \frac{1}{2}w - \frac{1}{6}w^{2} }}{{\left( {1 + w} \right)^{3} }}} \right],$$
(6)

kF is the Fermi wave number, and aB is the effective Bohr radius (aB = 1.5 nm for ZnSnN2). Assuming that the ionization ratio of ON was η = 30%, the total mobility was calculated using Matthiessen’s rule, μTotal = (1/μN + 1/μI)−1. The calculated neμtotal curve agrees well with the experimental neμH data, as shown in Fig. 4a. Furthermore, we grew an additional series of ZnSnN2−xOx epilayers under conditions similar to those mentioned above, and their neμH data are also plotted in Fig. 4a (triangles). Again, the experimental data was found to be located near the theoretical neμtotal curve. These findings led us to the conclusion that the unintentionally incorporated oxygen had an ionization rate as low as approximately ~ 30% and the unionized oxygen acted as neutral-impurity scattering centers that dominated electron transport in the heavily-doped ZnSnN2−xOx epitaxial films. The low μH values in the ZnSnN2−xOx epilayers resulted from the high concentration of neutral oxygen impurities. Frequently reported low electron mobilities in ZnSnN2 epitaxial films can be explained by the same reason. Recent study showed that μH of ZnSnN2−xOx films with very high oxygen contents (x > 0.4) was as low as < 5 cm2 V−1 s−1 [21]. That is, the more the oxygen content increases, the lower μH becomes. This supports the idea that electron scattering by the oxygen-related defects governs the electron-transport properties of ZnSnN2−xOx.

Conclusions

In summary, we analyzed the electron-transport properties of unintentionally oxygen-doped ZnSnN2 epitaxial layers. We confirmed that the incorporated oxygen impurities behaved as electron donors with the low ionization rate of 30%. The Hall- and Seebeck-effect measurements revealed that the compensated oxygen impurities, which are likely to form the electrically neutral ZnSn–2ON complexes, behaved as neutral-impurity scattering centers and further governed electron transport in the ZnSnN2−xOx. The low ionization rate led to the high concentration of neutral-impurity scattering centers. Therefore, we conclude that the low electron mobilities reported even in ZnSnN2 single-crystalline films are attributed to the high concentration of neutral oxygen impurities. Suppression of the oxygen concentration in ZnSnN2 is crucial not only to obtain nondegenerate ZnSnN2, but also to achieve high mobility.

Methods

Thin-film growth

Reactive radio-frequency (RF) magnetron sputtering was employed to grow ZnSnN2 epitaxial films on YSZ(111) single-crystalline substrates at the growth temperature of 300 °C. The base pressure of ~ 2 × 10–4 Pa was established prior to the film-growth. A disk of Zn0.5Sn0.5 alloy was used as a target (diameter of 10 cm and purity of 3 N). An RF power of 70 W was applied to the target. A mixture of Ar and N2 gas with various N2/(N2 + Ar) ≡ f(N2) ratios was introduced into the chamber through two independent mass flow controllers with a total flow rate of 5 sccm. The total pressure in the growth chamber (P) was held at 2.0 Pa during film growth. The nitrogen partial pressure was defined as PN2 = f(N2) × P. The growth time was adjusted to obtain films with thicknesses of 100–300 nm. Prior to the film growth, a 5 min-long sputter-etching with pure Ar gas was performed, followed by a 5 min-long pre-sputtering under the same condition with the film growth. After the film growth, the as-grown films were immediately stored in another vacuum chamber (the pressure was about 1 Pa) for subsequent measurements.

Characterization

A Rigaku ATX-G X-ray diffractometer with Cu Kα radiation was employed to perform out-of-plane (θ/ω) and in-plane (2θχ) scans to confirm the epitaxial growth. The compositions (Zn/(Zn + Sn) and x) of the ZnSnN2 films were examined by XPS (PHI Versa Probe), using monochromated Al Kα ( = 1486.6 eV) radiation. The XPS measurements were performed on 3-min Ar+-sputter-etched surface of the films. The relative sensitivity factor (RSF) approach was exploited to determine the compositions. It was confirmed that the compositions determined by the RSF method were consistent with those determined by Rutherford backscattering spectrometry [29, 32]. Hence, we believe that the compositions in this study are sufficiently reliable. The details of the RSF approach were already described elsewhere [18]. Electrical properties were determined by Hall-effect measurements in the van der Pauw configuration (Toyo Corp. Resitest 8200). Optical transmittance and reflectance were collected between 0.3 and 5.0 µm using a UV–Vis-NIR spectrophotometer (Shimadzu UV-3150) and FTIR spectrometer (Shimadzu IRAffinity-1).

Fitting analysis of reflectance spectra

As seen from Fig. 2a–d, free-electron reflection is clearly seen in the IR region. The well-known Drude dielectric model (Eqs. 1, 2) was employed to describe the optical response by the free electrons. In addition to the Drude model, the Tauc–Lorentz (TL) dispersion model was considered to describe the optical response across the whole spectral region. The explicit expression of the TL model is given in literature [25]. Deng et al. previously demonstrated that the dielectric response of ZnSnN2 in the UV to visible region can be reproduced by double TL functions [13]. Hence, we modelled the dielectric function of degenerate ZnSnN2−xOx as a sum of double TL and Drude functions. Theoretical reflectance spectra calculated via the Fresnel formulas combined with the dielectric function above were fitted to the experimental spectra. In the fitting analysis, the parameters in the TL functions were fixed for simplicity at constant values that were the same as the values reported by Deng et al. [13]. Thus, ωp and ΓD in the Drude function were used as the fitting parameters.

Availability of data and materials

All data generated or analyzed during this study are included in Additional files 19.

Abbreviations

μ H :

Hall mobility

n e :

electron density

ON :

impurity oxygen on nitrogen site in ZnSnN2

x :

oxygen content in ZnSnN2xOx

XPS:

X-ray photoelectron spectroscopy

XRD:

X-ray diffraction

ZnSn :

zinc atom on tin site in ZnSnN2

References

  1. Paudel TR, Lambrecht WRL. First-principles study of phonons and related ground-state properties and spectra in Zn–IV–N2 compounds. Phys Rev B. 2008;78:115204.

    Article  Google Scholar 

  2. Punya A, Lambrecht WRL, Schilfgaarde M. Quasiparticle band structure of Zn–IV–N2 compounds. Phys Rev B. 2011;84:165204.

    Article  Google Scholar 

  3. Narang P, Chen S, Coronel NC, Gul S, Yano J, Wang LW, Lewis NS, Atwater HA. Bandgap tunability in Zn(Sn, Ge)N2 semiconductor alloys. Adv Mater. 2014;26:1235–41.

    Article  CAS  Google Scholar 

  4. Shing AM, Coronel NC, Lewis NS, Atwater HA. Semiconducting ZnSnxGe1xN2 alloys prepared by reactive radio-frequency sputtering. APL Mater. 2015;3:76104.

    Article  Google Scholar 

  5. Coronel NC, Lahourcade L, Delaney KT, Shing AM, Atwater HA. Earth-abundant ZnSnxGe1−xN2 alloys as potential photovoltaic absorber materials. In: Proceedings of the 38th IEEE photovoltaic specialists conference. New York: IEEE; 2012. p. 3204–7.

  6. Feldberg N, Keen B, Aldous JD, Scanlon DO, Stampe PA, Kennedy RJ, Reeves RJ, Veal TD, Durbin SM. ZnSnN2: a new earth-abundant element semiconductor for solar cells. In: Proceedings of the 38th IEEE photovoltaic specialists conference. New York: IEEE; 2012. p. 2524–7

  7. Quayle PC, He K, Shan J, Kash K. Synthesis, lattice structure, and band gap of ZnSnN2. MRS Commun. 2013;3:135–8.

    Article  CAS  Google Scholar 

  8. Kawamura F, Yamada N, Imai M, Taniguchi T. Synthesis of ZnSnN2 crystals via a high-pressure metathesis reaction. Cryst Res Technol. 2016;51:220–4.

    Article  CAS  Google Scholar 

  9. Lahourcade L, Coronel NC, Delaney KT, Shukla SK, Spaldin NA, Atwater HA. Structural and optoelectronic characterization of RF sputtered ZnSnN2. Adv Mater. 2013;25:2562–6.

    Article  CAS  Google Scholar 

  10. Feldberg N, Aldous JD, Linhart WM, Phillips LJ, Durose K, Stampe PA, Kennedy RJ, Scanlon DO, Vardar G, Field RL, Jen TY, Goldman RS, Veal TD, Durbin SM. Growth, disorder, and physical properties of ZnSnN2. Appl Phys Lett. 2013;103:42109.

    Article  Google Scholar 

  11. Chen S, Narang P, Atwater HA, Wang LW. Phase stability and defect physics of a ternary ZnSnN2 semiconductor: first principles insights. Adv Mater. 2014;26:311–5.

    Article  CAS  Google Scholar 

  12. Fioretti AN, Zakutayev A, Moutinho H, Melamed C, Perkins JD, Norman AG, Al-Jassim M, Toberer ES, Tamboli AC. Combinatorial insights into doping control and transport properties of zinc tin nitride. J Mater Chem C. 2015;3:11017.

    Article  CAS  Google Scholar 

  13. Deng F, Cao H, Liang L, Li J, Gao J, Zhang H, Qin R, Liu C. Determination of the basic optical parameters of ZnSnN2. Opt Lett. 2015;40:1282–5.

    Article  CAS  Google Scholar 

  14. Martinez AD, Fioretti AN, Toberer E, Tamboli AC. Synthesis, structure, and optoelectronic properties of II–IV–V2 materials. J Mater Chem A. 2017;5:11418.

    Article  CAS  Google Scholar 

  15. Qin R, Cao H, Liang L, Xie Y, Zhuge F, Zhang H, Gao J, Javaid K, Liu C, Sun W. Semiconducting ZnSnN2 thin films for Si/ZnSnN2pn junctions. Appl Phys Lett. 2016;108:142104.

    Article  Google Scholar 

  16. Javaid K, Yu J, Wu W, Wang J, Zhang H, Gao J, Zhuge F, Liang L, Cao H. Thin film solar cell based on ZnSnN2/SnO heterojunction. Phys Status Solidi PRL. 2018;12:1700332.

    Google Scholar 

  17. Veal TD, Feldberg N, Quackenbush NF, Linhart WM, Scanlon DO, Piper LFJ, Durbin SM. Band gap dependence on cation disorder in ZnSnN2 solar absorber. Adv Energy Mater. 2015;5:1501462.

    Article  Google Scholar 

  18. Cao X, Kawamura F, Ninomiya Y, Taniguchi T, Yamada N. Conduction-band effective mass and bandgap of ZnSnN2 earth-abundant solar absorber. Sci Rep. 2017;7:14987.

    Article  Google Scholar 

  19. Fioretti AN, Stokes A, Young MR, Gorman B, Toberer ES, Tamboli AC, Zakutayev A. Effects of hydrogen on acceptor activation in ternary nitride semiconductors. Adv Electron Mater. 2017;3:1600544.

    Article  Google Scholar 

  20. Gaskill DK, Rowland LB, Doverspike K. Electrical transport properties of AlN, GaN and AlGaN. In: Edger JH, editor. Properties of group III nitrides, EMIS Data Reviews Series No. 11. London: Inspec; 1994. p. 105.

  21. Hamilton DC, Arca E, Pan J, Siol S, Young M, Lany S, Zakutayev A. Electron scattering mechanisms in polycrystalline sputtered zinc tin oxynitride thin films. J Appl Phys. 2019;126:035701.

    Article  Google Scholar 

  22. Yan M, Lane M, Kannewurf CR, Chang PH. Highly conductive epitaxial CdO thin films prepared by pulsed laser deposition. Appl Phys Lett. 2001;78:2342–4.

    Article  CAS  Google Scholar 

  23. Furubayashi Y, Yamada N, Hirose Y, Yamamoto Y, Otani M, Hitosugi T, Shimada T, Hasegawa T. Transport properties of d-electron based transparent conducting oxide: anatase Ti1xNbxO2. J Appl Phys. 2007;101:093705.

    Article  Google Scholar 

  24. Zhu QY, Ye ZZ, Yuan GD, Huang JY, Zhu LP, Zhao BH, Lu JG. Synthesis and characterization of Al–N codoped p-type ZnO epitaxial films using high-temperature homo-buffer layer. Appl Surf Sci. 2006;253:1903–6.

    Article  CAS  Google Scholar 

  25. Jellison GE Jr, Modine FA. Parameterization of the optical functions of amorphous materials in the interband region. Appl Phys Lett. 1996;69:371–3.

    Article  CAS  Google Scholar 

  26. Yamamoto A, Shin-ya T, Sugiura T, Hashimoto A. Characterization of MOVPE-grown InN layers on α-Al2O3 and GaAs substrates. J Cryst Growth. 1998;189–190:461–5.

    Article  Google Scholar 

  27. Rauch C, Tumisto F, King PDC, Veal TD, Lu H, Schaff WJ. Self-compensation in highly n-type InN. Appl Phys Lett. 2012;101:011903.

    Article  Google Scholar 

  28. Seeger K. Semiconductor physics; an introduction. New York: Springer; 2011.

    Google Scholar 

  29. Cao X, Yamaguchi Y, Ninomiya Y, Yamada N. Comparative study of electron transport mechanisms in epitaxial and polycrystalline zinc nitride films. J Appl Phys. 2016;119:025104.

    Article  Google Scholar 

  30. Erginsoy C. Neutral impurity scattering in semiconductors. Phys Rev. 1950;79:1013–4.

    Article  CAS  Google Scholar 

  31. Meyer JR, Bartoli FJ. Phase-shift calculation of electron mobility in n-type silicon at low temperatures. Phys Rev B. 1981;24:2089–100.

    Article  CAS  Google Scholar 

  32. Núñez CG, Pau JL, Hernández MJ, Cervera M, Ruiz E, Piqueras J. On the zinc nitride properties and the unintentional incorporation of oxygen. Thin Solid Films. 2012;520:1924–9.

    Article  Google Scholar 

Download references

Acknowledgements

The authors acknowledge M. Kawamura (Institute of Science and Technology Research, Chubu University) for providing useful advice on XRD and XPS measurements.

Funding

This work was supported by Japan Society for the Promotion of Science (JSPS) KAKENHI (Grant No. 16H04500). The funder had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

NY and FK designed the experiments. XC performed fabrication and measurements. NY carried out the theoretical analysis. TT, YN and NY were the PIs. XC and NY wrote the manuscript. All authors discussed the results and reviewed the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Naoomi Yamada.

Ethics declarations

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Additional files

Additional file 1.

Typical out-of-plane and in-plane XRD patterns for ZnSnN2 epitaxial films grown on YSZ (111).

Additional file 2.

Oxygen 1s core spectra of ZnSnN2 unintentionally doped with oxygen.

Additional file 3.

Zn/(Zn + Sn) atomic ratio and x in ZnSnN2−xOx.

Additional file 4.

Out-of-plane XRD patterns for ZnSnN2−xOx films with different x values and x-dependence of c-axis length.

Additional file 5.

Electron density and mobility as functions of temperature and oxygen impurity concentration, x.

Additional file 6.

Experimental reflectance and the best-fit curves using the TL and Drude combination model.

Additional file 7.

Electron density and Drude parameters (plasma frequency, ωp and broadening factor, ΓD in the Drude function) extracted from the best-fit spectra shown in Fig. 2a−e in the main text.

Additional file 8.

ne vs. μH data for ZnSnN2−xOx and InN epilayers. The data for InN were taken from Refs. [26, 27].

Additional file 9.

Theoretical mobility (ne vs. μ) and ne vs. S data shown in Fig. 4.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cao, X., Kawamura, F., Taniguchi, T. et al. Electron-transport properties of degenerate ZnSnN2 doped with oxygen. BMC Mat 2, 4 (2020). https://doi.org/10.1186/s42833-020-0010-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s42833-020-0010-6

Keywords