Effective Surface Passivation of InP Nanowires by Atomic-Layer

Sep 8, 2017 - The passivation scheme displays excellent long-term stability (>7 months) and is additionally shown to substantially improve the thermal...
0 downloads 17 Views 6MB Size
Subscriber access provided by GRIFFITH UNIVERSITY

Communication

Effective surface passivation of InP nanowires by atomic-layer-deposited AlO with PO interlayer 2

3

x

Lachlan E. Black, Alessandro Cavalli, Marcel Verheijen, Jos E.M. Haverkort, Erik P.A.M. Bakkers, and Wilhelmus (Erwin) M.M Kessels Nano Lett., Just Accepted Manuscript • DOI: 10.1021/acs.nanolett.7b02972 • Publication Date (Web): 08 Sep 2017 Downloaded from http://pubs.acs.org on September 10, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Nano Letters is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

Effective surface passivation of InP nanowires by atomic-layer-deposited Al2O3 with POx interlayer L. E. Black,1* A. Cavalli,1 M. A. Verheijen,1,2 J. E. M. Haverkort,1 E. P. A. M. Bakkers,1 and W. M. M. Kessels1* 1

Eindhoven University of Technology, Postbus 513, 5600 MB Eindhoven, The Netherlands

2

Philips Innovation Labs, High Tech Campus 11, 5656 AE Eindhoven, The Netherlands

KEYWORDS: Indium phosphide, semiconductor nanowires, surface passivation, atomic layer deposition, photoluminescence

ABSTRACT: III/V semiconductor nanostructures have significant potential in device applications, but effective surface passivation is critical due to their large surface-to-volume ratio. For InP such passivation has proven particularly difficult, with substantial depassivation generally observed following dielectric deposition on InP surfaces. We present a novel approach based on passivation with a phosphorus-rich interfacial oxide deposited using a low-temperature process, which is critical to avoid P-desorption. For this purpose we have chosen a POx layer deposited in a plasma-assisted atomic layer deposition (ALD) system at room temperature. Since POx is known to be hygroscopic and therefore

ACS Paragon Plus Environment

1

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 33

unstable in atmosphere, we encapsulate this layer with a thin ALD Al2O3 capping layer to form a POx/Al2O3 stack. This passivation scheme is capable of improving the photoluminescence (PL) efficiency of our state-of-the-art wurtzite (WZ) InP nanowires by a factor of ~20 at low excitation. If we apply the rate equation analysis advocated by some authors, we derive a PL internal quantum efficiency (IQE) of 75% for our passivated wires at high excitation. Our results indicate that it is more reliable to calculate the IQE as the ratio of the integrated PL intensity at room temperature to that at 10 K. By this means we derive an IQE of 27% for the passivated wires at high excitation (> 10 kW cm−2), which constitutes an unprecedented level of performance for undoped InP nanowires. This conclusion is supported by time-resolved PL decay lifetimes, which are also shown to be significantly higher than previously reported for similar wires. The passivation scheme displays excellent long-term stability (> 7 months), and is additionally shown to substantially improve the thermal stability of InP surfaces (> 300°C), significantly expanding the temperature window for device processing. Such effective surface passivation is a key enabling technology for InP nanowire devices such as nanolasers and solar cells.

TEXT III/V semiconductor nanostructures are the subject of increasing interest for device applications as researchers seek to overcome the intrinsic material limitations of Si while continuing device scaling and taking advantage of novel properties available at the nanoscale.1-4 However, the high surface-to-volume ratio of such nanostructures makes effective surface passivation critical for many device applications, and passivation of III/V surfaces has historically proven far from trivial.5-7 In this work, we aim to address this problem for the case of InP, a technologically important III/V semiconductor with

ACS Paragon Plus Environment

2

Page 3 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

applications in a variety of electronic and optoelectronic devices. InP is notable for its high electron mobility and velocity, ease of integration with important ternary and quaternary III/V compounds (including In0.53Ga0.47As and InxGa1−xAsyP1−y), and direct bandgap of 1.34 eV which is ideal for photovoltaic applications. InP/dielectric gate stacks have been employed in InGaAs buried-channel field-effect transistors which are the first to outperform Si-based devices.1,

8

InP nanowires can be

grown with good bulk material quality in both zinc-blende (ZB) and wurtzite (WZ) crystal phases, and InP nanowire solar cells have achieved some of the highest conversion efficiencies among nanowire photovoltaic devices,9,

10

while InP-based nanolasers hold promise as optical sources for silicon-

integrated photonics.11 Despite the widespread use of InP in devices, there exists no established method to stably and effectively passivate InP surfaces. It was early recognised12 that (in contrast to GaAs) the native InP surface exhibits a relatively low surface recombination velocity (on the order of ~102–104 cm/s depending on dopant concentration13-17), and this is frequently cited as an advantage for device applications. However, this level of passivation is still limiting for many devices, and the native surface displays poor thermal stability.18-20 More critically, this native surface passivation is generally not maintained following the deposition of standard dielectric layers (e.g. SiO2 or Al2O3) on the InP surface,18, 19 as required for the fabrication of many types of devices. More recent attempts to passivate InP nanowire surfaces have encountered similar difficulties. Wet chemical treatments have been shown to provide some degree of surface passivation,21, 22 but this is generally not stable. Münch et al.23 found that atomic layer deposition (ALD) of HfO2 significantly degraded surface passivation regardless of surface pre-treatment or deposition conditions. Dhaka et al.24 reported similar surface degradation following ALD of Al2O3, AlN, TiN, GaN, and TiO2 on InP nanowires and pillars. Some improvement was observed for very thin Al2O3 films below a critical thickness of 2–3 nm, though carrier lifetimes

ACS Paragon Plus Environment

3

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 33

remained well below 1 ns. Zhong et al.25 recently reported an increase in photoluminescence (PL) intensity and lifetime following plasma-enhanced chemical vapor deposition (PECVD) of SiNx on InP nanowires, however they also observed a significant decrease in the open-circuit voltage of nanowire solar cells with the same films, making the significance of this result somewhat ambiguous. In sum, despite the acknowledged importance of surface recombination in InP nanowire devices, there remains no clearly established, stable, and effective surface passivation method for this material. In attempting to devise an effective surface passivation scheme for InP nanowires, it is instructive to look back at historical work on the development of InP metal–insulator–semiconductor (MIS) devices. As noted above, early work showed that the deposition of standard dielectrics on InP tended to result in interfaces with poor electrical properties. It was conjectured that this might be linked to thermallyinduced degradation of the InP surface observed by various authors at temperatures ranging from 150°C to 225°C,18-20 and thought to be linked to the desorption of surface P known to occur at least above 200°C.26 Following this line of thought, or inspired by alternative structural considerations,27 a number of groups investigated the use of various P-rich layers, including deposition of P,28 POx,29 AlPxOy,27 InPxOy,30 PNx,31 or POxNy films (usually by PECVD),32 growth of P-rich anodic oxides,33,

34

or

deposition of other dielectrics in P overpressure,35, 36 to improve the interface characteristics of InP MIS structures. Such approaches appear to have been relatively successful, in most cases leading to significant improvements in MIS electrical characteristics. Since few of these groups employed PL characterization, it is not clear whether any of these approaches resulted in passivation improvements, or at least avoided degradation, relative to the native InP surface (those few who did perform such a comparison observed relative degradation31), but such approaches nevertheless show significant promise as a path towards effective surface passivation of InP.

ACS Paragon Plus Environment

4

Page 5 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

In this work, we adopt a novel approach inspired by this early work on InP MIS technology to develop an effective dielectric passivation scheme for InP nanowires. Specifically, we investigate the use of a deposited phosphorus oxide (POx) film to passivate the InP surface. Since phosphorus oxides are known to be hygroscopic and consequently unstable under ambient conditions (indeed uncapped POx films were observed to visibly degrade upon exposure to atmosphere), we encapsulate this layer with a thin ALD Al2O3 capping layer to form a POx/Al2O3 stack. We deposit both layers in an ALD reactor at room temperature (25°C) using a plasma-assisted process in order to avoid thermal degradation of the InP surface. Using this novel surface passivation scheme we are able to increase the PL internal quantum efficiency and PL lifetime to levels representing unprecedented performance for undoped InP nanowires. The investigated passivation scheme displays excellent long-term stability, and is additionally shown to improve the thermal stability of InP surfaces, significantly expanding the temperature window for InP device processing. The InP nanowires studied in this work were grown by catalyst-free selective-area vapor-phase epitaxy on Zn-doped InP (111)A substrates using a SiNx mask patterned by nanoimprint lithography.37 Growth was performed at 750°C from trimethyl indium (TMI) and phosphine (PH3) (PH3/TMI molar ratio = 82) for 20 minutes. The wires were not intentionally doped. High-resolution transmission electron microscopy (TEM) showed that the wires were pure wurtzite phase, consistent with photoluminescence spectroscopy. The wires had a length of ~2.7 µm and a hexagonal cross-section with an average diameter of ~180 nm, and were arranged in a square array with a 500 nm pitch. Fig. 1 shows top-down and tilted SEM images of the nanowire array.

ACS Paragon Plus Environment

5

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 33

Figure 1. a) Top-view and b) 30° tilt SEM images of as-grown WZ InP nanowire array. c) Bright field TEM image of single nanowire with POx/Al2O3 passivation stack. d) Bright field and e) HAADF scanning TEM images taken around the centre of the same nanowire, which show the bilayer structure of the deposited amorphous POx/Al2O3 stack. f) Expanded partial view of d).

Deposition of the POx/Al2O3 thin film passivation stacks was performed in an Oxford Instruments FlexAL ALD reactor. Immediately prior to loading samples for deposition, they were immersed for 1 minute in a 1% aqueous HF solution to remove the native oxide, followed by rinsing in deionized water. Phosphorus oxide (POx) deposition was performed at 25°C by exposing the samples alternately to

ACS Paragon Plus Environment

6

Page 7 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

trimethyl phosphate (TMP, (CH3)3PO4) and an O2 plasma, with separating N2 purges, in an ALD-like manner. Deposition of ALD Al2O3 capping layers was performed in situ immediately following POx deposition from trimethyl aluminum (TMA, Al(CH3)3) and O2 plasma at the same temperature. More information on the deposition processes is given in the Supplementary Information. The use of a roomtemperature deposition process was found to be critical to achieving the best passivation quality. Deposition of POx/Al2O3 stacks at higher temperatures (50°C or 100°C) resulted in consistently lower PL intensity and shorter PL lifetimes for both planar and nanowire InP samples compared to that obtained at room temperature (see Supporting Information). The thickness of the POx and Al2O3 layers was ~5 nm and ~16 nm respectively (each process was run for 100 cycles), as determined by in situ spectroscopic ellipsometry on planar InP surfaces. The POx layer appeared to be transparent within the measured range (< 5 eV) with a refractive index of 1.67 at 632 nm. Fig. 1 shows TEM images of an InP nanowire after deposition of such a POx/Al2O3 thin film stack. From the bright-field images the films appeared to be amorphous, as also indicated by X-ray diffraction measurements of similar film stacks on planar Si surfaces. The distinct bilayer structure of the passivation stack is clearly apparent in the contrast of the high-angle annular dark-field (HAADF) scanning TEM image. High-resolution TEM imaging of the POx/Al2O3 stack was unfortunately found to be impossible due to a pronounced beam-sensitivity of the POx layer, which was observed to undergo rapid partial crystallization and delamination from the nanowire surface when exposed to higher fluences of the electron beam (see TEM images in Supporting Information). This also precluded accurate analysis of the composition profile by energy-dispersive X-ray spectroscopy (EDX). To establish the composition of the deposited layers, depth-resolved (sputtered) X-ray photoelectron spectroscopy (XPS) measurements (Thermo Scientific K-Alpha system) were performed on POx/Al2O3 stacks deposited in parallel on HF-etched, polished (100) InP wafer substrates. Fig. 2a shows the

ACS Paragon Plus Environment

7

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 33

compositional depth profile determined from these measurements, while Fig. 2b shows the main photoelectron peaks at several representative depths. Consistent with the TEM measurements, the profile exhibits a distinct bilayer structure. Beneath the capping Al2O3 there is a P-rich oxide layer, as expected. Surprisingly however, this layer also appears to contain significant Al (Al:P:O ratio ≈ 0.5:1:4). This is despite the low deposition temperature and absence of post-deposition annealing which could promote Al diffusion. Near the substrate interface, indium oxides are also present in low concentrations, which can be attributed to the oxidizing effect of the O2 plasma during the initial deposition cycles when the POx film is still nucleating. The carbon concentration in the phosphate layer was ~0.7%, close to the detection limit. Note that the relative concentration of both In and Al in the POx layer is likely to be overestimated due to known selective sputtering of P with respect to both elements (this also accounts for the apparent non-stoichiometry of the InP substrate bulk: measurements of unsputtered InP wafers both as-received and after HF etching showed stoichiometric composition using the same sensitivity factors – see Supporting Information).27,

38

The apparent composition of the

phosphate layer, as well as the binding energy of the bulk P2p oxide peak (135.2 eV), are most consistent with a phosphorus-rich aluminum polyphosphate structure Al(PO3)x analogous with In(PO3)x, most likely incorporating excess oxygen in the form of OH groups (see Supporting Information for additional discussion of the XPS data).39-42

ACS Paragon Plus Environment

8

Page 9 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

Figure 2. a) Relative atomic composition determined by XPS as a function of sputtering time for the optimized POx/Al2O3 passivation stack on a polished InP (100) substrate. b) P2p, In3d5/2, Al2p, and O1s photoelectron spectra for sputter times of 20, 240, and 330s, approximately corresponding to the Al2O3 bulk, POx bulk, and near-interfacial regions, respectively.

Steady-state and time-resolved photoluminescence (PL) measurements were performed to characterize the electrical and optical quality of the nanowires and the effectiveness of the surface

ACS Paragon Plus Environment

9

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 33

passivation. In all cases the nanowires were measured as upright arrays on the growth substrate with excitation and detection parallel to the nanowire growth axis. A “positive aging” effect was noted for the passivated nanowire samples, whereby PL intensity and lifetime were observed to increase substantially over a period of weeks and months during sample storage under dark, ambient conditions. Such an effect has previously been noted for other semiconductor/dielectric interfaces, e.g. Al2O3passivated Si.43 It was found that this effect could be accelerated by post-deposition annealing at moderate temperatures. For instance, a 1 minute anneal at 250°C in N2 shortly after POx/Al2O3 deposition was sufficient to increase the PL lifetime to a level similar to that obtained after months of aging at room temperature (see Supporting Information). Unless otherwise mentioned, the PL measurements of the POx/Al2O3-passivated nanowires reported below represent stabilized values measured more than 7 months after POx/Al2O3 deposition, without annealing. The stabilized passivation layers were found to significantly enhance the PL efficiency of the nanowires. The integrated steady-state PL intensity at room temperature (295 K) was observed to increase by a factor of ~10–20 for the passivated wires compared to the as-grown wires depending on excitation intensity (Fig. 3b), with the greatest increase observed at low excitation. The PL spectrum, shown in Fig. 3a, is typical of pure WZ InP,44,

45

with an emission peak close to 1.42 eV at room

temperature. Interestingly, the dependence of PL intensity IPL on excitation intensity Iexc for both asgrown and passivated wires closely follows a power law (IPL ∝ Iexcn) over at least 7 orders of magnitude in PL intensity, with average exponents n of 1.52 and 1.41 respectively (the local value of n may be evaluated from the slope of log(IPL) vs log(Iexc)). These values can be considered as ideality factors in the one-diode equivalent circuit of the wires. While the physical interpretation of such apparent ideality factors should be treated with caution,46 increasing values in the range of 1 < n < 2 are consistent with increasing non-radiative recombination associated with bulk or surface states.

ACS Paragon Plus Environment

10

Page 11 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

In addition to the relative increase in PL efficiency apparent from Fig. 3b, it is desirable to quantify the absolute PL internal quantum efficiency (IQE). This quantity relates directly to the efficiency potential of light-emitting devices and to the open-circuit voltage potential of photovoltaic devices.37, 47 Yoo et al.48 have previously proposed that the PL IQE may be determined by fitting the roomtemperature excitation-dependent PL intensity at high excitation levels with an idealized “rate equation” describing the recombination dynamics. Using this approach, Gao et al.49 have reported an IQE of ~50% for as-grown, undoped WZ InP nanowires of similar diameter (200 nm) to our own, which constitutes the highest IQE reported for such structures. If we apply the same analysis to our own data at similar high excitation levels we derive an IQE of 61% for the as-grown wires and 75% for the passivated wires (see Supporting Information for details). Gratifying as such values would be, they are clearly not reliable given the much larger (~10 times) relative difference in PL intensity between the same samples, and thus serve rather to highlight the general unreliability of this method for determining the PL IQE. An alternative method of determining the PL IQE which is widely used in the literature is to take the ratio of integrated PL intensity at room temperature (295 K) to that at low temperature (10 K in our case). In this approach it is assumed that the efficiency of radiative recombination is 100% at low temperature, where non-radiative processes are strongly suppressed. This assumption, though not always justified,50 is supported in our case by the n = 1 power dependence of the integrated PL intensity on excitation at 10 K for both the as-grown and passivated nanowires (Fig. 3b). This provides strong evidence that the PL IQE determined in this way (Fig. 3c) is reliable in the present case. Resulting PL IQE values of 3.5% for the as-grown and 27% for the passivated nanowires are obtained at the highest excitation levels. The latter represents an unprecedented level of performance for undoped WZ InP nanowires, taking into account differences in nanowire dimensions, as illustrated in Fig. 3d.

ACS Paragon Plus Environment

11

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 33

Figure 3. a) Room-temperature (RT) steady-state PL intensity spectra of the InP nanowires, as-grown and after surface passivation with POx/Al2O3, measured at an excitation intensity of ~4.3×1018 photons cm−2 s−1. b) Integrated steady-state PL intensity of the same wires vs excitation flux at 10 K and 295 K. c) PL IQE and implied loss in solar cell open circuit voltage Voc.37, 47 at 295 K vs excitation flux,

ACS Paragon Plus Environment

12

Page 13 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

derived from the data of Fig. 3b. d) Peak PL internal quantum efficiency (IQE) at room temperature reported in the literature51-53 and in this work for undoped WZ InP nanowires and pillars versus wire or pillar diameter. Contour lines show expected IQE versus diameter for constant surface recombination velocity (SRV) assuming that all non-radiative recombination is surface-related, where each contour step towards the lower part of the figure represents a doubling of SRV.

Time-resolved PL (TRPL) measurements provide additional support for the effectiveness of the investigated surface passivation scheme. As shown in Fig. 4a, POx/Al2O3 passivation resulted in a significantly longer PL decay, evidencing an enhanced carrier lifetime due to reduced surface recombination. As observed previously for InP nanowires,54,

55

this decay is non-exponential. We

therefore choose to take the time constant characteristic of the initial, most rapid portion of the TRPL decay, which we designate τmin, as a conservative measure of the excess carrier lifetime at the peak injection level of the measurement. This yields PL decay lifetimes of 1.8 ns and 5.4 ns respectively for the as-grown and passivated nanowires, indicating a 3 times increase in carrier lifetime due to passivation. Conversely, nanowires with only the Al2O3 layer (i.e. without the POx interlayer) displayed a significantly reduced PL lifetime of 0.3 ns, showing the importance of the POx layer for passivation. The PL decay characteristics were observed to be dependent on excitation intensity (Fig. 4b). While such a dependence is expected, its exact form is rarely measured or reported. τmin was observed to first increase then decrease with increasing excitation intensity, with a qualitatively similar excitation dependence for both as-grown and passivated wires, but a significantly increased (by a factor of ~2–3) lifetime for the passivated wires over the whole excitation range. Peak PL lifetimes of 1.9 ns and 5.4 ns were measured for the as-grown and passivated wires, respectively. It should be noted that the behavior of the TRPL decay at longer times, subsequent to the initial decay, deviated from this excitation-

ACS Paragon Plus Environment

13

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 33

dependence at the lowest excitation intensities, becoming longer as excitation intensity was decreased (see Supplementary Information), which may indicate the presence of trapping.56

Figure 4. a) Room-temperature time-resolved PL decay at 870 nm of the nanowires as-grown and with either a surface-passivating POx/Al2O3 stack or Al2O3 only. Measurements were performed at an excitation intensity of ~1012 photons cm−2 pulse−1. Dashed lines show graphically the extraction of the initial minimum PL decay lifetime τmin from the data. b) τmin extracted from time-resolved PL measurements of the same as-grown and POx/Al2O3-passivated samples as a function of excitation intensity. Carrier lifetimes reported in the literature25, 49, 57-59 for undoped WZ InP nanowires of similar diameter to our own are shown for comparison.

In Fig. 4b we also show carrier lifetimes reported in the literature by various authors25,

49, 57-59

for

undoped WZ25, 49, 57, 59 or predominantly WZ58 InP nanowires of similar diameter to our own (160,58 200,49 220,57 240,25 and 335 nm59). None of these wires were deliberately passivated. It is interesting to

ACS Paragon Plus Environment

14

Page 15 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

observe that the highest values fall just below, but very close to, the data for our own as-grown wires. Since the dimensions of our wires are similar, this suggests that the bulk quality of these wires is comparable to the state-of-the-art. The large increase in lifetime observed for the passivated wires however, shows that the lifetime of the as-grown wires is still significantly surface-limited. This is contrary to the conclusions of some authors that surface recombination is practically negligible for such wires.49, 58 The lifetimes observed for the passivated wires appear to be the longest so far reported for InP nanowires by a significant margin. Li et al.45 have reported lifetimes of up to 7.4 ns for WZ InP micropillars, however these had a diameter greater than 1 µm, which significantly reduces the influence of surface recombination. Given the significantly larger (microscale) dimensions of these structures we do not include this result in the comparison of Fig. 4b. It is desirable to quantify the contribution of surface recombination in the passivated and unpassivated wires. This is not straightforward, because the contribution from bulk recombination processes for such WZ InP nanowires is not known. For a cylindrical geometry (and when diffusion is not limiting), the effective surface recombination velocity Seff is given by60 Seff = (D / 4) (τeff−1 – τb−1) where D is the nanowire diameter, τeff is the effective excess carrier lifetime, and τb is the bulk lifetime due to radiative and non-radiative processes. An upper limit on Seff may be derived by assuming that all recombination occurs at the nanowire surface (τb = ∞). If we take τeff as equal to the highest measured τmin for the passivated wires (6.5 ns, see Fig. 5), we derive an upper limit to Seff of 690 cm/s. For the as-grown wires (τmin = 1.9 ns) we similarly derive an upper limit to Seff of 2.4 × 103 cm/s. A lower limit to Seff for the asgrown wires may be derived by assuming that τb is equal to the maximum measured lifetime for the passivated samples. This yields a lower limit of Seff = 1.7 × 103 cm/s for the as-grown wires. This value is an order of magnitude higher than the value of 170 cm/s reported by Joyce et al.,58 which they derived from transient terahertz photoconductance measurements of as-grown WZ/ZB InP nanowires with a

ACS Paragon Plus Environment

15

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 33

range of diameters, but it is comparable to the value of 2.7 × 103 cm/s which may be derived in the same way from the diameter-dependent TRPL data of Tedeschi et al.59 for pure WZ wires (see Supporting Information). Note that in general Seff is expected to be dependent on dopant concentration.61 The passivated nanowires were found to be significantly more thermally stable than the as-grown wires. As mentioned earlier, the InP surface is known to be unstable at elevated temperatures, undergoing P-desorption and associated electrical degradation at temperatures above at least 200°C.18-20, 62

In fact, time-resolved PL measurements, which are extremely sensitive to surface quality, show that

the onset of electrical degradation occurs already between 100 and 150°C for “bare” (native oxide only) nanowire and planar InP surfaces, as shown in Fig. 5 for such samples annealed at various temperatures for 1 minute in N2. In contrast, the PL lifetime of POx/Al2O3-passivated samples was initially lower but was observed to increase with annealing temperature between 100 and 300°C, reaching values significantly higher (up to 6.5 ns) than those of the unannealed, unpassivated samples. Although the lifetimes of the passivated samples also declined at temperatures above 300°C, they remained higher than those of the unpassivated samples annealed at the same temperatures, indicating an improvement in the thermal stability of the surface. Fig. 5 also shows that the lifetime of samples coated with the same Al2O3 film without a POx interlayer remained well below those of the unpassivated surface regardless of annealing temperature up to 400°C, emphasizing the critical role of the POx interlayer to passivation. The fact that similar trends are observed both for the undoped WZ nanowires and for heavily doped ntype (100) ZB InP is a non-trivial result and suggests the effectiveness of the investigated passivation scheme for InP surfaces in general. The improved thermal stability of POx/Al2O3-passivated surfaces should allow the use of higher thermal budgets for InP device fabrication without compromising surface electrical properties, thereby expanding the temperature window for InP device processing.

ACS Paragon Plus Environment

16

Page 17 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

Figure 5. Initial PL decay lifetime τmin of a) undoped InP nanowires, and b) planar (100) n-type (Sdoped, Nd = 1.3×1018 cm−3) InP wafers, both as-grown/as-received (with native oxide) and with either a POx/Al2O3 passivation stack or Al2O3 only, as a function of annealing temperature (N2 ambient, 1 minute). Each data point represents a separate sample (i.e. samples were not annealed sequentially). An anneal temperature of zero indicates no anneal. Dashed lines show the decay time constant of the excitation pulse (“system response”) in a), which is the lifetime measured for the nanowires when the surface recombination is very high (as observed for samples with an Al2O3 film deposited at 200°C instead of room temperature), and in b) τmin measured for the same samples subjected to an HF dip followed by evaporation of a thin (~5 nm) Au coating. The latter has the effect of producing an effectively infinite surface recombination velocity,63 leading to diffusion-limited surface recombination, and thus represents a lower limit on the lifetime that can be obtained by degradation of the surface.

ACS Paragon Plus Environment

17

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 33

Finally, we note that the fact that we obtain good passivation of InP with a P-rich interfacial oxide runs counter to some recent thinking on InP surface passivation. It is not uncommon in the more recent InP passivation literature to find studies in which the goal is the elimination of P oxides, ostensibly for the purposes of surface passivation.64, 65 For example, based on comparison of capacitance–voltage and XPS measurements of the Al2O3 and HfO2 interfaces with InP, Galatage et al. concluded that the formation of P-rich oxides was probably harmful for InP passivation.66 Density functional theory studies have been interpreted in a similar light.67 Our results suggest that on the contrary, P-rich oxides may be beneficial for InP surface passivation. In conclusion, we have reported evidence of effective surface passivation of InP nanowire and planar surfaces by POx/Al2O3 thin-film stacks deposited at room temperature. The application of this passivation scheme allows us to achieve PL internal quantum efficiencies and lifetimes significantly higher than previously reported for wurtzite InP nanowires of similar diameter. The passivation displays excellent long-term stability, and is additionally shown to significantly improve the thermal stability of InP surfaces. Further improvements in passivation may be possible by optimizing the surface pretreatment and POx/Al2O3 film thicknesses in combination with post-deposition annealing conditions. Such passivation layers are a key enabling technology for InP nanowire device applications such as nanolasers and solar cells.

ASSOCIATED CONTENT Supporting Information. Details of POx/Al2O3 deposition process and POx optical properties, Instability of POx layers during TEM, XPS spectra of planar InP surfaces at various stages of the passivation process, Further discussion of XPS profile results, Experimental details of PL

ACS Paragon Plus Environment

18

Page 19 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

measurements, Evolution of POx/Al2O3 passivation with storage time, Influence of deposition temperature on passivation quality, Details of PL IQE extraction according to the method of Yoo et al., Calculation of expected diameter dependence of PL IQE, Detailed excitation dependence of timeresolved PL decay at low excitation levels, Extraction of Seff from the data of Tedeschi et al. AUTHOR INFORMATION Corresponding Authors * E-mail: [email protected]. * E-mail: [email protected]. Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Funding Sources This work was supported by the Gravitation/Zwaartekracht program “Research Centre for Integrated Nanophotonics” of the Netherlands Organization for Scientific Research (NWO). Notes The authors declare no competing financial interest. ACKNOWLEDGMENTS The authors gratefully acknowledge E. Smalbrugge and C.A.A. van Helvoirt for technical assistance, J. Greil, L. Gagliano, S. Assali, and A. Dijkstra for advice and assistance with PL measurements, and B.W.H. van de Loo for feedback on the draft manuscript. TEM measurements were performed at a facility funded by Solliance and the Dutch province of Noord-Brabant.

ACS Paragon Plus Environment

19

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 33

ABBREVIATIONS ALD, atomic layer deposition; PL, photoluminescence; ZB, zincblende; WZ, wurtzite; MIS, metal– insulator–semiconductor; PECVD, plasma-enhanced chemical vapor deposition; TMI, trimethyl indium; TEM, transmission electron microscopy; SEM, scanning electron microscopy; HF, hydrofluoric acid; TMP, trimethyl phosphate; SE, spectroscopic ellipsometry; TMA, trimethyl aluminum; HAADF, highangle annular dark-field; EDX, energy-dispersive X-ray spectroscopy; XPS, X-ray photoelectron spectroscopy; TRPL, time-resolved photoluminescence; IQE, internal quantum efficiency. REFERENCES (1)

del Alamo, J. A. Nature 2011, 479, 317–323.

(2)

Wallace, R. M.; McIntyre, P. C.; Kim, J.; Nishi, Y. MRS Bull. 2009, 34, 493–503.

(3)

Heyns, M.; Bellenger, F.; Brammertz, G.; Caymax, M.; Cantoro, M.; De Gendt, S.; De Jaeger,

B.; Delabie, A.; Eneman, G.; Groeseneken, G.; Hellings, G.; Houssa, M.; Iacopi, F.; Leonelli, D.; Lin, D.; Magnus, W.; Martens, K.; Merckling, C.; Meuris, M.; Mitard, J.; Penaud, J.; Pourtois, G.; Scarrozza, M.; Simoen, E. R.; Soree, B.; Van Elshocht, S.; Vandenberghe, W.; Vandooren, A.; Vereecke, P.; Verhulst, A.; Wang, W.-E. Shaping the future of nanoelectronics beyond the Si roadmap with new materials and devices. Proc. SPIE. 2010; p 764003. (4)

Takagi, S.; Zhang, R.; Suh, J.; Kim, S.-H.; Yokoyama, M.; Nishi, K.; Takenaka, M. Jpn. J. Appl.

Phys. 2015, 54, 06FA01. (5)

Wilmsen, C. W., Ed. Physics and chemistry of III-V compound semiconductor interfaces;

Springer US, 1985. (6)

Houssa, M.; Chagarov, E.; Kummel, A. MRS Bull. 2009, 34, 504–513.

ACS Paragon Plus Environment

20

Page 21 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

(7)

Robertson, J.; Guo, Y.; Lin, L. J. Appl. Phys. 2015, 117, 112806.

(8)

Radosavljevic, M.; Chu-Kung, B.; Corcoran, S.; Dewey, G.; Hudait, M. K.; Fastenau, J. M.;

Kavalieros, J.; Liu, W. K.; Lubyshev, D.; Metz, M.; Millard, K.; Mukherjee, N.; Rachmady, W.; Shah, U.; Chau, R. Advanced high-K gate dielectric for high-performance short-channel In0.7Ga0.3As quantum well field effect transistors on silicon substrate for low power logic applications. 2009 IEEE Int. Electron Devices Meet. (IEDM). 2009; pp 1–4. (9)

van Dam, D.; van Hoof, N. J. J.; Cui, Y.; van Veldhoven, P. J.; Bakkers, E. P. A. M.;

Gómez Rivas, J.; Haverkort, J. E. M. ACS Nano 2016, 10, 11414–11419. (10)

Wallentin, J.; Anttu, N.; Asoli, D.; Huffman, M.; Aberg, I.; Magnusson, M. H.; Siefer, G.; Fuss-

Kailuweit, P.; Dimroth, F.; Witzigmann, B.; Xu, H. Q.; Samuelson, L.; Deppert, K.; Borgström, M. T. Science 2013, 339, 1057–1060. (11)

Wang, Z.; Tian, B.; Pantouvaki, M.; Guo, W.; Absil, P.; Van Campenhout, J.; Merckling, C.;

Van Thourhout, D. Nat. Photonics 2015, 9, 837–842. (12)

Casey, H. C.; Buehler, E. Appl. Phys. Lett. 1977, 30, 247–249.

(13)

Rosenwaks, Y.; Shapira, Y.; Huppert, D. Phys. Rev. B 1991, 44, 13097–13100.

(14)

Bothra, S.; Tyagi, S. D.; Ghandhi, S. K.; Borrego, J. M. Surface recombination velocity and

lifetime in InP measured by transient microwave reflectance. IEEE Photovoltaic Spec. Conf.. 1990; pp 404–408 vol.1. (15)

Hoffman, C. A.; Jarašiunas, K.; Gerritsen, H. J.; Nurmikko, A. V. Appl. Phys. Lett. 1978, 33,

536–539.

ACS Paragon Plus Environment

21

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(16)

Page 22 of 33

Ahrenkiel, R. K. In Properties, processing and applications of indium phosphide; Pearsall, T. P.,

Ed.; Inst of Engineering & Technology, 2000; pp 105–110. (17)

Wang, X.; Bhosale, J.; Moore, J.; Kapadia, R.; Bermel, P.; Javey, A.; Lundstrom, M. IEEE

Journal of Photovoltaics 2015, 5, 282–287. (18)

Krawczyk, S.; Bailly, B.; Sautreuil, B.; Blanchet, R.; Viktorovitch, P. Electron. Lett. 1984, 20,

255–256. (19)

Sautreuil, B.; Viktorovitch, P.; Blanchet, R. J. Appl. Phys. 1985, 57, 2322–2324.

(20)

Chang, R. R.; Iyer, R.; Lile, D. L. J. Appl. Phys. 1987, 61, 1995–2004.

(21)

van Vugt, L. K.; Veen, S. J.; Bakkers, E. P. A. M.; Roest, A. L.; Vanmaekelbergh, D. J. Am.

Chem. Soc. 2005, 127, 12357–12362. (22)

Tajik, N.; Haapamaki, C. M.; LaPierre, R. R. Nanotechnology 2012, 23, 315703.

(23)

Münch, S.; Reitzenstein, S.; Borgström, M.; Thelander, C.; Samuelson, L.; Worschech, L.;

Forchel, A. Nanotechnology 2010, 21, 105711. (24)

Dhaka, V.; Perros, A.; Naureen, S.; Shahid, N.; Jiang, H.; Kakko, J.-P.; Haggren, T.;

Kauppinen, E.; Srinivasan, A.; Lipsanen, H. AIP Adv. 2016, 6. (25)

Zhong, Z.; Li, Z.; Gao, Q.; Li, Z.; Peng, K.; Li, L.; Mokkapati, S.; Vora, K.; Wu, J.; Zhang, G.;

Wang, Z.; Fu, L.; Tan, H. H.; Jagadish, C. Nano Energy 2016, 28, 106 – 114. (26)

Bayliss, C. R.; Kirk, D. L. J. Phys. D: Appl. Phys. 1976, 9, 233.

(27)

Meiners, L. Thin Solid Films 1984, 113, 85 – 92.

ACS Paragon Plus Environment

22

Page 23 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(28)

Nano Letters

Schachter, R.; Olego, D. J.; Baumann, J. A.; Bunz, L. A.; Raccah, P. M.; Spicer, W. E. Appl.

Phys. Lett. 1985, 47, 272–274. (29)

Pande, K. P.; Gutierrez, D. Appl. Phys. Lett. 1985, 46, 416–418.

(30)

Chang, H. L.; Meiners, L. G.; Sa, C. J. Appl. Phys. Lett. 1986, 48, 375–377.

(31)

Hirota, Y.; Kobayashi, T. J. Appl. Phys. 1982, 53, 5037–5043.

(32)

Hbib, H.; Bonnaud, O.; Gauneau, M.; Hamedi, L.; Marchand, R.; Quemerais, A. Thin Solid

Films 1997, 310, 1–7. (33)

Robach, Y.; Joseph, J.; Bergignat, E.; Commere, B.; Hollinger, G.; Viktorovitch, P. Appl. Phys.

Lett. 1986, 49, 1281–1283. (34)

Robach, Y.; Besland, M. P.; Joseph, J.; Hollinger, G.; Viktorovitch, P.; Ferret, P.; Pitaval, M.;

Falcou, A.; Post, G. J. Appl. Phys. 1992, 71, 2981–2992. (35)

Kobayashi, T.; Ichikawa, T.; Sakuta, K.; Fujisawa, K. J. Appl. Phys. 1984, 55, 3876–3878.

(36)

Choujaa, A.; Chave, J.; Blanchet, R.; Viktorovitch, P. J. Appl. Phys. 1986, 60, 2191–2193.

(37)

Mann, S. A.; Oener, S. Z.; Cavalli, A.; Haverkort, J. E. M.; Bakkers, E. P. A. M.; Garnett, E. C.

Nat. Nanotechnol. 2016, 11, 1071–1075. (38)

Malherbe, J.; Barnard, W. Surf. Sci. 1991, 255, 309 – 320.

(39)

Hollinger, G.; Bergignat, E.; Joseph, J.; Robach, Y. J. Vac. Sci. Technol., A 1985, 3, 2082–2088.

ACS Paragon Plus Environment

23

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(40)

Page 24 of 33

Faur, M.; Faur, M.; Jayne, D. T.; Goradia, M.; Goradia, C. Surf. Interface Anal. 1990, 15, 641–

650. (41)

Sferco, S. J.; Allan, G.; Lefebvre, I.; Lannoo, M.; Bergignat, E.; Hollinger, G. Phys. Rev. B

1990, 42, 11232–11239. (42)

Atuchin, V. V.; Kesler, V. G.; Pervukhina, N. V. Surf. Rev. Lett. 2008, 15, 391–399.

(43)

Dingemans, G.; Engelhart, P.; Seguin, R.; Einsele, F.; Hoex, B.; Sanden, M. C. M. v. d.;

Kessels, W. M. M. J. Appl. Phys. 2009, 106. (44)

Zilli, A.; De Luca, M.; Tedeschi, D.; Fonseka, H. A.; Miriametro, A.; Tan, H. H.; Jagadish, C.;

Capizzi, M.; Polimeni, A. ACS Nano 2015, 9, 4277–4287. (45)

Li, K.; Ng, K. W.; Tran, T.-T. D.; Sun, H.; Lu, F.; Chang-Hasnain, C. J. Nano Lett. 2015, 15,

7189–7198. (46)

McIntosh, K. R. Lumps, humps and bumps: Three detrimental effects in the current–voltage

curve of silicon solar cells. Ph.D. thesis, University of New South Wales, 2001. (47)

Cui, Y.; van Dam, D.; Mann, S. A.; van Hoof, N. J. J.; van Veldhoven, P. J.; Garnett, E. C.;

Bakkers, E. P. A. M.; Haverkort, J. E. M. Nano Lett. 2016, 16, 6467–6471. (48)

Yoo, Y.-S.; Roh, T.-M.; Na, J.-H.; Son, S. J.; Cho, Y.-H. Appl. Phys. Lett. 2013, 102, 211107.

(49)

Gao, Q.; Saxena, D.; Wang, F.; Fu, L.; Mokkapati, S.; Guo, Y.; Li, L.; Wong-Leung, J.;

Caroff, P.; Tan, H. H.; Jagadish, C. Nano Lett. 2014, 14, 5206–5211.

ACS Paragon Plus Environment

24

Page 25 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(50)

Nano Letters

Watanabe, S.; Yamada, N.; Nagashima, M.; Ueki, Y.; Sasaki, C.; Yamada, Y.; Taguchi, T.;

Tadatomo, K.; Okagawa, H.; Kudo, H. Appl. Phys. Lett. 2003, 83, 4906–4908. (51)

Li, K.; Sun, H.; Ren, F.; Ng, K. W.; Tran, T.-T. D.; Chen, R.; Chang-Hasnain, C. J. Nano Lett.

2014, 14, 183–190. (52)

Ren, F.; Ng, K. W.; Li, K.; Sun, H.; Chang-Hasnain, C. J. Appl. Phys. Lett. 2013, 102, 012115.

(53)

Tran, T.-T. D.; Sun, H.; Ng, K. W.; Ren, F.; Li, K.; Lu, F.; Yablonovitch, E.; Chang-

Hasnain, C. J. Nano Lett. 2014, 14, 3235–3240. (54)

Wagner, H. P.; Kaveh, M.; Gao, Q.; Tan, H.; Jagadish, C.; Langbein, W. Phys. Rev. B 2017, 95,

045305. (55)

Zhang, W.; Lehmann, S.; Mergenthaler, K.; Wallentin, J.; Borgström, M. T.; Pistol, M.-E.;

Yartsev, A. Nano Lett. 2015, 15, 7238–7244. (56)

Heinz, F. D.; Warta, W.; Schubert, M. C. Sol. Energy Mater. Sol. Cells 2016, 158, 107 – 114,

Proceedings of the 6th International Conference on Silicon Photovoltaics. (57)

Wang, F.; Gao, Q.; Peng, K.; Li, Z.; Li, Z.; Guo, Y.; Fu, L.; Smith, L. M.; Tan, H. H.;

Jagadish, C. Nano Lett. 2015, 15, 3017–3023. (58)

Joyce, H. J.; Wong-Leung, J.; Yong, C.-K.; Docherty, C. J.; Paiman, S.; Gao, Q.; Tan, H. H.;

Jagadish, C.; Lloyd-Hughes, J.; Herz, L. M.; Johnston, M. B. Nano Lett. 2012, 12, 5325–5330. (59)

Tedeschi, D.; De Luca, M.; Fonseka, H. A.; Gao, Q.; Mura, F.; Tan, H. H.; Rubini, S.;

Martelli, F.; Jagadish, C.; Capizzi, M.; Polimeni, A. Nano Lett. 2016, 16, 3085–3093.

ACS Paragon Plus Environment

25

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(60)

Page 26 of 33

Dan, Y.; Seo, K.; Takei, K.; Meza, J. H.; Javey, A.; Crozier, K. B. Nano Lett. 2011, 11, 2527–

2532. (61)

McIntosh, K. R.; Black, L. E. J. Appl. Phys. 2014, 116, 014503.

(62)

Lambrinos, M. F.; Besland, M. P.; Gagnaire, A.; Louis, P.; Callard, S.; Joseph, J. J.

Electrochem. Soc. 1997, 144, 2086–2095. (63)

Rosenwaks, Y.; Shapira, Y.; Huppert, D. Appl. Phys. Lett. 1990, 57, 2552–2554.

(64)

Brennan, B.; Dong, H.; Zhernokletov, D.; Kim, J.; Wallace, R. M. Appl. Phys. Express 2011, 4,

125701. (65)

Dong, H.; Santosh, K.; Qin, X.; Brennan, B.; McDonnell, S.; Zhernokletov, D.; Hinkle, C. L.;

Kim, J.; Cho, K.; Wallace, R. M. J. Appl. Phys. 2013, 114, 154105. (66)

Galatage, R. V.; Dong, H.; Zhernokletov, D. M.; Brennan, B.; Hinkle, C. L.; Wallace, R. M.;

Vogel, E. M. Appl. Phys. Lett. 2013, 102, 132903. (67)

KC, S.; Dong, H.; Longo, R. C.; Wang, W.; Xiong, K.; Wallace, R. M.; Cho, K. J. Appl. Phys.

2014, 115, 023703.

ACS Paragon Plus Environment

26

Page 27 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

For Table of Contents Only

ACS Paragon Plus Environment

27

(a) 1 2 3 4 5 6 7 8(c) 9 10 11 12 13 14 15 16 17 18 (e) 19 20 21 22 23 24 25 26 27

Nano Letters (b)

0.5 μm

Page 28 of 33

0.5 μm

(d)

1 μm

50 nm

(f) InP POx Al2O3 ACS Paragon Plus Environment

50 nm

10 nm

(a) 70

Page 29 of 33

Atomic %

60

O

Nano Letters

In (InP)

50

Counts (arb. units)

1 P (InP) 40 Al (ox.) 2 3 30 4 P (ox.) 20 5 6 10 In (ox.) 7 0 8 0 100 200 300 400 500 9 Sputter time (s) 10 (b)11 20s InP 12 P2p In3d5/2 240s 13 ox. 330s 14 ×50 15 InP 16 ox. 17 18 19 135 130 448 446 444 442 20 21 Al2p O1s 22 23 24 25 26 27 ACS Paragon Plus Environment 28 76 74 72 535 530 29 78 Binding Energy (eV) 30

(b)

Nano Letters

Integrated PL Intensity (arb. units)

(a)

1012

Excitation Intensity (kW Page cm−230 ) of 33 10−5 10−4 10−3 10−2 10−1 100 101

PL IQE

Implied Voc loss (mV)

PL IQE

PL Intensity (104 counts/s)

1 1.5 InP nanowires InP nanowires 2 RT 3 109 n=1 4 10 K 5 6 1 POx/Al2O3 7 n = 1.52 106 8 9 n = 1.41 295 K 100.5 11 103 POx/Al2O3 12 As-grown As-grown 13 14 0 800 850 900 950 1016 1017 1018 1019 1020 1021 1022 1023 15 750 16 Wavelength (nm) Excitation Flux (photons cm−2 s−1) 17 (c)18 (d) Excitation Intensity (kW cm−2) 19 −5 −4 −3 −2 −1 100 101 20100 10 10 10 10 10 0 100 21 InP nanowires 22 23 −1 295K 50 POx/Al2O3 10 24 POx/Al2O3 Li 25 Tran 26 100 27 −2 10 10−1 Ren 28 29 As-grown 150 30 As-grown 31 10−3 32 200 Increasing SRV 33 34 −2 ACS Paragon Plus 10 Environment 10−4 16 35 10 1017 1018 1019 1020 1021 1022 1023 102 103 36 −2 −1 Diameter (nm) 37 Excitation Flux (photons cm s )

(a)

(b)

Excitation Nano Letters

Page 31 of 33

10

InP nanowires RT

10−1

100

InP nanowires RT

POx/Al2O3

τmin (ns)

Normalized PL Intensity

1 2 3 4 5 6 710−1 8 9 10 11 12 13 14 10−2 0 15 16

10−2

10

0

Intensity (μJ cm−2 pulse−1)

τmin = 5.4 ns

As-grown τmin = 1.8 ns

102

POx/Al2O3

As-grown Gao Tedeschi Wang Zhong Joyce

1

Al2O3

101

τmin = 0.3 ns 5

10

15

20

Time (ns)

ACS Paragon Plus Environment 10 11

25

30

10

10

1012

1013

1014 −2

1015

Excitation Flux (photons cm pulse−1)

(a) 10

(b)

InP NW

POx/Al2O3

Nano Letters

10

InP (100) n-type

POx/Al2O3Page 32 of 33

τmin (ns)

τmin (ns)

1 As-grown 2 As-received 3 4 1 5 6 7 Al2O3 8 Al2O3 9 10 System response 1 Au-coated (diffusion-limited) 11 ACS Paragon Plus Environment 120.1 0 100 200 300 400 0 100 200 300 400 13 Anneal Temperature (°C) Anneal Temperature (°C) 14

1 2 3 4 5 6 7

Photoluminescence Intensity (arb.)

Al O PO

2 3 x Page 33 of 33

Nano Letters1.5

0.5 μm

InP

POx/Al2O3 1

0.5

ACS Paragon Environment InP POx Al2OPlus 3 As-grown 10 nm

0

800 850 900 Wavelength (nm)