Effect of Electric Dipoles on Fermi Level Positioning at the Interface

Aug 8, 2012 - Laboratorio MDM, IMM-CNR, via C. Olivetti 2, I-20864, Agrate Brianza (MB), Italy. ‡. Dipartimento di Scienza dei Materiali, UniversitÃ...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/JPCC

Effect of Electric Dipoles on Fermi Level Positioning at the Interface between Ultrathin Al2O3 Films and Differently Reconstructed In0.53Ga0.47As(001) Surfaces Carlo Grazianetti,*,†,‡ Alessandro Molle,*,† Grazia Tallarida,† Sabina Spiga,† and Marco Fanciulli†,‡ †

Laboratorio MDM, IMM-CNR, via C. Olivetti 2, I-20864, Agrate Brianza (MB), Italy Dipartimento di Scienza dei Materiali, Università degli Studi di Milano Bicocca, via R. Cozzi 53, I-20126, Milano (MI), Italy



ABSTRACT: The role played by electric dipoles at the interface Al2O3/In0.53Ga0.47As(001) has been investigated in a whole in situ approach by means of scanning tunneling microscopy/spectroscopy and X-ray photoelectron spectroscopy (XPS) in order to study local and macroscopic features, respectively. The initial pinned Fermi level (FL) shifts toward midgap after oxide deposition in both (2 × 4) and (4 × 2) reconstructions, while a reconstruction-dependent FL restoring is observed after annealing at 200 °C. This behavior is rationalized in terms of formation and suppression of positive charge in the as-grown oxide resulting from in situ XPS of the interface band line up.



INTRODUCTION Fermi level (FL) pinning is a physical effect which takes place at the surface of semiconductors whenever electrically active trap states partially or completely screen the semiconductor charge carriers from external electric field modulation.1 This phenomenon can occur in a number of physically different configurations, and it is particularly detrimental for highmobility III−V semiconductors which are intended as active channels for post-Si scaling of metal−oxide−semiconductor field effect transistors (MOSFETs).2 Among III−V semiconductors, In0.53Ga0.47As(001) shows high electronic mobility, direct band gap, and high breakdown voltage.3 However, the In0.53Ga0.47As surface, as other III−V semiconductors, is known to suffer from FL pinning caused by interface-induced gap states which may arise from native defects of the semiconductor. 4,5 Indeed, the key point to conceive an In0.53Ga0.47As-based MOSFET is to fabricate an unpinned oxide−semiconductor interface with low fixed charge and low trap density.6 In particular, a capping oxide should be able to passivate electrically active states in the In0.53Ga0.47As gap such as the dangling bonds or homopolar bonding.7 Since the electrical quality of the oxide/III−V semiconductor interface has been demonstrated to depend on the atomic details of the surface reconstruction and on the chemical bonding between the semiconductor and the oxide,7,8 it is important to investigate the oxide/In0.53Ga0.47As interface at the atomicscale level. To this scope, combining a molecular beam epitaxy (MBE) approach to the growth of ultrathin Al2O3 films with in situ atomically resolved scanning tunneling microscopy/spectroscopy (STM/STS) may offer a superior tool to manipulate and control the local properties at the interface level, as recently proposed for in situ oxidation of Ge.9 Correlating the latter properties with a macroscopic characterization of the electronic © 2012 American Chemical Society

band structure of Al2O3/In0.53Ga0.47As heterojunctions is here intended to provide a deeper insight into the physical details of the In0.53Ga0.47As-based interface which can be exploited in new methodologies for In0.53Ga0.47As surface passivation with Al2O3 interface layer10 and for integration of high-permittivity oxides on III−V semiconductors.11 In this respect, the present work reports on the study of the topographic and electronic features at the interface between MBE-grown ultrathin Al2O3 and two different In0.53Ga0.47As reconstructions based on in situ STM/ STS and X-ray photoelectron spectroscopy (XPS).



EXPERIMENTAL METHODS All experiments were carried out in an ultra-high-vacuum environment with base pressure in the 10−10 mbar range. Substrates used were 1 μm thick n-type (Si doping of 1 × 1017 cm−3) In0.53Ga0.47As(001) grown by MBE on n-type InP(001) wafers and capped with an As protective layer which was thermally desorbed in the temperature range 365−440 °C for 15 min. The As decapping procedure was in situ monitored by reflection high-energy electron diffraction and the surface reconstruction examined by in situ XPS and STM/STS. Preventing oxidation of the III−V surface during gate dielectric deposition could be essential to achieve FL unpinning.12 For this purpose, before Al2O3 deposition, one monolayer of Al has been grown by MBE in order to prevent In0.53Ga0.47As oxidation.13 Al2O3 films were grown at room temperature by means of the reactive deposition technique, i.e., a metallic Al flux is directed to the sample, while the growth chamber is filled Received: May 2, 2012 Revised: July 13, 2012 Published: August 8, 2012 18746

dx.doi.org/10.1021/jp3042318 | J. Phys. Chem. C 2012, 116, 18746−18751

The Journal of Physical Chemistry C

Article

by ultrapure O2 at 10−6 mbar pressure. XPS investigation of the Al2O3/In0.53Ga0.47As interface was carried out using standard Mg (1253.6 eV) and Al (1486.6 eV) Kα radiation sources with a pass energy of 20 eV and a take-off angle of 70°. Finally, postdeposition annealing (PDA) at 200 °C has been performed with O2 at a pressure of 10−6 mbar. Filled states STM images were acquired at −1.5 V bias with an etched tungsten tip and at a tunneling current set point of 0.3−0.5 nA. STS was performed using a lock-in amplifier with a sine wave reference signal of amplitude 50 mV and frequency ≈ 1.5 kHz. Several dI/dV curves are averaged together to establish the electronic structure of the surfaces.

reversed in a group III-rich (4 × 2) reconstruction.3,14 (2 × 4) surface reconstruction is made of a top row (along the [−110] direction) of dimerized As atoms which are bonded to In or Ga atoms. Conversely, the (4 × 2) reconstruction exhibits undimerized In/Ga atoms along the rows divided by trough regions containing two In/Ga dimers with a mean separation of 1.7 nm.3,14 After ∼3 nm thick Al2O3 deposition, the observed morphology is strongly different (Figure 1c and 1d) from the previous ones. The Al2O3 film on (2 × 4) reconstruction is characterized by large (tens of nanometers) three-dimensional islands. On the (4 × 2) reconstruction the growth mode is quite similar, but there are smaller (few nanometers) islands on the larger ones than on the (2 × 4) surface. In both cases the pristine In0.53Ga0.47As surface is fully covered by the Al2O3 film. STS experimentally provides the local electronic structure of the surface.15,16 In particular, STS may determine unambiguously the FL position with respect to the valence band maximum (VBM) and conduction band minimum (CBM). For a clean and unpinned n-type semiconductor surface, the FL should reside nearby the CBM. However, STS spectra in Figure 2a show that in both surface reconstructions of the n-type In0.53Ga0.47As the FL is located in the proximity of the VBM. This p-type character is attributed to FL pinning at the surface.6,8 While theoretical calculations show that surface FL pinning for (4 × 2) reconstruction is due to strained unbuckled dimers in the trough,8 so far neither a clear theoretical nor an experimental explanation has been proposed for the (2 × 4) case. Nonetheless, given the deeply different atomic arrangement of the two reconstructions (see Figure 1), a common physical origin of the FL pinning can be related to intrinsic defects in the proximity of the In0.53 Ga 0.47 As surface consistently with the theoretical prediction of As dangling bonds resonant at ∼0.1 eV above VBM. 5 After Al 2 O 3 deposition on both reconstructions a FL shift toward midgap is observed (green curves in Figure 2b and 2c). This shift can be estimated in ∼0.25 and ∼0.2 eV for the Al2 O3/ In0.53Ga0.47As-(2 × 4) and Al2O3/In0.53Ga0.47As-(4 × 2) interfaces, respectively. The presence of the ultrathin oxide films induces a little enlargement in the band-gap measurement, which is limited by the low thickness of the insulating film itself, thus ensuring the possibility to still perform STM and STS. A reconstruction-dependent behavior after PDA at 200 °C can be singled out. On the (2 × 4) reconstruction the FL slightly shifts toward the original condition (dark gray curve in Figure 2b), while on the other hand, on the (4 × 2) reconstruction the original FL pinning condition is fully restored (dark gray curve in Figure 2c), i.e., VB maxima of the STS spectra coincide. According to the (4 × 2) reconstruction, this behavior might be related to surface roughening as argued in the case of HfO2 on



RESULTS AND DISCUSSION The STM topographies in Figure 1a and 1b show As-rich (2 × 4) and group III-rich (4 × 2) In0.53Ga0.47As reconstructions,

Figure 1. STM topographies of In0.53Ga0.47As(001) (2 × 4) reconstruction (a), (4 × 2) reconstruction (b), Al2O3 thin film on (2 × 4) reconstruction (c), and on (4 × 2) reconstruction (d). All images are 70 × 70 nm2.

respectively. After decapping at 365 °C the resulting reconstruction is an As-rich (2 × 4) with periodicity 2× and 4×, respectively, along the [−110] and [110] directions. On the other hand, when annealed at 440 °C the configuration is

Figure 2. STS spectra for both clean reconstructions (2 × 4) (red line) and (4 × 2) (blue line) (a); after oxide deposition (green line) and postdeposition annealing (dark gray line) for (2 × 4) reconstruction (b) and (4 × 2) reconstruction (c). Fermi level position is at 0 eV. 18747

dx.doi.org/10.1021/jp3042318 | J. Phys. Chem. C 2012, 116, 18746−18751

The Journal of Physical Chemistry C

Article

Figure 3. XPS As 2p3/2, In 3d5/2, and Ga 2p3/2 lines upon Al2O3 deposition on (2 × 4) reconstruction (a) and after postdeposition annealing (b). Similar results have been obtained for the (4 × 2) case.

between the Al−Al bond and the Al−(InGaAs) bond.21 Nevertheless, the reactive deposition technique ensures a good reactivity of the Al metallic flux in the presence of O2. In fact, use of O2 leads to a reduced oxidation of the semiconductor substrate with respect the use of other more reactive oxidizers, e.g., atomic oxygen,13 and yields a chemical shift of the Al3+ component in the Al 2p line (with respect to the Al* component) of 2.17 and 1.67 eV for the (2 × 4) and (4 × 2) reconstruction, respectively, which is consistent with sesquioxide stoichiometry bonding (Figure 4). The Al* component does not vary its binding energy before and after PDA in the Al2O3/In0.53Ga0.47As(2 × 4) interface, whereas a different energetic position is found in the Al2O3/In0.53Ga0.47As(4 × 2) case. It could be possible that in the (2 × 4) reconstruction, i.e., an As-rich surface, the Al−(InGaAs) part of Al* may be dominated by Al−As bonding.21 This bonding could be stable after PDA, leaving the Al* peak unaltered. On the other hand, in the (4 × 2) reconstruction, i.e., a group IIIrich surface, the Al−(InGaAs) part of Al* could be more equally shared by In and Ga. Since Ga and Al have similar electronegativities,21 the slightly shift after PDA can be explained with a more balanced interplay between the Al− (InGaAs) and the Al−Al bonds, with an increase of this latter component. The reason why the interface is only partially pinned after Al2O3 deposition can be found in the limited In and Ga oxidations, which were recently found to generate a high density of electrically active traps in the midgap region.14 Nonetheless, partial FL unpinning observed from STS data can be elucidated by means of in situ XPS investigation of the electronic band structure at the Al2O3/In0.53Ga0.47As interface. Indeed, it is well known from Kraut’s determination of the interface band line up that the valence band offset (VBO) in an oxide−semiconductor heterojunction can be inferred by measuring the mutual energy separation between XPS lines of the semiconductor and of the oxide with respect to the respective valence band edges.22 In particular, as previously reported in the Gd2O3/Ge heterojunction,23 a process-dependent increase (decrease) of

In0.53Ga0.47As even if at higher annealing temperature (300 °C).7 In fact, also for Al2O3, a rms roughness is observed to increase from ∼1.3 to ∼1.8 nm after annealing at 200 °C for the Al2O3/In0.53Ga0.47As-(4 × 2) interface. On the other hand, in the Al2O3/In0.53Ga0.47As-(2 × 4) interface, the roughness value slowly decreases from ∼1.6 to ∼1.5 nm. Chemical analysis carried out by means of in situ XPS can shed light on the role played by the interface in both reconstructions. XPS spectra of the As 2p3/2, As 3d, In 3d5/2, Ga 2p3/2, and Al 2p lines have been collected for the as-grown Al2O3/In0.53Ga0.47As interfaces and subsequently PDA. Figure 3a shows the As 2p3/2, In 3d5/2, and Ga 2p3/2 lines related to the (2 × 4) reconstruction, and Figure 3b shows the same lines after PDA on the same reconstruction. Similar results have been obtained for the (4 × 2) case (data not shown). Spectra were decomposed with pseudo-Voigt functions after Shirley background removal. The As 2p3/2 lines of both interfaces are rid of As oxides irrespectively of the pristine surface reconstruction. This result is also confirmed by probing the As 3d line (Figure 4). The absence of As oxides is a good starting point in order to achieve effective passivation.17 On the contrary, In 3d5/2 and Ga 2p3/2 lines have an extra component along with the bulk one. For the In case, in both interfaces the extra component can be associated with the presence of two different chemical states In1+ (In2O) and In3+ (In2O3),18 with the amount of the former one being lower than the latter one. In the Ga 2p3/2 lines the extra component can be associated with the formation of Ga2O because the chemical shift is 0.6 eV for both reconstructions (Figure 3a).19 After PDA, the As 2p3/2 still has no oxides, the In2O content decreases, and In2O3 is fairly unaffected whereas the amount of Ga2O increases (Figure 3b) but is not assumed to be responsible for FL pinning.19 This increase in Ga2O content could be likely associated with the concomitant In2O decrease or the presence of O2 during PDA. In Figure 4 the Al 2p lines confirm in both reconstructions the presence of the Al3+ oxidation state at ∼75 eV, which is consistent with formation of Al2O3.20,21 An additional minor Al component at lower binding energy, ∼73 eV, denoted with Al* (green curves in Figure 4a and 4b), can be associated with an interplay 18748

dx.doi.org/10.1021/jp3042318 | J. Phys. Chem. C 2012, 116, 18746−18751

The Journal of Physical Chemistry C

Article

Figure 4. O 1s, Al 2p, and As 3d XPS lines for Al2O3 on (2 × 4) reconstruction as grown (top) and after annealing (bottom) (a) and for the (4 × 2) case (b). Doublets have been used to account for spin−orbit splitting (∼0.68 eV) in the As 3d XPS line.

Table 1. XPS Absolute Energya Differences for As-Grown Al2O3 on (2 × 4) Reconstruction (a), after Annealing at 200 °C (b), As-Grown Al2O3 on (4 × 2) Reconstruction (c), and after Annealing at 200 °C (d)

the absolute energy difference Δo‑s between an oxide and a semiconductor XPS line results in an increase (decrease) of the VBO provided that (a) no change in the composition and/or stoichiometry takes place in the oxide or at the interface level and (b) no band bending occurs in the semiconductor. It can be noticed that Δo‑s(O1s−As3d) = BE(O1s) − BE(As3d) undergoes a sensible decrease when measured after annealing with respect to the as-grown Al2O3, as evidenced in Figure 4a for the (2 × 4) reconstructed surface and Figure 4b for the (4 × 2) one using the As 3d and the O 1s lines as representative for the semiconductor and oxide, respectively. In detail, the quantitative amounts of Δo‑s as a function of the semiconductor lines, As 3d, Ga 2p3/2, and In 3d5/2, are reported in Table 1 for each process stage. A univocal trend appears irrespectively of the choice of the XPS lines. Indeed, from Table 1 it is possible to conclude that the decrease of the Δ(O1s−As3d) term is a common feature of all semiconductor lines which can be associated with a corresponding decrease of the VBO after annealing. This assignment is justified by the annealing-

sample

Δ(O1s− As3d), eV

Δ(O1s− In3d5/2), eV

Δ(Ga2p3/2− O1s), eV

Δ(O1s− Al2p), eV

Δ(Al2p− As3d), eV

a b c d

491.73 491.15 491.58 491.18

88.35 87.72 88.28 87.80

584.65 585.29 584.67 585.15

456.90 456.79 456.89 456.86

34.81 34.37 34.66 34.33

a

Uncertainty is ±0.08 eV.

independent observation of a constant energy separation between the O 1s and the Al 2p XPS lines, indicative of a stoichiometric invariance, and of the fixed position of all the semiconductor lines, indicative of no distortion of semiconductor electronic bands. A number of physical mechanisms can be invoked to interpret this annealing-induced VBO reduction; however, the more likely event can be related to 18749

dx.doi.org/10.1021/jp3042318 | J. Phys. Chem. C 2012, 116, 18746−18751

The Journal of Physical Chemistry C

Article

with MBE-grown ultrathin Al2O3 layers, the (2 × 4) reconstruction is more suitable to maintain a FL partially unpinned in a wider temperature range than the (4 × 2) reconstruction does, thus proving to be more favorable for application-oriented perspectives of In0.53Ga0.47As as an active channel in MOSFET devices.

relaxation of positively charged defects, e.g., Al dangling bonds24 or fixed positive charge.25 As similarly observed at the Al2O3/GaAs interface,25 accumulation of positive charges in the as-grown Al2O3 can produce an interface dipole which might result in the observed increase of the VBO by bending the oxide electronic bands with respect to a flat band condition. From this point of view, annealing Al2O3 restores the flat band condition which descends from suppression of the interface dipole and the decrease of the VBO compared to the as-grown Al2O3/In0.53Ga0.47As interface. It should be noticed that the assumed positive charge is intrinsic of the as-grown Al2O3 (and not other interfacial oxide species) because the VBO decrease is similarly reproduced by taking into account Al 2p as an oxiderelated XPS line instead of the O 1s line (see Table 1). The presence of positive charges in the as-grown Al2O3 can be also supported by the evidence of a process-dependent chemical shift between the Al3+ and the Al* components in the Al 2p line in Figure 4a [Figure 4b] which amount to 2.17 eV [1.67 eV] and 1.82 eV [1.59 eV] for the (2 × 4) reconstruction [(4 × 2) reconstruction] in the as-grown and annealed Al 2 O 3 , respectively. As the chemical shift can be influenced by Coulombic attraction between photoemitted electrons and residing charges, the higher chemical shift measured in the asgrown Al2O3 is consistent with an extra positive charge in the oxide which affects the core level photoemission of Al 2p electrons. Although oxide charging cannot be completely excluded, mainly due to the inability to exploit commonly used references,26 e.g., the C signal, it is noteworthy to underline that a comparative analysis of XPS level energy separation between samples with similar thickness has been considered with the aim to explain the presence of positive charges in the as-grown oxide. No precise determination of the VBO value is here intended because it is beyond the scope of the present work. A second relevant effect can be associated with the above picture which can elucidate the phenomenology of the STS evidence. Besides bending the Al2O3 electronic bands, an interface dipole generated by positive charges in the as-grown Al2O3 can electrically screen the hole accumulation observed in the as-prepared In0.53Ga0.47As surfaces, resulting in the partial FL unpinning reported in Figure 2. The back-shift of the interface FL after annealing can be regarded as a consequence of dipole suppression due to charge relaxation. As no reconstruction-dependent interface composition can be deduced from XPS inspection, the different extent of the backshift between the two In0.53Ga0.47As reconstructions necessarily claims for a mutually different annealing induced bond or structural rearrangement.27 This assumption is consistent with different values of the surface roughness extracted from the annealed Al2O3 grown on (2 × 4) and (4 × 2) reconstructed In0.53Ga0.47As, respectively.



AUTHOR INFORMATION

Corresponding Author

*Phone: +39 039 603 7452 (C.G.); +39 039 603 2884 (A.M.). Fax: +39 039 688 1175 (C.G.); +39 039 688 1175 (A.M.). Email: [email protected] (C.G.); alessandro. [email protected] (A.M.). Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We acknowledge C. Merckling and G. Brammertz (IMEC) for providing In0.53Ga0.47As substrates in the framework of the ARAMIS project and useful discussion with A. C. Kummel (UC San Diego).



REFERENCES

(1) Spicer, W. E.; Lindau, I.; Skeath, P.; Su, C. Y. J. Vac. Sci. Technol. 1980, 17, 1019. (2) Del Alamo, J. Nature 2011, 479, 317. (3) Shen, J.; Winn, D.; Melitz, W.; Clemens, J.; Kummel, A. C. ECS Trans. 2008, 16, 463. (4) Hasegawa, H.; Akazawa, M.; Domanowska, A.; Adamowicz, B. App. Surf. Sci. 2010, 256, 5698. (5) Komsa, H. P.; Pasquarello, A. Appl. Phys. Lett. 2010, 97, 191901. (6) Hale, M. J.; Yi, S. I.; Sexton, J. Z.; Kummel, A. C.; Passlack, M. J. Chem. Phys. 2003, 119, 6719. (7) Clemens, J. B.; Bishop, S. R.; Lee, J. S.; Kummel, A. C.; Droopad, R. J. Chem. Phys. 2010, 132, 244701. (8) Shen, J.; Clemens, J. B.; Chagarov, E. A.; Feldwinn, D. L; Melitz, W.; Song, T.; Bishop, S. R.; Kummel, A. C.; Droopad, R. Surf. Sci. 2010, 604, 1757. (9) Fleischmann, C.; Schouteden, K.; Merckling, C.; Sioncke, S.; Meuris, M.; Van Haesendonck, C.; Temst, K.; Vantomme, A. J. Phys. Chem. C 2012, 116, 9925. (10) Shekhter, P.; Kornblum, L.; Liu, Z.; Cui, S.; Ma, T. P.; Eizenberg, M. Appl. Phys. Lett. 2011, 99, 232103. (11) Chu, L. K.; Merkling, C.; Alian, A.; Dekoster, J.; Kwo, J.; Hong, M.; Caymax, M.; Heyns, M. Appl. Phys. Lett. 2011, 99, 042908. (12) Spicer, W. E.; Chye, P. W.; Garner, C. M.; Lindau, I.; Pianetta, P. Surf. Sci. 1979, 86, 763. (13) Fusi, M.; Lamagna, L.; Spiga, S.; Fanciulli, M.; Brammertz, G.; Merckling, C.; Meuris, M.; Molle, A. Microelectron. Eng. 2011, 88, 435. (14) Molle, A.; Lamagna, L.; Grazianetti, C.; Brammertz, G.; Merckling, C.; Caymax, M.; Spiga, S.; Fanciulli, M. Appl. Phys. Lett. 2011, 99, 193505. (15) Feenstra, R. M.; Stroscio, J. A.; Fein, A. P. Surf. Sci. 1987, 181, 295. (16) Bonnell, D. A. Prog. Surf. Sci. 1998, 57, 187. (17) Huang, M. L.; Chang, Y. C.; Chang, C. H.; Lee, Y. J.; Chang, P.; Kwo, J.; Wu, T. B.; Hong, M. Appl. Phys. Lett. 2005, 87, 252104. (18) Kirk, A. P.; Milojevic, M.; Kim, J.; Wallace, R. M. Appl. Phys. Lett. 2010, 96, 202905. (19) Hinkle, C. L.; Milojevic, M.; Brennan, B.; Sonnet, A. M.; Aguirre-Tostado, F. S.; Hughes, G. J.; Vogel, E. M.; Wallace, R. M. Appl. Phys. Lett. 2009, 94, 162101. (20) Lamagna, L.; Fusi, M.; Spiga, S.; Fanciulli, M.; Brammertz, G.; Merckling, C.; Meuris, M.; Molle, A. Microelectron. Eng. 2011, 88, 431.



CONCLUSIONS It has been demonstrated that Al2O3 can partially unpin the In0.53Ga0.47As (2 × 4) and (4 × 2) reconstructions and that PDA results in a reconstruction-dependent behavior which could be ascribed to a different rearrangement of the chemical bonds at the very interface. This behavior is rationalized in terms of an interface dipole induced by positive charges in the as-grown oxide which are suppressed upon annealing as results from in situ XPS investigation of the change in the electron band line up at the interface between the oxide and the semiconductor. From STM/STS and XPS data, when capped 18750

dx.doi.org/10.1021/jp3042318 | J. Phys. Chem. C 2012, 116, 18746−18751

The Journal of Physical Chemistry C

Article

(21) Brennan, B.; Milojevic, M.; Contreras-Guerrero, R.; Kim, H. C.; Lopez-Lopez, M.; Kim, J.; Wallace, R. M. J. Vac. Sci. Technol. B 2012, 30, 04E104. (22) Kraut, E.; Grant, R.; Waldrop, J.; Kowalczyk, S. Phys. Rev. Lett. 1980, 44, 1620. (23) Perego, M.; Molle, A.; Fanciulli, M. Appl. Phys. Lett. 2008, 92, 042106. (24) Shin, B.; Weber, J. R.; Long, R. D.; Hurley, P. K.; Van de Walle, C. G.; McIntyre, P. C. Appl. Phys. Lett. 2010, 96, 152908. (25) Hu, J.; Nainani, A.; Sun, Y.; Saraswat, K. C.; Wong, H. S. P. Appl. Phys. Lett. 2011, 99, 252104. (26) Perego, M.; Seguini, G. J. Appl. Phys. 2011, 110, 053711. (27) Shen, J.; Chagarov, E.; Feldwinn, D.; Melitz, W.; Santagata, N. M.; Kummel, A. C.; Droopad, R.; Passlack, M. J. Chem. Phys. 2010, 133, 164704.

18751

dx.doi.org/10.1021/jp3042318 | J. Phys. Chem. C 2012, 116, 18746−18751