an explanation of efficiency droop in ingan-based light emitting diodes: saturated radiative...

6
Journal of the Korean Physical Society, Vol. 58, No. 3, March 2011, pp. 503508 An Explanation of Efficiency Droop in InGaN-based Light Emitting Diodes: Saturated Radiative Recombination Rate at Randomly Distributed In-Rich Active Areas Jong-In Shim * and Hyunsung Kim Department of Electronics & Communication Engineering, Hanyang University, Ansan 426-791, Korea Dong-Soo Shin Department of Applied Physics, Hanyang University, Ansan 426-791, Korea Han-Youl Yoo Department of Physics, Inha University, Incheon 402-751, Korea (Received 12 January 2011) We present a comprehensive model of the dependence of the internal quantum efficiency (IQE) on both the temperature and the carrier density in InGaN-based blue and green light emitting diodes (LEDs). In our model, carriers are dominantly located and recombine both radiatively and nonradiatively inside randomly distributed In-rich areas of the InGaN quantum wells (QWs). In those areas, the carrier density is very high even at a small current density. We propose that the saturated radiative recombination rate is a primary factor determining the IQE droop of InGaN- based LEDs. In typical InGaN-based QWs, it is common for the total carrier recombination rate to be smaller than the carrier injection rate even at a small current density. This is mostly attributable to the saturation of the radiative recombination rate. The saturation of the radiative recombination rate increases carrier density in InGaN QWs, enlarges nonradiative carrier losses, and eventually gives rise to the large IQE droop with increasing current. We show how the radiative recombination rate saturates and the radiative recombination rate has influence on the IQE droop in InGaN-based QW LEDs. PACS numbers: 78.20.Bh, 78.30.Fs, 78.67.De Keywords: Internal quantum efficiency, Efficiency droop, Light emitting diode, Radiative recombination rate, In-rich area DOI: 10.3938/jkps.58.503 I. INTRODUCTION InGaN-based light emitting diodes (LEDs) operating especially at blue or green wavelengths suffer from a large reduction in the internal quantum efficiency (IQE) with increasing current, the so-called efficiency (or IQE) droop. Understanding and reducing the efficiency droop is an imperative need for the future of general light- ing because the droop is currently a limiting factor to more efficient LEDs under high current. The follow- ing experimental behaviors are commonly observed: The IQE droop is observed (1) in both blue and green LEDs [1], (2) in both resonant photoluminescence (PL) and electroluminescence (EL) measurements with very simi- lar excitation dependences [2], (3) under pulsed, as well as continuous-wave (CW), current injections [3,4], (4) * E-mail: [email protected]; Fax: +82-31-400-5179 in polar, non-polar, and semi-polar quantum-well (QW) structures [5–7], (5) for single QW, multiple QW, and bulk active layers [2, 5, 8], and (6) at all temperatures from 4 to 450 K [1–5]. Various mechanisms focusing on additional increases in the nonradiative recombination rates have been pro- posed for the efficiency droop. They include local junc- tion heating [3], carrier overflow enhanced by internal polarization field [4] or poor hole injection [9], carrier delocalization from In-rich regions and nonradiative re- combination at high defect sites [10,11], and Auger re- combination [1,2,12]. However, the mechanisms proposed so far can explain the aforementioned IQE droop behav- iors only partially. For example, at lower temperatures or at higher In-contents of InGaN active layer, the IQE droop is more pronounced; i.e., the IQE peak appears at a smaller current, and the IQE decreases more drastically towards higher current. These experimental observations -503-

Upload: yoo

Post on 08-Dec-2016

213 views

Category:

Documents


1 download

TRANSCRIPT

Journal of the Korean Physical Society, Vol. 58, No. 3, March 2011, pp. 503∼508

An Explanation of Efficiency Droop in InGaN-based Light Emitting Diodes:Saturated Radiative Recombination Rate at Randomly Distributed In-Rich

Active Areas

Jong-In Shim∗ and Hyunsung Kim

Department of Electronics & Communication Engineering, Hanyang University, Ansan 426-791, Korea

Dong-Soo Shin

Department of Applied Physics, Hanyang University, Ansan 426-791, Korea

Han-Youl Yoo

Department of Physics, Inha University, Incheon 402-751, Korea

(Received 12 January 2011)

We present a comprehensive model of the dependence of the internal quantum efficiency (IQE)on both the temperature and the carrier density in InGaN-based blue and green light emittingdiodes (LEDs). In our model, carriers are dominantly located and recombine both radiatively andnonradiatively inside randomly distributed In-rich areas of the InGaN quantum wells (QWs). Inthose areas, the carrier density is very high even at a small current density. We propose that thesaturated radiative recombination rate is a primary factor determining the IQE droop of InGaN-based LEDs. In typical InGaN-based QWs, it is common for the total carrier recombination rate tobe smaller than the carrier injection rate even at a small current density. This is mostly attributableto the saturation of the radiative recombination rate. The saturation of the radiative recombinationrate increases carrier density in InGaN QWs, enlarges nonradiative carrier losses, and eventuallygives rise to the large IQE droop with increasing current. We show how the radiative recombinationrate saturates and the radiative recombination rate has influence on the IQE droop in InGaN-basedQW LEDs.

PACS numbers: 78.20.Bh, 78.30.Fs, 78.67.DeKeywords: Internal quantum efficiency, Efficiency droop, Light emitting diode, Radiative recombinationrate, In-rich areaDOI: 10.3938/jkps.58.503

I. INTRODUCTION

InGaN-based light emitting diodes (LEDs) operatingespecially at blue or green wavelengths suffer from alarge reduction in the internal quantum efficiency (IQE)with increasing current, the so-called efficiency (or IQE)droop. Understanding and reducing the efficiency droopis an imperative need for the future of general light-ing because the droop is currently a limiting factor tomore efficient LEDs under high current. The follow-ing experimental behaviors are commonly observed: TheIQE droop is observed (1) in both blue and green LEDs[1], (2) in both resonant photoluminescence (PL) andelectroluminescence (EL) measurements with very simi-lar excitation dependences [2], (3) under pulsed, as wellas continuous-wave (CW), current injections [3, 4], (4)

∗E-mail: [email protected]; Fax: +82-31-400-5179

in polar, non-polar, and semi-polar quantum-well (QW)structures [5–7], (5) for single QW, multiple QW, andbulk active layers [2, 5, 8], and (6) at all temperaturesfrom 4 to 450 K [1–5].

Various mechanisms focusing on additional increasesin the nonradiative recombination rates have been pro-posed for the efficiency droop. They include local junc-tion heating [3], carrier overflow enhanced by internalpolarization field [4] or poor hole injection [9], carrierdelocalization from In-rich regions and nonradiative re-combination at high defect sites [10,11], and Auger re-combination [1,2,12]. However, the mechanisms proposedso far can explain the aforementioned IQE droop behav-iors only partially. For example, at lower temperaturesor at higher In-contents of InGaN active layer, the IQEdroop is more pronounced; i.e., the IQE peak appears ata smaller current, and the IQE decreases more drasticallytowards higher current. These experimental observations

-503-

-504- Journal of the Korean Physical Society, Vol. 58, No. 3, March 2011

cannot be understood easily because the proposed non-radiative recombination processes should decrease theo-retically with decreasing temperature and the materialconstants of InGaN active layers are not so sensitive toIn-compositions from blue to green wavelengths.

Recent experimental evidence suggests the existence ofIn-rich areas in the active layer of the InGaN-based blueLED, which can act as localization centers of carriersand behave more like quantum disks (Q-disks) or quan-tum dots (Q-dots) with a reduced density of states [13,14]. The existence of In-rich areas implies that the effec-tive active volume is smaller than the nominal volume ofthe QWs. In these In-rich areas, the actual carrier den-sity would be much higher than previously anticipated.We believe that this high carrier density in the In-richareas holds a key to understanding the efficiency droopphenomenon.

In this paper, to understand the efficiency droop, wefocus on the radiative recombination and how it sat-urates, rather than on the nonradiative recombinationprocesses as previously attempted. Starting from theimbalance of electrons and holes in k-space, combinedwith a reduced effective active volume, we show how sat-uration of radiative recombination happens. We presenta comprehensive model that can explain quantitativelythe IQE’s dependences both on temperature and current.

II. THEORETICAL MODEL

We consider the physical situation of an active QWin which the excess electron and hole concentrationsare equal and they are much larger compared to thoseat thermal equilibrium. The carrier injection efficiencywhich is defined as the fraction of carriers into the activeQW is assumed as one so that carrier leakages throughsurface or clad layers are neglected. Then, the carrierrate equation at steady state and the IQE can be writ-ten as

I = qVeff[A(T )N + B(T,N)N2], (1)

IQE =B(T,N)N2

A(T )N + B(T,N)N2, (2)

where A and B are the Shockley-Read-Hall (SRH) andthe radiative recombination coefficients, respectively. N ,I, and q are the carrier density, the current, and theelectron charge. Veff is the effective active volume, whichis different from the nominal active volume Va , as willbe discussed later.

In the well-known “carrier localization” model, carri-ers recombine radiatively in the In-rich areas and non-radiatively in defect areas like dislocations. Thus, theradiative and the nonradiative recombination processesare assumed to occur in spatially different areas. In ourpicture, however, most of the carriers recombine bothradiatively and nonradiatively inside Q-disk shaped In-rich areas. Thus, the nonradiative recombination rate issupposed to be much smaller at threading dislocationsthan at In-rich areas. This carrier recombination modelcan successfully explain the current-voltage (I-V) curvesof modern advanced LEDs. The carrier transport mech-anism in the forward current range of <0.1 A/cm2 isknown to be the tunneling-recombination with an ide-ality factor of >5 in the I-V curves [15]. The carriertunneling-recombination process happens through traplevels associated with electrically active dislocations likeV-defects. Since these tunneling shunts happen in a verysmall area and with high resistance, the tunneling cur-rent and the long-wavelength emission are rapidly satu-rated as current increases. At higher operating currents(>5 A/cm2), the electroluminescence emission from therandomly distributed In-rich areas becomes uniform overan entire chip. The ideality factors of the I-V curve inthis range are between 1 and 2 [16], which means that thenonradiative recombination process inside In-rich areascan be described by using the SRH recombination pro-cess.

Our goal is to find a phenomenological expression forB(T,N) so that the radiative carrier recombination rateper unit volume, Rr, can be simply described as BN2.Concerning the radiative recombination, only states withidentical k-vectors in the conduction band (kc) and thevalence band (kv) can form a transition pair satisfyingthe “k-selection rule”. In the free-carrier model [17], Rr

of an ideal 2-dimensional QW is given by

Rr = A◦|M |2∫ ∞

0

∫ ∞

0

fc(kc)fν(kν)kckνdkcdkνδ(kc − kν) = B(T,N)N2, (3)

where Ao is a constant. |M |2 represents an average ma-trix element, which is a function of temperature T . fc

and fv are the Fermi-Dirac distributions for electrons and

holes, respectively. The delta function δ(kc - kν) is zerounless kc = kν . We define a quantity Φcv(T,N) as

Φcv =∫ ∞

0

∫ ∞

0

fc(kc)fν(kν)kckνdkcdkνδ(kc − kν)/N2 (4)

An Explanation of Efficiency Droop in InGaN-based Light Emitting Diodes: · · · – Jong-In Shim et al. -505-

The radiative recombination coefficient B of an ideal 2-dimensional QW is obtained from Eqs. (3) and (4) as

B(T,N) = Bo(T )Φcν(T,N), whereBo(T ) = Ao|M(T )|2. (5)

Since In-rich areas that can act as the localization centerof carriers exist, the effective active volume, Veff, wherecarriers efficiently recombine, is smaller than the nomi-nal active volume, Va, of epitaxially grown InGaN layers.As Veff decreases, the electronic density of states in theconduction and the valence bands decreases, the carrierdensity at a certain current density increases. Moreover,the electron and the hole distributions in the electronwavevector k-space deviate much from each other due tothe different effective masses and occupation probabili-ties so that the average radiative recombination proba-bility of an electron-hole pair satisfying the k-selectionrule becomes less. Additionally, in highly excited media,the carrier scattering and the carrier screening becomeviolent, which eventually change the radiative recombi-nation probability and its spectrum from the free-carriermodel [18–20]. Because we used the free-carrier modelin evaluating Eq. (4) and the effective active volume isreduced to more like Q-disks or Q-dots of In-rich areasas discussed previously, we modify Eq. (5) of an ideal2-dimensional QW as

B(T,N) =Bo(T )Φcν(T,N)[1 + N/No(T )]n

. (6)

The additional saturation factor 1/(1 + N/No)n is toaccount for the reduced recombination probability of aQ-disk-shaped active In-rich area from that of an ide-ally uniform QW. Here, Bo, No, and the power n canbe experimentally determined. No is the carrier densityrepresenting the saturation of the B coefficient.

III. RESULTS AND DISCUSSION

Figure 1(a) shows an example of electron and hole dis-tributions in an ideal 2-dimensional QW as a function ofthe wavevector k for two different temperatures of 4 and300 K. The different distributions of electrons and holesin k-space mainly result from the different electron den-sity of states in the conduction and the valence bands. Inthis example, simple parabolic energy dispersion curvesare assumed and the effective masses in the conductionand valence bands are 0.17 mo and 0.85 mo, respectively,where mo is the electron rest mass. At a low tempera-ture of 4 K, electrons and holes are distributed in almostsame k-space so that the radiative recombination prob-ability is expected to be high. At a high temperatureof 300 K, however, electrons and holes spread out dif-ferently in k-space, which reduces the radiative recom-bination probability compared with the low-temperaturecase. The radiative recombination coefficient B(T,N) in

Fig. 1. (Color online) Electron and hole distributions alongthe electron wavevector k for different temperatures and (b)calculation result of Φcv in Eq. (4).

Eq. (6) is directly proportional to Φcv(T,N) in Eq. (4).Figure 1(b) shows Φcv(T,N) as a function of N for differ-ent temperatures. As temperature increases, the onsetof Φcv appears at higher N and its peak value decreases.Roughly speaking, Φcv(T,N) shows a similar to that ofthe IQE curves, as will be explained in Fig. 4.

With the reduced effective volume of randomly dis-tributed In-rich areas, the available electronic densityof states decreases, making the electronic density ofstates of In-rich areas smaller than that of an ideal 2-dimensional QW. In reality, it is difficult to accuratelycalculate the density of states of In-rich areas due to theirrandom shapes and sizes. Here, we show the effects of thereduced density of states on the radiative recombinationrate in Q-disk-like In-rich areas by considering an ideal2-dimensional QW, but with much reduced electron andhole effective masses. Since the 2-dimensional densityof states is proportional to the effective mass, reducingthe effective masses is equivalent to reducing the densityof states. Figures 2(a) and (b) show Φcv(T,N) for anideal 2-dimensional QW as a function of N for differenteffective masses in the conduction band at T = 4 and300 K, respectively. The effective masses in the valenceband are assumed to be five times larger than those of

-506- Journal of the Korean Physical Society, Vol. 58, No. 3, March 2011

Fig. 2. (Color online) Calculated results of Φcv(T, N) inEq. (4) for different effective masses in the conduction bandat (a) T = 4 K and (b) T = 300 K. The effective masses inthe valance band are assumed to be five times larger thanthose in the conduction band.

the conduction band. As the effective mass decreases,the onset of Φcv appears at smaller carrier density, andΦcv decreases at smaller carrier density. At a reducedtemperature, the same behaviors are observed, but withincreased peak values for Φcv. These observations indi-cate that the radiative recombination coefficient B(T , N)for a small effective active volume like In-rich areas canbe significantly decreased at very small carrier density,as can be seen in Eq. (6).

Figures 3(a) and (b) show the B coefficient in Eq. (6)and the radiative recombination rate BN2 normalized totheir peak values for different temperatures, respectively.The power n and the No in Eq. (6) are chosen so thatthe calculated IQE results can explain the experimentalIQE curves of the blue InGaN/GaN multiple QW LEDin Ref. 1. n is obtained to be 0.94 for all temperatures.No becomes smaller as the temperature decreases. Ascan be seen in Eq. (5), B(T,N) has a shape quite sim-ilar to that of Φcv(T,N) before the denominator of Eq.(6) plays a role. At high carrier density, the decrease in

Fig. 3. (Color online) (a) Radiative recombination coeffi-cient B and (b) radiative recombination rate BN2 as func-tions of carrier density for different temperatures. The pa-rameters are normalized to their peak values. The depen-dences on carrier density N are additionally plotted with thedashed lines.

B(T,N) is initiated by the factor 1/[1 + N/No(T )]n andis determined by both Φcv(T,N) and 1/[1 + N/No(T )]n.Thus, the radiative recombination rate BN2 shows sat-urated characteristics, as can be seen in Fig. 3(b). Thedashed lines in the figures guide the power dependenceson the carrier density. It is noted that as the carrierdensity increases, the B coefficient decreases, with theeventual power dependence being between 1 and 2 [Fig.3(a)]. This results in the radiative recombination rateBN2 saturating with a power dependence of Np, where0 < p < 1. This power dependence causes the droop toshow up with the saturated B only, without the nonra-diative process depending on N3.

Figure 4 compares the experimental IQEs in Ref. 1with the calculated ones based on the proposed model.They show good agreements for all temperatures. Twomeaningful results are found in this IQE fitting proce-dure. One is that the carrier density at the peak IQE(Np) is estimated to be anomalously high even though

An Explanation of Efficiency Droop in InGaN-based Light Emitting Diodes: · · · – Jong-In Shim et al. -507-

Fig. 4. (Color online) Experimental IQEs (symbols) andtheir fitted curves (solid lines) as functions of current for dif-ferent temperatures.

the peak current density (Jp) is very low. In Fig. 3, thevalues of (Jp, Np) in units of (A/cm2, cm−3) are esti-mated as (0.26, 1.5 × 1018), (2.12, 2.1 × 1019), (4.21,3.3 × 1019), and (5.9, 4.6 × 1019) for 4, 100, 200, and300 K, respectively. The other is that the effective ac-tive volume, Veff in Eq. (1), is quite different from thenominal active volume, Va. Va of the device is 290 µm× 290 µm × 2 nm. Veff can be obtained from Eq. (1)if either the A or the B coefficients is known. Recently,reported A and B coefficient values at room tempera-ture in InGaN QWs are in the ranges of 1 × 107 - 2 ×108 s−1 and 1 × 10−10 - 3 × 10−11 cm3/s, respectively[8, 21]. When an A coefficient of 7 × 107 s−1 is chosen,which is at the center of the reported values, the B coeffi-cient at the peak IQE is estimated as 1.3 × 10−10 cm3/s.From these assumptions, Veff is calculated to be approxi-mately one hundred times smaller than Va in order to fitthe IQE curves as a function of current. This indicatesthat only a part of the physical InGaN QW volume ef-fectively works as carrier recombination centers, where isconsistent with the experimental observation of In-richareas in the active layer.

The lateral sizes of the In-rich areas are reported tovary widely from a few nm to hundreds of nm [13,14],and the total volume of these Q-disks (or Q-dots) is muchsmaller than the physical volume of a QW, making Veff �Va. In fact, the IQE behavior’s dependences on tempera-ture and current for InGaN QW-based blue/green LEDsare quite similar to those of GaAs-based LEDs with ar-tificially grown active Q-dots, which also supports theclaim that the reduced active volume in the LEDs im-pacts very heavily on the IQE droop [22,23]. In Q-disk(or Q-dot) shaped In-rich areas, the carrier density be-comes very high at small current due to its small physicalvolume and electronic density of states [24]. The radia-tive recombination rate in those areas rapidly saturatesat small current injection due partly to the mismatch of

electron and hole distributions in k-space and partly tothe effects caused by high carrier density. On the otherhand, the SRH nonradiative recombination rate mono-tonically increases with increasing current.

Up to now, we have presented an IQE droop model inwhich carrier leakage from the QW to the clad layer isnegligible. In typical InGaN-based QWs, it is commonfor the total carrier recombination rate to be smaller thanthe carrier injection rate. This is mostly attributable tothe saturation of the radiative recombination rate at asmall current density. Thus, the limited radiative recom-bination rate forces carriers to pile-up rapidly in the QW,makes it easy for carriers to spill-over to the quantumbarrier, and eventually increases the IQE droop. Theproposed model strongly supports the idea that the ra-diative recombination rate is the primary factor to beconsidered for the reduction of the IQE droop in InGaN-based QW LEDs.

IV. CONCLUSIONS

We have investigated the dependence of the IQE be-havior on both the temperature T and the carrier densityN in blue and green InGaN/GaN multiple QW LEDs byformulating the radiative recombination coefficient B(T ,N). It is theoretically shown that the radiative recom-bination coefficient B(T , N) for a small effective activevolume like In-rich areas can be significantly decreasedat very small carrier density and low current level. Wepropose that the IQE droop originates from the limitedcarrier recombination rates, i.e., the saturated radiativerecombination rate and the monotonically increased non-radiative recombination rate, inside a Q-disk shaped In-rich active area with increasing current. The saturationof the radiative recombination rate increases carrier den-sity in InGaN QWs, enlarges nonradiative carrier losses,and eventually gives rise to the large IQE droop withincreasing current. We believe that our model presentsa comprehensive understanding of the dependence of theIQE on both the temperature and the carrier densityin InGaN-based LEDs and sheds light on how one canfurther improve the performance of blue or green LEDs.

ACKNOWLEDGMENTS

This work was supported by the Strategic Technol-ogy Development Project of the Ministry of KnowledgeEconomy (2010-0015297), Republic of Korea.

REFERENCES

[1] M. Peter, A. Laubsch, W. Bergbauer, T. Meyer, M.Sabathil, J. Baur and B. Hahn, Phys. Status Solidi A6, 1 (2009).

-508- Journal of the Korean Physical Society, Vol. 58, No. 3, March 2011

[2] A. Laubsch et al., Phys. Status Solidi C 6, S913 (2009).[3] A. A. Efremov, N. I. Bochkareva, R. I. Gorbunov, D. A.

Lavrinovich, Yu. T. Rebane, D. V. Taekhin and Yu. G.Shreter, Semiconductors 40, 605 (2006).

[4] M. H. Kim, M. F. Schubert, Q. Dai, J. K. Kim, E. F.Schubert, J. Piprek and Y. Park, Appl. Phys. Lett. 91,183507 (2007).

[5] K. Fujiwara, H. Jimi and K. Kaneda, Phys. Status SolidiC 6, S814 (2009).

[6] X. Li, J. Lee, M. Wu, U. Ozgur, H. Morkoc, T. Paskova,G. Mulholland and K. R. Evans, Appl. Phys. Lett. 95,121107 (2009).

[7] Y.-D. Lin, A. Chakraborty, S. Brinkley, H, C. Kuo, T.Melo, K. Fujito, J. S. Speck, S. P. DenBaars and S. Naka-mura, Appl. Phys. Lett. 94, 261108 (2009).

[8] N. F. Gardner, G. O. Muller, Y. C. Shen, G. Chen, S.Watanabe, W. Gotz and M. R. Krames, Appl. Phys.Lett. 91, 243506 (2007).

[9] X. Ni, Q. Fan, R. Shimada, U. Ozgur and H. Morkoc,Appl. Phys. Lett. 93, 071113 (2008).

[10] B. Monemar and B. E. Sernelius, Appl. Phys. Lett. 91,181103 (2007).

[11] J. Harder, J. V. Moloney and S. W. Koch, Appl. Phys.Lett. 96, 261106 (2010).

[12] J. Harder, J. V. Moloney, B. Pasenow, S. W. Koch, M.Sabathil, N. Linder and S. Lutgen, Appl. Phys. Lett. 92,261103 (2008).

[13] G. H. Gu, C. G. Park and K. B. Nam, Phys. Status SolidiRRL 3, 100 (2009).

[14] K. Okamoto, A. Scherer and Y. Kawakami, Appl. Phys.Lett. 87, 161104 (2005).

[15] J. M. Shah, Y.-L. Li, Th. Gessman and E. F. Schubert,J. Appl. Phys. Lett. 94, 2627 (2003).

[16] S. W. Lee et al., Appl. Phys. Lett. 89, 132117 (2006).[17] G. P. Agrawal and N. K. Dutta, Semiconductor Lasers,

2nd ed. (Van Norstrand Reinhold, New York, 1993).[18] W. W. Chou, S. W. Koch and M. Sargent III,

Semiconductor-Laser Physics (Spring-Verlag, BerlinHeidelberg, 1994).

[19] J. Hader, J. V. Moloney and S. W. Koch, Appl. Phys.Lett. 87, 201112 (2005).

[20] G. Bourdon, I. Robert, I. Sagnes and I. Abram, J. Appl.Phys. 92, 6595 (2002).

[21] H.-Y. Ryu, H.-S. Kim and J.-I. Shim, Appl. Phys. Lett.95, 081114 (2009).

[22] D. L. Huffaker and D. G. Deppe, Appl. Phys. Lett. 73,520 (1998).

[23] A. J. Bennett, P. N. Stavrinou, C. Roberts, R. Murray,G. Parry and J. S. Roberts, J. Appl. Phys. 92, 6215(2002).

[24] I. A. Pope, P. M. Smowton, J. D. Thomson, M. J. Kap-pers and C. J. Humphreys, Appl. Phys. Lett. 82, 2755(2003).