Hostname: page-component-7c8c6479df-24hb2 Total loading time: 0 Render date: 2024-03-28T08:13:04.483Z Has data issue: false hasContentIssue false

Dry and Wet Etching for Group III – Nitrides

Published online by Cambridge University Press:  13 June 2014

I. Adesida
Affiliation:
Department of Electrical and Computer Engineering, University of Illinois, Urbana-Champaign, IL 61801
C. Youtsey
Affiliation:
Department of Electrical and Computer Engineering, University of Illinois, Urbana-Champaign, IL 61801
A. T. Ping
Affiliation:
Department of Electrical and Computer Engineering, University of Illinois, Urbana-Champaign, IL 61801
F. Khan
Affiliation:
Department of Electrical and Computer Engineering, University of Illinois, Urbana-Champaign, IL 61801
L. T. Romano
Affiliation:
Xerox PARC, Palo Alto, CA 94304
G. Bulman*
Affiliation:
CREE Research, Inc.,Durham, NC 27713

Abstract

The group-III nitrides have become versatile semiconductors for short wavelength emitters, high temperature microwave transistors, photodetectors, and field emission tips. The processing of these materials is significant due to the unusually high bond energies that they possess. The dry and wet etching methods developed for these materials over the last few years are reviewed. High etch rates and highly anisotropic profiles obtained by inductively-coupled-plasma reactive ion etching are presented. Photoenhanced wet etching provides an alternative path to obtaining high etch rates without ion-induced damage. This method is shown to be suitable for device fabrication as well as for the estimation of dislocation densities in n-GaN. This has the potential of developing into a method for rapid evaluation of materials.

Type
Research Article
Copyright
Copyright © 1999 Materials Research Society

Introduction

The success of the synthesis and growth of the wide bandgap group-III nitrides over the last decade has made the realization of a wide range of new devices possible. The bandgap energies of the III-nitrides range from 1.9 eV for InN to 3.4 eV for GaN to 6.2 eV for AlN. Using these materials, bright light emitting diodes (LEDs) and laser diodes (LDs) [Reference Nakamura, Mukai and Senoh1,Reference Nakamura, Senoh, Nagahama, Iwasa, Yamada, Matsushita, Kiyoku and Sugimoto2] operating at short wavelengths have been demonstrated. Indeed, LDs with InGaN/AlGaN active layers having lifetimes greater than 10,000 hours have been demonstrated making the commercialization of these devices a certainty. The excellent electron transport characteristics of GaN coupled with the wide bandgap, the chemical stability, and the availability of AlGaN/GaN heterostructures also make III-nitrides suitable for high power, high temperatures transistors. AlGaN/GaN heterostructure field effect transistors (HFETs) on sapphire operating at frequencies greater than 70 GHz have been demonstrated [Reference Ping, Adesida, Boutros and Redwing3] and similar HFETs grown on SiC exhibiting power densities as high as 6.8 W/mm have also been fabricated [Reference Sheppard, Doverspike, Pribble, Allen, Palmour, Kehias and Jenkins4]. Most recently, AlGaN/GaN heterojunction bipolar transistors have been demonstrated [Reference McCarthy, Kozodoy, Rodwell, DenBaars and Mishra5].

Improvements in the performance of these devices depend on the quality of epitaxial materials and the development of device processing technologies. In particular, effective etching techniques are essential for forming facets for GaN LDs, defining mesas for photodetectors, and gate recessing for HFETs. Group-III nitrides have high bond energies compared to conventional III-V semiconductors. The bond energies are 7.7 eV/atom for InN, 8.9 eV/atom for GaN, and 11.5 eV/atom for AlN compared to 6.5 eV/atom for GaAs. The high bond strengths and wide bandgaps make them essentially chemically inert and highly resistant to bases and acids at room temperature. Therefore a wide range of dry and wet etching techniques have been investigated for the processing of III-nitrides. Since bond strengths are high for the III-nitrides, external energy is required to initiate and sustain the dissociation of the bonds. Sources of external energy include energetic ions, energetic electrons, and optical radiation for different etching methods. In this paper, we present these etching methods and discuss the progress made in applying some of them to III-nitrides.

Dry Etching

Various methods of dry etching involving ion-assisted mechanisms have been applied to the processing of III-nitrides. They include ion milling [Reference Adesida, Ping, Youtsey, Dow, Khan, Olson and Kuznia6,Reference Pearton, Abernathy, Ren and Lothian7], chemically assisted ion beam etching (CAIBE) [Reference Ping, Adesida and Khan8,Reference Ping, Youtsey, Adesida, Khan and Kuznia9,Reference Ping, Khan and Adesida55], reactive ion beam etching (RIBE) [Reference Lee, Park, Park, Yoo, Kim, Kim and Yeom10], reactive ion etching (RIE) [Reference Basak, Verdú, Montojo, Sánchez-García, Sánchez, Munoz and Calleja11-Reference Lin, Fan, Ma, Allen and Morkoς15], electron-cyclotron resonance reactive ion etching (ECR-RIE) [Reference Pearton, Abernathy, Ren, Lothian, Wisk and Katz16-Reference Vartuli, Pearton, Lee, Hong, MacKenzie, Abernathy and Shul21], and inductively-coupled-plasma reactive ion etching (ICP-RIE) [Reference Cho, Vartuli, Donovan, Mackenzie, Abernathy, Pearton, Shul and Constantine22-Reference Smith, Wolden, Bremser, Hanser, Davis and Lampert27]. Optical excitation sources with photon energies higher than the bandgap energies of the semiconductors have been applied to both dry and wet etching methods. The photoassisted dry etching method [Reference Leonard and Bedair28] involves optical radiation of the sample in the presence of reactive gases. Low energy electron-enhanced etching (LE4) [Reference Gillis, Choutov, Martin, Bremser and Davis29] is a dry etching method where electrons with energies < 15 eV are the source of external energy.

Ion milling rely on physical sputtering to achieve etching but this method is not practical for nitrides because of low etch rates and high ion-induced damage [Reference Adesida, Ping, Youtsey, Dow, Khan, Olson and Kuznia6,Reference Pearton, Abernathy, Ren and Lothian7]. Therefore, methods of dry etching involving chemical mechanisms in addition to physical sputtering are the most effective for device applications. Of these techniques, the CAIBE, RIE, ECR-RIE, and ICP-RIE have been the most widely investigated. Tools for ECR-RIE and ICP-RIE are high-density-plasma systems which use magnetic confinement of electrons to generate very high ion densities (> 5 × 1011 cm−3). Although, the methods for coupling power to the plasma in these systems are different, the plasmas have similar properties. It should also be noted that in these systems the rf power generators for controlling the ion flux and for fixing the ion energy are different. This de-coupling allows for the delivery of large ion fluxes at low energies (or biases) onto samples. This enhances etch rates and prevents excessive lattice damage in comparison to conventional RIE.

Etch Rates and Profiles

The chemistries for the dry etching of III-nitrides are mostly halogen-based with the most prevalent being chlorine-based. A summary of some of these chemistries along with the etching methods and accompanying etch rates are presented in Table I. It should be noted that a direct comparison of the etch rates in Table I cannot be made since material quality and etching apparatus can differ significantly.

The plasma chemistries in Table I are identical to those utilized for conventional compound semiconductors. It should be expected that etch products for the group-III elements should therefore be similar in both cases. For nitrides etched in Cl-based gases, the etch-products are GaClx, InClx, and AlClx for the Group III elements, while for nitrogen, it could be NCl3 or perhaps free N2. The volatility of these products is aided by ion bombardment. However, with the boiling point of InCl3 being high at ∼ 600 °C, other gas mixtures involving CH4 have been investigated. The potential products for these mixtures are methyl- or ethyl-based metal-organics for the metals along with NH3 for nitrogen. The boiling point of (CH3)3Ga, (CH3)3In, and (CH3)3Al are < 150 °C, and are therefore, more readily volatile. Notwithstanding the similarities in terms of etch chemistries, we note that the high bond energies of the nitrides degrade their etch rates in comparison to those of other compound semiconductors.

The etch rates reported for GaN using RIE with various etch chemistries range from 17 to 100 nm/min [Reference Basak, Verdú, Montojo, Sánchez-García, Sánchez, Munoz and Calleja11-Reference Lin, Fan, Ma, Allen and Morkoς15]. Etch rates were found to depend strongly on the plasma self-bias voltage, and essentially independent of the chamber pressure for pressures less than 80 mTorr [Reference Adesida, Mahajan, Andideh, Khan, Olson and Kuznia Appl13]. The higher etch rates were obtained at high plasma dc biases from −300 to −400 V. Anisotropic etch profiles were obtained in all cases but they were overcut which meant that physical mechanisms dominated the etching. In conventional RIE, physical and chemical components of etching cannot be independently controlled. This impacts the shape of etch profiles significantly especially in the case of III-nitrides where high ion energy is required to break the bonds.

Table I. Summary of etch rates for various dry etching methods.

Gas Chemistry Etching Technique Etch Rate (nm/min)
GaN AlN InN
Ar Ion Milling 110 500 eV6 29 500 eV7 61 500 eV7
HCl [Ar ion] CAIBE 190 500 eV8 - - - - - - - - - -
Cl2 [Ar ion] CAIBE 210 500 eV9 62 500 eV55 - - - - -
HCl RIBE 130 500 eV10 - - - - - - - - - -
Cl2 RIBE 150 500 eV10 - - - - - - - - - -
SF6 RIE 17 −400 V11 - - - - - - - - - -
CHF3, C2ClF5 RIE 45 500 W12 - - - - - - - - - -
SiCl4[w/Ar, SiF4] RIE 55 −400 V13 - - - - - - - - - -
HBr [w/Ar, H2] RIE 60 −400 V14 - - - - - - - - - -
BCl3 RIE 105 −230 V15 - - - - - - - - - -
CH4/H2/Ar ECR-RIE 9 −250 V16 2.5 −300 V16 10 −300 V16
CCl2F2/Ar ECR-RIE 20 −250 V16 18 −300 V16 18 −300 V16
BCl3/Ar ECR-RIE 30 −250 V16 17 −250 V16 17 −300 V16
HBr/H2 ECR-RIE 70 −150 V17 65 −150 V17 17 −150 V17
SiCl4/Ar ECR-RIE 75 −280 V18 - - - - - - - - - -
HI/H2 ECR-RIE 110 −150 V17 120 −150 V17 100 −150 V17
Cl2/H2/Ar ECR-RIE 200 −180 V19 110 −150 V19 150 −180 V19
IBr/Ar ECR-RIE 300 −170 V20 160 −170 V20 325 −170 V20
ICl/Ar ECR-RIE 1300 −275 V21 200 −275 V21 1150 −275 V21
Cl2/N2 ICP-RIE 65 −100 V22 39 −100 V22 30 −100 V22
Cl2/SF6 ICP-RIE 130 −250 V23 184 −250 V23 46 −250 V23
CH4/H2 /Ar ICP-RIE 140 −620 V24 30 −100 V24 110 −480 V24
Cl2/Ar/H2 ICP-RIE 688 −280 V25 - - - - - - - - - -
BCl3/Cl2 ICP-RIE 850 −120 V26 - - - - - - - - - -
Cl2/Ar ICP-RIE 980 −450 V27 670 −450 V27 150 −100 V22
HCl Photoassisted 0.004 nm/pulse28 - - - - - - - - - -
H2/Cl2 LE4 50-70 1-15eV29 - - - - - - - - - -

In CAIBE, an ion beam is directed onto a sample in a reactive gas ambient. The ion energy and beam current can be controlled while the flow of the reactive gas can also be controlled. Therefore, the physical and chemical etching components can be controlled independently. Adesida et al. [Reference Ping, Youtsey, Adesida, Khan and Kuznia9] have characterized CAIBE etching using Ar/Cl2. As shown in Fig. 1, the etch rates of GaN for 500 eV Ar and a flow of Cl2 incident on GaN increases with ion beam density at room temperature in curve (b). Etch rates as high as 160 nm/min was achieved; this is enhanced in comparison to the etch rates obtained with Ar ion beam as shown in curve (a). The enhancement is due to the chemical component brought about by the presence of the chlorine atoms. Another enhancement is observed for CAIBE at 200 °C where it has been ascertained that the thermal energy contributed primarily to the chemical etching mechanisms. CAIBE etch rates as high as 100 nm/min for GaN at room temperature have also been reported by Kneissl et al. [Reference Kneissl, Bour, Johnson, Romano, Krusor, Donaldson, Walker and Dunnrowicz30]. The trend for CAIBE etch rates of GaN has strong dependence on ion density and ion energy and moderate dependence on temperature and gas flow rate. The etch rates of AlxGa1−xN diminished linearly from x = 0 to 1 at room temperature [Reference Ping, Youtsey, Adesida, Khan and Kuznia9]. However, the etch rates for x < 0.1 were not significantly different, therefore heterostructures such as those utilized for lasers can essentially be etched at equi-etch rates. Ar/Cl2 CAIBE etching produced anisotropic but only near-vertical etch profiles [Reference Ping, Youtsey, Adesida, Khan and Kuznia9,Reference Binet, Duboz, Laurent, Bonnat, Collot, Hanauer, Briot and Aulombard31] at all substrate temperatures. The profiles were more vertical at higher temperatures due to increased chemical activities during etching [Reference Ping, Adesida and Khan8,Reference Ping, Youtsey, Adesida, Khan and Kuznia9]. In order to achieve the verticality necessary for laser facets, Binet [Reference Binet, Duboz, Laurent, Bonnat, Collot, Hanauer, Briot and Aulombard31] and Kneissl et al. [Reference Kneissl, Bour, Johnson, Romano, Krusor, Donaldson, Walker and Dunnrowicz30] tilted and rotated their samples while etching. InGaN/AlGaN laser diodes with CAIBE-etched facets have been fabricated and demonstrated using this method by Kneissl et al. [Reference Kneissl, Bour, Johnson, Romano, Krusor, Donaldson, Walker and Dunnrowicz30]. Highly vertical etch profiles have also been obtained by Ping et al. [Reference Ping, Adesida and Khan8] using Ar/HCl CAIBE at 300 °C with no tilting of samples. An example of such a vertical etch-profile produced in an AlGaN/GaN heterostructure is shown in Fig. 2. The ultra-smooth sidewall demonstrated in Fig. 2 was obtained using a regrown oxide masking process [Reference Khan, Youtsey and Adesida32]. The sidewall roughness was < 5 nm as measured in a scanning electron microscope.

Fig. 1. Etch rate of GaN vs Ar ion current

Fig. 2. GaN facet etched by CAIBE

High etch rates and highly anisotropic etch profiles have been obtained using high-density plasma reactive ion etching techniques. The high etch-rates produced by ECR-RIE and ICP-RIE methods are due to the higher plasma density available. The etch yields in the high-density plasma tools are the same as those in conventional RIE system, but the much larger ion fluxes in the former lead to higher etch rates. The higher efficiency of plasma generation also means that plasma can be generated and sustained in a higher vacuum environment than possible for conventional RIE tools. The independent biasing of the sample with 13.56 MHz generator provides for the control of the energy at which ions bombard the sample in high-density plasma etching tools. The directionality of low energy ions is preserved due to higher vacuum environment. This means that anisotropy can be achieved for etch profiles at lower ion energies.

A large body of work exists for ECR-RIE etching of III-nitrides [Reference Pearton, Abernathy, Ren, Lothian, Wisk and Katz16-Reference Vartuli, Pearton, Lee, Hong, MacKenzie, Abernathy and Shul21]. Various chemistries involving chlorine and methane-based gases have been investigated as shown in Table I. Etch rates for GaN ranging from 20 to 200 nm/min have been obtained for operating pressures between 1 to 10 mTorr and ion energies < −300 V. Corresponding etch rates for InN and AlN are comparable or lower than for GaN. For example, etch rates of 200, 150, and 110 nm/min were reported for GaN, InN, and AlN, respectively, using Cl2/H2/Ar plasma at a bias of −180 V [Reference Shul, Kilcoyne, Crawford, Parmeter, Vartuli, Abernathy and Pearton19]. The inclusion of H2 in the plasma was found to increase etch rates and also improve surface morphology. This was achieved by the removal of nitrogen through the formation of NHx products. In order to obtain even higher etch rates, novel gases such as ICl/Ar, IBr/Ar have been investigated [Reference Vartuli, Pearton, Lee, MacKenzie, Abernathy and Shul20,Reference Vartuli, Pearton, Lee, Hong, MacKenzie, Abernathy and Shul21]. High etch rates of 1300, 1150, and 200 nm/min in ICl/Ar plasma have been obtained for GaN, InN, and AlN, respectively. Highly anisotropic profiles were routinely obtained using ECR-RIE.

Another high-density plasma source is the ICP-RIE. ICP-RIE sources are easier to scale up than ECR sources and are more economical to operate [Reference Shul, McClellan, Casalnuovo, Rieger, Pearton, Constantine, Barratt, Karlicek, Tran and Schurman25]. These factors have led to the investigation of ICP-RIE for III-nitrides in various plasmas [Reference Cho, Vartuli, Donovan, Mackenzie, Abernathy, Pearton, Shul and Constantine22-Reference Smith, Wolden, Bremser, Hanser, Davis and Lampert27]. As expected, high etch rates and etch profiles with high anisotropy have been demonstrated. Etch rates of GaN as high as 688 nm/min at −280 V [Reference Shul, McClellan, Casalnuovo, Rieger, Pearton, Constantine, Barratt, Karlicek, Tran and Schurman25] and 980 nm/min at −450 V [Reference Smith, Wolden, Bremser, Hanser, Davis and Lampert27] have been reported for Cl2/H2/Ar and Cl2/Ar plasmas, respectively. A GaN etch rate of 850 nm was obtained using BCl3/Cl2 plasma at −120 V and 30 mTorr [Reference Lee, Kim, Yeom, Lee, Yoo and Kim26]. Etch rates for InN, AlN, and various mole fractions of AlInN and AlGaN have also been reported [Reference Cho, Vartulli, Donovan, Abernathy, Pearton, Shul and Constantine33]. Etch selectivities between these various materials have been reported for different gas mixtures including Cl2/SF6 [Reference Shul, Willison, Bridges, Han, Lee, Pearton, Abernathy, Mackenzie, Donovan, Zhang and Lester23]. Selectivities of 5 and 3 were reported for GaN on AlN and GaN on InN, respectively, using Cl2/Ar at −250 V [Reference Shul, Willison, Bridges, Han, Lee, Pearton, Abernathy, MacKenzie, Donovan, Zhang and Lester34]. Corresponding selectivity results for Cl2/SF6 at − 250 V were < 1 for GaN/AlN and 4 for GaN/InN, respectively [Reference Shul, Willison, Bridges, Han, Lee, Pearton, Abernathy, Mackenzie, Donovan, Zhang and Lester23]. It would be expected that the formation of AlF or InF should retard the etching of the AlN, InN and the ternaries, however, the dc biases used for the etching were relatively high. The high dc bias cause physical sputtering of the etch-stopping material leading to low selectivity values. The best reported selectivity for GaN on AlN of 38 was obtained using Cl2/Ar mixture at −20 V bias [Reference Smith, Wolden, Bremser, Hanser, Davis and Lampert27].

Highly anisotropic profiles with smooth sidewalls in GaN have been reported by Shul et al. [Reference Shul, McClellan, Casalnuovo, Rieger, Pearton, Constantine, Barratt, Karlicek, Tran and Schurman25]. A vertical etch profile with ultra-smooth sidewall in InGaN/AlGaN heterostructure obtained with a regrown oxide mask is shown in Fig. 3 [Reference Khan, Youtsey and Adesida32]. This was obtained using Cl2/Ar plasma at 2 mTorr and −160 V in a Plasmatherm SLR 790 ICP tool. These results have direct applications to laser facets.

Fig. 3. InGaN/AlGaN laser facet etched by ICP-RIE

Damage

Etch-induced damage can be manifested in different forms, all of which may affect the electronic and optical properties of devices fabricated on the etched materials. The forms in which damage can be categorized include: i) deposition of polymer, ii) creation of non-stoichiometric surfaces due to preferential depletion of one of the elements, iii) creation of near-surface lattice defects which can diffuse deep into the sample, and iv) implantation of etching species or hydrogen into the etched material. The few investigations that have been reported in this area measure electrical characteristics of the etched samples. Pearton et al. [Reference Pearton, Lee, MacKenzie, Abernathy and Shul35] performed Hall measurements on InN, InGaN, and InAlN exposed to Ar plasma under both ECR and conventional RIE conditions. It was found that the sheet resistances of the samples increased with increasing ion flux and ion energy. Ping et al. [Reference Ping, Schmitz, Khan, Chen, Yang and Adesida36] investigated the Schottky characteristics of Pd on etched n-GaN layers. The samples were etched by RIE in SiCl4 and Ar plasmas. Etching in SiCl4 plasmas degraded the Schottky barrier heights significantly for self-bias voltages above −200 V. The barrier height decreased from 0.93 eV to 0.41 eV. A more severe degradation to 0.38 eV occurred for Ar plasma at −100 V. Annealing at 700 °C was able to restore the barrier height of the SiCl4-etched sample.

The etch-induced degradation was utilized by Fang et al. [Reference Fang, Mohammad, Kim, Aktas, Botchkarev and Morkoς37] to improve the ohmic characteristics of Ti and Ti/Al on n-GaN layers. The samples were etched by RIE in BCl3 prior to metal deposition. It has also been found that etching with SiCl4 RIE improved ohmic characteristics under all investigated conditions [Reference Ping, Chen, Yang, Khan and Adesida38]. This can be explained by the depletion of nitrogen which leaves excess metallic Ga on the etched-surface. This change in stoichiometry renders the surface highly n-type, enhancing ohmic formation while degrading Schottky characteristics. The low bias etching that can be obtained using ICP-RIE and LE4 needs further investigations for realizing etched surfaces with low damage.

Wet Etching

Wet etching is an important complement to dry etching methods by providing low damage etching, low cost, and complexity. Conventional wet etching of GaN, AlN, and InN has been studied in base and acid solutions [Reference Morimoto39-Reference Mileham, Pearton, Abernathy and MacKenzie43]. Earlier studies [Reference Morimoto39] conducted on low quality GaN produced etch rates as high as 1 μm/min. However, recent studies by Mileham et al. [Reference Mileham, Pearton, Abernathy, MacKenzie, Shul and Kilcoyne40] did not produce any measurable etching for high quality GaN. Slow etch rates have also been recorded for InN. Pearton et al. [Reference Pearton, Abernathy, Ren, Lothian, Wisk and Katz41] found that InN etched very slowly in HCl/HNO3 solutions. Guo et al. [Reference Guo, Kato and Yoshida42] reported etch rates of ∼ 10 nm/min for InN using aqueous KOH and NaOH solutions at 60 °C. The etching of AlN was found to be highly dependent on the crystallinity of the sample [Reference Morimoto39]. Etch rates ranging from 10 to 1000 nm/min for AlN in KOH and AZ400K developer solution were reported by Mileham et al. [Reference Morimoto39,Reference Mileham, Pearton, Abernathy and MacKenzie43]. The lower end of the etch rates were obtained for high quality crystalline AlN. It can be concluded that chemical stability exhibited by III-nitrides has resulted in very low etch rates with conventional wet etchants.

Photoelectrochemical wet etching

A recent development is the demonstration of photoelectrochemical (PEC) wet etching which has resulted in significantly higher etch rates for GaN [Reference Minskey, White and Hu44-Reference Youtsey, Romano and Adesida53]. The PEC process utilizes photogenerated electron-hole pairs to enhance oxidation and reduction reactions taking place in an electrochemical cell. The etching of n-GaN proceeds through surface oxidation followed by dissolution in aqueous solutions. This process is enhanced by the photogenerated holes by converting surface atoms to higher oxidation states. Increasing absorption of incident optical radiation with energy greater than the bandgap energy increase the supply of holes at the surface, thereby enhancing the etch rates.

Minsky et al. [Reference Minskey, White and Hu44] first demonstrated PEC etching of n-GaN using KOH/H2O and dilute HCl solutions. They utilized HeCd laser at 325 nm wavelength for illuminating n-GaN samples at light intensities of ∼ 570 mW/cm2. The GaN sample was connected to a Pt cathode during etching with no external bias applied. Etch rates of ∼ 400 nm/min and 40 nm/min were obtained for the KOH and HCl solutions, respectively. No etching was observed in the absence of optical illumination. Youtsey et al. [Reference Youtsey, Adesida and Bulman45] have demonstrated the etching of n-GaN in KOH solutions using a broad-area Hg arc lamp. The electrochemical cell utilized by Youtsey et al. [Reference Youtsey, Adesida and Bulman45] is illustrated in Fig. 4. A Pt wire was used as the system cathode and a thin Ti (< 100 nm) metal was used as the mask. For a 0.04 M KOH solution. and for light intensities between 10 and 50 mW/cm2, etch rates were proportional to the light intensity and varied from 50 to 300 nm/min. Highly anisotropic etch profiles as shown in Fig. 5 were obtained with the rough surfaces attributed to defects in the sample. However, under conditions of very low KOH concentrations (< 0.01 M) and high light intensities, anisotropic etch profiles with very smooth surfaces were obtained by Youtsey et al. [Reference Youtsey, Adesida, Romano and Bulman46]. The reaction kinetics in the latter etching is believed to be diffusion-controlled. Using similar etching procedure in KOH solutions, Cho et al. [Reference Cho, Auh, Shul, Donovan, Abernathy, Lambers, Ren and Pearton47] reported etch rates greater than 100 nm/min and an activation energy of ∼ 0.8 kCal.mol−1 under diffusion controlled kinetics.

Fig. 4. Photoelectrochemical wet etching apparatus

Fig. 5. Highly anisotropic GaN structures by PEC etching

Lu et al. [Reference Lu, Wu and Bhat48] reported on photo-assisted anodic etching of n-GaN in solutions of tartaric acid/ethylene glycol at room temperature. Etch rates as high as 160 nm/min were obtained for Hg arc lamp illumination of ∼ 60 mW/cm2. Etch rates were found to be strongly dependent on the pH of the solutions. Peng et al. [Reference Peng, Chuang, Ho, Huang and Chen49] have also reported on the strong pH dependence of etch rates of n-GaN in aqueous H3PO4 and KOH solutions. Etch rates as high as 120 nm/min were obtained for H3PO4 solution of pH =1 and KOH solution of pH = 14. Current-controlled PEC etching of n-GaN in KOH solutions under HeCd laser illumination has been reported by Rotter et al. [Reference Rotter, Uffmann, Ackermann, Aderhold, Stemmer and Graul Mat50]. Etch rates of up to 8 μm/hr were obtained with smooth “mirror-like” etched surfaces.

Youtsey et al. [Reference Youtsey, Adesida and Bulman51] have demonstrated selective PEC etching of n-GaN on p-GaN in KOH solutions. No etching was detected for the p-GaN. However, O et al. [Reference Zory and Bour52] have demonstrated the etching of p-GaN and InGaN in an LED structure using pulsed electrochemical methods in H3PO4/ethylene glycol/H2O solutions.

In addition to the etching of GaN required for device fabrication, it has been shown that PEC can also reveal dislocations in n-GaN. Rotter et al. [Reference Rotter, Uffmann, Ackermann, Aderhold, Stemmer and Graul Mat50] observed etch pits with hexagonal symmetry and with a density in the range of 5 × 109 to 1 × 1010 cm−2, corresponding to the dislocation density of the films. Perhaps, the most striking results reported so far in this area pertains to the nanometer-scale “whisker-like” features obtained using border-line diffusion-controlled etching conditions [Reference Youtsey, Romano and Adesida53]. Figure 6 shows a scanning electron micrograph of “whiskers” obtained by etching n-GaN in 0.02 M KOH with a light intensity of 10 mW/cm2. The whiskers have diameters of ∼ 25 nm and lengths of ∼200 nm. Taller whiskers have been produced, however, at heights approaching 1 um, they coalesce forming tree like features. Figure 7 (a) shows a cross-sectional transmission electron micrograph (TEM) of the whiskers at low magnification. Both the whiskers and dislocations in the underlying unetched GaN are illustrated. At high magnification, propagation of dislocations from the unetched GaN into the etched whiskers is demonstrated in Fig. 7 (b). Both mixed (m) and edge (e) dislocations are associated with whisker formation as shown in the figure. The reduced etch rates at dislocations leading to whisker formation could arise due to a number of different effects [Reference Youtsey, Romano and Adesida53]. Based on the mechanism of etching, it can be proffered that a spatially varying concentration of photogenerated holes exists on the GaN surface. Weimann et al. [Reference Weimann, Eastman, Doppalapudi, Ng and Moustakas54] have modeled dislocations as negative charged coulombic centers in explaining the low transverse Hall mobilities in n-GaN. These negative charge centers can become sinks for photogenerated holes thereby locally depressing the concentration of holes that can participate in the etching. This would result in the lateral selective etching of the crystalline over dislocation areas in the n-GaN.

Fig. 6. Nanometer scale whiskers in GaN obtained by PEC etching

Fig. 7. (a) Cross-sectional TEM of etched GaN with whisker morphology, (b) higher magnification showing propagation of dislocations through the whiskers.

As seen, under certain PEC etching conditions, dislocations are isolated on n GaN. With the diameter of individual dislocation “whiskers” less than 50 nm, a plan view of a dislocation in a scanning electron microscope will be a spot or dot. Therefore an aerial image over a large area will result in the “star map” as shown Fig. 8. The counting of dots in a fixed area of the map yields the dislocation density of the material. Figure 8 shows the dislocation map of a high defect density MOCVD-grown n-GaN on SiC. The sample was etched in 0.004 M KOH under 30 mW/cm2 light intensity for 15 min. The dislocation density is estimated from the figure to be 3.2 × 109 cm−2. This estimation has been verified using TEM. It is seen that PEC method can be utilized for a rapid evaluation of dislocation densities in n-GaN materials. This method is less tedious than the conventional TEM method of assessing dislocation densities.

Fig. 8. ‘Star map’ of dislocations in GaN on SiC.

Summary

Dry and wet etching methods for III-nitrides have been reviewed. Although the high bonding energies have constituted obstacles to the etching of the nitrides, high-density-plasma etching methods have yielded etch rates that are suitable for device fabrication. The low bias voltages enabled by these high-density-plasma etching methods should allow low damage etching of nitride surfaces. More efforts are required in this area to quantify the processing latitude provided by techniques such as the ICP-RIE. The photoelectrochemical etching method has been shown to be an emerging method for device fabrication and material characterization. The PEC was shown to have the potential of becoming a rapid evaluation tool for dislocations in n-GaN.

Acknowledgments

This work was supported by at the University of Illinois by NSF grant No ECS 95-21671 and DARPA Grant No. F19628-96-C-0066; at CREE Research by DARPA Grant No. F19628-96-C-0066; and at Xerox by DARPA Grant No. MDA972-96-3-0014.

Footnotes

MRS Internet J. Nitride Semiconductor Res. 4S1, G1.4 (1999)

References

Nakamura, S., Mukai, T., and Senoh, M., Appl. Phys. Lett. 64, 1687 (1994).Google Scholar
Nakamura, S., Senoh, M., Nagahama, S., Iwasa, N., Yamada, T., Matsushita, T., Kiyoku, H., and Sugimoto, S., Appl. Phys. Lett. 68, 2105 (1996).Google Scholar
Ping, A. T., Adesida, I., Boutros, K., and Redwing, J., unpublished.Google Scholar
Sheppard, S., Doverspike, K., Pribble, W., Allen, S., Palmour, J., Kehias, L., Jenkins, T., Paper V.B.−5, Device Research Conference (Charlolltesville, VA, 1998).Google Scholar
McCarthy, L., Kozodoy, P., Rodwell, M., DenBaars, S., and Mishra, U., Compound Semiconductor 4 (8), 16 (1998).Google Scholar
Adesida, I., Ping, A.T., Youtsey, C., Dow, T., Khan, M. Asif, Olson, D.T., and Kuznia, J.N., Appl. Phys. Lett. 65, 889 (1994).Google Scholar
Pearton, S.J., Abernathy, C.R., Ren, F., and Lothian, J.R., J. Appl. Phys 76, 1210 (1994).Google Scholar
Ping, A.T., Adesida, I., and Khan, M. Asif, Appl. Phys. Lett. 67, 1250 (1995).Google Scholar
Ping, A.T., Youtsey, C., Adesida, I., Khan, M. Asif, and Kuznia, J.N., J. Electron. Mat. 24, 229 (1995).Google Scholar
Lee, J.-W., Park, H.-S., Park, Y.-J., Yoo, M.-C., Kim, T.-I, Kim, H.-S., and Yeom, G.-Y., MRS Symp. Proc. 468, 373 (1997).Google Scholar
Basak, D., Verdú, M., Montojo, M.T., Sánchez-García, M.A., Sánchez, F.J., Munoz, E., and Calleja, E., Semicond. Sci. Tech. 12, 1654 (1997).CrossRefGoogle Scholar
Lee, H., Oberman, D.B., and Harris, J.S. Jr, Appl. Phys. Lett. 67, 1754 (1995).Google Scholar
Adesida, I., Mahajan, A., Andideh, E., Khan, M. Asif, Olson, D.T., and Kuznia Appl, J.N.. Phys. Lett. 63, 2777 (1993).Google Scholar
Ping, A.T., Adesida, I., Khan, M. Asif, and Kuznia, J.N., Electron. Lett. 30, 1895 (1994).Google Scholar
Lin, M.E., Fan, Z.F., Ma, Z., Allen, L.H., and Morkoς, H., Appl. Phys. Lett. 64, 887 (1994).Google Scholar
Pearton, S.J., Abernathy, C.R., Ren, F., Lothian, J.R., Wisk, P.W., and Katz, A., J. Vac. Sci. Technol. A 11, 1772 (1993).Google Scholar
Pearton, S.J., Abernathy, C.R., and Vartuli, C.B., Electron. Lett. 30, 1985 (1994).Google Scholar
Zhang, L., Ramer, J., Brown, J., Zheng, K., Lester, L.F., and Hersee, S.D., Appl. Phys. Lett. 68, 367 (1996).Google Scholar
Shul, R.J., Kilcoyne, S.P., Crawford, M.H., Parmeter, J.E., Vartuli, C.B., Abernathy, C.R., and Pearton, S.J. , Appl. Phys. Lett. 66, 1761 (1995).Google Scholar
Vartuli, C.B., Pearton, S.J., Lee, J.W., MacKenzie, J.D., Abernathy, C.R., and Shul, R.J., J. Vac. Sci. Technol. B 15, 98 (1997).Google Scholar
Vartuli, C.B., Pearton, S.J., Lee, J.W., Hong, J., MacKenzie, J.D., Abernathy, C.R., and Shul, R.J., Appl. Phys. Lett. 69, 1426 (1996).Google Scholar
Cho, H., Vartuli, C.B., Donovan, S.M., Mackenzie, J.D., Abernathy, C.R., Pearton, S.J., Shul, R.J., and Constantine, C., J. Electron. Mat. 27, 166 (1998).Google Scholar
Shul, R.J., Willison, C.G., Bridges, M.M, Han, J., Lee, J.W., Pearton, S.J., Abernathy, C.R., Mackenzie, J.D., Donovan, S.M., Zhang, L., and Lester, L.F., J. Vac. Sci. Technol. A 16, 1621 (1998).Google Scholar
Vartuli, C.B., Lee, J.W., MacKenzie, J.D., Donovan, S.M., Abernathy, C.R., Pearton, S.J., Shul, R.J., Constantine, C., Barratt, C., Poyakov, A.Y., Shin, M., Skowronski, M., Mat. Res. Soc. Symp. Proc. 468, 393 (1997).Google Scholar
Shul, R.J., McClellan, G.B., Casalnuovo, S.A., Rieger, D.J., Pearton, S.J., Constantine, C., Barratt, C., Karlicek, R.F. Jr., Tran, C., and Schurman, M., Appl. Phys. Lett. 69, 1119 (1996).CrossRefGoogle Scholar
Lee, Y.H., Kim, H.S., Yeom, G.Y., Lee, J.W., Yoo, M.C., and Kim, T.I., J. Vac. Sci. Technol. A 16, 1478 (1998).Google Scholar
Smith, S.A., Wolden, C.A, Bremser, M.D., Hanser, A.D., Davis, R.F., and Lampert, W.V., Appl. Phys. Lett. 25, 3631 (1997).CrossRefGoogle Scholar
Leonard, R.T. and Bedair, S.M., Appl. Phys. Lett. 68, 794 (1996).CrossRefGoogle Scholar
Gillis, H.P., Choutov, D.A., Martin, K.P., Bremser, M.D., and Davis, R.F., J. Electron. Mat. 26, 301 (1997).Google Scholar
Kneissl, M., Bour, D.P., Johnson, N.M., Romano, L.T., Krusor, B.S., Donaldson, R., Walker, J., and Dunnrowicz, C., Apl. Phys. Lett. 72, 1539 (1998).Google Scholar
Binet, F., Duboz, J.Y., Laurent, N., Bonnat, C., Collot, P., Hanauer, F., Briot, O., and Aulombard, R.L., Apl. Phys. Lett. 72, 960 (1998).Google Scholar
Khan, F., Youtsey, C., and Adesida, I., unpublished.Google Scholar
Cho, H., Vartulli, C., Donovan, S., Abernathy, C., Pearton, S., Shul, R., and Constantine, C., J. Vac. Sci. Technol. A16, 1631 (1998).CrossRefGoogle Scholar
Shul, R., Willison, C., Bridges, M., Han, J., Lee, J., Pearton, S., Abernathy, C., MacKenzie, J., Donovan, S., Zhang, L., and Lester, L., J. Vac. Sci. Technol. A16, 1631 (1998).Google Scholar
Pearton, S.J., Lee, J.W., MacKenzie, J.D., Abernathy, C.R., and Shul, R.J., Appl. Phys. Lett. 67, 2329 (1995).Google Scholar
Ping, A.T., Schmitz, A.C., Khan, M. Asif, Chen, Q., Yang, J.W., and Adesida, I., J. Electron. Mater. 26, 266 (1997).CrossRefGoogle Scholar
Fang, Z., Mohammad, S.N., Kim, W., Aktas, O., Botchkarev, A.E., and Morkoς, H. [Appl. Phys. Lett. (USA) vol.68 (1996) p.1672]Google Scholar
Ping, A.T., Chen, Q., Yang, J.W., Khan, M. Asif, and Adesida, I., J. Electron. Mater. 27, 261 (1998).CrossRefGoogle Scholar
Morimoto, Y., J. Electrochem. Soc.121, 1384 (1974).Google Scholar
Mileham, J.R., Pearton, S.J., Abernathy, C.R., MacKenzie, J.D., Shul, R.J., and Kilcoyne, S.P., J. Vac. Sci. Technol. A14, 836 (1996).Google Scholar
Pearton, S., Abernathy, C., Ren, F., Lothian, J., Wisk, P., and Katz, A., J. Vac. Sci. Technol. 11, 1772 (1993).Google Scholar
Guo, Q.X., Kato, O., and Yoshida, A., J. Electrochem. Soc. 139, 2008 (1992).Google Scholar
Mileham, J.R., Pearton, S.J., Abernathy, C.R., and MacKenzie, J.D., Appl. Phys. Lett. 67, 1119 (1995)Google Scholar
Minskey, M.S., White, M., and Hu, E.L., Appl. Phys. Lett. 68, 1531 (1996).Google Scholar
Youtsey, C., Adesida, I., and Bulman, G., Appl. Phys. Lett. 71, 2151 (1997).CrossRefGoogle Scholar
Youtsey, C., Adesida, I., Romano, L.T., and Bulman, G., Appl. Phys. Lett. 72, 560 (1998).Google Scholar
Cho, H., Auh, K., Shul, R., Donovan, S., Abernathy, C., Lambers, E., Ren, F., and Pearton, S., J. Electron. Mater. (in press).Google Scholar
Lu, H., Wu, Z., and Bhat, I., J. Electrochem. Soc. 144, L8 (1997).Google Scholar
Peng, L. -H., Chuang, C., Ho, J., Huang, C. and Chen, C.-Y., Appl. Phys. Let. 72, 939 (1997)Google Scholar
Rotter, T., Uffmann, D., Ackermann, J., Aderhold, J., Stemmer, J., and Graul Mat, J.. Res. Soc. Symp. Proc. 482 (1997).Google Scholar
Youtsey, C., Adesida, I., and Bulman, G., J. Electron. Mater. 27, 282 (1998).Google Scholar
Zory, J. O, P.S. and Bour, D.P., SPIE Proc. 3002, 117 (1997).Google Scholar
Youtsey, C., Romano, L.T., and Adesida, I., Appl. Phys. Lett. 73, 797 (1998).CrossRefGoogle Scholar
Weimann, N.G., Eastman, L.F., Doppalapudi, D., Ng, H.M. and Moustakas, T.D., J. Appl. Phys. 83, 3656 (1998).Google Scholar
Ping, A.T., Khan, M. Asif, and Adesida, I., Semicond. Sci. Technol. 12, 133 (1997).Google Scholar
Figure 0

Table I. Summary of etch rates for various dry etching methods.

Figure 1

Fig. 1. Etch rate of GaN vs Ar ion current

Figure 2

Fig. 2. GaN facet etched by CAIBE

Figure 3

Fig. 3. InGaN/AlGaN laser facet etched by ICP-RIE

Figure 4

Fig. 4. Photoelectrochemical wet etching apparatus

Figure 5

Fig. 5. Highly anisotropic GaN structures by PEC etching

Figure 6

Fig. 6. Nanometer scale whiskers in GaN obtained by PEC etching

Figure 7

Fig. 7. (a) Cross-sectional TEM of etched GaN with whisker morphology, (b) higher magnification showing propagation of dislocations through the whiskers.

Figure 8

Fig. 8. ‘Star map’ of dislocations in GaN on SiC.