nonselective oxidation of gaas-based iii-v compound … · 2007. 10. 14. · gaasp/ingaasn mqw...

14
Nonselective oxidation of GaAs-based III-V compound semiconductor heterostructures for in-plane semiconductor lasers Di Liang, Jusong Wang, Douglas C. Hall* Dept. of Electrical Engineering, Univ. of Notre Dame, Notre Dame, IN USA 46556-5637 ABSTRACT A nonselective wet thermal oxidation technique for AlGaAs-containing heterostructures has been shown to enable the fabrication of a variety of novel high-efficiency, high-power GaAs-based in-plane laser devices. Applied in conjunction with a deep anisotropic dry etch, nonselective oxidation yields a simple, self-aligned high-index-contrast (HIC) ridge waveguide (RWG) structure. The native oxide grown directly on the waveguide ridge simultaneously provides excellent electrical insulation, passivation of the etch-exposed bipolar active region, and a low refractive index cladding, leading to numerous laser performance benefits. The resulting strong lateral optical confinement at the semiconductor/oxide interface (with refractive index contrast !n~1.7) enables half-racetrack ring resonator lasers with a record small 6 m bend radius. A nearly circularly-symmetric output beam is demonstrated on narrow w=1.4 m aperture width straight stripe-geometry lasers with single spatial and longitudinal mode total power output of ~180 mW at 228 mA (9x threshold). With the complete structural elimination of lateral current spreading, the excellent overlap of the optical field with the gain region provides high slope efficiency performance (ranging from >1.0 W/A at w=1.4 !m to 1.3 W/A for w=150 !m broad area stripes) for 300 K cw operation of unbonded, p-side up 808 nm InAlGaAs graded-index separate confinement heterostructure (GRINSCH) active region lasers. Using the direct thermal oxidation of a dilute nitride GaAsP/InGaAsN MQW active region, 1.3 m emission GaAs-based HIC RWG lasers exhibit a >2X threshold reduction and kink-free operation relative to conventional low-confinement devices. Other recent progress on the application of nonselective oxidation to GaAs-based semiconductor lasers will be reported. Keywords: high-index-contrast, ridge waveguide laser, thermal oxidation, half-racetrack ring resonator, surface passivation, dilute-nitride lasers, circular output beam 1. INTRODUCTION The well-known problem of beam asymmetry in the output from edge-emitting semiconductor diode lasers illustrates the challenges of forming a structure capable of providing adequate and dimensionally-comparable confinement of the optical field and the gain-providing carriers both in-plane and out-of-plane. While the evolution of epitaxially-grown heterostructures has provided the means for achieving a good optical confinement factor (i.e., overlap of optical field and gain) in the vertical growth direction, poor in-plane current and optical confinement has remained a chronic issue limiting diode-laser spatial mode quality and stability. Buried heterostructures employing layer disordering or etching and regrowth techniques have improved laser beam quality but not eliminated beam asymmetry. Vertical cavity surface emitting lasers can produce symmetric beams, but with much lower output power relative to edge-emitting devices. In this work, we review and further discuss our recent successful efforts to fabricate a high-index-contrast (HIC) ridge waveguide (RWG) structure for GaAs-based in-plane semiconductor lasers. Realized through the combination of a deep anisotropic etch through the heterostructure waveguide followed by a non-selective O 2 -enhanced wet thermal oxidation process, the HIC RWG completely eliminates lateral current spreading and provides a high lateral optical confinement factor. Efficient, high-power edge-emitting lasers, including devices with nearly-symmetric output beams, have been demonstrated. Benefiting from the strong optical confinement offered by a HIC interface, in-plane waveguides can also be sharply bent to route optical signals or form ring oscillators with negligible bending loss. There has been growing interest recently in HIC waveguides with !n >1 for integrated optics because of their potential to enable several orders of magnitude of growth in device integration density and complexity [1]. A major obstacle has been the greatly increased susceptibility of HIC waveguides to scattering loss, " s , directly proportional to the product " 2 (#n) 2 where " is the root-mean-square (RMS) sidewall surface roughness of a waveguide with core cladding effective index contrast #n; a more rigorous model predicts that " s increases in proportional to (#n) 3 [2]. In a deep-etched

Upload: others

Post on 27-Jan-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

  • Nonselective oxidation of GaAs-based III-V compound semiconductor

    heterostructures for in-plane semiconductor lasers

    Di Liang, Jusong Wang, Douglas C. Hall*

    Dept. of Electrical Engineering, Univ. of Notre Dame, Notre Dame, IN USA 46556-5637

    ABSTRACT

    A nonselective wet thermal oxidation technique for AlGaAs-containing heterostructures has been shown to enable the

    fabrication of a variety of novel high-efficiency, high-power GaAs-based in-plane laser devices. Applied in conjunction

    with a deep anisotropic dry etch, nonselective oxidation yields a simple, self-aligned high-index-contrast (HIC) ridge

    waveguide (RWG) structure. The native oxide grown directly on the waveguide ridge simultaneously provides excellent

    electrical insulation, passivation of the etch-exposed bipolar active region, and a low refractive index cladding, leading

    to numerous laser performance benefits. The resulting strong lateral optical confinement at the semiconductor/oxide

    interface (with refractive index contrast !n~1.7) enables half-racetrack ring resonator lasers with a record small 6 µm

    bend radius. A nearly circularly-symmetric output beam is demonstrated on narrow w=1.4 µm aperture width straight

    stripe-geometry lasers with single spatial and longitudinal mode total power output of ~180 mW at 228 mA (9x

    threshold). With the complete structural elimination of lateral current spreading, the excellent overlap of the optical field

    with the gain region provides high slope efficiency performance (ranging from >1.0 W/A at w=1.4 !m to 1.3 W/A for

    w=150 !m broad area stripes) for 300 K cw operation of unbonded, p-side up 808 nm InAlGaAs graded-index separate

    confinement heterostructure (GRINSCH) active region lasers. Using the direct thermal oxidation of a dilute nitride

    GaAsP/InGaAsN MQW active region, 1.3 µm emission GaAs-based HIC RWG lasers exhibit a >2X threshold reduction

    and kink-free operation relative to conventional low-confinement devices. Other recent progress on the application of

    nonselective oxidation to GaAs-based semiconductor lasers will be reported.

    Keywords: high-index-contrast, ridge waveguide laser, thermal oxidation, half-racetrack ring resonator, surface

    passivation, dilute-nitride lasers, circular output beam

    1. INTRODUCTION

    The well-known problem of beam asymmetry in the output from edge-emitting semiconductor diode lasers illustrates the

    challenges of forming a structure capable of providing adequate and dimensionally-comparable confinement of the

    optical field and the gain-providing carriers both in-plane and out-of-plane. While the evolution of epitaxially-grown

    heterostructures has provided the means for achieving a good optical confinement factor (i.e., overlap of optical field and

    gain) in the vertical growth direction, poor in-plane current and optical confinement has remained a chronic issue

    limiting diode-laser spatial mode quality and stability. Buried heterostructures employing layer disordering or etching

    and regrowth techniques have improved laser beam quality but not eliminated beam asymmetry. Vertical cavity surface

    emitting lasers can produce symmetric beams, but with much lower output power relative to edge-emitting devices. In

    this work, we review and further discuss our recent successful efforts to fabricate a high-index-contrast (HIC) ridge

    waveguide (RWG) structure for GaAs-based in-plane semiconductor lasers. Realized through the combination of a deep

    anisotropic etch through the heterostructure waveguide followed by a non-selective O2-enhanced wet thermal oxidation

    process, the HIC RWG completely eliminates lateral current spreading and provides a high lateral optical confinement

    factor. Efficient, high-power edge-emitting lasers, including devices with nearly-symmetric output beams, have been

    demonstrated. Benefiting from the strong optical confinement offered by a HIC interface, in-plane waveguides can also

    be sharply bent to route optical signals or form ring oscillators with negligible bending loss.

    There has been growing interest recently in HIC waveguides with !n >1 for integrated optics because of their potential

    to enable several orders of magnitude of growth in device integration density and complexity [1]. A major obstacle has

    been the greatly increased susceptibility of HIC waveguides to scattering loss, "s, directly proportional to the product

    "2(#n)

    2 where " is the root-mean-square (RMS) sidewall surface roughness of a waveguide with core cladding effective

    index contrast #n; a more rigorous model predicts that "s increases in proportional to (#n)3 [2]. In a deep-etched

  • structure which exposes the bipolar active region sidewall, passivation of non-radiative recombination centers from

    surface states or etch-damaged material is also critical to achieving good active semiconductor device performance. In

    this work, we review the demonstrated utility of using non-selective oxidation to effectively smooth and passivate the

    sidewalls of a deep-etched heterostructure ridge waveguide, overcoming the challenging issues of scattering loss and

    interface recombination. We demonstrate the improved optical mode quality and performance of HIC RWG lasers,

    including those with Al-free !~1.3 µm dilute nitride multi quantum well active regions, and present further data on the

    reduced current spreading and oxide thickness dependence of these devices.

    2. DEVICE FABRICATION

    We have elsewhere reported a simple process modification which enables a significantly enhanced oxidation rate for low

    Al-ratio AlxGa1-xAs (substantially reducing the oxidation rate selectivity to Al content) through the controlled addition of

    trace amounts of O2B [0-10000 ppm (1%) relative to N2B] to the process gas stream (N2+H2O vapor) [3, 4]. Low Al-ratio

    AlGaAs and even Al-free III-As waveguide core regions can now be oxidized laterally through this non-selective wet

    thermal oxidation technique without fully oxidizing the higher Al-ratio cladding layer, allowing a much higher, real

    lateral index step (!n~1.7) to be achieved. We have demonstrated a new self-aligned deep etch plus non-selective

    oxidation process for GaAs-based RWG laser fabrication, overcoming the typical limitations of deep-etched structures

    [5]. The laser fabrication process typically starts with the growth of a ~200 nm plasma-enhanced chemical vapor

    deposition (PECVD) SiNx mask layer to protect the p+-GaAs cap layer from a subsequent oxidation step. The

    waveguide stripe is patterned through conventional photolithography followed by two successive dry etching steps to

    translate the photoresist (PR) pattern to the SiNx layer and semiconductor epilayers. Unlike conventional dry etching

    which is stopped above the active layer so that defects introduced by etching are kept away from the active region, dry

    etching in this case reaches the lower cladding layer to form a waveguide with lateral dimensions close to that of the PR

    mask. Non-radiative recombination centers formed during this initial etching process are eliminated or largely reduced

    during the following non-selective thermal oxidation process, typically at 450 °C, through conversion to oxide or

    annealing, respectively. As shown in the schematic of Fig. 1(a) and the SEM cross-section image of Fig. 1(b), the oxide

    grown on the waveguide sidewalls (and base) results in a HIC (!n~1.7) semiconductor/oxide interface, enabling the

    realization of a HIC RWG providing strong optical confinement and capable of supporting very sharp bending, while

    simultaneously providing scaling from a conventional-lithography-defined ridge dimension ("1 "m) to the submicron

    dimensions required for HIC waveguide single-mode operation. Furthermore, instead of depositing PECVD SiO2 or

    SiNx for electrical confinement and surface passivation, the native oxide itself acts directly as the dielectric layer,

    providing a self-aligned process which eliminates the potential alignment errors, which unavoidably result from a second

    lithography step needed to open the contact window in the conventional fabrication process flow. A final dry etching

    procedure then selectively removes the dielectric masking stripe, using special care to prevent etch damage to the p+-

    GaAs cap layer, and the wafer is then thinned, metallized and cleaved into bars for laser characterization.

    (a) (b)

    Fig. 1. Schematic and SEM image of high-index-contrast ridge waveguide structure. Inset conduction band schematic shows location of graded-index waveguide core and quantum well.

  • To further highlight the advantages of this process, we note that the shallow etch in the conventional process flow can

    yield only a small lateral effective index step (!n"0.1), providing relatively weak optical mode confinement in the

    horizontal direction and leading to two undesirable effects: current spreading and output beam asymmetry. The

    significant current spreading (tens of microns) which plagues conventional index-guided RWG laser designs is

    prevented in this new device structure as current flow is effectively restrained to a vertical channel defined by the

    insulating oxide [6]. Uniform carrier distribution in the oxide-confined active region also substantially minimizes spatial

    mode-hopping, resulting in kink-free operation (i.e., output power vs. current device characteristics which are linear and

    free of kinks at currents above threshold) [6]. Strong optical mode confinement from the vertical oxide walls also offers

    a potential for overcoming the limitation of the asymmetric optical mode profile and corresponding output beam in-plane

    vs. out-of-plane far-field divergence angle asymmetry in edge-emitting lasers [7], a well-known disadvantage for

    numerous applications. We have elsewhere shown that the non-selective wet thermal oxidation process enables a

    significant reduction of semiconductor waveguide scattering loss through an effect known as oxidation smoothing

    where a thermal oxidation process smoothes the sidewall roughness as the oxidation front progresses inward [8].

    Compared with the lithography and etching processes typically required to achieve submicron feature sizes, non-

    selective oxidation is shown to be well-controllable for the formation of submicron structures simply by the tuning of

    several process parameters (e.g., temperature, O2 concentration and flow rate of N2 carrier gas), and can be carried out

    with much lower cost equipment. The proposed HIC process clearly can provide a significant improvement in the

    device performance/cost ratio.

    3. DEVICE PERFORMANCE

    The light output power vs. current (L-I) characteristics in Fig. 2 show the good kink-free laser performance for a narrow-

    stripe (w=5 µm) device with a cavity length of 361 µm under both pulsed current injection (1% duty cycle) and a

  • maximum (FWHM) and 1/e2 power width measured on the pattern at

    I=90 mA are 4.03 µm and 5.51 µm, respectively. These two values,

    both significantly smaller than the 8.9 µm aperture size, demonstrate

    the strong lateral optical confinement provided by the HIC interface.

    In Fig. 3 (b), single-mode operation is achieved on a w=5 µm device

    and well maintained at both current levels of 70 mA and 100 mA,

    with 1.99 µm FWHM and 3.41 µm 1/e2 width values measured at

    I=100 mA.

    Spatial single-mode operation in the directions both parallel and

    perpendicular to the junction plane is also further confirmed in the

    far-field pattern of the same w=5 µm device at different output

    power levels [Fig. 3(c)]. Mode simulations show that the higher-

    order modes of the same waveguide structure are cut-off at a width

    around w=1 µm so a passive RWG with w=5 µm is not expected to

    operate in the single-mode regime. However, single-mode operation

    for HIC waveguides in active laser devices are largely affected by

    mode competition where the fundamental mode with the lowest loss

    reaches stimulated emission first and consumes most of the carriers,

    suppressing the lasing probability for higher-order transverse modes.

    Higher order modes also penetrate further into the oxide cladding

    and can thus experience greater loss due to absorption by the metal

    on the outside of the ridge.

    4. REALIZATION OF SYMMETRIC OUTPUT BEAM

    The asymmetry of the divergence angles, defined as the ratio of

    !///!", is only 2.73 in Fig. 3(c), smaller than the typical values of !4

    observed in conventional shallow-etched index-guided RWG lasers.

    The strong dependence of the optical mode on the waveguide

    dimension offers a promising path to the further reduction of the

    output beam asymmetry by further scaling the aperture dimension.

    An asymmetric mode profile is clearly shown in the mode simulation

    of Fig. 4 (a) for a passive AlGaAs rib waveguide structure (w=4 µm)

    commonly employed for a conventional shallow-etched RWG laser.

    Due to the 1:2.4 ratio compression of the horizontal scale in the top

    of Fig. 4, the asymmetry for this representative conventional design

    is much worse (~27:1) than it appears. Reducing the rib waveguide

    width actually inversely increases the lateral dimension of the optical

    mode due to a loss of effective optical confinement as shown in Fig.

    4(b). By using a slightly broadened active region and squeezing the

    mode laterally with the low index (n~1.6) native oxide, however, a

    circular mode (1:1 aspect ratio) can be obtained in a HIC RWG [Fig.

    4(c)].

    Diode lasers with a stripe dimension "5 µm are fabricated to

    experimentally demonstrate the possibility of achieving a perfectly

    circular output beam. Fig. 5 shows the L-I characteristic of a HIC

    RWG straight laser showing a small aperture size of 1.39 µm. With

    threshold current of Ith=25.2 mA and laser cavity length of L=1107.1

    µm, the threshold current density is Jth=1644 A/cm2. This higher

    current density causes high junction temperature and a more

    noticeable red shift of the peak wavelength (#p=826.27 nm)

    measured at 189 mA true cw [Fig. 5 inset (a)]. At a true cw current

    (a)

    (b)

    (c)

    Fig. 3. Near-field patterns of 8.9 µm (b) and 5 µm (a)

    wide HIC RWG lasers at true cw currents of 60 and 90

    mA, and 70 and 100 mA, respectively. Multi-mode

    operation is observed on the w=8.9 µm device as

    additional intensity spots which appears when the

    current is increased from 60 mA to 90 mA. FWHM

    and power width at the 1/e2 level are 4.03 µm and 5.51

    µm, respectively. Single-mode operation is well

    maintained on the w=5 µm device at both current

    levels with 1.99 µm FWHM and 3.41 µm 1/e2 width

    values. Single-lobe far-field patterns at 20 mW output

    power in both directions parallel and perpendicular to

    the epilayers are shown in (c), indicating a FWHM

    beam asymmetry of 2.64.

  • of 180 mA and an output power close to 150 mW, the measured near-

    field pattern in Fig. 5 inset (b) still shows a smooth single-mode profile

    with a FWHM of 0.55 µm and 1/e2 width of 0.99 µm, demonstrating

    strong mode stability. The fast-dc LI measurement, limited to a 100

    mA injection current, shows no signs of thermal roll-over.

    The Rd=1.02 W/A slope efficiency in Fig. 5, corresponding to an

    !d=68% external quantum efficiency at a wavelength of 826.27 nm, is

    remarkably high considering the laser aperture is only 1.39 µm wide. An extrapolation of the efficiency data of devices

    with 90 µm, 40 µm, 10 µm, 7 µm and 5 µm aperture dimensions predicts an external quantum efficiency of only ~40%

    for a 1.39 µm wide aperture laser [9]. The unexpectedly high efficiency can be attributed to stopping of the dry etch in

    this device before penetration through the waveguide core and SQW active region, leading to a reduction in the number

    of etching-induced defects introduced into the active region. As shown by the dashed line in Fig. 4(c) which marks the

    level of the SQW, the oxide growth here still penetrates through a large part of waveguide core layer to achieve a HIC

    RWG. This thermal oxidation process converts much of the etch-damaged material to an amorphous oxide, eliminating

    a substantial numbers of the non-radiative recombination centers. Other remaining etch-induced defects in the

    semiconductor may also be annealed out during the next thermal oxidation process when considering that 350 °C anneals

    have been shown to promote some recovery from RIE etch damage [10].

    Lasers with other aperture dimensions are also fabricated and measured under the same conditions. Fig. 6 shows that for

    shrinking laser lateral dimensions (from 15 !m to 7, 4.67, 2.67, 1.62, and 1.39 µm), the FWHM divergence angle "//

    parallel to the junction plane increases due to increased light diffraction (from 5.5° to 8.8°, 15.0°, 15.94°, 22.05°, and

    Fig. 5. Pulsed and cw 300 K light-current characteristic of a HIC RWG

    straight with Ith=25.2 mA, active width w=1.39 !m, and laser cavity

    length L=1107.1 !m, leading to the threshold current density Jth=1644

    A/cm2. The differential slope and quantum efficiencies at I=100 mA are

    Rd=1.02 W/A and !d=68%, respectively. Inset (a): Linear spectrum

    measured at 189 mA true cw shows a peak wavelength of 826.27 nm.

    Inset (b): Near-field profile measured at 180 mA true cw shows a 0.55 µm

    FWHM and 0.99 µm 1/e2 width. Inset (c): SEM cross-section image of

    w=1.39 !m HIC RWG structure after etching (with 200 nm-thick PECVD

    SiNx mask layer on the ridge top) and 20 min, 450 °C nonselective

    oxidation. Dashed line shows SQW location.

    Fig. 4. (a,b) conventional and (c) native oxide-

    defined AlGaAs/GaAs passive waveguide

    structures.

  • 28.39°, respectively). The divergence increase is not linear but accelerates as the stripe width gets smaller and smaller.

    A weaker, opposite dependence of the divergence angle in the direction perpendicular to the junction plane on laser

    stripe width is also observed in Fig. 6, resulting in a noticeable reduction of the beam asymmetry !"/!// from >8 (at w=15

    µm) down to 1.23 (at w=1.39 µm). While the vertical dimension (i.e., thickness of the waveguide core layer) is not

    changed, the divergence angle !" still decreases from 47.1° to 35° as the stripe width is reduced from 15 µm to 1.39 µm.

    The variation of !" can be explained from an expression (Eqn. (1)) derived from Ref. [11] by Dumke [12], which is

    applicable not only to the case of a thin waveguide layer [11], but also that of a thick waveguide laser (i.e., large optical

    cavity):

    2 2

    / /

    1 ( / ) /1.2 1/ ( / ) /1.2

    ad d

    a d a d

    ! !"

    ! !#= =

    + + (1)

    where a=0.41#2(n1

    2-n2

    2), d, #, n1 and n2 are the thickness of the waveguide core layer, free-space wavelength and

    refractive indices of the waveguide core and cladding, respectively. Because Fig. 6 shows the divergence angle when all

    lasers are well above threshold under true cw mode operation without a heatsink, heat can easily build up inside the

    resonance cavity (i.e., the waveguide region here) but to a varying degree for lasers with stripe widths of 1.39, 1.62,

    2.67, 4.67, 7 and 15 µm. As the data trend in Fig. 6 shows, narrower devices consume more injection current to

    compensate the losses from non-radiative recombination and optical scattering, which results in a higher current density

    and a higher junction temperature. The narrower stripe lasers, therefore, experience more heat buildup than the wider

    ones since temperature is proportional to the current density. Because the material refractive index (real part) reduces

    with rising material temperature, the waveguide core index n1 for the 1.39 µm laser is smaller than that of the 15 µm

    laser. On the other hand, the cladding index n2 does not vary much since heat generation occurs mostly in the active

    region (within the waveguide) where, because the doping is the lowest, Joule (I2R) heating is the greatest and where non-

    recombination processes most likely occur due to bipolar activity. It thus turns out that the parameter a in Eqn (1)

    decreases (i.e., 1/a increases) from thermal effects as the device dimension narrows, leading to a smaller and smaller fast

    axis divergence angle !". Furthermore, the power area density at the laser emission facet is the highest for the w=1.39

    µm laser since the >50 mW/facet output power is distributed over the smallest area (w!d). This high power area density

    further increases the local temperature at the laser facet and consequently causes further narrowing of the fast axis

    divergence angle.

    The extrapolated curve fits in Fig. 6 predict that the

    a perfectly circular beam will require a submicron

    aperture (w=0.56 µm) device where the divergence

    angle is 32.4° in directions both parallel and

    normal to the junction plane. The passive

    waveguide mode simulation based on the same

    wafer and device structure (Fig. 6 inset) also shows

    a circularly symmetric mode profile when w=0.5

    µm. The small discrepancy may be due to the

    passive nature of the mode simulation which

    neglects carrier-dependent index variations present

    in active devices. Utilizing the good controllability

    of the relatively slow non-selective oxidation

    process, submicron aperture devices can potentially

    be patterned by conventional photolithography and

    dry etching when followed by a longer oxidation to

    achieve the target effective aperture size.

    The divergence angle of 32.4° projected for a

    perfectly symmetric output beam device is not

    optimal for butt-coupling to a commercial silica

    glass fiber whose index step of no more than 0.01

    gives an acceptance angle of only 11.9.° However,

    through use of a micro-lens system or tapered lens

    Fig. 6. A plot of the laser beam divergence angles and divergence angle

    ratio (i.e., beam asymmetry) vs. laser stripe width. The point of

    intersection of the curve fits predicts that a perfectly circular output beam

    with 32.4° divergence angles could be achieved at a stripe width of w=0.56 µm. Inset: passive waveguide mode simulation when w=0.5 µm.

  • fiber, such a laser may offer improved matching of the laser output mode to the fiber mode. The methods reported here

    for increasing the lateral optical confinement factor and eliminating current spreading are similar and complementary to

    recently reported slab-coupled optical waveguide approaches [13, 14]. It may be possible to further minimize or

    overcome the mode mismatch problem between a laser beam and optical fiber by employing an oxide-confined HIC

    device with a different, optimized epilayer design. The symmetric beam lasers described here may also offer a low

    astigmatism beam for high-brightness focused spot applications such as optical disk writing or laser printing.

    5. STUDIES OF LATERAL CURRENT SPREADING AND OPTICAL CONFINEMENT

    The unavoidable current spreading in conventional RWG lasers prevents the reduction of their lateral dimension while

    still maintaining low threshold operation. Since the laser beam is not well guided laterally outside the waveguide region,

    the carriers escaping out from the ridge prior to getting collected by the underlying quantum well region either dont

    contribute to the photon generation at all, or generate spontaneous emission in directions not amplified by the guided-

    mode stimulated emission. These lost photons essentially act as an electrical loss in the laser cavity and are likely to

    cause increased heating. Furthermore, negative impacts also include spatial mode distortion (e.g., a more elliptical far-

    field radiation pattern as in Fig. 4), mode hopping-induced kinks in the L-I curve, a perturbation of the longitudinal mode

    via lateral mode instability [15] and a rise in intermodulation distortion under direct modulation [16]. Tsang has done

    some early work to analyze the effects of lateral current spreading on stripe-geometry double-heterostructure lasers [17],

    followed by Hu et al. giving experimental data and a modified theoretical model for RWG lasers [18, 19]. Letal et al.

    have reported that up to 42% of the injection current escapes from the index-guided region at threshold in a w=2 µm

    InGaAsP/InP MQW laser [20], emphasizing why conventional RWG lasers are normally limited to stripe widths of w!5

    µm. A lateral spreading distance of up to 20 µm is reasonably estimated from simulations [15]. Direct observation of

    the current spreading using scanning voltage microscopy has also been recently reported [21], enabling an intuitive

    understanding from a visual perspective.

    Numerous approaches to device structure optimization have been demonstrated to minimize current spreading, including

    varying the residual upper cladding thickness (i.e., etch depth) [15, 22] and utilizing a buried heterostructure [23]. It is

    obvious that a reduced residual upper cladding layer thickness improves the lateral carrier confinement, however, ~3-4"

    higher optical scattering loss is also typically observed [22] since the guided light interacts with the waveguide sidewall

    roughness more intensely. Higher internal optical loss plus the probable extension of etch defects into the active region

    consequently leads to an overall higher threshold current [22]. A buried heterostructure employing an etch plus

    regrowth process to confine the etch-defined p-waveguide region by an adjacent regrown n-cladding material has been

    shown to reduce the current spreading through reverse biasing of the n-cladding/p-waveguide diode junction.

    Nevertheless, up to 30% of the threshold current in such single-mode devices still does not contribute to active region

    pumping [23]. Laser devices in which a buried current aperture is formed through selective lateral oxidation of a buried

    high Al-ratio AlxGa1-xAs layer (1 ! x ! 0.9) placed above the waveguide core layer [24-26], inspired from oxide-

    confined VCSELs, show a strong capability for suppressing the escape of carriers due to current spreading. This

    approach, however, doesnt provide much increase in the lateral index step to increase optical confinement, and is thus

    ill suited for achieving devices with smaller dimension or sharply bent features.

    The HIC RWG laser structure described herein simultaneously tackles the two limitations of conventional RWG devices:

    carrier confinement and device scalability. For comparative analysis, weak index-guided lasers employing only a

    shallow-etch have been fabricated along with HIC RWG lasers from the same 808 nm GRINSCH material.

    Approximately 925 nm of the combined 200 nm GaAs p+ cap layer plus 1.5 µm upper cladding layer is removed by RIE

    after patterning straight stripes with varying widths. A 20 min non-selective oxidation with the addition of 4000 ppm O2

    at 450 °C is then applied to grow about 200 nm of oxide, leaving a residual Al0.6Ga0.4As upper cladding layer thickness

    of 575 nm. A deeply etched sample with the same mask pattern is oxidized under the same conditions to closely match

    device dimensions. Fig. 7 shows the relationship of the threshold current density of weak index-guided lasers vs.

    reciprocal laser cavity length. Both narrow stripe (w=5, 7 µm) and broad-area devices (w=90 µm), usually immune to

    current spreading, experience a higher threshold current density to varying degrees. At L=500 µm, for instance, weak

    index-guided broad-area lasers need 492.75 A/cm2 to reach lasing threshold while the HIC broad-area devices require

    only 48.6% of this amount. Narrow-stripe devices with stripe widths of 5 and 7 µm exhibit 2608.7 and 1855.1 A/cm2

    threshold current density at the same cavity length, 2.85X and 2.53X higher, respectively, than their HIC counterparts.

  • The stripe-width dependent Jth data of Fig. 8 reveals a similar picture but further demonstrates that the threshold current

    density of the weak index-guided structure climbs much faster than that of the HIC structure as the laser stripe width is

    reduced, indicating a proportionately larger fraction of carriers are escaping from the ridge waveguide region in

    conventional devices. Unlike the shallow-etched devices, narrow stripe HIC devices can suffer somewhat from non-

    radiative recombination at the semiconductor/oxide interface, but the greater overall consequence of eliminating current

    spreading leads to a more than 2! performance improvement overall on w " 15 µm devices. Two sets of devices with

    similar cavity length are selected for comparison in Fig. 8 in

    order to minimize the influence of the distributed mirror

    loss. Fig. 8 also shows that even broad-area devices (w>50

    µm) fabricated with a HIC RWG structure consistently

    show improved performance (~24% lower Jth) due to the

    complete and significant elimination of current spreading.

    The HIC RWG achieved through this process enables very

    tight waveguide bends with low loss [27]. This has been

    demonstrated through the fabrication of half-racetrack-ring

    resonator lasers with a bend radius as low as r=6 !m, as

    shown here in Fig. 9 [28]. In addition to lasing for e-beam

    lithography defined devices with r=25, 10 and 8 !m

    reported in [28], Fig. 9 shows lasing with comparable

    performance for an r=6 !m, w=10 !m ridge width device.

    The slightly improved threshold current and efficiency is

    due to the r=6 !m devices shorter cavity length (636 !m

    vs. 1 mm for the others).

    6. OXIDE THICKNESS DEPENDENCE

    The native oxide formed at the etch-exposed active region

    is the paramount feature in this HIC structure, particularly

    for active devices, since it serves multiple key purposes

    including interface passivation, sidewall roughness

    Fig. 7. Comparison of weak index-guided and HIC RWG laser

    threshold current densities vs. inverse laser cavity length 1/L,

    showing the significantly higher threshold current density for (b)

    w=7 and (c) 5 µm weak index-guided devices vs. (e) 7 and (f) 5

    µm HIC RWG lasers. Even (a) w=90 µm weak index-guided

    broad-area lasers show notable degradation due to current

    spreading relative to (d) their HIC RWG counterpart.

    Fig. 8. Threshold current density vs. laser stripe width for HIC

    devices and weak index-contrast devices, both fabricated on

    808 nm GRINSCH structure and probe tested in room

    temperature and under pulsed mode (1% duty cycle) with p-

    side up. Up to 2.38X threshold current density reduction is

    achieved on w=5 µm HIC devices due to elimination of current

    spreading.

    0

    40

    80

    120

    160

    200

    240

    0 300 600 900 1200 1500

    Outp

    ut

    Po

    wer

    (mW

    )

    Current (mA)

    (a) r=25 µm

    (b) r=10 µm

    (c) r=8 µm

    Pulsed, 300 K (unbonded, p-side up)

    r=8 µm

    (d) r=6 µm

    Fig. 9. Total output power vs. current characteristics for w=10 !m wide HIC RWG lasers in half-racetrack-ring geometry with bend radii of (a)

    r=25 !m, (b) r=10 !m, (c) r=8 !m and (d) r=6 !m. Total resonator cavity lengths are ~1 mm for (a)-(c) and 636 !m for (d).

  • reduction, metal/semiconductor isolation and optical confinement. Each role is critical to the HIC RWG structure and

    failure in any aspect could derail the whole enterprise. It is therefore of great importance to explore if there is a

    minimum oxide thickness required for proper device operation, or even, perhaps, whether there is an optimal oxide

    thickness.

    HIC RWG lasers devices have been fabricated with two different oxide thicknesses. Fig. 10 exhibits significantly

    different levels of performance for lasers distinguished only by their oxide thickness at the QW active region: (a) 104

    and (b) 546 nm, corresponding to 10 and 40 min nonselective oxidations with addition of 4000 ppm O2, respectively.

    Two bars with comparable bar

    length are selected for a legitimate

    comparison and both of them

    represent typical performance (i.e.,

    neither the best or the worse)

    among all devices tested. In Fig.

    10(a), only broad-area devices with

    stripe widths of 120, 90, 60 and 40

    µm operate as lasers, albeit with

    fairly poor performance. The w=25

    µm device behaves strangely,

    followed by only spontaneous light

    emission on devices of w=15 and

    10 µm. Narrow w=7 and 5 µm

    stripe devices are still diodes but

    emit no light at all. In contrast,

    devices in Fig. 10 (b) show

    consistently high slope efficiency

    on broad-area devices with a

    maximum of 1.28 W/A observed

    on w=120 µm devices. The lower

    average slope efficiency for all

    widths of 1.05 W/A is attributed to

    the comparatively low slope

    efficiency (Rd < 1 W/A) of w=7 and

    5 µm narrow stripe devices, which

    sit near the edge of the bar and are

    thus also subject to poor facet

    cleaves. It is noted that the stripe

    widths in Fig. 10 refers to the photo

    mask dimension instead of the

    actual laser aperture size obtained

    by deducting the oxide thickness.

    Clearly, a 104 nm thin oxide is

    inadequate for achieving desirable

    device performance, due to a

    combination of insufficient

    interface passivation, inadequate

    oxidation smoothing, poor

    insulating property, or high optical

    loss owing to the evanescent wave

    penetration through the thin oxide

    layer to the absorbing sidewall

    metal. The failure of w=5 and 7

    µm devices in Fig. 10 (a) to emit

    0

    50

    100

    150

    200

    250

    300

    350

    400

    -1

    0

    1

    2

    3

    4

    0 200 400 600 800 1000 1200 1400

    w=120 µm

    w=90 µm

    w=60 µm

    w=40 µm

    w=25 µm

    w=15 µm

    w=10 µm

    w=7 µm

    w=5 µm

    To

    tal O

    up

    ut

    Po

    we

    r, P

    (m

    W)

    Vo

    ltag

    e, V

    (V)

    Injection Current, I (mA)

    pulsed, 300 Kp-side up, unbonded

    L=934 µm

    0.78 V

    QW 104 nm oxide

    (a)

    0

    50

    100

    150

    200

    250

    300

    350

    400

    0 200 400 600 800 1000 1200 1400

    w=120 µm

    w=90 µm

    w=60 µm

    w=40 µm

    w=25 µm

    w=15 µm

    w=10 µm

    w=7 µm

    w=5 µm

    To

    tal O

    utp

    ut

    Po

    we

    r, P

    (m

    W)

    Injection Current, I (mA)

    Rd: 0.92~1.28 W/A

    Ave. Rd=1.05 W/A

    pulsed, 300 Kp-side up, unbonded

    L=915 µm

    QW

    546 nm oxide

    (b)

    Fig. 10. Light-current characteristic of (a) a L=934 µm long bar with 104 nm non-selective

    oxide formed at the active region (inset) in 10 min, 450 °C oxidation with 4000 ppm O2

    participation, showing poor device performance, and (b) a L=915 µm long bar with 546 nm

    non-selective oxide grown at the active region (inset) in 40 min oxidation with 4000 ppm

    O2 participation, showing decent device performance with a slope efficiency as high as 1.28

    W/A on a w= 120 µm broad-area device and an average of 1.05 W/A for all devices. The

    current-voltage characteristic of the 5 µm wide device (not working) in (a) indicates

    substantial current leakage with a low turn-on voltage.

  • light altogether spotlights the factors of interface passivation and oxide insulating property, as there should be some

    weak, spontaneous emission even if the metal absorption or scattering was too high to achieve stimulated emission. The

    current-voltage characteristic of the w=5 µm device shown in Fig. 10 (a) indicates an anomalously low turn-on voltage

    (0.78 V) compared to the normal range of 1.6-1.7 V. An early turn-on may result from a high leakage current through

    the thin AlGaAs oxide, whose insulating property is just fair when compared [29] with other III-V compound

    semiconductor (InAlP) native oxides [30]. Preliminary studies on the electrical properties of non-selective oxides do

    show a negligible leakage of JL

  • InGaAs quantum wells for which we have

    also observed oxide growth (data not

    shown). It has been shown by Luo et al. [4]

    that the addition of O2 significantly

    enhances the oxidation rates of an undoped

    Al0.20Ga0.8As waveguide core containing a

    single 10 nm GaAs quantum well. Fig. 11

    now demonstrates that substantially thicker

    GaAs layers and even a dilute-nitride MQW

    structure can also be non-selectively

    oxidized.

    For HIC RWG laser fabrication, devices are

    deeply etched via RIE with a BCl3/Cl2/Ar

    plasma for 12 min to form a 1.8 µm high

    ridge. A 2 hour non-selective oxidation at

    450 °C with the addition of 7000 ppm O2 is

    then used to grow ~2.5 !m of oxide

    (measured at the etch-exposed

    GaAsP/InGaAsN MQW active region),

    resulting in an effective laser active aperture

    on all devices 5 µm smaller than the

    optically patterned laser stripe width. For

    comparison, conventional index-guided

    RWG lasers are fabricated with a shallow

    0.75 µm deep etch (requiring 8 min under

    the same dry etch conditions), followed by a

    short 30 min nonselective oxidation at 450

    °C with 7000 ppm O2 to convert part of the

    Al0.65Ga0.35As upper cladding layer to a 200

    nm native oxide for device isolation.

    Fig. 12 shows the pulsed (1% duty cycle) L-

    I characteristic of a broad-area device (c)

    with L=307 µm cavity length, w=85 µm

    (effective laser aperture dimension). A low

    threshold current of 169.6 mA and a high

    slope efficiency of 0.51 W/A, corresponding

    to a threshold current density of 650 A/cm2

    and a differential quantum efficiency of

    50.8% (at !peak=1.234 µm), respectively, are

    demonstrated. Fig. 12 also shows typical L-

    I characteristics for (a) HIC and (b)

    conventional RWG lasers with a w~10 µm

    narrow stripe laser effective active stripe

    width. It is well known that both poor

    optical confinement and carrier leakage via

    current spreading can lead to mode hopping

    in weakly-guided narrow-stripe lasers which

    in turn causes kinks in the L-I

    characteristics, as was above demonstrated

    for conventional shallow-etched 808 nm

    GRINSCH devices (Fig. 7). Such behavior

    is observed for the device of Fig. 12 as well,

    Fig. 11. SEM cross-section image of GaAsP/InGaAsN MQW structure, wet

    etched and nonselectively wet oxidized at 450 °C for 30 min with 7000 ppm

    added O2. The conduction band overlay schematically highlights the epitaxial

    structure in the image, showing ~115 nm and 40 nm oxide growth on GaAs

    waveguide core and GaAsP/InGaAsN MQW active region, respectively.

    Fig. 12. Pulsed L-I characteristics of typical w~10 !m stripe geometry lasers with

    (a) HIC and (b) conventional shallow-etched index-guided RWG structures. For

    (a) HIC diode laser with L=525 !m, Ith=45.73 mA, Jth=871 A/cm2, and the

    differential slope efficiency at 2Ith is Rd=0.451 W/A. For (b) conventional device

    with L=520 !m, Ith=86.18 mA and Jth=1657.3 A/cm2, and a kink indicative of

    mode hopping occurs (typical of most of the conventional devices). For (c) a

    L=307 µm HIC-type broad-area device with w=85 µm effective laser aperture,

    pulsed (1% duty cycle) L-I characteristic of showing Ith=169.6 mA and Rd=0.51

    W/A. Inset: spectrum of HIC diode laser operating at 100 mA pulsed current,

    showing a peak wavelength of 1.23 !m and SEM of a different dry etch-exposed

    waveguide sidewall nonselectively oxidized at 450 °C for 1 hour.

  • and is typical in most of our conventional devices. In contrast, the HIC RWG laser of Fig. 12 shows kink-free operation

    suggesting stable spatial-mode behavior. As shown in Fig. 12, laser operation with a low-threshold current (Ith=45.7

    mA) and high slope efficiency (Rd=0.45 W/A) is obtained without visible mode-hopping induced L-I kinks. The inset

    spectrum is measured at a pulsed injection current of 100 mA (~2.2!Ith), showing a peak wavelength of 1.23 µm. It has

    been shown elsewhere that growth modifications not incorporated in the structures used in this work can extend the

    wavelength to close to 1300 nm [35, 37]. The inset SEM cross-sectional image of the etch-exposed RWG sidewall

    oxidized under the same conditions but for a shorter time period of 1 hour, resulting in about 430-1220 nm of oxide

    growth in the MQW active region. An apparent superlinear lateral oxidation rate at the GaAsP/InGaAsN MQW active

    region observed from three samples oxidized for 30 min, 1 and 2 hours to thicknesses of 40 nm, 1220 nm and 2500 nm,

    respectively, can be attributed to the additional effect of inward oxidation of this more slowly oxidizing region from the

    surrounding faster-oxidizing GaAs and AlGaAs layers. The non-uniform oxidation observed in the AlGaAs cladding

    layers and GaAs waveguide p-n junction layer may be attributed to doping-related effects [40] and interface-enhanced

    oxidation [41] observed in other heterostructures.

    Furthermore, up to 2.3! lower threshold current density is achieved on native oxide-confined HIC w=5 µm narrow stripe

    devices relative to conventional, shallow-etched RWG devices with comparable cavity length because of the total

    elimination of current spreading (data not shown) [9]. A similar 2.38! threshold current density reduction on 808 nm

    GRINSCH HIC RWG lasers with a 5 µm stripe width was shown above in Fig. 8, indicating a comparable performance

    gain in both material structures.

    8. CONCLUSION

    In summary, we have reviewed recent progress in high-index-contrast ridge waveguide lasers offering improved spatial

    mode behavior and capability for a high-power symmetric output beam from edge-emitting laser devices. The

    performance benefits achieved through the complete elimination of current spreading are highlighted through a

    comparative analysis of weak-index-guided and HIC RWG devices. A comparison of device performance for different

    oxide thicknesses is reported. Application of the deep-etch plus non-selective oxidation fabrication process has been

    demonstrated for !=808 nm AlGaAs-GaAs GRINSCH and !~1.3 !m Al-free InGaAsN/GaAsP/GaAs active region

    lasers. In addition to providing a high lateral optical confinement factor for efficient performance of conventional edge-

    emitting lasers, the HIC RWG is shown to enable very-small radius waveguide bends with low bend loss, opening

    potential new vistas for dense photonic integration.

    ACKNOWLEDGEMENTS

    We gratefully acknowledge J. Y.-T. Huang, J.-Y. Yeh, and L. J. Mawst of the University of Wisconsin at Madison for

    growing the dilute nitride active region heterostructures used in this work. This work was supported in part by the

    National Science Foundation under Grants ECS-0123501 and ECS-0601702.

    REFERENCES

    [1] M. Smit, "Photonic integration," Telektronikk, vol. 2.2005, pp. 66-71, 2005.

    [2] S. Suzuki, M. Yanagisawa, Y. Hibino, and K. Oda, "High-density integrated planar lightwave circuits using

    SiO2/GeO2 waveguides with a high refractive index difference," IEEE Journal of Lightwave Technology, vol.

    12, pp. 790-796, 1994.

    [3] Y. Luo, "Properties of AlGaAs native oxides for integrated photonics and optoelectronics applications," Ph.D.

    Dissertation, Department of Electrical Engineering: University of Notre Dame, Indiana, 2001.

    [4] Y. Luo and D. C. Hall, "Nonselective Wet Oxidation of AlGaAs Heterostructure Waveguides Through

    Controlled Addition of Oxygen," IEEE Journal of Selected Topics in Quantum Electronics, vol. 11, pp. 1284-

    1291, 2005.

    [5] D. Liang, J. Wang, and D. C. Hall, "High-efficiency native-oxide-passivated high-index-contrast ridge

    waveguide lasers," Electronics Letters, vol. 42, pp. 349-350, 2006.

  • [6] D. Liang, D. C. Hall, J. Y.-T. Huang, J.-Y. Yeh, and L. J. Mawst, "High-Index-Contrast Oxide-Confined

    GaAsP/InGaAsN Multi-Quantum-Well Ridge Waveguide Lasers," presented at The 20th IEEE International

    Semiconductor Laser Conference, Big Island, Hawaii, USA, 2006.

    [7] D. Liang, J. Kulick, and D. C. Hall, "High-Efficiency Oxide-Confined Ridge Waveguide Laser with Nearly

    Symmetric Output Beam," presented at The 19th Annual Lasers and Electro Optics Society Meeting, Montreal,

    Canada, 2006.

    [8] D. Liang, D. C. Hall, and G. M. Peake, "Oxidation Smoothing of Sidewall Roughness in AlGaAs

    Heterostructure Waveguides," presented at The 18th Annual Lasers and Electro Optics Society Meeting,

    Sydney, Australia, 2005.

    [9] D. Liang, J. Wang, J. Y.-T. Huang, J.-Y. Yeh, L. J. Mawst, and D. C. Hall, "Deep-Etched Native-Oxide-

    Confined High-Index-Contrast AlGaAs Heterostructure Lasers With 1.3 !m Dilute-Nitride Quantum Wells,"

    Submitted to IEEE Journal of Selected Topics in Quantum Electronics, 2006.

    [10] H. F. Wong, D. L. Green, T. Y. Liu, D. G. Lishan, M. Bellis, E. L. Hu, P. M. Petroff, P. O. Holtz, and J. L.

    Merz, "Investigation of reactive ion etching induced damage in GaAs/AlGaAs quantum well structures,"

    Journal of Vacuum Science & Technology B: Microelectronics and Nanometer Structures, vol. 6, pp. 1906-

    1910, 1988.

    [11] H. C. Casey, Jr., M. B. Panish, and J. L. Merz, "Beam divergence of the emission from double-heterostructure

    injection lasers," Journal of Applied Physics, vol. 44, pp. 5470-5475, 1973.

    [12] W. Dumke, "The angular beam divergence in double-heterojunction lasers with very thin active regions," IEEE

    Journal of Quantum Electronics, vol. 11, pp. 400-402, 1975.

    [13] G. Ru, X. Yu, J. Cai, J. Yan, and F. S. Choa, "Robust slab-coupled buried-rib semiconductor laser with high

    fibre coupling efficiency," Electronics Letters, vol. 40, pp. 1538-1539, 2004.

    [14] R. K. Huang, J. P. Donnelly, L. J. Missaggia, C. T. Harris, J. Plant, D. E. Mull, and W. D. Goodhue, "High-

    power nearly diffraction-limited AlGaAs-InGaAs semiconductor slab-coupled optical waveguide laser," IEEE

    Photonics Technology Letters, vol. 15, pp. 900-902, 2003.

    [15] A. Martin and H. Amos, "Lateral current spreading in ridge waveguide laser diodes," Applied Physics Letters,

    vol. 74, pp. 1364-1366, 1999.

    [16] J. Piprek, J. K. White, and A. J. SpringThorpe, "What limits the maximum output power of long-wavelength

    AlGaInAs/InP laser diodes?," IEEE Journal of Quantum Electronics, vol. 38, pp. 1253-1259, 2002.

    [17] W. T. Tsang, "The effects of lateral current spreading, carrier out-diffusion, and optical mode losses on the

    threshold current density of GaAs-Al[sub chi ]Ga[sub 1 - chi ]As stripe-geometry DH lasers," Journal of

    Applied Physics, vol. 49, pp. 1031-1044, 1978.

    [18] S. Y. Hu, D. B. Young, A. C. Gossard, and L. A. Coldren, "The effect of lateral leakage current on the

    experimental gain/current-density curve in quantum-well ridge-waveguide lasers," IEEE Journal of Quantum

    Electronics, vol. 30, pp. 2245-2250, 1994.

    [19] S. Y. Hu, S. W. Corzine, K.-K. Law, D. B. Young, A. C. Gossard, L. A. Coldren, and J. L. Merz, "Lateral

    carrier diffusion and surface recombination in InGaAs/AlGaAs quantum-well ridge-waveguide lasers," Journal

    of Applied Physics, vol. 76, pp. 4479-4487, 1994.

    [20] G. J. Letal, J. G. Simmons, J. D. Evans, and G. P. Li, "Determination of active-region leakage currents in ridge-

    waveguide strained-layer quantum-well lasers by varying the ridge width," IEEE Journal of Quantum

    Electronics, vol. 34, pp. 512-518, 1998.

    [21] D. Ban, E. H. Sargent, K. Hinzer, J. D.-W. St, A. J. SpringThorpe, and J. K. White, "Direct observation of

    lateral current spreading in ridge waveguide lasers using scanning voltage microscopy," Applied Physics

    Letters, vol. 82, pp. 4166-4168, 2003.

    [22] R. K. Price, V. B. Verma, K. E. Tobin, K. C. Hsieh, V. C. Elarde, and J. J. Coleman, "Intrinsic parameter and

    modal characteristics of asymmetric cladding ridge waveguide lasers," presented at The 19th Annual Laser &

    Electro-Optics Society Meeting, Montreal, Canada, 2006.

    [23] G. Belenky, L. Shterengas, C. L. J. Reynolds, M. W. Focht, M. S. Hybertsen, and B. Witzigmann, "Direct

    measurement of lateral carrier leakage in 1.3-!m InGaAsP multiple-quantum-well capped mesa buried

    heterostructure lasers," IEEE Journal of Quantum Electronics, vol. 38, pp. 1276-1281, 2002.

    [24] S. Furst, C. Farmer, L. Hobbs, R. D. L. Rue, and M. Sorel, "Native oxidation of aluminum-containing III-V

    compound layers for increased current and optical confinement in semiconductor lasers," presented at The 19th

    Annual Lasers and Electro Optics Society Meeting, Montreal, Canada, 2006.

  • [25] G. A. Vawter, O. Blum, A. Allerman, and Y. Gao, "Highly-efficient laser with self-aligned waveguide and

    current confinement by selective oxidation," presented at The 12th Annual Lasers and Electro-Optics Society

    Meeting. LEOS '99. IEEE, 1999.

    [26] J. Heerlein, S. Gruber, M. Grabherr, R. Jager, and P. Unger, "Highly efficient laterally oxidized λ=950

    nm InGaAs-AlGaAs single-mode lasers," IEEE Journal of Selected Topics in Quantum Electronics, vol. 5, pp.

    701-706, 1999.

    [27] D. Liang, J. Wang, and D. C. Hall, "Oxide-Confined High Index Contrast Ridge Waveguide Curved Resonator

    Laser Diodes," presented at The 18th Annual Lasers and Electro Optics Society Meeting, Sydney, Australia,

    2005.

    [28] J. Wang, D. Liang, and D. C. Hall, "High-Performance Small-Radius Half-Racetrack-Ring-Resonator

    InAlGaAs Quantum Well Laser Diodes Fabricated via Non-Selective Wet Oxidation," presented at The 20th

    IEEE International Semiconductor Laser Conference, Big Island, Hawaii, USA, 2006.

    [29] P. J. Barrios, S.-K. Cheong, D. C. Hall, N. C. Crain, G. L. Snider, C. B. DeMelo, T. Shibata, B. A. Bunker, U.

    Chowdhury, R. D. Dupuis, G. Kramer, and N. El-Zein, "Residual As and the Electrical Characteristics of

    AlGaAs Native Oxides for MOS Applications," presented at The 42nd Electronic Materials Conference,

    Denver, Colorado, USA, 2000.

    [30] Y. Cao, J. Zhang, X. Li, T. H. Kosel, P. Fay, D. C. Hall, X. B. Zhang, R. D. Dupuis, J. B. Jasinski, and Z.

    Liliental-Weber, "Electrical properties of InAlP native oxides for metal--oxide--semiconductor device

    applications," Applied Physics Letters, vol. 86, pp. 062105, 2005.

    [31] S. W. Pang, W. D. Goodhue, T. M. Lyszczarz, D. J. Ehrlich, R. B. Goodman, and G. D. Johnson, "Dry etching

    induced damage on vertical sidewalls of GaAs channel," Journal of Vacuum Science & Technology B:

    Microelectronics and Nanometer Structures, vol. 6, pp. 1916-1920, 1988.

    [32] M. Heinbach, J. Kaindl, and G. Franz, "Lattice damage in III/V compound semiconductors caused by dry

    etching," Applied Physics Letters, vol. 67, pp. 2034-2036, 1995.

    [33] G. Morello, M. Quaglio, G. Meneghini, C. Papuzza, and C. Kompocholis, "Reactive ion etching induced

    damage evaluation for optoelectronic device fabrication," Journal of Vacuum Science & Technology B:

    Microelectronics and Nanometer Structures, vol. 24, pp. 756-761, 2006.

    [34] J.-Y. Yeh, N. Tansu, and L. J. Mawst, "Temperature-sensitivity analysis of 1360-nm dilute-nitride quantum-

    well lasers," IEEE Photonics Technology Letters, vol. 16, pp. 741-743, 2004.

    [35] N. Tansu, J.-Y. Yeh, and L. J. Mawst, "Improved photoluminescence of InGaAsN--(In)GaAsP quantum well by

    organometallic vapor phase epitaxy using growth pause annealing," Applied Physics Letters, vol. 82, pp. 3008-

    3010, 2003.

    [36] N. Tansu, N. J. Kirsch, and L. J. Mawst, "Low-threshold-current-density 1300-nm dilute-nitride quantum well

    lasers," Applied Physics Letters, vol. 81, pp. 2523-2525, 2002.

    [37] N. Tansu, Jeng-YaYeh, and L. J. Mawst, "Physics and characteristics of high performance 1200 nm InGaAs

    and 1300-1400 nm InGaAsN quantum well lasers obtained by metal-organic chemical vapour deposition,"

    Journal of Physics: Condense Matter vol. 16, pp. S3277-S3318, 2004.

    [38] S. R. Bank, L. L. Goddard, M. A. Wistey, H. B. Yuen, and J. S. Harris, Jr., "On the temperature sensitivity of

    1.5-!m GaInNAsSb lasers," IEEE Journal of Selected Topics in Quantum Electronics, vol. 11, pp. 1089-1098,

    2005.

    [39] Y. Qu, C. Y. Liu, and S. Yuan, "High-power 1.3-!m InGaAsN strain-compensated lasers fabricated with pulsed

    anodic oxidation," Applied Physics Letters, vol. 85, pp. 5149-5151, 2004.

    [40] F. A. Kish, S. A. Maranowski, G. E. Hofler, N. Holonyak, Jr., S. J. Caracci, J. M. Dallesasse, and K. C. Hsieh,

    "Dependence of doping type (p/n) of the water vapor oxidation of high-gap AlxGa1-xAs," Applied Physics

    Letters, vol. 60, pp. 3165-3167, 1992.

    [41] S. A. Maranowski, N. Holonyak, Jr., T. A. Richard, and F. A. Kish, "Photon-induced anisotropic oxidation

    along p-n-junctions in AlxGa1-xAs-GaAs quantum-well heterostructures," Applied Physics Letters, vol. 62, pp.

    2087-2089, 1993.