Gain-guided broad area quantum cascade lasers emitting 23 ... · 14. A. Bismuto, S. Blaser, R....

9
Gain-guided broad area quantum cascade lasers emitting 23.5 W peak power at room temperature I LIA SERGACHEV , 1,2,* RICHARD MAULINI , 1 ALFREDO BISMUTO, 1 STEPHANE BLASER, 1 TOBIAS GRESCH, 1 AND ANTOINE MULLER 1 1 Alpes Lasers SA, Avenue des Paquiers 1, CH-2075 St-Blaise, Switzerland 2 University of Basel, Department of Chemistry, Klingelbergstrasse 80, CH-4056 Basel, Switzerland * [email protected] http://www.alpeslasers.ch Abstract: We report gain-guided broad area quantum cascade lasers at 4.55 μm. The devices were processed in a buried heterostructure configuration with a current injector section much narrower than the active region. They demonstrate 23.5 W peak power at a temperature of 20 C and duty cycle of 1%, while their far field consists of a single symmetric lobe centered on the optical axis. These experimental results are supported well by 2D numerical simulations of electric currents and optical fields in a device cross-section. c 2016 Optical Society of America OCIS codes: (140.5965) Semiconductor lasers, quantum cascade; (140.3538) Lasers, pulsed; (230.7370) Waveguides. References and links 1. A. Lyakh, R. Maulini, A. Tsekoun, R. Go, and C. K. N. Patel, “Tapered 4.7 μm quantum cascade lasers with highly strained active region composition delivering over 4.5 watts of continuous wave optical power,” Opt. Express 20, 4382–4388 (2012). 2. Y. Bai, N. Bandyopadhyay, S. Tsao, S. Slivken, and M. Razeghi, “Room temperature quantum cascade lasers with 27% wall plug eciency,” Appl. Phys. Lett. 98, 181102 (2011). 3. R. Maulini, A. Lyakh, A. Tsekoun, and C. K. N. Patel, “λ 7.1 μm quantum cascade lasers with 19% wall-plug eciency at room temperature,” Opt. Express 19, 17203–17211 (2011). 4. Y. Bai, S. Slivken, S. R. Darvish, A. Haddadi, B. Gökden, and M. Razeghi, “High power broad area quantum cascade lasers,” Appl. Phys. Lett. 95, 221104 (2009). 5. B. Gökden, Y. Bai, N. Bandyopadhyay, S. Slivken, and M. Razeghi, “Broad area photonic crystal distributed feed- back quantum cascade lasers emitting 34 W at λ 4.36 μm,” Appl. Phys. Lett. 97, 131112 (2010). 6. D. Heydari, Y. Bai, N. Bandyopadhyay, S. Slivken, and M. Razeghi, “High brightness angled cavity quantum cascade lasers,” Appl. Phys. Lett. 106, 091105 (2015). 7. J. D. Kirch, C.-C. Chang, C. Boyle, L. J. Mawst, D. Lindberg, T. Earles, and D. Botez, “5.5 W near-diraction- limited power from resonant leaky-wave coupled phase-locked arrays of quantum cascade lasers,” Appl. Phys. Lett. 106, 061113 (2015). 8. F. Salin and J. Squier, “Gain guiding in solid-state lasers,” Opt. Lett. 17, 1352–1354 (1992). 9. M. N. Polyanskiy, “Refractive index database,” http://refractiveindex.info (2016). 10. V. Siklitsky, “Electronic archive: New semiconductor materials. characteristics and properties,” http://www.ioffe.ru/SVA/NSM/ (2001). 11. A. Lyakh, P. Zory, D. Wasserman, G. Shu, C. Gmachl, M. D’Souza, D. Botez, and D. Bour, “Narrow stripe-width, low-ridge high power quantum cascade lasers,” Appl. Phys. Lett. 90, 141107 (2007). 12. R. Terazzi and J. Faist, “A density matrix model of transport and radiation in quantum cascade lasers,” New J. Phys. 12, 033045 (2010). 13. K. Krishnaswami, B. E. Bernacki, B. D. Cannon, N. Ho, and N. C. Anheier, “Emission and Propagation Properties of Midinfrared Quantum Cascade Lasers,” IEEE Photon. Tech. Lett. 20, 306–308 (2008). 14. A. Bismuto, S. Blaser, R. Terazzi, T. Gresch, and A. Muller, “High performance, low dissipation quantum cascade lasers across the mid-IR range,” Opt. Express 23, 5477 (2015). Introduction Quantum cascade lasers (QCLs) have become the sources of choice for chemical sensing and directional infrared countermeasure applications in the mid-infrared spectral range (3-15 μm). Vol. 24, No. 17 | 22 Aug 2016 | OPTICS EXPRESS 19063 #267948 http://dx.doi.org/10.1364/OE.24.019063 Journal © 2016 Received 8 Jun 2016; revised 19 Jul 2016; accepted 29 Jul 2016; published 9 Aug 2016

Transcript of Gain-guided broad area quantum cascade lasers emitting 23 ... · 14. A. Bismuto, S. Blaser, R....

Page 1: Gain-guided broad area quantum cascade lasers emitting 23 ... · 14. A. Bismuto, S. Blaser, R. Terazzi, T. Gresch, and A. Muller, “High performance, low dissipation quantum cascade

Gain-guided broad area quantum cascadelasers emitting 23.5 W peak power at roomtemperature

ILIA SERGACHEV,1,2,* RICHARD MAULINI,1 ALFREDO BISMUTO,1

STEPHANE BLASER,1 TOBIAS GRESCH,1 AND ANTOINE MULLER1

1Alpes Lasers SA, Avenue des Paquiers 1, CH-2075 St-Blaise, Switzerland2University of Basel, Department of Chemistry, Klingelbergstrasse 80, CH-4056 Basel, Switzerland*[email protected]://www.alpeslasers.ch

Abstract: We report gain-guided broad area quantum cascade lasers at 4.55 μm. The deviceswere processed in a buried heterostructure configuration with a current injector section muchnarrower than the active region. They demonstrate 23.5 W peak power at a temperature of 20◦Cand duty cycle of 1%, while their far field consists of a single symmetric lobe centered onthe optical axis. These experimental results are supported well by 2D numerical simulations ofelectric currents and optical fields in a device cross-section.

c© 2016 Optical Society of America

OCIS codes: (140.5965) Semiconductor lasers, quantum cascade; (140.3538) Lasers, pulsed; (230.7370) Waveguides.

References and links1. A. Lyakh, R. Maulini, A. Tsekoun, R. Go, and C. K. N. Patel, “Tapered 4.7 μm quantum cascade lasers with highly

strained active region composition delivering over 4.5 watts of continuous wave optical power,” Opt. Express 20,4382–4388 (2012).

2. Y. Bai, N. Bandyopadhyay, S. Tsao, S. Slivken, and M. Razeghi, “Room temperature quantum cascade lasers with27% wall plug efficiency,” Appl. Phys. Lett. 98, 181102 (2011).

3. R. Maulini, A. Lyakh, A. Tsekoun, and C. K. N. Patel, “λ ∼7.1 μm quantum cascade lasers with 19% wall-plugefficiency at room temperature,” Opt. Express 19, 17203–17211 (2011).

4. Y. Bai, S. Slivken, S. R. Darvish, A. Haddadi, B. Gökden, and M. Razeghi, “High power broad area quantumcascade lasers,” Appl. Phys. Lett. 95, 221104 (2009).

5. B. Gökden, Y. Bai, N. Bandyopadhyay, S. Slivken, and M. Razeghi, “Broad area photonic crystal distributed feed-back quantum cascade lasers emitting 34 W at λ ∼ 4.36 μm,” Appl. Phys. Lett. 97, 131112 (2010).

6. D. Heydari, Y. Bai, N. Bandyopadhyay, S. Slivken, and M. Razeghi, “High brightness angled cavity quantumcascade lasers,” Appl. Phys. Lett. 106, 091105 (2015).

7. J. D. Kirch, C.-C. Chang, C. Boyle, L. J. Mawst, D. Lindberg, T. Earles, and D. Botez, “5.5 W near-diffraction-limited power from resonant leaky-wave coupled phase-locked arrays of quantum cascade lasers,” Appl. Phys. Lett.106, 061113 (2015).

8. F. Salin and J. Squier, “Gain guiding in solid-state lasers,” Opt. Lett. 17, 1352–1354 (1992).9. M. N. Polyanskiy, “Refractive index database,” http://refractiveindex.info (2016).10. V. Siklitsky, “Electronic archive: New semiconductor materials. characteristics and properties,”

http://www.ioffe.ru/SVA/NSM/ (2001).11. A. Lyakh, P. Zory, D. Wasserman, G. Shu, C. Gmachl, M. D’Souza, D. Botez, and D. Bour, “Narrow stripe-width,

low-ridge high power quantum cascade lasers,” Appl. Phys. Lett. 90, 141107 (2007).12. R. Terazzi and J. Faist, “A density matrix model of transport and radiation in quantum cascade lasers,” New J. Phys.

12, 033045 (2010).13. K. Krishnaswami, B. E. Bernacki, B. D. Cannon, N. Ho, and N. C. Anheier, “Emission and Propagation Properties

of Midinfrared Quantum Cascade Lasers,” IEEE Photon. Tech. Lett. 20, 306–308 (2008).14. A. Bismuto, S. Blaser, R. Terazzi, T. Gresch, and A. Muller, “High performance, low dissipation quantum cascade

lasers across the mid-IR range,” Opt. Express 23, 5477 (2015).

Introduction

Quantum cascade lasers (QCLs) have become the sources of choice for chemical sensing anddirectional infrared countermeasure applications in the mid-infrared spectral range (3-15 μm).

Vol. 24, No. 17 | 22 Aug 2016 | OPTICS EXPRESS 19063

#267948 http://dx.doi.org/10.1364/OE.24.019063 Journal © 2016 Received 8 Jun 2016; revised 19 Jul 2016; accepted 29 Jul 2016; published 9 Aug 2016

Page 2: Gain-guided broad area quantum cascade lasers emitting 23 ... · 14. A. Bismuto, S. Blaser, R. Terazzi, T. Gresch, and A. Muller, “High performance, low dissipation quantum cascade

Continuous-wave (cw) output powers up to ≈ 5 W at room temperature have been reported [1,2].While cw power is limited by heat dissipation due to the relatively low wall plug efficiency(≤20%) [3] of QCLs, much higher peak powers can be reached in pulsed mode operation byscaling the size of the device. Broad active region configurations are interesting for QCLs, be-cause, unlike diode lasers, they do not suffer from filamentation and, therefore, can operatestably in a single transverse mode at currents well above threshold [4]. However, if no specialprecaution is taken, because of the TM polarization facet reflectivity increases with mode indexand broad area QCLs lase on a high-order mode with a far field profile consisting of 2 lobespropagating at large angles from the optical axis [4]. Several approaches, using photonic crystalgratings [5], angled cavities [6] and phase locked arrays [7], attempt to overcome this problem.

In this work we achieved high peak powers at room temperature in lasers operating at zeroorder optical mode. For this, an unusual waveguide design was implemented: active region wasnot etched after growth - only the cladding above was. This created a very wide active region,compared to standard QCL designs, and a narrow current injector on top. Therefore currentdistribution in active region became strongly non-uniform, with a decay from laser center toits sides. As active region was not etched, usual optical mode confinement by refractive indexstep in horizontal direction was eliminated. Instead, the gain profile corresponding to horizontalcurrent density spread led to gain guiding and efficient optical mode selection [8].

Simulations

The effect of gain guiding was accurately simulated numerically. One symmetric half of a 2Dcross-section of a QCL waveguide was used in all simulations with layer geometries corre-sponding to ones in fabricated devices - see Fig. 1 and section about fabrication for details. Itis important to emphasize, that the depth of etching was not known precisely after fabrication;therefore, it was used as a fitting parameter. Simulations with different values of it demonstratedscaling of current density distribution width without changing its magnitude, and no qualitativechanges in device physics. A value of 7 μm, which corresponds to 1 μm of cladding thicknessleft unetched, was found to make the closest match between simulated and experimental dataand therefore was used. Refractive index and conductivity values for the materials were takenfrom [9] and [10]. No change in simulated laser properties was observed while varying currentinjector width from 8 to 16 μm; value of 10 μm is shown in Fig. 1.

First, the current density distributions in the whole cross-section were calculated for a rangeof voltages applied between top and bottom contacts. For the active region much wider thanin conventional index-guided QCLs, the current density in it could not be assumed uniformanymore, and the local dependency of conductivity in growth direction on electric field was ac-counted for. This dependency was obtained experimentally, as proposed in [11], from a standardmeasurement of IV curve of a narrow ridge device with same active region design, where thecurrent density is uniform. Even though the non-uniform conductivity added a complication ofcyclic dependency of input variable on the output solution for the numerical solver, the solu-tion still converged. The distributions of current density in the active region for a set of appliedvoltages are shown in (Fig. 2). They were integrated over the contacts of the laser as well tosimulate the IV curve (Fig. 3).

Second, the distributions of gain as function of current density were calculated for each ap-plied voltage. The function of gain was simulated according to [12].

Third, a boundary electromagnetic mode analysis was conducted for each distribution of gain,accounting for refractive indices of layers as well. For all voltages >16 V, the TM00 mode iseffectively confined by the gain distribution and its imaginary refractive index is higher, than forTM01, which indicates, that it also reaches the lasing threshold first (Fig. 4). Higher order modeshave negative net gain (= losses) in this simulation. An example of mode electric field intensityis shown in Fig. 1. The role of gain guiding was proven by running an additional simulation

Vol. 24, No. 17 | 22 Aug 2016 | OPTICS EXPRESS 19064

Page 3: Gain-guided broad area quantum cascade lasers emitting 23 ... · 14. A. Bismuto, S. Blaser, R. Terazzi, T. Gresch, and A. Muller, “High performance, low dissipation quantum cascade

0 5 10 15 20 25 30

Transverse direction, μm

0

2

4

6

8

10

12Growth

direction,μm

i-InP regrowth

n-InP

topclad

ding

n-InPActive region

Lower cladding

Substrate0.0

0.2

0.4

0.6

0.8

1.0

Fig. 1. Simulated TM00 optical mode intensity (color, a.u.) and current flow (red arrows)in the center of a cross-section of a gain-guided laser (one transverse symmetric half isshown). Applied voltage is 20 V.

0 50 100 150 200

Transverse direction, μm

0

1

2

3

4

5

Curren

tden

sity,kA/cm

2

25 V

23 V

21 V

19 V

17 V

15 V14 16 18 20 22 24 26

Voltage, V

1530456075

HW

HM,μm

Fig. 2. Simulated current density lateral distribution in the active region of a gain-guidedQCL as function of applied voltage (one transverse symmetric half is shown). Inset: halfwidth at half maximum (HWHM) of current density distribution as function of appliedvoltage (solid curve) for 1 μm of cladding thickness left unetched as in fabricated devices.Simulations for 2 μm (dashed) and 3 μm (dotted) provided for comparison as well.

Vol. 24, No. 17 | 22 Aug 2016 | OPTICS EXPRESS 19065

Page 4: Gain-guided broad area quantum cascade lasers emitting 23 ... · 14. A. Bismuto, S. Blaser, R. Terazzi, T. Gresch, and A. Muller, “High performance, low dissipation quantum cascade

0.0 0.5 1.0 1.5 2.0

Current, A

0

2

4

6

8

10

12

14

16

18

Voltage,

V

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

Totalpow

er(2

facets),

W

0 5 10 15 20

Current, A

0

5

10

15

20

25

Voltage,

V

0

5

10

15

20

25

Totalpow

er(2

facets),

W

Fig. 3. LIV (light-current-voltage) curves for index guided laser (left) and gain guided(right) at 20◦C, 300 ns pulse width, 1% duty cycle. Bold lines - measured power, thinnerlines - measured voltage, dashed - simulated voltage.

of same device geometry without gain variation: this resulted in high order mode operation.However, if then etching depth parameter was increased to reproduce the usual QCL geometrywith etched active region, horizontal index guiding and zero order mode selection were observedagain as expected, which became an additional proof of reliability of our model. A more precisemodel should have included thermally induced variation of refractive index profile during thecurrent pulse - that, however, would dramatically increase complexity of simulations.

18 20 22 24

Voltage, V

0

2

4

6

8

10

12

Net

modalgain,cm

−1

TM00

TM01

Fig. 4. Simulated modal gain as function of applied voltage in a gain-guided QCL. Highgain margin between the TM00 and TM01 modes makes first one always reach lasingthreshold first.

Finally, at 20 V applied voltage (corresponds to ≈ 14.3 A) far field of the laser was calculatedbased on the mode intensity distribution in the cross-section. A single gaussian-like lobe with56◦ vertical and 8◦ horizontal full angles of divergence was obtained. While vertical angle wasusual for a QCL [13], the horizontal one was few times smaller, which corresponded well to thestretched mode shape in the waveguide (Fig. 1), and what was observed later experimentally.

Vol. 24, No. 17 | 22 Aug 2016 | OPTICS EXPRESS 19066

Page 5: Gain-guided broad area quantum cascade lasers emitting 23 ... · 14. A. Bismuto, S. Blaser, R. Terazzi, T. Gresch, and A. Muller, “High performance, low dissipation quantum cascade

Fabrication

The QCL active region and top cladding were grown in a multi-wafer molecular beam epitaxy(MBE) reactor in a single growth step on n-doped substrates (InP:S, n ≈ 1−4 × 1017cm−3). Thegrowth started with a 1.5 μm-thick low doped InP layer (Si dopant, n = 3 × 1016 cm−3). The ac-tive region consisted of 35 stages (total thickness: 1.42 μm) of a strain-balanced In0.71Ga0.29As/Al0.75In0.25As (strain: +1.18%/-1.86%) two-phonon-resonant design with a sheet carrier den-sity per stage of 4.86 × 1010 cm−2. The upper cladding layer sequence, grown with MBE, frombottom to top was: InP (Si, 3 × 1016 cm−3) 1.5 μm, InP (Si, 1 × 1017 cm−3) 1.5 μm, InP (Si,8 × 1018 cm−3) 1.0 μm, InGaAs (Si, 2 × 1019 cm−3) 0.2 μm.

After this one of the wafers was processed in ridge-waveguide (RWG) configuration withridge width of 12 μm for cw operation using dry etching to obtain vertical sidewalls and thickgold deposition by electro-plating for improved heat dissipation. The results obtained with thiswafer were used as a reference point to evaluate the performance of the gain-guided devices.

Another wafer from the same epitaxial growth was processed in gain-guided broad area laserconfiguration using a process based on the one described in [14]. The upper cladding thicknesswas increased by growing with metalorganic chemical vapour deposition (MOCVD) additional4 μm of layers of InP (Si) with doping concentrations increasing from 2 × 1016 cm−3 at thebottom to 1× 1019 cm−3 at the top (layers between 9 and 13 μm of vertical axis on Fig. 1). Thenthe narrow current injection region was formed by etching highly conductive top cladding layerswith a wet solution of HBr/HNO3/H2O and regrowing the insulating InP:Fe. Several differentcurrent injector widths between 10 and 16 μm were fabricated on same wafer; this parametervariation did not affect laser performance. We left ≈ 1 μm of the low-doped (3 × 1016 cm−3)cladding layer (corresponds to 7 μm etching depth mentioned in simulation). The resultingdevice geometry corresponded to the one simulated (Fig. 1).

Both narrow ridge-waveguide and gain guided QCLs were cleaved in 6 mm-long chips andmounted epitaxial-side up on Cu carriers with In solder. Facets were left as cleaved.

Methods and experimental results

To optimally drive these lasers with high currents and voltages (up to 20 A and 25 V) in shortpulses (≥ 20 ns, ≈ 5 ns rise/fall), we have developed a special pulse driver board based onfast MOSFET (metal-oxide-semiconductor field-effect transistor) driven by a dedicated gatedriver (Fig. 5). Pulse amplitude was adjusted precisely with an onboard voltage regulator circuit;control of all operating parameters and pulse sequence as well as measurement of instantaneouslaser current and voltage was performed by an on-board MCU (micro controller unit). Thecompact size of the board allowed placing it very close to the laser, thus minimizing currentpath length and stray inductance, which is critical for keeping current pulse edges sharp.

All the measurements were taken at room temperature (20◦C) with 300 ns pulse width and1% duty cycle unless otherwise stated. Standard LIV (light-current-voltage) curves are shown inFig. 3. The peak power of a gain guided device reached 23.5 W at 20.4 A (limited by electronics,as laser rollover was not reached), which is over 7 times as large as maximum power of indexguided one at its rollover. 300 ns pulse width was chosen as an optimal value so that, on theone hand, measurements were not affected by ≈ 5 ns rise/fall times, and on the other - powerwas not reduced by heating. However, no change of LIV curves was observed for pulse widthsbetween 60 ns and 600 ns. Wall plug efficiency at these conditions reached 11% at 1.2 A forindex guided and 5.5% at 14 A for gain guided. At 20% duty cycle average total power emittedby gain guided device reached a maximum of 1.6 W. Cooling the lasers down to -30◦C ledto the power increase by 22% for index guided and 40% for gain guided compared to roomtemperature operation. Typical measured emission spectrum of the fabricated lasers is shown inFig. 6.

Vol. 24, No. 17 | 22 Aug 2016 | OPTICS EXPRESS 19067

Page 6: Gain-guided broad area quantum cascade lasers emitting 23 ... · 14. A. Bismuto, S. Blaser, R. Terazzi, T. Gresch, and A. Muller, “High performance, low dissipation quantum cascade

Fig. 5. Top: pulsed driver simplified schematics. Vin - input voltage, 1 - laser voltage con-trol, 2 - pulse generation, 3 - amplified pulses for driving MOSFET gate, 4 - laser voltagemeasurement, 5 - laser current measurement, 6 - external communication and synchroniza-tion, RS - sense resistor, AMP - operational amplifier. Solid line indicates the main currentpath, dashed ones - supplementary signals. Bottom: board photo, size: 88 × 33 mm.

4400 4450 4500 4550 4600 4650 4700

Wavelength, nm

0.0

0.2

0.4

0.6

0.8

1.0

Intensity,a.u.

Fig. 6. Typical measured emission spectrum of the fabricated lasers.

Vol. 24, No. 17 | 22 Aug 2016 | OPTICS EXPRESS 19068

Page 7: Gain-guided broad area quantum cascade lasers emitting 23 ... · 14. A. Bismuto, S. Blaser, R. Terazzi, T. Gresch, and A. Muller, “High performance, low dissipation quantum cascade

1.2 × Ith 1.9 × Ith 2.7 × Ith 3.4 × Ith

Fig. 7. Measured beam intensity profiles of a gain-guided QCL at a distance of 6 mm forvarious injection currents. Each image dimensions are 3.0 × 12.4 mm (initially 12.4 ×12.4, sides containing no signal were cropped). Ith ≈ 6 A.

The output beam was characterized using a pyroelectric camera with a 124 × 124 pixel, 100μm pitch focal plane array (FPA) and a mechanical chopper. The camera was first placed 6 mmaway from the laser facet to characterize the overall beam shape and check for the presence ofside lobes. In this configuration, the FPA covered an angular range of 90◦×90◦ with a resolutionof 0.95◦. Four beam pictures taken at various currents I between 1.2 × Ith and 3.4 × Ith areshown in Fig. 7. Duty cycle had to be lowered from 1% to 0.5% when reaching 1.9× Ith and thento 0.4% when reaching 3.4 × Ith in order not to saturate the camera sensor. As can be seen, thebeam consisted of a single bell-shaped lobe centered on the optical axis. In contrast to index-guided broad area QCLs [4] and phase-locked QCL arrays [7], no lateral lobes propagatingat large angles from the optical axis were observed. Furthermore, the beam retained its well-behaved shape up to the maximum current Imax = 3.4 × Ith .

After confirming the absence of side lobes, the camera was positioned 24 mm away from theoutput facet, in the far field of the laser, to measure the beam divergence. The intensity valuesmeasured by the individual camera pixels were summed column by column and fitted with agaussian line shape to determine the horizontal beam diameter (Fig. 8). Excellent agreementbetween the measured profile and the gaussian fit function was observed up to 2.5 × Ith . Abovethis value, the beam shape started to deviate from gaussian but remained single-lobe withoutdips. One of the reasons for that could be the thermally induced non-uniform increase of re-fractive index, which is stronger at higher currents and better confines modes, including highorder ones. In addition, when the electric field in the active region equals and then exceeds theresonant tunneling condition, optical gain saturates and then decreases with increasing currentdensity because most carriers are not injected into the upper laser level. This leads to a flatteningof the gain spatial profile and then in the emergence of a dip in the middle of the waveguide,resulting in a weaker guiding of the fundamental mode.

Beam propagation factor M2 was measured by collimating the beam with an aspheric lens(f = 1.873 mm) and measuring beam second moment width at different distances with samecamera. In the vertical direction M2

y = 1.271± 0.003, which is expectable [13]. However, in thehorizontal one the precise measurement could not be obtained with the available optics due to

Vol. 24, No. 17 | 22 Aug 2016 | OPTICS EXPRESS 19069

Page 8: Gain-guided broad area quantum cascade lasers emitting 23 ... · 14. A. Bismuto, S. Blaser, R. Terazzi, T. Gresch, and A. Muller, “High performance, low dissipation quantum cascade

−10 −5 0 5 10

Horizontal angle, ◦

0.0

0.5

1.0

1.5

2.0

Intensity,a.u.

20.4 A

18.0 A

14.8 A

11.6 A

8.8 A

8 12 16 20Current, I

2.5

4.5

6.5

θ d,◦

Fig. 8. Measured horizontal far field intensity of a gain-guided QCL at various currents(dots) along with gaussian fits (lines). Curves are shifted vertically with a step of 0.1 fordistinctness. Inset: half angle beam divergence (θd ) at 1/e2 of maximum value extractedfrom gaussian fits with black error bars representing standard deviation errors.

the strong ellipticity of the beam, and we could just estimate the M2x ≈ 6.

Additionally, a time evolution of far field profile of the beam was measured by capturingintensity variation at maximum current with a fast optical detector as a function of horizontalangle of observation (Fig. 9). The distance from laser facet to detector was 75 mm. Detectorsignal was captured with a fast oscilloscope synchronized with the laser pulse driver. It is clearlyvisible, that after first ≈ 40 ns of the pulse, where laser was operating in a pure zero order mode,higher order modes started to contribute, which resulted in beam steering and widening. Thisexplains a rather high value of measured M2

x . Nevertheless, the laser output averaged over time,like, for example, with camera measurements described above, remained gaussian-like singlelobe. Moreover, this lateral mode hopping behaviour has demonstrated a perfect pulse-to-pulserepeatability and high stability to variations of current and heatsink temperature, which couldbe another indication for that it is caused by thermal effects. At lower currents same pattern ofbeam variation was observed, but more stretched proportionally along time axis, in particularwith longer operation time in zero order time at the pulse start.

Conclusion

We have presented a novel configuration of broad area QCLs and did their extensive charac-terization and numerical modelling. Compared to a standard design having same active region,larger volume of gain medium pumped by a wide current distribution in this configuration hasenabled higher peak powers, while the corresponding shape of gain distribution provided effi-cient optical guiding. The lasers were driven with short high current pulses by a custom devel-oped pulsed driver. They demonstrate ≈ 7 times as high maximum power as ridge waveguidedevices with same active region design. Their output beam consists of a single lobe with a ratio

Vol. 24, No. 17 | 22 Aug 2016 | OPTICS EXPRESS 19070

Page 9: Gain-guided broad area quantum cascade lasers emitting 23 ... · 14. A. Bismuto, S. Blaser, R. Terazzi, T. Gresch, and A. Muller, “High performance, low dissipation quantum cascade

−10 −5 0 5 10Horizontal angle, ◦

0

50

100

150

200

250

300

Tim

e,ns

0.0

0.2

0.4

0.6

0.8

1.0

−10 −5 0 5 10Horizontal angle, ◦

0

50

100

150

200

250

300

Tim

e,ns

0 50 100 150 200 250 300Time, ns

3.5

5.5

7.5

θ d,◦

Fig. 9. Top left: measured instantaneous horizontal far field intensity evolution (color, a.u.)of a gain-guided QCL during 300 ns pulse at maximum current of 20 A showing transitionfrom TM00 to high order modes. Top right: same data averaged from 0 ns to a given time(50 ns, 100 ns, ...) shown with black dots along with gaussian fits (red lines). Bottom: halfangle beam divergence (θd ) at 1/e2 of maximum value extracted from gaussian fits withblack error bars representing standard deviation errors.

of ≈ 1 : 7 between horizontal and vertical axes, which can be efficiently reshaped into a circularone with, for instance, a pair of cylindrical lenses if required. Having such properties they canfind their applications in remote sensing, night vision and directional infrared countermeasuresystems.

Funding

European Commission Seventh Framework Programme (FP7 GA 607491 COMIQ).

Acknowledgements

Authors would like to thank Dr. Romain Terazzi for his help with the simulations.

Vol. 24, No. 17 | 22 Aug 2016 | OPTICS EXPRESS 19071