You are on page 1of 8

REVIEW ARTICLES | FOCUS

PUBLISHED ONLINE: 28 JUNE 2012 | DOI: 10.1038/NPHOTON.2012.143

Mid-infrared quantum cascade lasers


Yu Yao1†, Anthony J. Hoffman2†* and Claire F. Gmachl3

Mid-infrared quantum cascade lasers are semiconductor injection lasers whose active core implements a multiple-quantum-
well structure. Relying on a designed staircase of intersubband transitions allows free choice of emission wavelength and,
in contrast with diode lasers, a low transparency point that is similar to a classical, atomic four-level laser system. In recent
years, this design flexibility has expanded the achievable wavelength range of quantum cascade lasers to ~3–25  μm and the
terahertz regime, and provided exemplary improvements in overall performance. Quantum cascade lasers are rapidly becom-
ing practical mid-infrared sources for a variety of applications such as trace-chemical sensing, health monitoring and infrared
countermeasures. In this Review we focus on the two major areas of recent improvement: power and power efficiency, and
spectral performance.

Q
uantum cascade (QC) lasers, invented in 1994 by J.  Faist, the WPE remained of the order of 1% at room temperature for most
F. Capasso, and co-workers1, have progressed rapidly owing to QC lasers. A low WPE puts heavy demands on system designers;
their intrinsic design potential. These semiconductor injection for example, an application requiring 1 W of optical output power
lasers are based on intersubband transitions in multiple-quantum- from a laser with a WPE of 1% would require 100 W of input power.
well structures. The design of alternating wells and barriers, often The system requirements become even more demanding when one
numbering 500–1,000 (Fig.  1), allows great flexibility for a creative considers that any electrical energy not contributing to light output
engineer to explore2–9. Having exploited the benefits of mature, estab- must be dissipated as heat. As the number of proposed applications
lished materials — InGaAs/AlInAs alloys on InP and later GaAs/ for QC technology grew, so did the need for improving the WPE.
AlGaAs on GaAs — thanks to the communications and photonics Fig. 2a shows a plot of the reported WPE over time for both pulsed
industries, progress in the development of these new semiconductor and CW operation1,14–47.
lasers was determined first by the imagination of device researchers The WPE is a complicated parameter that depends on factors
and later by the pull of application requirements10. In the years fol- such as the temperature of the device, the quantum-mechanical
lowing the first demonstration of the QC laser 1, researchers repeat- structure of the energy levels and various device characteristics such
edly demonstrated the versatility enabled by quantum engineering, as length and waveguide loss48. The total efficiency of the device can
including demonstrations of lasers simultaneously emitting at mul- be approximated as the product of four device efficiencies: opti-
tiple wavelengths11, lasers with integrated sum-frequency nonlineari- cal, current, internal and voltage efficiencies. Optical efficiency
ties12 and lasers capable of broadband tuning 13. These new devices describes the emission of photons from the laser cavity before they
with extraordinary emission properties were created by changing are absorbed by the waveguide; current efficiency describes how far
the quantum-mechanical structure of the active region. Progress in above threshold the laser is operated; internal efficiency describes
output power and operating temperature was also made in conven- the properties of the quantum design, such as the radiative and
tional devices as the active region designs were refined14–19. Before non-radiative recombination rates of the designated optical tran-
2002, when continuous-wave (CW) operation at room temperature sition, thermal backfilling of the lower laser level and thermionic
was achieved, QC lasers were limited to low-duty-cycle pulsed opera- emission out of the quantum structure into the continuum states;
tion at room temperature20. However, even as device performance and voltage efficiency describes the fraction of the voltage drop
improved, most lasers exhibited very low (<1%) conversion efficien- over the device that contributes to photon generation48. Designing
cies between electrical power and optical power — a metric known lasers that maximize these four efficiencies is essential to improving
as the wall plug efficiency (WPE). Low WPEs limited the use of QC device performance.
lasers in applications such as portable sensors and infrared counter- These four device efficiencies are all functions of the photon
measures, where the power consumption of the laser was a major con- energy, which makes the WPE dependent on the wavelength. A
straint. For applications already employing QC lasers, high WPE was reduction in the WPE for longer-wavelength lasers was dem-
seen as a desirable improvement. Although low threshold and high onstrated both experimentally and theoretically by considering
optical output power were always clear goals (or at least desirable fea- the effect of wavelength on physical quantities such as oscillator
tures), only over the past five years has the power efficiency become strength, free-carrier absorption and optimal energy separation
an essential device parameter. This Review focuses first on power effi- between the lower laser level and the injector state48,49. The majority
ciency, as this generally also results in high optical output power and of the effort towards improving the WPE has been concerned with
is usually concomitant with a low-threshold current density. devices operating at wavelengths around 4.6 μm. Devices designed
for this wavelength window are driven primarily by security appli-
High power and power efficiency performance cations such as infrared countermeasures, and benefit from high-
The first QC laser had a WPE of less than 0.15% in pulsed operation quality growth of strain-balanced, InP-based heterostructures.
at 10 K, meaning that less than 0.15% of the input electrical power These are the devices that are emphasized in this Review.
was converted to optical power from the device1. Although WPE Significant advances in the WPE were achieved by further explor-
values improved slightly as better lasers were designed, little work ing the design space and re-examining common design strategies.
was done with the principal aim of improving the parameter, and One strategy focused on reducing the voltage defect, the voltage

1
School of Engineering and Applied Science, Harvard University, Cambridge, Massachusetts 02138, USA. 2Department of Electrical Engineering, University
of Notre Dame, Notre Dame, Indiana 46556, USA. 3Department of Electrical Engineering, Princeton University, Princeton, New Jersey 08544, USA. †These
two authors contributed equally to this work, and significantly more so than the third author. *e-mail: ajhoffman@nd.edu

432 NATURE PHOTONICS | VOL 6 | JULY 2012 | www.nature.com/naturephotonics


© 2012 Macmillan Publishers Limited. All rights reserved.
FOCUS | REVIEW ARTICLES
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2012.143 FOCUS | REVIEW ARTICLES
PUBLISHED ONLINE: 28 JUNE 2012 | DOI: 10.1038/NPHOTON.2012.143

a c

Injector
Electroplated gold
10 μm Active
region
SiN insulation
Gap in plating to Injector
facilitate cleaving
3

b Active core

Active

520 meV
region

60 nm

Figure 1 | Concept of a QC laser. a, Photograph of a laser bar with four QC lasers (left, courtesy of Frank Wojciechowski) and scanning electron microscopy
image of the front facet of a QC laser (right). b, High-resolution transmission electron microscopy image of a QC laser, showing four periods of active
regions and injectors. c, Simplified schematic of the conduction band structure for a basic QC laser, where the laser transition is between sub-bands 3 and 2.

drop in the active region that does not contribute to the genera- most clearly observed as an exponential increase in the threshold
tion of light34,39,43. One particularly successful design achieved this current and, somewhat less strongly, as an exponential decrease in
at both the laser turn-on and turn-off voltages43; by ensuring a low the optical slope efficiency of the laser. As a result, both the current
voltage defect at these two operating points, the overall electrical efficiency and internal efficiency are negatively impacted. The pri-
efficiency of the device was high and the differential resistance of mary causes for this reduction in laser performance are a reduction
the device was low. This study also achieved high differential gain in the gain of the laser due to backfilling of the lower laser level,
by employing a short period length, which allowed denser stack- thermionic emission from bound states to the continuum, and
ing of the amplifying quantum wells. The high differential gain increased phonon scattering rates53. Despite these additional chal-
enabled the fabrication of devices with larger mirror losses — and lenges, appropriate active-region designs can be used to mitigate
hence increased optical efficiency — while minimizing the impact some of the effects of higher operating temperatures.
of increased photon emission on the threshold current. This device The design considerations for lasers intended to operate at high
achieved a record WPE of more than 50% at 40 K in pulsed opera- temperatures or in CW mode are necessarily different from those
tion, thus generating more light than heat 43. operating at low temperatures. For example, small voltage defects
A second strategy developed in tandem with the above approach such as those used in pulsed, low-temperature lasers to achieve
also achieved a pulsed WPE of 50% under similar operating con- record WPEs are not viable design choices because the voltage
ditions by compensating for the effect of interface roughness on defect plays an important role in minimizing thermal back-filling
the transport of electrons inside the device44. Theoretical models of the lower laser level53,54. Instead, the active core must incorporate
using interface roughness as the relevant in-plane scattering time a larger voltage defect for devices operating at higher temperatures.
showed that broadening due to interface roughness was the lim- Strategies to increase the performance of these devices are multi-
iting factor in resonant tunnelling between the injector and the faceted: reduce the sensitivity of the laser to temperature changes
upper laser level, thus making it the limiting factor in laser per- and minimize temperature changes due to self-heating 42,55,56.
formance50. These results suggested that increasing the coupling Reducing the temperature sensitivity of QC laser performance
strength between the injector and the upper laser level would has been achieved using a variety of active-region designs. One par-
result in improved gain, which contradicted the previous notion ticularly effective design incorporates two strategies to achieve high
that increased coupling would reduce laser performance due to a performance. First, shallow wells and barriers are used to increase
broader gain spectrum. The conduction band energy diagram for the separation between the upper laser level and the energy state
such a device is shown in Fig. 2b. Light-current–voltage measure- immediately above it, thus reducing carrier injection into unwanted
ments and current-versus-WPE for an ultra-strong coupling laser states. Second, tall barrier inserts are used in the injector region to
developed by Liu et al.44 are shown in Fig. 2c and Fig. 2d, respec- increase the confinement of the optical transition and minimize
tively. These ultra-strong coupling lasers consistently achieved carrier loss into the continuum states. The effect of temperature
WPEs in the range of 40–50% below 160 K in pulsed mode, owing on the threshold of a QC laser can be described by J(T)  =  J0eT/T0,
to improvements in both the current and internal efficiencies44. where J0 is a constant and T0 is the characteristic temperature of
This new understanding of transport in QC lasers is significant to the device. High values of T0 are therefore desirable. This device has
the field because the results are broadly applicable and should lead a characteristic temperature (T0 ~ 383 K) that is greater than that
to improvements in most active-core designs at all temperatures50. of high-performance QC lasers incorporating more conventional
The aforementioned designs achieved record WPEs by operat- designs (T0 ~ 200 K)55. The highest-performing QC laser operating
ing at low temperatures in low-duty-cycle pulsed mode. However, at room temperature in CW mode has been demonstrated using this
many applications require lasers that operate at room temperature shallow-well architecture; the device operated with a peak WPE of
or in CW mode. Such requirements result in higher active-core 21% and had a maximum output power of 5.1 W (ref. 46). In addi-
temperatures due to an increase in heat sink temperature and/or tion to reducing the temperature sensitivity of the laser, these high-
self-heating 51–53, causing a degradation in laser performance that is performance devices also benefit from superb thermal packaging.

NATURE PHOTONICS | VOL 6 | JULY 2012 | www.nature.com/naturephotonics 433


© 2012 Macmillan Publishers Limited. All rights reserved.
REVIEW ARTICLES | FOCUS NATURE PHOTONICS DOI: 10.1038/NPHOTON.2012.143

a b 1,800
5 E = 102 kV cm−1
CW
40
45
1,400
20 85
Wall plug efficiency (%)

19 meV

Heat sink temperatue (K)


125
0 1,000

Energy (meV)
165

40 Pulsed 205
600
245
20
285
200
0 320
1993 1996 1999 2002 2005 2008 2011 0 15 30 45 60 75
Year Distance (nm)

c Current density (kA cm−2) d Current density (kA cm−2)


0 1 2 3 4 5 0 1 2 3 4 5
16 12
9K 50 9K
14 80 K
80 K 10
120 K 120 K
12 160 K 40 160 K
200 K 8 200 K
10
Voltage (V)

Power (W)

250 K 250 K
300 K 30 300 K
6
8
20
6 4

4 10
2
2
0 0
0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 1.5 2.0
Current (A) Current (A)

Figure 2 | WPE. a, Total WPE of selected lasers from the literature at various heat sink temperatures for pulsed and CW operation. Devices were selected
from highly cited articles in years 1994–2012. The colour of the symbol indicates the temperature of the heat sink. Symbols outlined in black represent
devices with one or both facets coated. For non-coated lasers, the output power from a single facet was doubled to obtain the total device efficiency. For
research that did not directly report the WPE, the value was estimated using the available data; estimates were required for many of the early devices. The
error for the estimated WPE is primarily due to an uncertainty in the voltage applied to the device at peak WPE, which is estimated to be about 30%. In
most cases, the estimated WPE is larger than the actual WPE because the ‘turn-on’ voltage, which is lower than the actual voltage applied to the device,
was used for the calculation. b, Conduction-band diagram for an ‘ultrastrong coupling’ QC laser. c, Pulsed light-current–voltage measurements for an
‘ultrastrong coupling’ laser at various heat sink temperatures. The device was 13.6 μm × 2.9 mm and the total output power was obtained by doubling the
output power from a single facet. d, WPE versus current for the device shown in b. Figure b–d reproduced with permission from ref. 44, © 2010 NPG.

A complementary strategy that has consistently been employed in and thus the power budget is limited. In such cases, it is desirable to
high-performance lasers is to minimize the temperature change by scale down the input and output powers of these high-performance
effectively removing heat from the active core35,37,41,42,46. The active core devices while maintaining a high WPE. Low-power-consumption
can be subject to significant heating during high-duty-cycle or CW QC lasers are realized using short, narrow cavities with low-power
operation, and proper thermal management is essential for ensuring active regions to ensure that the total power required for lasing
high-performance operation57. Laser performance when operating in is low 59–62. Because shorter cavities have larger photon emission
CW mode is greatly improved by fabricating the devices as buried- rates, the mirror loss of the laser must be modified to maintain a
ridge lasers using high-quality processing, employing proper growth low threshold current. One strategy for achieving this is to use high
conditions and using an InP regrowth technique (owing to the high reflectance and partial-high-reflectance coatings on the back and
thermal conductivity of InP)20,35,36,58. Mounting the device epitaxial- front facets, respectively. Such lasers have been demonstrated in
side-down on diamond or some other high-thermal-conductivity CW mode at room temperature with a threshold power consump-
heat sink further improves heat removal from the active core by mini- tion of 0.83  W. Conventional designs, in contrast, have threshold
mizing the thermal resistance between the heat-generating active core power consumptions of 1.5–3  W (ref.  60). Recently, even lower
and the heat-removing heat sink35. Finally, external cooling mecha- powers were reached by optimizing the doping density in the active
nisms such as thermoelectric coolers can be used to extract heat from region. These devices were fabricated as short-cavity, distributed
the package and regulate the temperature of the active core. feedback (DFB) lasers with antireflection coatings on the front and
A direct way of minimizing heating is to reduce the amount back facets to suppress Fabry–Pérot modes. These devices have
of power required by the laser. This strategy is particularly useful threshold power consumptions of around 0.7 W when operated in
for applications such as point sensors, where optical powers of the CW mode at room temperature. Their emission spectrum is single
order of a few milliwatts are sufficient, or for applications such as mode with a side-mode suppression ratio of 30  dB, which makes
portable network sensors, where the laser is powered by a battery them particularly useful for portable spectroscopic applications61.

434 NATURE PHOTONICS | VOL 6 | JULY 2012 | www.nature.com/naturephotonics


© 2012 Macmillan Publishers Limited. All rights reserved.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2012.143 FOCUS | REVIEW ARTICLES
a c 1.0

5
10 μm 1 mm Experiment Simulation
Surface grating

4
0.8

Intensity (normalized)
3
Highly doped cap layer
Upper cladding

2
Active region 0.6

Angle (°)
1
Bottom cladding

−5 −4 −3 −2 −1 0
Substrate
0.4
Buried grating Surface grating II
0.2

0
−5 −4 −3 −2 1 0 1 2 3 4 5
Angle (°)

b d
Laser waveguide top view

θ
Laser ridge
L+
Air hole P−
P+
1 mm L−

Laser facet
Intensity (log scale)

1
e
0.1 DFB master oscillator Single-pass power amplifier

0.01

1,200 1,150 1,100 1,050


Wavenumber (cm−1) Output facet with antireflection coating

Figure 3 | DFB QC lasers. a, Schematic of DFB QC laser structures. b, Scanning electron microscopy picture of a DFB QC laser array90 and output laser
spectra88. c, Scanning electron microscopy image of a two-dimensional ring cavity DFB QC laser array (top-left) and the far-field of the laser output beam
(bottom-right)89. d, Schematic of a photonic-crystal DFB QC laser. e, Schematic of a QC laser master oscillator power amplifier with simulated light
intensity distribution92. Figure reproduced with permission from: b (spectra) ref. 88, © 2009 IEEE; b (inset), ref. 90, © 2010 SPIE; c (left), ref. 89, © 2011
AIP; c (right), ref. 77, © 2010 AIP; e, ref. 92, © 2011 OSA.

High spectral performance CW operation at room temperature. An alternative surface grating


For many applications in spectroscopy, particularly trace-gas technique is to remove the metal from the grooves (Fig. 3a, bottom-
sensing, single-mode narrow-linewidth tunable mid-infrared right), which reduces the loss at the metal–semiconductor interface77.
lasers are desirable to achieve high selectivity, high sensitivity and The first room-temperature CW DFB QC laser was dem-
multiple-species detection. The most common approaches for onstrated using a buried grating in 200571. Since then, steady
achieving tunable single-mode operation are DFB QC lasers63,64 improvements have been made towards higher output powers,
and external cavity (EC) QC lasers65. Versatile cavity designs66–70 high-temperature operation and low power consumption72,78–81.
have also shown the potential to achieve single-mode lasing spec- Buried gratings were generally regarded as the only way to achieve
tra. This section focuses on DFB QC lasers and EC QC lasers, both room-temperature CW DFB QC lasers until recently, when the
of which have shown rapid progress in terms of laser performance design principle of surface grating was revisited82. Taking into
and tuning range. account the coupling between the dielectric waveguide mode and
DFB QC lasers are fabricated by integrating a Bragg grating into the grating surface plasmonic mode allowed researchers to predict
the laser waveguide. Repeated scattering from a Bragg grating favours the grating coupling coefficient and waveguide loss more accurately
a single wavelength — the Bragg wavelength — which is determined than previous models based on coupled-mode theory 63,67. The opti-
by the grating period. There are two ways to introduce gratings into mized structure exhibits a shallow grating with a complex-coupling
QC laser waveguides: a surface grating in the highly doped cap layer coefficient 83–86. Researchers have not only demonstrated room-tem-
(Fig. 3a, top)63; or a buried grating close to the active region, sand- perature CW operation, but also boosted the output power level
wiched between the waveguide core and the top cladding layers to above the watt-level47,86. However, the weak coupling strength
(Fig. 3a, bottom-left)64,71,72. Applying a grating to the laser ridge side- requires a high-quality antireflective coating on at least one facet
walls has also been attempted73–75 but is not as common. A surface to maintain single-mode operation at injection currents high above
grating is easier to fabricate than a buried grating, which requires the lasing threshold85.
high-quality upper waveguide cladding regrowth after fabrication. Another important feature of the DFB QC laser is that its emis-
However, surface gratings require larger grating depths to achieve sion wavelength can be continuously tuned by changing the heat
the necessary coupling strength, as the grating is further away from sink temperature or injection current 79,85, which in turn heats the
the laser mode centred in the active region. Furthermore, this type laser. Both methods of tuning lead to a change in the effective
of surface grating provides complex coupling; that is, a modulation refractive index of the laser waveguide and thus shift the reso-
in both refractive index and loss16,76. Although this helps to break the nance wavelengths of the Bragg gratings, with a tuning rate of
degeneracy of the two Bragg resonance modes and thus enforce sin- around 0.1–0.2 cm–1 K–1. This capability allows DFB QC lasers to be
gle-mode operation, the additional loss makes it difficult to achieve tuned across the absorption peaks of different gas species, thereby

NATURE PHOTONICS | VOL 6 | JULY 2012 | www.nature.com/naturephotonics 435


© 2012 Macmillan Publishers Limited. All rights reserved.
REVIEW ARTICLES | FOCUS NATURE PHOTONICS DOI: 10.1038/NPHOTON.2012.143

a b c Wavelength (μm)
12 11 10 9 8 7 6 5 4
0.2 2012
I E
DFB TPR
112
RDFB CB89 EC DAU
111
0.1 2011 EC CC108
M
5 EC DAU 111

0 2010 EC CB106

Energy (eV)
4 DFB BC88
EC TPR101
−0.1 1 2 3 2009 EC BC 5 stack105
CL lens
EC TPR97 EC BC97

Year
−0.2 2008
QCL DFB BC 115
2007 EC BC113
−0.3
0
2006 EC BC 2 stack
94
GR
−0.4
EC BC102
600 800 1,000 1,200 2005
EC BC114
d Distance (Å) 2004
10 5 CO NH3
1.0 CO2 NO
8 4 H2O NO2

Gain (normalized)
0.8
O3 N2O

(cm/mole ×e−19)
Intensity (a.u.)

6 0.6 3 CH4 SO2

Intensity
C2H4
4 0.4 2
0.2
2 1
0
0 0
1,950 2,000 2,050 2,100 2,150 2,200 2,250 1,000 1,500 2,000 2,500
Wavenumber (cm−1) Wavenumber (cm−1)

Figure 4 | EC QC lasers, broadband QC laser design and wavelength tuning ranges achieved in QC lasers. a, Schematic of EC QC laser in a ‘Littrow’
configuration (M: mirror; GR: grating reflector; CL: collimating lens). b, Conduction-band diagram of the continuum-to-continuum QC laser design45.
c, Top: Wavelength tuning ranges achieved with DFB arrays and EC lasers based on different QC laser designs. Red indicates CW operation; blue indicates
pulsed operation. Bottom: HITRAN simulated absorption spectra for several molecules in the mid-infrared wavelength range of 4–12 μm. TPR, BC, CB, CC
and DAU are different types of QC laser designs: TPR, two-phonon resonance; BC, bound-to-continuum; CB, continuum-to-bound; CC, continuum-to-
continuum; DAU, dual upper state. EC, DFB and RDFB are tuning methods of laser spectra: EC, external cavity; DFB, linear DFB laser array; RDFB, DFB ring
laser array. d, EC tuning laser spectra of the continuum-to-continuum QC laser and laser gain spectra at the same biased electric field108. The intensity of
the highest frequency component (grey) is magnified by a factor of 1,000 for clarity. Figure b reproduced with permission from ref. 45, © 2010 AIP.

providing high sensitivity in the detection of trace gases. The maxi- fabrication accuracy, this method is a very good solution for high-
mum tuning range of a single DFB QC laser is limited, however, by power brightness, single-mode lasers in pulsed mode operation.
both the technologically feasible temperature swing and the tuning However, the wide ridges lead to severe heating in CW operation
mismatch between the laser mode and the peak gain. Fully exploit- and are therefore unsuitable for CW room-temperature lasers
ing the gain bandwidth of a QC gain chip requires multiple DFB QC given the performance of today's laser gain chips. The QC laser
lasers with different grating periods to cover a broader wavelength master oscillator power amplifier incorporates a DFB QC laser as
range87. By integrating an array of 24 DFB lasers, a wavelength the master oscillator and a power amplifier at the output 92 (Fig. 3e).
tuning range of 220  cm−1 (8.0–9.8  μm) has been demonstrated The two sections are separately pumped, which allows the DFB
(Fig. 3b)88 using a bound-to-continuum active-region design. Using laser to be operated with moderate injection current to maintain
a two-dimensional array of ring cavity second order DFB QC lasers, single-mode emission while ensuring a large injection current in
a wavelength coverage of 180 cm−1 around 8.2 μm (1,220 cm–1) was the amplifier section. A high-quality antireflective coating is essen-
achieved with a surface-emitting laser beam at a very small diverg- tial for increasing the self-lasing threshold and allowing higher
ing angle of less than 3º (Fig. 3c)89. The availability of such on-chip transmission gain in the amplifier region.
widely tunable single-laser sources provides great opportunity for Although DFB QC lasers provide compact, reliable single-mode
achieving ultracompact mid-infrared spectrometers90. tunable laser emission, EC QC lasers are the most commonly used
Photonic-crystal DFB QC lasers91 and QC laser master oscil- widely tunable single-mode sources across numerous successful
lator power amplifiers92 have been employed to meet the require- applications94–100, owing to their rapid progress in tunability, output
ments of high peak power (tens of watts) and high brightness for power 101 and CW operation at room temperature101,102. Figure  4a
particular applications, with promising results. Figure  3d shows shows one typical configuration — the ‘Littrow’ configuration — for
the schematic of a photonic-crystal DFB QC laser, in which a pho- EC QC lasers. The set-up is carefully aligned to ensure that the col-
tonic crystal lattice is incorporated into the laser waveguide at a limated laser output beam is incident on the diffraction grating. The
tilted angle θ with respect to the laser facets. The four coupling first-order diffraction is then coupled back into the laser cavity while
directions (P+, P−, L+ and L−) cause the laser beam to propagate the zeroth-order diffraction provides the output laser beam. This sys-
in a zigzag pattern. The DFB selects the resonance wavelength and tem is essentially a coupled cavity system formed by two cavities; that
the lateral coupling helps to establish lateral coherence over a wide is, the Fabry–Pérot cavity formed by the laser’s front and back fac-
ridge width (several hundred micrometers, or even up to 1 mm)93. ets, and the grating cavity formed by the back facet and the grating.
A nearly diffraction-limited beam has been obtained with a total Detailed analyses have been performed into laser mode selection in
peak power of 34 W from a 400-μm-wide laser ridge91. With high such a configuration103,104.

436 NATURE PHOTONICS | VOL 6 | JULY 2012 | www.nature.com/naturephotonics


© 2012 Macmillan Publishers Limited. All rights reserved.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2012.143 FOCUS | REVIEW ARTICLES
By rotating the angle of the grating, one can change the grat- strongly coupled upper laser states and a very broad gain spectra
ing reflection spectra and thus tune the laser emission wavelength. (570 cm–1, around 40% of the emission frequency)109,110. Broad tun-
However, this results in step-wise tuning to discrete wavelengths at ing ranges have been demonstrated in both pulsed (321 cm–1) and
which the Fabry–Pérot cavity and grating-cavity mode coincide65. CW operation (248 cm–1)111.
Achieving continuous tuning requires complete mode tracking. A Figure 4c presents a summary of recent results for single-mode
piezo-activated mode-tracking system that provides independent broad-tuning QC lasers using an EC configuration or DFB arrays
control of the EC length and diffraction grating angle was imple- based on different QC laser designs88,89,94,97,101,102,104–106,108,111–115. Given
mented to achieve mode-hop-free tuning of EC QC lasers97,103. A the recent progress on broadband gain designs and tuning tech-
spectral resolution of <0.001 cm–1 (<30 MHz, limited by the laser niques, we believe there is still much room for expanding the tuning
linewidth) was obtained in CW operation, which is enough to range and improving the laser performance for single-mode tunable
resolve the narrow absorption spectra of trace gases97. mid-infrared light sources based on QC structures.
A major advantage of EC QC lasers is their potential to cover a
wavelength range as wide as that allowed by the gain medium using Outlook
a single QC laser chip. The maximum tuning range of a QC laser This Review has highlighted the rapid progress of QC lasers in areas
achieved so far is more than 430  cm–1 — over 39% of the centre of power efficiency, optical output power and spectral performance,
frequency — in an EC configuration employing a heterogeneous as determined by wide single-mode tunability. Canvassing the field
QC laser design105. The key to a wide tuning range, besides an opti- of mid-infrared QC laser research, an obvious speculation concerns
mized, well-aligned EC set-up, is a broad gain bandwidth. However, the areas of major future progress of the field. Several open ques-
the conventional QC laser gain medium has an intrinsically narrow tions invite major research investment. Fundamental issues include
linewidth because it is based on the intersubband transition. The the development of sub-picosecond, high-pulse-energy pulsed QC
heterogeneous design covers a broad wavelength range by integrat- lasers, and the development of QC lasers for (or, more specifically,
ing many sub-stacks at different wavelengths. This technique has the application of designed intersubband transitions to) materials
been used to achieve a supercontinuum lasing spectrum of 6–8 μm systems that are very different from conventional group iii–v semi-
in a single device13, thanks to its inhomogeneous broadening nature. conductor alloys. Applied challenges include developing QC lasers
This is an advantage for applications that require broadband laser whose power specifications are suitable for consumer electronics.
spectra. However, for single-mode widely tunable sources, a homo- Such achievements would highlight the significant potential of QC
geneously broadened gain medium is preferred because it helps to lasers and open up new fields for research and development.
maintain the single-mode emission at the selected wavelength, even
at high pumping currents. Another issue with the heterogeneous References
cascade design is that it sacrifices the laser gain at individual wave- 1. Faist, J. et al. Quantum cascade laser. Science 264, 553–556 (1994).
lengths because the confinement factor for each sub-stack becomes 2. Xie, F. et al. Room temperature CW operation of short wavelength quantum
cascade lasers made of strain balanced GaxIn1–xAs/AlyIn1–yAs material on InP
smaller as the number of sub-stacks increases, while the total core substrates. IEEE J. Sel. Top. Quant. 17, 1445–1452 (2011).
thickness of the laser is roughly maintained. Several techniques 3. Williams, B. S. Terahertz quantum-cascade lasers. Nature Photon. 1,
have been developed during the search for a homogeneously broad- 517–525 (2007).
ened QC gain medium with a wide gain spectra and high laser per- 4. Ulrich, J., Kreuter, J., Schrenk, W., Strasser, G. & Unterrainer, K. Long
formance, including bound-to-continuum19, continuum-to-bound/ wavelength (15 and 23 μm) GaAs/AlGaAs quantum cascade lasers.
continuum45,106 and dual-upper laser state designs107. Appl. Phys. Lett. 80, 3691–3693 (2002).
5. Revin, D. G. et al. InP-based midinfrared quantum cascade lasers for
Bound-to-continuum QC lasers have multiple transitions from wavelengths below 4 μm. IEEE J. Sel. Top. Quant. 17, 1417–1425 (2011).
the upper laser state to several lower laser states19. This design 6. Faist, J. et al. Short wavelength (λ ~ 3.4 μm) quantum cascade laser based on
simultaneously provides a short upper laser lifetime and a short strained compensated InGaAs/AlInAs. Appl. Phys. Lett. 72, 680–682 (1998).
depopulation time from the lower laser state — a compromise that 7. Colombelli, R. et al. Far-infrared surface-plasmon quantum-cascade lasers at
leads to a laser performance that comparable to the conventional 21.5 μm and 24 μm wavelengths. Appl. Phys. Lett. 78, 2620–2622 (2001).
two-phonon resonance design at longer wavelengths (>7–8  μm). 8. Cathabard, O., Teissier, R., Devenson, J., Moreno, J. C. & Baranov, A. N.
Quantum cascade lasers emitting near 2.6 μm. Appl. Phys. Lett. 96,
For wavelengths shorter than 5  μm, however, the smaller energy 141110 (2010).
gap between the injector states and the continuum above the quan- 9. Bismuto, A., Beck, M. & Faist, J. High power Sb-free quantum cascade laser
tum wells becomes detrimental to laser performance because the emitting at 3.3 μm above 350 K. Appl. Phys. Lett. 98, 191104 (2011).
upper laser state is close to the band edge of the barriers, which 10. Curl, R. F. et al. Quantum cascade lasers in chemical physics.
leads to a higher probability of carrier leakage into the continuum Chem. Phys. Lett. 487, 1–18 (2010).
states. Another challenge for achieving broad tuning in lasers based 11. Tredicucci, A. et al. A multiwavelength semiconductor laser. Nature 396,
350–353 (1998).
on bound-to-continuum design is the potentially unsymmetrical, 12. Owschimikow, N. et al. Resonant second-order nonlinear optical processes in
multi-peaked shape of the gain spectrum, due to the difference in quantum cascade lasers. Phys. Rev. Lett. 90, 043902 (2003).
oscillator strengths for individual transitions. In QC laser designs 13. Gmachl, C., Sivco, D. L., Colombelli, R., Capasso, F. & Cho, A. Y. Ultra-
based on continuum-to-bound106 and continuum-to-continuum45 broadband semiconductor laser. Nature 415, 883–887 (2002).
(Fig.  4b) active regions, ultra-strong coupling between injector 14. Faist, J. et al. High power mid-infrared (λ > 5 μm) quantum cascade lasers
states and the upper/lower laser states provides multiple laser tran- operating above room temperature. Appl. Phys. Lett. 68, 3680–3682 (1996).
15. Scamarcio, G. et al. High-power infrared (8-micrometer wavelength)
sitions, which contribute to a broadband gain spectrum. The strong superlattice lasers. Science 276, 773–776 (1997).
coupling facilitates ultrafast carrier transport from the injector to 16. Sirtori, C. et al. Mid-infrared (8.5 μm) semiconductor lasers operating at
the upper laser states, which compensates for the decrease in gain room temperature. IEEE Photon. Tech. Lett. 9, 294–296 (1997).
coefficient as a result of the broadening. Continuum-to-bound106 17. Tredicucci, A. et al. High performance interminiband quantum cascade lasers
and continuum-to-continuum45 designs are therefore promising with graded superlattices. Appl. Phys. Lett. 73, 2101–2103 (1998).
techniques for achieving high-performance QC lasers with a broad 18. Gmachl, C. et al. High temperature (T ≥ 425K) pulsed operation of quantum
cascade lasers. Electron. Lett. 36, 723–725 (2000).
gain spectrum. The homogeneous broadening mechanism is verified 19. Faist, J., Beck, M., Aellen, T. & Gini, E. Quantum-cascade lasers based on a
through the broad EC tuning range (350 cm–1, 90% of the gain spec- bound-to-continuum transition. Appl. Phys. Lett. 78, 147–149 (2001).
tral full width at half maximum, Fig. 4d) and strong gain-clamping 20. Beck, M. et al. Continuous wave operation of a mid-infrared semiconductor
effect 108. The dual-upper laser state design is characterized by two laser at room temperature. Science 295, 301–305 (2002).

NATURE PHOTONICS | VOL 6 | JULY 2012 | www.nature.com/naturephotonics 437


© 2012 Macmillan Publishers Limited. All rights reserved.
REVIEW ARTICLES | FOCUS NATURE PHOTONICS DOI: 10.1038/NPHOTON.2012.143

21. Faist, J. et al. Continuous-wave operation of a vertical transition quantum 51. Hoffman, A. J. et al. Lasing-induced reduction in core heating in high wall
cascade laser above T=80 K. Appl. Phys. Lett. 67, 3057–3059 (1995). plug efficiency quantum cascade lasers. Appl. Phys. Lett. 94, 041101 (2009).
22. Sirtori, C. et al. Quantum cascade laser with plasmon-enhanced wave-guide 52. Howard, S. S., Liu, Z. J. & Gmachl, C. F. Thermal and stark-effect roll-over of
operating at 8.4 μm wavelength. Appl. Phys. Lett. 66, 3242–3244 (1995). quantum-cascade lasers. IEEE J. Quant. Electron. 44, 319–323 (2008).
23. Faist, J. et al. High-power continuous-wave quantum cascade lasers. 53. Howard, S. S. et al. High-performance quantum cascade lasers: Optimized
IEEE J. Quant. Electron. 34, 336–343 (1998). design through waveguide and thermal modeling. IEEE J. Sel. Top. Quant. 13,
24. Page, H. et al. High peak power (1.1W) (Al)GaAs quantum cascade laser 1054–1064 (2007).
emitting at 9.7 μm. Electron. Lett. 35, 1848–1849 (1999). 54. Maulini, R., Lyakh, A., Tsekoun, A. & Patel, C. K. N. λ ~ 7.1 μm quantum
25. Slivken, S., Matlis, A., Rybaltowski, A., Wu, Z. & Razeghi, M. Low-threshold cascade lasers with 19% wall-plug efficiency at room temperature. Opt.
7.3 μm quantum cascade lasers grown by gas-source molecular beam epitaxy. Express 19, 17203–17211 (2011).
Appl. Phys. Lett. 74, 2758–2760 (1999). 55. Fujita, K. et al. Broad-gain (Δλ/λ0 ~ 0.4), temperature-insensitive (T0 ~ 510K)
26. Tredicucci, A. et al. High-performance quantum cascade lasers with quantum cascade lasers. Opt. Express 19, 2694–2701 (2011).
electric-field-free undoped superlattice. IEEE Photon. Tech. Lett. 12, 56. Shin, J. C. et al. Highly temperature insensitive, deep-well 4.8 μm
260–262 (2000). emitting quantum cascade semiconductor lasers. Appl. Phys. Lett. 94,
27. Green, R. P. et al. Room-temperature operation of InGaAs/AlInAs quantum 201103 (2009).
cascade lasers grown by metalorganic vapor phase epitaxy. Appl. Phys. Lett. 57. Chaparala, S. C., Xie, F., Caneau, C., Zah, C. E. & Hughes, L. C. Design
83, 1921–1922 (2003). guidelines for efficient thermal management of mid-infrared quantum
28. Yu, J. S., Slivken, S., Evans, A., Doris, L. & Razeghi, M. High-power cascade lasers. IEEE T. Compon. Pack. T. 1, 1975–1982 (2001).
continuous-wave operation of a 6 μm quantum-cascade laser at room 58. Beck, M. et al. Buried heterostructure quantum cascade lasers with a large
temperature. Appl. Phys. Lett. 83, 2503–2505 (2003). optical cavity waveguide. IEEE Photon. Tech. Lett. 12, 1450–1452 (2000).
29. Evans, A. et al. High-temperature, high-power, continuous-wave operation of 59. Blaser, S. et al. Low-consumption (<2W) continuous-wave singlemode
buried heterostructure quantum-cascade lasers. Appl. Phys. Lett. 84, quantum-cascade lasers grown by metal-organic vapour-phase epitaxy.
314–316 (2004). Electron. Lett. 43, 1201–1202 (2007).
30. Yu, J. S., Slivken, S., Darvish, S. R. & Razeghi, M. Short wavelength 60. Bai, Y., Darvish, S. R., Bandyopadhyay, N., Slivken, S. & Razeghi, M.
(λ ~ 4.3 μm) high-performance continuous-wave quantum-cascade lasers. Optimizing facet coating of quantum cascade lasers for low power
IEEE Photon. Tech. Lett. 17, 1154–1156 (2005). consumption. J. Appl. Phys. 109, 053103 (2011).
31. Diehl, L. et al. High-power quantum cascade lasers grown by low-pressure 61. Xie, F. et al. Continuous wave operation of distributed feedback quantum
metal organic vapor-phase epitaxy operating in continuous wave above cascade lasers with low threshold voltage and lower power consumption.
400 K. Appl. Phys. Lett. 88, 201115 (2006). Proc. SPIE 8277, 82770S (2012).
32. Gresch, T., Giovannini, M., Hoyer, N. & Faist, J. Quantum cascade lasers with 62. Xie, F. et al. High-temperature continuous-wave operation of low power
large optical waveguides. IEEE Photon. Tech. Lett. 18, 544–546 (2006). consumption single-mode distributed-feedback quantum-cascade lasers at
33. Evans, A. et al. Buried heterostructure quantum cascade lasers with high λ ~ 5.2 μm. Appl. Phys. Lett. 95, 091110 (2009).
continuous-wave wall plug efficiency. Appl. Phys. Lett. 91, 071101 (2007). 63. Faist, J. et al. Distributed feedback quantum cascade lasers. Appl. Phys. Lett.
34. Hoffman, A. J. et al. Low voltage-defect quantum cascade laser with 70, 2670–2672 (1997).
heterogeneous injector regions. Opt. Express 15, 15818–15823 (2007). 64. Gmachl, C. et al. Complex-coupled quantum cascade distributed-feedback
35. Bai, Y. et al. Room temperature continuous wave operation of quantum laser. IEEE. Photon. Tech. Lett. 9, 1090–1092 (1997).
cascade lasers with watt-level optical power. Appl. Phys. Lett. 92, 65. Luo, G. P. et al. Grating-tuned external-cavity quantum-cascade
101105 (2008). semiconductor lasers. Appl. Phys. Lett. 78, 2834–2836 (2001).
36. Bai, Y., Slivken, S., Darvish, S. R. & Razeghi, M. Room temperature 66. Fuchs, P. et al. Widely tunable quantum cascade lasers with coupled cavities
continuous wave operation of quantum cascade lasers with 12.5% wall plug for gas detection. Appl. Phys. Lett. 97, 181111 (2010).
efficiency. Appl. Phys. Lett. 93, 021103 (2008). 67. Semmel, J., Kaiser, W., Hofmann, H., Hofling, S. & Forchel, A. Single mode
37. Lyakh, A. et al. 1.6 W high wall plug efficiency, continuous-wave room emitting ridge waveguide quantum cascade lasers coupled to an active ring
temperature quantum cascade laser emitting at 4.6 μm. Appl. Phys. Lett. 92, resonator filter. Appl. Phys. Lett. 93, 211106 (2008).
111110 (2008). 68. Wakayama, Y., Iwamoto, S. & Arakawa, Y. Switching operation of lasing
38. Bai, Y. et al. High power broad area quantum cascade lasers. Appl. Phys. Lett. wavelength in mid-infrared ridge-waveguide quantum cascade lasers coupled
95, 221104 (2009). with microcylindrical cavity. Appl. Phys. Lett. 96, 171104 (2010).
39. Escarra, M. D. et al. Quantum cascade lasers with voltage defect of less than 69. Liu, P. Q., Wang, X. J., Fan, J. Y. & Gmachl, C. F. Single-mode quantum
one longitudinal optical phonon energy. Appl. Phys. Lett. 94, 251114 (2009). cascade lasers based on a folded Fabry–Pérot cavity. Appl. Phys. Lett. 98,
40. Katz, S., Vizbaras, A., Boehm, G. & Amann, M. C. High-performance 061110, (2011).
injectorless quantum cascade lasers emitting below 6 μm. Appl. Phys. Lett. 94, 70. Liu, P. Q., Sladek, K., Wang, X. J., Fan, J. Y. & Gmachl, C. F. Single-mode
151106 (2009). quantum cascade lasers employing a candy-cane shaped monolithic coupled
41. Lyakh, A. et al. 3 W continuous-wave room temperature single-facet emission cavity. Appl. Phys. Lett. 99, 241112 (2011).
from quantum cascade lasers based on nonresonant extraction design 71. Blaser, S. et al. Room-temperature, continuous-wave, single-mode quantum-
approach. Appl. Phys. Lett. 95, 141113 (2009). cascade lasers at λ ≈ 5.4 μm. Appl. Phys. Lett. 86, 041109 (2005).
42. Bai, Y. et al. Highly temperature insensitive quantum cascade lasers. Appl. 72. Yu, J. S. et al. High-power, room-temperature, and continuous-wave
Phys. Lett. 97, 251104 (2010). operation of distributed-feedback quantum-cascade lasers at λ ~ 4.8 μm.
43. Bai, Y. B., Slivken, S., Kuboya, S., Darvish, S. R. & Razeghi, M. Quantum Appl. Phys. Lett. 87, 041104 (2005).
cascade lasers that emit more light than heat. Nature Photon. 4, 73. Kennedy, K. et al. High performance InP-based quantum cascade distributed
99–102 (2010). feedback lasers with deeply etched lateral gratings. Appl. Phys. Lett. 89,
44. Liu, P. Q. et al. Highly power-efficient quantum cascade lasers. Nature Photon. 201117 (2006).
4, 95–98 (2010). 74. Slight, T. J. et al. λ ~ 3.35 μm distributed-feedback quantum-cascade lasers
45. Yao, Y., Wang, X. J., Fan, J. Y. & Gmachl, C. F. High performance ‘continuum- with high-aspect-ratio lateral grating. IEEE Photon. Tech. Lett. 23,
to-continuum’ quantum cascade lasers with a broad gain bandwidth of over 420–422 (2011).
400 cm–1. Appl. Phys. Lett. 97, 081115 (2010). 75. Golka, S., Pflugl, C., Schrenk, W. & Strasser, G. Quantum cascade lasers with
46. Bai, Y., Bandyopadhyay, N., Tsao, S., Slivken, S. & Razeghi, M. Room lateral double-sided distributed feedback grating. Appl. Phys. Lett. 86,
temperature quantum cascade lasers with 27% wall plug efficiency. 111103 (2005).
Appl. Phys. Lett. 98, 181102 (2011). 76. Finger, N., Schrenk, W. & Gornik, E. Analysis of TM-polarized DFB laser
47. Lu, Q. Y., Bai, Y., Bandyopadhyay, N., Slivken, S. & Razeghi, M. 2.4 W room structures with metal surface gratings. IEEE J. Quant. Electron. 36,
temperature continuous wave operation of distributed feedback quantum 780–786 (2000).
cascade lasers. Appl. Phys. Lett. 98, 181106 (2011). 77. Mujagić, E. et al. Ring cavity induced threshold reduction in single-mode
48. Faist, J. Wallplug efficiency of quantum cascade lasers: Critical parameters surface emitting quantum cascade lasers. Appl. Phys. Lett. 96,
and fundamental limits. Appl. Phys. Lett. 90, 253512 (2007). 031111 (2010).
49. Yang, Q. K. et al. Wall-plug efficiency of mid-infrared quantum cascade 78. Darvish, S. R., Slivken, S., Evans, A., Yu, J. S. & Razeghi, M. Room-temperature,
lasers. J. Appl. Phys. 111, 053111 (2012). high-power, and continuous-wave operation of distributed-feedback quantum-
50. Khurgin, J. B. et al. Role of interface roughness in the transport and lasing cascade lasers at λ ~ 9.6 μm. Appl. Phys. Lett. 88,
characteristics of quantum-cascade lasers. Appl. Phys. Lett. 94, 091101 (2009). 201114 (2006).

438 NATURE PHOTONICS | VOL 6 | JULY 2012 | www.nature.com/naturephotonics


© 2012 Macmillan Publishers Limited. All rights reserved.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2012.143 FOCUS | REVIEW ARTICLES
79. Wittmann, A. et al. Distributed-feedback quantum-cascade lasers at 9 μm 100. Weidmann, D. & Wysocki, G. High-resolution broadband (>100 cm-1)
operating in continuous wave up to 423 K. IEEE Photon. Tech. Lett. 21, infrared heterodyne spectro-radiometry using an external cavity quantum
814–816 (2009). cascade laser. Opt. Express 17, 248–259 (2009).
80. Xie, F. et al. High-temperature continuous-wave operation of low power 101. Maulini, R. et al. Widely tunable high-power external cavity quantum cascade
consumption single-mode distributed-feedback quantum-cascade lasers at laser operating in continuous-wave at room temperature. Electron. Lett. 45,
λ < 5.2 μm. Appl. Phys. Lett. 95, 091110 (2009). 107–108 (2009).
81. Zhang, J. C. et al. Low-threshold continuous-wave operation of distributed- 102. Maulini, R., Yarekha, D. A., Bulliard, J. M., Giovannini, M. & Faist, J.
feedback quantum cascade laser at λ ~ 4.6 μm. IEEE Photon. Tech. Lett. 23, Continuous-wave operation of a broadly tunable thermoelectrically
1334–1336 (2011). cooled external cavity quantum-cascade laser. Opt. Lett. 30,
82. Carras, M. & De Rossi, A. Photonic modes of metallodielectric periodic 2584–2586 (2005).
waveguides in the midinfrared spectral range. Phys. Rev. B 74, 235120 (2006). 103. Wysocki, G. et al. Widely tunable mode-hop free external cavity quantum
83. Carras, M. et al. Top grating index-coupled distributed feedback quantum cascade laser for high resolution spectroscopic applications. Appl. Phys. B 81,
cascade lasers. Appl. Phys. Lett. 93, 011109 (2008). 769–777 (2005).
84. Carras, M. et al. Room-temperature continuous-wave metal grating 104. Hugi, A., Maulini, R. & Faist, J. External cavity quantum cascade laser.
distributed feedback quantum cascade lasers. Appl. Phys. Lett. 96, Semicond. Sci. Tech. 25, 083001 (2010).
161105 (2010). 105. Hugi, A. et al. External cavity quantum cascade laser tunable from 7.6 to
85. Lu, Q. Y., Bai, Y., Bandyopadhyay, N., Slivken, S. & Razeghi, M. 2.4 W room 11.4 μm. Appl. Phys. Lett. 95, 061103 (2009).
temperature continuous wave operation of distributed feedback quantum 106. Yao, Y. et al. Broadband quantum cascade laser gain medium based on a
cascade lasers. Appl. Phys. Lett. 98, 181106 (2011). ‘continuum-to-bound’ active region design. Appl. Phys. Lett. 96,
86. Lu, Q. Y., Bai, Y., Bandyopadhyay, N., Slivken, S. & Razeghi, M. Room- 211106 (2010).
temperature continuous wave operation of distributed feedback quantum 107. Fujita, K., Edamura, T., Furuta, S. & Yamanishi, M. High-performance,
cascade lasers with watt-level power output. Appl. Phys. Lett. 97, 231119 (2010). homogeneous broad-gain quantum cascade lasers based on dual-upper-state
87. Wittmann, A. et al. Room temperature, continuous wave operation of design. Appl. Phys. Lett. 96, 241107, (2010).
distributed feedback quantum cascade lasers with widely spaced operation 108. Yao, Y., Tsai, T., Wang, X. J., Wysocki, G. & Gmachl, C. F. Broadband
frequencies. Appl. Phys. Lett. 89, 141116 (2006). quantum cascade lasers based on strongly-coupled transitions with an
88. Lee, B. G. et al. Broadband distributed-feedback quantum cascade laser array external cavity tuning range over 340 cm–1. 2011 Conf. on Lasers and Electro-
operating from 8.0 to 9.8 μm. IEEE Photon. Tech. Lett. 21, 914–916 (2009). Optics (2011).
89. Mujagić, E. et al. Two-dimensional broadband distributed-feedback quantum 109. Fujita, K. et al. High-performance quantum cascade lasers with wide
cascade laser arrays. Appl. Phys. Lett. 98, 141101 (2011). electroluminescence (~600 cm–1), operating in continuous-wave above
90. Capasso, F. High-performance midinfrared quantum cascade lasers. Opt. Eng. 100 °C. Appl. Phys. Lett. 98, 231102 (2011).
49, 111102 (2010). 110. Fujita, K., Edamura, T., Furuta, S. & Yamanishi, M. High-performance,
91. Gokden, B., Bai, Y., Bandyopadhyay, N., Slivken, S. & Razeghi, M. Broad area homogeneous broad-gain quantum cascade lasers based on dual-upper-state
photonic crystal distributed feedback quantum cascade lasers emitting 34 W design. Appl. Phys. Lett. 96, 241107 (2010).
at λ ~ 4.36 μm. Appl. Phys. Lett. 97, 131112 (2010). 111. Dougakiuchi, T. et al. Broadband tuning of external cavity dual-upper-state
92. Menzel, S. et al. Quantum cascade laser master-oscillator power-amplifier quantum-cascade lasers in continuous wave operation. Appl. Phys. Express 4,
with 1.5 W output power at 300 K. Opt. Express 19, 16229–16235 (2011). 102101 (2011).
93. Vurgaftman, I. & Meyer, J. R. Photonic-crystal distributed-feedback quantum 112. Gokden, B., Tsao, S., Haddadi, A., Bandyopadhyay, N. & Slivken, S. Widely
cascade lasers. IEEE J. Quant. Electron. 38, 592–602 (2002). tunable, single-mode, high-power quantum cascade lasers. SPIE Proc.
94. Maulini, R., Mohan, A., Giovannini, M., Faist, J. & Gini, E. External cavity Integrated Photonics: Materials, Devices, and Applications 8069,
quantum-cascade laser tunable from 8.2 to 10.4 μm using a gain element with 806905 (2011).
a heterogeneous cascade. Appl. Phys. Lett. 88, 201113 (2006). 113. Mohan, A. et al. Room-temperature continuous-wave operation of
95. Phillips, M. C., Myers, T. L., Wojcik, M. D. & Cannon, B. D. External cavity an external-cavity quantum cascade laser. Opt. Lett. 32,
quantum cascade laser for quartz tuning fork photoacoustic spectroscopy of 2792–2794 (2007).
broad absorption features. Opt. Lett. 32, 1177–1179 (2007). 114. Maulini, R., Beck, M., Faist, J. & Gini, E. Broadband tuning of external
96. Mukherjee, N. & Patel, C. K. N. Molecular fine structure and transition cavity bound-to-continuum quantum-cascade lasers. Appl. Phys. Lett. 84,
dipole moment of NO2 using an external cavity quantum cascade laser. 1659–1661 (2004).
Chem. Phys. Lett. 462, 10–13 (2008). 115. Lee, B. G. et al. Widely tunable single-mode quantum cascade laser source for
97. Wysocki, G. et al. Widely tunable mode-hop free external cavity quantum mid-infrared spectroscopy. Appl. Phys. Lett. 91, 231101 (2007).
cascade lasers for high resolution spectroscopy and chemical sensing.
Appl. Phys. B 92, 305–311 (2008). Acknowledgements
98. Hancock, G., van Helden, J. H., Peverall, R., Ritchie, G. A. D. & Walker, R. J. The authors acknowledge collaborations with colleagues at Princeton University and
Direct and wavelength modulation spectroscopy using a CW external cavity associated with the NSF Engineering Research Center MIRTHE. A.J.H. thanks S. Howard
quantum cascade laser. Appl. Phys. Lett. 94, 201110 (2009). for valuable discussions. They also acknowledge partial support by MIRTHE (NSF-ERC)
99. Weidmann, D., Tsai, T., Macleod, N. A. & Wysocki, G. Atmospheric and DTRA.
observations of multiple molecular species using ultra-high-resolution
external cavity quantum cascade laser heterodyne radiometry. Opt. Lett. 36, Additional information
1951–1953 (2011). Correspondence and requests for materials should be addressed to A.J.H.

NATURE PHOTONICS | VOL 6 | JULY 2012 | www.nature.com/naturephotonics 439


© 2012 Macmillan Publishers Limited. All rights reserved.

You might also like