Fact-checked by Grok 2 weeks ago

Quantum-cascade laser

A quantum cascade laser (QCL) is a that achieves light emission through intersubband transitions in a series of coupled quantum wells, enabling tunable output wavelengths in the mid- to far-infrared range (typically 3–300 μm) independent of the material's bandgap energy. Unlike conventional lasers, which rely on interband recombination of electrons and holes, QCLs operate unipolarly with electrons alone, cascading sequentially through multiple active regions to generate photons at each stage. This design, based on quantum confinement in ultrathin layers grown by , allows precise band structure engineering for and optical gain. Invented in 1994 by Jerome Faist and colleagues at Bell Laboratories, the first QCL demonstrated pulsed lasing at 4.2 μm with peak powers exceeding 8 mW, marking a breakthrough in by extending capabilities to longer wavelengths where traditional sources falter due to Auger recombination and thermal limitations. Subsequent advancements rapidly achieved continuous-wave room-temperature operation by the late 1990s, with mid-infrared QCLs delivering multi-watt powers, wall-plug efficiencies up to 22% (as of 2025), and single-mode outputs via distributed feedback structures. Materials systems such as /AlInAs on InP substrates dominate mid-IR designs, while GaAs/AlGaAs supports (THz) QCLs, which emerged in 2002 and extend operation to 1–5 THz frequencies. QCLs excel in applications requiring compact, high-brightness sources, including high-resolution for detection, free-space optical communications, and biomedical . Their ability to produce frequency combs spanning octaves has revolutionized dual-comb and , while THz variants enable non-invasive screening and astronomical observations. Ongoing challenges, such as enhancing THz efficiency and integrating QCLs into photonic circuits, promise further expansion into sensing platforms and quantum technologies.

Fundamentals

History

The theoretical proposal for what would become the quantum cascade laser originated in 1971, when Robert F. Kazarinov and R. A. Suris described the potential for amplification of electromagnetic waves via intersubband transitions in a semiconductor superlattice structure. This concept remained unrealized for over two decades until 1994, when the first experimental demonstration was achieved at Bell Laboratories by Federico Capasso, Jerome Faist, Deborah L. Sivco, Carlo Sirtori, Albert L. Hutchinson, and Alfred Y. Cho. Their device, based on GaInAs/AlInAs heterostructures, operated in pulsed mode at cryogenic temperatures (around 4 K) and emitted at a wavelength of 4.2 μm, marking a fundamental departure from traditional interband semiconductor lasers. Early quantum cascade lasers faced substantial hurdles, requiring cryogenic cooling due to inefficient carrier injection, thermal backfilling, and non-radiative processes such as recombination and carrier leakage over heterobarriers, which severely limited threshold currents and gain at elevated temperatures. Progress accelerated in the late 1990s, with room-temperature pulsed operation first demonstrated around 5 μm in 1998 using improved designs that enhanced injection and reduced losses. By 2002, continuous-wave room-temperature lasing was realized at wavelengths near 9 μm, enabling practical applications through better thermal management and buried heterostructure geometries that dissipated heat more effectively. Further advancements pushed the performance envelope: by 2007, the short-wavelength emission limit reached 3 μm using strained InAs/AlSb material systems with deep quantum wells to achieve the necessary intersubband energy spacing, though operation remained challenging at ambient temperatures due to increased non-radiative losses. In parallel, the technology expanded to the regime in 2002, when Rüdeger Köhler and colleagues demonstrated the first terahertz quantum cascade laser at 67 μm (4.4 THz) using GaAs/AlGaAs superlattices with bound-to-continuum transitions for broadband gain. Marking the 30th anniversary in 2024, quantum cascade lasers have profoundly influenced mid-infrared , evolving from laboratory prototypes to compact, high-power sources integral to , sensing, and free-space communications, with ongoing refinements in efficiency and wavelength versatility.

Intersubband versus interband transitions

Conventional lasers, such as lasers, operate via interband transitions, where electrons recombine from the conduction band to the valence band across the material's bandgap. In direct bandgap like GaAs, with a bandgap of approximately 1.42 , these transitions are limited to near-infrared or visible wavelengths, for example, around 0.87 μm emission. This bandgap constraint restricts interband lasers from efficiently accessing mid- to far-infrared spectral regions without using narrow-bandgap materials that suffer from severe non-radiative losses. In contrast, quantum-cascade lasers (QCLs) rely on intersubband transitions, in which electrons jump between quantized subbands within the conduction band of quantum wells. These transitions enable engineered energies spanning from about 3 μm in the mid-infrared to the range, unbound by the host material's bandgap. The unipolar nature of these transitions involves only electrons, eliminating the need for injection and associated complications in devices. The in QCLs consists of alternating thin layers of materials, such as GaAs and AlGaAs or InGaAs and AlInAs, grown via techniques like to create confined electron states. The transition energy E = \hbar \omega is determined by the well width, barrier height, and layer composition, allowing precise tuning through band rather than material selection. This design offers key advantages for emission: wavelengths are scalable by adjusting quantum confinement, enabling operation in regions where interband lasers falter due to non-radiative recombination. In QCLs, the unipolar intersubband process avoids the dominant losses prevalent in mid- interband devices, as it lacks electron-hole pairs that trigger such processes. A of the levels typically depicts a series of coupled quantum wells forming a "staircase" potential, where an injected undergoes multiple intersubband emissions, each releasing a of \hbar \omega, before tunneling to the next stage for reuse.

Operating principles

Rate equations

The rate equations for quantum cascade lasers (QCLs) describe the dynamics of carrier populations in the upper (N₃) and lower (N₂) laser subbands, as well as the density (S) in the , capturing the interplay between electrical injection, intersubband transitions, and stimulated/. These equations are derived from the balance of scattering rates and radiative processes in the , assuming a three-level model where electrons are injected into the upper subband, undergo to the lower subband, and are extracted for the next stage. For the upper subband population, the is given by \frac{dN_3}{dt} = \eta_\text{inj} \frac{J}{e} - \frac{N_3}{\tau_3} - g \Gamma (N_3 - N_2) S, where \eta_\text{inj} is the injection efficiency, J is the current density, e is the elementary charge, \tau_3 is the lifetime of the upper subband (dominated by phonon scattering and tunneling), g is the gain coefficient (related to the intersubband transition dipole moment), \Gamma is the optical confinement factor, and S is the photon density. The first term represents carrier injection from the injector region, the second accounts for non-radiative decay, and the third describes stimulated emission depleting the population inversion \Delta N = N_3 - N_2. For the lower subband, the equation is \frac{dN_2}{dt} = g \Gamma (N_3 - N_2) S - \frac{N_2}{\tau_2} + \text{extraction terms}, where \tau_2 is the lower subband lifetime (typically shorter due to fast to the ), and the extraction terms model tunneling to the next region, ensuring unipolar operation. The photon density evolution follows \frac{dS}{dt} = \Gamma g (N_3 - N_2) S - \frac{S}{\tau_\text{ph}} + \beta R_\text{sp}, with \tau_\text{ph} the lifetime (determined by mirror and losses), \beta the spontaneous emission coupling factor (usually small, ~10⁻³–10⁻⁵ in QCLs), and R_\text{sp} the spontaneous emission rate into the lasing mode (proportional to N_3 / \tau_\text{sp}, where \tau_\text{sp} is the spontaneous lifetime). These coupled equations highlight how drives optical gain while buildup clamps the inversion above threshold. In steady-state operation (dN_3/dt = dN_2/dt = dS/dt = 0), the simplifies to \Delta N \approx (\eta_\text{inj} J \tau_3 / e) / (1 + g \Gamma S \tau_3), assuming fast extraction such that N_2 \ll N_3 and neglecting minor terms; this leads to the modal G = \Gamma g \Delta N, which balances losses for lasing. The current density is then J_\text{th} = (e / \eta_\text{inj} \tau_3) \cdot (1 / g \Gamma) \cdot (1 / \tau_\text{ph}), emphasizing the role of injection efficiency \eta_\text{inj} (typically 0.5–0.8 in optimized designs) in minimizing and maximizing wall-plug . This condition marks the point where equals total losses, enabling net . Transient behaviors arise from the coupled carrier-photon , leading to relaxation oscillations in the and as the system approaches after . Numerical solutions of the rate equations reveal an initial buildup time scaling with \tau_\text{ph} \ln(J / J_\text{th}) and damped oscillations at frequencies ~–10 GHz, depending on current and ; however, the fast intersubband (\tau_3, \tau_2 \sim 0.1–1 ) often heavily damps these oscillations compared to interband lasers, resulting in smoother transients with overshoots on scales.

Active region designs

The active region of a quantum cascade laser is composed of repeated stages, each featuring a basic cascade unit structured as a three-level system. In this configuration, electrons resonantly from an / (level 1) into the upper laser level (level 3), undergo to the lower laser level (level 2) while emitting a , and then tunnel to the (level 1) of the subsequent stage. This sequential process repeats across multiple stages, allowing a single to generate multiple photons as it cascades through the structure. One of the earliest designs is the bound-to-bound configuration, where both the upper (level 3) and lower (level 2) levels are discrete, bound states confined within . This design yields a narrow linewidth due to the well-defined separation between levels and was employed in the initial demonstrations of QCLs. The for such intersubband transitions is approximated as \mu \approx e \cdot L_w / 2, where e is the and L_w is the width, influencing the optical strength. To address limitations in lower level lifetime and temperature sensitivity, bound-to- designs were developed, in which the upper laser level remains bound while the lower level forms part of a miniband . This enables rapid depopulation of level 2 via (with lifetime \tau_2 < 0.3 ps), minimizing electron accumulation that could clamp the population inversion and broadening the gain spectrum for applications requiring wider tunability. These designs combine the selectivity of bound states with the fast extraction of superlattice-like structures. Diagonal designs introduce spatial separation between the upper and lower laser levels, with their wavefunctions overlapping at an angle \theta that tunes the inter-level coupling. This reduces the oscillator strength of the radiative transition compared to vertical bound-to-bound setups, which helps suppress thermal backfilling and enhances temperature stability, particularly in terahertz operating above 100 K. The diagonal geometry balances gain and depopulation rates, improving overall device performance under high-temperature conditions. For applications demanding simultaneous emission at multiple wavelengths, incorporate tapered or step-graded quantum wells across the active region, enabling dual or broadband emission by varying the well widths and thus the transition energies within stages. Strain-balanced heterostructures, such as those using GaAs/Al_{0.15}Ga_{0.85}As with compensated layers, are integrated to reduce lattice mismatch defects and maintain structural integrity during epitaxial growth. These approaches facilitate compact sources for spectroscopy without requiring external tuning elements. The overall active region typically comprises N_\text{stages} \approx 30--$50repeated units to achieve sufficient optical gain while managing electrical properties. This results in a threshold voltageV_\text{th} \approx N_\text{stages} \cdot (E_{32} + E_{21})/e \approx 10--&#36;20 V, where E_{32} and E_{21} are the energy differences across the lasing and extraction transitions, respectively; the population dynamics of these levels, as modeled by rate equations, underpin the voltage scaling with stage count.

Material systems and fabrication

Semiconductor materials

Quantum-cascade lasers (QCLs) operating in the mid-infrared range (3-20 μm) predominantly utilize InP-based heterostructures, such as GaInAs/AlInAs, which are lattice-matched to InP substrates for strain-free growth and reliable performance. These materials provide a conduction band offset (ΔE_c) of approximately 0.5 eV in the lattice-matched configuration, enabling efficient intersubband transitions while maintaining structural integrity. For enhanced conduction band offsets and reduced carrier leakage, strain-compensated designs incorporating InGaAs/AlInAs layers are employed, allowing operation at higher temperatures and powers by increasing ΔE_c beyond the lattice-matched limit. In the far-infrared and terahertz regime (20 μm to 1 THz), GaAs/AlGaAs heterostructures dominate due to the low effective electron mass (m* ≈ 0.07 m_e) in the Γ-valley, which supports long-wavelength transitions with sufficient oscillator strength. The AlGaAs barriers typically incorporate 15-30% aluminum content to tune the barrier height (ΔE_c ≈ 150-300 meV), balancing confinement and minimizing alloy disorder scattering while enabling phonon-assisted depopulation. At the short-wavelength limit near 3 μm, achieving sufficient ΔE_c (>1 eV) requires high-barrier materials like InGaAs/AlAs, where the conduction band offset approaches 1 eV, allowing radiative transitions with energies exceeding those in standard InP-based systems. Alternative material systems include SiGe heterostructures for integration with , leveraging compatible growth on Si substrates to enable compact, CMOS-compatible THz sources despite challenges in valley scattering. Recent advances include epitaxial growth of InAs/AlSb type-II QCLs on Si substrates, enabling integration with for mid-IR applications. Type-II superlattices like InAs/AlSb offer low optical losses and high gain due to their broken-gap alignment, facilitating mid-infrared QCLs with reduced waveguide absorption compared to type-I systems. Doping is confined to n-type levels in the injector regions, typically around 10^{17} cm^{-3}, to provide sufficient supply for cascading without excessive . Active regions remain undoped to minimize free carrier absorption losses, which would otherwise degrade optical confinement and increase threshold currents. A key challenge in short-wavelength QCLs is interface roughness , which broadens linewidths, enhances non-radiative recombination, and limits net gain by increasing the lower lifetime and promoting leakage currents.

Growth techniques

Quantum cascade lasers (QCLs) are primarily fabricated using epitaxial growth techniques that enable the precise layering of semiconductor heterostructures with thicknesses on the order of 1-10 nm per layer. () serves as the predominant method due to its ability to achieve atomic-level precision in environments, where elemental fluxes are controlled to deposit materials like /AlInAs on InP substrates. Growth rates typically range from 0.5 to 1 monolayer per second, allowing for the construction of complex active regions comprising dozens of cascaded stages. In-situ monitoring via reflection high-energy electron diffraction (RHEED) ensures interface quality and layer uniformity during deposition. An alternative to MBE is metal-organic chemical vapor deposition (MOCVD), which offers higher throughput for larger wafers and is suitable for industrial-scale production of QCLs. In MOCVD, precursors such as trimethylindium, trimethylgallium, and trimethylaluminum are used to grow GaInAs/AlInAs layers on InP substrates at temperatures around 720°C, with V/III ratios of approximately 116 for InGaAs and 21 for InAlAs to optimize material quality. Growth interruptions of about 3 seconds between layers enhance interface sharpness, though residual precursor flows can lead to slight compositional gradients and a red-shift in emission wavelength by 0.5-1 μm compared to MBE-grown structures. While MOCVD enables efficient, repeatable fabrication, it generally introduces more defects than MBE, though optimized conditions can mitigate this for high-performance devices. Following epitaxial growth, device processing involves defining optical waveguides and electrical contacts. Ridge waveguides are typically formed by (ICP-RIE) using gases like Ar/SiCl₄ to create double-trench structures with controlled depths for optical confinement. Top contacts are metallized with / stacks (e.g., 20 nm and 200 nm ), often thickened via , while bottom contacts use Ge//Ni/ after mechanical thinning of the substrate to around 120-200 μm. For certain designs, such as type-II configurations, full substrate removal may be performed, and facets are cleaved or coated with high-reflectivity (HR, ~95%) and anti-reflectivity (AR, ~5%) layers to form Fabry-Pérot cavities. Key challenges in QCL fabrication include maintaining thickness uniformity below 1 across 3-inch wafers to ensure consistent intersubband transition energies, and keeping defect densities under 10^6 ^{-2} to minimize non-radiative recombination. Cleaving facets requires precise control to avoid cracks in the brittle heterostructures, and sidewall roughness from must be minimized to reduce optical losses. Packaging is tailored to mitigate thermal and optical issues, particularly for QCLs, where epoxy-free bonding is essential to prevent losses. Devices are often mounted epi-layer up on heatspreaders like or using , with for electrical connections, and thermoelectric coolers enable continuous-wave operation up to temperatures of 129 K (as of 2024).

Emission characteristics

Wavelength ranges

Quantum cascade lasers (QCLs) primarily operate in the mid-infrared (mid-IR) spectral range, achieving emission wavelengths from approximately 3 to 20 μm through intersubband transitions in engineered quantum wells. As of 2023, shorter wavelengths down to 2.6 μm are possible by employing deep quantum wells that provide large conduction band offsets (ΔE_c > 0.4 eV), as seen in InAs/AlSb heterostructures where emission at 2.6–3.0 μm has been realized, including continuous-wave room-temperature operation at 3.02 μm. This range is standard for most devices, with peak powers exceeding 1 W demonstrated at 4–10 μm using /AlInAs material systems on InP substrates. The short-wavelength limit is fundamentally set by the maximum barrier height in the material system, beyond which carrier confinement becomes insufficient. In the far-infrared (far-IR) regime, QCLs extend to 20–100 μm by using shallower quantum wells and reduced aluminum content in the barriers to lower the transition energies. However, output power decreases significantly in this range due to the smaller intersubband , which scales proportionally with (μ ∝ λ), leading to reduced oscillator strengths. For example, emission at 24.4 μm has been achieved at 240 K using InGaAs/AlInAs designs, though cryogenic cooling is often required for higher performance. As of 2016, the longest far-IR emission reached ~28 μm. The (THz) range, corresponding to 1–5 THz or 60–300 μm wavelengths (with recent extensions up to ~5.4 THz as of 2023), is accessed via phonon-confined active region designs that avoid resonant coupling with longitudinal optical () phonons, which cause rapid carrier depopulation. In GaAs/AlGaAs systems, the LO-phonon energy limits operation to below approximately 8 THz, as becomes dominant at longer wavelengths. Metal-metal waveguides enable confinement with losses α ≈ 10–100 cm⁻¹, supporting peak powers over 2 W at low temperatures and operation up to 200 K. Recent demonstrations include emission spanning 1.9–4.5 THz and extensions to lower frequencies down to approximately 1.6 THz using bound-to-continuum transitions. Wavelength tuning in QCLs is facilitated by temperature variations, with shifts of dλ/dT ≈ 0.1–1 /K due to and changes, or by current injection, leveraging broad gain spectra of ~100 cm⁻¹ for single-chip tunability exceeding 1000 cm⁻¹ in external-cavity configurations. These mechanisms, combined with design flexibility in widths, enable versatile spectral coverage across the IR and THz domains.

Optical waveguides

In quantum cascade lasers (QCLs), optical waveguides are essential for confining the electromagnetic mode to the , ensuring efficient interaction between the light and the intersubband transitions while minimizing losses. These structures leverage the high contrast between the core and surrounding claddings to guide the transverse magnetic (TM)-polarized emission, which is inherent to intersubband processes. The design optimizes the overlap of the optical field with the gain medium, balancing confinement, thermal dissipation, and fabrication complexity across mid-infrared and wavelengths. The waveguide is a prevalent in QCLs, formed by a typically 5-10 μm wide into the epitaxial structure, followed by deposition of a top metal cladding for and optical confinement. This geometry provides lateral confinement through at the ridge sidewalls, with a vertical confinement factor Γ ≈ 0.3-0.5, enabling single-mode operation when the ridge width suppresses higher-order lateral modes. However, losses can arise from sidewall or , particularly in deeply etched structures, impacting currents and output power. For improved performance, buried heterostructure () waveguides incorporate semi-insulating InP regrown laterally around the etched ridge, encapsulating the active region to reduce and enhance thermal management by isolating heat from the . This regrowth, often using metal-organic (MOCVD) or hydride vapor phase (HVPE) on InP-based QCLs, yields smoother interfaces and lower leakage currents compared to simple ridge designs, enabling higher wall-plug efficiencies and continuous-wave operation at . The BH approach is particularly beneficial for high-power devices, as it minimizes current spreading and improves heat extraction. In terahertz QCLs, double-metal waveguides employ thin layers sandwiched between top and bottom metal layers, exploiting modes for near-unity confinement (Γ > 0.8) across the entire thickness. This configuration, fabricated via and substrate removal, achieves strong vertical overlap but incurs metallic losses typically in the range of 20–100 cm⁻¹ in the THz regime, which can vary with frequency and materials, necessitating short lengths to maintain low thresholds. Despite these losses, the high Γ compensates by enhancing overlap, enabling up to frequencies around 5 THz. Substrate removal techniques are employed in QCLs on low-index materials like GaAs to suspend the , facilitating surface emission or integration with plasmonic structures. The process involves selective to thin the substrate, often followed by to a host wafer, which exposes the bottom metal or for improved mode control and reduced from the high-index GaAs (n ≈ 3.6). This enables vertical in THz devices and enhances far-field patterns, though it adds fabrication complexity and potential mechanical fragility. The mode overlap, or confinement factor Γ, quantifies the efficiency of light-matter interaction and is defined vertically as \Gamma = \frac{\int_{\text{active}} |E|^2 \, dz}{\int_{\text{total}} |E|^2 \, dz}, where |E| is the amplitude, integrated over the and total cross-section. Horizontally, confinement is tuned by ridge width to favor the fundamental mode, suppressing higher orders that reduce effective . Optimizing Γ across designs is critical for balancing losses and performance, with values approaching 1 in plasmonic THz structures but typically lower in mid-infrared ridge guides.

Device configurations

Fabry–Pérot lasers

The Fabry–Pérot quantum cascade laser (QCL) is the foundational device configuration for QCLs, employing a linear defined by the uncoated, naturally cleaved facets of the chip as end mirrors. These facets arise from the crystalline planes of materials like GaAs or InP, providing reflectivities of approximately 0.3 due to Fresnel reflection at the air- interface. Cavity lengths typically range from 1 to 3 mm to balance threshold current and output power, with the effective around 3.3 leading to longitudinal mode spacings of Δν = c / (2 n L) ≈ 15–45 GHz, where c is the , n is the , and L is the length. Fabrication begins with epitaxial growth of the QCL heterostructure using (MBE) or metalorganic (MOCVD) to form the and layers, followed by lithographic definition of waveguides (typically 10–20 μm wide) and metal contacts for injection. The is then cleaved along {110} planes to create the Fabry–Pérot facets, diced into bars, and mounted epi-up or epi-down on a heatsink using indium for . This process is straightforward but results in relatively high mirror losses (α_m ≈ 10–20 cm⁻¹) from the uncoated facets, limiting efficiency compared to coated or grating-based designs. Operation above occurs at current densities J_th ≈ 1–5 kA/cm² for mid-infrared Fabry–Pérot QCLs at , depending on and design; for example, values of 1.5–2.8 kA/cm² have been reported for 4.3–9.3 μm in continuous-wave mode. The output power from one facet follows the standard expression P = \frac{\alpha_m}{\alpha_\text{total}} (I - I_\text{th}) \frac{\eta_i}{q}, where α_total is the total optical loss, I is the drive current, I_th is the threshold current, η_i is the internal (often 0.5–0.8), and q is the ; peak powers exceeding 1 per facet are achievable in pulsed mode at . The inherent to intersubband transitions in QCLs results in multimode , with 10–20 longitudinal modes typically lased simultaneously and a spectral width of ~10–50 cm⁻¹ (or up to 120 cm⁻¹ in designs), centered on the target . This multimodality arises from the Fabry–Pérot supporting multiple frequencies within the envelope, leading to a clustered without external selection. The ease of fabrication without additional optical elements makes Fabry–Pérot QCLs valuable for of novel active regions and material systems. However, the lack of mode selection yields poor side-mode suppression ratios (often <20 dB) and wavelength drifts of ~0.3–1 cm⁻¹/K with temperature, limiting applications requiring narrow-linewidth or stable emission.

Distributed feedback lasers

Distributed feedback (DFB) quantum cascade lasers incorporate periodic gratings within the waveguide to provide wavelength-selective feedback, enabling stable single-mode operation essential for applications requiring narrow linewidths and precise spectral control. Unlike Fabry-Pérot configurations, which rely on cleaved facets for broadband feedback, DFB designs integrate the grating directly into the structure to suppress unwanted modes and select a specific longitudinal mode. The seminal demonstration of DFB QCLs occurred in 1997, achieving pulsed single-mode emission above room temperature at wavelengths of 5.4 μm and 8 μm. The grating structure typically employs a first-order etched into the top layers of the semiconductor waveguide, with a period given by \Lambda = \frac{\lambda}{2 n_{\text{eff}}}, where \lambda is the emission wavelength and n_{\text{eff}} is the effective refractive index of the guided mode. This period ensures the Bragg condition for reflection at the desired wavelength, providing distributed feedback along the cavity length. The coupling coefficient, which quantifies the strength of the grating's feedback, is approximated as \kappa \approx \frac{\pi \Delta n}{\lambda} for an index perturbation \Delta n induced by the grating etch. Single-mode selection in DFB QCLs arises from the grating's wavelength-dependent feedback, which strongly amplifies one longitudinal mode while suppressing others through destructive interference. This results in a sidemode suppression ratio (SMSR) exceeding 30 dB, ensuring high spectral purity even under varying operating conditions. Design variants enhance performance by tailoring the coupling mechanism. Imaginary-coupled gratings, achieved via metal overlays that perturb the gain profile, provide feedback through absorption or gain modulation rather than purely refractive index changes. Complex-coupled designs combine index and gain perturbations, yielding a broader stopband and improved mode stability compared to purely index-coupled gratings. The spectral linewidth of a DFB QCL is described by \Delta \nu \approx \frac{h \nu n_{\text{sp}} (1 + \alpha_c^2)}{8 \pi P_{\text{out}} \tau_p}, where h \nu is the photon energy, n_{\text{sp}} is the spontaneous emission factor, \alpha_c is the linewidth enhancement factor, P_{\text{out}} is the output power, and \tau_p is the photon lifetime. High output powers reduce the linewidth to below 1 kHz, as observed in stabilized mid-infrared DFB QCLs, enhancing coherence for interferometric applications. Fabrication of DFB gratings often utilizes electron-beam lithography to define the submicron periodic structure with high precision, followed by techniques such as epitaxial regrowth for buried heterostructures or metal overlay deposition to avoid regrowth while maintaining coupling strength. In performance, DFB QCLs have demonstrated continuous-wave (CW) output powers exceeding 100 mW at 4.6 μm, with wavelength tuning achieved by varying injection current or temperature, offering tuning coefficients around 0.3–0.5 nm/K.

External cavity and tunable lasers

External cavity quantum cascade lasers (EC-QCLs) extend the tunability of standard Fabry-Pérot QCLs by incorporating external optical elements, such as diffraction gratings, to select and control the lasing wavelength. Common configurations include the Littrow and Littman-Metcalf setups, where the QCL output is collimated and directed to a rotatable diffraction grating or microelectromechanical system (MEMS) mirror for wavelength selection. In the Littrow configuration, the first-order diffracted beam is retro-reflected directly back into the laser facet, while the Littman-Metcalf design uses an additional mirror to direct the diffracted beam, enabling finer angular control and reduced sensitivity to misalignment. These setups typically feature a total cavity length exceeding 1 cm, resulting in a narrow free spectral range (FSR) of approximately 10 GHz, which supports dense mode spacing for precise tuning. Tuning in EC-QCLs is primarily achieved by rotating the grating or mirror, allowing access to a broad range exceeding 100 cm⁻¹, as demonstrated in systems operating around 9-10 μm with ranges up to 110 cm⁻¹. This tuning is limited by the QCL's gain curve, which defines the spectral region of sufficient amplification, and by mode-hopping, where the laser jumps between longitudinal modes. Mode-hop-free operation over segments of 120 cm⁻¹ or more can be realized by ramping the injection current, which shifts the gain peak and refractive index to track the desired wavelength continuously without discontinuities. For example, in a with a 10 cm cavity, continuous tuning of 110 cm⁻¹ has been reported with output powers suitable for spectroscopic applications. Vernier tuning enhances coverage by employing two Fabry-Pérot cavities with slightly mismatched FSRs, typically differing by ΔFSR/FSR ≈ 0.1%, enabling supermode selection across a broad spectrum. This configuration allows quasi-continuous coverage of 100-200 cm⁻¹ by aligning the comb-like mode spectra of the two cavities, where only wavelengths satisfying the Vernier condition lase. Such external or integrated vernier schemes have been implemented in mid-infrared QCLs to achieve tuning ranges up to approximately 200 cm⁻¹ without mechanical movement, leveraging the superposition of cavity modes for extended operational bandwidth. Coupled cavity designs integrate a Fabry-Pérot section with a distributed feedback (DFB) section, where tuning is accomplished by injecting current into one section to control the phase relationship between the cavities. This alters the effective refractive index, enabling single-mode selection and continuous tuning over tens of cm⁻¹ by adjusting the current in the FP section while using the DFB as a wavelength reference. Such hybrid configurations provide stable, widely tunable output with side-mode suppression ratios exceeding 30 dB, suitable for applications requiring precise wavelength control. Linewidth narrowing in EC-QCLs is often achieved by incorporating etalons within the cavity to filter unwanted modes, reducing the spectral width to below 100 kHz. Additional stabilization uses feedback loops to a piezoelectric transducer (piezo) on the grating or mirror, locking the wavelength to a reference such as a Fabry-Pérot etalon or frequency comb, yielding frequency stabilities of 100 kHz over integration times of seconds. These techniques suppress frequency noise, enabling high-resolution operation with linewidths as narrow as 8 kHz in optimized systems. EC-QCLs with external cavities and tuning capabilities are particularly valuable for spectroscopy applications requiring broad wavelength scans, such as trace gas detection in environmental monitoring and breath analysis. Their ability to perform rapid, mode-hop-free sweeps over hundreds of cm⁻¹ with narrow linewidths facilitates high-sensitivity absorption measurements of molecular fingerprints in the mid-infrared.

Performance and applications

Power, efficiency, and limitations

Quantum cascade lasers (QCLs) have achieved significant output powers, particularly in pulsed operation, with peak powers exceeding 10 W demonstrated at wavelengths between 4 and 9 μm. In continuous-wave (CW) mode at room temperature, output powers greater than 1 W have been reported, for example, 1.15 W at approximately 10.3 μm. The slope efficiency, defined as η_s = dP/dI, typically ranges from 1 to 3 W/A, reflecting the incremental power increase per unit current, as observed in devices with 2.16 W/A near threshold at room temperature. Wall-plug efficiency (WPE), the ratio of output optical power to electrical input power expressed as WPE = (P_out / P_elec) × 100%, has reached up to 31% at approximately 4.9 μm in room-temperature pulsed operation (as of 2020), with recent advancements achieving 29.3% in pulsed mode (as of 2025). This efficiency is fundamentally limited by the operating voltage, which arises from the number of stages N_stages and the total energy per stage E_total, yielding V ≈ N_stages × E_total / e, typically 5-10 V per stage due to intersubband transition energies. Temperature performance in mid-infrared QCLs allows maximum operating temperatures T_max exceeding 400 K, as achieved in devices emitting at 3.3-3.5 μm. The characteristic temperature T0, which quantifies threshold current sensitivity to heat, ranges from 100 to 200 K, with values like 128 K reported for high-power devices. Performance degrades via thermal rollover, where output power saturates due to carrier leakage and increased non-radiative processes at higher temperatures. Key limitations include interface roughness, characterized by height fluctuations Δ of 0.2-0.5 monolayers (ML), which broadens energy levels ΔE and enhances scattering, reducing gain and increasing threshold currents. Auger scattering, involving carrier-carrier interactions, contributes to non-radiative recombination and limits high-temperature operation by shortening upper-state lifetimes. Free-carrier absorption, with coefficient α_fc proportional to λ² N_dop (where N_dop is doping density), introduces optical losses that scale with wavelength and carrier concentration, further constraining efficiency in longer-wavelength devices. Notable trade-offs exist between power and efficiency, as designs optimizing for high output often increase voltage defects or losses, reducing WPE. Shorter wavelengths enable higher T0 due to larger subband separations but typically yield lower power because of reduced gain per stage and higher interband absorption.

Practical applications

Quantum cascade lasers (QCLs) have found widespread use in mid-infrared absorption spectroscopy for trace gas detection, leveraging their ability to target specific molecular absorption lines in the 3-5 μm range. For instance, QCLs operating at 3.3 μm enable sub-ppb sensitivity for (CH₄) detection in portable sensors suitable for environmental monitoring of greenhouse gases. Similarly, at 4.3 μm, QCLs achieve high-precision measurements of (CO₂) isotopes with sensitivities down to 0.02‰ for δ¹³C(CO₂) and 0.4 ppb for CH₄, supporting atmospheric and isotopic analysis. These systems facilitate real-time, selective detection in field-deployable configurations, enhancing applications in air quality assessment and emission tracking. In the terahertz regime, QCLs enable imaging for non-destructive testing and security screening, where their compact size and high power support real-time visualization of concealed objects. THz QCL-based systems have been demonstrated for standoff detection of explosives, with arrays providing spectral fingerprints for identification at distances up to several meters. These applications benefit from QCLs' ability to penetrate non-conductive materials like clothing or packaging, making them valuable for airport security and industrial quality control without ionizing radiation. QCLs also serve as transmitters in free-space optical communication systems, particularly in the 3-5 μm atmospheric window, where low absorption by water vapor and aerosols allows for high-speed data links over several kilometers. Devices modulated at gigabit-per-second rates have been reported, offering secure, interference-free alternatives to radio frequency systems in military and remote sensing scenarios. In medical diagnostics, QCLs support breath analysis for non-invasive detection of disease biomarkers, such as acetone at levels indicative of diabetes management. Spectroscopy systems using QCLs at around 8 μm have measured exhaled acetone in type 1 diabetes patients with sub-ppm accuracy, enabling real-time monitoring without blood sampling. Fiber-coupled QCL configurations further extend to biomedical sensing, facilitating attenuated total reflection measurements for tissue analysis in potential endoscopic applications. Industrially, QCLs are integrated into process control systems for petrochemical refining, where they monitor hydrogen sulfide (H₂S) and sulfur emissions with ppb-level sensitivity to ensure compliance and safety. In pollution monitoring, tunable QCL analyzers provide continuous, multi-gas detection for stack emissions and ambient air, supporting regulatory standards in manufacturing environments. Commercial QCL modules have been available since 2005, with companies like Daylight Solutions and Pranalytica offering rugged, turnkey systems for spectroscopy and sensing. The global QCL market exceeded $400 million by 2025, driven by demand in defense, environmental, and industrial sectors.

Recent developments

High-power and efficiency advances

Since 2020, significant strides have been made in enhancing the power output and efficiency of (QCLs), particularly in the mid-infrared range, through refined active region designs, advanced epitaxial growth, and improved thermal management. These advances have enabled room-temperature continuous-wave (CW) operation with multi-watt powers and wall-plug efficiencies (WPE) exceeding 20%, addressing longstanding limitations in heat dissipation and voltage drop per stage. Key strategies include optimizing injector regions to minimize the energy separation ΔE21 between subbands, reducing intersubband losses, and implementing low-voltage architectures that operate below 4 V per stage, which collectively boost overall device efficiency while scaling output power. High-power records have been pushed forward, with single-facet CW outputs reaching 5.6 W at room temperature around 5 μm wavelength, representing a benchmark for mid-infrared QCLs. In pulsed operation, peak powers surpass 100 W using nanosecond pulses, facilitating applications requiring intense bursts without excessive thermal buildup. For further scaling, multi-emitter arrays, such as 8-element configurations at 4.6 μm, deliver peak powers of approximately 12 W at low duty cycles (e.g., 4%), enabling average powers in the watts range under quasi-CW conditions. Spectral beam combining of multiple QCL emitters has also emerged as a pathway to kilowatt-level systems, though primarily demonstrated in fiber laser hybrids, with ongoing adaptations for mid-infrared QCLs showing promise for high-brightness outputs. Thermal management plays a crucial role in these achievements, with buried ridge structures bonded epi-down onto synthetic diamond heatspreaders achieving maximum operating temperatures (T_max) over 500 K, significantly extending CW thresholds by enhancing heat extraction from the active region. Efficiency improvements have centered on WPE enhancements, with pulsed-mode records exceeding 30% at room temperature near 5 μm, achieved via optimized injectors that reduce ΔE21 and minimize carrier leakage. efficiencies have reached 22% at 20°C, with cryogenic operation pushing beyond 40% (e.g., 41% at 80 K), underscoring the impact of low-voltage designs under 4 V/stage that lower power dissipation. Pulsed versus trade-offs are evident: nanosecond pulses enable >100 W peaks at high duty cycles (>50% for quasi-), where average powers approach levels but with reduced , while full demands stringent cooling to maintain efficiency. These gains build on baseline efficiencies from prior designs but emphasize incremental refinements in stage count and doping for practical deployment. In 2024–2025, metal-organic chemical vapor deposition (MOCVD) growth has yielded notable advances, including 9.4 μm QCLs with aluminum-compensated InAlAs barriers to correct lattice mismatch and improve interface quality, delivering 1.26 W and 2.08 W pulsed outputs with 7.4% and 10.1% WPE, respectively. Short-pulse regimes have seen WPE surpass 40%, as in cryogenic demonstrations, while overall records for power (5.6 W) and room-temperature WPE (22%) highlight optimized thermal and electrical designs for sustained high-brightness operation. These developments prioritize scalability for applications like and sensing, with MOCVD enabling cost-effective production of compensated structures. A 2025 review highlights further advances, including output powers exceeding 12 W at 80 K for mid-IR QCLs and wall-plug efficiencies over 41% under cryogenic conditions.

Emerging technologies

Recent advancements in quantum-cascade laser (QCL) technology have introduced innovative architectures that enhance control, scalability, and integration capabilities. One notable variant is the transistor-injected QCL (TI-QCL), which integrates a (HBT) structure within the QCL active region to provide independent gate control over carrier injection. This three-terminal design decouples the lasing energy from the injection current, allowing precise tuning of the emission wavelength via base-collector bias while modulating through the emitter current. Demonstrated in InP-based devices, TI-QCLs have achieved near-infrared lasing at approximately 1.6 μm and mid-wave infrared detection beyond 6 μm, with pulsed operation at using low duty cycles such as 1%. QCL arrays and beam-combining techniques represent another frontier, enabling higher brightness and multi-wavelength operation for applications like and systems. Monolithic arrays of distributed-feedback (DFB) QCLs, integrated with arrayed gratings, have achieved beam combining across multiple emitters operating near 4.65 μm, delivering powers up to 1 W per element in pulsed mode. Two-dimensional arrays facilitate parallel outputs for , while coherent combining methods, such as those using tilted-cavity designs, produce high-quality, common-aperture beams with continuous-wave tunability. These configurations scale power without sacrificing beam quality, supporting compact sources for standoff detection. In integrated , hybrid integrations of QCLs with platforms are advancing on-chip mid-infrared sensing. Heterogeneous of III-V QCLs onto silicon-on-nitride or germanium-on- waveguides enables phase-matched for compact spectrometers. Examples include a single-mode DFB QCL at 4.3 μm integrated on a SONOI platform via molecular , QCLs operating between 5.7 and 5.9 μm integrated via flip-chip on GoS platforms for multi-gas detection, and dual DFB QCLs at ~7 μm via self-aligned flip-chip supporting low-loss propagation. These hybrids leverage 's scalability for devices, reducing size and cost for portable sensors. Efforts to lower thresholds through incorporation, such as quantum dashes in cascade designs, have shown broadband emission but remain focused on mid-infrared rather than direct hybrids in recent demonstrations. Terahertz (THz) QCL developments continue to push boundaries in single-mode operation and output power. Recent designs have achieved continuous-wave (CW) emission with powers exceeding 100 mW at frequencies around 3.9 THz, using optimized phonon-depopulation schemes for low thresholds. Single-mode THz QCLs at approximately 2 THz have demonstrated CW powers above 1 mW under , with surface-emitting configurations yielding narrow divergences of 4.4° for efficient coupling. Integration with metasurfaces, such as all-dielectric THz metalenses, enhances focusing by achieving numerical apertures up to 0.99 and sub-wavelength spot sizes, improving beam delivery for without bulky . These advances enable , achromatic focusing from 0.2 to 0.9 THz, broadening THz QCL utility in . Looking ahead, THz QCLs hold promise for room-temperature in networks, leveraging their narrow linewidths for high-data-rate links. QCL-based systems are emerging for autonomous vehicles, offering weather-resilient ranging with THz penetration through fog. In , compact THz QCL probes enable non-invasive for tissue imaging, detecting water content variations in real-time. Market analyses project the QCL sector to grow to approximately $739 million by 2035, fueled by demand in climate monitoring (e.g., sensing) and healthcare diagnostics.

References

  1. [1]
    Quantum Cascade Laser - Science
    A semiconductor injection laser that differs in a fundamental way from diode lasers has been demonstrated. It is built out of quantum semiconductor structures.
  2. [2]
    30 years of the quantum cascade laser | Communications Physics
    Dec 4, 2024 · In a Quantum Cascade Laser (QCL) the photon emission is obtained via quantum jumps among electronic states created by quantum confinement in ...
  3. [3]
    Quantum cascade lasers: 20 years of challenges
    ### Summary of Quantum Cascade Lasers
  4. [4]
    Possible amplification of electromagnetic waves in a semiconductor ...
    ... intersubband transitions in a repeated stack of semiconductor multiple quantum well heterostructures (Kazarinov, 1971) . The devices are carefully designed ...
  5. [5]
    Laser Frequency Combs with Fast Gain Recovery: Physics and ...
    Dec 16, 2021 · When first demonstrated in 1994, mid-infrared QCLs were multimode, low-power lasers operating in pulsed mode at cryogenic temperatures.1 Introduction · 1.2 Linewidth Enhancement... · 3 Harmonic Frequency Combs<|control11|><|separator|>
  6. [6]
    Continuous Wave Operation of a Mid-Infrared Semiconductor Laser ...
    Continuous wave operation of quantum cascade lasers is reported up to a temperature of 312 kelvin. The devices were fabricated as buried heterostructure lasers.
  7. [7]
    (PDF) Quantum Cascade Lasers Emitting Below 3 μm - ResearchGate
    Aug 6, 2025 · Quantum cascade lasers emitting below 3 μ m are demonstrated. The lasers based on the In As / Al Sb material system emit at 2.95–2.97 μ m in ...
  8. [8]
    Terahertz semiconductor-heterostructure laser - Nature
    May 9, 2002 · Here we report a monolithic terahertz injection laser that is based on interminiband transitions in the conduction band of a semiconductor (GaAs/AlGaAs) ...Missing: paper | Show results with:paper
  9. [9]
    1 Quantum devices | Quantum Cascade Lasers - Oxford Academic
    It describes interband and intersubband transitions, which are transitions between electrons and holes, and between the electron (or holes) confined states, ...<|control11|><|separator|>
  10. [10]
    Gallium Arsenide - an overview | ScienceDirect Topics
    Gallium arsenide has a direct bandgap of 1.424 eV at room temperature and the temperature dependency is typically assumed to follow the relation. Sign in to ...
  11. [11]
  12. [12]
    Active-region designs in quantum cascade lasers - IOPscience
    This paper analyses the development of active-region designs in quantum cascade lasers. Active-region designs have been demonstrated to date that employ ...
  13. [13]
    Quantum-cascade lasers based on a bound-to-continuum transition
    Jan 8, 2001 · A quantum-cascade structure combining the advantages of the three-quantum well and superlattice active regions is demonstrated.
  14. [14]
    [PDF] Step-Tapered Active-Region Mid-Infrared Quantum Cascade Lasers ...
    The first room-. Page 16. 3 temperature operation of QCLs in pulsed mode was demonstrated in 1996 [9] by utilizing a low-doped, funnel-injector and the vertical ...<|separator|>
  15. [15]
    Multi-wavelength distributed feedback quantum cascade lasers for ...
    Jul 5, 2019 · We review the progress of multi-wavelength distributed feedback (DFB) quantum cascade lasers (QCLs) as sources for broadband mid-infrared (mid-IR) spectroscopy.
  16. [16]
    Quantum Cascade Lasers - RP Photonics
    The operation voltage can easily be of the order of 10 V, whereas a few volts are sufficient for ordinary laser diodes. schematic of quantum cascade laser ...
  17. [17]
    [PDF] New frontiers in quantum cascade lasers: high performance room ...
    May 26, 2020 · Here we review recent research that has led to a new class of QCL light sources that has overcome these limitations leading to room temperature ...
  18. [18]
    Volume Fabrication of Quantum Cascade Lasers on 200 mm-CMOS ...
    Apr 10, 2020 · InAs based QCL are also good candidates in particular for short wavelengths around 3 µm. More recently, Baranov and co-workers have ...<|separator|>
  19. [19]
    Recent advances in mid infrared (3-5µm) Quantum Cascade Lasers
    Oct 10, 2013 · The GaInAs/AlInAs on InP has a conduction band offset of only 0.52 eV in lattice matched condition. This value can be increased to as high as ...
  20. [20]
    High-power, high-wall-plug-efficiency quantum cascade lasers with ...
    Jul 25, 2025 · In this paper, we are going to review the development of QCLs and present the strategies and methodologies to achieve high-power, high-WPE ...Qcl Design And Simulation · Mid-Infrared Qcl · High Brightness Lwir Qcls At...
  21. [21]
    Continuous-wave GaAs/AlGaAs quantum cascade laser at 5.7 THz
    Jan 16, 2024 · For the GaAs/AlGaAs system, we consider electron transport from carriers in the Γ-valley with an effective mass of 0.067m 0, where m 0 is the ...
  22. [22]
    Barrier Height Tuning of Terahertz Quantum Cascade Lasers for ...
    Oct 17, 2018 · The standard material system for THz QCLs GaAs/AlxGa1-xAs, which remains a direct band gap semiconductor up to 45% aluminum, allows the ...
  23. [23]
    Wavelength tuning of GaAs/AlGaAs terahertz quantum cascade ...
    Experimental details. A set of five structures with Al barrier concentrations varying from 13.5% to 15.5% was grown. · Results and discussion. Fig. · Conclusion.
  24. [24]
    (PDF) High-power short-wavelength quantum cascade lasers
    Aug 9, 2025 · We have investigated intersubband transitions in strain-compensated AlAs/(In,Ga)As heterostructures, demonstrating both absorption and quantum- ...
  25. [25]
    Short-wavelength (λ≈3.3μm) InP-based strain-compensated ...
    Nov 22, 2006 · The authors describe the design and implementation of a short-wavelength quantum-cascade laser emitting at approximately 3.3μm at 80K.Missing: AlGaAs | Show results with:AlGaAs
  26. [26]
    Room temperature operation of n-type Ge/SiGe terahertz quantum ...
    We compare the temperature dependence of the terahertz gain properties with an equivalent GaAs/AlGaAs quantum cascade laser design. In the Ge/SiGe case, the ...
  27. [27]
    Mid-infrared III–V semiconductor lasers epitaxially grown on Si ...
    Jun 1, 2022 · Antimonides have a number of advantages for use in quantum-cascade lasers (QCLs) compared with other material systems. The conduction band ...
  28. [28]
    InAs-Based Quantum Cascade Lasers with Extremely Low Threshold
    Oct 10, 2022 · We report InAs-based quantum cascade lasers (QCLs) operating near 14 μm with a threshold current density J th as low as 0.6 kA/cm 2 at room temperature.Missing: II | Show results with:II
  29. [29]
    Doping in quantum cascade lasers. I. - AIP Publishing
    Aug 17, 2006 · A companion paper presented the influence of the doping concentration on lasing performance for QCL's emitting in the terahertz region. Further ...
  30. [30]
    Rebalancing of internally generated carriers for mid-infrared ...
    Dec 13, 2011 · Mid-infrared interband cascade lasers (ICLs)8,9 generally have the advantage of much longer lifetimes characteristic of the interband active ...Missing: avoid | Show results with:avoid
  31. [31]
    Influence of doping on the performance of terahertz quantum ...
    Mar 5, 2007 · The doping concentration has a strong impact on the performance of a terahertz QCL, as it determines the number of free carriers inside the ...
  32. [32]
    Leakage current in quantum-cascade lasers through interface ...
    Oct 14, 2013 · The impact of interface roughness (IFR)-scattering on the quantum efficiency of quantum-cascade lasers (QCLs) is demonstrated and analyzed ...
  33. [33]
    Linewidth Broadening in Short-Wavelength Quantum Cascade Lasers
    Jan 18, 2022 · The decreased broadening due to IR scattering is not due to very smooth interfaces, but rather due to the finite carrier concentration and the ...
  34. [34]
    [PDF] Fabrication, Characterisation and Applications of Mid-Infrared ...
    Quantum cascade lasers (QCLs) are unipolar semiconductor lasers based on intersubband transition, mainly operate in the mid-infrared (mid-IR), far-infrared (FIR) ...<|control11|><|separator|>
  35. [35]
    [PDF] Low threshold InAs-based interband cascade lasers grown by MBE
    Reflection high energy electron diffraction (RHEED) was used for in-situ monitoring of the as-grown sample surface. All the InAs substrates were thermally ...
  36. [36]
    Quantum cascade lasers grown by MOCVD
    Oct 15, 2023 · This review summarizes the recent progress of QCLs grown by MOCVD. Material quality and the structure design together determine the device performance.
  37. [37]
    Etching Processing of InGaAs/InAlAs Quantum Cascade Laser - MDPI
    The ridge waveguide table structure of the quantum cascade laser can be prepared by dry etching or wet etching. Bewley found that a ridge waveguide prepared by ...Missing: fabrication steps
  38. [38]
    An Overview on Quantum Cascade Lasers: Origins and Development
    Apr 5, 2017 · This chapter presents an introductory review on quantum cascade lasers (QCLs). An overview is prefaced, including a brief description of their beginnings and ...Quantum Cascade Lasers · 2. Fundamentals And... · 5. State Of The Art And...
  39. [39]
    Compact, thermoelectrically cooled surface emitting THz QCLs ...
    Jun 30, 2025 · We present a 3.9 THz quantum cascade laser operating in a compact high heat load (HHL) housing. The device features a loop mirror back reflector ...Missing: epoxy- challenges
  40. [40]
    [1911.06582] Thermoelectrically cooled THz quantum cascade laser ...
    Nov 15, 2019 · Abstract: We present a \MF{terahertz} quantum cascade laser operating on a thermoelectric cooler up to a record-high temperature of 210.5 K.Missing: packaging epoxy- challenges
  41. [41]
    Very short wavelength (λ=3.1–3.3μm) quantum cascade lasers
    Quantum cascade lasers emitting at wavelengths as short as 3.1–3.3μm are reported. Such high intersubband emission energies (up to 400meV) have been ...
  42. [42]
    Quantum cascade lasers: from tool to product
    The quantum cascade laser is a powerful tool for spectroscopy. The wide spectral coverage, combined with narrow linewidth output and high power enable a wide ...
  43. [43]
    Ultrabroadband Heterogeneous THz Quantum Cascade Laser - NIH
    Dec 21, 2022 · The devices lase in pulsed and continuous-wave operation and emit in a spectral range from 1.9 to 4.5 THz, covering a bandwidth of 1.37 octaves.
  44. [44]
    Quantum cascade laser: 30 years of discoveries - PMC
    Oct 21, 2025 · The first quantum cascade laser was demonstrated in Bell Laboratories in 1994. This device featured characteristics that distinguished it ...
  45. [45]
    Quantum cascade lasers: 20 years of challenges
    Feb 20, 2015 · We review the most recent technological and application advances of quantum cascade lasers, underlining the present milestones and future ...2. Mid-Ir Quantum Cascade... · 2.2 Frequency Comb Operation... · 3. Terahertz Quantum Cascade...
  46. [46]
    Mid-infrared quantum cascade laser waveguides with non-vertical ...
    Confinement factor as a function of the sidewall angle in the ridge waveguide of the width of d = 9.5 μm for the first three lowest-order transversal modes.Mid-Infrared Quantum Cascade... · Highlights · Abstract
  47. [47]
    Buried heterostructure quantum cascade lasers with high ...
    Aug 13, 2007 · The authors report on the development of λ∼4.7μm strain-balanced InP-based quantum cascade lasers with high wall plug efficiency and room ...
  48. [48]
    Buried heterostructure quantum cascade lasers - SPIE Digital Library
    (BH-QC) lasers. The buried heterostructure is formed by regrowth of InP lateral on the side walls and on top of the InAlAs/InGaAs laser structure by ...
  49. [49]
    [PDF] Terahertz quantum-cascade laser at ␭É100 ␮m using metal ...
    We report lasing at ⬃3.0 THz (␭⬇98– 102 ␮m) in a quantum-cascade structure in which mode confinement is provided by a double-sided metal waveguide.
  50. [50]
    Double-metal waveguide terahertz difference-frequency generation ...
    Oct 15, 2018 · High THz mode confinement in the laser waveguide enabled by the double-metal waveguide configuration increases the mode interaction with the ...
  51. [51]
    Design and fabrication of terahertz quantum cascade laser with ...
    quantum cascade laser (THz QCL) with double metal waveguide via low-temperature In-Au wafer bonding followed by SI-GaAs substrate removal. ICP RIE in BCl3 ...
  52. [52]
    Correction to the effective refractive index and the confinement factor ...
    Abstract. The equations for the effective medium refractive index and for the confinement factor in the waveguide design for quantum cascade lasers are derived.
  53. [53]
    High-Power Fabry-Perot QCLs on D-Mounts - Thorlabs
    Each D-mount QCL has a high-reflection (HR) coated back facet and a partial-reflection (PR) coated front facet (see the Appearance tab for details). While these ...Missing: cleaved threshold
  54. [54]
    [PDF] High-performance midinfrared quantum cascade lasers
    Nov 22, 2010 · The first QCL was invented and demonstrated3 at Bell. Labs in 1994 by Faist, Capasso, Sivco, Sirtori, Hutchinson, and Cho. The first ...
  55. [55]
    (PDF) Multimode regimes in quantum cascade lasers - ResearchGate
    A theoretical and experimental study of multimode operation regimes in quantum cascade lasers (QCLs) is presented. It is shown that the fast gain recovery ...
  56. [56]
    Key issues associated with low threshold current density for InP ...
    For the multi-mode F–P QCLs, CW operation at room temperature with a threshold current densities of 1.5–2.8 kA/cm2 at a wavelengths range between 4.3 and 9.3 μm ...
  57. [57]
    Quantum cascade lasers: from tool to product
    ### Summary of Fabry-Perot Configurations in Quantum Cascade Lasers
  58. [58]
    Room temperature continuous wave, monolithic tunable THz ...
    Mar 24, 2016 · We present a strong-coupled strain-balanced quantum cascade laser design for efficient THz generation based on intracavity difference frequency generation.
  59. [59]
    Quantum and Interband Cascade Lasers (QCLs and ICLs), 3 - 11 µm
    Some of Thorlabs' DFB quantum and interband cascade lasers are uniquely suited for gas sensing and analysis. Select high heat load QCLs and ICLs are designed to ...Missing: short | Show results with:short
  60. [60]
    Quantum cascade lasers for mid-infrared spectroscopy - ScienceDirect
    (a) Multi-mode emission spectrum of a GaAs/AlGaAs QCL measured at T=77 K. The laser is exhibiting more than 40 longitudinal modes centered at 1000 cm−1 (λ=10 μm) ...
  61. [61]
    Quantum Cascade Lasers (QCLs): Distributed Feedback, Two-Tab ...
    where λ is the wavelength, neff is the effective refractive index, and Λ is the grating period. By changing the effective index, the lasing wavelength can ...
  62. [62]
    High-performance uncooled distributed-feedback quantum cascade ...
    Mar 13, 2012 · Single-mode emission with a side-mode suppression ratio (SMSR) about 30 dB and in the entire current range is observed. We also measured the ...
  63. [63]
    US7349456B2 - Gain-coupled distributed quantum cascade laser ...
    Purely gain-coupled diffraction gratings may be realized for use in QCLs and other edge emitting lasers that lack a typical p-n junction.
  64. [64]
  65. [65]
    Observing the Intrinsic Linewidth of a Quantum-Cascade Laser
    The intrinsic linewidth of a quantum-cascade laser was found to be a few hundred hertz, much lower than predicted by standard models.
  66. [66]
    Kilohertz linewidth from frequency-stabilized mid-infrared quantum ...
    Frequency stabilization of mid-IR quantum cascade (QC) lasers to the kilohertz level has been accomplished by use of electronic servo techniques. With this ...
  67. [67]
    Distributed-feedback quantum cascade laser emitting at 3.2 μm
    Fabrication was done using e-beam lithography. The devices showed a low beam divergence (8°). QCLs show good thermal qualities, output stabilities and tuning ...
  68. [68]
    Frontiers | Advances in narrow linewidth and wide tuning range external-cavity wavelength-swept lasers
    ### Summary of External Cavity Quantum Cascade Lasers (EC-QCLs) from Frontiers Article
  69. [69]
    [PDF] Broadly Tunable External Cavity Quantum Cascade Laser ...
    In the Littman-Metcalf external cavity design, the collimated laser beam is directed to a diffraction grating. The first-order diffraction beam is then ...Missing: Littrow | Show results with:Littrow
  70. [70]
  71. [71]
  72. [72]
    Widely tunable quantum cascade lasers with coupled cavities for ...
    Nov 5, 2010 · The authors report the fabrication of widely tunable monolithic quantum cascade lasers (QCLs) with coupled Fabry–Pérot (FP) cavities on ...
  73. [73]
    Absolute spectroscopy near 7.8 μm with a comb-locked extended ...
    Jan 22, 2018 · The locking is obtained by slow feedback to the EC-QCL piezo with a 100 Hz servo bandwidth, which results in a 100 kHz frequency stability over ...
  74. [74]
    External cavity tunable quantum cascade lasers and their ...
    Tunable external cavity (EC)-QCLs in particular offer narrow linewidths, wide ranges of tunability, and stable power outputs, which open up new possibilities ...
  75. [75]
    [PDF] ~8.5 μm-emitting InP-based quantum cascade lasers grown on ...
    Oct 26, 2022 · Razeghi12 reported high peak-pulsed power (>10 W from both facets). InGaAs ... In conclusion, we demonstrate the 8lm-emitting InP-based quantum cascade laser ...
  76. [76]
    Watt-Level Room Temperature Continuous-Wave Operation of ...
    Feb 25, 2013 · CW output power of 1.15 W is demonstrated at λ ~ 10.3 μm for a QCL which has a 6-mm-long cavity and a 14-μm-wide stripe.
  77. [77]
    Watt level performance of quantum cascade lasers in room ...
    Oct 1, 2010 · The slope efficiency of the tested QCL near threshold is 2.16 W/A at RT. The cw characteristic temperature, T 0 is calculated to be 128 K ...
  78. [78]
    Analysis of limitations to wallplug efficiency and output power for ...
    Jun 26, 2006 · We have carried out detailed optical and thermal simulations of quantum cascade laser (QCL) operation as a function of material and device ...<|control11|><|separator|>
  79. [79]
    Analysis of the Thermal Roll-Over of Quantum Cascade Lasers
    A description of thermal roll-over in room temperature continuous-wave operating quantum cascade lasers is presented. Models predicting maximum operating ...
  80. [80]
    (PDF) Impact of interface roughness distributions on the operation of ...
    Due to interface roughness scattering, the lower laser state depopulation lifetime of Quantum Cascade lasers can be significantly reduced from 0.25 to 0.09 ...Missing: ML | Show results with:ML
  81. [81]
    Aspects of the internal physics of InGaAs∕InAlAs quantum cascade ...
    Jun 5, 2006 · Carrier-carrier scattering (the Auger interaction) is also an important mechanism in QCLs, particularly in those devices with closely spaced ...
  82. [82]
    Free-carrier absorption in quantum cascade structures | Phys. Rev. B
    Feb 22, 2012 · We show that the free-carrier absorption in quantum cascade lasers (QCLs) is very small and radically different from the classical Drude ...Missing: Auger | Show results with:Auger
  83. [83]
    High-Power Mid-Infrared (λ∼3-6 μm) Quantum Cascade Lasers
    Feb 25, 2022 · 201103. [19] J. C. Shin, “Tapered active-region quantum cascade laser,” Ph.D. ... Botez, “10 W near- diffraction-limited pulsed power from 0.98 μm- emitting ...
  84. [84]
    Sub-ppb-level CH 4 detection by exploiting a low-noise differential ...
    A highly sensitive CH4 sensor was demonstrated, implementing an interband cascade laser operating at 3.3 µm and a low-noise differential PAS cell. An Al ...Missing: CO2 | Show results with:CO2
  85. [85]
    [PDF] High precision laser spectrometer for multiple greenhouse gas ...
    Aug 13, 2020 · ... 4.3 µm QCL to further reduce optical feedback. High-current ... CO2, 0.02 ‰ for δ13C(CO2), 0.4 ppb for CH4. 370 and 0.1 ppb for N2O ...
  86. [86]
    Application of quantum cascade lasers to trace gas analysis
    Dec 8, 2007 · Quantum cascade (QC) lasers are virtually ideal mid-infrared sources for trace gas monitoring. They can be fabricated to operate at any of a very wide range of ...
  87. [87]
    [PDF] Quantum Cascade Lasers (QCLs) for Standoff Explosives Detection
    This LDRD was undertaken to investigate the benefit of the possible use of quantum cascade lasers (QCLs) in standoff explosives detection equipment. Standoff ...Missing: destructive screening
  88. [88]
    Terahertz imaging using quantum cascade lasers—a review of ...
    Aug 28, 2014 · In this paper we review the technological approaches to THz QCL-based imaging and the key advancements within this field.
  89. [89]
    Free‐Space Communications Enabled by Quantum Cascade Lasers
    Aug 19, 2020 · A promising source for mid-IR free-space communications is the semiconductor quantum cascade laser (QCL), which can be directly modulated at a high speed.Introduction · Summary of Reported QCL... · Experimental Setup of a Mid...
  90. [90]
    Study of the Exhaled Acetone in Type 1 Diabetes Using Quantum ...
    The acetone concentration exhaled in the breath of three type 1 diabetes patients (two minors and one adult) and one healthy volunteer is studied using a ...Missing: endoscopy | Show results with:endoscopy
  91. [91]
    Infrared Spectroscopy with a Fiber-Coupled Quantum Cascade ...
    Nov 23, 2019 · Infrared Spectroscopy with a Fiber-Coupled Quantum Cascade Laser for Attenuated Total Reflection Measurements Towards Biomedical Applications.Missing: endoscopy | Show results with:endoscopy
  92. [92]
    Implementation of a quantum cascade laser-based gas sensor ... - NIH
    Sep 17, 2016 · Sensitive and selective detection of hydrogen sulfide (H2S) is essential for production control and environmental monitoring purposes in the ...Missing: pollution | Show results with:pollution
  93. [93]
    Continuous Emissions Monitoring with Quantum Cascade Laser ...
    Apr 5, 2023 · Keith Linsley shares how Quantum Cascade Laser & Tunable Laser Diode technologies together enable reliable Continuous Emissions Monitoring.Missing: petrochemicals | Show results with:petrochemicals
  94. [94]
    Best-In-Class Mid-Infrared, Quantum Cascade Laser Provider
    DRS Daylight Solutions, a line of business under Leonardo DRS, provides photonics solutions to researchers, industry allies, and system manufacturers.Careers · Products · About Us · Contact Us
  95. [95]
    [PDF] January 2012 - Pranalytica
    Jan 1, 2012 · Quantum Cascade Lasers (QCLs) represent a fundamentally new class of semiconductor lasers that are capable of generating light in the mid-.
  96. [96]
    Quantum Cascade Laser Market | Global Market Analysis Report
    Oct 22, 2025 · Quantum Cascade Laser Market was worth USD 455.8 million in 2025, and is predicted to grow to USD 1011.8 million by 2035, with a CAGR of ...Missing: modules 2005
  97. [97]
    Continuous wave quantum cascade lasers with 5.6 W output power ...
    May 19, 2020 · We demonstrate a high continuous wave wall-plug efficiency of an InP-based quantum cascade laser reaching ∼41% with an output power of ∼12 W from a single ...
  98. [98]
    High power RT quasi-CW 4.6um quantum cascade laser (QCL) arrays
    Mar 15, 2023 · An 8-emitter laser array, epi-down mounted on copper heat sink, outputs ~12 watts peak optical power @ 20us pulse width 4% Duty Cycle, and ~8 Watts with 100us ...
  99. [99]
    9.4 μm emitting quantum cascade lasers grown by MOCVD
    Jan 15, 2025 · This study undertakes an exhaustive analysis of QCL devices, comparing those with and without aluminum content compensation, through a combination of ...
  100. [100]
    (Invited) Transistor-Injected Quantum-Cascade Devices - IOP Science
    To overcome these limitations, the transistor-injected quantum-cascade laser (TI-QCL) has been proposed [2]. The TI-QCL is a three-terminal device in which ...
  101. [101]
  102. [102]
    Monolithic beam combined DFB quantum cascade laser arrays with ...
    Mar 19, 2025 · The chip combines 5 lasers operating near 4.65 µm wavelength in pulsed mode. The peak power of the best QCL element reached 1 W with a 2 kA/cm² ...
  103. [103]
    Beam Combining of a Tunable Tilted Quantum Cascade Laser Array
    Jul 17, 2025 · A tunable quantum cascade laser (QCL) array utilizing tilted waveguides, spanning 4.51–4.61 μm, is reported. By integrating S-bends, Y-branches, ...
  104. [104]
    Innovative Integration of Dual Quantum Cascade Lasers on Silicon ...
    Aug 22, 2024 · We demonstrate the hybrid integration of dual distributed feedback (DFB) quantum cascade lasers (QCLs) on a silicon photonics platform using an innovative 3D ...
  105. [105]
    Hybrid Si/III-V Quantum Cascade Lasers Integrated on a Phase ...
    Aug 26, 2025 · In this work, we demonstrate the heterogeneous integration of a QCL onto a novel high-index-contrast, phase-matched silicon photonic platform ...
  106. [106]
    [PDF] Hybrid Integration of Quantum Cascade Lasers with Germanium-on ...
    Abstract: We present a novel scheme for hybrid integration of quantum cascade laser bars with germanium-on-silicon waveguides operating in the mid-infrared.
  107. [107]
    Continuous-wave terahertz quantum cascade laser based on a ...
    In this work, we designed a hybrid active structure by tailoring quantum transport process to improve the quantum efficiency and threshold for THz QCLs.
  108. [108]
    High brightness terahertz quantum cascade laser with near ... - Nature
    Aug 16, 2024 · The quantum cascade surface-emitting laser is capable of delivering an output peak power over 185 mW with a narrow beam divergence of 4.4° × 4.4° at 3.88 THz.Missing: review | Show results with:review
  109. [109]
    High focusing efficiency metalens with large numerical aperture at ...
    The experimental results show that the designed metalens is able to increase the focusing efficiency from 22.8% to 41.7%. The full widths at half maxima (FWHMs) ...
  110. [110]
    Terahertz Lasers Are the Key to Ultra-Fast Wireless Communication
    Aug 29, 2025 · The ability to operate at room temperature and the potential for integration with silicon photonics platforms make THz lasers especially ...Missing: autonomous | Show results with:autonomous
  111. [111]
    Autonomous cars drive terahertz research - SPIE
    Mar 1, 2021 · Terahertz signals can also be used as a form of radar, as a substitute for or complement to cameras and lidar for imaging other vehicles, people ...Missing: temperature QCL endoscopy
  112. [112]
    [PDF] Terahertz endoscopy of hard-to-access objects in the context of ...
    206. Nazarov, M. et al. A flexible terahertz waveguide for delivery and filtering of quantum-cascade laser radiation. Applied Physics Letters. 113, 131107 ...
  113. [113]
    Quantum Cascade Laser Market Size, Share, Growth & Forecast 2035
    Sep 11, 2025 · Quantum Cascade Laser Market size was over USD 466.92 million in 2025 and is projected to reach USD 739.11 million by 2035, witnessing around 4 ...