Voltage-Controlled and Injector Layer Thickness-Dependent Tuning of Quantum Cascade Laser for Terahertz Spectroscopy (2025)

1. Introduction

Click to copy section linkSection link copied!

Over the past few decades, terahertz radiation has emerged as a promising area of research, filling the gap between the microwave and infrared regions of the electromagnetic spectrum. Its unique properties, including its nonionizing nature and ability to penetrate various materials, make it an ideal candidate for a wide range of applications. One example of such an application is spectroscopy, which allows the analysis of the chemical composition of materials with exceptional sensitivity. (1,2) Terahertz spectroscopy is also used in construction and architecture, (3−5) environmental monitoring, (6−10) pharmaceuticals, (11,12) medicine, (13−15) and even in the field of art conservation. (16−18) The ability to precisely identify the molecular structure of a substance without causing any damage makes THz spectroscopes an indispensable part of airport security systems (19−21) or the main tool for obtaining information about the internal structure of materials or objects. (22,23) Such systems are based on the fact that forces arising from intra- and intermolecular interactions and their radiofrequency sensitivity to atomic structure. (24) Intermolecular bonds cause rotation of the rigid body, movement of the center of mass and internal vibrational modes, which under the influence of THz radiation become the source of unique spectral features. These features allow the chemical identification of samples and the so-called chemical mapping in two (25) or even three geometric dimensions (26) in the time (27,28) for frequency (29,30) domain. An example of such a system is shown in Figure 1. The target object is raster scanned using the THz-TDS (Terahertz Time Domain Spectroscopy) technique. (31−35) This technique uses short electromagnetic pulses of the order of picoseconds, allowing the measurement of signals with very high resolution and the observation of the dynamics of phenomena occurring in the tested materials over very short time intervals. The THz-TDS system consists of several key components that work together to provide precise analysis of the target. The main source of terahertz radiation is the QCL (Quantum Cascade Laser) whose beam is directed to the beam splitter. This element splits the beam into two parts: one part goes to the reference chamber, where the radiation parameters are analyzed under reference conditions, while the other part goes to the object under study and then to the THz camera, which acts as a detector. The object under investigation is located in the beam path between the beam splitter and the THz camera, which allows the analysis of its optical and structural properties. The reference chamber is connected to the control electrical units, which allow to control the measurement conditions and to verify the results against the reference standard signal.

Figure 1

High Resolution Image

Download MS PowerPoint Slide

The power supply and control system includes several key elements: Impulse Power Supply, which supplies power to the controller and the QCL module (terahertz laser emitter), Control Electric Units, which work with PC software to regulate measurement parameters and operating conditions of the reference chamber, PC software, which acts as a central control system and enables communication with the Impulse Power Supply, controller, Control Electric Units, and the THz camera. All system elements are interconnected in a way that allows precise control and data acquisition. The PC software acts as a master control module, enabling bidirectional communication with the Impulse Power Supply, Controller, Control Electric Units and THz Camera. Impulse Power Supply has a unidirectional connection with Controller and QCL Module, ensuring stable power supply to the system. In turn, the THz camera is powered by the power supply, which enables precise detection of the terahertz signal. The entire system allows precise imaging of the object under investigation, providing high quality data on its internal structure and optical properties.

One of the main elements of the system described here is a quantum cascade laser, which has a relatively short but rich history. (36−38) The schematic operation of this device is illustrated in Figure 2.

Figure 2

High Resolution Image

Download MS PowerPoint Slide

The device consists of many nanometer-sized alternating semiconductor layers with two different lattice constants (e.g., AlGaAs–GaAs), which form modules containing quantum wells (see Figure 2b). Such structures are called superlattices (SL) and are an essential part of QCL. The idea is that a one electron is the source of a cascade transition through the superlattice structure combined with the emission of many photons (see Figure 2a). Most importantly, it has become a leading emitter in the systems described here due to its relatively easy tunability and its ability to provide high spectral power densities at target frequencies above 2–4 THz, where many materials exhibit spectral absorption properties. As a result, research centers around the world are continuously working on improving and extending the capabilities of existing QCL structures, (39−41) as well as finding new design solutions. (42,43) A special role is played by work on increasing the operating temperature of lasers and their tuning ranges, (44−46) which is crucial for improving their energy efficiency and enabling their operation at room conditions without the need for complicated cooling. In addition, the development of new semiconductor materials and advanced heterostructuring techniques allows the design of QCL structures with increased quantum efficiency, opening the way to their applications in spectroscopy, optical communication and biosecurity. (47−49) An indispensable tool in these activities are numerical models and simulation programs, (50−53) which significantly reduce the huge costs of developing and implementing innovative low-dimensional structures and also speed up the process of reaching the final design solutions of modern quantum devices.

Our research group has developed several complementary numerical models (54,55) that can be used to simulate low-dimensional structures acting as emitters or detectors of electromagnetic radiation. The use of a specific model results from the nature of the research and the desire to optimize its duration. For example, the fast models IMSL (also known as Wannier Function Method - WFM) (56) and FMSL (57) are most often used for preliminary simulations for specific power supply conditions of the structure, in order to determine the range of computational parameters important from the point of view of the research. Then, for such a determined range, one can use the very versatile, but also time-consuming RSM. (58)

This paper presents numerical studies carried out to determine the tunability of the selected QCL structure described in, (59,60) which could be used in a terahertz imaging system. The above structure was optimized by varying the thickness of the semiconductor layers responsible for extracting carriers from the active region of the laser and delivering them to the active region of the next device module (injector region). The calculations included checking the possibility of emitting THz radiation over a wide range of supply voltages, as well as determining the influence of temperature on the current–voltage characteristics of the device, taking into account all the necessary dissipation mechanisms.

2. Numerical Models of QCL

Click to copy section linkSection link copied!

The simulation of quantum cascade lasers is a very complex process that requires a lot of computational resources and can take a very long time, depending on the models and objectives adopted. Therefore, the strategy adopted in this work is to use fast but less accurate models (IMSL, FMSL) to check the possibility of radiation emission in a wide range of supply voltage variations for different variants of injector layer thickness. On the other hand, the accurate RSM model has been used to determine accurate maps of laser radiation for selected, most interesting configurations of supply parameters and injector layer thickness.

2.1. IMSL and FMSL Models

To observe the location of the most important quantum states from the perspective of laser action, effective IMSL and optimized FMSL (61) models were employed. The first one assumes infinite dimensions of the model and periodic boundary conditions. It uses the properties of Bloch and Wannier functions to create a basis of quantum states spanning three superlattice periods and relies on the nonequilibrium Green’s function formalism (NEGF) to determine the transport parameters of the structure. This approach involves solving the Dyson and Keldysh equations to obtain the delayed Green’s function, GR(E), and the Green’s correlation function, G<(E). The Dyson equation is written as

(EIHΣR)GR=I

(1)

where H represents the device Hamiltonian, GR is the retarded Green’s function matrix, and ΣR is the self-energy matrix. Initially, the self-energies, ΣR were treated as constant diagonal elements iη, with the parameter η defined by

GR(E)=limη0+G(Z=E+iη)

(2)

The Keldysh equation takes the following form:

G<=GRΣ<GR

(3)

where G< is the correlation Green’s function matrix, and Σ< denotes the self-energy matrix. In the initial state (thermodynamic equilibrium), this is a diagonal matrix with elements iη·fn (E), where

fn(E)=1/[exp((EEF)/kBT)1]

(4)

where the EF values are obtained using either the step function or a Büttiker probe-inspired (62,59) approximation. In contrast, our approach introduces a novel method for calculating the GR (E) and G<(E) functions, as well as the self-energies matrix, which, to the best of our knowledge, is unique among superlattice models. It is worth emphasizing that in the method the Hamiltonian device matrix is transformed from the energy and position representation to the pure energy representation. Thanks to this, its size is reduced by 2 to 4 orders, which significantly speeds up the calculations and makes it the fastest model we use. Therefore, a large number of simulations of structures with different configurations of supply voltages and injector layer thicknesses were performed on it.

The advantage of FMSL model is the use of polynomials to approximate the charge and potential distribution functions in a single layer of the superlattice. This allows for semianalytical solutions of the Schrödinger and Poisson equations, taking into account the changing potential in a single layer of the structure without the need to divide it into many subregions after applying a voltage to the system. Schrödinger equation in this approach has the form:

2ψu2+[1Wj(u)]ψ(u)=0

(5)

where

Wj(u)=k=0Mdj,kuk,dj,k=bj,kEσjk,σj=2mejeE

(6)

The dimensionless form of eq 5 was obtained after introducing dimensionless variables:

uj=2mejeE(zzj),Wj(u)=Vj(z)E

(7)

where Vj(z) is the potential function in the j-th layer of the superlattice and olutions are represented as a power series:

2ψu2+[1Wj(u)]ψ(u)=0

(8)

Poisson’s equation in the form

ddzε(z)dVS(z)dz=eρ(z)

(9)

where ε is dielectric function and VS(z) denotes the potential derived from impurities and unbalanced charge carriers, with the continuity conditions:

Vj(zj+1)=Vj+1(zj+1)

(10)

εjdVjdz|zj+1=εj+1dVj+1dz|zj+1

(11)

is solved by assuming that the right-hand side of eq 9 is approximated by an N-degree polynomial in the jth layer of the superlattice in the form

ρj(z)=k=0Naj,k(zzj)k

(12)

The approximation procedure consists in comparing the charge density function calculated in the previous iteration step with the function described by formula 12 at N + 1 selected points (nodes) of each structural layer. This leads to algebraic equation systems of N + 1 degree, the solution of which gives the values of the coefficients aj,k A detailed description of the method mentioned here, together with sample calculation results, is described in the works. (56,57)

This model has been used to simulate the most interesting cases of input parameter configuration (supply voltage and injector thickness), selected after analyzing the results obtained by the IMSL method. The FMSL model allows taking into account the mutual interaction of a larger number of superlattice modules (IMSL considered 3 periods), which allows a better approximation of the structure consisting of several tens of modules.

Example simulations performed using the IMSL and FMSL models are shown in Figure 3, which shows the test result of the QCL structure initially calculated for three superlattice periods (IMSL) and extended to four superlattice modules. These modules are marked as #0 in the above paper. The quantum states visible here were obtained for a temperature of T = 50 K and a voltage of U = 50 mV per period using the Modified Bisection Method (MBM) with a parameter of NL = 50. Such a value of the NL parameter is responsible for the number of analyzed compartments in the unit energy range, which provides high accuracy in the determination of eigenstates with optimal computational time. The algorithm of the method, together with a detailed description of its other parameters, has been published in the ref (61).

Figure 3

High Resolution Image

Download MS PowerPoint Slide

The choice of temperature was dictated by the well-known fact that for 50 K all dopants are ionized. In this case, it was possible to detect 60 eigenstates grouped in the range of 12 minibands. The most important eigenstates for the operation of the laser were then plotted. The position of the states in the energy domain was calculated with the accuracy defined by two parameters of the Transfer Matrix (TM) procedure, in which the wave functions are represented by power series. The first of these parameters (nKS = 30) determines the number of terms in the power series within the procedure for calculating the eigenenergies, while the second (nKD = 100) refers to the number of terms in the power series within the procedure for calculating the waveforms of the wave functions. A detailed description of the method and its parameters has been published in. (57) Furthermore, the influence of the charge originating from the 5 nm doped areas located in the injector cavity (marked in gray) was taken into account, which was determined by the self-consistent solution of the Schrödinger and Poisson equations. Simulations in this regard were carried out according to the algorithm described in the paper. (57)

In the active region of the laser, which consists of two narrow (8.9 and 8.15 nm) quantum wells (numbers 1 and 2), two different states are distinguished: a high (d) state and a medium (c) state. These states, which are powered by electrons from the phonon well injector (3”) of the previous period, serve as the source of the photon transition d→c with energy = Edc. The electrons are then transported by the electric field to the extractor region (3), which also serves as an injector for the d′ and c′ states in the subsequent period of the structure. In the area of the phonon well (3), for the given supply voltage, there is a ba a transition with energy Eba = 35.1 meV. It is noteworthy that the energy of the b′a′ phonon transition is Eb′a′ = 35.7 meV, while for the b″a″ transition it is Eb″a″ = 36 meV. Consequently, we are dealing with an energy spectrum of phonon transitions analogous to the case of photon transitions. In the latter, the photon energy dc has an energy of Edc = 15.1 meV. However, for the d′c′ transition, we have Ed′c′ = 14.2 meV, while the transition d″c″ gives the energy Ed″c″ = 16.4 meV. It can thus be demonstrated that max, which corresponds to the energy of the transition b″a″, is distinguishable from min, which represents the energy of the transition d′c′. The distinction between these values provides insight into the width of the emitted energy spectrum, as illustrated in Figure 1b. This width, which is referred to in the work as Δν̂ (cm–1), is commonly understood as the wavenumber. The calculated value of the wavenumber using the FMSL for four modules of the tested structure supplied with voltage U = 50 mV/period is Δν̂ = 1610 cm–1. In contrast, the same parameter calculated for a structure containing ten modules increases to Δν̂ = 7017 cm–1 and remains relatively constant with further increases in the number of QCL modules. The data obtained in this manner, validated by more precise calculations using the RSM, constituted one of the input parameters for subsequent design calculations, which will be described in the subsequent chapter.

Laser action is contingent upon population inversion, wherein the electron density in the d miniband exceeds their concentration in the c miniband. This phenomenon can be observed in the energy map of electron concentration shown in the lower left corner of Figure 2, marked as (a), which was prepared for the structure #0 supplied with a voltage of U = 50 mV/period at T = 50 K. The results demonstrate a markedly higher electron density around the d state relative to the c state, which ensures dc photon transitions. Furthermore, the high electron density observed in the 3″ quantum well, resulting from the 5 nm doping layer at Nd = 6 × 1022 m–3, ensures that the energy levels of the d miniband are occupied by electrons injected by the electric field from the a miniband. Tunnelling processes occur between the injector well 3 and the well of the active region 2, which supply electrons to miniband b. These electrons are then transferred to miniband a through phonon transitions, which subsequently occupy the states. This miniband supplies the d′ state in the subsequent module. The sequence of charge carrier transport combined with multiple photon emission by the same electrons is therefore possible, and its quantitative value can be accurately calculated using the RSM, which will be described in the following chapters of the paper.

2.2. RSM Model

RSM is a semi-infinite model in which the multiquantum well alignment of the polarized superlattice structure is continued in the derivatives. It is assumed that the conduction band edge does not change outside the analyzed part of the structure. In this approach, the Keldysh and Dyson transport equations are solved assuming that the retarded Green’s functions can be written as

GR=f(z,z,k||,E)

(13)

where z and z′ are the real space coordinates. This is the main difference of this model from IMSL, as it does not require initial calculations of allowed minibands in the simulated structure. In contrast to IMSL, the quantum states in the superlattice can be obtained directly using the NEGF formalism. But on the other hand, the Hamiltonian matrices describing the device reach enormous sizes, which greatly increases the time of solving the transport equations and requires very large memory resources. Detailed differences between both models are described in. (54)

Example QCL simulations using RSM are presented in Figure 4, were obtained using the parameters given in Table 1. The calculations take into account electron scattering due to the crystal lattice disorder (AD), interface roughness (IR), scattering on impurity ions (ID), acoustic and optical phonons (AP and OP, respectively) and electron–electron (E–E) interactions according to their implementation method, as described in the works. (58,63,64) The IV characteristics of the tested QCL are presented in Figure 4 a, where the dependencies of the current density J (kA/cm–2) on the supply voltage U (V/period) are plotted. These were calculated for temperatures T = 200 K and T = 50 K, respectively, and are plotted in green and red. The first of the aforementioned characteristics was used to compare the simulation with real measurements made in the work. (59)

Table 1. Basic Parameters of the QCL Simulation for the Results Presented in Figures 36

GaAs/Al045Ga0,55As QCLwellbarrier
effective mass, m*0.0670.104
bandgap, Eg (eV)0.841.84
relative permittivity, εr12.8513.8
structurelayers,(nm) (bariersinbold)#0: 4.3, 8.9, 2.46, 8.15, 4.1, 16.0
#1: 4.3, 8.9, 2.46, 8.15, 4.1, 15.4
#2: 4.3, 8.9, 2.46, 8.15, 4.1, 16.6
bandoffset(eV), ΔEC0.12
ndop (cm–3)6×1016
LO-phonon energy (eV)0.036
deformation potential (eV)5.89
screeninglength(nm), lDebye32
no. of periods QCL236
temperature (K)50, 200

Figure 4

High Resolution Image

Download MS PowerPoint Slide

After calculating the point values for the given structure power conditions, it was found that the obtained simulation and measurement results converged, confirming a good fit of the model input parameters. The threshold value of current density JTH ≅ 1.195 kA/cm2 was determined by taking into account losses in the waveguide as described in the aforementioned work, where gsb ≅ 35 cm–1 was assumed as the threshold value. In the graph for T = 50 K, where the threshold current is JTH ≅ 1.025 kA/cm2, the portion of the characteristic exhibiting negative dynamic resistance (NDR) is noteworthy, which is not observed during measurements (see work in ref (59)). The physical reasons for such a phenomenon resulting from the assumption of perfectly repeatable geometric dimensions of each superlattice module for the simulation have been described in detail in ref (65). Nevertheless, this does not significantly impact the subject of our research.

The map of the laser optical gain (α (cm–2)) visible in Figure 4b, calculated for a voltage of U = 45 mV/period and a temperature of T = 50 K, plotted against the background of the potential of the structure, reveals several crucial aspects of laser operation. (i) It can be employed to present electronic transitions between quantum states in both the energy and position domains in a quantitative manner. (ii) It illustrates the areas of radiation emission (positive value of the α parameter) and its absorption (negative value parameter α). (iii) It provides information indicating that photon transitions are essentially limited to the active area of the laser and that their intensity is the highest in the energy range corresponding to the position of the d and c states.

The values of the α parameter with respect to the position can be summed to plot the characteristics shown in Figure 4c. This figure illustrates that the values of energy for positive optical gain are in the range maxmin = 10 meV, which gives Δν̂ = 8065 cm–1. This approximately corresponds to the calculations obtained using FMSL, but in this case, it was achieved with much less computational time and computer memory. However, the precise results visible in Figure 4c allow the determination of the optical gain peak, which in this case occurs for mG = 11 meV and has a value of gmax = 74 cm –1. Based on the mG, trend lines for the change in laser radiation caused by changes in the thickness of the injector and the power supply conditions of the structure were determined and will be described in the following chapters of the work.

3. QCL Tuning

Click to copy section linkSection link copied!

As demonstrated in the work, (59) the range of waves emitted by terahertz lasers can be modified by adjusting the power supply conditions in an appropriate manner. Based on the aforementioned experiments, the operation of the selected structure was simulated in a wide range of supply voltages to assess its tunability in terms of use in THz-TDS imaging systems. The results of these simulations are presented in Figure 5, which were carried out for T = 50 K. Figure 5 a illustrates the borders (in the energy domain) of the highest (max) and lowest (min) values of photon transitions depending on the structure’s supply voltage.

Figure 5

High Resolution Image

Download MS PowerPoint Slide

The calculations were conducted using FMSL with the input parameters specified in Section 2 for ten superlattice modules spanning the supply voltage range from 0 to 80 mV/period. Two additional scales, f (THz) and F (kV/cm), have been introduced in the chart to facilitate the conversion of the results obtained for Δ (meV) and U (mV/period), respectively. The supply voltage range responsible for tuning the laser above the JTH threshold has been specified, which is marked as QCLTR#0F (Tunning Range #0 FMSL). The graph illustrates that in the QCLTR#0F range, there are significant changes in the width of the radiation spectrum.

These changes are illustrated in part (b), where the dependence of Δ (meV) on the supply voltage U (mV/period) is plotted. These changes were converted into wavenumber values that can be read using an additional axis Δν̂ (cm–1). As illustrated in the graph, the width of the radiation spectrum of the tested laser reaches Δν̂max#0F = 6800 cm–1 (see for U = 60 mV/period), which corresponds to wavelength changes in the range from λmin = 77.5 μm to λmax = 154.9 μm (Δλ = 77.4 μm). As the supply voltage is increased beyond a value of U = 60 mV/period, the radiation spectrum narrows. However, the narrowest spectrum is observed in the range of U = 46–52 mV/period. The width of the spectrum drops even to Δν̂min#0 = 323 cm–1 (for U = 49 mV/period), which corresponds to changes in wavelength in the range from λmin = 91.8 to λmax = 94.6 μm (Δλ = 2.8 μm). The entire QCLTR#0F range allows the laser to be tuned to a value of fTR#0F = 1.9–4.4 THz, which corresponds to waves in the range λTR#0F = 68.1–157.8 μm.

A comprehensive illustration of radiation in specific energy ranges is shown in Figure 5c, which shows the gain map gain (cm–1) with respect to the device supply voltage U (mV/period). Calculations were performed using RSM, taking into account all major scattering mechanisms (AD, IR, ID, AP, OP, and EE) and assuming an energy resolution of 1 meV and the parameters given in Table 1. The chart shows that the highest radiation (gain > 100 cm–1) occurs for voltages U = 54–71 mV/period, where the device can be tuned in the range fTR#0R = 2.9–4 THz, which corresponds to the wavelength λTR#0R = 74.9–103.3 μm. The maximum width of the radiation spectrum is observed near around the voltage U = 55 mV/period, where Δν̂max100g = 995.4 cm–1 (for gain > 100 cm–1) and Δν̂max40g = 5503.4 cm–1 (for gain > 40 cm–1). The results obtained are consistent, with a small discrepancy, with those predicted by FMSL, as the calculations do not take into account the losses occurring in the device’s waveguide and all scattering mechanisms. Consequently, it can be concluded that the FMSL is a useful tool for rapidly estimating the trend and tuning range of the QCL, which is then corroborated by the trend line (shown in black) visible in map (c) and designated as mG, relating to the optical gain maxima (see Figure 4c). The collapse observed around U = 55 mV/period is directly related to the narrowing of the spectrum, as illustrated in Figure 4a.

4. Modeling of QCL Injector Region

Click to copy section linkSection link copied!

In their work, (59) the authors report that the thickness of the injection and extraction barriers was optimized prior to fabrication of the device based on experiments conducted using simple Monte Carlo (MC) and Matrix Density (MD) models. From our perspective, the critical location of state b (see Figure 6), which is connected to states d and c by a small but visible coupling, may act as a bottleneck between the active area and the injector.

Figure 6

High Resolution Image

Download MS PowerPoint Slide

Therefore, it seems logical to investigate its impact on the tunability of the structure. Consequently, a study was conducted within the framework of the present article to evaluate the impact of b-miniband changes caused by varying the width of the injector cavity Qw on the character of the radiation spectrum of the laser structure.

The methodology of the research entailed the execution of expeditious and simplified simulations of the structure for reduced and increased widths of Qw (±10%) in comparison to the original layer thickness (structure #0). Then, the most interesting cases were selected for detailed analysis, with precise calculations to include the most significant electron scattering mechanisms (AD, IR, ID, AP, OP, E-E). The two most interesting structures were created by changing the Qw width by approximately two monolayer (±0.6 nm). These structures are designated as #1 and #2 in the work, and the most significant results presented in Figure 7 pertain to these structures.

Figure 7

High Resolution Image

Download MS PowerPoint Slide

Figure 7 a depicts the boundaries of the radiation spectrum, expressed by the energy of the emitted photons (meV), as a function of the supply voltage U (mV/period), for the narrowed (in blue) and widened (in red) wells of the injector. The boundaries are represented by the values of min and max, which have been converted to their corresponding frequencies f (THz). Furthermore, the supply voltage has been converted to values of electric field and presented on an additional axis, F (kV/cm). The results were compared with those obtained when the injector well was represented by the original size (#0 in gray). The simulations in this section were conducted using FMSL with input parameters consistent with those described in Section 2, including a doping charge of ndop = 6 × 1016 cm–3 and a temperature of T = 50 K. The graphs illustrate the tuning areas (QCLTR#0F, QCLTR#1F and QCLTR#2F), which were defined based on the IV characteristics observed in Figure 6b.

The threshold currents and corresponding supply voltages were observed to be limited from below the range of supply voltages that can be used to tune the laser. As illustrated, narrowing the injector well by one monolayer increased the laser threshold current to JTH = 1.195 kA/cm2, while widening it in the same dimension lowered the threshold current to JTH = 0.815 kA/cm2. Neither of these changes resulted in any visible alterations to the nature of the relationship, as evidenced by the persistence of two distinct areas of negative differential resistance (NDR) in both cases. It should be noted that the aforementioned current values were determined under the assumption of a temperature of T = 50 K and losses in the instrument’s waveguide at a level of gsb ≅ 36 cm–1.

Narrowing the injector well by one monolayer (Qw – 0.6 nm) resulted in a slight shift of the radiation spectrum in the higher frequency range and its broadening particularly for the high bias voltages (for U = 65–76 mV/period). This can be observed, for example, in Figure 6c or the voltage U = 70 mV/period, where these changes extend in the range fmin#1F = 2.90 THz to fmax#1F = 4.7 THz, which gives the width of the radiation spectrum Δf#1F = 1.8 THz. It is maximum width of the radiation spectrum in this tuning area, converted to wavenumber this gives Δν̂max#1F = 6000 cm–1. For comparison, Δν̂0F = 4275 cm–1 (see the gray graph for the same voltage) when converted gives us Δf#0F = 1.28 THz, which corresponds to frequency changes fmin#0F = 2.83 THz to fmax#0F = 4.11 THz. In the remaining area of the QCLTR#1F (for U = 43 – 65 mV/period) we observe a band cut in the frequency range f ≈ 2 – 2.5 THz. Maximum width of the radiation spectrum in this tuning area occurs for U = 57 mV/period, where we have Δν̂max#1F = 5149 cm–1 and this corresponds to changes f′in#1F = 2.47 THz to f′max#1F = 4.01 THz. Throughout the tuning region of the QCLTR#1F, radiation from 2.47 (for U = 57 mV/period) to 5.21 THz (for U = 78 mV/period) can be obtained.

The most visible effect of widening the quantum well of the injector by one monolayer (Qw + 0.6 nm) is the narrowing of the radiation spectrum. Interestingly, in a large part of the QCLTR#2F (for U = 55 – 75 mV/period), the radiation spectrum of structure #2 does not extend beyond the boundaries of the radiation spectrum of structure #0. The shift of the spectrum toward lower frequencies (relative to Figure 6c), can only be seen for narrow intervals at the beginning and end of the tuning area (for U = 35–38 mV/period and U = 75–78 mV/period) as well as for supply voltages in the U = 45–55 mV/period range. For example, for the supply voltage U = 78 mV/period in Figure 6d) one can read the changes in the emitted radiation frequencies in the range fmin#2F = 3.12 THz to fmax#2F = 3.41 THz. That is, below the range fmin#0F = 3.6 THz to fmax#0F = 3.85 THz, characteristic of the original laser structure. However, the width of the radiation spectrum for the same voltage is Δf#2F = 0.29 THz (in conversion Δν̂#2F = 968 cm–1), which is more than twice the value of Δf#0F = 0.12 THz (in conversion Δν̂#0F = 403 cm–1). The broadest spectrum in the QTR#2F tuning range is observed at U = 76 mV/period, where the device emits radiation in the range of f#2F = 2.95 – 4.69 THz, resulting in Δfmax#2F = 1.74 THz (converted Δν̂max#2F = 5807 cm–1). This spectrum completely includes the Δν̂#0F = 2134 cm–1 which extends in the frequency range f#0F = 3.48 – 4.12 THz for original structure. The entire QTR#2F tuning region covers radiation from 1.85 THz (for U = 35 mV/period) to 4.52 THz (for U = 77 mV/period).

A general review of the results obtained with the FMSL (see Table 2) shows that narrowing the injector well by one monolayer did not significantly change the trend of changes in emitted radiation as a function of supply voltage. However, a greater width of the frequency tuning range was obtained, which is represented here by the parameter TRfw (tuning range frequency width) defined as the difference of the highest and lowest frequency in the tuning region (TRfw = 5.21 – 2.47 = 2.74 THz).

Table 2. Selected Tuning Parameters of Tested QCL Structures Calculated Using FMSL

structure#1structure#0structure#2
tuning areaQCLTR#1FQCLTR#0QCLTR#2F
frequecy range (THz)2.47–5.211.9–4.41.85–4.52
TRfw (THz)2.742.52.67
wavelengths (μm)121to57.5158to6.8.1162to66.3
tunning voltage range (mV/period)42–7838–7835–78
TRSI, (mV/period)364043
Δν̂max (cm–1)600068005807

For comparison, in structure #0, we get TRfw = 4.4 – 1.9 = 2.5 THz. On the other hand, the range of the supply voltage controlling the QCLTR#1F area has been slightly reduced, as indicated by the TRSI (tuning range supply interval) parameter defined as the difference of the largest and smallest value of the supply voltage in the tuning area under consideration. For structure #1 it is TRSI = 78 – 42 = 36 mV/period for comparison in structure #0 we have TRSI = 78 – 38 = 40 mV/period. Similarly, the largest observed spectral width of radiation (Δν̂max) in the QCLTR#1F area exhibits a 12% decrease in value when compared to the parameter’s value in the QCLTR#0 area.

The situation is somewhat different for the injector well broadening. It turns out that in this case we have a slightly smaller broadening of the tuning region in the frequency range (TRSI = 2.67 THz), but with an increased control voltage interval (TRSI = 43 mV/period) and a slightly smaller maximum width of the radiation spectrum (Δν̂max = 5807 cm–1). The results also showed that there were three distinct changes in this trend in the QCLTR#2F region (around U = 44, 60, 70 mV/period), so calculations using RSM were necessary to verify it thoroughly. The simulation results in this range are shown in Figure 7c,d, where there are maps of optical gain as a function of supply voltage and frequency of emitted photons calculated for the temperature T = 200 K. In addition, selected parameters of the tuning regions of the tested QCL structures calculated with RSM are listed in Table 3.

Table 3. Selected Tuning Parameters of Tested QCL Structures Calculated Using RSM

structure#1structure#0structure#2
tuning areaQCLTR#1RQCLTR#0QCLTR#2aQCLTR#2bQCLTR#2c
frequecy range (THz)2.47–4.411.9–4.42.31–3.743.42–4.242.0–2.4
TRfw (THz)1.942.51.430.820.4
wavelengths (μm)121to68.0158to68.1130to80.287.7to70.7150to125
tunning voltage range (mV/period)51–7238–7843–6062–7258–64
TRSI (mV/period)214017106
Δν̂max (cm–1)53376800417018351334

For the narrowed injector well, the trend line of changes in mG (see Figure 7c) does not differ significantly from what was observed for the original structure (see Figure 5c). Assuming gsb ≅ 40 cm–1, the tuning region QCLTR#1R can be determined, which covers the frequency range fTR#1R = 2.47 – 4.41 THz (for U = 51–72 mV/period), which corresponds to waves with lengths λTR#1R = 68–121 μm. The maximum width of the radiation spectrum occurs here for U = 60 mV/period and is Δν̂max#1R = 5337 cm–1, which, in relation to the previously given value calculated using FMSL, informs us about a slight shift of the maximum spectrum width toward a higher supply voltage. At the same time, by comparing the QCLTR#0R and QCLTR#1R areas, it can be concluded that, as predicted by FMSL, there was a shift in the laser tuning range toward higher frequencies. Finally, after careful calculations taking into account all major electron scatters, it is found that narrowing the injector area reduces the frequency tuning range (TRfw = 1.94 THz) and the control voltage interval (TRSI = 21 mV/period) with respect to the original #0 structure.

For the widened injector well, the mG trend line shows significant changes compared to that observed for the original structure, as can be seen in Figure 7d. This is because the optical gain map shows three separate regions for which g > 40 cm–1 (QCLTR#2a, QCLTR#2b and QCLTR#2c). They refer to changes in the trend of the dependence of the emission frequency on the supply voltage QCL, predicted using FMSL (see Figure 7a) and they occur for similar voltages as in the mentioned case, so we have peaks of optical gain for U = 44, 60, 77 mV/period. This makes the laser tuning for this case more complex and covers three ranges with different spectra and light intensity.

The largest of the above ranges is QCLTR#2a, within which the laser can be tuned in the range fTR#2a = 2.3 – 3.74 THz (for U = 43–60 mV/period), corresponding to wavelengths λTR#2a = 80.2–130.3 μm. This gives TRfw = 1.43 THz for the control voltage range TRSI = 17 mV/period. The maximum width of the radiation spectrum is Δν̂max#2a = 4170–1 for U = 48 mV/period (FMSL predicts Δν̂#2F = 4274 cm–1 for U = 44 mV/period). The tuning range of QCLTR#2b covers the frequencies fTR#2b = 3.42 – 4.24 THz (for U = 62 – 72 mV/period), which corresponds to the wavelengths λTR#2b = 70.7 – 87.7 μm. This results in a TRfw = 0.82 THz for the control voltage range of TRSI = 10 mV/period. The maximum width of the radiation spectrum occurs here for U = 66 mV/period and is Δν̂max#2b = 1835 cm–1. The trend change predicted by FMSL occurs for U = 78 mV/period at Δν̂#2F = 1613 cm–1, but is as high as Δν̂#2F’’ = 5081 cm–1 for U = 76 mV/period. Such a significant difference in spectral width observed in the FMSL results can be attributed to the presence of an additional tuning range (QCLTR#2c). Assuming gsb ≅ 40 cm–1, the frequency of fTR#2c changes from 2 to 2.4 THz (TRfw = 0.4 THz), corresponding to wavelengths λTR#2c = 124.9–149,9 μm at a supply voltage U = 58–64 mV/period (TRSI is only 6 mV/period). As can be seen, the range of supply voltages associated with tuning in the QCLTR#2c area partially covers the ranges of control voltages of the QCLTR#2a and QCLTR#2b areas, which limits the possibility of seeing this phenomenon only in optical gain energy maps. The maximum spectral width of the radiation of the QCLT#2c area occurs at U = 61 mV/period and is equal to Δν̂max#2c = 1334 cm–1 (predicted by FMSL for U = 60 mV/period, where Δν̂#2F’’’ = 1631 cm–1).

5. Experiment Hint

Click to copy section linkSection link copied!

In order to verify the simulation of the tested QCL device against its measurements, a comparison was made between the measured and simulated current–voltage characteristics. The results are shown in Figure 8, where the black solid line shows the characteristics obtained by the experiment described in the paper, (59) and the curve of the corresponding I/U relationship (green line + symbol plot) calculated using the RSM model. As can be seen, the calculated current–voltage characteristic differs from the experimental one. This fact is well-known and described in the literature 66 and 67 and results from the existence of parasitic effects such as series resistance or voltage drop at the metal–semiconductor interface.

Figure 8

High Resolution Image

Download MS PowerPoint Slide

After taking these effects into account (red line + symbol graph), a good agreement between simulations and measurements was obtained. However, it should be emphasized that such a situation is only possible in the subthreshold range, since the RSM model based on NEGF does not take into account light-matter interactions.

Simulation results for the original structure #0 also showed that the maximum gain for the temperature T = 200 K corresponds to a frequency in the range of 3.4–3.6 THz and agrees with the measured frequency of ∼3.6 THz, and the value of this gain is close to the standard mirror and waveguide losses in this type of laser. This means that the temperature T = 200 K is the emission limit temperature of this laser as measured by (59) (Tmax = 199.5 K).

6. Conclusions

Click to copy section linkSection link copied!

The conducted tests confirmed the apparent sensitivity of the QCL radiation designed by Kumar et al. (60) to changing the width of the injector well. Depending on the thickness of the injector layer, it is possible to obtain either homogeneous (QCLTR#0, QCLTR#1) or nonhomogeneous (QCLTR#2) tuning regions with a linear or nonlinear trend of mG changes. Despite apparent changes in the width of the radiation spectrum due to changes in the width of the injector well, no large changes were observed in the frequency range of the emitted wavelengths. It is a well-known feature of terahertz lasers that it is difficult to achieve a significant shift in the radiation spectrum without significant changes in the superlattice structure. However, two examples of the device (structures #1 and #2) were proposed, which with a small change in the width of the injector layer (by one monolayer) can have different tuning characteristics. A homogeneous one with a wide radiation spectrum can be used in systems for detecting and imaging a wide variety of objects, while a nonhomogeneous one with several smaller radiation spectral widths can be used for more precise imaging of selected targets, but this case requires a complex laser power control system. The work performed showed the important role of the developed numerical models (FMSL and RSM), which allowed effective and accurate simulations of demanding QCL structures. In summary, the main achievement of our work is the proposal of a precise modeling of the influence of the injector layer on the emission wavelength control, which has not been analyzed in such detail in the literature.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • Mariusz Mączka - Department of Electronics Fundamentals,Faculty of Electrical and Computer Engineering Rzeszow University of Technology, 35-959 Rzeszow, Poland; Voltage-Controlled and Injector Layer Thickness-Dependent Tuning of Quantum Cascade Laser for Terahertz Spectroscopy (1)https://orcid.org/0000-0003-4137-4829; Email: [emailprotected]

  • Author
    • Grzegorz Hałdaś - Department of Electronics Fundamentals,Faculty of Electrical and Computer Engineering Rzeszow University of Technology, 35-959 Rzeszow, Poland

  • Author Contributions

    Conceptualization: M.M. and G.H. Methodology: M.M. and G.H. Software: M.M. and G.H. Validation: M.M. and G.H. Formal analysis: M.M. and G.H. Investigation: M.M. and G.H. Resources: M.M. and G.H. Writing─original draft preparation, M.M. Writing─review and editing, M.M.

  • Funding

    Projekt: PB22.EP.24.001─Automatyka, elektronika i elektrotechnika. Ministerstwo Nauki i Szkolnictwa Wyższego

  • Notes

    The authors declare no competing financial interest.

Abbreviations

Click to copy section linkSection link copied!

QCL

quantum cascade laser

FMSL

finite model of superlattice

MBM

modified bisection method

NDR

negative dynamic resistance

RSM

real space model

THz-TDS

terahertz time-domain spectroscopy

WFM

Wannier function method

References

Click to copy section linkSection link copied!

This article references 65 other publications.

  1. 1

    Sun, J.; Ding, J.; Liu, N.; Yang, G.; Li, J. Detection of multiple chemicals based on external cavity quantum cascade laser spectroscopy, Spectrochim Acta. Part A Mol. Biomol. Spectrosc. 2018, 191, 532538, DOI: 10.1016/j.saa.2017.10.059

    Google Scholar

    There is no corresponding record for this reference.

  2. 2

    Park, S.; Son, J.; Yu, J.; Lee, J. Standoff Detection and Identification of Liquid Chemicals on a Reflective Substrate Using a Wavelength-Tunable Quantum Cascade Laser. Sensors 2022, 22, 3172, DOI: 10.3390/s22093172

    Google Scholar

    There is no corresponding record for this reference.

  3. 3

    Abina, A.; Puc, A.; Jeglic, U.; Zidansek, A. Applications of Terahertz Spectroscopy in the Field of Construction and Building Materials. Appl. Spectrosc. Rev. 2015, 50, 279303, DOI: 10.1080/05704928.2014.965825

    Google Scholar

    There is no corresponding record for this reference.

  4. 4

    Chrzanowski, L.; Beckmann, J.; Marchetti, B.; Ewert, U.; Schade, U. Terahertz Time Domain Spectroscopy for Non-Destructive Testing of Hazardous Liquids. Materials Testing 2012, 54, 444450, DOI: 10.3139/120.110351

    Google Scholar

    There is no corresponding record for this reference.

  5. 5

    Krügener, K.; Ornik, J.; Schneider, L. M.; Jäckel, A.; Koch-Dandolo, C. L.; Castro-Camus, E.; Riedl-Siedow, N.; Koch, M.; Viöl, W. Terahertz Inspection of Buildings and Architectural Art. Appl. Sci. 2020, 10, 5166, DOI: 10.3390/app10155166

    Google Scholar

    There is no corresponding record for this reference.

  6. 6

    Zang, Z.; Li, Z.; Wang, J.; Lu, X.; Lyu, Q.; Tang, M.; Cui, H.-L.; Yan, S. Terahertz spectroscopic monitoring and analysis of citrus leaf water status under low temperature stress. Plant Physiology and Biochemistry 2023, 194, 5259, DOI: 10.1016/j.plaphy.2022.10.032

    Google Scholar

    There is no corresponding record for this reference.

  7. 7

    Kwaśny, M.; Miczuga, M. Ł. Gas detection systems utilizing cascade lasers. Elektronika-konstrukcje, technologie, zastosowania 2009, 50, 8690

    Google Scholar

    There is no corresponding record for this reference.

  8. 8

    Wysocki, G.; Lewicki, R.; Curl, R. Widely tunable mode-hop free external cavity quantum cascade lasers for high resolution spectroscopy and chemical sensing. Appl. Phys. B: Laser Opt. 2008, 92, 305311, DOI: 10.1007/s00340-008-3047-x

    Google Scholar

    There is no corresponding record for this reference.

  9. 9

    Zhang, X.; Wu, X.; Xiao, B.; Qin, J. Terahertz determination of imidacloprid in soil based on a metasurface sensor. Opt. Express 2023, 31, 3777837788, DOI: 10.1364/OE.503624

    Google Scholar

    There is no corresponding record for this reference.

  10. 10

    Zhan, H.; Zhao, K.; Bao, R.; Xiao, L. Monitoring PM2.5 in the Atmosphere by Using Terahertz Time-Domain Spectroscopy. Journal of Infrared, Millimeter, and Terahertz Waves 2016, 37, 929938, DOI: 10.1007/s10762-016-0283-8

    Google Scholar

    There is no corresponding record for this reference.

  11. 11

    Huang, S.; Deng, H.; Wei, X.; Zhang, J. Progress in application of terahertz time-domain spectroscopy for pharmaceutical analyses. Frontiers in Bioengineering and Biotechnology 2023, 11, 1219042, DOI: 10.3389/fbioe.2023.1219042

    Google Scholar

    There is no corresponding record for this reference.

  12. 12

    Sulovská, K.; Křesálek, V. Examining pharmaceuticals using terahertz spectroscopy. Proceedings Volume 9651, Millimetre Wave and Terahertz Sensors and Technology VIII 2015, 965103, DOI: 10.1117/12.2194585

    Google Scholar

    There is no corresponding record for this reference.

  13. 13

    MacPherson, E.; Gallerano, G. P.; Park, G.-S.; Hintzsche, H.; Wilmink, G. J. Terahertz Imaging and Spectroscopy for Biology and Biomedicine. IEEE J. Biomed. Health Inf. 2013, 17, 765767, (Guest Editorial) DOI: 10.1109/JBHI.2013.2257333

    Google Scholar

    There is no corresponding record for this reference.

  14. 14

    Yang, S.; Ding, L.; Wang, S.; Du, C.; Feng, L.; Qiu, H.; Zhang, C.; Wu, J.; Fan, K.; Jin, B.; Chen, J.; Wu, Ph. Studying Oral Tissue via Real-Time High-Resolution Terahertz Spectroscopic Imaging. Physical Review Applied 2023, 19, 19, DOI: 10.1103/PhysRevApplied.19.034033

    Google Scholar

    There is no corresponding record for this reference.

  15. 15

    Shi, J.; Guo, Z.; Chen, H.; Xiao, Z.; Bai, H.; Li, X.; Niu, P.; Yao, J. Artificial Intelligence-Assisted Terahertz Imaging for Rapid and Label-Free Identification of Efficient Light Formula in Laser Therapy. Biosensors 2022, 12, 826, DOI: 10.3390/bios12100826

    Google Scholar

    There is no corresponding record for this reference.

  16. 16

    Ma, X.; Wang, K.; Fauquet, F.; Roux, M.; Dandolo, C. L. K.; Mounaix, P.; Guillet, J.-P. Terahertz frequency modulated continuous wave imaging for non-destructive evaluation of painting and multilayer parts. Proceedings, Volume 10531, Terahertz, RF, Millimeter, and Submillimeter-Wave Technology and Applications XI , March 20, 2018; 1053122. DOI: 10.1117/12.2291016 .

    Google Scholar

    There is no corresponding record for this reference.

  17. 17

    Krügener, K.; Ornik, J.; Schneider, L. M.; Jäckel, A.; Koch-Dandolo, C. L.; Castro-Camus, E.; Riedl-Siedow, N.; Koch, M.; Viöl, W. Terahertz Inspection of Buildings and Architectural Art. Appl. Sci. 2020, 10, 5166, DOI: 10.3390/app10155166

    Google Scholar

    There is no corresponding record for this reference.

  18. 18

    Fukunaga, K.; Ogawa, Y.; Hayashi, S.; Hosako, I. Terahertz spectroscopy for art conservation Ieice. Electronic Express 2007, 4, 258263, DOI: 10.1587/elex.4.258

    Google Scholar

    There is no corresponding record for this reference.

  19. 19

    Praveena, A.; Ponnapalli, V. A. S.; Umamaheswari, G. Terahertz Antenna Technology for Detection of Explosives and Weapons: A Concise Review. In Malik, P. K., Lu, J., Madhav, B. T. P., Kalkhambkar, G., Amit, S., Eds.; Terahertz Antenna Technology for Detection of Explosives and Weapons: A Concise Review. Smart Antennas; EAI/Springer Innovations in Communication and Computing; Springer, 2022. DOI: 10.1007/978-3-030-76636-8_25 .

    Google Scholar

    There is no corresponding record for this reference.

  20. 20

    Palka, N.; Szala, M.; Czerwinska, E. Characterization of prospective explosive materials using terahertz time-domain spectroscopy. Appl. Opt. 2016, 55, 45754583, DOI: 10.1364/AO.55.004575

    Google Scholar

    There is no corresponding record for this reference.

  21. 21

    Shi, W.; Dong, C.; Hou, L.; Xing, Z.; Sun, Q.; Zhang, L. Investigation of aging characteristics in explosive using terahertz time-domain spectroscopy. International Journal of Modern Physics B 2019, 33, 1950272, DOI: 10.1142/S0217979219502722

    Google Scholar

    There is no corresponding record for this reference.

  22. 22

    Mieloszyk, M.; Majewska, K.; Ostachowicz, W. THz spectroscopy for inspection and evaluation of internal structure of sandwich samples. International Journal of Structural Integrity 2018, 9 (6), 793803, DOI: 10.1108/IJSI-11-2017-0068

    Google Scholar

    There is no corresponding record for this reference.

  23. 23

    Komies, S.; Sun, J.; Hodges, C.; Lynch, S. A.; Watson, J. S.; Lucyszyn, S. Low-Cost Thermal-Infrared ‘THz-Torch. Spectroscopy, in IEEE Access 2024, 12, 2186821885, DOI: 10.1109/ACCESS.2024.3359054

    Google Scholar

    There is no corresponding record for this reference.

  24. 24

    Yang, R.; Dong, X.; Chen, G.; Lin, F.; Huang, Z.; Manzo, M.; Mao, H. Novel Terahertz Spectroscopy Technology for Crystallinity and Crystal Structure Analysis of Cellulose. Polymers 2021, 13 (6), 6, DOI: 10.3390/polym13010006

    Google Scholar

    There is no corresponding record for this reference.

  25. 25

    Seliya, P.; Bonn, M.; Grechko, M. On selection rules in two-dimensional terahertz-infrared-visible spectroscopy. J. Chem. Phys. 2024, 160 (3), 034201, DOI: 10.1063/5.0179041

    Google Scholar

    There is no corresponding record for this reference.

  26. 26

    Bowman, T.; Wu, Y.; Gauch, J. Terahertz Imaging of Three-Dimensional Dehydrated Breast Cancer Tumors. J. Infrared Milli Terahz Waves 2017, 38, 766786, DOI: 10.1007/s10762-017-0377-y

    Google Scholar

    There is no corresponding record for this reference.

  27. 27

    Fischer, B; Hoffmann, M; Helm, H; Modjesch, G; Jepsen, P U. Chemical recognition in terahertz time-domainspectroscopy and imaging. Semicond. Sci. Technol. 2005, 20, S246, DOI: 10.1088/0268-1242/20/7/015

    Google Scholar

    There is no corresponding record for this reference.

  28. 28

    Lu, M.; Shen, J.; Li, N.; Zhang, Y.; Zhang, C.; Liang, L.; Xu, X. Detection and identification of illicit drugs usingterahertz imaging. J. Appl. Phys. 2006, 100, 103104, DOI: 10.1063/1.2388041

    Google Scholar

    There is no corresponding record for this reference.

  29. 29

    Watanabe, Y.; Kawase, K.; Ikari, T.; Ito, H.; Ishikawa, Y.; Minamide, H. Component spatial pattern analysis of chemicals using terahertz spectroscopic imaging. Appl. Phys. Lett. 2003, 83 (4), 800802, 28 DOI: 10.1063/1.1595132

    Google Scholar

    There is no corresponding record for this reference.

  30. 30

    Kawase, K.; Ogawa, Y.; Watanabe, Y.; Inoue, H. Non-destructive terahertz imaging of illicit drugs usingspectral fingerprints. Opt. Express 2003, 11, 254954, DOI: 10.1364/OE.11.002549

    Google Scholar

    There is no corresponding record for this reference.

  31. 31

    Davies, A. G.; Burnett, A. D.; Fan, W.; Linfield, E. H.; Cunningham, J. E. Terahertz spectroscopy of explosivesand drugs. Mater. Today 2008, 11, 1826, DOI: 10.1016/S1369-7021(08)70016-6

    Google Scholar

    There is no corresponding record for this reference.

  32. 32

    Chan, W. L.; Deibel, J.; Mittleman, D. M Imaging with terahertz radiation. Rep. Prog. Phys. 2007, 70, 1325, DOI: 10.1088/0034-4885/70/8/R02

    Google Scholar

    There is no corresponding record for this reference.

  33. 33

    Shimosato, H.; Ashida, M.; Itoh, T.; Saito, S.; Sakai, K. Ultrabroadband Detection of Terahertz Radiation from 0.1 to 100 THz with Photoconductive Antenna in Ultrafast. Optics V; Watanabe, S., Midorikawa, K., Eds.; Springer: New York, 2007. DOI: 10.1007/978-0-387-49119-6_41 .

    Google Scholar

    There is no corresponding record for this reference.

  34. 34

    Dean, P.; Saat, N. K.; Khanna, S. P.; Salih, M.; Burnett, A.; Cunningham, J.; Linfield, E. H.; Davies, A. G. Dual-frequency imaging using an electrically tunable terahertz quantum cascade laser. Opt. Express 2009, 17, 2063120641, DOI: 10.1364/OE.17.020631

    Google Scholar

    There is no corresponding record for this reference.

  35. 35

    Singh, K.; Aalam, U.; Mishra, A.; Dixit, N.; Bandyopadhyay, A.; Sengupta, A. Spectroscopic and imaging considerations of THz-TDS and ULF-Raman techniques towards practical security applications. Opt. Express 2024, 32, 13141324, DOI: 10.1364/OE.507941

    Google Scholar

    There is no corresponding record for this reference.

  36. 36

    Faist, J.; Capasso, F.; Sirtori, C.; Sivco, D. L.; Baillargeon, J. N.; Hutchinson, A. L.; Chu, S. N. G.; Cho, A. Y. High power mid-infrared (λ∼5 μm) quantum cascade lasers operating above room temperature. Appl. Phys. Lett. 1996, 68, 36803682, DOI: 10.1063/1.115741

    Google Scholar

    There is no corresponding record for this reference.

  37. 37

    Callebaut, H.; Kumar, S.; Williams, B. S.; Hu, Q.; Reno, J. L. Analysis of transport properties of terahertz quantum cascade lasers. Appl. Phys. Lett. 2003, 83 (2), 207209, DOI: 10.1063/1.1590749

    Google Scholar

    There is no corresponding record for this reference.

  38. 38

    Kosiel, K., AlGaAs/GaAs quantum cascade lasers for gas detection systems. International Conference on Infrared, Millimeter, and Terahertz Waves; IEEE, 2011; pp 12. DOI: 10.1109/irmmw-THz.2011.6105118 .

    Google Scholar

    There is no corresponding record for this reference.

  39. 39

    Lever, L.; Hinchcliffe, N. M.; Khanna, S. P.; Dean, P.; Ikonic, Z.; Evans, C. A.; Davies, A. G.; Harrison, P.; Linfield, E. H.; Kelsall, R. W. Terahertz ambipolar dual-wavelength quantum cascade laser. Opt. Express 2009, 17, 1992632, DOI: 10.1364/OE.17.019926

    Google Scholar

    There is no corresponding record for this reference.

  40. 40

    Freeman, J. R.; Marshall, O. P.; Beere, H. E.; Ritchie, D. A. Electrically switchable emission in terahertz quantum cascade lasers. Opt. Express 2008, 16, 198305, DOI: 10.1364/OE.16.019830

    Google Scholar

    There is no corresponding record for this reference.

  41. 41

    Qin, Q.; Williams, B. S.; Kumar, S.; Reno, J. L.; Hu, Q. Tuning a terahertz wire laser. Nature Photon 2009, 3, 7327, DOI: 10.1038/nphoton.2009.218

    Google Scholar

    There is no corresponding record for this reference.

  42. 42

    LanderGower, N.; Levy, S.; Piperno, S.; Addamane, S. J.; Reno, J. L.; Albo, A. Doping engineering: Next step toward room temperature performance of terahertz quantum cascade lasers. Journal of Vacuum Science & Technology. B 2024, 42, 10801, DOI: 10.1116/6.0003160

    Google Scholar

    There is no corresponding record for this reference.

  43. 43

    Ushakov, D.; Afonenko, A.; Khabibullin, R.; Fadeev, M.; Dubinov, A. Phosphides-Based Terahertz Quantum-Cascade Laser. Physica Status Solidi (RRL) - Rapid Research Letters 2024, DOI: 10.1002/pssr.202300392

    Google Scholar

    There is no corresponding record for this reference.

  44. 44

    Vitiello, M. S.; Tredicucci, A. Tunable emission in THz quantum cascade lasers. IEEE Trans. THz Sci. Technol. 2011, 1, 7684, DOI: 10.1109/TTHZ.2011.2159543

    Google Scholar

    There is no corresponding record for this reference.

  45. 45

    Wysocki, G.; Curl, R.F.; Tittel, F.K.; Maulini, R.; Bulliard, J.M.; Faist, J. Widely tunable mode-hop free external cavity quantum cascade laser for high resolution spectroscopic applications. Appl. Phys. B 2005, 81, 769777, DOI: 10.1007/s00340-005-1965-4

    Google Scholar

    There is no corresponding record for this reference.

  46. 46

    Mohan, A.; Wittmann, A.; Hugi, A.; Blaser, S.; Giovannini, M.; Faist, J. Room-temperature continuous-wave operation of an external-cavity quantum cascade laser. Opt. Lett. 2007, 32, 27922794, DOI: 10.1364/OL.32.002792

    Google Scholar

    There is no corresponding record for this reference.

  47. 47

    Pierściński, K.; Pierścińska, D.; Pluska, M.; Gutowski, P.; Sankowska, I.; Karbownik, P.; Czerwinski, A.; Bugajski, M. Room temperature, single mode emission from two-section coupled cavity InGaAs/AlGaAs/GaAs quantum cascade laser. J. Appl. Phys. 2015, 118 (13), 133103, DOI: 10.1063/1.4932141

    Google Scholar

    There is no corresponding record for this reference.

  48. 48

    Janczak, M.; Sarzała, R. P.; Dems, M.; Kolek, A.; Bugajski, M.; Nakwaski, W.; Czyszanowski, T. Threshold performance of pulse-operating quantum-cascade vertical-cavity surface-emitting lasers. Opt. Express 2022, 30, 4505445069, DOI: 10.1364/OE.474086

    Google Scholar

    There is no corresponding record for this reference.

  49. 49

    Motyka, M.; Ryczko, K.; Dyksik, M.; Sęk, G.; Misiewicz, J.; Weih, R.; Dallner, M.; Höfling, S.; Kamp, M. On the modified active region design of interband cascade lasers. J. Appl. Phys. 2015, 117 (8), 084312, DOI: 10.1063/1.4913391

    Google Scholar

    There is no corresponding record for this reference.

  50. 50

    Birner, S.; Zibold, T.; Andlauer, T.; Kubis, T.; Sabathil, M.; Trellakis, A.; Vogl, P. Nextnano: General Purpose 3-D Simulations. IEEE Trans. Electron Devices 2007, 54 (9), 21372142, DOI: 10.1109/TED.2007.902871

    Google Scholar

    There is no corresponding record for this reference.

  51. 51

    Kubis, T.; Yeh, C.; Vogl, P.; Benz, A.; Fasching, G.; Deutsch, C. Theory of nonequilibrium quantum transport and energy dissipation in terahertz quantum cascade lasers. Phys. Rev. B 2009, 79 (19), 195323, DOI: 10.1103/PhysRevB.79.195323

    Google Scholar

    There is no corresponding record for this reference.

  52. 52

    Shashurin, V. D.; Vetrova, N. A.; Filyaev, A. A. Evaluation of channel transmission of nanoelectronic devices on low-dimensional structures with quantum confinement. J. Phys.: Conf. Ser. 2020, 1560, 012048, DOI: 10.1088/1742-6596/1560/1/012048

    Google Scholar

    There is no corresponding record for this reference.

  53. 53

    Pereira, M.; Lee, S. C.; Wacker, A. Effect of Coulomb corrections and mean field on gain and absorption in Quantum Casscade Lasers. Phys. Status Solidi C: current topics in Solid State Physics 2005, 2 (8), 30273030, DOI: 10.1002/pssc.200460712

    Google Scholar

    There is no corresponding record for this reference.

  54. 54

    Mączka, M.; Hałdaś, G. Calculations of transport parameters in semiconductor superlattices based on the Green’s functions method in different Hamiltonian representations. Bull. Polym. Acad. Sci. Technol. Sci. 2019, 67, 631641, DOI: 10.24425/bpasts.2019.129661

    Google Scholar

    There is no corresponding record for this reference.

  55. 55

    Mączka, M.; Pawłowski, S. Efficient method for transport simulations in quantum cascade lasers. In Proceedings of the International Conference on Semiconductor Nanostructures for Optoelectronics and Biosensors (IC SeNOB), Rzeszow, Poland, May 22–25, 2016; EPJ Web of Conferences, 2016; p 133.

    Google Scholar

    There is no corresponding record for this reference.

  56. 56

    Mączka, M. Effective Simulations of Electronic Transport in 2D Structures Based on Semiconductor Superlattice Infinite Model. Electronics 2020, 9, 1845, DOI: 10.3390/electronics9111845

    Google Scholar

    There is no corresponding record for this reference.

  57. 57

    Mączka, M.; Pawłowski, S. A Polynomial Approximation to Self Consistent Solution for Schrödinger - Poisson Equations in Superlattice Structures. Energies 2022, 15, 760, DOI: 10.3390/en15030760

    Google Scholar

    There is no corresponding record for this reference.

  58. 58

    Hałdaś, G. Implementation of non-uniform mesh in non-equilibrium Green’s function simulations of quantum cascade lasers. J. Comput. Electron. 2019, 18, 14001406, DOI: 10.1007/s10825-019-01386-4

    Google Scholar

    There is no corresponding record for this reference.

  59. 59

    Fathololoumi, S.; Dupont, E.; Chan, C. W. I.; Wasilewski, Z. R.; Laframboise, S. R.; Ban, D.; Mátyás, A.; Jirauschek, C.; Hu, Q.; Liu, H. C. Terahertz quantum cascade lasers operating up to ∼ 200 K with optimized oscillator strength and improved injection tunneling. Opt. Express 2012, 20, 38663876, DOI: 10.1364/OE.20.003866

    Google Scholar

    There is no corresponding record for this reference.

  60. 60

    Kumar, S.; Hu, Q.; Reno, J. L. 186 K operation of terahertz quantum-cascade lasers based on a diagonal design. Appl. Phys. Lett. 2009, 94, 131105, DOI: 10.1063/1.3114418

    Google Scholar

    There is no corresponding record for this reference.

  61. 61

    Pawłowski, S.; Mączka, M. Optimisation of QCL Structures Modelling by Polynomial Approximation”. Materials 2022, 15, 5715, DOI: 10.3390/ma15165715

    Google Scholar

    There is no corresponding record for this reference.

  62. 62

    Büttiker, M. Four-Terminal Phase-Coherent Conductance. Phys. Rev. Lett. 1986, 57, 17611764, DOI: 10.1103/PhysRevLett.57.1761

    Google Scholar

    There is no corresponding record for this reference.

  63. 63

    Lake, R.; Klimeck, G.; Bowen, R. C.; Jovanovic, D. Single and multiband modeling of quantum electron transport through layered semiconductor devices”. J. Appl. Phys. 1997, 81, 78457869, DOI: 10.1063/1.365394

    Google Scholar

    There is no corresponding record for this reference.

  64. 64

    Hałdaś, G.; Kolek, A.; Tralle, I. Modeling of Mid-Infrared Quantum Cascade Laser by Means of Nonequilibrium Green’s Functions. IEEE J. Quantum Electron. 2011, 47, 878885, DOI: 10.1109/JQE.2011.2130512

    Google Scholar

    There is no corresponding record for this reference.

  65. 65

    Hałdaś, G.; Kolek, A. Quantum-mechanical modeling of nanoelectronic devices. 40th International Spring Seminar on Electronics Technology (ISSE), Sofia, Bulgaria; IEEE, 2017; pp 16. DOI: 10.1109/ISSE.2017.8000968 .

    Google Scholar

    There is no corresponding record for this reference.

Cited By

Click to copy section linkSection link copied!

This article has not yet been cited by other publications.

Download PDF

Voltage-Controlled and Injector Layer Thickness-Dependent Tuning of Quantum Cascade Laser for Terahertz Spectroscopy (2025)
Top Articles
Latest Posts
Recommended Articles
Article information

Author: Carmelo Roob

Last Updated:

Views: 5544

Rating: 4.4 / 5 (45 voted)

Reviews: 84% of readers found this page helpful

Author information

Name: Carmelo Roob

Birthday: 1995-01-09

Address: Apt. 915 481 Sipes Cliff, New Gonzalobury, CO 80176

Phone: +6773780339780

Job: Sales Executive

Hobby: Gaming, Jogging, Rugby, Video gaming, Handball, Ice skating, Web surfing

Introduction: My name is Carmelo Roob, I am a modern, handsome, delightful, comfortable, attractive, vast, good person who loves writing and wants to share my knowledge and understanding with you.