Abstract
The temperature dependences of the threshold current and slope efficiency, as represented by their respective characteristic temperature coefficients T0 and T1, are discussed for quantum cascade lasers (QCLs) emitting in the 3.0–3.8 μm, 3.9–5.0 μm, 8–10 μm, and 12–16 μm wavelength ranges. Carrier-leakage mechanisms are treated with emphasis on shunt-type leakage within active regions (ARs); the dominant leakage path in state-of-the-art devices. Carrier-leakage suppression, best evidenced by the T1 value, is shown to have been the key to effectively doubling the room-temperature pulsed and continuous-wave (CW) wallplug efficiencies for 4.5–5.0 μm emitting QCLs. By employing deep-well and/or tapered-active (TA)-type AR designs, for carrier-leakage suppression, T0 values as high as 278 K at λ = 4.8 μm and 242 K at λ = 8.4 μm have been achieved for devices of moderately high injector doping, as required for watt-range room-temperature CW operation. Similarly, TA-type QCLs have led to record-high T1 values: 797 K at λ = 4.8 μm, and 561 K at λ = 8.8 μm, for low-threshold (~1.6 kA cm−2) devices at room temperature. Step-taper TA (STA) AR designs for 8.4 and 8.8 μm emitting QCLs have resulted in both carrier-leakage suppression as well as fast and efficient carrier extraction. That, in turn, led to internal-differential-efficiency values in the 85–90% range; that is, 30–40% higher than for any previously reported 7–10 μm emitting QCLs. We further show that 4.6 μm emitting STA-type devices hold the potential for room-temperature CW wallplug efficiency values in excess of 27%. Should the internal differential efficiency reach theoretical limits (87–89%) at λ = 4.6 μm, single-facet, room-temperature CW wallplug efficiency values in excess of 40% become possible.
Original content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.
1. Introduction
The study of the temperature dependence of the electro-optical characteristics of semiconductor lasers has a long history, since it has been a crucial factor towards devices achieving room-temperature (RT) operation, continuous-wave (CW) high-power operation, and CW operation with high electrical-to-optical power-conversion efficiency, so-called wallplug efficiency. Thus, it is an important subject, as it holds the keys to long-term reliable operation at high CW output powers. Many of the lessons learned from the optimization of highly efficient, high-CW-power near-infrared (IR) interband-transition semiconductor lasers are applicable right now to advancing the quasi-CW or CW performance of intersubband-transition lasers emitting the mid-IR wavelength range.
The key parameters are the characteristic temperature coefficient of the pulsed threshold-current density Jth, T0 [1], and the characteristic temperature coefficient of the pulsed slope efficiency ηs, T1 [2] which are defined as follows:
where Tref + ΔT is the heatsink temperature, Tref is the reference heatsink temperature, and ΔT is the range in heatsink temperature over which the temperature dependence of Jth or ηs can be approximated by an exponential function.
With the advent of the implementing the intersubband-transition laser concept [3] via the achievement in 1994 of the so-called quantum cascade laser (QCL) [4], the T0 value, for Jth values above liquid-nitrogen temperature, was found rightaway [5] to clearly separate QCLs from interband-transition mid-IR lasers. That is, T0 was found to be ~110 K, above 80 K heatsink temperature, by comparison to only ~30–50 K for interband-transition GaSb-based or lead-salt lasers, an indication that Auger effects were, as predicted, negligible in QCLs. Thus, QCLs held the promise to reach lasing at room temperature. Next, it was found that T0 is a strong function of the energy difference Δinj between the lower laser level and the ground energy state, g, of the injector miniband in the next period of a QCL structure [6]. That is, for low Δinj values (50–70 meV) thermally excited electrons from level g readily fill the lower laser level, an effect called backfilling; in turn, that lowers the population inversion and causes a fast increase in Jth with temperature. The remedy was to design structures with Δinj ⩾ 100 meV, which led to breakthroughs in performance: CW operation, at liquid-nitrogen temperature, in 1995 [7] and RT pulsed operation in 1996 [8]. After the achievement of RT CW operation in 2002 [9] it slowly and steadily became apparent that what affects the T0 value, at and above RT, is primarily carrier leakage from or within the active regions of the QCL structure. Since about 2009 carrier-leakage suppression, via conduction-band engineering of the active regions [10–13], has resulted in large increases in the T0 value above RT (e.g. from ~140 to ~250 K for moderately doped, 4.5–4.9 μm emitting devices).
However, the T1 parameter has been by and large neglected, since the ηs value is not affected by backfilling, and thus it hardly varies with temperature up to ~300 K. In fact, the nature of the T1 parameter is not at all covered even in the most recent books on QCLs. Yet the T1 value, just like in interband-transition lasers [14], is the very signature of carrier leakage [2, 15], whose suppression is key to high CW power [14, 16] and high CW wallplug efficiency [11, 16]. Multidimensional conduction-band (CB) engineering [17] has led to both high T0 (⩾250 K) and T1 (300–800 K) values [18–20] which, in turn, have resulted in record-high RT CW output power (5.1 W) and RT CW wallplug efficiency (21%) for QCLs [19]. However, those record-high values were obtained at the price of doubling the thermal-resistance value [17] a penalty that, as described below, can be avoided, and then CW wallplug efficiencies as high as 40% can be achieved.
Now that QCLs, at least in the mid-IR spectral range, have reached a high degree of maturity, it is high time for their CW operational characteristics to be fully understood as well as for CW reliable operation to become the subject of scientific research.
This article reviews the mechanisms behind the T0 and T1 values, with emphasis on the carrier leakage at and above RT, for QCLs emitting in the 3–16 μm wavelength range. The history of steady improvement in the T0 and T1 values via advanced CB engineering is covered in detail as well as what are the prospects for mid-IR QCLs to achieve their ultimate potential as long-term reliable sources of watt-range CW or average power.
Section 2 covers the carrier-leakage mechanisms both in early devices of relatively short barriers and in state-of-the-art devices of relatively tall barriers; such that there is at least one active-region energy state residing above the upper laser level. Then, shunt-type carrier leakage, through that energy state, is discussed in detail. Sections 3 and 4 discuss the temperature dependences of the basic electro-optical characteristics (threshold current, slope efficiency, maximum wallplug efficiency) in both pulsed and CW operation. At the end of section 4 the CW output power is addressed, and plots of RT CW output power and wallplug efficiency vs drive current, for 4.9 μm-emitting QCLs, are presented and compared for several sets of relevant parameters. Projected ultimate maximum CW power and wallplug-efficiency values are provided for 4.6 and 4.9 μm emitting devices. Section 5 covers highly temperature-insensitive QCL structures emitting in the 3.9–5.0 μm wavelength range. The progression of CB engineering towards complete carrier-leakage suppression is presented and key high-performance device designs are compared. Section 6 covers the temperature sensitivity of QCLs emitting in the 8–10 μm range with emphasis on recent designs that have achieved both carrier-leakage suppression as well as record-high internal differential efficiencies. Sections 7 and 8 cover the temperature sensitivity of QCLs emitting in the 12–16 μm and 3.0–3.8 μm wavelength ranges, respectively. In particular, intrinsic limitations of QCLs emitting below 3.5 μm are addressed, and a potential solution is presented.
2. Carrier-leakage mechanisms
2.1. Short-barrier devices
Early work on QCLs involved InP-based, lattice-matched 4.5–5.0 μm emitting devices or GaAs-based 8–9 μm emitting devices. Due to the relatively short barriers, the upper laser level was the highest energy state in the active region (AR). Initially, for very short barriers (i.e. Al0.33Ga0.67As), carrier leakage occurred primarily from the injector states to the continuum [21] (figure 1(a)) which was later identified as leakage to both Γ and X valleys [22, 23], even at temperatures as low at 77 K. With an increase in barrier height, via employing Al0.45Ga0.55As barriers, room-temperature operation was achieved [24], and, as shown in figure 1(b), carrier leakage was found to basically consist of phonon-assisted electron excitation from the upper level to the upper Γ miniband and from there to the continuum [25, 26]. As mentioned above, later it was shown that a small part of the 300 K leakage is due to scattering to the X miniband as well. Ortiz et al [25] treated leakage simply as thermionic leakage, based on the following expression for the escape/leakage time:
where is the wavefunction spatial extent in the upper laser level, kbT is the thermal energy, m* is the effective mass and, ΔEact is the energy difference between the upper level and the upper Γ miniband (figure 1(b)). Then, for the Jth expression, the upper-level lifetime, τul, was taken to be:
where τph is the optical-phonon scattering time of electrons in the upper level. Although such a simple treatment did not consider the wavefunctions' overlap or scattering to the continuum from the injector states, it worked well enough to both explain the T0 value (i.e. 140 K) for Al0.45Ga0.55As-barrier devices up 300 K and show that with increased barrier height the T0 value increases (e.g. a T0 value of 330 K was obtained from AlAs-barrier devices, up to 220 K, limited by device shutoff due the negative-differential-resistance effects). Later, Howard et al used a similar approach for the analysis of 8.2 μm emitting, InP-based lattice-matched devices [27]. Jin et al [26] further confirmed that carrier leakage, for Al0.45Ga0.55As-barrier devices, is primarily due to thermal excitation to the upper Γ miniband. For yet better carrier confinement, the insertion of AlAs spikes in the barriers was proposed and demonstrated [25], but, while the T0 value increased from 140 to 190 K, the devices stopped lasing at 250 K. Another relevant development was the finding that the carrier leakage is enhanced by the fact that the electrons in the injector ground state, g, and in the upper laser level, ul, are 'hot'; that is, they have a temperature significantly higher than the lattice temperature [28–31].
More recently, Shi and Knezevic [32], by considering nonequilibrium phonons, did show excellent agreement between experimental and theoretical Jth versus temperature curves, up to 240 K, for Al0.45Ga0.55As-barrier QCLs. Furthermore, nonequilibrium phonons were shown to increase the electronic temperatures in all subbands at both 77 and 300 K.
For lattice-matched InP-based 4.5–5.0 μm emitting, room-temperature carrier leakage turned out to be massive, due to the upper-level's closeness to the tops of the barrier layers; thus, very similar to the leakage in GaAs/Al0.33Ga0.67As QCLs (figure 1(a)). In 2002, an AlAs layer was inserted in the exit AR barrier by Yang et al [33] for the purpose of blocking carrier leakage, but, due to the severity of the leakage, it only improved the pulsed slope efficiency, while the threshold-current density and the T0 values remained basically the same [10]. Yet, that experiment provided the first evidence that the slope efficiency in QCLs is strongly affected by carrier leakage, a well-known fact in interband-transition devices [2, 14, 15, 34]. For QCLs this fact was analytically identified in 2009 [10] and the slope-efficiency equation was modified to account for it in 2010 [11].
Carrier leakage in those early QCLs constituted true leakage to the continuum, just as in interband-transition lasers. However, as seen below, for state-of-the-art QCLs carrier leakage is primarily a shunt-type leakage mostly occurring within the AR.
2.2. Tall-barrier devices
In order to suppress carrier leakage, relatively tall barriers were introduced for InP-based 3.5–5.5 μm emitting devices by the use of strain-compensated (SC) structures [35, 36]. 8–9 μm emitting, lattice-matched InP-based devices already had relatively tall barriers, which allowed the first demonstration of RT CW operation [9]. A more recent development has been the use of SC InP-based structures for 7–9 μm emitting QCLs [37–40], which, in turn, resulted in significantly improved performance due to some degree of carrier-leakage suppression.
Tall barriers bring about stronger quantum confinement and, in turn, several energy states emerge, in the AR, above the upper laser level. Then, as shown in figure 2, carrier leakage mainly consists of thermal excitation of injected carriers from the upper laser level (level 4 in the figure) to the high-energy AR states, followed by relaxation to the low-energy AR states [11, 41]. That is, carrier leakage for such devices is 'de facto' a shunt-type leakage current within the AR, not leakage to the continuum. Actually, there is some leakage to the continuum, in that some of the carriers thermally excited to the highest-energy AR state (i.e. state 6 in figure 2) can be further excited to the upper-Γ-miniband, but that is a negligible part of the total leakage current [11, 16]. Also, as shown by Pflügl et al [42] if the barriers are shallower than in state-of-the-art devices and the energy differential between the high-energy states is of the order of the phonon energy, part of the carriers injected directly into the those states (shown with thin lines in figure 2) will leak to the continuum, as identified from V–I curves [42]. Nonetheless, even in such cases the leakage at and near threshold is primarily shunt leakage through states 5 and 6. The fact that the shunt leakage may in part be due to carriers thermally excited from state g directly to the high-energy AR states has been suggested by Masselink et al for high-CB-offset 3.9 μm emitting QCLs [43]. However, for conventional 4.5–5.5 μm emitting devices, that is likely to be negligible at threshold, since there is poor overlap between wavefunctions of the g and the high-energy AR states. Yet, at high drive levels above threshold, as the state-g energy approaches that of the upper laser level, such leakage currents become stronger and may well in part explain the drooping of pulsed L–I curves at high drive levels [17].
Download figure:
Standard image High-resolution imageIt should be stressed, just as in the case of carrier leakage to the continuum, that thermal excitation to higher energy states is significantly enhanced by the fact that the injected carriers are 'hot' [11, 44]. This fact has been confirmed, in that experimental T0 values have been accurately predicted for both conventional [11, 45] and deep-well [11] 4.5–5.0 μm emitting QCLs only when considering hot electrons in the upper laser level. More recently, electrons in the upper level were found to be hot in 8–9 μm emitting QCLs as well [46, 47], which, as we shall see below in section 6, are responsible for enhanced carrier leakage as well.
For a four-quantum-well (QW) AR, as shown in figure 2, the leakage-current density at threshold, , can be written as such:
where q is the electron charge, n5 and n6, and τ5,leak and τ6,leak are the sheet densities and electron-leakage lifetimes, respectively, corresponding to the AR's upper states 5 and 6. More specifically and , where is the lifetime corresponding to thermal excitation from state 6 to all the states in the upper Γ miniband (see equation (5)), and the other lifetimes correspond to relaxation from states 5 and 6 to states 3, 2, and 1, respectively. The sheet densities n5 and n6 are defined by the following equations [11]:
where τ5,tot and τ6,tot are the net lifetimes corresponding to electron scattering from state 5 to states 1–4 and 6, and from state 6 to states 1–5 and to states in the upper Γ miniband, respectively, and n4 is the sheet density in the upper laser level, state 4. The lifetime corresponding to thermal excitation of electrons from a lower-energy state i to a higher-energy state j, τij, which is predominantly due to LO-phonon absorption scattering for large-energy separations, is approximated by the following expression [11]:
where is the energy difference between states j and i, is the longitudinal-optical (LO) phonon energy (), is the occupation number of phonons (assumed to be in thermal equilibrium with the lattice at temperature Tl), Tei is the electronic temperature for state i, which under very low duty-cycle operation (i.e. negligible Joule heating) is obtained from:
where αe–l is the electron–lattice coupling constant [28, 29]. Applying this equation to scattering from the upper laser level ul to the next higher active-region energy state ul +1 and taking into account that typically the value of the quantity is in the 0.15–0.20 range [11], and , equation (5) is well approximated by the following [48]:
Then, the scattering rate is quite similar to the case when carriers are in equilibrium with the lattice, except for the fact that the lattice temperature is replaced by the electronic temperature of the upper laser level. From equation (7) it is clear that to effectively suppress carrier leakage, the Eul+1,ul and τul+1,ul values should be increased as much as possible. As for minimizing the Teul value, that can be done by making sure that electrons are kept 'cool' in state ul either via strong quantum confinement in structures of high conduction-band (CB) offset [30, 44] or by minimizing the Jth value.
We have found [11, 16] that for E54 ⩾ 50 meV the primary leakage path is through state 5. Then, leakage through state 6 can be neglected, and a simplified carrier-leakage diagram is considered (figure 3) for which relaxation to the extractor states penetrating into the AR is shown as well. By employing the rate equations for a four-state system, the pumping efficiency term [17, 20, 41] , which is the measure of the impact of carrier leakage on the Jth and ηs values, can be shown [49] to be given by:
for which transitions the extractor states penetrating the AR (e.g. states 1' and 2' in figure 3) should be considered as well [17]. It is clear that a low scattering rate 1/τ45 from the upper level, state 4, to the next higher AR energy level, state 5, is key to minimizing the effect of carrier leakage on the Jth and the ηs values.
Download figure:
Standard image High-resolution imageThe above equations are for inelastic scattering. However, when calculating T0 values one has to take into account the effect of elastic scattering on both the electroluminescence (EL) linewidth [50] and the various lifetimes involved [51, 52]. For instance, for accurate estimates of the T0 values for conventional and deep-well 4.6–4.8 μm QCLs of double-phonon-resonance (DPR) lower-level depopulation scheme [11, 16] we used the temperature dependence of the EL linewidth from the EL data, for DPR QCLs, modelled and experimentally validated by Wittman et al [50]. As for the temperature dependence of the upper-state lifetime τul, we considered the effect of elastic scattering on its value by using the 1.5 ps elastic-scattering lifetime validated by Pflügl et al from their study of 4.6 μm emitting structures [42]. More specifically, Pflügl et al obtained best fits to current-density versus temperature curves of 4.6 μm emitting QCL structures only when considering τul having both inelastic- and elastic-scattering components, with a temperature-independent elastic-scattering lifetime of 1.5 ps. Later, Flores et al [53] revealed the existence of interface-roughness (IFR)-assisted carrier leakage at cryogenic temperatures, where (equilibrium)-phonon-assisted scattering is negligible. That work, in particular, was an excellent demonstration of the existence of hot electrons in the upper level at high current densities, and of the direct impact carrier leakage has on the external differential quantum efficiency [11, 20]. More recently, Masselink et al [54] have shown that IFR-mediated leakage, at room temperature, while mildly affected by the AR QW/barrier geometry, is significantly reduced as increases. That finding may reflect the fact that in QCLs designed for large values [17, 41] the wavefunction overlap between the ul and the ul + 1 states is poor (i.e. large values). That is, QCL structures of high and values, such as tapered active-region (TA) QCLs [17, 20], suppress carrier leakage triggered by both elastic and inelastic scattering.
Finally, carrier scattering to satellite valleys, such as the X or L valleys of InGaAs QWs, was blamed in the past for carrier leakage in mid-IR (3.5–5.0 μm) QCLs, and more recently for carrier leakage in 4.0 μm emitting QCLs [55]. However, experimental data [56] have shown that the X and L valleys, for the InGaAs QWs used in 4.0–5.0 μm range, are much higher in energy above the Γ valleys than previously thought. Then, Flores et al [41] showed that at λ = 3.9 μm the L valley is above the high-energy AR states; thus, not a factor in leakage. In addition, Aldukhayel et al [57] have shown that at λ = 3.5 μm carrier leakage to the L valley amounts to only about 3% of the threshold current at room temperature. Thus, it clearly appears that shunt-type leakage current through the high-energy AR states is primarily responsible for carrier leakage in QCLs of emission wavelengths >3.5 μm.
3. Electro-optical characteristics in pulsed operation
3.1. Threshold-current density
3.1.1. Definitions.
The threshold-current density, Jth, is the sum of Jth in the absence of backfilling and electron leakage, J0,th, and the current densities required at threshold to compensate for backfilling, , and electron leakage, , [11]:
which, considering the pumping-efficiency term [17, 20, 41] , can be rewritten:
Note that ηp, as defined here, is the efficiency with which electrons injected into the upper laser level are used for lasing at threshold, when taking the tunnel-injection efficiency to be unity. We also can define a total injection efficiency for which, to a very good approximation, is taken to be a factor in the denominators of J0,th [6] and Jbf,th, when the actual is close to unity. It just happens that ηp, as used in equation (9b), is equal to the differential pumping efficiency in lasing, at and slightly above threshold [11, 17, 20, 49], as is the case for leakage currents in interband lasers [34].
In the case of 4.5–5.5 μm emitting lasers of DPR lower-level depopulation scheme, as that shown for a conventional QCL in figure 4, is given by equation (3), while the other components are defined as follows [11]:
where is the global 'effective' upper-state lifetime [17] (due to both inelastic and elastic scattering) in the absence of carrier leakage, and taking to be unity, with τ4g and τ3g being the global lifetimes in the upper and lower levels; αtot is the sum of the mirror losses (αm) and actual waveguide losses (αw); gc is the gain cross-section [58]; ns is the electron sheet density in the injector; Δinj is the energy difference between the lower laser level and the ground state in the next injector region; Teg is the electronic temperature in the injector miniband [44]; is the lifetime corresponding to electron relaxation from state 5/6 to state 3; is the lifetime in the lower laser level (state 3). τ4g and τ3g are global lifetimes, in that they correspond to transitions to both lower AR states (e.g. to states 3, 2, 1 from state 4) and to extractor states penetrating into the AR (e.g. the three blue-coloured extractor states penetrating the AR in figure 4). That is, the global upper-level lifetime can be expressed as:
Download figure:
Standard image High-resolution imageThe expression for is derived [16] by writing a rate equation for the lower laser state, which is populated (backfilled) via LO-phonon reabsorption from states 1 and 2 (assumed to be strongly coupled to the injector states and thus sharing a common electronic temperature Teg) and by LO-phonon emission from states 5 and 6. Taking Teg = (1 + δeg) Tl one obtains for equation (9d) the following expression:
For 4.5–5.5 μm emitting QCLs, around room temperature, the δeg quantity takes values in the 0.15–0.20 range depending on the Jth value considered. Then, as seen from equation (11), the numerator in the exponential part of the expression hardly changes, and what mostly affects the backfilling current is the denominator. The non-exponential term turns out to be negligible compared to the exponential term [11]. Thus, equation (11) can be well approximated by:
For a device with Jth and Δinj values of 1.5 kA cm−2 and 150 meV one obtains δeg = 0.18, and then, from equation (11), the effect of having hot electrons in the injector basically doubles the Jbf value by comparison to the case when they are not considered. That is, hot electrons strongly affect backfilling. This may explain why, for devices designed with Δinj = 130–150 meV, the derived Jbf value was found to be significantly higher than the calculated value [59] when considering that Teg is equal to with lattice temperature (i.e. equation (12a) with Tl instead of Teg). For short-cavity (1–2 mm) devices [46, 60] or for devices of low Δinj (20 meV) [61] the backfilling current at RT can be as high as 70% of the Jth value. While for the latter case it is clear why the RT Jbf value is large, the former can be understood from the fact that for short-cavity devices the RT Jth value is relatively high, which will increase the Teg value and, in turn, both Jbf and Jleak will increase, leading to high so-called 'transparency' currents. Revin et al [60] have shown that, if the lower level is not depopulated via resonant LO-phonon transitions to lower AR energy states, the backfilling current becomes > 70% the RT Jth value, since most electrons in the injector reside in the lower level. As a result T0 decreases from a typical value of 137 K for moderately high doped, conventional 5.5 μm emitting devices, to a value of only 106 K.
For devices with Δinj ≈ 100 meV [38, 40] equation (12a) with Tl instead of Teg was found to overestimate the Jbf value. Then, a model involving backfilling from all subbands in the injector has been proposed by Maulini et al [38], but, while it provides lower Jbf values than when using the conventional formula, it implies much higher injector-doping values than indicated by the maximum current-density Jmax value. For instance, in [40] the thus estimated ns value was 5.7 × 1011 cm−2, while the Jmax value of ~10 kA cm−2 indicates an actual ns value of ~2.4 × 1011 cm−2, in agreement with typical values for 8–9 μm emitting QCLs [50]. We find that Maulini et al's Jbf formula predicts well the Jbf value only if one assumes a Teg value ~100 K higher than the lattice temperature at RT, which agrees with the findings that electrons are hot in 8–9 μm QCL as well [46, 47]. That is, reasonably accurate Jbf values are obtained with Maulini's derived expression, only when considering hot electrons:
where Ninj stands for the number of subbands in the injector lower miniband. In either case, the electron temperature in the injector lower miniband has a strong effect on the Jbf value; thus, lowering the Teg value, for instance by lowering the overall Jth value, is likely to significantly decrease the backfilling current.
Currently, for characterizing device performance, the following Jth expression is used:
where αw,eff is an effective waveguide loss, Γ is the transverse optical-confinement factor to the core region, and g is the differential gain. Using equations (9b) and (9c) and comparing to equation (13), the following relationships can be written:
That is, from a Jth versus 1/L plot, where L is the cavity length, the derived value for the modal differential gain, Γg, reflects not only the transverse optical-mode confinement (i.e. Γ), but also the degree of carrier leakage, as expressed by the pumping-efficiency term ηp, which for conventional mid-IR QCLs is ~0.85 at 300 K and ~0.77 at 360 K [11]. In effect, ηp represents a carrier-confinement term, as expected given the shunt-current nature of Jleak. By defining, from (14a), a modified effective upper-state lifetime τ'up,g = ηpτup,g, and using equations (9b) and (12a), the Jth equation becomes:
in a form that includes all optical losses, the backfilling term as well as takes into account the carrier leakage through the use of rather than τup,g. The difference is significant, especially for conventional devices in which case carrier leakage increases [11] the Jth value by ~18% at 300 K and ~30% at 360 K. As for the experimentally derived value of the effective waveguide loss (equation (14b)), it contains the actual waveguide loss plus a term reflecting backfilling, αbf, sometimes called 'resonant' waveguide loss [59]. Due to backfilling, αw,eff is a strongly temperature-dependent quantity, which represents the actual waveguide loss αw only at cryogenic temperatures. In order to extract αw values at room temperature one has to compare the slope efficiencies of devices of different αm values [40, 59].
As mentioned above, we have found that for E54 ⩾ 50 meV the primary leakage path is through state 5 [11, 16]; thus, carrier leakage can be assumed to occur only via state 5 (figure 3). Then, from the rate equations [58] with J = J4 + J5,leak, where J4 = J0 + Jbf = qn4/τ4g and J5,leak = qn5/τ5,leak = qn4τ5,tot/τ45τ5,leak, an expression for ηp is derived (equation (8)) which, in turn, provides the following expression for effective upper-state lifetime :
That is, due to carrier leakage, the required value for the upper-state carrier-sheet-density n4 at threshold is higher than in the case of no leakage. Note that for a short-barrier case (i.e. no level 5; thus, leakage to the continuum) τ45 becomes τ4,leak and the equation reduces to equation (2b).
Finally, in the Jth equation some authors [62, 63] call the term which is added to the J0,th term, transparency-current density Jtr, since backfilling is thought of as corresponding to 'resonant' intersubband (ISB) absorption [59]. However, the leakage-current density at and above room temperature is by no means a negligible term for QCLs in general [11, 41] and the shunt current through the high-energy AR states contributes to the backfilling current (equation (9d)). Thus, considering the Jth − J0,th quantity to be a transparency-current density is incorrect, since carrier leakage and carrier-leakage-induced backfilling have nothing to do with absorption.
3.1.2. Temperature dependence.
Conventional high-performance QCLs emitting in the 4.5–5.5 μm range have suffered from severe carrier leakage, which manifested itself as low characteristic temperature T0 values (130–150 K) [11, 45, 60, 65]. Such devices had moderately high-doped injector regions, in order to obtain the large dynamic range necessary for high-power operation. If the devices have low- doped (i.e. ns = (0.5–0.7) × 1011 cm−2) injectors, the T0 value is ~200 K, as the backfilling current is significantly reduced [42, 66–68]; thus, such devices are useful mostly for low-power-dissipation, low-output-power applications.
By suppressing the carrier leakage via multidimensional CB engineering [17], in moderately high-doped devices, has led to T0 values in the 230–285 K range [10, 18–20]. More specifically, the carrier leakage was suppressed by increasing the value, from ~44 meV for conventional devices to values in the 60–90 meV range, and the value, from ~0.22 ps for conventional devices to values in the 0.32–0.72 ps range, by means outlined in section 5. For low-doped devices with suppressed carrier leakage a T0 value of 363 K was recorded [13] due to reduced backfilling and, due to a strongly diagonal transition, a relatively large, highly temperature-insensitive Jth,0 term. Good agreement with experiment was obtained for calculated T0 values for both deep-well (DW) and conventional moderately high-doped, 4.6–4.8 μm emitting devices, of DPR lower-laser-level depopulation scheme, by using the above outlined definitions for Jth and an iterative process outlined in [11, 16] (table 1). For calculating Tei values we used αe–l = 35 K cm2 kA−1, as measured for Teg in 4.8 µm emitting QCLs [44]. Since state g is strongly coupled to the upper level, state 4, it was assumed that at threshold Te4 ≈ Teg, which was recently validated to be accurate from Monte-Carlo based calculations of GaAs QCLs [32]. We also used Te5 ≈ Te6 ≈ Te4, although such assumptions may well not be accurate. In any event, since leakage primarily occurs through state 5, the Te5 and Te6 values are of little relevance.
Table 1. Measured and calculated parameters for conventional and deep-well QCLs.
Measured T0 | Jleak / Jth | Calculated T0 | ||
---|---|---|---|---|
Temperatures | 300–360 K | 300 K | 360 K | 300–360 K |
Conventional QCL | 143 K | 0.15 | 0.23 | 167 K |
Deep-well QCL | 253 K | 0.08 | 0.13 | 234 K |
Similar good agreement between experimental and theoretical T0 values (i.e. 143 K versus 147 K, over the 260–300 K temperature range) has been obtained for 4.6 μm emitting devices of single-phonon-resonance (SPR) lower-level depopulation scheme [45], by assuming hot electrons in the upper level. The T0 values are lower than for DPR-type devices of ~70 meV value, most likely due to higher contribution from backfilling currents, since Δinj is significantly smaller than for DPR devices (i.e. ~125 meV [58] versus ~150 meV).
For 3.9 μm emitting SPR devices [41], suppression of carrier leakage, by increasing by design the value to ~120 meV, has resulted in a relatively high T0 value (i.e. 175 K over the 150–300 K temperature range), not as high as for optimized ~3.8 μm emitting SPR devices [69]; possibly due to a relatively low Δinj value (135 meV) and/or higher injector doping level.
Many of the T0 values mentioned above are from 3 mm long, uncoated-facets devices. However, T0 is a function of mirror loss, generally decreasing slowly with increasing cavity length [41, 60]. This can be understood from equations (9b) and (9c); in that, as decreases the relatively temperature-insensitive portion of Jth (i.e. Jth,0) decreases; thus, Jth becomes more temperature sensitive for longer-cavity devices.
As expected, another factor that affects the T0 value is the electronic temperature in the upper level, Teul. Since at threshold and a lattice temperature, Tl, Teul = Tl + αe–l ⋅ Jth, it follows that, if Jth significantly decreases, Teul will decrease which will lead to lower backfilling and leakage currents; thus, to higher T0 values. A good example is what happened in the case of 4.6 μm emitting devices of non-resonant-extraction (NRE) lower-level depopulation scheme when going from 3 mm long, uncoated devices of αw ~ 1.2 cm−1 [70] to 3 mm long, high-reflectivity (HR)-coated devices of αw ~ 0.5 cm−1 [63]. Then, the RT Jth value decreased from ~1.5 kA cm−2 to ~0.95 kA cm−2. In turn, the T0 value increased from 140 K [70] to 168 K over the 290–350 K temperature range [55]. However, it is worth noting that although the value was relatively high (~63 meV); thus, comparable to those for DW and TA devices [17], the T0 value for 3 mm long, uncoated devices, with αw ~1.2 cm−1 is much smaller than for DW/TA devices (i.e. 140 K versus 230–255 K). We attribute this difference in large part to the fact that, unlike in the case of DW and TA devices, the increase in value is not accompanied by an increase in the value, since there is strong overlap between the wavefunctions of the upper level and the next higher AR energy state [71]; thus, the leakage current is similar to that in conventional QCLs. Further evidence that carrier leakage is significant in NRE-type QCLs comes, in part, from the fact that the experimentally [63] measured value for the internal differential efficiency per period [17] is significantly smaller (i.e. ~50%) than that (i.e. ~70%) for 4.9 μm emitting 'shallow-well' QCLs of suppressed carrier leakage [19], which have been shown [17] to, de facto, be TA-SPR-type QCLs. (In fact, the NRE-QCL value is smaller than that for conventional 4.65 μm emitting DPR devices (i.e. ~60%) [72]). Another reason for the lower value in NRE versus TA-SPR devices is a lower laser-transition efficiency [73] ηtr value (i.e. 0.81 versus 0.89), as calculated from device modelling including elastic scattering [58].
Similarly, 4.0 μm emitting NRE-type QCLs with In0.73Ga0.27As QWs [55] most likely owe their relatively low T0 and T1 values (135 K and 168 K, respectively) to carrier leakage through state 5, since significantly higher Jth(300 K) values translate into higher Teul values than for NRE-type 4.6 μm emitting QCLs. The suggested reason: leakage to indirect valleys, is highly unlikely, given that the most recent measurements of the L-valley position in strained In0.73Ga0.27As QWs [74] found it to be 40 meV higher than expected above the Γ valley; thus, for the QCLs in [55], the L valley lies above state 5, similar to what Flores et al found for their 3.9 μm emitting QCLs [41].
Another example of the effect of the Teul value on the T0 value is what happened for SPR-type 4.7 μm emitting devices of strong coupling [75]. Although the value was high (71 meV) [58], T0 was only moderately high (188 K) since the RT Jth value was high (~2.5 kA cm−2) which, in turn, increased the Teul value, and thus the leakage current. Evidence of carrier leakage similar to that in conventional DPR 4.7 μm QCLs is that the measured T1 had a rather small value of 149 K [76].
3.2. External differential quantum efficiency
3.2.1. Definition.
The external differential quantum efficiency ηd for a QCL having Np periods can be written:
where ηp, as derived in [49], is the differential pumping efficiency at and close to threshold [11, 49] and is given by:
ηtr is the laser-transition differential efficiency [73]; αm and αw are the mirror loss and actual waveguide loss (due to free-carrier absorption (FCA) and ISB absorption), respectively. The product ηpηtr can be thought of as the internal differential efficiency per period [16], when the tunnel-injection efficiency (into the upper level) is taken to be unity. However, in the case is less than unity, due for example to electron scattering from the injector ground state in the next higher energy state(s) above the upper level [42, 43], it becomes a multiplying factor in equation (17a), since it accounts for another leakage path [34]. Then, the total injection efficiency, in equation (17a), is equal to ηp, basically as used in equation (9b). For QCLs, the validity of using the ηp term in the equation (17a) was confirmed by the interpretation [10] of Yang et al's experimental data from QCLs of tall exit barrier [33], and by Flores et al's experimental analysis of carrier leakage via IFR [53].
3.2.2. Temperature dependence.
Severe carrier leakage in conventional high-performance QCLs emitting in the 4.5–5.0 μm range has also manifested itself as low characteristic-temperature T1 values (140–170 K) for the slope efficiency over wide heatsink-temperature ranges above RT [11, 42, 65]. Since ηtr is basically temperature independent, it is obvious from equation (17a) that the T1 parameter is primarily determined by the variations with temperature of both ηp and αw [11]. Note that, just as in the case of interband-transition lasers [14], carrier leakage strongly affects the T1 value. That is, T1 is a much better indicator of carrier leakage than the T0 parameter, since T0 reflects the temperature dependence of the backfilling current as well.
While T0 values have been successfully predicted theoretically for both conventional [11, 45] and CB-engineered 4.6–4.8 μm QCLs [11], that has not been the case for T1 values. As shown in [11], the use of the calculated Jleak values that worked in predicting well the experimental T0 values, resulted in T1 values significantly higher than the experimental ones. We concluded that an increase with temperature of the αw value had to account for this difference; a reasonable assumption since αw of state-of-the-art 4.6–4.8 μm emitting devices have been reported [70] to be primarily (~90%) due to ISB absorption. Since the ISB-absorption portion of αw, αISB, increases with temperature [77, 78], αw should also increase with temperature. While, for devices emitting in the 4.5–5.5 μm wavelength range, there are no experimental data as far as the temperature dependence of αw, such data have been published for 8.4 μm emitting QCLs [79]. We plot in figure 5(a) the αw and αw,eff variations with increasing temperature, extracted from figure 4.8 of [79], for DPR- and bound-to-continuum (BC)-type devices [50]. Since the free-carrier-absorption part of αw is weakly temperature dependent, and αISB can be as much as 73% of αw [80], it becomes clear that the increase of αw with temperature (e.g. from 7.6 to 11.3 cm−1 over the 243–400 K range, for DPR devices) is primarily due to ISB absorption for both device types. The much stronger increase with temperature of αw,eff is due to increased backfilling, and it is quite similar to the behaviour of αw,eff for 8.2 µm emitting DPR-type devices [78].
Download figure:
Standard image High-resolution imageFigure 5(b) shows, for the 8.4 μm emitting DPR QCLs [50], the relative variations with temperature of the slope efficiency ηsl and of the quantity αm/(αm + αw), with αw values taken from figure 4.8 of [79], with respect to their 20 °C values. While ηsl decreases by ~33% over the 20–103 °C range; thus, giving a T1 value of 263 K, the αm/(αm + αw) quantity decreases by only 18%. The difference; that is, an additional 18% decrease, is due to the temperature variation of the ηpηtr product. Since ηtr is basically temperature insensitive, the ~18% decrease reflects the increase in relative carrier leakage (i.e. Jleak/Jth) with increasing temperature, similar to what happens in 4.5–5.0 μm QCLs [11]. Such a relatively high increase in Jleak/Jth is expected, in view of the fact that electrons have been found to be 'hot' in the upper level of 8–9 μm emitting QCLs as well [46, 47]. Such leakage, at and above RT, explains the relatively low values for T0 and T1: ~160 K [50] and ~263 K, respectively. Further proof of carrier leakage is the relatively low RT value (~0.86) obtained from more advanced modelling of the same devices [81]. We calculate for the 8.4 μm DPR devices [50] values of ~0.98 and ~0.89 for and ηp which give an value of ~0.87; in very good agreement with the 0.86 value deduced in [81]. As we shall see in section 6.2.3, the carrier leakage can be suppressed by employing a step-taper TA (STA)-type AR design quite similar to that proposed for high-performance 4.5–5.0 μm QCLs [17].
Carrier-leakage suppression via conduction-band engineering of 3.9–5.0 μm QCL [17, 41], has led to T1 values in the 285–800 K range [13, 18–20, 41, 49]. For 4.9 μm, TA-STR-type ridge-guide devices [13, 19], increasing the average doping per period and the period number from 30 to 40 led to an increase in αw from ~0.5 cm−1 to ~0.75 cm−1, consistent with the observation that for state-of-art-devices αw is mostly due to ISB losses in the core [70] and with that fact that αISB is proportional with the number of carriers in the injector (i.e. out of the 50% increase only 14% is due to the higher Γ value, with the rest being mostly due to the ~25% increase in doping). Since, as mentioned above, αISB increases with temperature, an increase in the αw value with temperature will, in turn, lead to a decrease in the T1 value. Indeed, for the same mirror-loss value, T1 decreased from 645 to 343 K; thus, confirming, for 4.5–5.0 μm QCLs as well, that the αw increase with temperature is, in a significant way, partly responsible for the T1 value. Note also that, if αw is relatively cavity-length (L)-independent, the T1 value will increase with decreasing L, as long as the carrier leakage does not significantly increase.
As mentioned above in the threshold-current section, significantly lowering the Jth value causes carrier-leakage suppression due reduced electronic temperature in the upper level, as evidenced by an increase in the T0 value of NRE-type devices as Jth was lowered. Similarly, for those same NRE devices, significantly Jth lowering led to an increase in the T1 value from 140 to 295 K [55]. Since for those devices the ratio of αw to αm is basically the same, it can be clearly seen from equations (3), (7) and (17a) that the increase in T1 value can be primarily attributed to better carrier-leakage suppression through reduced electronic temperature. Nonetheless, as discussed in section 3.3, there is still a fair amount of carrier leakage, as evidenced from lower values than in conventional 4.6 μm emitting. DPR-type QCLs (i.e. 50% versus 60%).
3.3. Maximum wallplug efficiency
3.3.1. Definition.
The expression for the pulsed maximum wallplug efficiency ηwp,max can be written as [11]:
where ηs reflects the deviation from a straight line of the pulsed L–I curve at the ηwp,max point [11], αm,opt is the optimal mirror loss [63], Jwpm is the current density at the ηwp,max point, hν is the photon energy and Vwpm is the voltage at Jwpm. The value for αm,opt has been found both theoretically [63] and experimentally [19, 63] to be ~2.2 cm−1. Analysis of published data reveals that the Jwpm value, in both pulsed and CW operation, is an approximate fixed multiple of the pulsed Jth value at RT, Jth (300 K); that is: Jwpm ≅ BJth (300 K). For instance, for devices with significant carrier leakage [64, 71] we found the B value to be ~2.5 [11]. However, for devices of moderately high doping and suppressed carrier leakage [19] the B value increases to ~3.0. As for ηs, its value is found to be ~0.90 at 2.5 × threshold for devices with typical carrier leakage [63]. For devices with suppressed carrier leakage [19] ηs is ~0.95 at 2.5 × threshold and ~0.92 at 3 × threshold. The higher ηs value at the same drive level most likely reflects carrier-leakage suppression. That is, Jleak continues to increase about threshold, since it is proportional with n4, which increases above threshold, and, as observed by Pflugl et al [42], at high drives above threshold there is leakage due to carrier thermally activation from the injector ground state to high-energy AR states. Thus, the higher the energy difference between the upper level ul and/or state g and the state ul + 1, the less droop is expected for the L–I curve at high drives above threshold.
3.3.2. Temperature dependence.
At a heatsink temperature Th above room temperature (i.e. above 300 K) ηwp,max can be written as such:
With increasing heatsink temperature the pulsed ηwp,max is clearly affected by the T1 and T0 values. We define a characteristic temperature T2 for ηwp.max from the relation:
Taking both the T0 and T1 values for conventional, 3 mm long, back-facet HR-coated high-performance QCLs [11, 42, 82] to be 143 K over the 300–360 K temperature range, and B = 2.5, we obtain T2 = 71 K. In sharp contrast, for TA-type devices [20] with T0 = 231 K, T1 ≅ 750 K and B = 3, the calculated T2 value is 249 K. This means that over the 300–360 K temperature range ηwp.max would vary 2.66 times slower than for conventional QCLs. For TA-type devices there are experimental results for the temperature dependence of ηwp,max for the so-called shallow-well QCLs [19]. Given T0 = 244 K, T1 = 343 K and B = 3 in that case, the calculated T2 value is 186 K in very good agreement with the 195 K value that can be extracted from the experimental data [19]; thus, confirming the accuracy of the approximation in equation (19). Then, for 'shallow-well' TA-type devices the ηwp,max value varies 2.15 times slower than in conventional devices; reflecting the beneficial effect that carrier-leakage suppression brings about as far as low temperature sensitivity of the ηwp,max value. Furthermore, carrier-leakage suppression is the major factor leading to the almost doubling of the single-facet, pulsed ηwp,max value at ~300 K; that is, from 15% in 4.6 μm emitting NRE-type [71] and SPR-type devices [45, 75] to 27% in 4.9 μm emitting TA-SPR devices [19]. More specifically, out of the 1.8 factor improvement in ηwp,max over the NRE-QCL value, 1.55 is primarily due to carrier-leakage suppression and higher ηtr value (i.e. 1.40 from higher value and 1.11 from the higher drive level at which ηwp,max occurs) with the rest of 1.16 being mostly due to a lower Vwpm value. Furthermore, the 27% ηwp,max value at λ = 4.9 μm corresponds to an ~29% value at λ = 4.6 μm [73]; thus, there is virtual doubling of the ηwp,max value at a given wavelength.
However, it is worth noting that the highest reported pulsed value for ηwp,max (i.e. 27% for 4.9 μm emitting devices) was obtained from devices with an value of only 70%; in that, with virtually suppressed carrier leakage and a calculated ηtr value of ~0.89 [58] one would expect values of ~86%. The 27% ηwp,max value is in good agreement with theoretical predictions by Faist [73], but only because those predictions consider ≈ 70%. That is, we find that the solid-line prediction curve in figure 1 of [73], derived with key parameters from a 9 μm emitting QCL ([12] of that work), corresponds to ≈ 70%. The difference between the experimentally obtained value and the expected ultimate value may be due to IFR-mediated carrier leakage, since the TA-SPR device uses has high CB offsets in the injector region which, in turn, are conducive to IFR-assisted leakage from state g to the high-energy AR states [43, 54]. Such leakage means that the value is less than unity, which, in turn, decreases the value in equation (17a). In sharp contrast, recently developed TA-type 8.7–8.8 μm emitting devices have provided values in the 85–90% range [83]; thus, approaching theoretical limits. Since for 8–9 μm QCLs the CB offsets are significantly smaller than for 4.5–5.0 μm QCLs, it is quite possible that IFR-mediated carrier leakage has a negligible impact on 8–9 μm QCLs, unless their core regions consist of heavily strain-compensated structures, in which case drops again to relatively low values (58%) [40].
4. Electro-optical characteristics in CW operation
4.1. Threshold current
Under CW operating conditions the device-core temperature increases by the quantity ΔTact defined as:
where Tl is the lattice temperature, Rth is the thermal resistance, Pel is the electrical power dissipated in CW operation (i.e. Pel = A J V, where A is the pumped area), Popt is the CW output power and ηwp,cw is the wallplug efficiency in CW operation.
At threshold, Popt is zero, so the CW threshold is given by:
with:
where ΔTact,th is the device-core temperature rise at lasing threshold, Vth is the voltage at threshold and Gth is the thermal conductance. Then, the key factors for minimizing the Jth increase from pulsed to CW conditions are maximizing the T0 and Gth values. As pointed out above, T0 can be maximized by carrier-leakage suppression, which significantly reduces both the Jleak and Jbf values, as well as by low doping which reduces the Jbf influence on T0. As for maximizing Gth, heat can be most effectively removed by using buried-heterostructure (BH) geometry devices mounted in the epi-down configuration on high thermal-conductance heatsinks. From published data on BH 4.6–4.8 μm emitting, conventional, 40-period QCLs mounted episide-down on diamond [64, 72] we find that, as the buried-ridge width w varies from 8.6 μm to 11.6 μm, Gth is approximately inversely proportional with , in good agreement with theory [84]. This is expected, since a significant portion of heat removal occurs laterally, away from the buried-core region. Therefore, w needs to be significantly lowered. However, too narrow a ridge increases the series resistance and decreases the output power. For high CW power, a good compromise for the w value appears to be 8 μm, as recently used in devices of the highest CW power and single-facet ηwp,max reported to date [19]. In addition, w ≈ 8 μm also ensures, for λ = 4.5–5.0 μm, stable, single-spatial-mode operation.
It has been shown [7] that the maximum heatsink temperature for CW operation is obtained from equation (21) by setting dTh/dTl = 0, which, in turn, for an initial Th value of 300 K, gives:
Thus, in order to maximize the CW-operating temperature high T0 and Gth values are required, which, as mentioned above, mean carrier-leakage suppression, low injector-doping level and narrow-ridge, BH devices mounted on high-Gth submounts. In addition, the power dissipation at threshold (i.e. Jth (300 K) × Vth) needs to be minimized. Generally high-temperature CW operation is desirable for DFB lasers of large tuning range. A good example is a BH-type, 9 μm emitting DFB laser [80], mounted episide-down on AlN submounts and purposely low doped (ns = 0.7 × 1011 cm−2) for low-power dissipation. As a result, the device had a relatively high T0 value (189 K) and achieved the highest CW-operating temperature reported to date (i.e. 150 °C) for any QCL.
4.2. Maximum wallplug efficiency
The maximum, single-facet CW wallplug efficiency, at a heatsink temperature Th of a value close to 300 K, can be expressed as [17]:
where ηd,opt (Th) is the (pulsed) differential quantum efficiency for an optimal mirror loss αm,opt. ΔTact,wpm is the core temperature rise at the ηwp,max point:
where Tl,wpm is the lattice temperature at CW ηwp,max, Rth is the thermal resistance, Pel,wpm is the electrical power dissipated in CW operation at the ηwp,max point (i.e. Pel,wpm = A Jwp,maxVwpm, where A is the pumped area). As evident from equations (24a) and (24b), both the CW ηwp,max and ΔTact,wpm values are strong functions of the T1 and Gth values and weak functions of T0. Therefore, to maximize ηwp,max and minimize ΔTact,wpm, carrier leakage has to be suppressed (to obtain high T0 and T1 values), the operating voltage needs to be minimized, and the Gth value needs to be increased as much as possible. As discussed above, for BH-type, 4.5–5.0 μm emitting QCLs a good compromise for the ridge-guide width, w, that provides relatively high Gth values and high CW powers, is 8 μm.
Equation (24a) can be simplified by considering the slope efficiency ηsl = ηd (hν/q):
Specific examples of high-CW-WPE devices are: the TA-STR-type 4.9 μm emitting QCL [19] and the NRE-type 4.6 μm emitting QCL [71]. For the TA-STR device the T0 and T1 values were high (i.e. 244 K and 343 K, respectively) while the Rth value (~3.4 K/W, as extracted from CW-curve fitting in the next subsection) was about two times higher than for conventional 5 mm long devices of similar buried-ridge width and diamond submount (i.e. ~1.6 K/W) [64], most likely due to the insertion of seven tall barriers (i.e. AlAs) per period. More specifically, the Gth value is about two times higher for conventional than for TA-STR, 8 μm wide-ridge BH devices (i.e. ~1450 W cm−2 K−1 versus ~735 W cm−2 K−1). The virtual halving of the Gth value is very likely primarily due to significantly more interfaces per period [44, 85] (i.e. 34 versus 22 interfaces) with the rest likely associated with highly strained layers [86]. The increased Rth value gives ΔTact,wpm ~53 K; a rather high value. Using equation (25) and the experimentally obtained pulsed slope efficiency, we calculate a CW ηwp,max value of 20.6%, quite close to the experimentally value of 21%. Similarly, for pulsed operation the calculated value: 26.4%, is in good agreement with the experimental value (27%). For NRE-type 3 mm long, back-facet HR-coated devices the T0 and T1 values were smaller (i.e. 168 and 295 K), due to carrier leakage, but the Rth value was significantly smaller (2.4 K/W) considering that the devices were shorter (i.e. 3 mm versus 5 mm). Then, given a smaller Jwpm value because of lower pulsed Jth (300 K) value (i.e. 0.925 kA cm−2 versus 1.31 kA cm−2), the calculated ΔTact,wpm is also significantly smaller: 20.3 K. Using equation (25) we calculate an ηwp,max value of 12.9%, quite close to the experimentally obtained value of 12.8%. Thus, equation (25) proves to be a good approximation. The main factors that cause TA-STR devices to have much higher ηwp,max value than NRE-type devices of same αm,opt value are: (a) significantly higher slope efficiency (5.72 W/A versus 4.1 WA−1) because of 40% higher value, in significant part due to carrier-leakage suppression; and (b) lower Vwpm value (13.3 V versus 14.6 V).
Then, the obvious question is: Can devices with suppressed carrier leakage be improved to obtain even higher CW ηwp,max values? Clearly, the key is to obtain carrier-leakage suppression at little to no penalty in the Rth value. There is evidence that can be achieved, since deep-well TA devices have demonstrated high T0 and T1 values (i.e. 231 K and ~750 K, respectively) [20] from structures with no additional interfaces compared to conventional QCLs, and of similar number of interfaces per period as conventional QCLs; thus, unlikely to affect the Rth value. Taking the estimated Rth value for conventional 5 mm long BH devices of 8 μm wide buried-ridge width and mounted episide down on diamond (i.e. ~1.6 K/W) and T0 and T1 values of 244 K and 750 K, respectively the ΔTact,wpm value drops to ~23 K and the projected CW ηwp,max value, when J = 3 × Jth (300 K), is ~25%. Of course, with less core heating the ηwp,max value is likely to occur at somewhat higher drive level than 3 × Jth (300 K); thus, ηwp,max values somewhat higher than 25% are achievable (see figure 6(b)). Therefore, carrier-leakage suppression holds the potential to basically double the maximum RT CW wallplug efficiency.
Download figure:
Standard image High-resolution imageFurther increases in wallplug efficiency for 4.5–5.0 μm emitting QCLs can be achieved by either increasing the internal differential efficiency from the current highest reported value of 70% to theoretical limits in the 87–89% range and/or by using designs [87] for significantly lowering the Vwpm value. The former may be realized by reducing IFR-mediated carrier leakage from injector states through the use low-CB-offset injector barriers, just as in the case of the proposed 4.7–4.8 μm emitting STA-SPR QCL (see section 5.3.2). In addition, in the case of no carrier leakage the pulsed L–I curve becomes linear; that is, becomes unity in equation (25). Then, the maximum, single-facet pulsed wallplug efficiency could reach values of ~37%, which, in turn, would result in CW ηwp,max values of ~34.5% (see figure 6(b)). As for significantly reducing the Vwpm value, it has been theoretically and experimentally shown that can be done by using the 'pocket injector' concept [87]. That is, carrier injection into the upper laser level occurring from the first excited state of the injector region, by taking advantage of the fact that, since electrons are hot in the injector miniband, that state is equally populated as the injector ground state [87]. Then, single-facet, CW wallplug efficiencies in excess of 30% may be possible. Ultimately, by increasing both the value to theoretical limits and decreasing the Vwpm value via a pocket-type injector, one may well be able to achieve room-temperature CW ηwp,max values in excess of 40%, when λ = 4.6 μm (see next subsection).
4.3. Output power
The CW optical power can be written as such:
taking ηs(J), the droop in the pulsed L–I curve, only at Th, and with ΔTact being given by:
where
The key parameters for high-CW-power operation are high values for T1 and low values for Rth, just as for interband devices [14]. Similarly, for maximizing the average powers of 10 μm emitting QCLs in quasi-CW operation, Monastyrskyi et al [88] have emphasized the importance of increasing the T1 value and recommended T1 ≈ 2 Th. We have matched the CW L–I curve from [19], by using T0 = 244 K, T1 = 343 K, λ = 4.9 μm and curves fitting the experimental pulsed L–I and V–I curves. Best fit (i.e. 5 W CW output at 1.73 A drive current) occurs for Rth = 3.4 K/W (figure 6(a)). As seen in figure 6(b), the maximum CW and pulsed wallplug-efficiency values match the experimental values: 21% and 27%, respectively. Then, we have run several cases for comparison. If Rth = 8 K/W, as may be the case for episide-up mounting, the CW power at I = 1.73 A drops from 5 W to 3.4 W, and the maximum drops from 21% to 14.5%. If the T0 and T1 values are lowered to 143 K, as in conventional QCLs, while keeping the same Rth value, the CW power at I = 1.73 A drops from 5 W to 3.6 W, and the maximum drops from 21% to 15%. Thus, low T1 values have very similar effects as high Rth values. Increasing the T0 value to 244 K made little difference, as far as the maximum-CW-power and maximum-ηwp,cw values, since, just as in interband devices [14], at high CW drive levels above threshold (⩾ 2.5 Ith) the Jth variation with temperature is not relevant. Next we consider a device with T0 = 244 K, T1 = 750 K and Rth = 1.6 K/W, as may be the case for 4.7 μm emitting STA-SPR devices (section 5.3.2); that is, devices of virtually complete carrier-leakage suppression at no price in Rth value, since no AlAs barrier inserts, throughout the injector region, are needed. Then, the CW power at I = 1.73 A increases from 5 to 6 W, and the maximum increases from 21% to 25.2%. These significant increases reflect the more than halving of the value, at I = 1.5 A, from 53.3 to 23.5 K.
Finally, as mentioned above, the value experimentally obtained in [19] (i.e. 70%) is short of the theoretical limits of 87–89%, achieved only for 8.7–8.8 μm STA-type devices [83] and, as mentioned above, possible for 4.7 μm emitting STA-SPR devices. In addition, also mentioned above, in the case of no carrier leakage becomes unity. We plot the case T0 =244 K, T1 =750 K and Rth = 1.6 K/W with = 87% and , as a limiting case for 4.5–5.0 μm QCLs. Then, the CW power at I = 1.73 A increases from 5 to 8.2 W, and the maximum wallplug-efficiency values are 36.6% pulsed and 34.6% CW. Furthermore, , at I = 1.5 A, drops to 21.2 K (at 6.6 W CW output power). Note that the 34.6% CW value is 65% higher than current record value (21%) at λ = 4.9 μm [19].
Extrapolating to λ = 4.6 μm is nontrivial, but, from the pulsed- prediction curve by Faist [73], an ~8% improvement is expected. Then, STA-SPR devices can reach CW in excess of 27%; that is truly doubling the single-facet values with respect to conventional 4.6 μm devices. Similarly, for devices with = 87%, the pulsed value becomes 39.5% with a respective CW value of ~37%. Furthermore, employing of the pocket-injector concept [87] should result in an ~10% decrease in operational voltage. In turn, the values could reach ~41% and ~43.5% in CW and pulsed operation, respectively. Thus, by using an optimized AR design, the amount of heat to be dissipated would significantly decrease. For instance, at 4.25 W CW output power (i.e. at the point in [19]) the heat to be dissipated would drop from 16 W to 8.4 W; that is, by almost 50%. Taking again the CW point (21%) from [19] as reference (i.e. 4.25 W at 1.5 A current), we can compare the values at the same CW power (4.25 W) for two relevant cases. While for the [19] case at the point is 53.3 K, it drops to ~21 K for the case T0 = 244 K, T1 = 750 K, Rth = 1.6 K/W with = 70%; and to ~16.5 K for the same case with = 87% and . Thus, at ~4 W CW output power becomes ~18 K, a value close to the ~15 K value for the 4.6 μm emitting QCLs that have demonstrated long-term reliability [65].
5. Highly temperature insensitive QCL structures (λ = 3.9–5.0 μm)
Below we discuss all types of QCL devices employing means for carrier-leakage suppression; thus, highly temperature insensitive devices. As a base for comparison, we show in figure 7(a) the conduction band diagram and relevant wavefunctions for a conventional QCL emitting in the 4.6–4.7 μm range [82].
Download figure:
Standard image High-resolution image5.1. Deep-well devices and the effects of strong quantum confinement
Deep-well (DW) devices [10, 18] were the first type of QCLs that demonstrated significant decreases in the temperature sensitivity of the threshold and slope efficiency (i.e. high T0 and T1 values), at heatsink temperatures above RT. By definition, DW QCLs are devices for which the QWs in the active region of each period are lower in energy than the QWs in the injector region. Such devices were also designed for significantly shortening the emission wavelength of GaAs-based QCLs (i.e. from 9.4 to 6.7 μm) [31, 89], at no penalty in room-temperature performance. For carrier-leakage suppression, the conduction-band diagram of a typical DW InP-based DPR-type structure is shown in figure 7(b). The QWs in the active region, being composed of In0.68Ga0.32As, are deeper in energy than the In0.60Ga0.40As QWs in the injector region. The deep wells and associated tall barriers, in the AR, brought about a significant increase in the E54 value: from ~44 meV, typical of conventional QCLs (figure 7(a)), to ~57 meV. As a result, carrier leakage was substantially suppressed which, in turn, led to high T0 and T1 values. T0 is 260 K and 243 K over the 20–60 °C and 60–90 °C ranges in heatsink temperature, respectively, while T1 is 285 K over the 20–90 °C range (figure 8), at no penalty in the RT Jth value [11]. The inset shows that T0 ≈ 315 K over the 80–293 K range in heatsink temperature. T0 values as high as 278 K were also recorded, up to 90 °C [18].
Download figure:
Standard image High-resolution imageThe analysis of DW-QCL behaviour [11, 16] led to the insight that the main carrier-leakage path is from the upper laser level ul through the next higher-energy AR state, ul +1, followed by relaxation to the lower AR states. Thus, it become apparent that further carrier-leakage suppression can be achieved by both maximizing the energy difference between the two levels, , as well as minimizing the wavefunction overlap between the two levels by maximizing the value. In turn, that led to the Tapered-Active (TA) concept [12, 90], that was subsequently applied to DPR QCLs [20]; the so-called 'shallow-well' design [13] which, as shown in [17], 'de facto' constitutes the TA concept applied to an SPR-like structure without deep wells; and to the step-TA (STA) QCL concept [17, 83, 91]. The TA-DPR, TA-SPR and STA QCL designs are discussed below.
Although carrier leakage was suppressed at RT (i.e. 300 K) in DW devices, the value for Jth (300 K) stayed basically the same as for conventional devices. The reason for this behaviour stems from the wavefunctions behaviour for the two device types. As seen from figure 7(a), for conventional QCLs the state-1 wavefunction peaks away from the state-4 wavefunction, such that the τ41 lifetime is about twice as long as each of the τ42 and τ43 lifetimes [17]. Then, the corresponding τ4 value is relatively large: 1.67 ps. In contrast, for DW QCLs the strong quantum confinement brought about by using deep wells and tall barriers causes the state-1 wavefunction to be pulled towards the injection barrier and thus overlap well with the state-4 wavefunction, such that τ41 decreases and becomes comparable to τ43 [17]. Then, the corresponding τ4 value is smaller than in conventional QCLs: 1.50 ps (see the comparison in table 2). As for global τ4 values, penetration of extractor states into the AR causes in conventional devices a τ4g value of 1.34 ps; that is, a 20% decrease compared to the case when considering relaxation to only the lower AR states 3, 2, and 1. For DW devices τ4g =1.21 ps; that is, also a 20% decrease compared to the τ4 value. Finally, calculated τup,g values (table 2) reveal that the DW QCLs have a τup,g value ~11% lower than conventional QCLs.
Table 2. Deep-well versus conventional QCL.
Parameter | DW QCL | Conventional QCL |
---|---|---|
τ4 | 1.50 ps | 1.67 ps |
τ4g | 1.21 ps | 1.34 ps |
τ3g | 0.25 ps | 0.21 ps |
τup,g | 1.11 ps | 1.25 ps |
Jleak / Jth at 300 K | ~8.5% | ~15% |
However, for DW devices, carrier leakage is substantially reduced both because of higher E54 value (57 meV versus 44 meV) and τ54 value (0.32 ps versus 0.22 ps); such that the RT relative leakage-current density value /Jth decreases from ∼ 15% to ∼ 8.5% [11]. In turn, the effect of lower τup,g value is offset by the lower /Jth value and the net result is basically the same Jth (300 K) value for both conventional and DW QCLs [11] (when considering only LO-phonon- assisted transitions). Nevertheless, the substantially higher T0 and T1 values should lead to higher temperatures for CW lasing as well as higher CW power and wallplug efficiencies than when using conventional QCLs.
Another attribute of DW devices is the relative depth of the upper level with respect to the top of the barriers in the AR; that is, the degree of carrier quantum confinement to the upper level. We consider, as a measure the strength of electron confinement to the upper level, the energy difference between the upper level and the top of the second barrier in the AR, E2b,ul, which can be thought of as an effective conduction-band offset ΔEc,eff. There is experimental evidence that the electron–lattice coupling αe−l decreases with increased ΔEc,eff [30, 44]. That means, as can be inferred from equation (6), that for a given Jth value, the larger the ΔEc,eff value is, the closer the electronic temperature in the upper level, Teul, is likely to be to the lattice temperature Tl. For DW QCLs ΔEc,eff ≈ 500 meV, while for conventional QCLs ΔEc ≈ 250–300 meV [55, 82]. In turn, the quantity is likely to be smaller in DW than in conventional QCLs, which may well explain why the predicted T0 value [11], when using the αe-l value measured on low-ΔEc,eff (~250 meV) devices [44, 92], is somewhat lower than the experimental T0 value. Further indication that likely decreases with increased ΔEc,eff comes from comparing lattice-matched to SC 8–9 μm emitting QCLs [37–39, 93]. That is, even though they have basically the same E54 value (i.e. 60 meV) as conventional devices, the high-ΔEc,eff SC devices were found to have higher T0 and T1 values than conventional devices [37, 93]; thus, indicating reduced carrier leakage. That, in turn, was clear indirect evidence of hot electrons in the upper level of conventional 8–9 μm QCLs, which has been recently theoretically found to be the case for 8.5 μm emitting QCLs [47] and agrees with experimental results from 7.6 μm [46] and 9 μm emitting [94] devices indicating that, at around room temperature, is at least ~100 K.
However, due to interface roughness (IFR), the large CB offsets associated with large-ΔEc,eff devices have a significant impact on both the EL linewidth and the upper-state lifetime [50–52, 95]. Thus, the CB offset of 4.8 μm emitting DW devices (~970 meV), being larger than that for conventional 4.6–4.8 μm emitting devices (~800 meV), is likely to have decreased the τup,g value, but that appears to have been compensated for by a lower value. Therefore, even when considering the effects of IFR, the conclusion remains the same: the room-temperature Jth values are basically the same for DW and conventional QCLs. Conversely, introducing low-CB- offset barriers in the AR portion where the lasing transition occurs [13, 95], in order to minimize the IFR influence on device performance, will likely result in large values which, however, can be compensated for by CB-engineered high-E54 devices such as TA-type QCLs [12, 13, 41] since hot electrons in the upper level hardly affect carrier leakage for E54 values ⩾ 80 meV. Cases in point are the TA-like QCLs by Bai et al [19] and Flores et al [41], which have displayed high T1 values (343 K and 550 K, respectively) in spite of low CB offsets (~500 meV) where the lasing transition primarily occurred. Of course, hot electrons in the upper level cause, due to strong coupling to the injector ground state, hot electrons in the injector miniband which, in turn, will increase backfilling (see equation (12a)), especially given that in those cases Δinj had relatively small values (125–135 meV). As a result, the Jth (300 K) values are relatively high [19, 41], in spite of carrier-leakage suppression. Nevertheless, as discussed below, such low-CB-offset TA-type QCLs have achieved high T1 values, and record-high CW output power and wallplug efficiency [19].
5.2. Tapered-active region (TA) devices
The basic concept is to progressively increase the barrier heights in the AR from the injection to the exit barrier (e.g. a linear-taper device, as shown in figure 9(a)), the so-called tapered active-region (TA) design [12, 49, 90]. Specifically, in figure 9(a), the Al concentration x of the AlxInxAs barriers increases linearly, across the AR, from 0.59 to 0.74. As demonstrated in [17], linear tapering causes significant increases in the and values due to: (a) higher energy separation between the excited states of a coupled QW (CQW) pair in asymmetric versus symmetric CQW structures; (b) increased spatial separation between antisymmetric and symmetric energy states with increased asymmetry in CQW structures. The former is somewhat dampened by more Stark shift for asymmetric than for symmetric CQW structures, but that can be alleviated by using step-like tapering in the AR [17]. Below we discuss experimental results from both TA-type devices with deep wells and DPR lower-level depopulation scheme [20, 49], and TA devices without deep wells and SPR lower-level depopulation scheme [13, 19]. Step-taper TA (STA) devices are treated in section 5.3.
Download figure:
Standard image High-resolution imageDownload figure:
Standard image High-resolution image5.2.1. TA double-phonon-resonance (DPR) devices.
Figure 9(b) shows the so-called moderate-taper TA design [17]; that is, a stepwise approximation of the linear-taper design shown in figure 9(a). The design was implemented for the purpose of easing fabrication. The introduced asymmetry causes antisymmetric states 2 and 5 to be 'pulled' to the right which, in turn, results in higher τ4g and τ54 values, respectively, than in DW devices (see table 3). In particular, as seen from table 4, where we compare moderate-taper to conventional QCLs, τ4g recovers to values close to those in conventional QCLs. The net effect is that the effective upper-level lifetime τup,g is pretty much the same for both cases. In addition, due to barrier-height tapering, the E54 value has increases to 67 meV versus 44 meV in conventional QCLs, and the τ54 value is two times larger than in conventional QCLs (table 3). In turn, carrier-leakage suppression is expected to be significantly more pronounced than in DW devices. By using the same approach as in [11] for calculating the leakage-current density, and considering global values for both τ5,leak and τ5,tot lifetimes, we obtain a relative leakage-current density value Jleak/Jth of ∼ 4%, that is, less than half that for DW devices and almost four times less than for conventional devices. Note that the assumption of halving the 300 K τ4g values in order to take into account elastic scattering [11, 52] is still basically valid, since for both TA and conventional QCLs the CB offsets, where the lasing transition occurs, are similar (~800 meV), and τ4 involves diagonal transitions to only one lower energy level. In turn, since τup,g is basically the same, an ∼ 10% decrease in the Jth (300 K) value is expected, primarily due to carrier-leakage suppression. Thus, tapering the barrier heights in the AR significantly suppresses carrier leakage and, unlike in DW devices, it also reduces the Jth (300 K) value.
Table 3. Comparison of various types of QCLs.
Parameter | DW | Moderate taper | Linear taper | Conventional |
---|---|---|---|---|
τ54 | 0.32 ps | 0.44 ps | 0.49 ps | 0.22 ps |
τ4g | 1.21 ps | 1.29 ps | 1.25 ps | 1.34 ps |
τup,g | 1.11 ps | 1.21 ps | 1.19 ps | 1.25 ps |
E54 | 57 meV | 67 meV | 77 meV | 44 meV |
Table 4. Moderate-taper versus conventional QCL.
Parameter | Moderate taper QCL | Conventional QCL |
---|---|---|
τ4g | 1.29 ps | 1.34 ps |
τ3g | 0.21 ps | 0.21 ps |
τup,g | 1.21 ps | 1.25 ps |
Jleak / Jth at 300 K | ~4.0% | ~15% |
Thirty-period QCL devices of active-region design shown in figure 9(b) were grown by MOCVD. Details of the growth, the material characterization, and the structural design are provided in [20] and [96]. Initial devices [49, 97] involved inadvertently heavy-doped injector regions and resulted, due to increased backfilling, in high Jth values (~2 kA cm−2) and relatively low T0 values (~180 K). However, the T1 values were quite high (454 K); thus, confirming that TA QCLs suppress carrier leakage much more than DW QCLs. As seen below, for medium-doped TA devices the T1 values are much higher (~750 K), as expected since the ISB part of αw is a basically proportional with the carrier concentration in the core region and thus with the injector-doping level [78, 80]. A similar behaviour is observed for TA-SPR devices of different doping levels (i.e. the T1 value decreases with increasing the injector-doping level) [13, 19]. In figure 10(a), we show, for medium-doped devices, the light–current (L–I) curves in pulsed operation (100 ns, 2 kHz) for 3 mm long, 19 μm wide ridge, uncoated-facets devices as the heatsink temperature is varied. An inset shows that the devices lased at ∼4.8 μm. Jth (300 K) is 1.58 kA cm−2, a value ∼15% less than for conventional QCLs of same geometry [11]. 10% of that difference is due to carrier-leakage suppression, with the rest of 5% most likely due to lower backfilling current in these TA QCLs versus conventional QCLs, as a result of lower injector-region sheet-doping density in the former (i.e. 0.83 × 1011 cm−2 versus 1011 cm−2).
The carrier-leakage suppression is evidenced by high T0 and T1 values (figure 10(b)). The T0 values are ∼ 230 K, typical of devices with suppressed carrier leakage and moderately high injector-doping level [11, 19], as needed for high-CW-power performance. T1 values as high as 797 K have been measured which, to the best of our knowledge, is the highest T1 value reported to date from any narrow-gain, mid-IR QCL. On average, the T1 values are ∼ 750 K, that is, almost three times larger than for DW QCLs [11, 18] and more than twice the T1 value (i.e. 343 K) for TA QCLs of SPR design and similar injector-doping level [19]. However, the latter is mainly due to a stronger influence, in the case of TA-SPR QCLs, of the temperature dependence of the ISB part of αw on the temperature dependence of , since αm and the FCA-related part of αw are significantly smaller than those for the TA DPR devices reported in [20].
We note, from table 3, that the linear-taper TA QCL has significantly higher E54 and τ54 values than moderate-taper TA QCL, as expected due to increased barrier-height tapering [17]. In turn, further carrier-leakage suppression will occur (Jleak/Jth of ∼ 3%) [49] and higher T0 and T1 values are expected. Furthermore, there is resonant extraction from both lower AR energy states 3 and 2 (i.e. miniband-type extraction [83]), which, in turn, will provide higher ηtr and slope-efficiency values than for moderate-taper TA QCL. However, miniband extraction while keeping a vertical transition (i.e. a z43 value > 15 Å) comes at a price in τup,g; thus, τup stays basically the same as for moderate-taper devices. With respect to conventional QCLs there is almost full recovery of the τup,g value, while achieving substantial carrier-leakage suppression; that is, one is likely to obtain high- T0 and -T1 QCLs with RT thresholds ~10–12% less than conventional QCLs of same injector-doping level as well as significantly higher RT slope efficiencies due to increases in both ηp and ηtr. Next we discuss the application of linear AR-barrier tapering to QCLs of SPR lower-level depopulation scheme and miniband-type extraction.
5.2.2. TA single-phonon-resonance (SPR) devices.
Figure 11(a) shows the conduction-band diagram, calculated at the operating point, for the so-called 'shallow-well' QCL [13]; which is a TA-type QCL, since a shallow barrier (Al0.48In0.52As) (adjacent to a shallow well), a normal-height barrier in the AR centre (Al0.64In0.36As) and an exit barrier composed of Al0.64In0.36As layers sandwiching a thin (6 Å) AlAs layer, form a linear-taper TA-type AR structure [17]. In addition, there are six AlAs layers inserted in the barriers of the extractor/injector region, in order to ensure a large enough minigap to prevent carrier escape from the energy state right above the upper level, state 4, to the upper Γ miniband; thus, to the continuum. Furthermore, we calculate that the six AlAs layers ensure the high value for the energy difference between the upper level and state 4, E43 (i.e. ~85 meV rather than ~69 meV in the case without the AlAs layers) [98]. This 4.9–5.0 μm emitting TA-type QCL was reported [13] as having an AR structure of SPR lower-level depopulation scheme. Actually, it is not the 'classical' SPR scheme involving lower-level depopulation to one energy state and resonant extraction from that state. Rather, there is resonant extraction from both the lower laser level, state 2, and state 1; thus, extraction happens from a miniband [58] similar to the bound-to-continuum (BC) AR design. This miniband-type extraction was recently also employed for 8.7–8.8 μm emitting STA-type QCLs [83], and is a key feature of the 4.5–5.0 μm emitting, STA-type QCLs treated below in section 5.3.
Download figure:
Standard image High-resolution imageAs pointed out by Faist [58], the shallow well and shallow barrier, around where the lasing transition occurs in the AR, cause a significant decrease in the negative influence of IFR on the τup,g value. Furthermore, the effect of sharply reduced IFR scattering appears to compensate for the increase in alloy-disorder scattering [51] associated with stronger upper-level wavefunction penetration in the barriers as well as with the presence of layers of mole-fraction compositions close to 0.50. This is reflected by the fact that when comparing the TA-SPR device to conventional-height barrier devices [58] the former has the highest (inelastic + elastic) τup,g value. In addition, by using values calculated in [58], we note that the ratio of τup,g values considering both inelastic and elastic scattering to τup,g values considering only inelastic scattering is basically the same (i.e. ~0.22) for the TA-SPR device and an SPR device of conventional-height barriers [45]. This means that while for TA-SPR devices the decrease in IFR scattering, by lowering the barriers where the transition occurs, is compensated in part by alloy-disorder scattering, for conventional-barrier SPR devices the higher IFR scattering is compensated in part by lower alloy-disorder scattering.
The structure has strong coupling between the injector ground state, g and the upper level, state 3; in that, the splitting at resonance between those levels is ~14 meV (at a field of 128 kV cm−1) by comparison to typical splitting values of 4–6 meV for QCLs emitting in the 4.0–5.5 μm range. Then, the operating point is generally taken such that the energy difference between the upper level, state 3, and the injector g state, E3g, is about one kT larger than the splitting value [45, 75]. Specifically, we chose the operating point at 87 kV cm−1, since there, with our k ⋅ p code, E3g is 41 meV and Δinj is 123 meV, similar to the 124 meV value Faist considered when calculating this same device [58]. Then, as shown in figure 11(a), the energy difference between the upper level and the next higher energy level, E43, is 85 meV, in agreement with the 90 meV value reported by Razeghi et al [99]. In addition, τ43 has a quite high value: 0.72 ps; ensuring strong carrier-leakage suppression. Thus, this steep linear-taper TA device follows the same trends seen in table 3 for TA-type devices. However, in sharp contrast to those devices, the strong carrier-leakage suppression comes at the price of virtual halving the thermal conductance, due to the presence of the seven additional AlAs barriers in the injector region (see section 4.2).
Table 5 provides key parameters of the device, and is used for comparing to other designs of > 80 meV and miniband-type extraction (i.e. 4.5–5.0 μm emitting STA-type QCLs), discussed below in the next subsection. Since the transition is diagonal, the dipole matrix element, z32, is relatively small (10.6 Å) and the calculated τup,g is relatively large (2.14 ps). By comparison, for conventional [82] and TA-DPR devices [20] vertical transitions give z43 ~15.6 Å and τup,g ~1.25 ps. In addition, the EL linewidth of vertical-transition devices is at least 1.35 times smaller than that of diagonal-transition devices [75]. Considering a conventional DPR device of same geometry [72] we calculate a Jth (300 K) value of ~1.20 kA cm−2. Since the ratio of values between this device and conventional ones [72] is ~1.14, the estimated value of the Jth (300 K) value, minus the difference in leakage-current densities, is ~1.05 kA cm−2. In contrast, for the TA-SPR device the Jth (300 K) value is ~25% higher (i.e. 1.31 kA cm−2). The difference is likely due to lower z32, larger EL linewidth and backfilling current, in spite of higher τup,g value, even when interface roughness and alloy scattering are taken into account [58]. Nonetheless, this fits the classical tradeoff of higher threshold for higher CW power, although such a tradeoff does not necessarily apply for obtaining maximum CW wallplug efficiency [17].
Table 5. Comparison of TA-types of QCLs ( > 80 meV).
Parameter | TA-SPR | STA-DPR | STA-SPR |
---|---|---|---|
E54 | 85 meV | 100 meV | 87 meV |
τ54 | 0.72 ps | 0.84 ps | 0.99 ps |
ztr | 10.6 Å | 15.3 Å | 11 Å |
τup,g | 2.14 ps | 1.15 ps | 1.81 ps |
Δinj | 123 meV | 154 meV | 127 meV |
Gth (W cm−2 K−1) | ~735 | ~1450 | ~1450 |
The carrier-leakage suppression is evidenced by high T0 and T1 values (figure 11(b)). The T0 value is 244 K and the T1 value, measured from the published data, is 343 K. Further proof of strong carrier-leakage suppression is an value of 70% [19], the highest value reported to date from 4.5–5.0 μm emitting QCLs. However, as pointed above in section 3.3.2, the theoretical expected value is ~86%; thus, another carrier-leakage mechanism is likely at work, that decreases the injection efficiency and, in turn, lowers the value. Nevertheless, the above-mentioned attributes led to the achievement of record-high values for maximum pulsed and CW wallplug efficiency (27% and 21%, respectively) and maximum CW power (5.1 W) [19]. The same AR design was used by Metaferia et al [100], when employing a novel method for quick fabrication of BH-type devices, for obtaining 2.4 W single-facet CW power, in a single spatial mode with no beam steering. High T0 and T1 values were also obtained from those devices: 275 K and 442 K, respectively [101]. The somewhat higher values than those for the devices reported in [19] are most likely due to somewhat lower doping level, as evidenced by a lower Jmax value [101]. However, as mentioned above in section 4.2, the seven highly strained layers inserted in the extractor/injector region cause a thermal resistance value more than twice that of conventional 4.5–5.0 μm QCLs, and, in turn, a rather high ΔTact value (~50 K) is obtained at the CW ηwp,max point. Thus, long-term reliable operation, which is well known to be a strong function of ΔTact in QCLs [102], is doubtful. Therefore, for a practical device one needs to implement TA-type or STA-type designs [17] that negligibly affect the device thermal resistance value.
5.3. Step-taper tapered-active region (STA) devices
As a part of a study of TA-type devices, we have shown [17] that step-like tapering of the AR barrier heights provides higher E54 and τ54 values than for linear-taper TA devices, as a result of reduced Stark effect and stronger asymmetry. The STA design can be applied to structures with either DPR of SPR lower-level depopulation schemes and miniband-type extraction.
5.3.1. STA DPR devices.
We have presented [17] an STA-DPR design of high E54 and τ54 values (97 meV and 1.17 ps, respectively) which, in turn, resulted in negligible Jleak/Jth values (~1%) and a projected CW ηwp,max value of 27%, since the Gth value was estimated to be similar to that for conventional QCLs. However, that DPR-type design involves resonant extraction only from state 2. Although, carrier depopulation of state 1 via phonon absorption to state 2 may solve the issue (i.e. a pocket-extractor approach similar to the pocket-injector approach [87]), that is not a certainty. That is the reason why we present below an STA-DPR device with resonant extraction from both levels 2 and 1 as well as from the lower laser level; that is, miniband-type extraction, as in the cases of 4.9–5.0 μm emitting TA-SPR [13, 19] and 8.7–8.8 μm emitting STA-DPR QCLs [83].
The structure of the STA-DPR device with miniband extraction is shown on figure 12. The AR has deep wells and two 'normal' height barriers (Al0.59In0.41As and Al0.65In0.35As) followed by a tall (Al0.93In0.07As) third barrier. The design provides very high values for both E54 and τ54 (100 meV and 0.84 ps), somewhat higher than for TA-SPR devices, but, unlike in that case, without the need of additional AlAs layers inserted throughout the extractor/injector region. Key device parameters are listed in table 5. Due to Stark-effect reduction [17], the E54 value is significantly higher than for the linear TA-DPR device shown in figure 9(a) (i.e. 100 meV versus 77 meV) and the τ54 value is almost twice that for linear TA-DPR devices (i.e. 0.84 ps versus 0.49 ps (table 3)). Then, the calculated Jleak/Jth value is negligible (~1%). However, the constraint of resonant extraction from all three lower AR states, lowers the τup,g value to 1.15 ps; that is, τup,g is ~8% lower than for conventional QCLs. Nonetheless, due to virtual complete carrier-leakage suppression, the Jth (300 K) value is calculated to be ~7% less than that for conventional QCLs.
Download figure:
Standard image High-resolution imageThe transition is basically vertical, given a z43 value of 15.3 Å, basically the same as in conventional and TA-DPR QCLs (15.6 Å); thus, much higher than for TA-SPR QCLs (10.6 Å). The figure of merit; that is, ⋅ τup,g, where ztr is the lasing-transition dipole matrix element, considering only inelastic scattering, is somewhat higher than for TA-SPR QCLs (i.e. 268 Å2 ps versus 239 Å2 ps). When considering interface roughness, the τup,g value of TA-SPR devices will decrease much less than for STA-DPR devices, due to much lower CB offsets in the part of the AR where the lasing transition occurs (i.e. ~500 meV versus ~700 meV). However, the EL linewidth will be significantly narrower in STA-DPR devices, due to much less transition diagonality than in TA-SPR devices. To that effect, when calculating the IFR factor, summed over all transition energies, affecting the EL linewidth (i.e. equation (1) in [50]), we find it to be 59% that for the TA-SPR QCL. In any event, while Jth (300 K) for STA-DPR devices is expected to be ~7% less than in conventional QCLs, we have found above that for TA-SPRs devices Jth (300 K) is ~25% higher than in conventional QCLs; thus, STA-DPR devices are expected to have Jth (300 K) values ~26% less than for TA-SPR devices. The significantly lower Jth (300 K) value is likely to compensate for lower Vwpm values in SPR devices that in DPR devices, due to lower Δinj values (i.e. ~124 meV versus ~150 meV). Thus, the difference may well come down to the much higher Gth values, due to fewer interfaces per period and less strain, for STA-DPR devices compared to TA-SPR devices (table 5). Then, as pointed out in section 4.3, the devices are projected to reach CW ηwp,max values of at least 25%, while the core-temperature rise will be less than half that of TA-SPR devices (i.e. ΔTact ~20 K) at the ηwp,max point. Such devices are expected to be reliable at high (3–4 W) CW powers, since, to date, the only lifetest data showing long-term QCL reliability [65] have been obtained from (low-power) QCLs with ΔTact ~15 K.
It should be pointed out that a 4.8 μm emitting STA-like DPR QCL has been reported [44, 92]; in that, both AR barriers and QWs were step-tapered via inserting submonolayer-thick AlAs and InAs spikes in the barriers and QWs. While the thermal conductivity did improve compared to that of conventional QCLs, due to potential interface broadening [44], the STA-like AR design resulted in carrier-leakage suppression only below 0 °C [18]. That is, the inserted submonolayer-thick layers led only to modest increases in E54 and τ54 compared to conventional QCLs: ~51 meV versus 44 meV, and 0.30 ps versus 0.22 ps [98], because their overall effect was minor as far as increasing the effective CB offsets. In turn, while T0 had a high value of ~240 K below 0 °C, it decreased to 154 K above 0 °C, and T1 was only 143 K above 0 °C [18]. The T1 value, in particular, indicates that, above 0 °C, the device had basically the same high degree of carrier leakage as conventional QCLs [11].
The STA-DPR concept with miniband extraction has been demonstrated for 8.4 μm and 8.8 μm emitting QCLs [83] and resulted in very high T0 and T1 values, for low-Jth 8–9 μm emitting devices, as well as in record-high values (85–90%) for any type of QCL. The results are discussed below in section 6.2.3.
5.3.2. STA SPR devices.
Semtsiv et al [95] have used a conventional 3.8 μm emitting device with tall AR barriers (AlAs) and miniband-type extraction [103] for which they replaced the first barrier within the AR with a relatively shallow barrier (i.e. Al0.48In0.52As) (figure 13(a)). The result was a 3.9 μm emitting device, which, because of much less IFR influence on the EL linewidth and the upper-state lifetime, resulted in significantly reduced Jth (300 K) values. Employing within the AR a shallow barrier followed by two tall barriers is equivalent to using a STA-type structure. In turn, due to strong asymmetry and Stark-effect reduction [17], the E43 value was very high (i.e. ~120 meV). As a result, carrier leakage was suppressed [41] which resulted in a high T0 value of 175 K for 3.8–4.0 μm emitting QCLs, given the relatively high injector sheet-doping density (5 × 1011 cm−2), and, in particular, a high T1 value of 550 K for any type of QCL, albeit below 300 K. More recently Flores [104] found the carrier leakage in such devices to be due to both IFR and LO-phonon scattering from two injector ground states to the AR energy state situated above the upper level. Note that in that case, in equation (9b), ; thus, .
Download figure:
Standard image High-resolution imageFor emission in the 4.5–5.0 μm range, we present in figure 13(b), a STA-SPR device with miniband extraction, just as employed in TA-SPR [13] and STA-DPR devices [83]. The AR has deep wells and two 'normal'-height barriers (Al0.58In0.41As and Al0.62In0.38As) followed by a tall (Al0.93In0.07As) third barrier, which is also the exit barrier. The structure is strongly coupled; that is, 15.4 meV splitting at resonance (133 kV cm−1) between energy levels g and 3; thus, similar to the TA-SPR design [13]. Key parameters are listed in table 5. Step tapering brings about two advantages over the TA-SPR device: (1) the key energy differential E43 has a similar high value (87 meV), but without the need of AlAs additional layers inserted throughout the extractor/injector region; (b) the τ43 value is significantly higher: 0.99 ps versus 0.72 ps, giving an estimated value of 0.98; and, in turn, an ultimate limit of ~87% for the value. Therefore, carrier leakage is virtually completely suppressed at no price in Gth; thus, holding the potential, just like the STA-DPR design, to result in long-term reliable operation at watt-range powers.
The transition is diagonal (z32 = 11 Å) with a correspondingly long, inelastic-scattering upper-state lifetime: 1.82 ps. The figure of merit, only for inelastic scattering, is somewhat smaller than that for TA-SPR QCLs (i.e. 219 Å2 ps versus 239 Å2 ps). However, when calculating the IFR factor, summed over all transition energies, affecting the EL linewidth (i.e. equation (1) in [50]) we find it to be 76% of that for the TA-SPR devices. Thus, a smaller EL linewidth is likely to partly compensate for less τup,g reduction due to higher CB offset(s), where the transition occurs in the AR (i.e. ~700 meV versus ~500 meV). Further compensation for the reduction of the IFR-scattering effect on τup,g (when using low CB offset(s)) comes, as discussed in section 5.2.2, from less alloy-disorder scattering when using conventional-height barriers in the AR (where the lasing transition occurs). All in all, given similar Δinj values, the Jth (300 K) value should basically be the same as for the moderately doped TA-SPR device (~1.3 kA cm−2) [19]. Then, given that the Vwpm value will be similar to that for the TA-SPR QCL and that Gth is likely to be about twice that of the TA-SPR QCL (section 4.2), the devices are very likely to reach RT CW ηwp,max values of at least 25%, with relatively low core-temperature rise (~20 K).
6. Temperature sensitivity of InP-based QCLs emitting in the 8–10 μm range
6.1. Lattice-matched structures
6.1.1. Conventional devices.
Conventionally QCLs emitting in the 8–10 μm range, with low enough Jth values to operate CW at room temperature, have been grown with cores composed of InGaAs/AlInAs superlattices lattice-matched to InP. Typical RT Jth and ηsl/period values have been [50, 105] ~1.7 kA cm−2 and ~25 mW/A, respectively. Due to the smaller lasing-transition energies involved, compared to 4.5–5.5 μm emitting SC QCLs, carriers in the upper level have stronger quantum confinement; thus, resulting in relatively large E54 values (56–60 meV). In turn, above 300 K, typical T0 (160–217 K) [50, 59, 80, 105–109] and T1 values (192–260 K) [50, 110] have been higher than for conventional 4.5–5.5 μm emitting QCLs. (A T0 value of 231 K was reported [111], over the 250–300 K temperature range, from a 9.8 μm QCL, but from short-cavity, high-Jth (3.1 kA cm−2 at 300 K) devices; a similar behaviour to that for short-cavity, high-Jth 4.5–5.5 μm QCLs (section 3.1.2)). Nonetheless carrier leakage can be significant, with the main path being shunt leakage through the next higher energy state above the upper level [27], just like in the case of 4.5–5.5 μm emitting QCLs [11]. A highly sophisticated analysis of 8.4 μm emitting devices [50], taking into account elastic scattering, did fairly accurately predict the T0 value (~160 K), but the predicted ηs values were virtually temperature insensitive, in stark contrast with the experiment. (i.e. T1 ~260 K) (figure 5(b)).
Furthermore, estimates of the electronic temperature were conflicting. On the on hand, analysis based on density-matrix modelling of 8.4 μm emitting devices [81] concluded that electrons are in equilibrium with the lattice; although, as before [50], there was good agreement with experiment only as far as T0, not as far as T1. That indicates that carrier leakage was not taken into account, a fact also inferred from the relatively low derived value (see section 3.2.2). On the other hand, experiments on 7.8 μm [46] and 9.0 μm emitting [94] devices led to estimates of 100–160 K differences between the electronic and lattice temperatures at RT. Recently, as already mentioned above, a comprehensive analysis of 8.5 μm QCLs [47] has found that electrons in the upper laser level can have an electronic temperature Teul ~500 K; that is, ~200 K above the lattice temperature at RT. That analysis combined with recent experimental results from strain-compensated devices (section 6.2) clearly pointing to some degree of carrier-leakage suppression associated with increasing CB offsets, that can only be explained as a consequence of decreasing Teul values, indicate that for 8–9 μm QCLs hot electrons are as much of a reality as for 3.9–5.0 μm QCLs. Furthermore, the fact that the recent implementation of the STA concept for 8.4 and 8.8 μm emitting QCLs [83] has resulted in dramatic increases in both the T0 and T1 values, at little to no penalty in Jth, has established that conventional 8–9 μm emitting QCLs have significant carrier leakage as a result of hot electrons in the upper level.
An extensive analysis by Liu et al [78] of the temperature dependence of optical gain and effective waveguide loss, αw,eff, in 8–10 μm emitting QCLs found that for vertical-transition devices αw,eff does increase with temperature over the 200–300 K range, and attributed the increase to increases in both free-carrier absorption and intersubband absorption. Another part of the increase in αw,eff may well have been due to the backfilling current, as clearly evident from Wittman's study [79] of the temperature dependence of αw and αw,eff for 8.4 μm emitting QCLs over the 240–400 K temperature range (figure 5(a)). The increase with temperature of the αw value, together with carrier leakage explain the relatively low T1 value (i.e. 192 K) [110] for 8.2 μm emitting QCLs [105]. However, for diagonal-transition devices emitting around 10 μm, αw,eff was found to decrease with temperature, a behaviour believed to be due to the complex nature of resonant absorption in those devices.
6.1.2. Unconventional-pumping or -extraction devices.
Very high T0 and T1 values have been reported from 8–9 μm emitting QCLs of unconventional upper-state(s) populating mechanisms [112–114]; however, at a price in Jth and/or ηs. 8.0 μm emitting, indirect-pump scheme devices [112] provided high T0 values: 243 K and 303 K, above RT, from 4 mm and 1.5 mm long chips, respectively, due to suppression of electron populations in the injector regions. However, they had relatively high RT Jth values of 2.7 kA cm−2 and 3.3 kA cm−2, respectively. Dual-upper-state (DAU), broad-gain QCLs [113, 114], have displayed 306–510 K T0 values, mainly because of weak temperature-dependence of the peak gain. However, those occur for RT Jth = 2.1–2.6 kA cm−2, which are relatively high values for 40-period, low-injector-doping devices. The ηs value was found to be virtually temperature independent above RT, with a typical ηs/period value of ~25 mW/A (figure 14). The DAU-design devices thus appear to be quite suitable for low-power (~100 mW), broadband-tuning applications.
Download figure:
Standard image High-resolution imageAn extraction scheme involving an SPR structure with depopulation of the lower level to a miniband, so-called single phonon-continuum (SPC), has been implemented in both 7.9 μm [115] and 8.6 μm emitting [116] devices, for high-yield mass production. Devices at both wavelengths displayed lower-level lifetimes of ~0.2 ps; similar to those in DPR-scheme devices. The former showed a relatively high T0 value (181 K) for moderately high-doped devices, as a result a carrier-leakage suppression (E43 ~75 meV) balanced by relatively high backfilling (Δinj = 130 meV). The latter, being somewhat lower doped, showed, for 33-period uncoated-facets devices, a T0 value of ~200 K, and a high T1 value of ~400 K above RT, which the authors attribute to relatively low carrier-excitation from the upper level to the next higher AR energy state. However, the RT ηs/period value was relatively modest: 22.5 mW/A.
6.2. Strain-compensated structures
6.2.1. Conventional and NRE-type devices.
Strain compensation in 8–10 μm QCLs for the purpose of improving their performance is a relatively new development. In 2010 Leavitt et al [37] compared SC versus lattice-matched 7.7–7.8 μm QCLs and found that, while E54 was basically the same (~60 meV), the strained devices had clearly superior performance; that is, lower CW RT Jth values (1.59 kA cm−2 versus 1.77 kA cm−2), higher T0 values (~222 K versus 194 K), and, most notably, higher CW slope efficiency (0.76 W/A versus 0.56 W/A). The significant slope increase, in particular, is indicative of some reduction in carrier leakage in spite of similar E54 value. Now that is clearly appears that electrons can be quite hot in 8.0–9.0 μm QCLs [47], one likely explanation for such improved behaviour is that the value is reduced in these high CB-offset devices, in agreement with the observation by Spagnolo et al [30] that with increased CB offset the value decreases, since there is stronger coupling between electrons and the lattice. Another factor that may also have contributed to the increased slope efficiency is reduced ISB loss in high CB-offset devices.
Other indications of somewhat reduced carrier leakage have come from SC 7.1 μm [38], 9.0 μm [40] and 10.3–10.7 μm [117] emitting QCLs. For the 7.1 μm device, in spite of a relatively low T0 value (159 K) due to large backfilling, the T1 value was high (411 K); albeit from uncoated-facet devices for which the temperature variation of the low αw value (~1.7 cm−1 at 300 K) had a negligible impact on T1. This low-loss design led to record-high, both-facets pulsed ηwp,max value (18.9%) at 7.1 μm emission wavelength, slightly above theoretically predicted limits when assuming ≈ 70% [73]. However, the value was only ~63.5%; thus, the good match with predicted limits is likely due to much lower Δinj value: 100 meV versus 150 meV [73] as well as to a high ηtr value due to efficient extraction via the triple-phonon-resonance design [118]. The 9.1 μm device also involved a low-loss design of similar E54 value (~60 meV). Similarly, although due to significant backfilling the pulsed Jth value was relatively high (2.1 kA cm−2) (thus, T0 had a conventional-device value (169 K) [93]), an uncoated-facets device had a low αw value (1.6 cm−1) and a high T1 value (331 K) [93]; indicating some carrier-leakage reduction. Just like for 7.1 μm devices, given a high mirror-loss value, the temperature dependence of the low αw value had little impact on the T1 value. The reduction in carrier leakage is also indicated by the fact that the differential pumping efficiency [11] ηp (equations (8) and (17a)) increased from an extracted value of ~62% from 4.6 μm emitting devices [63], considering an ηtr derived from lifetime calculations [58], to 70%. Just as for the above-mentioned SC devices, given the same E54 value, the carrier-leakage reduction may well have been due to a reduction in the value with increasing CB offsets. A record-high, both-facets pulsed ηwp,max value (16%) at 9 μm wavelength was achieved, matching theoretically predicted limits with ≈ 70% [73]. However, the value was only ~58%; thus, just like for the 7.1 μm devices, the good match may be due to much lower Δinj value: 93 meV versus 150 meV [73]. 10.7 μm emitting, uncoated-facets QCLs [117] demonstrated a T1 value of 290 K; which may be indicative of some degree of carrier-leakage suppression. A pulsed ηwp,max value of 10.4% at λ = 10.7 μm was obtained, also matching theoretical predictions when assuming ≈ 70% [73], as expected, given that the extracted value is 64%.
6.2.2. Unconventional-pumping devices.
Fujita et al [39] have reported on 8.7 μm emitting SC QCLs with anticrossed DAU laser states. Just like for the other 8–10 μm SC devices, carrier-leakage suppression occurs even though the energy difference between the upper level and the next higher energy state, E54, is basically the same as for lattice-matched devices: 60 meV. The authors acknowledge that a reduced Teul value may account for carrier-leakage suppression. Due to both optical-absorption quenching in the injector and carrier-leakage suppression, record-high T0 values were achieved: 466 K, to heatsink temperatures as high as 400 K from low-doped devices. However, the Jth (300 K) value is somewhat high (~2.3 kA cm−2) for 3 mm long, HR-coated 40-period devices; thus, leading to a relatively small dynamic range (i.e. lasing up to ~1.5 × Jth). Due to a strong decrease in the αw value with increasing temperature, while carrier leakage is suppressed, the slope efficiency increases with temperature; thus, resulting in negative T1 values: −330 K. Yet, the highest ηs/period value, obtained at 400 K heatsink temperature, is only 0.25 W/A. Both T0 and T1 values are better that for lattice-matched DAU lasers of same E54 value (~60 meV); most likely due to significant carrier-leakage suppression in the SC structures. Such devices appear quite useful for low-power applications such as broadband tuning involving an external cavity.
6.2.3. STA-DPR devices with miniband extraction.
The STA-DPR approach has been implemented for 8.38 μm and 8.8 μm emitting QCLs [83]. The schematic band diagram and relevant wavefunctions are shown in figure 15 for the 8.38 μm device. The design starting point was a published lattice-marched 8.2 μm emitting DPR device [105]. The AR barriers are stepwise tapered, in that the first two barriers are Al0.51In0.49As layers and the third and exit barriers are taller: Al0.58In0.42As layers. In addition, the third and fourth wells in the AR are In0.60Ga0.40As layers, which, having lower conduction-band energy than wells in the injector, are so-called deep wells [10]. These deep wells cause an enhancement of the step tapering of the AR-barrier heights. The results are both a higher E54 value: 75–76 meV versus 57 meV, as well as a significant decrease in the spatial overlap between states 5 and 4, which results in a higher τ54 value: 0.73 ps [83] versus 0.37 ps. Then, we estimate that, due to drastic carrier-leakage suppression, Jleak/Jth is at least four times less than in the conventional QCL [105], and that ηp reaches values as high as ~97%.
Download figure:
Standard image High-resolution imageThe 8.8 μm emitting devices were relatively low doped (~0.7 × 1011 cm−2), while the 8.38 μm emitting devices were moderately high doped (~1.6 × 1011 cm−2). Light–current (L–I) characteristics (200 ns-wide pulses, 100 Hz repetition rate) as a function of heatsink temperature are shown in figure 16(a) for 8.8 μm devices. The L–I curves are superlinear since above ~1.3 × threshold there is resonant extraction from the lower laser level, which, in turn, gives a high ηtr value: ~90% [83]. For the 8.38 μm emitting devices, resonant extraction occurs right from threshold, providing a linear L–I curve (figure 16(b)) of a relatively low Jth value (1.88 kA cm−2) and reaching a maximum single-facet peak power of 3.3 W. The 8.38 μm devices also have a high ηtr value: ~89%. The net result of virtually complete carrier-leakage suppression and high ηtr, due to lower-level resonant extraction, are record-high extracted values for : 85–90%, in good agreement with calculated values (86–87%) [83]. In contrast, the values for conventional 7–10 μm emitting devices are significantly lower: ~63.5% for 7.1 μm NRE SC devices [38]; 58% for 9.0 μm NRE SC devices [40]; and 64% for 10 μm DPR SC devices [117]. The significantly higher achieved values reflect the fact that this STA design takes advantage of both the good properties of the TA design [17] (significant carrier-leakage suppression) and of the miniband design with resonant extraction from the lower laser level [118] (short (~0.1 ps) lower-level lifetimes, as in superlattice-type QCLs [119, 120]).
Download figure:
Standard image High-resolution imageThe drastic carrier-leakage suppression is reflected in very high T0 and T1 values for both low-doped (figure 17(a)) and moderately doped (figure 17(b)) devices. That is, for low-doped devices the average T0 and T1 values are 277 K and 551 K, respectively. By comparison, for similarly doped [110] 3.5 mm long, HR-coated QCLs, emitting at 8.2 μm, the T0 and T1 values were: 217 K and 192 K [110]; and for 3.0 mm long, HR-coated 8.8 μm emitting QCLs [108] the T0 was 200 K. For moderately high doped devices the average T0 and T1 values are 241 K and 272 K, respectively. The decrease in average T0 value is due to larger backfilling with increased doping, yet the value is significantly higher than for lower-doped (~1.2 × 1011 cm−2) 8.4 μm emitting conventional QCLs (i.e. ~160 K) [50]; thus, reflecting the carrier-leakage suppression in STA devices. The decrease in average T1 value reflects the increase in ISB loss with increased doping. Again, the value is superior to the one from lower-doped, 8.4 μm emitting conventional QCLs (i.e. ~260 K) [50]; reflecting the carrier-leakage suppression. Further proof of carrier-leakage suppression is that, in spite of similar αw RT values, the RT slope efficiency (1.15 W/A) is significantly higher than that (0.9 W/A) of same-geometry 8.4 μm emitting QCLs [50].
Download figure:
Standard image High-resolution imageThe achievement of values 30–40% higher than for any type of 7–10 μm-emitting QCLs demonstrates not only the superior performance of STA-type devices, but also that values close to theoretical limits can be achieved, as long as carrier leakage is effectively suppressed and there is fast, miniband-type carrier extraction out of the active region.
7. Temperature sensitivity of InP-based QCLs emitting in the 12–16 μm range
By employing GaAs-based superlattice QCLs [121] RT lasing was achieved beyond 12 μm, but the T0 value was relatively low (124 K), most likely due to carrier leakage to the continuum, just as in the case of DPR, 9.4 μm emitting devices [26]. InP-based, BC-scheme QCLs were the first to demonstrate RT operation at ~16 μm wavelength [122], with relatively high T0 value (234 K), but with a high RT Jth (9 kA cm−2) due to high αw values (~30 cm−1). More recently, by using the IDP scheme, Fujita et al [123] achieved RT operation at ~15 μm wavelength with a relatively low Jth value (3.45 kA cm−2). In addition, the IDP scheme brought about a very high T0 value (450 K) above RT. A relatively high slope efficiency of ~346 mW A−1 was also achieved, almost an order of magnitude higher than in the early work reported in [122], apparently due to both low FCA loss as well as reduced ISB loss in the injectors, as a result of substantial decreases in injector population. However, above ~340 K the T1 value dropped to ~48 K, due to both shunt leakage current through AR-state 5 as well as resonant absorption from upper level 3 to state 5; thus, confirming that T1 is determined by both carrier leakage and an increase in αw with increasing temperature [11].
A remarkable development was the achievement of low-RT-Jth (2 kA cm−2) and high-T0 (306 K) operation from 3 mm, HR-coated ~14 μm emitting QCLs [124]. The devices employ a diagonal transition and a 'two-phonon-continuum' lower-level depopulation scheme. The latter ensures a DPR-like lower-level depopulation scheme for which electrons from an AR state lying one LO-phonon energy below the lower laser level relax to a wide miniband (~70 meV), like in the BC-type devices [58]; thus, ensuring efficient extraction from the AR. The scheme is similar to the SPC depopulation scheme involving a wide (~50 meV) AR miniband for extraction, in 8.6 μm emitting QCLs [116]; resulting in similarly low (~0.2 ps) lower-level lifetimes. The difference though is that, for the same number of periods (70), the SPC device had less-than-half RT Jth values (i.e. 0.97 kA cm−2) due mainly to a vertical transition of high (30 Å) dipole matrix element. Yet, as discussed below, the diagonal transition in the two-phonon-continuum device helps with the carrier-leakage suppression that led to T0 values in excess on 300 K. As shown in figure 18, the T0 value was only 189 K below 240 K heatsink temperature, and then became 306 K up to 390 K heatsink temperature. The behaviour was linked to the value of the energy difference between the lower laser level and the AR state just below it, called El-d. When El-d < 34 meV, the LO-phonon energy, phonon absorption from the lower AR state backfills the lower level, but the lower-level depopulation via phonon emission is very weak; thus, strong backfilling causes a relatively low T0 value. However, when El-d ⩾ 34 meV the lower level is readily depopulated via phonon emission, while backfilling via phonon absorption rapidly degrades; thus, lower-level backfilling diminishes which, in turn, allows for the relatively high T0 value. The same behaviour, as in figure 18, was observed for 4.2 μm emitting, low-voltage-defect QCLs [125] as well as for 8 μm emitting short-injector QCLs [126]. The latter, in particular, showed high T0 values (~250 K) which were believed to be due to improved depopulation of the lower laser states above 240 K. Similarly, Revin et al [60] reported higher T0 values and less backfilling for DPR-scheme 5.5 μm emitting QCLs than from same-wavelength QCLs without resonant LO-phonon assisted depopulation. The likely reasons for the high T0 value (above 240 K) in the 14 μm emitting QCLs are, on the one hand, carrier-leakage suppression, and, on the other hand, highly temperature insensitive lifetimes and EL linewidth due to the transition being dominated by elastic scattering in long-wavelength devices [52]. Carrier leakage is suppressed due to two factors: (a) the upper level is ~100 meV deeper in the QWs than in 8–9 μm emitting devices, which, in turn, reduces the electrons temperature (i.e. small αe–l value) [30]; (b) the diagonal transition ensures poor overlap between the wavefunctions of the upper level and the next higher AR energy state; thus, long values are likely obtained. Then, as seen from equations (6) and (7), given that the Teul value decreases due to low Jth and αe–l values, the scattering rate from state ul to state ul + 1 is minimized which, in turn, significantly reduces carrier leakage. The use of high thermal-conductivity InP cladding layers should allow such devices to reach RT average powers of ~100 mW, just as in the case of 8.6 μm emitting, 70-period SPC devices [116].
Download figure:
Standard image High-resolution image8. QCLs emitting in the 3.0–3.8 μm range
8.1. Temperature sensitivity of InP-based QCLs emitting in the 3.0–3.8 μm range
Initially SPR-BC-type composite-barrier InGaAs/(In)AlAs SC devices of high CB offset [103], similar in design to the first SC QCL [35] that provided 3.4 μm lasing, were used to obtain 3.8 μm lasing. However, RT Jth values were high (~5 kA cm−2) and T0 = 119 K, most likely due to both carrier leakage and backfilling, as the injector sheet-doping density was high (~5 × 1011 cm−2). Later, by using highly strained structures, work on lowering the lasing wavelength of conventional DPR-type devices below 4 μm [65, 127], resulted in strong carrier leakage for high-Jth devices (e.g. T0 and T1 values of 107 K and of 83 K, respectively, above RT) [65]. In addition, episide-down mounted 3.8 μm InGaAs/AlInAs QCLs have been shown to have Gth values ~70% those for 4.3 μm QCLs [86], presumably due to poorer interfaces with increased layers' strain. That was later confirmed, as large EL linewidths (~50 meV) were recorded for 3.76 μm SPR-type devices [69]. As a result, the maximum RT CW power and wallplug efficiency were relatively low: ~150 mW and 1%, respectively, for ridge-guide devices [127] and ~400 mW and 2.7%, respectively, for narrow (3.5 μm) buried-ridge BH-type devices [65]. Further wavelength lowering, by increased layer strain, led to 3.5 μm emitting DPR-type devices [65] of even larger carrier leakage (i.e. T0 and T1 values of 116 and 67 K respectively), possibly due to the onset of carrier leakage to satellite valleys [57]. In turn, the RT CW power was limited to ~20 mW.
High CB-offset (1.6 eV) strain-compensated InGaAs/AlAs(Sb) QCL structures have also been employed for achieving lasing in 3.0–3.8 μm range. The use of AlAs barriers in the AR, instead of AlAsSb barriers, dramatically improved the device performance, apparently due to significantly better interface quality [128]. Up to 2.6 W peak pulsed power per facet and a pulsed ηwp,max value of 10.7% (both facets) were achieved at RT from 3.7 μm emitting, 10 μm aperture devices [128], which are record-high values as far as 3.7–3.8 μm emitting, narrow-ridge QCLs. However, high injector-doping levels (~5 × 1011 cm−2) appear to have resulted in relatively high RT Jth values (3.3 kA cm−2) and low T0 values (108 K), most likely due to significant backfilling. Further work by Commin et al led to 3.3 μm and 3.5 μm emitting devices [129] with RT Jth values of 3.5 kA cm−2 and 2.5 kA cm−2, respectively. T0 values were relatively low: 106 K and 101 K, over the 240–400 K temperature range, respectively, likely due to backfilling and, above ~300 K, to carrier leakage to satellite valleys as well. More recently, a detailed study of 3.5 μm emitting QCLs, using high hydrostatic pressure [57], has identified carrier leakage to the L valley as the likely reason why T0 decreases, above 220 K heatsink temperature, from 264 K to 154 K.
A breakthrough in the performance of ~3.8 μm emitting QCLs occurred with the use of the SPR-type lower-level depopulation scheme coupled with miniband-type extraction [69], quite similar to the AR design later used for record-high CW performance at 4.9 μm [58] and record-high values at 8.7–8.8 μm [83]. The energy difference between the upper laser level, state 3, and the next higher energy level did increase from the typical ~44 meV value for DPR devices to 72 meV [98]. Similarly the scattering time from the higher AR energy state to the upper laser level, τ43, also increased with respect to conventional DPR-type devices: from 0.22 ps to 0.48 ps [98]. In turn, carrier leakage was significantly suppressed as reflected by high T0 values (~200 K). Maximum CW power of ~1 W was achieved at RT; thus, significantly higher than for previous 3.8 μm DPR-type devices, yet about a factor of two lower than from 30-period, HR-coated 4.6 μm DPR-type devices [130]. Similarly, the CW ηwp,max value obtained (6%) was basically a factor of two lower than for 30-period, HR-coated 4.6 μm DPR-type devices [130]. The likely reasons for lower CW performance than for 4.6 μm devices are: (a) the measured Gth value (i.e. 1015 W cm−2 K−1) was ~70% of the value for similar buried-ridge width (8 μm) conventional 4.6 μm devices mounted on diamond submounts [58, 84]; (b) the Jth value was ~15% higher than 4.6 μm devices [130], in spite of carrier-leakage suppression; and (c) the threshold voltage Vth was ~30% higher than 4.6 μm devices (i.e. 12.5 V versus 9.5 V) [130]. The lower Gth value is understandable, given that the EL linewidth was found to be 50 meV compared to ~30 meV for 4.6 μm devices; thus, reflecting poorer interface quality with increased strain [86]. The higher Jth value, in spite of reduced carrier leakage, also mainly reflects significantly higher EL linewidth than for 4.6 μm devices. As for Vth being higher, that reflects the higher-energy transition involved. The net effect is that ΔTact at threshold was about twice that for 4.6 μm devices, which eventually resulted in significantly lower maximum CW power and wallplug efficiency.
Using the same SPR-miniband design, shorter-wavelength (3.4–3.56 μm) devices [131] were obtained by further increasing the strain in the QWs and the barriers. Higher T0 values (152–166 K above RT) than for DPR-type devices [65] were, as expected, obtained. The T1 value was reasonably high (~190 K) for 3.56 μm emitting devices, but it dropped to ~116 K for 3.39 μm emitting devices, most likely due to the onset of leakage to satellite valleys [57]. Similarly, 3.3 μm emitting SPR-BC-type devices [132] have displayed low T1 values (~71 K), over the 250–300 K temperature range, indicating strong carrier leakage to satellite valleys [133]. T0 values were also low (100 K) above 250 K, due to carrier leakage and possibly strong backfilling, given the relatively high injector-doping level. Studies of 3.1 μm emitting InGaAs/AlInAs QCLs [74] did identify the exact location of indirect valleys in strained In0.73Ga0.27As QWs which, in turn, clarified that the reason for the sudden drop in T0 value, above ~125 K, in 3.05 μm emitting QCLs [134] was strong carrier leakage to such valleys. In spite of such high leakage, RT lasing was obtained from the high-CB-offset InGaAs/AlAsSb-based QCLs [135], although with high Jth values (19 kA cm−2) and, as expected, low T0 value (90 K). Even further strain increases in InGaAs/AlInAs structures led to 3.0–3.2 μm emitting QCLs of low RT pulsed Jth values (~2 kA cm−2) [136], with low T0 (102–108 K) and T1 (33–90 K) values, most likely due to drastic increases in carrier leakage. CW operation was obtained, but, due to high voltages and low T0 and T1 values, the output powers were low (3–20 mW), in spite of using narrow (3 μm) buried-ridge BH devices.
8.2. Temperature sensitivity of InAs-based QCLs emitting in the 3.0–3.4 μm range
The InAs/AlSb material system has the highest CB offset (2.1 eV) of all materials used for fabricating QCLs [137]. That has allowed lasing to the shortest wavelength for QCLs (2.6 μm). By reducing thermal backfilling, 3.33 μm RT lasing with Jth and T0 values of 3 kA cm−2 and 175 K (up to 370 K), respectively, were achieved [138]. Such results are comparable to the best obtained from InP-based QCLs, at the same wavelength. The slope efficiency decreased fast with temperature above 300 K (i.e. a T1 value of only ~93 K over the 300–340 K temperature range); thus, indicating significant carrier leakage. That is, very high CB offsets, while preventing carrier leakage to the Γ-band continuum, cannot prevent shunt-type leakage through high-energy states within the AR [11] (see figure 2) and/or leakage to satellite valleys. Devices lasing in the 3.0–3.2 μm range have RT Jth and T0 values of ~3.3 kA cm−2 and 150 K, respectively [137]. The T0 value around RT, in particular, is higher than for similar-wavelength InP-based QCLs, which may reflect cooler electrons in the upper level, due to much higher CB offsets [30]. Carrier leakage is stated to occur primarily through the L valley, although the band diagram for 3.12 μm emitting devices [137] indicates that leakage through the AR energy level right above the upper laser level is likely to occur as well.
8.3. Metamorphic-buffer-layer (MBL)-based QCL structures for lasing in the 3.0–3.8 μm range
As discussed in section 8.1, to accommodate the large transition energies needed for achieving short-wavelength lasing, deeper wells (i.e. high In content) and taller barriers (i.e. high Al content) are necessary, in order to prevent excessive carrier leakage as well as to achieve lower electron temperatures [30]. However, as the layer critical thicknesses for strain relaxation are approached, it is anticipated that device reliability will deteriorate through thermally activated relaxation processes. In addition, highly strained layers lead to increased interfacial roughness and poor thermal conductivity of the superlattice (SL) core-region materials. In fact, highly strained barrier materials (Δa/a ~3%) have been used for short-wavelength (3.0 μm) QCLs on InP [136], although, as pointed out in section 8.1, the performance was severely degraded relative to lower-strained, longer-wavelength (4.6 μm) devices. As the QW In content is increased to shorten the emission wavelength, λ, below 3.5 μm, significant carrier leakage to satellite valleys (L, X) is expected [57, 134] to contribute towards high Jth values as well as reduced T0, T1, and values.
As a means to mitigate strain and open up the parameter design space for short-λ QCLs as well as to avoid leakage to satellite valleys for λ ⩽ 3.5 μm, a metamorphic buffer layer (MBL) may be employed as a virtual substrate of a designed lattice constant [139, 140]. By choosing an MBL lattice constant with a value between that of GaAs and InP, very deep InGaAs/AlInAs QWs with relatively low strain can be achieved. While an MBL is desirable from a design standpoint, significant challenges arise in their practical implementation. The localization of dislocations to the internal interfaces of the MBL serves to substantially reduce threading dislocations in the overlying layers of the MBL and the subsequently grown device structures. Ultra-thick (~20 μm) compositionally graded InxGa1−xAs MBLs on GaAs substrates have been demonstrated using hydride vapour-phase-epitaxy growth [140] with threading-dislocation density, in the top layers of the MBL, in the range 105–106 cm−2, which is considered acceptably low for device applications. In particular, QCLs being unipolar devices, do not suffer from minority-carrier recombination via dislocations. However, the strain relaxation within the MBL gives rise to a surface with tilt and a rough 'cross-hatched' morphology, which can have significant impact on devices employing thin, strained-layer SL materials, such as those used in the QCL core regions. The use of Chemical-Mechanical Polishing (CMP) followed by a wet chemical etching step was found to be effective for improving the surface morphology of compositionally graded InxGa1−xAs MBLs on GaAs substrates, while maintaining an epi-appropriate surface chemical composition and structure [141, 142]. As a result, low-strain In0.37Ga0.63As/Al0.8In0.2As QCL SL core regions have been realized on MBLs with lattice constants intermediate between those of GaAs and InP, and have demonstrated electroluminescence near 3.6 μm [140].
Employing an MBL lattice constant of 0.574 nm, which corresponds to relaxed In0.22Ga0.78As (representing the top surface of the MBL), deep-well (DW) step-tapered-active (STA) QCL designs on MBLs with λ as short as 3.1 μm were reported [140]. In fact, employing such a lattice-constant value allows for the design of low-strain QCL active regions covering the entire 3.0–4.0 μm wavelength range. Figure 19 illustrates a DW STA-QCL design on a virtual substrate for emission at 3.1 μm, and employing a DPR lower-level depopulation scheme. The energy difference between the upper laser level, state 4, and the top of the exit barrier is extremely large (490 meV); thus, much larger than that for conventional short-λ (3.8 μm) QCLs (~225 meV) [127] grown on InP. Moreover, the 3.1 μm emitting STA-QCL design on MBL (figure 19) exhibits a large energy difference between the upper laser level, state 4, and state 5: E54 = 75 meV, and a very long τ54 value (1.80 ps), which will ensure suppression of carrier leakage via state 5. Note that this STA-QCL design can be further optimized, as discussed in section 6.2.3, for miniband-type extraction. The strain differentials between the wells and barriers within the injector regions have relatively small values (~1%) and should result in thermal-conductance Gth values similar to those for conventional 4.6–4.8 μm QCLs. As shown in table 6, the strain × layer-thickness products for the wells and barriers of active and injector regions of STA-type 3.1 μm QCLs are comparable to those for TA-type QCLs emitting at 4.8 μm [20] and much smaller than those for conventional 3.0 μm emitting QCLs [136]. The significantly lower values for the MBL-based QCL designs, compared to those of conventional short-λ QCLs, are expected to result in much improved performance and reliability for 3.0–4.0 μm emitting QCLs. Aside from lower strain, the short-λ MBL-based QCL designs employ wells of low In content (i.e. In0.46Ga0.54As), which is significantly lower than those used in conventional short-λ QCLs on InP. This, in turn, is expected to eliminate carrier leakage to satellite valleys (L, X) within the QWs; thus, further improving the performance of MBL-based QCLs compared to that of a highly strained QCLs grown on InP substrates.
Table 6. Comparison of strain × layer thickness products (Angstroms).
Region | DW-TA (4.8 μm) | DW-STA (3.1 μm) | Conventional (3.0 μm) |
---|---|---|---|
Injector well | −0.103 | −0.103 | −0.445 |
Injector barrier | 0.140 | 0.139 | 0.531 |
Active well | −0.400 | −0.330 | −0.570 |
Active barrier | 0.099 | 0.093 | 0.614 |
Download figure:
Standard image High-resolution imageChallenges remain to realize such QCL devices on MBLs. Typically, 30–40 periods, using the injector + active regions shown in figure 19, would be used in a QCL device. Achieving sufficient planarity of the strained SL core region requires removal, via polishing, of the cross-hatched surface morphology present atop the as-grown MBL. Also, since the MBL lattice constant is not that of InP, ternary materials such as InxGa1−xP, which can be lattice-matched to the top of the MBL, must be utilized for the cladding/waveguide layers in an actual QCL device. That, in turn, will diminish the device Gth value. A careful choice of MBL lattice constant can be used to minimize these effects, allowing for the InGaP-cladding composition to be closer to InP.
9. Conclusions
For state-of-the-art mid-IR QCLs carrier leakage is primarily shunt-type leakage within the active regions (ARs), as a result of thermal excitation of hot electrons from AR upper laser levels and/or injector ground states to high-energy states within the ARs, followed by relaxation to lower AR energy states. Suppression of carrier leakage is key to achieving high CW output powers and wallplug efficiency, and ultimately long-term reliable operation at watt-range power levels.
Multidimensional conduction-band (CB) engineering, for carrier-leakage suppression in the ARs of mid-IR QCLs, has primarily consisted of tapering the barrier heights across the AR, from the injection to the exit barriers. Such tapered-active (TA) devices have provided the current records in CW power and wallplug efficiency (5.1 W and 21% at λ = 4.9 μm) and internal differential efficiency (85–90% at λ = 8.4 μm and 8.8 μm).
For 3.9–5.0 μm emitting QCLs, progressive AR-barriers tapering, from moderate to linear and stepwise, has resulted in virtual complete carrier-leakage suppression, as evidenced by record-high T0 and T1 values: 278 K and 797 K at λ = 4.8 μm, respectively, for the moderately high-doped devices needed for high-power operation. However, the record 21% CW wallplug- efficiency value (λ = 4.9 μm) has come at the price of basically doubling the devices thermal resistance values compared to conventional mid-IR QCLs. Designs are presented for step-taper TA (STA)-type devices that can achieve CW wallplug efficiencies in excess of 25% at λ = 4.9 μm, and in excess of 27% at λ = 4.6 μm, at no expense in thermal resistance. In turn, core-temperature rises at 3–4 W CW powers become relatively low (15–20 °C); thus, allowing for long-term reliable operation. Should means be found to raise the value to theoretical limits (~87%), such as suppressing interface-roughness-assisted carrier leakage, and should operating voltages be reduced via the pocket-injector approach, projected maximum CW and pulsed wallplug efficiencies, at λ = 4.6 μm, are ~41% and 43.5%, respectively.
For 8–10 μm emitting QCLs, STA-type devices have resulted in both carrier-leakage suppression as well as fast, efficient carrier extraction out of the ARs, due to resonant extraction from the lower laser level. In turn, record-high T0 and T1 values: 283 K and 561 K, have been obtained from 8.7–8.8 μm emitting devices, and, due to high (~90%) differential transition efficiencies, the values have, for the first time for QCLs, reached theoretical predicted limits (~87%). Devices based on the dual-upper-state (DAU) AR design have demonstrated T0 values in excess of 300 K and virtually temperature-insensitive slope efficiencies, which makes them attractive for low-power (~100 mW), broadband-tuning applications.
InP-based QCLs emitting in the 3.0–3.8 μm wavelength range suffer from excessive core-region heating, due to high-strain-induced poor interfaces quality and, for λ ⩽ 3.5 μm, from carrier leakage to satellite valleys. Best results have been obtained from SPR-type 3.7 μm devices with miniband-type extraction. A promising solution, for both strain reduction and eliminating leakage to satellite valleys, is the use on metamorphic buffer layers (MBLs) as virtual substrates.
Multidimensional CB engineering for optimizing the CW performance of mid-IR QCLs can be realized by any advanced crystal-growth technique, since it has been demonstrated by using both metal-organic chemical vapour deposition and gas-source molecular-beam epitaxy (MBE) as well as MBE employing multiple effusion cells as sources of the necessary elements.