Current Gain Enhancement at High-Temperature Operation of Triple-Quantum-Well Heterojunction Bipolar Light-Emitting Transistor For Smart Thermal Sensor Application

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

896 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 71, NO.

1, JANUARY 2024

Current Gain Enhancement at High-Temperature


Operation of Triple-Quantum-Well
Heterojunction Bipolar Light-Emitting
Transistor for Smart Thermal
Sensor Application
Mukul Kumar , Graduate Student Member, IEEE, Shu-Yun Ho , Graduate Student Member, IEEE,
Shu-Jui Hsu , Graduate Student Member, IEEE, Pin-Chia Li , Student Member, IEEE,
Shu-Wei Chang , Senior Member, IEEE, and Chao-Hsin Wu , Member, IEEE

Abstract— In this article, we present an investigation HBLETs or TQW-HBTs or TQW-LETs) at varying ambient
into the amplification characteristics of triple-quantum-well temperatures. Our analysis involves a modified
heterojunction bipolar light-emitting transistors (TQW- charge-control model that integrates the idea of thermionic
emission of electrons from the TQWs, which are strate-
gically positioned within the base region of HBTs. Both
Manuscript received 10 August 2023; revised 5 November 2023;
accepted 28 November 2023. Date of publication 19 December 2023; experimental measurements and simulations demonstrate
date of current version 2 January 2024. This work was supported in that the minority carrier within the TQWs acquires greater
part by the Ministry of Science and Technology, Taiwan, under Grant energy at higher operating temperatures, facilitating its
110-2224-E-992-001, Grant 110-2622-8-002-021, Grant 110-2622- rapid escape from the TQWs due to a reduced escape time.
8-A49-008-SB, Grant 111-3114-E-002-001, Grant 111-2119-M-002- Interestingly, this phenomenon leads to a distinct aug-
008, Grant 111-2119-M-002-009, Grant 111-2622-8-002-001, Grant mentation in the current gain, contrary to the conventional
111-2221-E-002-051-MY3, Grant 111-2218-E-002-025, Grant 111- behavior observed in HBTs. Specifically, we observed a
2622-8-002-032, Grant 111-2622-8-002-021, Grant 112-2923-E-002- remarkable increase in current gain of approximately 200%,
008-MY3, Grant 112-2119-M-002-015, and Grant 112-2119-M-002-013; when the operating temperature of TQW-HBTs is elevated
in part by the National Chung-Shan Institute of Science and Technology
under Grant NCSIST-ACOM-111-6712002 and Grant NCSIST-ACOM- from 25 ◦ C to 85 ◦ C. Notably, the experimental findings align
112-6712002; in part by AU Optronics, Hsinchu, Taiwan, under Grant consistently with the outcomes obtained through simula-
110HZT9B001; in part by the National Taiwan University System (NTUS) tions. Consequently, our study presents a compelling case
Innovation Cooperation, Taiwan, under Grant 11112071002; in part for the application of light-emitting transistors (LETs) in the
by National Taiwan University under Grant 112L7860; and in part by design of cutting-edge smart thermal sensors, illustrating
the Center for Quantum Science and Engineering, National Taiwan their promising potential in the front end of such systems.
University. The work of Shu-Wei Chang was supported in part by the
Research Center for Applied Sciences, Academic Sinica, Taiwan; in Index Terms— InGaP/GaAs, light-emitting transistors
part by the Academic Sinica, Taiwan, through the Innovative Materials (LETs), quantum-well single-heterojunction bipolar tran-
and Analytical Technology Exploration (iMATE) Program under Grant sistors (QW-HBTs), temperature, temperature-dependent
AS-iMATE-111-41; and in part by the National Science and Technology current gain, thermal sensor, triple-quantum-well (TQW),
Council, Taiwan, under Grant MOST-108-2221-E-001-018-MY3, Grant TQW heterojunction bipolar light-emitting transistors
MOST-110-2224-E-992-001, and Grant MOST-111-2221-E-001-008.
(TQW-HBLETs), TQW heterojunction bipolar transistors
The review of this article was arranged by Editor Y. Xu. (Corresponding
author: Chao-Hsin Wu.) (TQW-HBTs).
Mukul Kumar and Pin-Chia Li are with the Graduate Institute of Pho-
tonics and Optoelectronics, National Taiwan University, Taipei 10617,
Taiwan (e-mail: mukulkumar.research@gmail.com; r11941081@ I. I NTRODUCTION
ntu.edu.tw).
COMPREHENSIVE investigation has been conducted
Shu-Yun Ho and Shu-Jui Hsu are with the Graduate Institute of Elec-
tronics Engineering, National Taiwan University, Taipei 10617, Taiwan
(e-mail: r11943069@ntu.edu.tw; r11943076@ntu.edu.tw).
A over the past few decades to explore the impact
of temperature on the electrical I –V characteristics of
Shu-Wei Chang is with the Research Center for Applied Sciences,
Academia Sinica, Taipei 11529, Taiwan, and also with the Department InGaP/GaAs heterojunction bipolar transistors (HBTs) [1], [2],
of Photonics, National Chiao Tung University, Hsinchu 30010, Taiwan [3]. In emitter–base heterojunctions, the valance band energy
(e-mail: swchang@sinica.edu.tw). discontinuity 1E V efficiently prevents the back injection of
Chao-Hsin Wu is with the Graduate Institute of Photonics and Opto-
electronics, the Graduate Institute of Electronics Engineering, and the hole currents from the emitter side, resulting in a substantial
Graduate School of Advanced Technology, National Taiwan University, current gain. However, at elevated temperatures, 1E V between
Taipei 10617, Taiwan (e-mail: chaohsinwu@ntu.edu.tw). the emitter and the base decreases, leading to a slight reduction
Color versions of one or more figures in this article are available at
https://doi.org/10.1109/TED.2023.3339084. in the current gain of conventional HBTs due to increased
Digital Object Identifier 10.1109/TED.2023.3339084 thermal energy [4].

0018-9383 © 2023 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See https://www.ieee.org/publications/rights/index.html for more information.
Authorized licensed use limited to: ANNA UNIVERSITY. Downloaded on February 09,2024 at 06:10:49 UTC from IEEE Xplore. Restrictions apply.
KUMAR et al.: CURRENT GAIN ENHANCEMENT AT HIGH-TEMPERATURE OPERATION 897

InGaP/GaAs HBTs commonly encounter power dissipation


issues due to their high currents and relatively low thermal
conductivity of the GaAs substrate, which causes self-heating.
Previous literature indicates that the self-heating effect in
HBTs is primarily governed by the base–emitter junction volt-
age or forward bias voltage [5], [6], [7]. This thermoelectrical
property has found applications in various temperature sensors,
including thermistors, thermocouples, diode thermal sensors,
resistance temperature detectors, and transistor thermal sensors
based on CMOS (with advantages of small area and low
cost) and bipolar junction transistor (BJT) technology [8], [9], Fig. 1. (a) Schematic epitaxial structure of the TQW-HBT. (b) Energy
band diagram of n-p-n InGaP/GaAs TQW-HBLET. (c) Device top
[10]. Commercially, PIN diodes have long been employed as view layout of TQW-HBLET. The emitter cross-sectional area of the
temperature sensors due to their temperature-sensitive forward TQW-HBLET is 40 µm × 40 µm.
bias I –V characteristics, operating within a temperature range
of 1.4–500 K [11]. These sensors exhibit high sensitivity (hun- II. D EVICE FABRICATION AND E XPERIMENTAL R ESULTS
dreds of mV/K), generate high signal levels (ranging from hun-
dreds of millivolts to volts), and maintain exceptional stability In this article, we present our study on the epitaxial
and reproducibility [12]. In addition, silicon-based transistors structures of n-p-n TQW-HBTs, which were fabricated on
excel as a temperature sensor, particularly when high stability, a semi-insulating (S.I.) GaAs substrate through the applica-
low cost, and high sensitivity are required for a specific tion of metal–organic chemical vapor deposition (MOCVD)
temperature range of application (−55 ◦ C to −150 ◦ C) [13]. process. The epitaxial layer stack comprised several layers.
In 2004, Feng et al. [14], [15] presented a ground-breaking It started with a heavily doped n-type GaAs buffer layer
innovation in the field of optoelectronics by introducing with a thickness of 5000 Å, succeeded by bottom cladding
a novel type of heterojunction bipolar light-emitting tran- layers comprising a 634-Å-thick n-type Al0.40 Ga0.60 As layer,
sistors (HBLETs), also known as light-emitting transistors an oxidizable layer of Al0.95 Ga0.05 As with a thickness of
(LETs). This innovative device incorporates a quantum well 5000 Å, and an oxide buffer layer of Al0.40 Ga0.60 As with
(QW) within its base region of HBT, setting it apart from a thickness of 150 Å. The layer sequence continued with a
conventional transistors. This LET not only functions as a 200-Å-thin subcollector layer of heavily doped n-type GaAs,
QW single-HBTs (QW-HBTs) but also serves as an ultrafast followed by a 120-Å etching stop layer of In0.49 Ga0.51 P,
light source with three input ports (one electrical input, one and a 600-Å collector layer of undoped GaAs. The active
electrical output, and one optical output) for optical commu- layer design comprised an 1120-Å GaAs base layer with an
nication purposes, acting as a receiver or transceiver of optical average p-type doping concentration of 3 × 1019 cm−3 , and
data [16]. it included three undoped 70-Å-thin InGaAs QWs (tailored
This novel transistor class exhibits an extremely short for a wavelength λ around 980 nm) positioned side by side
spontaneous recombination lifetime (in the range of tens with a separation of 35 Å. The epitaxial TQW-HBT structure
of picoseconds), allowing for rapid signal modulation at was finalized by growing a 250-Å-wide-gap n-type layer com-
frequencies up to 4.3 GHz by eliminating the slow carri- posed of In0.49 Ga0.51 P, serving as the emitter layer. Moreover,
ers [17]. By incorporating an optical cavity and confining the upper cladding layers replicated the composition of the
layer in the layer structure, the QW-HBT can be further lower ones. Finally, a 1000-Å contact layer of heavily doped
transformed into a transistor laser (TL) [18]. The LETs sen- n-type GaAs was added to cap the epitaxial structure of the
sitivity to ambient temperature can be significantly enhanced TQW-HBTs. Fig. 1(a) illustrates the finished device epitaxial
by implementing specific layer structure designs that induce structure of the TQW-HBTs. During the fabrication process of
strong thermionic emission within the QW [19] or multiple- the TQW-HBT, wet etching was initially utilized to create the
quantum-wells (MQWs) [20] of the HBTs base region. emitter and the base mesas, and then, isolation etching was
Previous studies have primarily concentrated on the opti- carried out until reaching the S.I. GaAs substrate. Following
cal properties of the device, including optical modulation that, standard lithographic techniques and metallization were
and laser thermal characteristics while giving limited con- employed to deposit the contacts of the emitter (E), base (B),
sideration to its electrical characteristics [21], [22], [23]. and collector (C). The fabrication process was completed with
In this research work, we introduce novel triple-quantum- planarization, via hole etching, and pad metallization. The
well HBLETs (TQW-HBLETs or TQW-LETs) or TQW-HBTs, energy band diagram, as depicted in Fig. 1(b), represents a
examining the temperature-dependent current–voltage (I –V ) TQW-HBLET with an n-type In0.49 Ga0.51 P/p+ -type GaAs/n-
characteristics of InGaP/GaAs TQW-HBTs. We conducted type GaAs configuration. In addition, the top view of the
both experimental and theoretical demonstrations showcasing device is presented in Fig. 1(c).
the influence of ambient temperature variations on the cur- To evaluate the device’s performance, it was placed on a
rent gain modulation. This finding emphasizes the potential Peltier temperature-controlled stage. Approximately 20 min
utilization of TQW-HBTs for designing highly sensitive and were required for the Peltier stage to achieve a steady-state
temperature-dependent current–voltage (I –V ) characteristics temperature before stable temperature measurements were
for future smart thermal sensors’ front-end components. taken. An Agilent E5270B was utilized to supply the electrical

Authorized licensed use limited to: ANNA UNIVERSITY. Downloaded on February 09,2024 at 06:10:49 UTC from IEEE Xplore. Restrictions apply.
898 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 71, NO. 1, JANUARY 2024

Fig. 2. (a) Experimental characteristics: collector current IC versus


collector-to-voltage VCE of n-p-n TQW-HBT at varied base currents IB
from 0.2 to 1 mA, investigated at 25 ◦ C (solid line) and 85 ◦ C (dashed
line). (b) Experimental analysis: current gain (βTQW ) of TQW-HBT as a
function of externally controlled substrate temperature Text .

DC bias and current. During the parameter sweep, the base


current I B was systematically varied from 0.2 to 1 mA in Fig. 3. Schematic of minority charge distribution (ρ) in the base region
0.20 mA increments, while increasing the collector–emitter of the TQW-HBT.
voltage VCE from 0 to 2 V. The TQW-HBTs collector current
IC versus collector-to-emitter voltage VCE characteristics were Applying reverse bias to the base–collector junction creates
measured for different bias currents I B at substrate temper- a high electric field that drives electrons into the collector,
atures Text of 25 ◦ C and 85 ◦ C, as depicted in Fig. 2(a). resulting in a near-zero electron density at the collector. The
When biased at I B = 1 mA and VCE = 2 V, the collector basic charge-control model of LET was previously reported
current IC increased from 0.210 to 0.631 mA at Text = 25 ◦ C by Feng et al. [24]. In addition, our research team conducted
and Text = 85 ◦ C, respectively, in the forward-active region. further modifications to the charge control of HBLET and
This unique temperature-dependent collector current is a key investigated its thermal characteristics in 2019 [25]. This
parameter in designing a smart thermal sensor based on the research work focuses on analyzing the behavior of elec-
IC (T) characteristics for temperature sensing. Moreover, it was trons surrounding the TQWs and the temperature-dependent
observed that under the same bias conditions of I B and VCE , behaviors of current gain in TQW-HBT. Consequently, we pri-
the collector-to-emitter offset voltage decreased from 0.64 to oritize analytical formulations that explicitly account for the
0.56 V at Text = 25 ◦ C and Text = 85 ◦ C, respectively, indicat- time-dependent parameters in our research.
ing the presence of intrinsic carrier surges as the temperature Fig. 3 presents the charge control model for the distribution
increased. of minority charges (ρ) within the base region of TQW-HBT.
Fig. 2(b) illustrates the unique thermally enhanced current The base charge distribution of TQW-HBT can be modeled as
gain of TQW-HBT, displaying the relationship βTQW and a combination of four triangular charge populations referred
controlled temperature Text for different base currents I B to as Q 0 , Q 1 , Q 2 , and Q 3 . Within the base, the total volume
ranging from 25 ◦ C to 85 ◦ C. The current gain of the LETs charge density of minority carriers comprises seven compo-
increased with higher temperature Text or base current I B . The nents: Q 0 , Q 1 , Q 2 , Q 3 , Q QW1 , Q QW2 , and Q QW3 . Among
QW conduction band offset played a crucial role in confining these, Q 1 , Q 2 , and Q 3 account for the electrons that diffuse
the electron, causing those injected into the base region to from the emitter to QW1 , QW2 , and QW3 , respectively, for
be localized or attracted around the QWs. Higher tempera- spontaneous or stimulated recombination. Q 0 maintains the
tures led to increased thermal energy in the QWs reservoir, diffusion of electrons toward the B–C electrical collector,
enabling electrons in the base to acquire the necessary energy thereby forming the collector current (IC ). In addition, Q QW1 ,
to overcome the quantum well barrier (QWB) and escape Q QW2 , and Q QW3 are associated with the capture of minority
from the QW reservoir. As a result, a considerable number carriers by QW1 , QW2 , and QW3 , respectively.
of minority carriers rejoined the diffusion flow toward the Expanding upon our prior derivation for the current gain
collector, reducing the effect of carrier capture into the QW of MQW-HBLETs, we extend our investigation to assess the
at elevated temperatures. This led to an enhanced current impact of temperature on the current gain of TQW-HBT
gain under reverse bias, as more electrons were collected by using (18) as referenced in [26]. In this analysis, we delib-
the base–collector junction. In addition, the higher current erately omit consideration of quantum-well coupling and
gain βTQW observed with an increased base current I B might quantum-tunneling current, given their relatively minor influ-
potentially be linked to the thermal effect, as the injection of ence on current gain in contrast to the significant role played
current could elevate the internal temperature of the device. by thermionic emission. The current gain of TQW-HBTs,
denoted by βTQW , can be determined using [26, eq. (18)], with
III. M ODIFIED C HARGE -C ONTROL M ODEL N set to 3, resulting in the following expression:
A. Triple-Quantum-Well Base Region Current Gain in
βEC
HBTs βTQW ≈ . (1)
1 + (1 + βEC )(V1 + V2 + V3 )
QW-HBTs exhibit distinct behavior compared with separate
confinement heterojunction (SCH) due to the tilting of the Here, βEC = τb /τt,EC represents the current gain corresponding
minority carrier distribution during forward-active operation. to fictitious HBTs with the base width W B . τb denotes the

Authorized licensed use limited to: ANNA UNIVERSITY. Downloaded on February 09,2024 at 06:10:49 UTC from IEEE Xplore. Restrictions apply.
KUMAR et al.: CURRENT GAIN ENHANCEMENT AT HIGH-TEMPERATURE OPERATION 899

bulk recombination lifetime of electrons in the base, and τt,EC B. Modified Thermionic Emission Model
represents the transit time of electrons through the GaAs base The investigation into the escape time of electrons from
region of the transistor. A notable observation is that increasing the QWs was carried out using the classical thermionic
the number of QWs in the base region of HBTs results in theory by Schneider and Klitzing [27] and Nagarajan [28].
a reduced current gain, attributed to the accelerated carrier According to this model, the distribution of carriers in the
recombination within the TQWs. Alternatively, if the TQWs QW follows Boltzmann-like statistics, and escape currents
are absent (V1 +V2 +V3 = 0), the current gain simply becomes are related to the carrier distribution tail located at the top
βEC of the HBT with the same base width W B . In addition, of the energy barrier. However, this method for estimating
V1 = F1 /F0 , V2 = F2 /F0 , and V3 = F3 /F0 are calculated the escape time is limited to QW widths of a few hundred
using the relation Vl = F l /F0 from [26, eq. (15)], where Angstroms. In our experimental device, we have three QWs,
Fl is obtained from [26, eq. (13)]. In Fig. 3, the terms F0 , each with a width of 70 Å. To address this, we utilize a mod-
F1 , F2 , and F3 represent the charge flux of minority carriers ified expression for the thermionic emission lifetime proposed
corresponding to Q 0 , Q 1 , Q 2 , and Q 3 , respectively. by Nelson et al. [29], which takes into account thermionic
With [26, eq. (13)], we may represent the terms F1 , F2 , and emission, thermally assisted tunneling, and direct tunneling
F3 as follows: in the QW. In our study, we employ a comparable approach,
as described in [29], to estimate the escape time of electrons.
F1 ∼
= U1C F0 (2a) The first step involves calculating the injection current density
∼ U2C F0 + U21 F1
F2 = (2b) (Je ) by integrating the product of carrier concentration in the
= U3C F0 + U31 F1 + U32 F2 .
F3 ∼ (2c) QW, transmission probability, and carrier escape velocity over
the energy range. Next, we integrate the carrier population
Now, with [26, eq. (12)], the above terms U1C , U2C , U3C , U21 , present in the confined states of different sub-bands and the
U31 , and U32 are defined as follows: unconfined states on the top QW. Finally, we evaluated the
thermionic emission lifetime using the following equation:
1
1 + β1E1
!
τb τcap,1 1 d1 ne d
U1C =
τQW,1 τ 1 + τ 1 1 β W1C
 (3a) τesc
te
= . (4)
QW,1 esc,1
1 + βEC 1C 2 Je
1
1 + β1E2
!
τb τcap,2 1 d2 In thermionic emission lifetime calculations, the classical
U2C = (3b) Maxwell–Boltzmann distribution is selected for its simplicity,
τQW,2 τ 1 + τ 1 1 β W2C

1 + βEC 2C 2
QW,2 esc,2 making it suitable for analyzing macroscopic systems at ele-
1
1 + β1E3
!
τb τcap,3 1 d3 vated temperatures. However, it is important to acknowledge
U3C = (3c) that its accuracy may decrease as quantum effects become
τQW,3 τ 1
+ τesc,3 1 + βEC β3C W23C
1 1

QW,3 more pronounced. As an alternative approach, we employ the
1
1 + β1E2
!
τb τcap,2 1 d2 classical Boltzmann distribution instead of the Fermi–Dirac
U21 = (3d) distribution to reformulate the expression for thermionic emis-
τQW,2 τ 1
+ τesc,2 1 + β E1 β21 W221
1 1

QW,2
sion lifetime, as detailed in (4), resulting in the following
1
1 + β1E3
!
τb τcap,3 1 d3 analytical expression:
U31 = (3e)
τQW,3 τ 1 + τ 1 1 β W31

1 + 31  Nb ( E −E )
β E1 2π m ∗ 2m ∗ k B T −1 X

QW,3 esc,3 2 c,b i
1 τesc =
te
e kB T
1 + β1E3
!
τb τcap,3 1 d3 ℏ ℏ2
U32 = . (3f) i=1
τQW,3 τ 1
+τ 1
1+ β 1 β32 W232 m ∗ d 2m ∗ 1E B −1/2
 
QW,3 esc,3 E2
+
ℏ ℏ2
Here, we consider several key parameters related to the
N
πℏ X π ℏd 2m 1E B 1/2
b    ∗ 
recombination and carrier dynamics in the TQW-HBT. The E c,b − E i
≈ exp +
recombination lifetime of electrons in each QW is denoted k B T i=1 kB T 2π 1E B ℏ2
by τQW,1 , τQW,2 , and τQW,3 for QW1 , QW2 , and QW3 ,
(5)
respectively. The density of minority carriers in the virtual
conduction states of each QW is represented as n VS,1 , n VS,2 , where the effective barrier height E B is determined by the
and n VS,3 for QW1 , QW2 , and QW3 , respectively. Furthermore, difference in band edge energies between the conduction band
the capture time of minority carriers transitioning from their in the GaAs barrier (E c,b ) and the first sub-band of the InGaAs
respective virtual states to the bound states in each QW QW (E 1 ). The width of the QW is denoted by d, while
is denoted as τcap,1 , τcap,2 , and τcap,3 for QW1 , QW2 , and m ∗ represents the effective mass of electrons. The Boltzmann
QW3 , respectively. In addition, τesc,1 , τesc,2 , and τesc,3 represent constant is represented as k B , and T denotes the temperature
the escape time of electrons from QW1 , QW2 , and QW3 , in kelvins. [ℏ2 /2m ∗ 1E B ]1/2 is the length scale determined
respectively, into the doped base region. It is important to by 1E B . Nb represents the number of bound sub-bands, and
note that the width of each QW in our fabricated TQW-HBTs E i denotes the band edge energy of the ith bound sub-
device is kept the same (d1 = d2 = d3 = d). The remaining band. In this modified thermionic emission model, we also
terms, such as β E3 , β E2 , β E1 , β32 , β31 , β3C , β2C , β1C , and β21 , consider the effect of the carrier-longitudinal optical (LO)
are calculated using the formulation proved in [26, eq. (10)]. phonon interaction due to polar scattering on the escape rate.

Authorized licensed use limited to: ANNA UNIVERSITY. Downloaded on February 09,2024 at 06:10:49 UTC from IEEE Xplore. Restrictions apply.
900 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 71, NO. 1, JANUARY 2024

Fig. 4. Temperature-dependent escape times for different base currents from the QW in TQW-HBT. (a) Escape time from QW3 (τesc,3 ). (b) Escape
time from QW2 (τesc,2 ). (c) Escape time from QW1 (τesc,1 ). (d) Capture time τcap versus temperature T for various base currents ranging from 0.2 to
1 mA. (e) Recombination lifetime τQW versus temperature T for different base currents ranging from 0.2 to 1 mA. Note: (d) and (e) assume almost
equal capture (τcap,3 ≈ τcap,2 ≈ τcap,1 ) and recombination (τQW,3 ≈ τQW,2 ≈ τQW,1 ) times in all the QWs. (f) Base transit time (τt,EC , black solid line)
and bulk recombination lifetime (τb , red solid line) as a function of temperature T. The inset of (a)–(c) illustrates the position of sub-band energy in
the TQW.

At elevated temperatures, electrons can readily escape from the temperature increase, the capture time substantially increases
QW through the absorption of an LO phonon. To account for because carriers transitioning from bulk states to the quantum
this effect, we include the additional temperature-dependent well states become blocked at the accessible states. In addition,
term [exp(ℏωLO /k B T ) − 1]−1 in the escape rate, known as the the capture time is also influenced by the operating current–
Bose–Einstein distribution function [30], [31]. Consequently, voltage. With an increase in the base current, the scattering rate
the escape time due to thermionic emission can be expressed decreases, leading to an increase in capture time due to the
as follows: increased Fermi wavevector, as depicted in Fig. 4(d). We also
 −1 extracted the QW lifetime as a function of temperature at
1 1 1 different base currents, as shown in Fig. 4(e). In comparison to
τesc ≈  te +  . (6)
τesc τLO exp ℏωLO − 1 the three timescales, the QW recombination time and capture
 
kB T time exhibit less sensitivity to temperature changes than the
Here, τLO represents the electron escape time from the bound escape time in the QW. Furthermore, the escape time τesc
states to the unconfined state caused by the absorption of LO varies more significantly than the bulk recombination time τb
phonons, and ℏωLO is the energy of LO phonons. and transit time τt,EC as the temperature increases.
We extracted the capture time (τcap ), escape time (τesc ),
and recombination time (τQW ) as functions of temperature C. Recombination and Transit Time of Carriers for
at different base currents I B by fitting the experimental cur- TQW-HBTs
rent gain to the theoretical current gain (βTQW ) from (1) Carrier recombination and transit time play crucial roles
for TQW-HBT. These results are illustrated in Fig. 4(a)–(e). in the performance of TQW-HBTs. The base region of
As the device’s electron temperature rises, the extensive carrier TQW-HBTs is heavily doped with a p-type dopant concen-
distribution over the QW is impacted by the band filling tration of approximately 3 × 1019 cm−3 . As a consequence,
p
effect. As a result, the escape time decreases exponentially the minority carrier mobility µe is primarily influenced by
with increasing temperature and base current, as depicted in ionized impurities. However, its influence on carrier scatter-
Fig. 4(a)–(c). The escape time differs for QW1 , QW2 , and ing becomes negligible. The influence of lattice scattering,
QW3 due to their distinct first sub-band energy level in the attributed to phonon scattering, is proportional to the tem-
QW, resulting from the specific layer structure of TQW-HBT. perature. Therefore, the minority carrier mobility decreases
We have also incorporated self-heating effects, which indicate with increasing temperatures, albeit with less influence at
that the junction temperature in our TQW-HBTs is lower than higher temperatures [32], [33]. To analyze the base transit
that in conventional HBT. However, as the base current and time τt,EC as a function of temperature, we employ Einstein’s

Authorized licensed use limited to: ANNA UNIVERSITY. Downloaded on February 09,2024 at 06:10:49 UTC from IEEE Xplore. Restrictions apply.
KUMAR et al.: CURRENT GAIN ENHANCEMENT AT HIGH-TEMPERATURE OPERATION 901

p
relationship (Dn = µe k B T /q) and consider the temperature
p
dependence of minority carrier mobility µe . In Fig. 4(f) (black
solid line), we observe that the transit time reduces by 23.25%
as the temperature rises from 300 to 400 K, indicating a weak
p
dependency of minority carrier mobility µe on temperatures.
Next, we investigate the changes in the bulk recombination
time within a heavily p-doped base with respect to tem-
perature T, as illustrated in Fig. 4(f) (red solid line). The
lifetime can be described through the models of Shockley–
Read–Hall (SRH) recombination, radiative recombination, and
Auger recombination, outlined as follows:
 −1
1 1 1 −1
τrec = + + = A0 + B0 P + C0 P 2 .
τSRH τrad τAuger
(7)
Here, P denotes the doping concentration in the base region;
A0 , B0 , and C0 are the SRH, spontaneous, and auger recom- Fig. 5. Theoretical simulation of the temperature-dependent ratios
based on [26, eqs. (22) and (23)]. (a) Ratio of QQW3 −Q0 , (b) QQW2 −Q0 ,
bination coefficient, respectively; τSRH , τrad , and τAuger are the (c) QQW1 − Q0 , and (d) Q(QW3 +QW2 +QW1 ) − Q0 . The total electron stored
SRH, radiative, and Auger recombination times respectively. in the TQWs is denoted by Q(QW3 +QW2 +QW1 ) .
In a heavily doped GaAs semiconductor, recombination caused
by the Auger is a significant issue due to the high impurity
density and carrier density. The threshold kinetic energy E th
of high-energy electrons in the nondegenerate semiconduc-
tor is significantly influenced by temperature, especially in
the context of band-to-band Auger recombination. At higher
temperatures, the energy bandgap E g shrinks, leading to a
decrease in the threshold kinetic energy E th of high-energy
electrons and potentially increasing the Auger recombination
coefficient C0 [34]. On the other hand, at room temperature,
the recombination process is mainly governed by band-to-
band radiative recombination of the direct band gap GaAs,
rather than the exciton process. The lifetime of radiative
recombination τrad is linked to the spontaneous emission
coefficient B0 , which is proportionally related to T −3/2 for
the bulk materials with parabolic bands; and T −1 for quantum
wells with separated bands [35], [36]. In general, the SRH
recombination time τSRH is significantly longer than that of Fig. 6. Temperature dependence of current gain βTQW in TQW-HBT
for varied base currents (0.2–1 mA): experimental (star solid) and
radiative and Auger recombination. As B0 is treated as a simulation (solid line) results.
constant in our temperature-dependent charge-control model,
the bulk lifetime τb varies by approximately 41.55% when because it is closer to the emitter side than other QWs.
the temperature T rises from 300 to 400 K, as depicted in As the temperature increases, electrons gain sufficient energy
Fig. 4(f) (red solid line). These findings suggest that auger to escape more readily from the QW, and these additional
recombination dominates over bulk recombination lifetime at escaping electrons contribute to the base charge region Q 0 .
higher temperatures. Our theoretical simulation results are shown in Fig. 5(a)–(d)
using the same simulation parameters as shown in Fig. 6.
D. Charge Analysis in the Quantum Well and Base These simulations demonstrate that the ratio of Q QW3 –Q 0 ;
Region Q QW2 –Q 0 ; Q QW1 –Q 0 ; and Q (QW3 +QW2 +QW1 ) –Q 0 , all decrease
with increasing temperature and base current. Consequently,
The existing formula used for the theoretical calcula-
at higher temperature T, a larger number of carriers escape
tion of temperature-dependent charge analysis, as presented
from the QW, moving toward the collector, which results in an
in [26, eqs. (22) and (23)], seems to underestimate the charge
amplified current gain. The amplified current gain is attributed
analysis in the TQWs and base region. The ratio of total charge
to the higher electron escape rate, leading to a consequential
stored in the TQWs to Q 0 defined a follows:
impact on the base charge region Q 0 .
Q (QW3 +QW2 +QW1 ) Q QW3 Q QW2 Q QW1
= + + . (8)
Q0 Q0 Q0 Q0 IV. S IMULATION R ESULTS AND D ISCUSSION
In our fabricated TQW-HBT, QW3 captures a higher num- In our fabricated TQW-HBT device, the three QWs could be
ber of electrons compared with QW2 and QW1 , primarily considered as a bucket that captures electrons transmitted from

Authorized licensed use limited to: ANNA UNIVERSITY. Downloaded on February 09,2024 at 06:10:49 UTC from IEEE Xplore. Restrictions apply.
902 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 71, NO. 1, JANUARY 2024

TABLE I sensing when integrated into optoelectronic integrated circuits


A NALYZING THE C URRENT G AIN (βTQW ) AT 300 K W ITH (OEICs) in the near future.
THE F OLLOWING PARAMETERS

V. C ONCLUSION
In conclusion, this research work successfully demonstrates,
both experimentally and through simulations, that the cur-
rent gain of TQW-HBTs exhibits an increase with operating
substrate temperature. The incorporation of InGaAs triple-
quantum-wells into the transistor base region renders the tran-
sistor’s current gain temperature sensitive. Notably, the current
gain is remarkably enhanced by approximately 200% by
designing a specific TQW-HBTs layer structure and increasing
the temperature from 25 ◦ C to 85 ◦ C. Our analysis thoroughly
considers the impact of TQWs and provides an explanation
for the phenomenon of thermoelectric enhancement of current
gain using our developed temperature-dependent modified
the emitter to the collector side. As the temperature increases, charge-control model based on thermionic emission theory.
the escape time τesc diminishes exponentially with 1/k B T , The theoretical calculations align well with the experimental
causing the electrons stored in the TQW to gain sufficient trend of current gain in TQW-HBTs, validating the accuracy
energy for escaping. Consequently, at higher temperatures, of our approach. This TQW-HBT is a strong candidate for
these escaping electrons contribute to the base charge Q 0 . designing a high-resolution temperature sensor as a front-end
Under the influence of B–C reverse bias, they are swept to the component due to its strong high current-to-temperature signal.
collector, leading to an increase in both the collector current Finally, this work sheds light on the potential of TQW-HBTs as
IC = AQ0 /τt,EC and the current gain βTQW . This results in promising devices with enhanced thermal characteristics and
a notable temperature variation in the current gain and the lays the foundation for further advancements in the field of
collector current IC . Fig. 6 displays the experimental current OEICs and temperature sensing technologies.
gain as a function of temperature T, represented by family
curves at VBC = 0, and compares it with the corresponding
simulated results at various base currents. The simulation R EFERENCES
trends of temperature-dependent current gain for TQW-HBTs [1] W. Liu, S.-K. Fan, T. Henderson, and D. Davito, “Temperature
dependences of current gains in GaInP/GaAs and AlGaAs/GaAs het-
at a base current of I B = 0.2 mA closely matches the exper- erojunction bipolar transistors,” IEEE Trans. Electron Devices, vol. 40,
imental data. Table I presents the corresponding parameters no. 7, pp. 1351–1353, Jul. 1993, doi: 10.1109/16.216446.
used in the simulation at a temperature of 300 K. [2] E. S. Yang, Y.-F. Yang, C.-C. Hsu, H.-J. Ou, and H. B. Lo, “Tem-
Moreover, we made efforts to fit the experimental measure- perature dependence of current gain of GaInP/GaAs heterojunction
and heterostructure-emitter bipolar transistors,” IEEE Trans. Electron
ment data at various base currents, not limited to 0.2 mA, Devices, vol. 46, no. 2, pp. 320–323, Feb. 1999, doi: 10.1109/16.740896.
by adjusting specific parameters that linearly change with [3] D. A. Ahmari, G. Raghavan, Q. J. Hartmann, M. L. Hattendorf,
base current and temperature. However, due to the internal M. Feng, and G. E. Stillman, “Temperature dependence of InGaP/GaAs
heterojunction bipolar transistor DC and small-signal behavior,” IEEE
heat-up of the device at higher base currents, a more metic- Trans. Electron Devices, vol. 46, no. 4, pp. 634–640, Apr. 1999, doi:
ulous experimental calibration was necessary to achieve a 10.1109/16.753694.
theoretical fit with the experimental current gain. Nonetheless, [4] E. S. Yang, C. C. Hsu, H. B. Lo, and Y.-F. Yang, “Modeling of current
gain’s temperature dependence in heterostructure-emitter bipolar tran-
the simulation results were still satisfactory. Notably, as the sistors,” IEEE Trans. Electron Devices, vol. 47, no. 7, pp. 1315–1319,
electrons thermionically escape from the TQWs, the current Jul. 2000, doi: 10.1109/16.848270.
gain at I B = 1 mA increases by approximately 200% with the [5] W. Liu and A. Khatibzadeh, “The collapse of current gain in multi-finger
heterojunction bipolar transistors: Its substrate temperature dependence,
temperature rising from 25 ◦ C to 85 ◦ C. This indicates that instability criteria, and modeling,” IEEE Trans. Electron Devices,
at higher temperature T, more carriers are released from the vol. 41, no. 10, pp. 1698–1707, Oct. 1994, doi: 10.1109/16.324577.
QWs and reach the collector, ultimately increasing the current [6] W. Liu, H.-F. Chau, and E. Beam, “Thermal properties and ther-
mal instabilities of InP-based heterojunction bipolar transistors,” IEEE
gain. The sensitivity of TQW-HBTs is defined as the change in Trans. Electron Devices, vol. 43, no. 3, pp. 388–395, Mar. 1996, doi:
the collector current with respect to temperature (1I C /1T ), 10.1109/16.485651.
which distinguishes it from the sensitivity (change in voltage [7] S. P. Marsh, “Direct extraction technique to derive the junction tem-
perature of HBT’s under high self-heating bias conditions,” IEEE
with respect to temperature, 1V /1T ) typically defined in Trans. Electron Devices, vol. 47, no. 2, pp. 288–291, Feb. 2000, doi:
conventional BJT or PIN diode-based thermal sensors. In our 10.1109/16.822269.
future work, we are planning to design more complex circuits [8] A. Bakker, “CMOS smart temperature sensors—An overview,”
in Proc. IEEE Sensors, vol. 2, Jun. 2002, pp. 1423–1427, doi:
incorporating passive components. This will enable us to 10.1109/ICSENS.2002.1037330.
redefine the sensitivity as 1V /1T , allowing for a more [9] K. A. A. Makinwa, “Smart temperature sensors in standard
meaningful comparison with existing technology. Notably, CMOS,” Proc. Eng., vol. 5, pp. 930–939, Jan. 2010, doi:
the average current-to-temperature signal of TQW-HBLET 10.1016/j.proeng.2010.09.262.
[10] K. Souri, Y. Chae, Y. Ponomarev, and K. A. A. Makinwa, “A precision
is 7 µA/◦ C. This substantial current-to-temperature signal DTMOST-based temperature sensor,” in Proc. ESSCIRC, Sep. 2011,
holds great promise of achieving high-resolution temperature pp. 279–282, doi: 10.1109/ESSCIRC.2011.6044961.

Authorized licensed use limited to: ANNA UNIVERSITY. Downloaded on February 09,2024 at 06:10:49 UTC from IEEE Xplore. Restrictions apply.
KUMAR et al.: CURRENT GAIN ENHANCEMENT AT HIGH-TEMPERATURE OPERATION 903

[11] I. M. Dmitrenko, S. P. Logvinenko, N. I. Ivanov, and Z. M. Kolot, “Ther- [24] M. Feng, N. Holonyak, H. W. Then, and G. Walter, “Charge control
mometric characteristics of semiconductor diodes,” Cryogenics, vol. 6, analysis of transistor laser operation,” Appl. Phys. Lett., vol. 91, no. 5,
no. 6, pp. 357–358, Dec. 1966, doi: 10.1016/0011-2275(66)90137-8. Jul. 2007, Art. no. 053501, doi: 10.1063/1.2767172.
[12] J. K. Krause, and B. C. Dodrill, “Thermometry employing gal- [25] Y.-H. Chang, Y.-L. Chou, S.-W. Chang, and C.-H. Wu, “Thermally-
lium aluminum arsenide diode sensor,” U.S. Patent 4 643 589, enhanced current gain of quantum-well heterojunction bipolar transis-
Feb. 17, 1987. tor,” J. Appl. Phys., vol. 126, no. 1, Jul. 2019, Art. no. 014503, doi:
[13] G. C. M. Meijer, “Thermal sensors based on transistors,” Sens. 10.1063/1.5091050.
Actuators, vol. 10, nos. 1–2, pp. 103–125, Sep. 1986, doi: 10.1016/0250- [26] M. Kumar, L.-C. Hsueh, S.-W. Cheng, S.-W. Chang, and C.-H. Wu,
6874(86)80037-3. “Analytical modeling of current gain in multiple-quantum-well het-
[14] M. Feng, N. Holonyak, and W. Hafez, “Light-emitting transistor: erojunction bipolar light-emitting transistors,” IEEE Trans. Electron
Light emission from InGaP/GaAs heterojunction bipolar transistors,” Devices, early access, 2023, doi: 10.1109/TED.2023.3289930.
Appl. Phys. Lett., vol. 84, no. 1, pp. 151–153, Jan. 2004, doi: [27] H. Schneider and K. V. Klitzing, “Thermionic emission and Gaussian
10.1063/1.1637950. transport of holes in a GaAs/Alx Ga1−x As multiple-quantum-well struc-
[15] M. Feng, N. Holonyak, and R. Chan, “Quantum-well-base heterojunc- ture,” Phys. Rev. B, Condens. Matter, vol. 38, no. 9, pp. 6160–6165,
tion bipolar light-emitting transistor,” Appl. Phys. Lett., vol. 84, no. 11, Sep. 1988, doi: 10.1103/physrevb.38.6160.
pp. 1952–1954, Mar. 2004, doi: 10.1063/1.1669071. [28] R. Nagarajan, “Carrier transport effects in quantum well lasers:
[16] G. Walter, C. H. Wu, H. W. Then, M. Feng, and N. Holonyak, “4.3 GHz An overview,” Opt. Quantum Electron., vol. 26, no. 7, pp. S647–S666,
optical bandwidth light emitting transistor,” Appl. Phys. Lett., vol. 94, Jul. 1994, doi: 10.1007/bf00326653.
no. 24, Jun. 2009, Art. no. 241101, doi: 10.1063/1.3153146. [29] J. Nelson, M. Paxman, K. W. J. Barnham, J. S. Roberts, and
[17] C. H. Wu, G. Walter, H. W. Then, M. Feng, and N. Holonyak, “Scal- C. Button, “Steady-state carrier escape from single quantum wells,”
ing of light emitting transistor for multigigahertz optical bandwidth,” IEEE J. Quantum Electron., vol. 29, no. 6, pp. 1460–1468, Jun. 1993,
Appl. Phys. Lett., vol. 94, no. 17, Apr. 2009, Art. no. 171101, doi: doi: 10.1109/3.234396.
10.1063/1.3126642. [30] C.-Y. Tsai, L. F. Eastman, Y.-H. Lo, and C.-Y. Tsai, “Breakdown
[18] M. Feng, N. Holonyak, G. Walter, and R. Chan, “Room temperature of thermionic emission theory for quantum wells,” Appl. Phys. Lett.,
continuous wave operation of a heterojunction bipolar transistor laser,” vol. 65, no. 4, pp. 469–471, Jul. 1994, doi: 10.1063/1.112339.
Appl. Phys. Lett., vol. 87, no. 13, Sep. 2005, Art. no. 131103, doi: [31] C.-Y. Tsai, C.-Y. Tsai, Y.-H. Lo, R. M. Spencer, and L. F. Eastman,
10.1063/1.2058213. “Nonlinear gain coefficients in semiconductor quantum-well lasers:
[19] L.-C. Hsueh, H.-Y. Lin, M. Kumar, and C.-H. Wu, “Thermal Effects of carrier diffusion, capture, and escape,” IEEE J. Sel. Top-
dynamic performance and integrated optoelectronic system with ics Quantum Electron., vol. 1, no. 2, pp. 316–330, Jun. 1995, doi:
InGaP/GaAs quantum well light-emitting transistors (LETs),” in Proc. 10.1109/2944.401211.
Asia Commun. Photon. Conf. (ACP), Oct. 2021, pp. 1–3, doi: [32] S. Tiwari and S. L. Wright, “Material properties of p-type GaAs at large
10.1364/ACPC.2021.M5D.2. dopings,” Appl. Phys. Lett., vol. 56, no. 6, pp. 563–565, Feb. 1990, doi:
[20] M. Kumar, L.-C. Hsueh, S.-W. Cheng, S.-Y. Ho, S.-J. Hsu, 10.1063/1.102745.
and C.-H. Wu, “Effect of quantum-well number on the current [33] M. L. Lovejoy, M. R. Melloch, and M. S. Lundstrom, “Temperature
gain of heterojunction bipolar light-emitting transistors,” in Proc. dependence of minority and majority carrier mobilities in degener-
Conf. Lasers Electro-Optics (CLEO), May 2023, pp. 1–2, doi: ately doped GaAs,” Appl. Phys. Lett., vol. 67, no. 8, pp. 1101–1103,
10.1364/cleo_at.2023.jw2a.43. Aug. 1995, doi: 10.1063/1.114974.
[21] M. Feng, N. Holonyak, and A. James, “Temperature dependence of a [34] P. T. Landsberg, “The band-band Auger effect in semiconductors,”
high-performance narrow-stripe (1 µm) single quantum-well transistor Solid State Electron, vol. 30, no. 11, pp. 1107–1115, Nov. 1987, doi:
laser,” Appl. Phys. Lett., vol. 98, no. 5, Jan. 2011, Art. no. 051107, doi: 10.1016/0038-1101(87)90074-8.
10.1063/1.3528206. [35] G. W. ’T. Hooft, M. R. Leys, and H. J. Talen-v.d. Mheen, “Tem-
[22] F. Tan, R. Bambery, M. Feng, and N. Holonyak, “Transistor laser perature dependence of the radiative recombination coefficient in
with simultaneous electrical and optical output at 20 and 40 Gb/s GaAs-(Al,Ga)As quantum wells,” Superlattices Microstruct, vol. 1,
data rate modulation,” Appl. Phys. Lett., vol. 99, no. 6, Aug. 2011, no. 4, pp. 307–310, Jan. 1985, doi: 10.1016/0749-6036(85)90092-8.
Art. no. 061105, doi: 10.1063/1.3622110. [36] Y. Arakawa, H. Sakaki, M. Nishioka, J. Yoshino, and T. Kamiya,
[23] W. Huo et al., “Fabrication and characterization of deep ridge “Recombination lifetime of carriers in GaAs-GaAlAs quantum wells
InGaAsP/InP light emitting transistors,” Opt. Exp., vol. 22, no. 2, near room temperature,” Appl. Phys. Lett., vol. 46, no. 5, pp. 519–521,
pp. 1806–1814, Jan. 2014, doi: 10.1364/oe.22.001806. Mar. 1985, doi: 10.1063/1.95578.

Authorized licensed use limited to: ANNA UNIVERSITY. Downloaded on February 09,2024 at 06:10:49 UTC from IEEE Xplore. Restrictions apply.

You might also like