KimSoomin1
ChoSeongjae*
-
(Department of Electronic and Electrical Engineering, Ewha Womans University, 52 Ewhayeodae-gil,
Seodaemun-gu, Seoul 03760, Korea)
Copyright © The Institute of Electronics and Information Engineers(IEIE)
Index Terms
1T DRAM, 3C-SiC, quantum well, sensing margin, data retention, high temperature
I. Introduction
Dynamic random-access memory (DRAM) is a crucial product in the global semiconductor
industry. A conventional DRAM is composed in the one-transistor and one-capacitor
(1T1C) configuration. As CMOS device technology pursues scaling to the nanoscale regime,
DRAM has faced several limitations. Most of the limitations of DRAM stem from constraints
on the scaling and integration process of capacitors [1,2]. To address this problem, capacitorless DRAM, or 1T DRAM, is proposed. Since 1T DRAM
can overcome the limitations caused by capacitor, it is one of the most promising
candidates for more embedded-system-oriented DRAM [3-9]. On the other hand, relatively short retention time has been considered as the weakness
of 1T DRAM, compared with 64 ms, an industry standard [10]. The short retention stems from the annihilation of stored charges by either recombination
or fast diffusion. While the capacitor in the conventional DRAM structure is designed
to have a high enough capacitance near 30 fF/cell [11], it is hard to reach a comparably large capacitance in a 1T DRAM in which the effective
capacitance is defined usually by the floating body. Although novel device structures
including silicon-on-insulator (SOI) have been introduced [12-15], issues in retention and self-heating problem make it difficult to replace the conventional
DRAM. However, 1T DRAM is a solution to realize a high-speed memory embedded in the
processing unit through a batch integrated processing, for higher cost effectiveness
and product yield. Also, as applications in the power electronics and automotive platform
are rapidly increasing, memory component technologies based on SiC need to be developed
in pace. Recent studies also show the advances in SiC memories [16-20]. Among the several phases of SiC, 3C-SiC has a high thermal conductivity, higher
critical breakdown field, and a wider energy bandgap compared with Si [21,22]. Owing to these advantages, 3C-SiC has been used as an ideal material for high-performance
power devices. Furthermore, the study on 3C-SiC is essential in the field of memory
devices since it has the potential to overcome the limitations of conventional silicon-based
memory technologies. In this work, a 1T DRAM made with 3C-SiC/Si/3C-SiC for the source/channel/drain
has been designed and characterized. A quantum-well (QW) structure formed by 3C-SiC/Si/3C-SiC
enhances data retention by effectively confining the mobile charges. Also, a double-gate
structure and a vertical channel were applied to improve the program/erase operation
efficiencies and to achieve higher integration density. An effective confinement of
holes in the hold state was confirmed. Also, quantitative evaluations of sensing margin
and retention time of the proposed 1T DRAM were carried out at elevated temperatures
up to 500 K.
II. Device Structure
Fig. 1(a) shows a schematic of the proposed 1T DRAM with a 3C-SiC/Si/3C-SiC QW. The dependencies
of memory operation characteristics were investigated by varying channel length (L$_{\mathrm{ch}}$)
and channel thickness (T$_{\mathrm{ch}}$) by using HfO$_{2}$ as the gate oxide material
with a thickness (t$_{\mathrm{ox}}$) of 3 nm. As shown in Fig. 1(b), the difference in the energy-band diagrams between the all-Si and the 3C-SiC/Si/3C-SiC
QW-embedded channels are confirmed. The large valence-band offset (VBO) in the QW
1T DRAM occurs because of the bandgap of 3C-SiC, 2.2 eV, much larger than that of
Si, 1.12 eV. Due to this VBO, holes are effectively confined in the QW, resulting
in a significant elongation of retention time. Also, an enhanced gate controllability
is facilitated by the double-gate structure, enabling faster program and erase operations
at smaller voltages. p$^{+}$ poly-Si gate is used, and the doping concentrations for
the drain, channel, and source junctions are n-type 10$^{20}$ cm$^{-3}$, p-type 10$^{17}$
cm$^{-3}$, and n-type 10$^{20}$ cm$^{-3}$, respectively. The band-to-band tunneling
mechanism dominates during a write-1 operation to transfer electrons from the source
valence band into the channel, avoiding device degradation by eliminating the need
for repeated impact ionizations. For a write-0 operation, electrons are injected into
the QW body and recombine with the holes [23]. Energy bandgap, effective masses of electrons and holes, electron affinity, electron
and hole mobilities, and the effective density of states in the conduction band and
valence band for 3C-SiC used in these simulations are 2.2 eV, 0.6m$_{0}$, 0.68m$_{0}$,
3.83 eV, 40 cm$^{2}$/V·s, 1,000 cm$^{2}$/V·s, 1.559${\times}$10$^{19}$ cm$^{-3}$,
and 1.559${\times}$10$^{19}$ cm$^{-3}$, respectively [24,25].
Fig. 1. Device structure: (a) Schematic of the 1T DRAM; (b) Comparison of energy-bands
for all-Si and the proposed cells.
III. Simulation Results and Discussion
1. Operation Voltage Scheme
To find the optimal operation voltage conditions for the proposed structure, technology
computer-aided design (TCAD) simulations were conducted with L$_{\mathrm{ch}}$ of
100 nm and T$_{\mathrm{ch}}$ of 20 nm. Fig. 2(a) and (b) illustrate the hole concentrations in the QW body after the write-1 and write-0
operations, respectively, at various operation voltages. As briefly mentioned, the
write-1 operation is performed by band-to-band tunneling and the conditions for the
write-1 and other operations are summarized in Table 1:
Table 1. Voltage Schemes for Respective Memory Operations
|
VGS [V]
|
VDS [V]
|
Time [ns]
|
Write 1
|
-1.0
|
-3.0
|
10
|
Write 0
|
1.0
|
0.5
|
20
|
Read
|
0.3
|
0.3
|
10
|
Hold
|
-0.6
|
0.0
|
-
|
Fig. 2. Hole concentrations in the QW after: (a) write-1; (b) write-0 operations under
the different operation voltages.
condition 1: (V$_{\mathrm{GS}}$, V$_{\mathrm{DS}}$) = (-1.0 V, -2.0 V), condition
2: (V$_{\mathrm{GS}}$, V$_{\mathrm{DS}}$) = (-2.0 V, -2.0 V), condition 3: (V$_{\mathrm{GS}}$,
V$_{\mathrm{DS}}$) = (-1.0 V, -3.0 V), and condition 4: (V$_{\mathrm{GS}}$, V$_{\mathrm{DS}}$)
= (-2.0 V, -3.0 V) were simulated for write-1 operation. Under the conditions 1 and
2 in Fig. 2(a), holes did not accumulate in the QW body. On the other hand, under the conditions
3 and 4, holes were accumulated effectively through band-to-band tunneling. Among
the simulated conditions, condition 4 shows a slightly higher hole concentration by
a write-1 operation than condition 3. However, when a hold operation starts, the excess
holes eventually escape from the QW under the both conditions, keeping the remaining
hole concentration similar. Consequently, condition 3 can be chosen for a low-voltage
operation. The write-0 operation relies primarily on electron-hole recombination within
the floating QW body, and the deep QW constructed by the large VBO provides the predominance
to recombination rather than drift and diffusion out of the QW. Thus, a write-0 operation
exhibits less voltage dependency, unlike 1T DRAMs with all-Si or alternate material
configurations, requiring a sufficient time to erase the holes as can be predicted
by Fig. 1(b). When performing a read operation, it is necessary to ensure that the stored holes
are not affected. In actual operation schemes for the 1T DRAM, the read operation
should have as less effect as possible on the memory state for minimizing the use
of refresh operation. Various voltage conditions were evaluated for optimizing the
read scheme: condition 1: (V$_{\mathrm{GS}}$, V$_{\mathrm{DS}}$) = (0.3 V, 0.3V),
condition 2: (V$_{\mathrm{GS}}$, V$_{\mathrm{DS}}$) = (0.3 V, 0.5 V), condition 3:
(V$_{\mathrm{GS}}$, V$_{\mathrm{DS}}$) = (0.3 V, 1.0 V), condition 4: (V$_{\mathrm{GS}}$,
V$_{\mathrm{DS}}$) = (0.5 V, 0.5 V), and condition 5: (V$_{\mathrm{GS}}$, V$_{\mathrm{DS}}$)
= (0.5 V, 1.0 V). Among these conditions, condition 1 exhibited the smallest change
in hole concentration after a read operation, thus showing the least disruptive behavior.
When a negative V$_{\mathrm{GS}}$ is applied for hold operation, the depth of QW for
hole confinement goes deeper, which makes it hard for the holes to escape from the
QW and results in a longer retention time. In this work, a negative V$_{\mathrm{GS}}$
of -0.6 V was applied for conducting hold operation. By controlling V$_{\mathrm{GS}}$
only, no energy is consumed for hold operation. Fig. 3(a) shows the distribution of hole concentration in the hold state after a write-1 operation.
The relative locations with reference to T$_{\mathrm{ch}}$ = 20 nm and L$_{\mathrm{ch}}$
= 100 nm are depicted in the horizontal and vertical axes, respectively. A high hole
concentration of 10$^{20}$ cm$^{-3}$ is observed beneath the gates on both sides of
the channel and the hole concentration at the channel center is as relatively low
as 10$^{17}$ cm$^{-3}$. When a write-0 operation is performed, electrons are injected
from the source junction into the body so that the number of holes beneath the gate
is momentarily reduced down to 10$^{17}$ cm$^{-3}$ level, as shown in Fig. 3(b). Table 1 summarizes the voltage schemes for the respective memory operations of the proposed
1T DRAM.
Fig. 3. Hole distributions in the QW body after: (a) write-1; (b) write-0 operations.
2. Dimension Optimization
To explore how the sensing margin and retention time are influenced by the scaling
of device dimensions, L$_{\mathrm{ch}}$ of the 1T DRAM cell was adjusted to 100 nm,
70 nm, 50 nm, and 30 nm, while keeping T$_{\mathrm{ch}}$ constant at 20 nm. Fig. 4(a) through (d) shows sensing margin and data retention at different channel lengths
of 100 nm, 70 nm, 50 nm, and 30 nm, respectively. Here, retention time was defined
as the elapsed time until the initial difference in read currents between states 1
and 0 is reduced to half so that the original data can no longer be correctly identified.
As L$_{\mathrm{ch}}$ gets shorter, the sensing margin increases, but the amount of
increase is not large. On the other hand, the retention time drastically decreases
with scaling down. As scaling progresses, the electron potential energy of the QW
decreases, and consequently, the read current of state 1 increases and the effective
depth of QW is reduced, making it easier to lose the confined holes, which shortens
the retention time. Nevertheless, with a sufficiently long L$_{\mathrm{ch}}$, the
1T DRAM cell can maintain hole storage for over 1 second, demonstrating a robust retention
capability at a L$_{\mathrm{ch}}$ of 100 nm. Fixing L$_{\mathrm{ch}}$ at 100 nm, the
effects of T$_{\mathrm{ch}}$ on the sensing margin and retention time are examined
in Fig. 5. The retention time is primarily influenced by the VBO, and thus, remains without
a notable change. However, the sensing margin is determined by the amount of current,
or the number of mobile carriers that make up the read current for state 1, the sensing
margin is rapidly reduced. In the following simulations, both L$_{\mathrm{ch}}$ and
T$_{\mathrm{ch}}$were fixed at 100 nm and 20 nm, respectively. In Fig. 4(a) through (d) and Fig. 5, retention time is defined as the moment of time at which the initial difference
in read currents of state 1 and 0 is reduced to the half value. In the actual array
operation, the retention time can be determined by a reference current in an absolute
value considering the capability of the sensing circuits, rather than reading by ratio.
The sensing margin threatened by the high integration density should be considered
in many aspects. Narrowing of sensing margin becomes inevitable as the array density
increases, or equivalently, as the number of wordlines (WLs) increases. The critical
factors determining the sensing margin in a high-density DRAM array might vary depending
on types of sensing schemes. If the conventional voltage sensing scheme is applied,
the ratio of cell capacitance to bitline (BL) capacitance can be a critical factor.
The cell capacitance of the heterojunction 1T DRAM cell in this work is determined
by the change in the amount of charge stored and released by program and erase operations.
Thus, the design of QW needs to be further optimized. If the read operation is based
on the current sensing scheme, one way of obtaining a clear memory window comes with
the minimization of the line resistances. As the portion of potential drop is allocated
to the series resistance in the array interconnect with more weight, the memory window
genuinely determined by a DRAM cell is buried by the effect of line resistance and
gets narrower. The QW volume should be increased for accommodating a larger number
of holes for increasing the effective cell capacitance, which can be a smaller technological
hurdle since the footprint is not sacrificed with the help of vertical structure of
the proposed DRAM cell, which can be a substantial solution in the cell level approach.
Fig. 4. Sensing margin and data retention over time for different Lch's: (a) 100 nm; (b) 70 nm; (c) 50 nm; (d) 30 nm (Tch = 20 nm).
Fig. 5. Sensing margin and retention time as a function of channel thickness at a
fixed Lch of 100 nm.
3. High-temperature Operations
As mentioned earlier, a key motivation for this study is to explore the applicability
of SiC/Si/SiC QW 1T DRAM in the environments that require high-temperature stability,
such as in automotive and power integrated circuits. This is supported by the high
melting point and strong breakdown field of 3C-SiC, along with the deep QW created
by the large VBO across the 3C-SiC/Si heterojunction. Fig. 6 depicts the transient simulation results over one cycle of the 1T DRAM operations
at different temperatures, including 300 K, 358 K, 400 K, and 500 K: write-1/hold/read-1/hold/write-0/hold/read-0/hold.
All the currents from write-1, read-1, write-0, and read-0 operations increase as
the operating temperature is elevated, and the rate of increase in read-0 current
is significantly faster than the rate of increase in read-1 current. Consequently,
the large I$_{\mathrm{on}}$/I$_{\mathrm{off}}$ ratio of about 10$^{10}$ at 300 K decreases
as temperature increases. At a high temperature, a sharp increase in read-0, or off-state
current, is commonly observed. However, the SiC/Si/SiC QW structure successfully limits
the read-0 current, even at 500 K, by the precise control over the critical dimensions.
As shown in Fig. 7(a) through (d), higher temperatures lead to an increase in the rate of generation and
more diffusion of stored holes out of the QW, leading to faster deterioration of retention
time and I$_{\mathrm{on}}$/I$_{\mathrm{off}}$ current ratio, accompanying increase
in leakage current. However, the deep QW formed by 3C-SiC/Si/3C-SiC heterojunction
suppresses the escape of stored holes to source and drain junctions, even at high
temperatures. In the QW 1T DRAM, memory state breaking gets faster at a high temperature
mainly due to the accelerated diffusion of the carriers occupying the higher energy
states in the Fermi-Dirac distribution tail. However, with the 3C-SiC/Si/3C-SiC QW,
the programmed holes are more effectively confined and less affected by the temperature
elevation. At temperatures above 300 K, even a small gate voltage of -0.6 V can effectively
constrain the holes within the QW, resulting in a longer retention time. At 300 K,
the proposed device exhibits approximately 30 times wider sensing margin and about
3 times longer retention time compared to Si/Si$_{\mathrm{0.6}}$Ge$_{\mathrm{0.4}}$/Si
QW 1T DRAM. At 358 K, the device has more than 10 times larger sensing margin compared
with an all-Si 1T DRAM [26], and 8 times longer retention time compared with a Si/Si$_{\mathrm{0.6}}$Ge$_{\mathrm{0.4}}$/Si
QW device [27].
Fig. 6. One-cycle operations at 300 K, 358 K, and 500 K.
Fig. 7. Data retention time and Ion/Ioff ratio at different temperatures of (a) 300 K; (b) 358 K; (c) 400 K; (d) 500 K.
4. Viability of the Device Fabrication Processing
A viable process integration can be suggested for higher practicability of a device
designed by the series of simulation works. It is crucial to grow high-quality SiC
and Si layers on the Si substrate for higher carrier mobility, perfect construction
of the quantum well, and absence of interfacial charge effects. There have been a
variety of epitaxial growth methods for SiC on the Si platform and chemical vapor
deposition (CVD)-based epitaxy has been widely used [28]. A presumable series of integrated processes can be schemed for the fabrication of
the heterojunction QW 1T DRAM cell in this work as schematically shown in Fig. 8(a) through (f). In Fig. 8(a), the process begins with a Si wafer followed by an epitaxial growth of n$^{+}$ 3C-SiC
layer. The SiC layer acts as the source junction and the degenerate n-type doping
can be performed in the in-situ manner. Si can be epitaxially grown in the following
sequence, with a thickness of 100 nm and a doping concentration of p-type 10$^{17}$
cm$^{-3}$ (in-situ doping with the epitaxy), as shown in Fig. 8(b). Another SiC layer is epitaxially grown for constructing the drain junction (Fig. 8(c)). A thin vertical fin channel can is constructed by an anisotropic dry etch leaving
the source junction region for contact as shown in Fig. 8(d). Then, gate oxide and p$^{+}$ poly-Si are sequentially deposited over the vertical
fin channel structure as shown in Fig. 8(e). The deposited poly-Si is etched back with a larger amount than the deposited thickness
to separate the poly-Si gate and put double gates on both sides of the vertical channel
as shown in Fig. 8(f). Here, the etch-back needs to be performed with a high accuracy for determining the
location of gate edge, not making an excessive gate underlap from the drain junction
end [29].
Fig. 8. A presumable process integration for fabricating the proposed heterojunction
1T DRAM cell: (a) Epitaxial growth of n+ 3C-SiC on the Si substrate for source junction; (b) p-type Si epitaxy for constructing
the vertical channel; (c) Epitaxial growth of n+ 3C-SiC on the Si substrate for drain junction; (d) Construction of the vertical thin
channel in the fin shape by dry etching; (e) Gate oxide and gate poly-Si depositions;
(f) Etch-back of the poly-Si for obtaining a double-gate structure.
IV. Conclusion
In this work, a vertical double-gate 3C-SiC/Si/3C-SiC QW 1T DRAM has been designed
and characterized. By introducing a 3C-SiC/Si/3C-SiC QW structure, due to the depth
created by a significantly large VBO between 3C-SiC and Si, extended retention time
and reduced data loss at high temperatures have been achieved. A permissibly optimized
device demonstrated retention times of up to 2 s at room temperature and 0.7 ms at
500 K, proving that the proposed 3C-SiC/Si QW 1T DRAM is a promising candidate for
advanced DRAM technology. Moreover, the proposed 1T DRAM device showed remarkably
improved characteristics superior to those of devices made of other material combinations
making up QW structures.
ACKNOWLEDGMENTS
This work was supported by National R&D Program through the National Research
Foundation of Korea (NRF) funded by the Ministry of Science and ICT of Korea (MSIT)
under Grants 2021M3F4A6A01048300 and 2022M3I7A1078936. The EDA tool was supported
by IC Design Education Center (IDEC).
References
H. El Dirani, K. H. Lee, M. S. Parihar, J. Lacord, S. Martinie, J.-Ch. Barbe, X. Mescot,
P. Fonteneau, J.-E. Broquin, G. Ghibaudo, Ph. Galy, F. Gamiz, Y. Taur, Y.-T. Kim,
S. Cristoloveanu, and M. Bawedin, “Ultra-low power 1T-DRAM in FDSOI technology,” Microelectron.
Eng., vol. 178, pp. 245-249, Jun. 2017.
C. Y. Lim and M.-W. Kwon, “Multi-gate BCAT Structure and Select Word-line Driver in
DRAM for Reduction of GIDL,” J. Semicond. Technol. Sci., vol. 22, no. 6, pp. 452-458,
Dec. 2022.
S. Okhonin, M. Nagoya, J. M. Sallese, and P. Fazan, “A Capacitor-less SOI 1T DRAM
concept,” Proc. IEEE Int. SOI Conf., pp. 153-154, Oct. 2001.
H. Kim, I. M. Kang, S. Cho, W. Sun, and H. Shin, “Analysis of operation characteristics
of junctionless poly-Si 1T-DRAM in accumulation mode,” Semicond. Sci. Technol., vol.
34, no. 10, pp. 1-8, Sep. 2019.
G. Kim, S. W. Kim, K. C. Ryoo, J. H. Oh, M. C. Sun, H. W. Kim, D. W. Kwon, J. S. Jang,
S. Jung, J. H. Kim, and B.-G. Park, “Split-gate-structure 1T DRAM for retention characteristic
improvement,” J. Nanosci. Nanotechnol., vol. 11, no. 7, pp. 5603-5607, Jul. 2011.
S. H. Jang and T. W. Kim, “Retention Time Analysis in a 1T-DRAM With a Vertical Twin
Gate and p+/i/n+ Silicon Nanowire,” IEEE Trans. Electron Devices, vol. 69, no. 9,
pp. 4909-4913, Sep. 2022.
S. Go, S. Kim, D. K. Lee, J. Y. Park, S. Park, D. H. Kim, G. Kim, and S. Kim, “Design
optimization of heterojunction 1T DRAM cell with SiGe body/drain for high performance,”
Semicond. Sci. Technol., vol. 37, no. 12, pp. 1-7, Nov. 2022.
G. Giusi, “Floating body DRAM with body raised and source/drain separation,” Electron.,
vol. 10, no. 6, pp. 1-12, Mar. 2021.
J. Choe, “Memory Technology 2021: Trends & Challenges,” Proc. International Conf.
Simulation of Semiconductor Process and Devices, pp. 111-115, Sep. 2021.
J. Liu, B. Jaiyen, R. Veras, and O. Mutlu, “RAIDR: Retention-aware intelligent DRAM
refresh,” Proc. 39th Annu. Int. Symp. Comput. Archit., pp. 1-12, Jun. 2012.
A. Nitayama, Y. Kohyama, and K. Hieda, “Future directions for DRAM memory cell technology,”
Proc. IEDM Tech. Dig., pp. 355-358, Dec. 1998.
Y. J. Yoon, J. H. Seo, S. Cho, J.-H. Lee, and I. M. Kang, “A polycrystalline-silicon
dual-gate MOSFET-based 1T-DRAM using grain boundary-induced variable resistance,”
Appl. Phys. Lett., vol. 114, no. 18, pp. 1-5, May 2019.
J. H. Seo, Y. J. Yoon, E. Yu, W. Sun, H. Shin, I. M. Kang, J. H. Lee, and S. Cho “Fabrication
and characterization of a thin-body poly-si 1T DRAM with charge-trap effect,” IEEE
Electron Device Lett., vol. 40, no. 4, pp. 566-569, Apr. 2019
S.-W. Lee, S. Cho, I. H. Cho, and G. Kim, “1T DRAM with Raised SiGe Quantum Well for
Sensing Margin Improvement,” J. Semicond. Technol. Sci., vol. 23, no. 1, pp. 64-70,
Feb. 2023.
E. Yu, S. Cho, H. Shin, and B.-G. Park, “A band-engineered one-transistor DRAM with
improved data retention and power efficiency,” IEEE Electron Device Lett., vol. 40,
no. 4, pp. 562-565, Apr. 2019.
D. Berbara, M. A. Abid, M. Hebali, M. Benzohra, and D. Chalabi, “An Ultra-low Power,
High SNM, High Speed and High Temperature of 6T-SRAM Cell in 3C-SiC 130 nm CMOS Technology,”
J. Nano Electron. Phys., vol. 12, no. 4, pp. 1-4, Aug. 2020.
L. Chen, K. Len, Z. Ma, X. Yu, Z. Shen, J. You, W. Li, J. Xu, L. Xu, K. Chen, and
D. Feng, “Tunable Si dangling bond pathway induced forming-free hydrogenated silicon
carbide resistive switching memory device,” J. Phys. Chem. Lett., vol. 11, no. 19,
pp. 8451-8458, Sep. 2020.
A. Mazurak, R. Mroczyński, D. Beke, and A. Gali, “Silicon-carbide (SiC) nanocrystal
technology and characterization and its applications in memory structures,” Nanomater.,
vol. 10, no. 12, pp. 1-11, Nov. 2020.
R. Mroczyński, “Incorporation of silicon-carbide (SiC) nanocrystals in the MIM structures
based on pulsed-DC reactively sputtered HfOx layers,” Proc. 2021 Joint International
EUROSOI Workshop and International Conference on Ultimate Integration on Silicon,
pp. 1-4, Sep. 2021.
A. Mazurak and R. Mroczyński, “Memory effect in MIS structures with embedded all-inorganic
colloidal silicon carbide (SiC) nanocrystals,” Proc. 2019 IEEE 14th Nanotechnology
Materials and Devices Conference, pp. 1-4, Oct. 2019.
E. M. Huseynov, “Neutron Irradiation Effects on the Temperature Dependencies of Electrical
Conductivity of Silicon Carbide (3C-SiC) Nanoparticles,” Silicon, vol. 10, no. 3,
pp. 995-1001, May 2018.
G. Colston and M. Myronov, “Electrical properties of n-type 3C-SiC epilayers in situ
doped with extremely high levels of phosphorus,” Semicond. Sci. Technol., vol. 33,
no. 11, pp. 1-6, Nov. 2018.
A. Biswas and A. M. Ionescu, “1T Capacitor-Less DRAM Cell Based on Asymmetric Tunnel
FET Design,” IEEE J. Electron Devices Soc., vol. 3, no. 3, pp. 217-222, May 2015.
M. Mehregany, C. A. Zorman, N. Rajan, and C. H. Wu, “Silicon Carbide MEMS for Harsh
Environments,” Proc. IEEE, vol. 86, no. 8, pp. 1594-1609, Aug. 1998.
A. Arvanitopoulos, N. Lophitis, S. Perkins, K. N. Gyftakis, M. B. Guadas, and M. Antoniou,
“Physical parameterization of 3C-Silicon Carbide (SiC) with scope to evaluate the
suitability of the material for power diodes as an alternative to 4H-SiC,” Proc. IEEE
11th International Symposium on Diagnostics for Electrical Machines, Power Electronics
and Drives, pp. 565-571, Aug. 2017.
M. H. R. Ansari and S. Cho, “Performance Improvement of 1T DRAM by Raised Source and
Drain Engineering,” IEEE Trans. Electron Devices, vol. 68, no. 4, pp. 1577-1584, Apr.
2021.
A. Lahgere and M. J. Kumar, “1-T Capacitorless DRAM Using Bandgap-Engineered Silicon-Germanium
Bipolar I-MOS,” IEEE Trans. Electron Devices, vol. 64, no. 4, pp. 1583-1590, Apr.
2017.
C. Langpoklakpam, A.-C. Liu, K.-H. Chu, L.-H. Hsu, W.-C. Lee, S.-C. Chen, C.-W. Sun,
M.-H. Shih, K.-Y. Lee, and H.-C. Kuo, “Review of Silicon Carbide Processing for Power
MOSFET,” Cryst., vol. 12, no. 2, pp. 245-1-245-17, Feb. 2022.
S. Kim and S. Cho, “Analysis on Performances of Ultra-Thin Vertical MOSFET Depending
on Position of Gate-Drain Misalignment,” Jpn. J. Appl. Phys., vol. 63, no. 5, pp.
054002-1-054002-6, May 2024.
Soomin Kim received the B.S. degree in Electronic and Electrical Engineering from
Ewha Womans University, Seoul, Korea, in 2023. She is currently pursuing the M.S.
degree at Ewha Womans University. Her current research interests include novel semiconductor
logic and memory devices in the nanoscale region, low-power synaptic devices and circuits,
and scalable solid-state power source for CMOS integration.
Seongjae Cho received the B.S. and the Ph.D. degrees in electrical engineering
from Seoul National University, Seoul, Korea, in 2004 and 2010, respectively. He worked
as an Exchange Researcher at the National Institute of Advanced Industrial Science
and Technology (AIST), Tsukuba, Japan, in 2009. He worked as a Postdoctoral Researcher
at Seoul National University in 2010 and at Stanford University, Palo Alto, CA, from
2010 to 2013. Also, he worked as a faculty member at the Department of Electronic
Engineering, Gachon University, from 2013 to 2023. He is currently working as an Associate
Professor at the Division of Convergence Electronic and Semiconductor Engineering,
Ewha Womans University, Seoul, Korea from 2023. His current interests include emerging
memory deviecs, advanced nanoscale CMOS devices, and novel devices for future computing.