Observation of hopping transitions for delocalized electrons by temperature-dependent conductance in silicon junctionless nanowire transistors
Guo Yang-Yan1, 2, Han Wei-Hua1, 2, †, Zhao Xiao-Song1, 2, Dou Ya-Mei1, 2, Zhang Xiao-Di1, 2, Wu Xin-Yu1, 2, Yang Fu-Hua1, 2, 3
Engineering Research Center for Semiconductor Integrated Technology & Beijing Engineering Center of Semiconductor Micro-Nano Integrated Technology, Institute of Semiconductors, Chinese Academy of Sciences, Beijing 100083, China
Center of Materials Science and Optoelectronics Engineering, University of Chinese Academy of Sciences, Beijing 100049, China
State Key Laboratory for Superlattices and Microstructures, Institute of Semiconductors, Chinese Academy of Sciences, Beijing 100083, China

 

† Corresponding author. E-mail: weihua@semi.ac.cn

Abstract

We demonstrate transitions of hopping behaviors for delocalized electrons through the discrete dopant-induced quantum dots in n-doped silicon junctionless nanowire transistors by the temperature-dependent conductance characteristics. There are two obvious transition platforms within the critical temperature regimes for the experimental conductance data, which are extracted from the unified transfer characteristics for different temperatures at the gate voltage positions of the initial transconductance gm peak in Vg1 and valley in Vg2. The crossover temperatures of the electron hopping behaviors are analytically determined by the temperature-dependent conductance at the gate voltages Vg1 and Vg2. This finding provides essential evidence for the hopping electron behaviors under the influence of thermal activation and long-range Coulomb interaction.

1. Introduction

The prospect of atom-scale transistor has been initially indicated by single atom transistor, in which a single phosphorus atom as a quantum dot is high-precisely positioned between source and drain leads by the probe lithography.[15] Many notable approaches toward atomic electronic devices have also been explored.[69] As the channel width of the silicon transistor is scaled down to several nanometers, few ionized dopant atoms randomly distributed in the channel can work as quantum dots (QDs) and play a significant role in the electron transport behaviors.[1014] In recent years, single-electron tunneling through the dopant-induced QD array has attracted much attention in the study of the silicon junctionless nanowire transistors (JNTs), which may provide a one-dimensional bulk channel with an adjustable width by the gate electric field.[7, 15, 16] The conducting path in the center of the silicon nanowire, which is effectively confined by the surface depletion potentials, can be gradually broadened to the whole conduction channel region with the increase of the gate voltage.[17] Few dopant atoms would be discrete in the extremely narrow channel at the initial gate voltages. Therefore, it is very necessary to further understand the thermally activated electron hopping through discrete dopant-induced QDs in the extremely narrow channel. At low temperatures, the silicon JNTs show that the conductance features evolve from clear oscillatory peaks to several steps with the channel broadening, which reflect the electrons successively passing through the impurity levels of the dopant-induced QDs and the conduction subbands of the quantum wire.[18, 19] With temperature increasing, the conductance features can be smeared due to thermal broadening by the scattering of thermally activated electrons, which are delocalized from the ionized dopant atoms.[20] In this paper, we demonstrate transitions of hopping behaviors for delocalized electrons through the discrete dopant-induced quantum dots in n-doped silicon junctionless nanowire transistors by the temperature-dependent conductance characteristics. We find two obvious transition platforms within the critical temperature regimes for the experimental conductance data, which are extracted from the unified transfer characteristics for different temperatures at the gate voltage positions of the initial transconductance gm peak in Vg1 and valley in Vg2. The crossover temperatures of electron hopping behaviors are analytically determined by the temperature-dependent conductance at the gate voltages Vg1 and Vg2.

2. Device fabrication and characterization

The schematic structure of the silicon JNT device for investigation is provided in Fig. 1(a). The fabrication of the silicon JNT device started from a boron-doped (1015 cm−3) (100)-oriented silicon-on-insulator (SOI) wafer with 55-nm-thick top silicon layer and 145-nm-thick buried oxide layer. After growing a 18-nm-thick thermal oxidation layer, the top silicon layer was uniformly doped by phosphorus ion implantation with a dose of 2 ×1012 cm−2 at an energy of 33 keV. A silicon nanowire was defined along direction by electron beam lithography (EBL) and inductively coupled plasma (ICP) etching, followed by a sacrificial oxidation to eliminate the etching-induced damages. Then 22-nm-thick gate oxide layer was grown by thermal oxidation at the temperature of 900 °C in dry oxygen. The 200-nm-thick polysilicon layer was deposited by low-pressure chemical vapor deposition (LPCVD). After heavily doped by boron ion implantation, the sample was rapidly annealed in nitrogen ambient at 1000 °C for 10 s. The polysilicon gate with the length of 280 nm was then defined by electron beam overlay exposure and dry etching, conformally wrapping the Si/SiO2 core–shell nanowire. Finally, it was followed by the deposition of 200 nm-thick SiO2 passivation layer, standard metal contact formation, and sintering. The SEM image of the single-channel JNT in Fig. 1(b) shows the width of 58 nm for the Si/SiO2 core–shell nanowire. As a result, we may estimate that the silicon core has the physical width of about 14 nm and the height of about 26 nm. The dopant concentration in the silicon channel after several times annealing is estimated to be 4.16×1017 cm−3 according to the implantation dose of phosphorus ions. The fabricated device was measured in a vacuum chamber which can be cooled down to the low temperature of 6 K with the help of Lakershore-340 temperature controller.

Fig. 1. (a) The schematic structure of silicon JNT. (b) The top-view SEM image of silicon JNT after gate formation.
3. Results and discussion

We study the thermally activated electrons transport behaviors through the discrete dopant-induced QDs in the impurity band of the silicon JNT within the temperature range from 6 K to 250 K by the temperature-dependent conductance characteristics of the silicon JNT. The thermally activated electron transport is dominated by phonon-assisted hopping through several dopant-induced QDs.[21] Mott believed that a hopping electron with activation energy Ea would always try to find a lower energy state around the Fermi level by the variable range hopping (VRH) in the absence of long-range Coulomb interaction.[22] At high temperatures, the temperature dependence of conductance G for the nearest neighbor hopping (NNH) exhibits as , where kB is the Boltzmann constant.[23] If the activation energy at low temperatures is reduced as large as the Coulomb interaction energy, the VRH conductance which has been predicted by Efros and Shklovskii (ES) should obey the 1/2 exponent law given by , where TES is the temperature-independent ES coefficient.[24] Therefore, with temperature increasing, three kinds of electron hopping transports would be observed in the following consequence: ES-type VRH, Mott-type VRH, and NNH.[21] In our experiment, we find two obvious transition platforms within the critical temperature regimes of A and B for the experimental conductance data, which are extracted from the unified transfer characteristics at the gate voltage positions of the initial transconductance gm peak in Vg1 and valley in Vg2. One crossover temperature TA in the higher temperature regime A corresponds to the transition of thermally-activated-electron transport behaviors from NNH to Mott-type VRH.[25] The other crossover temperature TC in the lower temperature regime B corresponds to the VRH transition for the delocalized electrons from Mott law to the ES law under the influence of Coulomb interaction.[26] The quantitative analysis on the experimental temperature-dependent conductance data has been taken by linear fitting of for NNH, for Mott-type VRH, and for ES-type VRH.

Figure 2(a) presents the temperature-dependent conductance G (i.e., ) curves obtained from the transfer characteristics at low source–drain bias VDS = 1 mV. The clear quantized current steps below the temperature of 75 K indicate the population of the individual sub-bands caused by the quantum confinement effect.[27] Above the temperature of 75 K, the thermal energy is greater than the subbands spacing, resulting in the smearing of the current steps. Figure 2(b) illustrates the corresponding transconductance gmVG characteristics at different temperatures. We define the onset gate voltage Vgt as the initial point of the first gm peak, where the drain current in the conduction channel is at the onset state. The onset gate voltage Vgt increases from 6 K to 20 K and decreases from 100 K to 250 K, resulting from the interaction of the induced image charges in the dielectric interface with the impurity and the subband states in the channel.[13] To explore the conductance characteristics under the same filled energy level, we take the onset gate voltage Vgt to unify the transfer characteristics in Fig. 2(a) for the alignment of energy levels at different temperatures. In order to clarify the electron hopping transport in the impurity band, figure 2(c) shows the IDSVg curves (upper part) and the corresponding transconductance gmVg curves (lower part) at the temperature of 6 K under the bias VDS varying from 1 mV to 10 mV. The clear oscillatory current with several splitting peaks (upper part) identifies the coupling of dopant-induced QDs in the impurity band. With the gate voltage increasing, the Fermi energy level of electrons in the quantum confined channel is allowed to enter the conduction subbands, resulting in the current steps.[28, 29]

Fig. 2. (a) The temperature-dependent conductance G (i.e., G = IDS/VDS) curves of the device at VDS = 1 mV within the temperature range from 6 K to 250 K. (b) The temperature-dependent transconductance gmVG characteristics at VDS = 1 mV, where the transconductance is given by . (c) IDSVg curves (upper part) and the corresponding transconductance gmVg curves (lower part) at the temperature of 6 K under the bias VDS varying from 1 mV to 10 mV. (d) The Arrhenius conductance plots (G vs. 1/T) on logarithmic scale at the gate voltages of Vg1 and Vg2 within the temperature range from 6 K to 250 K, The inset shows the corresponding temperature-dependent conductance data (G vs. T), which are extracted from panel (a) after unified by the onset gate voltage Vgt at different temperatures.

In order to study the temperature-dependent conductance G characteristics in Fig. 2(a), we firstly extract the experimental conductance data for different temperatures according to the initial gm peak at the gate voltage Vg1, in which the effective mobility of hopping electrons is the highest in the impurity band. For comparison, we also provide the temperature-dependent conductance data according to the initial gm valley at the gate voltage Vg2, in which the electron hopping behavior is suppressed by the Coulomb interaction. Figure 2(d) provides the Arrhenius conductance plots (G vs. 1/T) at the gate voltages of Vg1 and Vg2 within the temperature range from 6 K to 250 K. Interestingly, it is found that the conductance at the temperature of 6 K is much larger than that of 10 K, which may be related with the Coulomb interaction. As the temperature increases, the incomplete ionized donors in the channel are gradually transformed into ionized donors, which result in stronger impurity scattering to reduced the electron mobility.[30] The inset GT diagrams in Fig. 2(d) show two apparent platforms of the conductance for the gate voltages Vg1 and Vg2 within the temperature regions of around 200 K and around 100 K, as the color marked A and B. The platform of the temperature-dependent conductance for the gate voltage Vg1 within the temperature range from 50 K to 100 K is more evident than that of the gate voltage Vg2 within the temperature range from 75 K to 125 K, resulting from the stronger quantum confinement at the initial stage of the conduction channel in the silicon JNT.

In order to precisely determine the transition temperatures TA in temperature region A, we replot the Arrhenius curves of temperature-dependent conductance at high temperature regime as the curves of ln G vs. 1/T in Fig. 3 for the gate voltages Vg1 and Vg2 respectively. The inset shows the Mott-type VRH temperature dependence of the conductance in ln G vs. (1/T)1/4 scales. Here, we are interested in the hopping behaviors of the thermally activated electrons from the ionized dopant atoms. The linearly fitting by is shown in Fig. 3 for the thermally activated electrons by NNH above the crossover temperature TA, in which the electrons have enough thermal activation energy Ea to overcome the potential barriers between the nearest neighbor states. The hopping distance r in NNH is supposed to be equivalent to the mean distance d between neighbored dopant atoms, which can be determined by the doping concentration Nd as

According to the doping concentration , the hopping distance between the nearest neighbored dopant atoms is estimated to be 13.4 nm. According to the linearly fitting of the experimental conductance data in Fig. 3 within the temperature range from 175 K to 250 K, the activation energies Ea for the hopping electrons are extracted to be 45.2 meV and 33.7 meV respectively for the curves at the gate voltages Vg1 and Vg2.[31] According to Mott’s theory, the activation energy Ea for electron hopping is related to the constant density of states (DOS) g0 at the Fermi level and the hopping distance r as[32]
The probability P for hopping is proportional to the conductance G of the device, which depends on the overlap integral of the wavefunctions within the localization radius a and the activation energy Ea,[26] i.e.,
Taking the expression of Ea (Eq. (2)) into Eq. (3), one may get the relation between the probability and the hopping distance, . After taking the derivative of the hopping probability , the optimal electron hopping distance is obtained to be[33]
where aM is the localization length in the Mott hopping regime. At the crossover temperature TA, the electron hopping distance of VRH is assumed to be equal to that of NNH,
By taking the expression of (Eq. (4)) into Eq. (5), we deduce the expression
which is determined by the localization length aM and the density of states g0. In order to determine the localization length aM, we may use the Mott-type VRH conductance expression by substituting Eqs. (2) and (4) into Eq. (3), i.e., . The temperature factor TM is related to the localization length aM by . According to the fitting results from 175 K to 250 K of ln G vs. scales in the inset of Fig. 3, the temperature factor TM extracted from the slopes is 38.784 K for the gate voltage Vg1 and 28.94 K and for the gate voltage Vg2. Therefore, the localization length aM can be obtained from the expression of TM be about 3.47 nm and 4.66 nm respectively for the gate voltages Vg1 and Vg2. The density of states g0 at the Fermi energy can be estimated to be and from Eqs. (2) and (5) by taking the activation energies of 45.2 meV and 33.7 meV respectively for the gate voltages Vg1 and Vg2, both of which have the same magnitude as reported in silicon nanowires.[19] As expected from Eq. (6), we obtain the crossover temperatures of TA1 = 203 K and TA2 = 202 K respectively for the gate voltages Vg1 and Vg2, which are consistent with the experimental observation in the conductance platform A of Fig. 2(d). The result shows that the crossover temperature TA from NNH to VRH is independent of the gate electric field, which may result from the stronger interactions of the thermally activated electrons.

Fig. 3. Arrhenius plot of conductance in ln G vs. 1/T scales for NNH within the temperature range from 175 K to 250 K. The inset shows Mott-type VRH temperature dependence of the conductance in ln G vs. (1/T)1/4 scales.

In order to understand the transition of the conductance at temperature TC, we provide the Arrhenius plot of conductance in ln G vs. (1/T)1/2 scales by linear fitting in Fig. 4. With the temperature decreasing below 150 K, the activation energy extracted from Fig. 2(d) is further reduced to be about 0.84 meV and 0.68 meV respectively for the gate voltages Vg1 and Vg2, by which the VRH at the vicinity of Fermi level EF is associated with the long-range Coulomb interaction of the dopant-induced QD array. If the energy of each dopant site depends on distribution of other charged dopants, the electrons localized on the states near the Fermi level will be redistributed. The long-range Coulomb interaction energy could reduce the DOS for the electrons localized on the states of QDs around the Fermi level, in which the number of states is given by . In order to estimate the DOS near the Fermi level, ES-type VRH model considers that the Coulomb interaction energy is comparable to the activation energy . Then, the DOS near the Fermi level is given by[21]

whose curve is shown in the inset of Fig. 4. At the crossover temperature TC for the VRH transition from Mott-type law to ES-type law, the activation energy Ea should be equivalent to the Coulomb interaction energy , i.e.,
The DOS at the condition of would remain constant as g0 in Mott VRH, which is given by
Then, the hopping distance of ES-type VRH can be estimated by the maximum hopping probability[33] of Eq. (3) as
in which the localized length in ES hopping regime indicates the average modulation length of long-range Coulomb interaction between the dopant-induced QDs in the conduction channel. In order to determine the localization length , we substitute Eqs. (8) and (10) into Eq. (3) to get the ES-type VRH conductance expression
The temperature factor TES is related to the localization length by
The temperature factor TES extracted from the slopes is 34.8 K for the gate voltage Vg1 and 31.4 K for the gate voltage Vg2. Therefore, the localization length for ES-type VRH can be obtained from Eq. (12) to be about 323 nm and 358 nm respectively for the gate voltages Vg1 and Vg2. The result indicates that the localization length is enhanced with the increase of the gate voltage due to screening of the trapping potentials of the dopant atoms. Considering the expressions from Eq. (7) to Eq. (10), we finally obtain the expression of TC,
which is determined by the constant DOS g0 and the localized length aES. As a result, the crossover temperature of the VRH conductance transition from Mott-type law to ES-type law can be estimated to be TC1 = 85 K at the gate voltage Vg1 and TC2 = 126 K at the gate voltage Vg2. Both the two crossover temperatures TC1 and TC2 are consistent with the experimental observation of conductance platform B of Fig. 2(d). The crossover temperature TC2 of 126 K is much higher than TC1 of 85 K, which may result from the stronger Coulomb interaction at the gate voltage Vg2 of the gm valley.

Fig. 4. Arrhenius plot of conductance in ln G vs. (1/T)1/2 scales for ES-type VRH, with the inset for the density of states near the Fermi energy level.
4. Conclusion

We present an experimental evidence of hopping transition for the delocalized electrons in silicon JNT by the temperature-dependent conductance characteristics. The theoretical models of Mott-type VRH and ES-type VRH agree well with the experimental data of temperature-dependent conductance, which are extracted from transfer characteristics for different temperatures at the gate voltage positions of the initial transconductance gm peak in Vg1 and the valley in Vg2. The crossover temperature TA from NNH to Mott VRH is analytically determined to be 203 K and 202 K constantly for the gate voltages Vg1 and Vg2 by the conductance transition of the thermally activated electrons. Another crossover temperature TC of VRH behavior from Mott-type law to ES-type law is theoretically determined to be 85 K and 126 K respectively for the gate voltages Vg1 and Vg2 by considering the Coulomb interactions. As expected, stronger Coulomb interaction at the gate voltage Vg2 of the gm valley leads to the obvious increase of the crossover temperature TC for the VRH transition behavior. Our finding provides essential evidence for the hopping electron behaviors under the influence of thermal activation and long-range Coulomb interaction.

Acknowledgment

The authors acknowledge Dr. Hao Wang, Dr. Liuhong Ma, and Mr. Xiaoming Li for their supports in device fabrication.

Reference
[1] Samanta A Muruganathan M Hori M Ono Y Mizuta H Tabe M Moraru D 2017 Appl. Phys. Lett. 110 093107
[2] Jehl X Niquet Y M Sanquer M 2016 J. Physics.: Condens. Matter 28 103001
[3] Tabe M Moraru D Ligowski M Anwar M Jablonski R Ono Y Mizuno T 2010 Phys. Rev. Lett. 105 016803
[4] Pierre M Wacquez R Jehl X Sanquer M Vinet M Cueto O 2010 Nat. Nanotechnol. 5 133
[5] Moraru D Udhiarto A Anwar M Nowak R Jablonski R Hamid E Tarido J C Mizuno T Tabe M 2011 Nanoscale Research Letters 6 479
[6] Dagesyan S A Shorokhov V V Presnov D E Soldatov E S Trifonov A S Krupenin V A 2017 Nanotechnol. 28 225304
[7] Uddin W Georgiev Y M Maity S Das S 2017 J. Phys. D: Appl. Phys. 50 365104
[8] Enrico Prati M H Filippo Guagliardo Giorgio Ferrari Takahiro Shinada 2012 Nat. Nanotechnol. 7 443
[9] Hamid E Moraru D Tarido J C Miki S Mizuno T Tabe M 2010 Appl. Phys. Lett. 97 262101
[10] Hori M Shinada T Ono Y Komatsubara A Kumagai K Tanii T Endoh T Ohdomari I 2011 Appl. Phys. Lett. 99 062103
[11] Sellier H Lansbergen G P Caro J Rogge S Collaert N Ferain I Jurczak M Biesemans S 2006 Phys. Rev. Lett. 97 206805
[12] Tabe M Moraru D Ligowski M Anwar M Yokoi K Jablonski R Mizuno T 2010 Thin. Solid Films 518 S38
[13] Wang H Han W Ma L Li X Hong W Yang F 2014 Appl. Phys. Lett. 104 133509
[14] Su L N Lv L Li X X Qin H Gu X F 2015 Chin. Phys. Lett. 32 047301
[15] Rueß F J Micolich A P Pok W Goh K E J Hamilton A R Simmons M Y 2008 Appl. Phys. Lett. 92 052101
[16] Pichon L Jacques E Rogel R Salaun A C Demami F 2013 Semicond. Sci. Technol. 28 025002
[17] Colinge J P Lee C W Afzalian A Akhavan N D Yan R Ferain I Razavi P O’Neill B Blake A White M Kelleher A M McCarthy B Murphy R 2010 Nat. Nanotechnol. 5 225
[18] Ma L Han W Wang H Hong W Lyu Q Yang X Yang F 2015 J. Appl. Phys. 117 034505
[19] Wang H Han W Li X Zhang Y Yang F 2014 J. Appl. Phys. 116 124505
[20] Barrett C Lederer D Redmond G Xiong W Colinge J P Quinn A J 2010 Solid-State Electron. 54 1273
[21] Shlimak I 2015 Is Hopping A Science Singapore World Scientific p. 25
[22] Kaveh M F N F M 1983 Philos. Mag. B 47 577
[23] Boris I Shklovskii A L E 1984 Electronic Properties of Doped Semiconductors Berlin Springer-Verlag p. 191
[24] Efros A L Shklovskii B I 1975 J. Phys. C: Solid State Phys. 8 L49
[25] Qiu H Xu T Wang Z Ren W Nan H Ni Z Chen Q Yuan S Miao F Song F Long G Shi Y Sun L Wang J Wang X 2013 Nat. Commun. 4 2642
[26] Yu D Wang C Wehrenberg B L Guyot-Sionnest P 2004 Phys. Rev. Lett. 92 216802
[27] Ma L H Han W H Wang H Yang X Yang A F 2015 IEEE Trans. Elec. Dev. 36 941
[28] Je M Han S Kim I Shin H 2000 Solid-State Electron. 40 2207
[29] Zhao X H Guo Y Y Dou Y M Yang F H 2018 Chin. Phys. B 27 097310
[30] Li X Han W Wang H Ma L Zhang Y Du Y Yang F 2013 Appl. Phys. Lett. 102 223507
[31] Wang W Xu G Chowdhury M D H Wang H Um J K Ji Z Gao N Zong Z Bi C Lu C Lu N Banerjee W Feng J Li L Kadashchuk A Jang J Liu M 2018 Phys. Rev. B 98 245308
[32] Mott N F 1969 Philos. Mag. 19 835
[33] Zhang Y Dai O Levy M Sarachik M P 1990 Phys. Rev. Lett. 64 2687