Next Article in Journal
Active Accumulation of Spherical Analytes on Plasmonic Hot Spots of Double-Bent Au Strip Arrays by Multiple Dip-Coating
Next Article in Special Issue
Emerging 2D Materials and Their Van Der Waals Heterostructures
Previous Article in Journal
Effect of Ultrasonic Surface Rolling Process on the Hot Compression Behavior of Inconel 718 Superalloy at 700 °C
Previous Article in Special Issue
Designing a Novel Monolayer β-CSe for High Performance Photovoltaic Device: An Isoelectronic Counterpart of Blue Phosphorene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Graphene Schottky Junction on Pillar Patterned Silicon Substrate

by
Giuseppe Luongo
1,2,*,
Alessandro Grillo
1,
Filippo Giubileo
2,
Laura Iemmo
1,2,
Mindaugas Lukosius
3,
Carlos Alvarado Chavarin
3,
Christian Wenger
3,4 and
Antonio Di Bartolomeo
1,2,5,*
1
Physics Department “E. R. Caianiello”, University of Salerno, via Giovanni Paolo II n. 132, 84084 Fisciano, Italy
2
CNR-SPIN Salerno, via Giovanni Paolo II n. 132, 84084 Fisciano, Italy
3
IHP–Leibniz Institut fuer innovative Mikroelektronik, Im Technologiepark 25, 15236 Frankfurt (Oder), Germany
4
Brandenburg Medical School Theodor Fontane, 16816 Neuruppin, Germany
5
Interdepartmental Centre NanoMates, University of Salerno, via Giovanni Paolo II n. 132, 84084 Fisciano, Italy
*
Authors to whom correspondence should be addressed.
Nanomaterials 2019, 9(5), 659; https://doi.org/10.3390/nano9050659
Submission received: 29 March 2019 / Revised: 19 April 2019 / Accepted: 22 April 2019 / Published: 26 April 2019

Abstract

:
A graphene/silicon junction with rectifying behaviour and remarkable photo-response was fabricated by transferring a graphene monolayer on a pillar-patterned Si substrate. The device forms a 0.11 eV Schottky barrier with 2.6 ideality factor at room temperature and exhibits strongly bias- and temperature-dependent reverse current. Below room temperature, the reverse current grows exponentially with the applied voltage because the pillar-enhanced electric field lowers the Schottky barrier. Conversely, at higher temperatures, the charge carrier thermal generation is dominant and the reverse current becomes weakly bias-dependent. A quasi-saturated reverse current is similarly observed at room temperature when the charge carriers are photogenerated under light exposure. The device shows photovoltaic effect with 0.7% power conversion efficiency and achieves 88 A/W photoresponsivity when used as photodetector.

1. Introduction

The discovery of two-dimensional (2D) materials such as graphene [1], MoS2 [2,3], WSe2 [4,5], phosphorene and so on [6], has attracted the interests of the scientific community in the recent years. Graphene is still one the most studied materials for its 2D honeycomb structure, high electron mobility, high electrical and thermal conduction, low optical absorption coefficient and easy fabrication methods [1,7,8]. Large graphene layers can be easily synthesized by chemical vapor deposition (CVD) and integrated into the existing semiconductor device technologies. These properties make graphene the perfect candidate to realize a new generation of transistors [9,10,11,12,13,14], diodes [15,16,17,18,19,20], chemical-biological sensors [21,22,23], photodetectors and solar cells [24,25,26,27,28,29,30]. In the recent years, a lot of activity has been focused on the graphene/silicon junction (gr/Si) as one of the simplest graphene devices offering the possibility to study the physical phenomena that occur at the interface between 2D and 3D materials [31]. The gr/Si junction usually forms a Schottky barrier and behaves as a rectifier with a current-voltage (I-V) characteristic similar to that of a metal/semiconductor Schottky diode [31,32]. Because of its particular band structure, graphene possesses low electron density of states close to the Dirac point, hence the Fermi level is highly dependent on charge transfer to or from it. In the gr/Si junction, the application of a bias affects the charge transfer process and the consequent shift of the graphene Fermi energy modulates the gr/Si Schottky barrier height, which becomes therefore bias dependent [31,32]. Indeed, adding such a feature into the standard thermionic emission (T.E.) theory provides an accurate model to describe the gr/Si experimental I-V characteristics [31,33]. Gr/Si Schottky diodes are characterized by a higher ideality factor ( n > 2 ) than metal/semiconductor devices ( n 1.3 ) [31]. The higher n arises because native oxide layers are generally formed at the interface during the graphene transfer process along with silicon interface trap states and/or metallic contamination [34,35]. Obviously, the ideality of the junction can be improved by reducing the interface defects, for instance through a suitable patterning of the substrate. Indeed, the gr/Si tip junctions that we presented in a previous work showed an ideality factor of 1.5 as the patterning of the Si substrate in a tip-array geometry reduces the probability of finding defects or contaminates at the junction, compared to a planar junction of the same area [17]. In addition to that, the tip geometry amplifies the electric field close to the junction, inducing a potential that shifts the graphene Fermi level even at low bias. We exploited such a feature to realize a bias-tunable graphene-based Schottky barrier device [17].
Modifying the substrate geometry is a viable approach to improve the gr/Si device performance or its photoresponse when used as a photodetector. We remark that the photoresponsivity of the gr/Si junction has been also improved by acting on the device structure. One possible way is to reduce the oxide layer underneath the graphene in order to create a metal/oxide/semiconductor (MOS) capacitor next to the gr/Si junction perimeter. Indeed, such an MOS capacitor plays an important role in the photo-charge collection process, by providing photogenerated carriers from the Si substrate to the junction [16,18,19,20,36].
In this work, we combine the tip geometry and the MOS capacitor approach, by fabricating a graphene/silicon junction on Si pillars to realize a bias-tunable Schottky diode that can be used also for photovoltaic and photodetection applications. The pillar perimeter works similarly to the nanotips in enhancing the electric field at the junction but is easier to fabricate and provides a better control of the MOS capacitor areas. We present an extensive analysis of the I-V characteristics of gr/Si pillar junction and evaluate the relevant parameters using the T.E. model and the Cheung and Cheung (C.C.) method [31,32,37]. We also investigate the photo response and the photovoltaic effect of the device using white LED light at different intensities.

2. Materials and Methods

Figure 1a shows the schematic view of the gr/Si-pillar junction. Starting from a highly n-doped silicon substrate ( 10 18   cm 3 ) three pillars with the height of 500   nm and square sections of 30   μ m , 50   μ m and 100   μ m per side were patterned by photolithography (Figure 1b). In a gr/Si junction the Schottky barrier is controlled by the sharper geometries, that is by the pillar perimeter in our case. As the three pillars have similar perimeter/area ratio ( 10%), we expect that they contribute in a similar way to the junction properties. A SiO2 layer was CVD-deposited until it covered the silicon pillars. Chemical-mechanical polishing (CMP) was then used to remove the oxide layer on the top of the pillars. After that, a graphene layer was transferred from Cu foil on the pillars with a method detailed elsewhere [35].
The Raman spectrum of the graphene measured on the SiO2 and Si pillars is shown in Figure 1c. The plot shows two clear peaks at 1568   cm 1 and 2680   cm 1 which indicates that graphene is a good quality monolayer.
A gold contact (anode) was evaporated on the sample through a shadow mask. The other contact (cathode) was formed by coating silver paste on the scratched back-side of the Si substrate. The I-V measurements were performed with a Keithley Semiconductor Characterization System 4200 (SCS-4200) connected to a Janis probe station. During the measurements the sample was kept in dark and at a pressure of 1 mbar.

3. Results

Figure 1d shows the I-V characteristics measured for the gr/Si-pillar junction at different temperatures in the range 200 400   K . From low to room temperature the gr/Si junction shows an exponential reverse current which is typical of gr/Si junctions [17]. At higher temperatures, after the initial fast growth of the ohmic regime at low bias, the reverse current exhibits a gradual weaker dependence on the bias until it becomes quasi-saturated. The I-V characteristic at room temperature shows a rectification factor of two orders of magnitude at ± 1.5   V .
The exponential reverse current growth at lower temperatures in Figure 1d can be explained considering the Fermi level shift due to the graphene low density of states, which reduces the Schottky barrier in reverse bias [31]. The variation of the barrier can be contributed also by the geometry and doping level of the substrate through the image-force barrier lowering. The pillar geometry magnifies the electric field around the perimeter where a wider depletion layer is created. Such a depletion layer is mirrored by charges in graphene, which cause an up-shift of the Fermi level and a reduction of the Schottky barrier. The high doping of the Si substrate can further contribute to barrier lowering through the image force effect. Conversely, the change of behaviour at higher temperatures indicates that the augmenting thermal generation rate in the depletion layer dominates the reverse leakage current which becomes less sensitive to the bias. The slight deviation of such current from saturation can be ascribed to image force barrier lowering [38,39].
To determine the Schottky diode parameters, we use the T.E. model with voltage dependent Schottky barrier height q ϕ B [31], expressed by the equations:
I = I 0 e q V n k T ( 1 e q V k T ) ,
I 0 = A A * T 2 e q ϕ B k T ,
where I0 is the reverse saturation current, q the electron charge, n > 1 the ideality factor, k the Boltzmann constant, T the temperature, A the junction area, A * = 4 π m e * k 2 h 3 = 112   A   cm 2   K 2 the Richardson constant for n-type Si ( m e * is the electron effective mass and h is the Plank constant) [40]. For q V > n k T , Equations (1) and (2) can be rewritten as:
ln ( I ) = ln ( I 0 ) + q V n k T   ,
ln ( I 0 T 2 ) = ln ( A A * ) q ϕ B k T .
According to Equation (3), the straight-line fitting of the ln(I)-V characteristics for q V > > k T can be used to extrapolate the reverse current I 0 at zero bias and to estimate the ideality factor n . The so-obtained ideality factor as a function of temperature is shown in Figure 2a. The ideality factor at room temperature is n 2.6 and is a monotonic decreasing function of the temperature because several non-idealities manifest more at lower temperatures. These non-idealities include metal residues consequence of the etching process (Cu in this case) which form carrier recombination centers, interface states at the junction which lead to charge trapping and detrapping, and the presence of a native oxide layer [31,34]. The zero-bias current, I 0 , is used in the Richardson plot, ln ( I 0 / T 2 ) vs 1 T , shown in Figure 2b, which, according to Equation (4), yields a Schottky barrier at zero-bias of 0.11   eV and ln ( A A *   ) = 33.72 . Since the effective gr/Si junction contact area is   1.34 10 2   mm 2 , the Richardson constant is A *   =   1.68 · 10 9   Acm 2 K 2 . A possible explanation for the low Richardson constant and the ideality factor n > 2 is the presence of a thin oxide layer [16]. Taking into account the native oxide thickness, Equation (2) can be modified by adding a tunnelling factor as:
I 0 = A A * exp ( χ 1 2 δ ) exp ( q ϕ B k T ) ,
where δ (expressed in Å) is the oxide layer thickness and χ 3   eV is the differences between the energy Fermi level and the conduction band minimum of SiO2. From Equation (5), we estimated an oxide layer of 15 Å, which is thin enough to allow a tunnelling current, but can sustain a voltage drop and affect the I-V characteristic of the junction.
At higher positive bias ( V   0.8   V ) , the thermionic emission current is limited by the series resistance R S , which is the lump sum of contact, graphene and substrate resistances. By taking it into account, Equation (1) can be rewritten as
I = I 0 e q ( V I R S ) n k T ,
And from Equation (6), two new equations can be derived when V I R S > n k T / q [37]:
d V d ( ln ( I ) ) = I R S n k T q ,
H ( I ) = I R S + n ϕ B ,
where H ( I ) is defined as:
H ( I ) = V n k T q ln ( I A A * T 2 ) .
Accordingly, the series resistance and the ideality factor can be extrapolated from the slope and the intercept of the d V d ( ln ( I ) ) vs I plot (Figure 3a), respectively, while the Schottky barrier can be estimated from the intercept of H ( I ) vs I plot (Figure 3b). Using this method, at room temperature, we obtain 10   M Ω series resistance and ideality factor 3 . Figure 3c,d display the series resistance, the ideality factor and q ϕ B measured at different temperatures. The decreasing series resistance with increasing temperature shows the typical semiconductor behaviour. This behaviour cannot be attributed to silicon, Au or Ag paste in this temperature range [41,42,43]. Therefore, it can only be caused by the graphene layer. The resistance drop at high temperature and the negative d R S / d T has been reported for both exfoliated and CVD grown graphene [44,45,46]. The graphene semimetal behaviour has been attributed mainly to the thermally activated transport through the inhomogeneous electron-hole puddles, the formation of which is favoured by the transfer process of CVD-grown graphene [35,46].
Using Equations (7)–(9), we estimate q ϕ B at different temperatures (Figure 3d); in particular q ϕ B 0.11   eV at room temperature which is in agreement with the previous evaluation. The temperature growing q ϕ B is an indication of possible spatial inhomogeneities. The homogeneity of the barrier will be discussed later. In Figure 4a,b we show the Richardson plot at given forward and reverse biases. In forward bias, the temperature dependence of the current has a linear behaviour, which is in agreement with the T.E. theory. Contrarily, in reverse bias, the evolving behaviour of the current, from exponential to saturation trend, is reflected in the Richardson plot (Figure 4b), which for T ≤ 300 K is similar to the forward bias one (Figure 4a), while at higher temperature shows rising converging curves. Because of this, we consider only the lower temperature part of the curves in Figure 4b ( T     300   K ) to determinate the Schottky barrier and the ln ( A A * ) , which are displayed in Figure 4c. We highlight that the Schottky barrier increases with the applied voltage, as expected. In forward bias, the graphene Fermi energy shifts down with respect to the semiconductor energy bands, thus increasing the Schottky barrier, while the opposite occurs in reverse bias. The relative shift, and therefore the barrier variation, is enhanced by the magnified electric field of the pillar and is made possible by the depinning of the Fermi level caused by the thin interfacial oxide layer [17,40].
Because of the CMP treatment (see Section 2), there is a possibility that the pillar top surface is not homogeneous and there could be points where the Schottky barrier is higher or lower. Following Refs. [17,40,47], we assume that the spatial variation of the Schottky barrier can be described by a Gaussian distribution. Therefore, the temperature dependence of the barrier is expressed as:
q ϕ B = q ϕ B M q σ 2 2 k T   ,
where q ϕ B M is the maximum Schottky barrier and σ is the standard deviation of the Gaussian distribution. σ characterizes the inhomogeneity of the Schottky barrier and can be extracted from a plot of q ϕ B vs 1 / 2 k T (Figure 4d). We obtain σ = 45   meV , which is lower than those reported in literature for CVD grown graphene [48,49]. Since the graphene was CVD grown, the low standard variation can be considered as a remarkable advantage of the patterning of the substrate.
Finally, we measured the gr/Si response to light. Figure 5a shows the semi-logarithmic I-V curves of the device measured under different white LED light intensities. The responsivity = ( I l i g h t I d a r k ) / P o p t ( I l i g h t and I d a r k are the current measured at 1 V under illumination and in dark, respectively) as a function of the incident light power P o p t is shown in Figure 5b. The device presents a responsivity with a maximum of 88   A / W at 10 5 10 4   Wcm 2 , which decreases at higher intensities. The reduction of the responsivity at higher intensities is due to the raising recombination rate. Indeed, at high illumination, the increasing of electron-hole pair density in the depletion layer enhances the recombination rate thus making the photocurrent deviate from its linearly behaviour as shown Figure 5c.
Remarkably, the device achieves a reverse current that can be greater than the forward one. The high reverse current measured at high illumination confirms that there is a contribution to the junction current from the photogeneration occurring in the substrate areas where graphene forms a MOS capacitor with Si, as explained in previous works [16,18,36,50]. Furthermore, we note that the photogeneration has the same effect as the thermal generation in shaping the I-V curves of the device. Figure 5d shows the I-V measured in dark and under illumination at 5   mW / cm 2 in linear scale. A photovoltaic effect with an open circuit voltage around 0.19   V , which is close to the estimated Schottky barrier height, and a short circuit current of 1.8   nA , corresponding to 0.7 % power conversion efficiency, can be clearly observed. The conversion efficiency can be improved by lowering the doping of the Si substrate, which would result in an extended depletion layer for enhanced light absorption, and by reducing the shunt and series resistance that would increase the cell fill factor.

4. Conclusions

In conclusion, we fabricated a gr/Si pillar junction that possesses both a bias-tunable Schottky barrier, remarkable photoresponse and appreciable power conversion efficiency. The reverse current grows exponentially with reverse bias at lower temperatures, while it shows a saturation at higher temperatures or under illumination. Such behaviour has been explained by taking into account the thermo- and photo- generated charges both at the gr/Si junction and in the surrounding regions.

Author Contributions

A.D.B. conceived the experiment, G.L., A.G., M.L. and C.A.C. performed the experiment, G.L., L.I., F.G. and A.D.B. analysed the data, A.D.B., F.G. and C.W. contributed reagents/materials/analysis tools, G.L., F.G. and A.D.B. wrote the article.

Funding

We acknowledge the economic support of POR Campania FSE 2014–2020, Asse III Ob. specifico l4, D.D. n. 80, 31/05/2016 and CNR-SPIN SEED Project 2017 and Project PICO & PRO, ARS S01_01061, PON “Ricerca e Innovazione” 2014–2020.

Acknowledgments

We thank Technology Department, IHP-Microelectronics, Frankfurt Oder, Germany for the fabrication of the devices.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Geim, A.K.; Novoselov, K.S. The rise of graphene. Nat. Mater. 2007, 6, 183–191. [Google Scholar] [CrossRef] [PubMed]
  2. Ganatra, R.; Zhang, Q. Few-Layer MoS2: A promising layered semiconductor. ACS Nano 2014, 8, 4074–4099. [Google Scholar] [CrossRef]
  3. Di Bartolomeo, A.; Genovese, L.; Giubileo, F.; Iemmo, L.; Luongo, G.; Foller, T.; Schleberger, M. Hysteresis in the transfer characteristics of MoS2 transistors. 2D Mater. 2017, 5, 015014. [Google Scholar] [CrossRef]
  4. Li, H.; Lu, G.; Wang, Y.; Yin, Z.; Cong, C.; He, Q.; Wang, L.; Ding, F.; Yu, T.; Zhang, H. Mechanical exfoliation and characterization of single- and few-layer nanosheets of WSe2, TaS2, and TaSe2. Small 2013, 9, 1974–1981. [Google Scholar] [CrossRef]
  5. Di Bartolomeo, A.; Urban, F.; Passacantando, M.; McEvoy, N.; Peters, L.; Iemmo, L.; Luongo, G.; Romeo, F.; Giubileo, F. A WSe 2 vertical field emission transistor. Nanoscale 2019, 11, 1538–1548. [Google Scholar] [CrossRef] [PubMed]
  6. Carvalho, A.; Wang, M.; Zhu, X.; Rodin, A.S.; Su, H.; Castro Neto, A.H. Phosphorene: From theory to applications. Nat. Rev. Mater. 2016, 1, 16061. [Google Scholar] [CrossRef]
  7. Castro, N.A.H.; Guinea, F.; Peres, N.M.R.; Novoselov, K.S.; Geim, A.K. The electronic properties of graphene. Rev. Modern Phys. 2009, 81, 109–162. [Google Scholar] [CrossRef] [Green Version]
  8. Xia, F.; Perebeinos, V.; Lin, Y.; Wu, Y.; Avouris, P. The origins and limits of metal–graphene junction resistance. Nat. Nanotechnol. 2011, 6, 179–184. [Google Scholar] [CrossRef]
  9. Di Bartolomeo, A.; Santandrea, S.; Giubileo, F.; Romeo, F.; Petrosino, M.; Citro, R.; Barbara, P.; Lupina, G.; Schroeder, T.; Rubino, A. Effect of back-gate on contact resistance and on channel conductance in graphene-based field-effect transistors. Diam. Relat. Mater. 2013, 38, 19–23. [Google Scholar] [CrossRef] [Green Version]
  10. Bartolomeo, A.D.; Giubileo, F.; Romeo, F.; Sabatino, P.; Carapella, G.; Iemmo, L.; Schroeder, T.; Lupina, G. Graphene field effect transistors with niobium contacts and asymmetric transfer characteristics. Nanotechnology 2015, 26, 475202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Giubileo, F.; Di Bartolomeo, A. The role of contact resistance in graphene field-effect devices. Prog. Surf. Sci. 2017, 92, 143–175. [Google Scholar] [CrossRef] [Green Version]
  12. Di Bartolomeo, A.; Giubileo, F.; Iemmo, L.; Romeo, F.; Russo, S.; Unal, S.; Passacantando, M.; Grossi, V.; Cucolo, A.M. Leakage and field emission in side-gate graphene field effect transistors. Appl. Phys. Lett. 2016, 109, 023510. [Google Scholar] [CrossRef]
  13. Yang, H.; Heo, J.; Park, S.; Song, H.J.; Seo, D.H.; Byun, K.-E.; Kim, P.; Yoo, I.; Chung, H.-J.; Kim, K. Graphene barristor, a triode device with a gate-controlled schottky barrier. Science 2012, 336, 1140–1143. [Google Scholar] [CrossRef]
  14. Mehr, W.; Dabrowski, J.; Scheytt, J.C.; Lippert, G.; Xie, Y.-H.; Lemme, M.C.; Ostling, M.; Lupina, G. Vertical graphene base transistor. IEEE Electr. Dev. Lett. 2012, 33, 691–693. [Google Scholar] [CrossRef]
  15. Chen, C.-C.; Aykol, M.; Chang, C.-C.; Levi, A.F.J.; Cronin, S.B. Graphene-silicon schottky diodes. Nano Lett. 2011, 11, 1863–1867. [Google Scholar] [CrossRef]
  16. Di Bartolomeo, A.; Luongo, G.; Giubileo, F.; Funicello, N.; Niu, G.; Schroeder, T.; Lisker, M.; Lupina, G. Hybrid graphene/silicon Schottky photodiode with intrinsic gating effect. 2D Mater. 2017, 4, 025075. [Google Scholar] [CrossRef] [Green Version]
  17. Di Bartolomeo, A.; Giubileo, F.; Luongo, G.; Iemmo, L.; Martucciello, N.; Niu, G.; Fraschke, M.; Skibitzki, O.; Schroeder, T.; Lupina, G. Tunable Schottky barrier and high responsivity in graphene/Si-nanotip optoelectronic device. 2D Mater. 2016, 4, 015024. [Google Scholar] [CrossRef] [Green Version]
  18. Luongo, G.; Giubileo, F.; Genovese, L.; Iemmo, L.; Martucciello, N.; Di Bartolomeo, A. I-V and C-V characterization of a high-responsivity graphene/silicon photodiode with embedded MOS capacitor. Nanomaterials 2017, 7, 158. [Google Scholar] [CrossRef]
  19. Luongo, G.; Di Bartolomeo, A.; Giubileo, F.; Chavarin, C.A.; Wenger, C. Electronic properties of graphene/p-silicon Schottky junction. J. Phys. D Appl. Phys. 2018, 51, 255305. [Google Scholar] [CrossRef] [Green Version]
  20. Luongo, G.; Giubileo, F.; Iemmo, L.; Di Bartolomeo, A. The role of the substrate in Graphene/Silicon photodiodes. J. Physi. Conf. Ser. 2018, 956, 012019. [Google Scholar] [CrossRef] [Green Version]
  21. Kim, H.-Y.; Lee, K.; McEvoy, N.; Yim, C.; Duesberg, G.S. Chemically modulated graphene diodes. Nano Lett. 2013, 13, 2182–2188. [Google Scholar] [CrossRef]
  22. Singh, A.; Uddin, M.A.; Sudarshan, T.; Koley, G. Tunable Reverse-biased graphene/silicon heterojunction schottky diode sensor. Small 2014, 10, 1555–1565. [Google Scholar] [CrossRef]
  23. Fattah, A.; Khatami, S.; Mayorga-Martinez, C.C.; Medina-Sánchez, M.; Baptista-Pires, L.; Merkoçi, A. Graphene/silicon heterojunction schottky diode for vapors sensing using impedance spectroscopy. Small 2014, 10, 4193–4199. [Google Scholar] [CrossRef] [PubMed]
  24. Riazimehr, S.; Bablich, A.; Schneider, D.; Kataria, S.; Passi, V.; Yim, C.; Duesberg, G.S.; Lemme, M.C. Spectral sensitivity of graphene/silicon heterojunction photodetectors. Solid State Electron. 2016, 115, 207–212. [Google Scholar] [CrossRef] [Green Version]
  25. Bonaccorso, F.; Sun, Z.; Hasan, T.; Ferrari, A.C. Graphene photonics and optoelectronics. Nat. Photonics 2010, 4, 611–622. [Google Scholar] [CrossRef] [Green Version]
  26. An, X.; Liu, F.; Jung, Y.J.; Kar, S. Tunable graphene–silicon heterojunctions for ultrasensitive photodetection. Nano Lett. 2013, 13, 909–916. [Google Scholar] [CrossRef] [PubMed]
  27. Ferrari, A.C.; Bonaccorso, F.; Fal’ko, V.; Novoselov, K.S.; Roche, S.; Bøggild, P.; Borini, S.; Koppens, F.H.L.; Palermo, V.; Pugno, N.; et al. Science and technology roadmap for graphene, related two-dimensional crystals, and hybrid systems. Nanoscale 2015, 7, 4598–4810. [Google Scholar] [PubMed] [Green Version]
  28. An, X.; Liu, F.; Kar, S. Optimizing performance parameters of graphene–silicon and thin transparent graphite–silicon heterojunction solar cells. Carbon 2013, 57, 329–337. [Google Scholar] [CrossRef]
  29. Behura, S.K.; Nayak, S.; Mukhopadhyay, I.; Jani, O. Junction characteristics of chemically-derived graphene/p-Si heterojunction solar cell. Carbon 2014, 67, 766–774. [Google Scholar] [CrossRef]
  30. Ruan, K.; Ding, K.; Wang, Y.; Diao, S.; Shao, Z.; Zhang, X.; Jie, J. Flexible graphene/silicon heterojunction solar cells. J. Mater. Chem. A 2015, 3, 14370–14377. [Google Scholar] [CrossRef]
  31. Di Bartolomeo, A. Graphene Schottky diodes: An experimental review of the rectifying graphene/semiconductor heterojunction. Phys. Rep. 2016, 606, 1–58. [Google Scholar] [CrossRef] [Green Version]
  32. Liang, S.-J.; Hu, W.; Di Bartolomeo, A.; Adam, S.; Ang, L.K. A modified Schottky model for graphene-semiconductor (3D/2D) contact: A combined theoretical and experimental study. In Proceedings of the 2016 IEEE International Electron Devices Meeting (IEDM), San Francisco, CA, USA, 3–7 December 2016; pp. 14.4.1–14.4.4. [Google Scholar]
  33. Tung, R.T. Electron transport at metal-semiconductor interfaces: General theory. Phys. Rev. B 1992, 45, 13509–13523. [Google Scholar] [CrossRef]
  34. Zhang, X.; Zhang, L.; Ahmed, Z.; Chan, M. Origin of nonideal graphene-silicon schottky junction. IEEE Transact. Electr. Dev. 2018, 65, 1995–2002. [Google Scholar] [CrossRef]
  35. Lupina, G.; Kitzmann, J.; Costina, I.; Lukosius, M.; Wenger, C.; Wolff, A.; Vaziri, S.; Östling, M.; Pasternak, I.; Krajewska, A.; et al. Residual metallic contamination of transferred chemical vapor deposited graphene. ACS Nano 2015, 9, 4776–4785. [Google Scholar] [CrossRef] [PubMed]
  36. Riazimehr, S.; Kataria, S.; Bornemann, R.; Haring Bolívar, P.; Ruiz, F.J.G.; Engström, O.; Godoy, A.; Lemme, M.C. High photocurrent in gated graphene–silicon hybrid photodiodes. ACS Photonics 2017, 4, 1506–1514. [Google Scholar] [CrossRef] [PubMed]
  37. Cheung, S.K.; Cheung, N.W. Extraction of Schottky diode parameters from forward current-voltage characteristics. Appl. Phys. Lett. 1986, 49, 85–87. [Google Scholar] [CrossRef]
  38. Zhang, X.; Zhang, L.; Chan, M. Doping enhanced barrier lowering in graphene-silicon junctions. Appl. Phys. Lett. 2016, 108, 263502. [Google Scholar] [CrossRef]
  39. Di Bartolomeo, A.; Grillo, A.; Urban, F.; Iemmo, L.; Giubileo, F.; Luongo, G.; Amato, G.; Croin, L.; Sun, L.; Liang, S.-J.; et al. Asymmetric schottky contacts in bilayer MoS 2 field effect transistors. Adv. Funct. Mater. 2018, 28, 1800657. [Google Scholar] [CrossRef]
  40. Sze, S.M.; Ng, K.K. Physics of Semiconductor Devices, 3rd ed.; Wiley-Interscience: Hoboken, NJ, USA, 2007; ISBN 978-0-471-14323-9. [Google Scholar]
  41. Li, S.S.; Thurber, W.R. The dopant density and temperature dependence of electron mobility and resistivity in n-type silicon. Solid State Electron. 1977, 20, 609–616. [Google Scholar] [CrossRef]
  42. Jacoboni, C.; Canali, C.; Ottaviani, G.; Alberigi Quaranta, A. A review of some charge transport properties of silicon. Solid State Electron. 1977, 20, 77–89. [Google Scholar] [CrossRef]
  43. Swirhun, S.E.; Swanson, R.M. Temperature dependence of specific contact resistivity. IEEE Electr. Dev. Lett. 1986, 7, 155–157. [Google Scholar] [CrossRef]
  44. Shao, Q.; Liu, G.; Teweldebrhan, D.; Balandin, A.A. High-temperature quenching of electrical resistance in graphene interconnects. Appl. Phys. Lett. 2008, 92, 202108. [Google Scholar] [CrossRef]
  45. Skákalová, V.; Kaiser, A.B.; Yoo, J.S.; Obergfell, D.; Roth, S. Correlation between resistance fluctuations and temperature dependence of conductivity in graphene. Phys. Rev. B 2009, 80, 153404. [Google Scholar] [CrossRef] [Green Version]
  46. Heo, J.; Chung, H.J.; Lee, S.-H.; Yang, H.; Seo, D.H.; Shin, J.K.; Chung, U.-I.; Seo, S.; Hwang, E.H.; Das Sarma, S. Nonmonotonic temperature dependent transport in graphene grown by chemical vapor deposition. Phys. Rev. B 2011, 84, 035421. [Google Scholar] [CrossRef] [Green Version]
  47. Werner, J.H.; Güttler, H.H. Barrier inhomogeneities at Schottky contacts. J. Appl. Phys. 1991, 69, 1522–1533. [Google Scholar] [CrossRef]
  48. Parui, S.; Ruiter, R.; Zomer, P.J.; Wojtaszek, M.; van Wees, B.J.; Banerjee, T. Temperature dependent transport characteristics of graphene/n-Si diodes. J. Appl. Phys. 2014, 116, 244505. [Google Scholar] [CrossRef] [Green Version]
  49. Tomer, D.; Rajput, S.; Hudy, L.J.; Li, C.H.; Li, L. Inhomogeneity in barrier height at graphene/Si (GaAs) Schottky junctions. Nanotechnology 2015, 26, 215702. [Google Scholar] [CrossRef] [Green Version]
  50. Di Bartolomeo, A.; Luongo, G.; Iemmo, L.; Urban, F.; Giubileo, F. Graphene–Silicon Schottky Diodes for Photodetection. IEEE Transact. Nanotechnol. 2018, 17, 1133–1137. [Google Scholar] [CrossRef]
Figure 1. (a) Two-dimensional (2D) schematic view of the gr/Si-pillar device. (b) Optical microscope image of the pillars. (c) Raman spectroscopy of the graphene on SiO2 and Si. (d) The current-voltage (I-V) characteristic of the device measured from 200   K to 400   K .
Figure 1. (a) Two-dimensional (2D) schematic view of the gr/Si-pillar device. (b) Optical microscope image of the pillars. (c) Raman spectroscopy of the graphene on SiO2 and Si. (d) The current-voltage (I-V) characteristic of the device measured from 200   K to 400   K .
Nanomaterials 09 00659 g001
Figure 2. (a) Ideality factor vs the temperature extracted from the thermionic emission (T.E.) model (b) Richardson plot of the ln ( I 0 / T 2 ) versus 10 3 / T .
Figure 2. (a) Ideality factor vs the temperature extracted from the thermionic emission (T.E.) model (b) Richardson plot of the ln ( I 0 / T 2 ) versus 10 3 / T .
Nanomaterials 09 00659 g002
Figure 3. Cheung’s plot of (a) d V / d ln ( I ) vs I and (b) H ( I ) vs I at 300 K . (c) Devices series resistance, (d) ideality factor and the Schottky barrier extracted from the Cheung and Cheung (CC) method versus the temperature.
Figure 3. Cheung’s plot of (a) d V / d ln ( I ) vs I and (b) H ( I ) vs I at 300 K . (c) Devices series resistance, (d) ideality factor and the Schottky barrier extracted from the Cheung and Cheung (CC) method versus the temperature.
Nanomaterials 09 00659 g003
Figure 4. (a) Richardson plot of ln ( I / T 2 ) vs 10 3   T in forward and (b) in reverse bias. (c) Schottky barrier and ln ( A A * ) respect the bias. (d) Schottky barrier height at zero bias as a function of temperature.
Figure 4. (a) Richardson plot of ln ( I / T 2 ) vs 10 3   T in forward and (b) in reverse bias. (c) Schottky barrier and ln ( A A * ) respect the bias. (d) Schottky barrier height at zero bias as a function of temperature.
Nanomaterials 09 00659 g004
Figure 5. (a) I-V characteristic in semilogarithmic scale of the gr/Si pillar device measured at different intensity illumination level. (b) Responsivity of the gr/Si pillar device as function of the light intensity. (c) Photocurrent measured at 1 V and at different light intensities in logarithmic scale. (d) I-V characteristic measured in dark (black line) and at 5 mWcm−2 (red line).
Figure 5. (a) I-V characteristic in semilogarithmic scale of the gr/Si pillar device measured at different intensity illumination level. (b) Responsivity of the gr/Si pillar device as function of the light intensity. (c) Photocurrent measured at 1 V and at different light intensities in logarithmic scale. (d) I-V characteristic measured in dark (black line) and at 5 mWcm−2 (red line).
Nanomaterials 09 00659 g005

Share and Cite

MDPI and ACS Style

Luongo, G.; Grillo, A.; Giubileo, F.; Iemmo, L.; Lukosius, M.; Alvarado Chavarin, C.; Wenger, C.; Di Bartolomeo, A. Graphene Schottky Junction on Pillar Patterned Silicon Substrate. Nanomaterials 2019, 9, 659. https://doi.org/10.3390/nano9050659

AMA Style

Luongo G, Grillo A, Giubileo F, Iemmo L, Lukosius M, Alvarado Chavarin C, Wenger C, Di Bartolomeo A. Graphene Schottky Junction on Pillar Patterned Silicon Substrate. Nanomaterials. 2019; 9(5):659. https://doi.org/10.3390/nano9050659

Chicago/Turabian Style

Luongo, Giuseppe, Alessandro Grillo, Filippo Giubileo, Laura Iemmo, Mindaugas Lukosius, Carlos Alvarado Chavarin, Christian Wenger, and Antonio Di Bartolomeo. 2019. "Graphene Schottky Junction on Pillar Patterned Silicon Substrate" Nanomaterials 9, no. 5: 659. https://doi.org/10.3390/nano9050659

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop