Open access peer-reviewed chapter

Ge0.83Sn0.17 P-Channel Metal-Oxide-Semiconductor Field- Effect Transistors: Impact of Sulfur Passivation on Gate Stack Quality

Written By

Dian Lei and Xiao Gong

Submitted: 08 November 2017 Reviewed: 26 January 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.74532

From the Edited Volume

Design, Simulation and Construction of Field Effect Transistors

Edited by Dhanasekaran Vikraman and Hyun-Seok Kim

Chapter metrics overview

1,377 Chapter Downloads

View Full Metrics

Abstract

The effect of sulfur passivation of the surface of Ge0.83Sn0.17 is investigated. X-ray photoelectron spectroscopy (XPS) was used to examine the interfacial property between HfO2 and Ge0.83Sn0.17. Sulfur passivation is effective in reducing both the Ge oxides and Sn oxides formation and the Sn atoms segregation. In addition, sulfur passivation reduces the interface trap density Dit at the HfO2/Ge0.83Sn0.17 interface from the valence band edge to the midgap. After the implementation of sulfur passivation, Ge0.83Sn0.17 p-MOSFETs show improved subthreshold swing S and effective hole mobility μeff. 25% μeff enhancement can be observed in Ge0.83Sn0.17 p-MOSFETs with sulfur passivation at a high inversion carrier density Ninv of 1 × 1013 cm−2.

Keywords

  • GeSn
  • p-MOSFETs
  • sulfur passivation
  • XPS
  • dangling bond

1. Introduction

Materials with high carrier mobilities such as germanium (Ge) could replace Silicon (Si) as the channel material in metal-oxide-semiconductor field-effect transistors (MOSFETs) for future high performance logic applications [1, 2, 3, 4, 5, 6, 7, 8, 9, 10]. Recently, germanium-tin (GeSn) was reported to have a higher hole mobility than Ge and is a promising channel material for p-channel MOSFETs (p-MOSFETs) [11, 12, 13, 14, 15, 16, 17, 18].

Theoretical calculation [11] shows that the light hole effective mass of GeSn decreases with increasing Sn composition. It is also demonstrated experimentally [19] that increasing Sn composition in GeSn p-MOSFETs increases the effective hole mobility. However, due to the low surface energy and large covalent radius of Sn, Sn atoms may segregate to the surface during growth of GeSn [20, 21, 22]. Hence, the thermal stability may be worse at a higher Sn composition. Li et al. reported that a Sn-rich surface layer would form when Ge0.922Sn0.078 is annealed at 620°C [23]. A similar phenomenon occurs on Ge0.915Sn0.085 surface after annealing at 500°C [24]. For Sn composition as high as 17%, self-assembled Sn wires can form at an annealing temperature as low as 280°C [25]. Severe Sn segregation may reduce carrier mobility and degrade the drive current in MOSFETs. Therefore, in order to achieve high performance GeSn p-MOSFETs with high Sn composition, a fabrication process with low thermal budget may be required to maintain a good quality of the GeSn channel material and the GeSn/high-k dielectric interface.

Various passivation techniques have been demonstrated to be effective in improving the gate stack quality of both Ge and GeSn channel p-MOSFETs, such as Si2H6 passivation [12, 17], Ge capping [26], GeSnOx passivation [27, 28], and sulfur passivation [15, 29, 30]. Among these passivation techniques adopted for GeSn p-MOSFETs fabrication, Si2H6 passivation and GeSnOx passivation require a process temperature higher than 370°C and 400°C, respectively. It has already been reported that Sn can segregate out to the GeSn surface during Si passivation process at a temperature of 370°C and degrades the device performance [17]. Therefore, low temperature passivation technique was investigated in this work for the fabrication of GeSn p-MOSFETs with Sn composition of 17%.

The adsorption of sulfur atoms is a promising route to chemically and electrically passivate the highly reactive Ge and GeSn surface [15, 29, 30, 31]. Compared with other passivation techniques, room temperature sulfur passivation using (NH4)2S solution has several advantages: (1) GeSn surface could be effectively passivated through the formation of covalent S-Ge and S-Sn bond. This will reduce oxide formation which degrades device performance; (2) The formed sulfur passivation layer is very thin with very little increase on the effective oxide thickness (EOT); (3) Sn segregation can be suppressed during the passivation process owning to a lower thermal budget. Sulfur passivation has already been implemented into Ge0.947Sn0.053 p-MOSFETs fabrication and demonstrated enhanced peak hole mobility as compared with Si2H6 passivation [15]. However, the mechanism of the effect of sulfur passivation on the GeSn/HfO2 interface quality has not been investigated. In addition, the impact of sulfur passivation on the reduction of Dit was not quantified.

In this chapter, sulfur passivation of GeSn surface at room temperature was investigated and implemented in the fabrication of Ge0.83Sn0.17 p-channel MOSFETs. To study the impact of sulfur passivation on the quality of high-k dielectric/GeSn interface, extensive X-ray photoelectron spectroscopy (XPS) analysis was carried out. Sulfur passivation is found to be effective in suppressing the formation of Sn oxides and Ge oxides, and Sn surface segregation. In addition, sulfur passivation helps to reduce the high-k dielectric/GeSn interface trap density Dit as extracted using the conductance method. Material study of nickel stanogermanide [Ni(GeSn)] contact formation at low temperatures was also performed for low resistivity [Ni(GeSn)] S/D contact. The sulfur-passivated Ge0.83Sn0.17 p-MOSFETs exhibit smaller subthreshold swing S, higher intrinsic transconductance Gm,int, and higher effective hole mobility μeff as compared to the non-passivated control. At a high inversion carrier density of 1 × 1013 cm−2, sulfur passivation enhances μeff by 25% as compared with the non-passivated control.

Advertisement

2. Experiment

2.1. Material growth and characterization of Ge0.83Sn0.17 substrate

The high quality Ge0.83Sn0.17 sample was grown using molecular beam epitaxy (MBE). 4-inch (001)-oriented Ge wafers with n-type doping concentration of 5 × 1016 cm−3 were used as the starting substrates. After the cyclic cleaning of Ge substrates using dilute hydrofluoric acid (DHF) (HF:H2O = 1:50) and deionized (DI) water, the unintentionally p-type doped Ge0.83Sn0.17 film was grown on the Ge substrate using the solid source low temperature MBE system [32, 33]. The growth temperature was set at 100°C. 99.9999% pure Ge and 99.9999% pure Sn were used as Ge and Sn sources, respectively. The growth chamber has a base pressure of 3 × 10−10 Torr. Ge0.83Sn0.17 film with the thickness of 10 nm was grown on the Ge substrates. Figure 1 shows the 5 × 5 μm AFM scan of the as-grown Ge0.83Sn0.17 surface. The surface is very smooth with a root-mean-square (RMS) roughness as small as 0.198 nm. High-resolution transmission electron microscopy (HRTEM) was employed to analyze the crystalline quality of the as-grown Ge0.83Sn0.17 sample. Figure 2(a) shows a low magnification cross-sectional TEM (XTEM) image of an as-grown Ge0.83Sn0.17 sample, indicating that the GeSn surface is smooth. The GeSn layer thickness is ~10 nm. The high resolution TEM (HRTEM) image in Figure 2(b) shows the smooth Ge0.83Sn0.17 surface. In addition, very clear lattice fringes and defect-free GeSn/Ge interface can be observed, as shown in the HRTEM image of Figure 2(c).

Figure 1.

5 × 5 μm AFM scan of the as-grown Ge0.83Sn0.17 substrate. The GeSn surface is very smooth with a RMS roughness as small as 0.198 nm.

Figure 2.

(a) Low magnification XTEM image of an as-grown Ge0.83Sn0.17 sample showing the smooth GeSn surface. High magnification XTEM images of the Ge0.83Sn0.17 sample shows (c) the zoom-in view of smooth GeSn surface and (d) defect-free Ge0.83Sn0.17/Ge interface.

High resolution X-ray diffraction (HRXRD) was also used to analyze the Sn composition and strain property of the as-grown Ge0.83Sn0.17 substrate. Figure 3(a) shows the (004) ω-2θ rocking scan of the as-grown sample. Both Ge and GeSn peaks are well-defined. The peak at smaller 2θ value is the GeSn peak. The relative broad full-width-half-maximum (FWHM) is due to the thin GeSn layer thickness (~10 nm). (115) reciprocal space mapping (RSM) of an as-grown Ge0.83Sn0.17/Ge (001) sample is shown in Figure 3(b). The GeSn film is fully strained to the Ge substrate and the substitutional Sn composition is calculated to be 17% from XRD measurement.

Figure 3.

(a) (004) rocking scan of the as-grown sample shows both the Ge0.83Sn0.17 and Ge peaks. The well-defined GeSn peak indicates the good crystalline quality of the GeSn layer. The peak is relatively broader than the Ge peak because of the thin layer thickness of the GeSn layer. (b) (115) RSM showing that the Ge0.83Sn0.17 film is fully strained to the Ge (001) substrate. The substitutional Sn composition is calculated to be 17%. Device fabrication of Ge0.83Sn0.17 p-MOSFETs.

Figure 4(a) summarizes the key process steps for fabricating Ge0.83Sn0.17 p-MOSFETs. After the MBE growth of ~10 nm Ge0.83Sn0.17 film on the lightly n-type doped Ge (100) substrate, pre-gate cleaning using DHF (HF:H2O = 1:50) and DI water was performed. Two splits were introduced: one with 10 minutes sulfur passivation using (NH4)2S solution (24% by weight) at room temperature (25°C) and the other one without sulfur passivation. After that, the samples were loaded into the atomic layer deposition (ALD) chamber immediately to avoid surface oxidation due to air exposure. Surface treatment was done using Trimethylaluminum (TMA) as precursor with pulse duration of 30 ms. This was followed by deposition of 3 nm-thick hafnium dioxide (HfO2) at 250°C using Tetrakis (dimethylamido) hafnium and H2O as precursors. The total ALD process duration including the pumping and venting steps is ~ 15 min. After that, 110 nm-thick tantalum nitride (TaN) was deposited using a reactive sputtering system. The metal gate was then patterned by photolithography and etched using chlorine (Cl2)-based plasma. A 10 nm-thick nickel (Ni) was then deposited using e-beam evaporator and the self-aligned Ni(GeSn) metallic contact was formed by rapid thermal annealing (RTA) at 250°C for 30 s in the N2 ambient. Finally, excess Ni was removed by selective wet etch using concentrated sulfuric acid (H2SO4) (96% by weight). The maximum processing temperature of the whole fabrication process was 250°C to limit out-diffusion of Sn to the channel surface or into the gate dielectric. A top-view scanning electron microscopy (SEM) image of a completed Ge0.83Sn0.17 p-MOSFET with a gate length LG of 4 μm is shown in Figure 4(b). Figure 4(c) shows a HRTEM image of the transistor along the dash line A-A’ indicated in Figure 4(b).

Figure 4.

(a) Process flow for fabricating Ge0.83Sn0.17 channel p-MOSFETs where a sulfur passivation step was introduced prior to the deposition of high-k gate dielectric (HfO2). The maximum processing temperature is 250°C. (b) Top-view SEM image showing a completed Ge0.83Sn0.17 p-MOSFET with self-aligned NiGeSn S/D contacts. (c) HRTEM image of a Ge0.83Sn0.17 channel p-MOSFET as seen in a cross-section along the dash line AA’ in (b).

Advertisement

3. Results and discussion

3.1. Sulfur-passivated gate stack study

The (001) surface of a diamond-structure semiconductor has two dangling bonds per surface atom. GeSn grown on Ge (001) surface has a (001) surface as shown in the atomic structure in Figure 5(a) viewed into the [110] direction. One monolayer (ML) of a Group VI element can passivate all the dangling bonds by occupying the bridge site in a (1 × 1) geometry [34, 35]. Sulfur atoms could obtain an ideal (1 × 1) termination of the bivalent (001) surfaces of silicon and Ge. Although sulfur could desorb from the Si surface at room temperature or diffuse into the Si bulk during heating [36], Weser et al. found that an ordered (1 × 1) structure with one sulfur atom bonded on a bulk-like bridge site could be achieved by introducing elemental sulfur atoms on the Ge (001) surface under ultrahigh vacuum (UHV) condition [34, 35]. The formation of Ge-S-Ge bridge bonds after a treatment in (NH4)2S solution has also been reported based on various characterization techniques, such as photoelectron spectroscopy [37], ion scattering spectroscopy [38], as well as X-ray standing wave measurements [39]. Similarly, sulfur passivation should also be able to passivate the GeSn (001) surface through the formation of S-Ge and S-Sn covalent bonds which suppress the formation of Ge and Sn oxides at the surface, as illustrated in the atomic schematic shown in Figure 5(b). In this Section, the effectiveness of sulfur passivation on the gate stack of GeSn p-MOSFETs is investigated using XPS. The interface trap density value is also extracted using conductance method and compared with the non-passivated control.

Figure 5.

Side view into the [110] direction of (a) non-passivated (1 × 1) and (b) the sulfur-passivated GeSn (001) surfaces.

3.2. XPS study on the effect of sulfur passivation

To investigate the interfacial property between the high-k dielectric and Ge0.83Sn0.17 after sulfur passivation, XPS measurement was carried out to study the change of the interfacial chemical bonds. Two blanket Ge0.83Sn0.17 samples were prepared for the measurement. After the cyclic DHF (1:50) cleaning, one of the sample went through 10 minutes aqueous (NH4)2S solution (24% by weight) and the other one did not. After that, an ultra-thin (~1 nm) HfO2 layer was deposited by ALD on these two samples. The HfO2 layer thickness should be smaller than the XPS information depth [40]. XPS characterization was then performed using a VG ESCALAB 220i–XL imaging XPS system. Monochromatic aluminum (Al) Kα X-ray (1486.7 eV) was used to obtain the core level spectra of these samples. Binding energy was calibrated with standard samples for some pure metals. The binding energy of Carbon (C) 1 s from adventitious hydrocarbon surface contamination was set at 285.0 eV for further charge correction.

In order to confirm the incorporation of S after the (NH4)2S passivation, core level XPS spectra of the S 2p peak were captured for HfO2-capped Ge0.83Sn0.17 blanket samples with and without sulfur passivation, as shown in Figure 6(a). The black curve represents the S 2p signal obtained from the non-passivated Ge0.83Sn0.17 sample. The circles are the raw data points obtained from the sulfur-passivated sample. Gaussian and Lorentzian line shapes with a Shirley background subtraction were used to fit the raw data. The blue line is the overall fitting of the core level spectra and the gray lines are the fitted peak components. For the S 2p core level spectra, the well-resolved two peaks correspond to S 2p1/2 (163.4 ± 0.02 eV) and S 2p3/2 (162.0 ± 0.02 eV) [41]. The S 2p signal obtained from the sulfur-passivated Ge0.83Sn0.17 sample indicates that S is introduced onto the GeSn surface by the (NH4)2S treatment and is still present after the deposition of HfO2. Figure 6(b) shows N 1 s core level spectra for both samples. The N 1 speak (399.0 ± 0.02 eV) [41] on the sulfur-passivated sample is not observed, indicating that nitrogen is not incorporated. The sulfur passivation layer thickness is calculated using two different photoelectron peaks of the same element at kinetic energies E1 and E2 (E1 and E2 have sufficiently large differences in λ) [42, 43]. Ge 3d and Ge 2p3 data (not shown) are used and the equation is shown below:

d = cos θ λ A E 1 λ A E 2 λ A E 1 λ A E 2 ln I B ' E 1 I B ' E 2 , E1

where λA(E1) and λA(E2) are the attenuation length of Ge 3d and Ge 2p3, I’B(E1) and I’B(E2) are corrected photoelectron intensities, and θ is take-off angle. The sulfur passivation layer thickness is calculated to be 0.44 nm using the Ge 3d and Ge 2p3 XPS signals.

Figure 6.

(a) S 2p core level XPS spectra for ~1 nm HfO2-capped Ge0.83Sn0.17 samples with and without sulfur passivation. The circles, blue lines and gray lines are the raw data points, the overall fitting curves, and the S 2p1/2 or S 2p3/2 peak components, respectively. The black curve represents the S 2p signal obtained from the non-passivated sample. (b) N 1s core level spectra for both samples.

Figure 7(a) and (b) show the Sn 3d core level spectra of HfO2-capped Ge0.83Sn0.17 blanket samples with and without sulfur passivation, respectively. The circles, blue lines, and gray lines shown in Figure 7 are the raw data points, the overall fitting curves, and fitted peak components, respectively. Due to spin orbit splitting, two separated Sn 3d peaks (Sn 3d3/2 and Sn 3d5/2) can be observed on both samples. The left shoulders, binding energy at 486.7 ± 0.2 eV and 494.9 ± 0.2 eV, can be attributed to the formation of Sn oxides (SnOx) [41]. Similarly, Figure 7(c) and (d) show the Ge 3d core level spectra of HfO2-capped Ge0.83Sn0.17 blanket samples with and without sulfur passivation, respectively. Ge oxides (GeO2 and GeOx) can also be observed in both samples. The Ge and Sn oxide signals can be detected on both samples and could come from two sources: (1) ALD HfO2 deposition as H2O pulses were introduced in the chamber with a temperature of 250°C, (2) sample transfer before loading into the XPS chamber as the samples were exposed to the air ambient. However, the intensities of both Ge oxides and SnOx are reduced significantly after the sulfur passivation, indicating the effectiveness of sulfur passivation in suppressing both Ge oxides and SnOx formation.

Figure 7.

Sn 3d and Ge 3d core-level spectra of (a) and (c) sulfur passivated and (b) and (d) non-passivated Ge0.83Sn0.17 samples obtained by XPS. A 1 nm-thick HfO2 was deposited on both samples.

The native Ge and Sn oxides formation at the high-k/GeSn interface could result in high Dit value and gate leakage current. Lee et al. reported that the native Ge oxide could react with Ge at the interface and form GeO which is easily desorbed during thermal processing [44].

GeO 2 + Ge 2 GeO . E2

This could generate a huge amount of interface states which degrade the gate stack quality [45]. The Sn oxide could also be detrimental to the GeSn gate stack as SnO2 is known to exhibit metallic behaviour, which leads to high gate leakage current [46, 47]. Therefore, suppressing the Ge and Sn oxides formation is important for achieving good gate quality for Ge0.83Sn0.17 p-MOSFETs.

To quantify the impact of sulfur passivation on Ge oxides and SnOx formation at the high-k/GeSn interface, angle-resolved XPS (ARXPS) was performed. Both SnOx and Ge oxide signals can be detected. The ratio of oxidized Sn (or Ge) atoms to the total Sn (or Ge) atoms can be calculated using

γ Sn O = A Sn O A Sn Sn + A Sn O , E3
γ Ge O = A Ge O A Ge Ge + A Ge O , E4

where ASn-O, ASn-Sn, AGe-O, and AGe-Ge are the normalized Sn-O peak area, normalized Sn-Sn peak area (also including Sn-Ge bonding), normalized Ge-O peak area, and normalized Ge-Ge peak area (also including Ge-Sn bonding), respectively [48]. With consideration of Scofield photoionization cross-sections and the transmission function of the spectrometer, γSn-O and γGe-O can be plotted as a function of the photoelectron take-off angle θ and are shown in Figure 8. The inset in Figure 8 illustrates the definition of θ. θ of 0°, 30°, 45°, and 60° were used. It is observed that γSn-O and γGe-O increase with increasing θ. This is due to the fact that more Ge and Sn atoms at the surface get oxidized than those at the sub-surface. For larger θ, the information depth is smaller and ARXPS becomes more surface sensitive. For Ge oxides, the oxide percentages of the sulfur-passivated Ge0.83Sn0.17 sample increase from 20 to 42% when θ increases from 0 to 60°. However, all the values are 10–20% smaller than those of the non-passivated one. A similar trend is also observed for SnOx, and the sulfur-passivated Ge0.83Sn0.17 sample shows more than 50% reduction in SnOx percentage than the non-passivated one at all take-off angles. The reduction of oxide formation is more obvious in Sn atoms than Ge atoms. This reveals that sulfur passivation is more effective in suppressing Sn oxide formation than Ge oxide formation. The reduction of both Ge and Sn oxides can be ascribed to the formation of S-Ge and S-Sn bonds on the sample surface. Since both samples went through the DHF treatment, most native oxides were removed and the sample surface becomes H-terminated. As a result, the Ge0.83Sn0.17 sample surface has Ge-H, Sn-H bonds, and possibly Ge-O and Sn-O bonds due to the incomplete surface oxide removal in DHF [49]. After sulfur passivation, Ge-H bond (bond energy: 263 kJ/mol [50]) and Sn-H bond (bond energy: 264 kJ/mol) are replaced by more stable Ge-S bond (bond energy: 534 kJ/mol) and Sn-S bond (bond energy: 467 kJ/mol), respectively. The S passivation layer formed at the GeSn surface can suppress the further oxidation of sub-surface Ge and Sn atoms.

Figure 8.

γSn-O and γGe-O calculated from angle-resolved XPS measurement for both the sulfur-passivated and non-passivated GeSn samples as a function of photoelectron take-off angle θ. The inset shows the definition of θ, which is set to be 0°, 30°, 45°, or 60° in the measurements.

To further investigate the effect of sulfur passivation on the interface quality between the high-k dielectric and Ge0.83Sn0.17, the extent of surface segregation of Sn atom was analyzed using the obtained ARXPS data. The Ge and Sn atomic concentrations can be calculated using the stabilized Ge 3d and Sn 3d spectra. The atomic concentrations of Sn (γSn/(Ge + Sn)) and Ge (γGe/(Ge + Sn)) can be expressed as

γ Sn / Ge + Sn = A Sn Sn + A Sn O A Sn Sn + A Sn O + A Ge Ge + A Ge O , E5
γ Ge / Ge + Sn = A Ge Ge + A Ge O A Sn Sn + A Sn O + A Ge Ge + A Ge O , E6

and plotted as a function of θ. Figure 9 shows γGe/(Ge + Sn) and γSn/(Ge + Sn) near the surface of Ge0.83Sn0.17 as a function of θ for both the sulfur-passivated and non-passivated samples. The calculated γSn/(Ge + Sn) increases with the increase of θ, indicating that surface segregation of Sn occurred in both samples. The Sn composition of the non-passivated Ge0.83Sn0.17 sample even exceeds 20% at θ of 60°. This is because Sn tends to segregate toward the surface, with the severity increasing at higher Sn compositions. Wang et al. reported that Sn segregation can occur at a temperature as low as 200°C for strained Ge0.915Sn0.085 grown on Ge [24]. Since our Ge0.83Sn0.17 sample went through the ALD deposition process with a temperature of 250°C, Sn segregation could also occur. Although the segregation of Sn occurs on both GeSn samples, the calculated γSn/(Ge + Sn) of the sulfur-passivated GeSn sample is smaller than that of the non-passivated one at all take-off angles. The S passivation layer appears to suppress the underlying Sn atoms from segregating to the surface and from further oxidation during ALD. The good integrity of high-k dielectric/GeSn interface maintained by sulfur passivation through the prevention of Sn out-diffusion and interfacial oxidation may help to improve the carrier transport characteristics in transistors.

Figure 9.

The Ge and Sn atomic concentrations at surface region of Ge0.83Sn0.17 as a function of photoelectron take-off angle θ, as determined by angle-resolved XPS.

3.3. Extraction of interface trap density

In order to extract the Dit of HfO2/Ge0.83Sn0.17 interfaces with and without sulfur passivation, Ge0.83Sn0.17 MOS capacitors (MOSCAPs) with 4 nm-thick HfO2 were fabricated. TaN and Al were deposited as the front gate and backside metals by reactive sputtering, respectively. Low temperature C-V measurement with frequencies ranging from 10 kHz to 1 MHz was performed on the Ge0.83Sn0.17 capacitors. Dit was extracted using the conductance method [51]. At a particular gate bias, the peak of the Gp versus frequency curve can be obtained at one sweeping frequency and is referring to the maximum of per-cycle energy loss. The per-cycle energy loss is due to charge trapping and detrapping at certain oxide-semiconductor interface and its maximum occurs when the energy level of the trap states is aligned with the semiconductor surface Fermi-level. The value of Dit can be extracted using the following equation:

D it = 2.5 G p / ω max qA , E7

where (Gp)max is the peak energy loss value, q is the electronic charge, and A is the area of capacitor. The band-gap of fully compressively strained Ge0.83Sn0.17 on Ge (100) substrate is ~0.45 eV [52]. Dit values from the valence band edge to the midgap of GeSn for both sulfur-passivated and non-passivated GeSn capacitors are extracted and plotted as a function of energy in the GeSn band-gap as shown in Figure 10. For the sample with sulfur passivation, Dit of 1013 cm−2·eV−1 was obtained at E-Ev of 0.13 eV. This is much smaller as compared with the non-passivated sample which has Dit of 6 × 1013 cm−2·eV−1. In addition, sulfur passivation also leads to significant reduction in Dit near the valence band edge. Sulfur passivation suppresses Ge and Sn oxide formation and Sn out-diffusion, leading to the reduction of Dit. As a result, S of the sulfur-passivated GeSn p-MOSFETs is improved as compared with the non-passivated sample. In terms of the Ge0.83Sn0.17 p-MOSFETs, high density interface traps near the valence band edge can be charged when the device is biased to strong inversion, and degrade the effective hole mobility. In order to further improve the effective hole mobility of the Ge0.83Sn0.17 p-MOSFETs, further optimization and significant improvement are needed to reduce the Dit near the valence band for the high-k/GeSn gate stack.

Figure 10.

Dit distribution from valence band edge to the midgap of GeSn as a function of energy. The sulfur-passivated GeSn sample demonstrates reduced midgap Dit as compared to the non-passivated control.

3.4. Electrical characterization of Ge0.83Sn0.17 p-MOSFETs

IDS - VGS curves of Ge0.83Sn0.17 p-MOSFETs with and without sulfur passivation are shown in Figure 11(a). The blue circles represent the sulfur-passivated sample and the black circles are the data points of the non-passivated one. Both devices have LG of 4 μm and gate width WG of 100 μm. The sulfur-passivated GeSn p-MOSFET exhibits S of 100 mV/decade. This is also the smallest reported S for any GeSn p-MOSFETs (non-passivated control shows S of 118 mV/decade). Figure 11(b) shows the output characteristics of the same devices in Figure 11(a). 10% on-state current enhancement was demonstrated by sulfur-passivated Ge0.83Sn0.17 p-MOSFET as compared to non-passivated control. Drive current of 32 μA/μm was achieved at a gate over drive of −1.0 V and VDS of −1.0 V by the sulfur-passivated device. Table 1 benchmarks S of the sulfur-passivated Ge0.83Sn0.17 p-MOSFETs realized in this work with other GeSn p-MOSFETs reported using various passivation techniques [12, 13, 14, 15, 16, 17, 18, 53]. Despite the highest Sn composition, the Ge0.83Sn0.17 p-MOSFET realized in this work shows the smallest S as compared with the other GeSn p-MOSFETs. This could be attributed to the relative low Dit at the mid-gap (3.4 × 1012 cm−2 eV−1) as compared with the other passivation techniques (7–9 × 1012 cm−2 eV−1) [17]. This indicates the high quality of the Ge0.83Sn0.17 film grown by MBE which was maintained throughout the fabrication process using low processing temperatures, as well as the Ge and Sn oxides formation between Ge0.83Sn0.17 and high-k dielectric enabled by sulfur passivation. However, the Dit value near the valence band is still high (~1 × 1013 cm−2 eV−1) as shown in the Dit plot in Figure 10. This may degrade the effective hole mobility which will be discussed in the following sections.

Figure 11.

(a) IDS - VGS curves of the sulfur-passivated Ge0.83Sn0.17 p-MOSFET show S of 100 mV/decade and ION/IOFF ratio of more than 3 orders of magnitude. S of the sulfur-passivated sample is smaller than that of the non-passivated one. (b) IDS - VDS curves of the same devices in (a).

Work Passivation Technique Sn composition (%) S value
[12] Si passivation 5.3 250
[15] Sulfur passivation 4.2 220
[53] Si passivation 3 113
[13] No passivation 3 250
[16] Si passivation 4.2 135
[17] Si passivation 3 158
[14] Si passivation 8 198
[18] Ge capping 9 160
This work Sulfur passivation 17 100

Table 1.

S values of GeSn p-MOSFETs with different Sn compositions and passivation techniques.

Figure 12 shows the capacitance C as a function of gate voltage VGS for the sulfur-passivated Ge0.83Sn0.17 p-MOSFET (LG = 8 μm, WG = 100 μm) measured at frequency of 50 kHz, 100 kHz, and 1 MHz. The schematic in the inset illustrates the configuration for C-V measurement. A quantum-mechanical C-V simulator [54] was used to fit the measured inversion C-V curve at 100 kHz and the simulated data were plotted using solid line in Figure 12. In the C-V simulator, the C-V characteristics are obtained through the calculation of hole and electron distributions by solving Schrödinger’s and Poisson’s equations self-consistently with the Fermi-Dirac distribution. In the simulation, the heavy hole effective mass of 0.27 m0 and light hole effective mass of 0.025 m0 (m0 is the free electron mass) were used for Ge0.83Sn0.17 channel [11]. From the simulated C-V curve, the equivalent oxide thickness (EOT) is extracted to be 7.5 Å. Figure 13 shows the forward and backward inversion C-V sweeps of one Ge0.83Sn0.17 p-MOSFET (LG = 8 μm, WG = 100 μm) measured at a frequency of 100 kHz. The hysteresis is small, which indicates good dielectric quality with low density of oxide traps.

Figure 12.

C vs. VGS plot of a sulfur-passivated Ge0.83Sn0.17 p-MOSFET measured at frequency of 50 kHz, 100 kHz, and 1 MHz. The measured data points are plotted as symbols. The solid curve is obtained using a quantum-mechanical C-V simulator to fit 100 kHz C-V curve. The inset shows the C-V measurement configuration.

Figure 13.

Forward and backward inversion C-V sweeps at 100 kHz for one Ge0.83Sn0.17 p-MOSFET with a LG of 8 μm. The hysteresis is small.

The Gm,int curves versus VGS at VDS of −0.05 V for both sulfur-passivated and non-passivated devices are shown in Figure 14. LG is 4 μm. Gm,int is extracted using:

G m , in t = G m , ext 1 0.5 R SD G m , ext , E8

where Gm,ext is the measured extrinsic transconductance and RSD is the source/drain resistance. Higher peak Gm,int was achieved in sulfur-passivated Ge0.83Sn0.17 p-MOSFET as compared with that of the non-passivated control. Improvement in Gm,int is attributed to better HfO2/GeSn interface achieved using sulfur passivation.

Figure 14.

Intrinsic transconductance Gm,int vs. VGS characteristics for Ge0.83Sn0.17 p-MOSFETs with and without sulfur passivation at VDS = −0.05 V. The LG of the device is 4 μm and WG is 100 μm.

The μeff of Ge0.83Sn0.17 p-MOSFETs with and without sulfur passivation is extracted using the split C-V method:

μ eff = 1 W G Q inv Δ R Total Δ L G , E9

where Qinv is the inversion charge density in the GeSn channel, and ΔRTotal/ΔLG is the slope of total resistance (RTotal) versus LG. Qinv can be obtained by integrating the measured inversion C-V curve as shown in Figure 12. Using this approach, the impact of RSD on extraction of hole mobility is taken out. Figure 15 shows the extracted μeff versus the inversion carrier density (Ninv) for both the sulfur-passivated and non-passivated Ge0.83Sn0.17 p-MOSFETs. The sulfur-passivated Ge0.83Sn0.17 p-MOSFET shows a peak hole mobility of 478 cm2/V·s at Ninv of ~2 × 1012 cm−2. At Ninv of 1 × 1013 cm−2, 25% higher hole mobility is achieved by the sulfur-passivated Ge0.83Sn0.17 p-MOSFET as compared with the non-passivated one. This is consistent with the peak intrinsic transconductance results shown in Figure 14.

Figure 15.

μeff vs. Ninv for Ge0.83Sn0.17 p-MOSFETs with and without sulfur passivation. The impact of RSD on μeff extraction was taken out using the inset equation through the total resistance slope method.

Advertisement

4. Conclusion

Sulfur passivation and low temperature process modules are developed and used in the fabrication of Ge0.83Sn0.17 p-MOSFET. Reduction in S and improvement in peak Gm,int and μeff are observed for the sulfur-passivated Ge0.83Sn0.17 p-MOSFETs as compared with the non-passivated control. This is attributed to the effective suppression of Ge and Sn oxides formation, and suppression of Sn surface segregation by sulfur passivation. In addition, the effect of sulfur passivation on Dit reduction is also investigated. It is observed that sulfur passivation reduces the Dit from the valence band edge to midgap of GeSn. As a result, the lowest S of 100 mV/decade is achieved by the sulfur-passivated Ge0.83Sn0.17 p-MOSFETs. Dit level of 1013 cm−2 eV−1 in the sulfur-passivated sample is still very high. Further improvement to significantly reduce Dit is needed to increase the hole mobility.

References

  1. 1. Pillarisetty R. Academic and industry research progress in germanium nanodevices. Nature. 2011;479(7373):324-328
  2. 2. Pillarisetty R, Chu-Kung B, Corcoran S, Dewey G, Kavalieros J, Kennel H, Kotlyar R, Le V, Lionberger D, Metz M, Mukherjee N, Nah J, Rachmady W, Radosavljevic M, Shah U, Taft S, Then H, Zelick N, Chau R. High mobility strained germanium quantum well field effect transistor as the p-channel device option for low power (Vcc = 0.5 V) III-V CMOS architecture. IEEE IEDM Technical Digest. 2010:150-153
  3. 3. Mitard J, Jaeger BD, Leys FE, Hellings G, Martens K, Eneman G, Brunco DP, Loo R, Lin JC, Shamiryan D, Vandeweyer T, Winderickx G, Vrancken E, Yu CH, Meyer KD, Caymax M, Pantisano L, Meuris M, Heyns MM. Record ION/IOFF performance for 65nm Ge pMOSFET and novel Si passivation scheme for improved EOT scalability. IEEE IEDM Technical Digest. 2008:873-876
  4. 4. Weber O, Bogumilowicz Y, Ernst T, Hartmann JM, Ducroquet F, Andrieu F, Dupre C, Clavelier L, Royer CL, Cherkashin N, Hytch M, Rouchon D, Dansas H, Papon AM, Carron V, Tabone C, Deleonibus S. Strained Si and Ge MOSFETs with high-k/metal gate stack for high mobility dual channel CMOS. IEEE IEDM Technical Digest. 2005:137-140
  5. 5. Xie R, Phung TH, He W, Sun Z, Yu M, Cheng Z, Zhu C. High mobility high-k/Ge PMOSFETs with 1 nm EOT-new concept on interface engineering and interface characterization. IEEE IEDM Technical Digest. 2008:393-396
  6. 6. Chern W, Hashemi P, Teherani JT, Yu T, Dong Y, Xia G, Antoniadis DA, Hoyt JL. High mobility high-k-all-around asymmetrically-strained germanium nanowire trigate p-MOSFETs. IEEE IEDM Technical Digest. 2012:387-390
  7. 7. Hashemi P, Chern W, Lee HS, Teherani JT, Zhu Y, Gonsalvez J, Shahidi GG, Hoyt JL. Ultrathin strained-Ge channel p-MOSFETs with high-k/metal gate and sub-1-nm equivalent oxide thickness. IEEE Electron Device Letters. 2012;33(7):943-945
  8. 8. Wu H, Si M, Dong L, Zhang J, Ye PD. Ge CMOS: Breakthroughs of nFETs (Imax = 714 mA/mm, Gmax = 590 mS/mm) by recessed channel and S/D. IEEE Symposium on VLSI Technology Digest. 2014:96-97
  9. 9. Gong X, Zhou Q, Owen MHS, Xu X, Lei D, Chen S-H, Tsai G, Cheng C-C, Lin Y-R, Wu C-H, Ko C-H, Yeo Y-C. InAlP-capped (100) Ge nFETs with 1.06 nm EOT: Achieving record high peak mobility and first integration on 300 mm Si substrate. IEEE IEDM Technical Digest. 2014:231-234
  10. 10. Lee CH, Lu C, Nishimura T, Nagashio K, Toriumi A. Thermally robust CMOS-aware Ge MOSFETs with high mobility at high-carrier densities on a single orientation Ge substrate. IEEE Symposium on VLSI Technology Digest. 2014:144-145
  11. 11. Low KL, Yang Y, Han G, Fan W, Yeo Y-C. Electronic band structure and effective mass parameters of Ge1-xSnx alloys. Journal of Applied Physics. 2012;112(10):103715
  12. 12. Han G, Su S, Zhan C, Zhou Q, Yang Y, Wang L, Guo P, Wang W, Wong CP, Shen ZX, Cheng B, Yeo Y-C. High-mobility germanium-tin (GeSn) p-channel MOSFETs featuring metallic source/drain and sub-370 °C process modules. IEEE IEDM Technical Digest. 2011:402-404
  13. 13. Gupta S, Chen R, Magyari-Kope B, Lin H, Yang B, Nainani A, Nishi Y, Harris JS, Saraswat KC. GeSn technology: Extending the Ge electronics roadmap. IEEE IEDM Technical Digest. 2011:398-401
  14. 14. Liu M, Han G, Liu Y, Zhang C, Wang H, Li X, Zhang J, Cheng B, Hao Y. Undoped Ge0.92Sn0.08 quantum well pMOSFETs on (001), (011) and (111) substrates with in situ Si2H6 passivation: High hole mobility and dependence of performance on orientation. IEEE Symposium on VLSI Technology Digest. 2014:100-101
  15. 15. Wang L, Su S, Wang W, Gong X, Yang Y, Guo P, Zhang G, Xue C, Cheng B, Han G, Yeo Y-C. Strained germanium–tin (GeSn) p-channel metal-oxide-semiconductor field-effect-transistors (p-MOSFETs) with ammonium sulfide passivation. Solid-State Electronics. 2013;83:66-70
  16. 16. Gong X, Han G, Bai F, Su S, Guo P, Yang Y, Cheng R, Zhang D, Zhang G, Xue C, Cheng B, Pan J, Zhang Z, Tok ES, Antoniadis D, Yeo Y-C. Germanium-tin (GeSn) p-channel MOSFETs fabricated on (100) and (111) surface orientations with sub-400 °C Si2H6 passivation. IEEE Electron Device Letters. 2013;34(3):339-341
  17. 17. Guo P, Han G, Gong X, Liu B, Yang Y, Wang W, Zhou Q, Pan J, Zhang Z, Tok ES, Yeo Y-C. Ge0.97Sn0.03 p-channel metal-oxide-semiconductor field-effect transistors: Impact of Si surface passivation layer thickness and post metal annealing. Journal of Applied Physics. 2013;114(4):044510
  18. 18. Huang YS, Huang CH, Lu FL, Lin CY, Ye HY, Wong IH, Jan SR, Lan HS, Liu CW, Huang YC, Chung H, Chang CP, Chu SS, Kuppurao S. Record high mobility (428cm2/V-s) of CVD-grown Ge/strained Ge0.91Sn0.09/Ge quantum well p-MOSFETs. IEEE IEDM Technical Digest. 2016:822-825
  19. 19. Liu Y, Yan J, Wang H, Zhang Q, Liu M, Zhao B, Zhang C, Cheng B, Hao Y, Han G. Strained GeSn p-channel metal-oxide-semiconductor field-effect transistors with in situ Si2H6 surface passivation: Impact of Sn composition. IEEE Transactions on Electron Devices. 2014;61(11):3639-3645
  20. 20. Gurdal O, Desjardins P, Carlsson JRA, Taylor N, Radamson HH, Sundgren J-E, Greene JE. Low-temperature growth and critical epitaxial thicknesses of fully strained metastable Ge1-xSnx (x ≤ 0.26) alloys on Ge(001)2×1. Journal of Applied Physics. 1998;83(1):162-170
  21. 21. Oehme M, Kostecki K, Schmid M, Oliveira F, Kasper E, Schulze J. Epitaxial growth of strained and unstrained GeSn alloys up to 25% Sn. Thin Solid Films. 2014;557:169-172
  22. 22. Johll H, Samuel M, Koo RY, Kang HC, Yeo Y-C, Tok ES. Influence of hydrogen surface passivation on Sn segregation, aggregation, and distribution in GeSn/Ge(001) materials. Journal of Applied Physics. 2015;117(20):205302
  23. 23. Li H, Cui YX, Wu KY, Tseng WK, Cheng HH, Chen H. Strain relaxation and Sn segregation in GeSn epilayers under thermal treatment. Applied Physics Letters. 2013;102(25):251907
  24. 24. Wang W, Li L, Zhou Q, Pan J, Zhang Z, Tok ES, Yeo Y-C. Tin surface segregation, desorption, and island formation during post-growth annealing of strained epitaxial Ge1-xSnx layer on Ge(001) substrate. Applied Surface Science. 2014;321:240-244
  25. 25. Wang W, Li L, Tok ES, Yeo Y-C. Self-assembly of tin wires via phase transformation of heteroepitaxial germanium-tin on germanium substrate. Journal of Applied Physics. 2015;117(22):225304
  26. 26. Merckling C, Sun X, Shimura Y, Franquet A, Vincent B, Takeuchi S, Vandervorst W, Nakatsuka O, Zaima S, Loo R, Caymax M. Molecular beam deposition of Al2O3 on p-Ge(001)/Ge0.95Sn0.05 heterostructure and impact of a Ge-cap interfacial layer. Applied Physics Letters. 2011;98(19):192110
  27. 27. Zhao M, Liang R, Wang J, Xu J. Improved electrical properties of Ge metal-oxide-semiconductor devices with HfO2 gate dielectrics using an ultrathin GeSnOx film as the surface passivation layer. Applied Physics Letters. 2013;102(14):142906
  28. 28. Kato K, Taoka N, Asano T, Yoshida T, Sakashita M, Nakatsuka O, Zaima S. Formation of high-quality oxide/Ge1-xSnx interface with high surface Sn content by controlling Sn migration. Applied Physics Letters. 2014;105(12):122103
  29. 29. Frank MM, Koester SJ, Copel M, Ott JA, Paruchuri VK, Shang H, Loesing R. Hafnium oxide gate dielectrics on sulfur-passivated germanium. Applied Physics Letters. 2006;89(11):112905
  30. 30. Xie R, Zhu C. Effects of sulfur passivation on germanium MOS capacitors with HfON gate dielectric. IEEE Electron Device Letters. 2007;28(11):976-979
  31. 31. Sioncke S, Ceuppens J, Lin D, Nyns L, Delabie A, Struyf H, De Gendt S, Müller M, Beckhoff B, Caymax M. Atomic layer deposition of Al2O3 on S-passivated Ge. Microelectronic Engineering. 2011;88(7):1553-1556
  32. 32. D’Costa VR, Wang W, Zhou Q, Tok ES, Yeo Y-C. Above-bandgap optical properties of biaxially strained GeSn alloys grown by molecular beam epitaxy. Applied Physics Letters. 2014;104(2):022111
  33. 33. D’Costa VR, Wang W, Zhou Q, Chan TK, Osipowicz T, Tok ES, Yeo Y-C. Compositional dependence of optical critical point parameters in pseudomorphic GeSn alloys. Journal of Applied Physics. 2014;116(5):053520
  34. 34. Weser T, Bogen A, Konrad B, Schnell R, Schug C, Steinmann. Photoemission surface core-level study of sulfur adsorption on Ge(100). Physical Review B. 1987;35(15):8184
  35. 35. Weser T, Bogen A, Konrad B, Schnell R, Schug C, Moritz W, Steinmann W. Chemisorption of sulfur on Ge(100). Surface Science. 1988;201(1-2):245-256
  36. 36. Han M, Luo Y, Camillone N, Osgood R. Reaction of H2S with Si(100). The Journal of Physical Chemistry B. 2000;104(28):6576-6583
  37. 37. Maeda T, Takagi S, Ohnishi T, Lippmaa M. Sulfur passivation of Ge(001) surfaces and its effects on Schottky barrier contact. Materials Science in Semiconductor Processing. 2006;9(4):706-710
  38. 38. Fleischmann C, Sioncke S, Couet S, Schouteden K, Beckhoff B, Müller M, Hönicke P, Kolbe M, Van Haesendonck C, Meuris M, Temst K, Vantomme A. Towards passivation of Ge(100) surfaces by sulfur adsorption from a (NH4)2S solution: A combined NEXAFS, STM and LEED study. Journal of the Electrochemical Society. 2011;158(5):H589-H594
  39. 39. Lyman P, Sakata O, Marasco D, Lee T-L, Breneman K, Keane D, Bedzyk M. Structure of a passivated Ge surface prepared from aqueous solution. Surface Science. 2000;462(1):L594-L598
  40. 40. Watts JF, Wolstenholme J. Compositional depth profiling. In: An Introduction to Surface Analysis by XPS and AES. England: John Wiley & Sons, Ltd; 2005. pp. 79-111
  41. 41. NIST X-ray Photoelectron Spectroscopy Database, Version 4.1. National Institute of Standards and Technology, Gaithersburg 2012; http://srdata.nist.gov/xps/
  42. 42. Bernstein RW, Grepstad JK. XPS intensity analysis for assessment of thickness and composition of thin overlayer films: Application to chemically etched GaAs(100) surfaces. Surface and Interface Analysis. 1989;14(3):109-114
  43. 43. Mihailova T, Velchev N, Krastev V, Marinova T. XPS study of residual oxide layers on p-GaAs surfaces. Applied Surface Science. 1997;120(3):213-219
  44. 44. Lee CH, Tabata T, Nishimura T, Nagashio K, Kita K, Toriumi A. Ge/GeO2 interface control with high pressure oxidation for improving electrical characteristics. ECS Transactions. 2009;19(1):165-173
  45. 45. Kita K, Takahashi T, Nomura H, Suzuki S, Nishimura T, Toriumi A. Control of high-k/germanium interface properties through selection of high-k materials and suppression of GeO volatilization. Applied Surface Science. 2008;254(19):6100-6105
  46. 46. Schulte-Braucks C, von den Driesch N, Glass S, Tiedemann A, Breuer U, Besmehn A, Hartmann J-M, Ikonic Z, Zhao Q, Mantl S. Low temperature deposition of high-k/metal gate stacks on high-Sn content (Si)GeSn-alloys. ACS Applied Materials & Interfaces. 2016;8(20):13133-13139
  47. 47. Kılıç Ç, Zunger A. Origins of coexistence of conductivity and transparency in SnO2. Physical Review Letters. 2002;88(9):095501
  48. 48. Wang L, Wang W, Zhou Q, Pan J, Zhang Z, Tok ES, Yeo Y-C. Post-growth annealing of germanium-tin alloys using pulsed excimer laser. Journal of Applied Physics. 2015;118(2):025701
  49. 49. Onsia B, Caymax M, Conard T, De Gendt S, De Smedt F, Delabie A, Gottschalk C, Heyns MM, Green M, Lin S, Mertens P, Tsai W, Vinckier C. On the application of a thin ozone based wet chemical oxide as an interface for ALD high-k deposition. Solid State Phenomena. 2005;103:19-22
  50. 50. Luo Y-R. Comprehensive Handbook of Chemical Bond Energies. CRC press; 2007
  51. 51. Nicollian EH, Goetzberger A. The Si-SiO, interface-electrical properties as determined by the metal-insulator-silicon conductance technique. The Bell System Technical Journal. 1967;46(6):1055-1033
  52. 52. Yahyaoui N, Sfina N, Lazzari JL, Bournel A, Said M. Band engineering and absorption spectra in compressively strained Ge0.92Sn0.08/Ge (001) double quantum well for infrared photodetection. Physica Status Solidi C. 2014;11(11-12):1561-1565
  53. 53. Liu B, Zhan C, Yang Y, Cheng R, Guo P, Zhou Q, Kong EYJ, Daval N, Veytizou C, Delprat D, Nguyen BY, Yeo Y-C. Germanium multiple-gate field-effect transistor with in situ boron-doped raised source/drain. IEEE Transactions on Electron Devices. 2013;60(7):2135-2141
  54. 54. Yang K, King Y-C, Hu C. Quantum effect in oxide thickness determination from capacitance measurement. IEEE Symposium on VLSI Technology Digest. 1999:77-78

Written By

Dian Lei and Xiao Gong

Submitted: 08 November 2017 Reviewed: 26 January 2018 Published: 05 November 2018