OPEN Nanoscale Nitrogen Doping in Silicon by Self...

26
1 SCIENTIFIC REPORTS | 5:12641 | DOI: 10.1038/srep12641 www.nature.com/scientificreports Nanoscale Nitrogen Doping in Silicon by Self-Assembled Monolayers Bin Guan 1 , Hamidreza Siampour 1 , Zhao Fan 1 , Shun Wang 2 , Xiang Yang Kong 3 , Abdelmadjid Mesli 4 , Jian Zhang 5 & Yaping Dan 1 This Report presents a nitrogen-doping method by chemically forming self-assembled monolayers on silicon. Van der Pauw technique, secondary-ion mass spectroscopy and low temperature Hall effect measurements are employed to characterize the nitrogen dopants. The experimental data show that the diffusion coefficient of nitrogen dopants is 3.66 × 10 15 cm 2 s 1 , 2 orders magnitude lower than that of phosphorus dopants in silicon. It is found that less than 1% of nitrogen dopants exhibit electrical activity. The analysis of Hall effect data at low temperatures indicates that the donor energy level for nitrogen dopants is located at 189 meV below the conduction band, consistent with the literature value. e successful development of the complementary metal-oxide-semiconductor (CMOS) technology in the past decades has generated a huge impact on the human society by offering devices with higher performances at lower cost 1 . e main driving force of this development is the constant downscaling of CMOS transistors, which, however, has become increasingly difficult. One of the main issues is the short channel effect that the transistor gate loses control as the size becomes smaller, resulting in malfunctions. Currently, the CMOS industry solves this issue by using a 3-dimensional “FIN” structure that allows the gate grabbing around the channel to enhance its control 2 . is solution unfortunately increases the complexity of the fabrication processes. e cost per device increases instead of continuously decreasing as Moore’s Law predicts 3 . A second solution to alleviate the short channel effect is to form ultra-shallow junctions in the source and drain regions of the transistor. is is one of the main motivations of “gold rush” for developing tran- sistors based on the 2-dimensional atomic monolayers such as graphene and MoS 2 , which will facilitate the formation of atomically thin junctions 4–6 . Silicon, as the most commonly used substrate in CMOS industry, is unlikely to be replaced by any new materials in the foreseeable future 7 . With silicon as the substrate, it has been difficult to form ultra-shallow junctions on the silicon surface by the traditional ion implantation doping process. is is because the physical bombardment of accelerated ions turns the single-crystalline silicon surface (about 10 nm thick) into low quality amorphous silicon, resulting in significant degradation in the device performances 8 . In recent years, doping by self-assembled monolay- ers (SAMs) has proven to be a mild and controllable doping process 9,10 . e process consists of chemical assembly of dopant-containing molecules on silicon surfaces and subsequent thermal treatment that drives dopants to diffuse into the substrate. To form ultra-shallow junctions, spike thermal treatment is oſten required for this process 11 . A sub-10 nm junction can be formed for fast diffusing dopants like phosphorus by limiting the annealing time to as short as a few seconds. However, a very short pulse of high temperature might lead to problems such as low reproducibility. It is more reliable to form 1 University of Michigan - Shanghai Jiao Tong University Joint Institute. 2 Key Laboratory of Artificial Structures and Quantum Control (Ministry of Education), Department of Physics and Astronomy. 3 School of Materials Science and Engineering, Shanghai Jiao Tong University, Shanghai, 200040, China. 4 Institut Matériaux Microélectronique Nanosciences de Provence, UMR 6242 CNRS, Université Aix-Marseille, 13397 Marseille Cedex 20, France. 5 Faculty of Medicine, Shanghai Jiao Tong University, Shanghai, 200040, China. Correspondence and requests for materials should be addressed to J.Z. (email: [email protected]) or Y.D. (email: [email protected]) Received: 12 February 2015 Accepted: 06 July 2015 Published: 31 July 2015 OPEN

Transcript of OPEN Nanoscale Nitrogen Doping in Silicon by Self...

Page 1: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

1Scientific RepoRts | 5:12641 | DOi: 10.1038/srep12641

www.nature.com/scientificreports

Nanoscale Nitrogen Doping in Silicon by Self-Assembled MonolayersBin Guan1, Hamidreza Siampour1, Zhao Fan1, Shun Wang2, Xiang Yang Kong3, Abdelmadjid Mesli4, Jian Zhang5 & Yaping Dan1

This Report presents a nitrogen-doping method by chemically forming self-assembled monolayers on silicon. Van der Pauw technique, secondary-ion mass spectroscopy and low temperature Hall effect measurements are employed to characterize the nitrogen dopants. The experimental data show that the diffusion coefficient of nitrogen dopants is 3.66 × 10−15 cm2 s−1, 2 orders magnitude lower than that of phosphorus dopants in silicon. It is found that less than 1% of nitrogen dopants exhibit electrical activity. The analysis of Hall effect data at low temperatures indicates that the donor energy level for nitrogen dopants is located at 189 meV below the conduction band, consistent with the literature value.

The successful development of the complementary metal-oxide-semiconductor (CMOS) technology in the past decades has generated a huge impact on the human society by offering devices with higher performances at lower cost1. The main driving force of this development is the constant downscaling of CMOS transistors, which, however, has become increasingly difficult. One of the main issues is the short channel effect that the transistor gate loses control as the size becomes smaller, resulting in malfunctions. Currently, the CMOS industry solves this issue by using a 3-dimensional “FIN” structure that allows the gate grabbing around the channel to enhance its control2. This solution unfortunately increases the complexity of the fabrication processes. The cost per device increases instead of continuously decreasing as Moore’s Law predicts3.

A second solution to alleviate the short channel effect is to form ultra-shallow junctions in the source and drain regions of the transistor. This is one of the main motivations of “gold rush” for developing tran-sistors based on the 2-dimensional atomic monolayers such as graphene and MoS2, which will facilitate the formation of atomically thin junctions4–6. Silicon, as the most commonly used substrate in CMOS industry, is unlikely to be replaced by any new materials in the foreseeable future7. With silicon as the substrate, it has been difficult to form ultra-shallow junctions on the silicon surface by the traditional ion implantation doping process. This is because the physical bombardment of accelerated ions turns the single-crystalline silicon surface (about 10 nm thick) into low quality amorphous silicon, resulting in significant degradation in the device performances8. In recent years, doping by self-assembled monolay-ers (SAMs) has proven to be a mild and controllable doping process9,10. The process consists of chemical assembly of dopant-containing molecules on silicon surfaces and subsequent thermal treatment that drives dopants to diffuse into the substrate. To form ultra-shallow junctions, spike thermal treatment is often required for this process11. A sub-10 nm junction can be formed for fast diffusing dopants like phosphorus by limiting the annealing time to as short as a few seconds. However, a very short pulse of high temperature might lead to problems such as low reproducibility. It is more reliable to form

1University of Michigan - Shanghai Jiao Tong University Joint Institute. 2Key Laboratory of Artificial Structures and Quantum Control (Ministry of Education), Department of Physics and Astronomy. 3School of Materials Science and Engineering, Shanghai Jiao Tong University, Shanghai, 200040, China. 4Institut Matériaux Microélectronique Nanosciences de Provence, UMR 6242 CNRS, Université Aix-Marseille, 13397 Marseille Cedex 20, France. 5Faculty of Medicine, Shanghai Jiao Tong University, Shanghai, 200040, China. Correspondence and requests for materials should be addressed to J.Z. (email: [email protected]) or Y.D. (email: [email protected])

Received: 12 February 2015

accepted: 06 July 2015

Published: 31 July 2015

OPEN

Page 2: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

www.nature.com/scientificreports/

2Scientific RepoRts | 5:12641 | DOi: 10.1038/srep12641

ultra-shallow junctions using dopants with low thermal diffusion coefficient, because such dopants will allow extended periods of annealing time. Nitrogen, similar to phosphorus, is also a group V element acting as a donor-type dopant in silicon. Previous research on nitrogen diffusion in silicon has showed that nitrogen dopants introduced by implantation and rapid thermal annealing have low diffusivity12,13. Nitrogen doping by SAMs can potentially form ultra-shallow or even atomically thin junctions in silicon. It thus may provide a better solution to the short channel effect and allow for the continuous downscaling of the CMOS technology.

In this Report we investigate the properties of nitrogen doping by SAMs that carry doping elements such as phosphorus and nitrogen. The chemical composition of the SAMs and the doped silicon substrate were examined using X-ray photoelectron spectroscopy (XPS) and secondary-ion mass spectrometry (SIMS). Van der Pauw and Hall effect measurements were used to characterize the electronic properties of the dopants. We demonstrate the successful doping of nitrogen and phosphorus in intrinsic silicon wafer, extract the diffusion coefficients for both dopants and estimate the donor energy levels in the monolayer doped silicon samples.

Results and DiscussionsThe preparation of SAMs on silicon followed the so-called hydrosilylation strategy, as shown in Fig. 1. The hydrosilylation process on silicon, first reported by Linford and Chidsey14,15, refers to the addition of silicon hydride onto an unsaturated carbon-to-carbon bond. It has been considered the most robust method for producing covalent Si− C bound organic monolayers on silicon16–18. Hydride-terminated

Figure 1. Schematic illustration of surface modification (SAMs formation) and monolayer doping on silicon.

Page 3: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

www.nature.com/scientificreports/

3Scientific RepoRts | 5:12641 | DOi: 10.1038/srep12641

surfaces were reacted with a nitrogen-containing alkene species 1 to form a silicon-carbon bound mon-olayer under UV light radiation. The strategies to modify the silicon surfaces with other molecules are shown in the supplementary Fig. S1. Figure 2 shows XPS spectra acquired on the silicon surface with alk-ene species 1 modification. The survey spectrum indicates the presence of Si, C and O, which is in good agreement with the presence of an organic monolayer on the silicon substrate. The oxygen 1s emission at 532 eV is ascribed to oxygen from the monolayer, as well as adventitiously adsorbed oxygen. Oxygen does not usually give reliable signals in XPS due to the contamination by the ambient atmosphere. However, both carbon and nitrogen can provide useful information on the nature of surface-bound species. The narrow scan of the C 1s region (Fig. 2b) shows a broad band at 285 eV (1.4 eV FWHM), attributed to aliphatic carbon-carbon (C− C) bonding. The side shoulder of the C 1s band at 286.5 eV (1.6 eV FWHM) is assigned to carbon-oxygen (C− O) and carbon-nitrogen (C− N) bonds from the molecule 1. The peak at 289 eV (1.6 eV FWHM) accounts for the carbonyl (C= O) in the terminal group. These observed bind-ing energies are consistent with those reported for carbon species on Si(100) surfaces19. Moreover, the representative peak area ratio of C− C to C− O/C− N shows an experimental value of 2.88:1, which is in agreement with those predicted by the stoichiometry of the atoms on the putative surface product (2.5:1). The narrow scan spectrum of N 1s in Fig. 2c displays one component at 400.5 eV (1.6 eV FWHM), corre-sponding to the nitrogen atom at tert-Butoxycarbonyl (t-Boc) group of alkene 1. Therefore, it is suggested that an alkene species 1 monolayer is formed on the silicon surface. We also employed XPS to analyse other SAMs formed on silicon according to the scheme in supplementary Fig. S1. Some oxides have been detected on the modified silicon surface (supplementary Fig. S2), although it was reported that an oxide-free surface after hydrosilylation on Si(100) could be achieved20–22. The data and discussions can be found in the supplementary Fig. S2 and Fig. S3.

The monolayer modified samples were then coated with a 20-nm thick layer of silicon dioxide by atomic layer deposition (ALD) and subsequently annealed at 1050 °C for 2 min (Fig.  1). To evaluate the nitrogen doping process, Van der Pauw technique was employed to measure the sheet resistance of the silicon samples at room temperature. Table 1 lists the experimental results of the silicon samples with and without nitrogen doping, as well as the sample with both nitrogen and phosphorus doping

Figure 2. (a) XPS survey spectra of the alkene species 1 monolayer on silicon. (b) Narrow scan of the C 1s region. (c) Narrow scan of the N 1s region.

Page 4: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

www.nature.com/scientificreports/

4Scientific RepoRts | 5:12641 | DOi: 10.1038/srep12641

(for the I-V curves, refer to the supplementary Fig. S4). The sheet resistance of the original undoped silicon substrate was measured to be 143 ± 21 kΩ . A piece of silicon wafer labeled as “blank sample” was selected to undergo all the capping and annealing processes except the SAM formation. The sample shows a sheet resistance as high as 121 ± 13 kΩ , indicating negligible contamination from the capping SiO2 and the ambient environment of the annealing chamber. A control sample with the absence of doping atoms was prepared to test the possible influence of C, O and H to the doping process. No significant change was observed from the control sample with only C monolayer modification (see the scheme in Supplementary Fig. S1). This means that C and O atoms, when sitting at their natural sites, substitutional for carbon and interstitial for oxygen, are electrically inactive in silicon. The samples that were modified with the monolayers containing N or both N and P (Supplementary Fig. S1) experience a significant drop in the sheet resistance. Clearly, the drop in the sheet resistance is solely contributed by N or both N and P that are carried by the SAMs. Electrostatic force microscopy (EFM) was also applied on the monolayer doped silicon samples to obtain the static charge profile across the samples. The EFM images in Supplementary Fig. S5 showed no clear aggregation of dopants across the samples, suggesting a relatively uniform doping.

To access to the doping profiles, the doped samples were analysed by SIMS, as shown in Fig.  3. In sample #1 with only nitrogen as dopant, the dopant concentration drops rapidly from ~4 × 1019 cm−3 near the surface to ~1017 cm−3 within a distance of 100 nm. In sample #2 with both nitrogen and phos-phorus doping, nitrogen shows the same trend of diffusion residing in the top 100 nm of the substrate. Phosphorus dopants, however, diffuse deeper than nitrogen to 200 nm below the interface, leading to a lower dopant concentration near the surface than nitrogen as required by the mass conservation. These dopants are carried by the SAMs, the nature of which limits the total number of initial dopants on the surface. After high temperature annealing, the dopants will follow the limited-source diffusion process, which is governed by the equation below:

π( , ) =

( )

N x tN

DtxDt

exp2 1

02

where N0 is the initial surface concentration, x the diffusion distance, t the annealing time and D the diffusion constant which is temperature dependent.

By fitting equation  (1) into the curves in Fig.  3, we find that the diffusion coefficient for nitrogen is 3.66 × 10−15 cm2 s−1 in sample #1 and 1.39 × 10−14 cm2 s−1 in sample #2. The diffusion coefficient for phosphorus in sample #2 is 1.69 × 10−13 cm2 s−1 which is comparable to the widely observed value for phosphorus in silicon23,24, but two orders of magnitude larger than that of nitrogen in intrinsic silicon (sample #1). The diffusion of nitrogen follows a complex behavior in which defects distribution and some other factors during annealing process play key roles25,26. It is likely that the diffusion of nitrogen dopants can be enhanced by interaction with the faster diffusing phosphorus dopants. This is in line with the aforementioned values showing that the diffusion coefficient of nitrogen in the phosphorus-contained silicon (sample #2) is relatively larger than that in sample #1 without phosphorus.

High temperature annealing aims at accomplishing two steps: (i) driving the dopant atoms from the surface towards the bulk; (ii) activating them electrically by allowing doping atoms to find substitutional

Si(100) Samples Sheet Resistance Rs

(kΩ)

Blank sample: intrinsic silicon sample without any surface modification 121 ± 13

Control: sample modified with alkyl

monolayer

116 ± 12

Sample #1: modified with monolayer containing nitrogen

35.6 ± 0.2

Sample #2: modified with monolayer containing both nitrogen and phosphorus

36.6 ± 0.3

Table 1. Sheet resistances measured by Van der Pauw technique.

Page 5: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

www.nature.com/scientificreports/

5Scientific RepoRts | 5:12641 | DOi: 10.1038/srep12641

sites in the lattice. The first step is verified by SIMS, while the second is measured by Van der Pauw tech-nique and Hall effect measurement. The comparison of the two approaches allows for informing about the activated versus total fraction of introduced dopants.

From the SIMS data in Fig. 3, we can calculate the total amount of dopant atoms that diffuse from the surface into the bulk. In sample #1 the average nitrogen dopant concentration (nd) per unit area is 2.36 × 1013 cm−2, and the average nitrogen and phosphorus dopant concentration per unit area in sample #2 are 9.35 × 1012 cm−2 and 4.3× 1012 cm−2, respectively. The amount of nitrogen diffusing into substrate in sample #2 outnumbers that of phosphorus by a factor of around 2, which is consistent with the fact that ~65% of phosphate molecules are coupled onto nitrogen modified surface (see supplementary Fig. S3). If the nitrogen dopants were 100% electrically active in sample #1, the sheet resistance of the sample would be around 188 Ω . This is lower by a factor of about 200 than the measured sheet resistance listed in Table 1, indicating that the nitrogen dopants were not fully electrically activated. The activation rate can be estimated by the ratio 188/35517 = 0.53%. Similar conclusions can be reached for sample #2. However, it is not straightforward to find the exact activation rate of the dopants from the resistivity measurements, because the carrier mobility is dependent on the concentration of the activated dopants that are highly non-uniform in our case. We therefore turn to the Hall effect measurements that can directly provide the carrier concentration without the information of carrier mobility.

For Hall effect measurements, the Hall voltage (VH) is governed by the following equation:

= ( )VR IB

d 2HH

where RH is the Hall coefficient which is equal to −ne1 for n-type semiconductors with the electron con-

centration n and unit charge e, I the current flow, B the magnetic field, and d the thickness of the sample.

Considering that the electron concentration is non-uniform from the surface to the bulk, Hall resist-ance can be written in the following expression in which nc is the electron concentration per unit area.

∫ ∫= = − = −

( )= −

( )= −

( )

VI

Rd

Bned

Bn x edx

Be n x dx

B Ben

1 1 1

3

H Hd d

c0 0

Figure 4 displays the Hall effect measurement results on both samples in Hall bar geometry at 300 K. The Hall resistance is linearly correlated with the magnetic field. It is worth pointing out that the meas-ured value of Hall resistance is influenced by the longitudinal misalignment and the thermoelectric volt-age between the two Hall contacts. Hence in Fig. 4 Hall resistance shows a background number when the magnetic field is zero, which however does not influence the extraction of the free carrier concentration. From the slope of the linear correlation, the free electron concentration per unit area (nc) is extracted to be 1.36 × 1011 cm−2 in sample #1 and 2.1× 1011 cm−2 in sample #2. Compared to the SIMS data, we find that the activation rate for N in sample #1 is only 0.58% (a value very close to the crude estimation 0.53% given above), and the average activation rate for N and P in sample #2 is approximately 1.54%. Less than 1% activation rate of nitrogen in our case is consistent with the research in 1970s27, as the fraction of substitutional nitrogen is small. It is known that phosphorus implanted silicon has a high degree of electrical activity compared to nitrogen. However, it was reported that the presence of nitrogen could retard phosphorus activation by forming inactive complexes of phosphorus and nitrogen atoms at a ratio

Figure 3. Nitrogen and phosphorus doping profiles in sample #1 (a) and sample #2 (b). Inset graphs show the nitrogen and phosphorus concentration in log scale.

Page 6: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

www.nature.com/scientificreports/

6Scientific RepoRts | 5:12641 | DOi: 10.1038/srep12641

of 1:128. That might be the cause of low activation rate for sample #2. It also has been shown previously that P dopants carried by SAMs (not containing N) have an ionization rate over 90%9. We conclude that the low activation rate is not intrinsic to this method, but highly dependent on other impurities in the substrate. For example, it was shown that the activation rate of N dopants by ion implantation could be significantly increased in a boron doped silicon substrate29. Those boron atoms residing at the lattice sites of silicon can be displaced by silicon self-interstitials and later nitrogen atoms can fill the vacancies to render nitrogen donors. As a result, the activation rate of nitrogen could be significantly improved. The work using boron-doped silicon substrates to increase the electrical activation of nitrogen dopants by the SAM method is under investigation.

To further investigate the properties of dopants, we performed low temperature Hall effect measure-ments, from which the activation energy of the dopants can be found. We first extracted the average free electron concentrations (nc) at different temperatures from the curves of Hall resistance dependence on magnetic field for both N doped sample and N&P doped sample (see the supplementary Table S1), which are similar to those in Fig. 4 except that they show some nonlinearity at low temperatures. The details are illustrated in the supplementary Fig. S6 and S7. Then we plot the surface concentration of nc as a function of kT in Fig. 5, where k is the Boltzmann constant and T is the absolute temperature. The dopant distribution is highly non-uniform within the thin-doped layer near the surface. But the average concentration of electrons (nc), holes (pc) and ionized dopants (nD

+) will always remain charge neutral. On the assumption that the holes are negligibly small in concentration compared to the electrons, which is true in our case, we can derive the electron concentration nc that is correlated with the temperature as the following expression (see the supplementary materials for details):

( )( )

=− + +

( )

Δ

Δn

N N N N exp

exp

8

4 4c

c c c DE

kT

EkT

2

Figure 4. Hall resistance dependence on the magnetic field at 300 K for sample #1 (a) and sample #2 (b).

Figure 5. Temperature dependence of the average free electron concentration for sample #1 and #2 .

Page 7: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

www.nature.com/scientificreports/

7Scientific RepoRts | 5:12641 | DOi: 10.1038/srep12641

where Nc is the effective density of states function which is defined as ( )≈ = ( )π //

N w kT2cm kTh

2 3 23 2n

2 with w being the constant related to the band structure of the semiconductor, ND the concentration of donors, and ΔE the activation energy which is equal to (Ec − Ed) with Ec and Ed being the conductance band edge and the donor energy level, respectively.

By fitting equation  (4) into the curve in Fig.  5 for N doped sample, we can obtain the activation energy of the nitrogen dopants as 189 meV and some other parameters as listed in Table  2. In the N and P doped sample, there are two types of donors to generate free electrons. As phosphorus dopants in silicon have much lower ionization energy than nitrogen, which is ~44 meV30, P atoms that replace Si lattice remain complete ionization within the temperature range of our measurements. For the N and P sample, equation (4) can be rewritten as

=− + +

+( )

Δ

Δn

N N N N exp

expN

8

4 5c

c c c DE

kTE

kT

P

2

where all the parameters have the same meaning as in equation (4) except that NP is the average concen-tration of completely ionized P donors.

The fitting results of equation  (5) are also displayed in Fig.  5 and Table  2. The concentration of N donors (ND) and the activation energy for N atoms (ΔE) derived from the two equations show similar results, ~3.4× 1011 cm−2 and ~189 meV, respectively. The activation energy we obtained here shows a deep energy level existing for N dopants in Si, which is in good agreement with the literature31.

According to the fitting values in Table 2, the concentration of N donors in N-doped sample is 2-order magnitude lower than the total nitrogen concentration detected by SIMS, indicating that only about 1% of the nitrogen atoms diffusing into silicon bulk have substituted the Si lattice. This is again consistent with the report that more than 95% implanted nitrogen in silicon is trapped on some non-substituional sites or defects in the form of molecular N2, silicon nitride precipitates or randomly distributed atomic nitrogen32. Similarly, only 1.4% of phosphorus atoms have substituted the Si lattice. The rest of P atoms may form the inactive nitrogen-phosphorus complexes with N atoms as mentioned previously28.

In conclusion, we have successfully demonstrated that the nitrogen-containing SAMs on intrinsic silicon surface leads to nitrogen doping in the substrate with the assistance of high temperature anneal-ing. The activation rate and diffusion coefficient of nitrogen dopants were found to be consistent with literature values. As both parameters are related to defects such as silicon interstitials in the substrate, nitrogen doping with a higher activation rate might be achieved on p-type Si substrates. Based on the Van der Pauw and Hall effect measurements, a deep donor energy level at 189 meV below the conduction band was found for N doped silicon, which is associated with substitutional nitrogen. In future, DLTS (deep level transient spectroscopy) measurements will be employed to obtain further information about dopants and defects in nitrogen-doped silicon.

MethodsWafer cleaning. Float Zone single-side polished silicon wafers, (100) surface orientation ((100) ± 0.05°), 500 ± 25 μ m thick, 9000–15000 Ω cm in resistivity, were cleaved into 1 cm by 1 cm pieces and cleaned with acetone and ethanol of CMOS grade in a sonication bath for 5 min, respectively. After rinsed with DI water, the Si samples were immersed in “piranha solution” (98% H2SO4:30% H2O2, 3:1 (v/v)) for 30 minutes at 90 °C, followed by rinsing with DI water again. The wafers were then etched in 2.5% hydrofluoric acid (HF) solution for 90 seconds to remove the oxide layer and rinsed with DI water, followed by drying under N2 stream.

UV-mediated hydrosilylation and surface functionalization. The freshly etched Si(100) sam-ples were immediately placed in the reaction tube, covered with 30–50 μ L of 50 mg/mL tert-Butyl N-Allylcarbamate (TCI America, > 98%) methanol solution and were exposed to the irradiation of a 254 nm UV lamp (0.36 mW/cm2) in an inert nitrogen atmosphere for 3 h, affording sample #1. Control

Fitting results Parameter ValueStandard

error

N doped sample (sample #1), fitted with equation (4)

ND (× 109 cm−2) 334 30

ΔE (eV) 0.189 0.003

N and P dope sample (sample #2), fitted with equation (5)

ND (× 109 cm−2) 348 47

ΔE (eV) 0.188 0.007

NP (× 109 cm−2) 61.8 4.6

Table 2. Fitting results of equations (4) and (5) to the plots of nc as the function of kT.

Page 8: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

www.nature.com/scientificreports/

8Scientific RepoRts | 5:12641 | DOi: 10.1038/srep12641

experiment was conducted on freshly prepared Si samples with neat undecylenic acid (J&K Scientific Ltd., > 97%) under 254 nm UV radiation overnight. The samples were then rinsed with copious dichlo-romethane and ethanol, and dried under nitrogen gas. To obtain samples with phosphorus functionaliza-tion (sample #2, see Supplementary Fig. S1), sample #1 was subjected to 25% trifluoroacetic acid (TFA) in dichloromethane for 3 h, followed by a 10 min rinse in 10% ammonium hydroxide (NH4OH) to form the primary amine terminated surface. The surface was immersed in an ethanol solution of bifunctinal crosslinker dicyclohexylcarbodiimide (DCC, 40 mΜ , Aldrich) and mono-Dodecyl phosphate (5 mΜ , Aldrich) for 6 h, affording sample #2.

Silicon dioxide deposition and thermal annealing. SiO2 films were grown by atomic layer depo-sition on Si samples with tris(dimethylamino)silane (TDMAS) and ozone as precursors. Thermal anneal-ing at 1050 °C for 120 s, with a ramp temperature of 100 °C/min, starting from 800 °C, was performed in an argon environment in a tube furnace (Thermo scientific Lindberg/Blue, USA). After annealing, the Si samples were immersed again in 2.5% HF solution to remove SiO2 film on the surfaces.

Surface Characterization. X-ray photoelectron spectroscopy (XPS) was carried on a Kratos AXIS UltraDLD spectrometer with a monochromated Al Kα source (1486.6 eV), a hybrid magnification mode analyser and a multichannel detector at a takeoff angle of 90° from the plane of the sample surface. Analysis chamber pressure is less than 5 × 10−9 Torr. All energies are reported as binding energies in eV and referenced to the C 1s signal (corrected to 285.0 eV) for aliphatic carbon on the analysed sample surface. Survey scans were carried out selecting 250 ms dwell time and analyser pass energy of 160 eV. High-resolution scans were run with 0.1 eV step size, dwell time of 100 ms and the analyser pass energy set to 40 eV. After background subtraction using the Shirley routine, XPS spectra were fitted with a convolution of Lorentzian and Gaussian profiles by using software Casa XPS. Secondary-ion Mass Spectroscopy (SIMS) was conducted to obtain dopant profile at the top 300 nm of substrate at Evans Analytical Group, NJ, USA. EFM experiments were performed on a Nanonavi E-Sweep system (Seiko, Japan) operated in tapping mode.

Electrical Characterization. The metal contacts on silicon for electrical measurements were realized by evaporating 200-nm aluminum or chromium/gold films in a thermal evaporation system (Angstrom Engineering, Canada). Van der Pauw measurements were performed on square-shaped samples on which the metal contacts are exactly located at the four corners. The custom-made probe station is equipped with four solid tungsten probe tips (the tip size < 1 μ m). Keithley 2400 sourcemeter units and a custom-written Labview script were employed to generate and collect current/voltage data. Hall effect measurements were performed on samples in Hall bar geometry in a Physical Property Measurement System (PPMS, Quantum Design, USA).

References1. Rabaey, J. M., Chandrakasan, A. & Nikolic, B. Digital integrated circuits:a design perspective. 2nd edn, (Publishing House of

Electronics Industry, 2012).2. Choi, Y.-K., Tsu-Jae, K. & Chenming, H. Nanoscale CMOS spacer FinFET for the terabit era. IEEE Electron Device Lett. 23, 25–27

(2002).3. Barbara, S. No Moore? A golden rule of microchips appears to be coming to an end, < http://www.economist.com/news/21589080-

golden-rule-microchips-appears-be-coming-end-no-moore> (2013)(Date of access: 17/10/2014).4. Lemme, M. C., Echtermeyer, T. J., Baus, M. & Kurz, H. A Graphene Field-Effect Device. IEEE Electron Device Lett. 28, 282–284

(2007).5. Liu, H., Neal, A. T. & Ye, P. D. Channel Length Scaling of MoS2 MOSFETs. ACS Nano 6, 8563–8569 (2012).6. Liu, H. et al. Phosphorene: An Unexplored 2D Semiconductor with a High Hole Mobility. ACS Nano 8, 4033–4041 (2014).7. Thompson, S. E. & Parthasarathy, S. Moore’s law: the future of Si microelectronics. Mater. Today 9, 20–25 (2006).8. Gable, K. A., Robertson, L. S., Jain, A. & Jones, K. S. The effect of preamorphization energy on ultrashallow junction formation

following ultrahigh-temperature annealing of ion-implanted silicon. J. Appl. Phys. 97, 044501 (2005).9. Ho, J. C. et al. Controlled nanoscale doping of semiconductors via molecular monolayers. Nat. Mater. 7, 62–67 (2008).

10. Longo, R. C., Cho, K., Schmidt, W. G., Chabal, Y. J. & Thissen, P. Monolayer Doping via Phosphonic Acid Grafting on Silicon: Microscopic Insight from Infrared Spectroscopy and Density Functional Theory Calculations. Adv. Funct. Mater. 23, 3471–3477 (2013).

11. Ho, J. C. et al. Wafer-Scale, Sub-5 nm Junction Formation by Monolayer Doping and Conventional Spike Annealing. Nano Lett. 9, 725–730 (2009).

12. Itoh, T. & Abe, T. Diffusion coefficient of a pair of nitrogen atoms in float‐zone silicon. Appl. Phys. Lett. 53, 39–41 (1988).13. Voronkov, V. V. & Falster, R. Multispecies nitrogen diffusion in silicon. J. Appl. Phys. 100, 083511 (2006).14. Linford, M. R. & Chidsey, C. E. D. Alkyl monolayers covalently bonded to silicon surfaces. J. Am. Chem. Soc. 115, 12631–12632

(1993).15. Linford, M. R., Fenter, P., Eisenberger, P. M. & Chidsey, C. E. D. Alkyl Monolayers on Silicon Prepared from 1-Alkenes and

Hydrogen-Terminated Silicon. J. Am. Chem. Soc. 117, 3145–3155 (1995).16. Sung, M. M., Kluth, G. J., Yauw, O. W. & Maboudian, R. Thermal Behavior of Alkyl Monolayers on Silicon Surfaces. Langmuir

13, 6164–6168 (1997).17. Sieval, A. B. et al. Highly Stable Si− C Linked Functionalized Monolayers on the Silicon (100) Surface. Langmuir 14, 1759–1768

(1998).18. Puniredd, S. R., Assad, O. & Haick, H. Highly Stable Organic Monolayers for Reacting Silicon with Further Functionalities: The

Effect of the C− C Bond nearest the Silicon Surface. J. Am. Chem. Soc. 130, 13727–13734 (2008).

Page 9: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

www.nature.com/scientificreports/

9Scientific RepoRts | 5:12641 | DOi: 10.1038/srep12641

19. Lin, Z. et al. DNA Attachment and Hybridization at the Silicon (100) Surface. Langmuir 18, 788–796 (2002).20. Ciampi, S. et al. Functionalization of Acetylene-Terminated Monolayers on Si(100) Surfaces: A Click Chemistry Approach.

Langmuir 23, 9320–9329 (2007)21. Lavi, A., Cohen, H., Bendikov, T., Vilan, A. & Cahen, D. Si-C-bound alkyl chains on oxide-free Si: towards versatile solution

preparation of electronic transport quality monolayers. Phys. Chem. Chem. Phys. 13, 1293–1296 (2011).22. Peng, W., DeBenedetti, W. J. I., Kim, S., Hines, M. A. & Chabal, Y. J. Lowering the density of electronic defects on organic-

functionalized Si(100) surfaces. Appl. Phys. Lett. 104, 241601 (2014).23. Uematsu, M. Simulation of boron, phosphorus, and arsenic diffusion in silicon based on an integrated diffusion model, and the

anomalous phosphorus diffusion mechanism. J. Appl. Phys. 82, 2228–2246 (1997).24. Bracht, H., Silvestri, H. H., Sharp, I. D. & Haller, E. E. Self- and foreign-atom diffusion in semiconductor isotope heterostructures.

II. Experimental results for silicon. Phys. Rev. B 75, 035211 (2007).25. Hockett, R. S. Anomalous diffusion of nitrogen in nitrogen-implanted silicon. Appl. Phys. Lett. 54, 1793–1795 (1989).26. Schultz, P. A. & Nelson, J. S. Fast through-bond diffusion of nitrogen in silicon. Appl. Phys. Lett. 78, 736–738 (2001).27. Palvlov, P. V., Zorin, E. I., Tetelbaum, D. I. & Khokhlov, A. F. Nitrogen as Dopant in Silicon and Germanium. Phys. Status Solidi

(a) 35, 11–36 (1976).28. Chang, R.-D. & Lin, C.-H. Deactivation of phosphorus in silicon due to implanted nitrogen. Phys. Status Solidi (c) 11, 24–27

(2014).29. Tetelbaum, D. I., Zorin, E. I. & Lisenkova, N. V. Anomalous solubility of implanted nitrogen in heavily boron-doped silicon.

Semiconductors 38, 775–777 (2004).30. Pantelides, S. T. & Sah, C. T. Theory of localized states in semiconductors. I. New results using an old method. Phys. Rev. B 10,

621–637 (1974).31. Tokumaru, Y., Okushi, H., Masui, T. & Abe, T. Deep Levels Associated with Nitrogen in Silicon. Jpn. J. Appl. Phys. 21, L443

(1982).32. Mitchell, J. B., Pronko, P. P., Shewchun, J., Thompson, D. A. & Davies, J. A. Nitrogen− implanted silicon. I. Damage annealing

and lattice location. J. Appl. Phys. 46, 332–334 (1975).

AcknowledgementsThis research was financially supported by the national “1000 Young Scholars” program of the Chinese central government, the National Science Foundation of China (61376001), the “Pujiang Talent Program” of the Shanghai municipal government, and the 54th general fund of China Postdoctoral Science Foundation. We thank Dr. Limin Sun (Instrumental Analysis Center, Shanghai Jiao Tong University) for her help with XPS, Mr. Ang Wang and Mr. Xiaofei Zhou (University of Michigan - Shanghai Jiao Tong University Joint Institute) for wire bonding.

Author ContributionsY.D. conceived the idea and supervised the research. J.Z. advised on the monolayer surface modification. B.G. designed and performed the experiments and collected the data. B.G., Y.D. and H.S. analysed the data. F.Z. conducted the EFM experiment. X.Y.K. performed the ALD experiment. S.W. contributed to interpreting the results of Hall effect measurement. A.M. contributed to discussion of the results. B.G. and Y.D. wrote the manuscript. All authors reviewed the manuscript.

Additional InformationSupplementary information accompanies this paper at http://www.nature.com/srepCompeting financial interests: The authors declare no competing financial interests.How to cite this article: Guan, B. et al. Nanoscale Nitrogen Doping in Silicon by Self-Assembled Monolayers. Sci. Rep. 5, 12641; doi: 10.1038/srep12641 (2015).

This work is licensed under a Creative Commons Attribution 4.0 International License. The images or other third party material in this article are included in the article’s Creative Com-

mons license, unless indicated otherwise in the credit line; if the material is not included under the Creative Commons license, users will need to obtain permission from the license holder to reproduce the material. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/

Page 10: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

Supporting  Information  for    

Nanoscale  Nitrogen  Doping  in  Silicon  by  Self-­‐Assembled  Monolayers  

 

Bin  Guan  1,  Hamidreza  Siampour  1,  Zhao  Fan  1,  Shun  Wang2,  Xiangyang  Kong3,  

Abdelmadjid  Mesli  4,  Jian  Zhang  5*,  Yaping  Dan  1*  

1  University  of  Michigan  -­‐  Shanghai  Jiao  Tong  University  Joint  Institute,  

2Key  Laboratory  of  Artificial  Structures  and  Quantum  Control  (Ministry  of  

Education),  Department  of  Physics  and  Astronomy,  

3School  of  Materials  Science  and  Engineering,  

Shanghai  Jiao  Tong  University,  Shanghai,  200040,  China.  

4  Institut  Matériaux  Microélectronique  Nanosciences  de  Provence,  UMR  6242  

CNRS,  Université  Aix-­‐Marseille,  13397  Marseille  Cedex  20,  France  

5  Faculty  of  Medicine,  

Shanghai  Jiao  Tong  University,  Shanghai,  200040,  China.  

Page 11: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

 

Figure   S1.   Surface   modification   routes   for   preparing   monolayers   without  

nitrogen   (control   sample)   and   with   both   nitrogen   and   phosphorus  

functionalization  (sample  #1  and  #2).  The  freshly  etched  hydride-­‐terminated  Si  

(100)   samples   were   immersed   in   neat   deoxygenated   undecylenic   acid   3   and  

were   reacted   under   the   irradiation   of   a   254   nm  UV   lamp   in   an   inert   nitrogen  

atmosphere   for   over   16   h.   The   samples   were   then   rinsed   with   copious  

dichloromethane   and   ethanol,   and   dried   under   argon   gas,   yielding   control  

sample.  Sample  #2  was  prepared  with  further  modification  on  alkene  species  1  

modified   sample   #1.   Following   by   amine   deprotection   in   25%   trifluoroacetic  

acid,   the   sample   was   derivatized   with   phosphate   2   in   the   presence   of   cross-­‐

linker   DCC,   providing   monolayers   containing   both   nitrogen   and   phosphorus  

atoms.  

 

 

H H H H

Si (100) Si (100)

1, 3 h

254 nm UV

HN

O

O

25% TFA, 2.5 h

NH4OH, 10 min

HOP

O

OHO

9

NH

P

OO

HO

Si (100)

2, DCC, 6 h

DCC: dicyclohexylcarbodiimide

N C N

3, overnight254 nm UV

OH

O7

H H H HH H

NH

OO

NH

OO

Si (100)

HH H

NH2 NH2

H H

Si (100)

H H

OHO

7

OHO

7

1 2 3

sample 1 sample 2

control sample

Page 12: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

 

Figure  S2.  XPS  spectra  of  silicon  sample  modified  with  undecylenic  acid  (control  

sample).  a)  Survey  scan;  b)  high-­‐resolution  scan  of   the  Si  2p  region,   showing  a  

clear   Si   2p1/2−Si   2p3/2  spin−orbit-­‐split   doublet   (99.9   and   99.3   eV,   respectively)  

from   the  bulk   silicon   and   some   silicon  oxides   in   the   range  of   102-­‐104   eV.   The  

fractional  monolayer   coverage   of   oxidized   silicon,   calculated   directly   from   the  

oxidized-­‐to-­‐bulk   Si   2p   peak   area   ratio   according   to   a   conventional   model,1   is  

approximately  0.5  monolayer   equivalents.  The   silicon  oxide   is   likely   caused  by  

the  unreacted   silicon  hydride  bonds  underneath   the  monolayer  being  oxidized  

again  when  exposed   to   air. As  Si(100)   surface   is   terminated  with  a  mixture  of  

SiH,  SiH2  and  SiH3  species,  the  orientation  of  Si(100)  may  lead  to  the  formation  of  

self-­‐assembled   monolayers   (SAMs)   not   as   densely   packed   as   that   on   Si(111).  

Several  studies  have  shown  that  densely  packed  SAMs  on  silicon  can  prevent  the  

oxidation   of   silicon  more   efficiently.2-­‐4   Therefore,   it   is   easier   for   the   unreacted  

silicon   hydride   on   Si(100)   to   be   oxidized.   Nevertheless,   forming   an   oxide-­‐free  

surface   on   Si(100)   after   hydrosilylation   is   still   possible   as   reported   by   other  

Page 13: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

researchers.5-­‐7;   c)  C  1s  narrow  scan,   revealing  an  asymmetric  peak  at  285.0  eV  

attributed  to  aliphatic  C−C  carbons,  a  second  peak  at  286.5  eV  assigned  to  carbon  

adjacent  to  the  carbonyl  moiety  (−CC(O)OH)  and  a  peak  at  290.0  eV  for  carbonyl  

–C(O)OH.   Experimentally   the   peak   area   ratio   of   aliphatic   C−C   and   two   other  

peaks   is   8   :   1   :   1,   which   is   close   to   the   value   (9   :   1   :   1)   predicted   by   the  

stoichiometry  of  the  atoms  of  undecylenic  acid.  

 

 

Figure   S3.   XPS   spectra   of   chemically  modified   Si   surfaces.   a)   The   nitrogen   1s  

spectrum  of  the  N-­‐Boc-­‐allylamine  monolayer  on  silicon,  showing  a  single  peak  at  

400.5  eV.  After  treatment  in  25%  TFA  (b)  two  peaks  are  recorded.  The  NH3+  and  

Page 14: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

the  primary  amine  NH2  have  been  assigned  to  the  peaks  at  402.0  eV  and  400.2  eV,  

respectively.   This   indicates   amine-­‐terminated   surface   has   been   formed.   The  

surface  further  reacted  with  phosphate  shows  a  peak  at  400.6  eV  corresponding  

to  newly  formed  N−P  bonding  and  a  peak  at  402.0  eV  suggesting  the  unreacted  

amine   group   on   the   surface   (c).   According   to   the   ratio   of   two   peak   areas,   the  

phosphate   coupling   efficiency   can   be   estimated   as   65%.   Hence,   the   ratio   of  

nitrogen   atom   and   phosphorus   atom   on   sample   #2   is   about   1.54.   d)   The  

phosphorus  2s  spectra  of  the  N-­‐Boc-­‐allylamine  modified  sample  displays  a  broad  

peak  at  185.0  eV,  corresponding  to  silicon  plasmon  loss8.  After  the  treatment  of  

25%  TFA  and  phosphate,   a  new  peak  appears  on   the  shoulder  at  190.9  eV   (e),  

assigned  to  the  phosphorus  from  N−P  bonding  on  the  surface.  Note  that  the  peak  

area   of   N−P   component   in   N   1s   spectrum   is   equal   to   the   peak   area   in   P   2s  

spectrum.   The   above   results,   consistent   with   literature,9   demonstrate   the  

successful  modification  on  silicon  surfaces.    

 

Page 15: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

Van  der  Pauw  measurement  

 

The  shape  of  the  samples  for  resistivity  measurement  is  square  with  point  ohmic  

electrodes  contacted  at  the  four  corners.  A  voltage  is  applied  to  flow  current  IAB  

along  one  side  of  the  square  sample  and  the  voltage  along  the  opposite  side,  VCD,  

is   measured.   The   current   IAB   and   voltage   VCD   are   then   plotted   to   obtain   the  

resistance  for  one  side  RAB,CD.  After  repeating  the  process  to  the  four  sides  of  the  

square   sample,   the   sheet   resistance   of   the   sample   can   be   deduced   from   the  

following  equations.    

exp −πRx

Rs

"

#$

%

&'+ exp −

πRy

Rs

"

#$

%

&'=1    (1)  

Rx =12RAB,CD + RCD,AB( ),Ry =

12RBC,AD + RAD,BC( )      (2)  

 

Page 16: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

 

Figure   S4.   I-­‐V   curves   from   Van   der   Pauw   measurements   on   blank,   control  

samples,  sample  #1  and  sample  #2.    

 

Page 17: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

 

Figure  S5.  Electrostatic  force  microscope  images  on  control  (left)  and  monolayer  

nitrogen-­‐doped   samples   (right).   The   top   two   images   were   recorded   when   no  

voltage  was  applied  on  the  probe  tip.  The  bottom  images  were  taken  at  positive  

bias  of  1  V.  Under  positive  bias,  the  sample  with  different  doping  concentration  

patterns  induced  different  static  force  towards  the  probe  tip  and  thus  rendered  a  

phase  shift  profile.  The  image  for  nitrogen-­‐doped  sample  with  +  1  V  bias  shows  

no  abrupt  change  in  the  profile,  suggesting  the  even  distribution  of  dopants.  

 

Page 18: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

Low-­‐temperature  Hall  measurement  

 

Figure   S6.  Dependence  of   the  Hall  resistance  of   the  nitrogen-­‐doped  sample  #1  

(a)  and  both  N  and  P  doped  sample  #2  (b)  on  magnetic  flux  density  at  different  

temperature  from  300  K  to  200  K.  

 

At   low  temperatures,   the  hall  measurement  results   from  sample  #1  show  non-­‐

linear  correlation  between  measured  resistance  and  magnetic  field  in  Fig.  S6.  It  

was   due   to   the  magnetoresistance   effect,   as   the   resistance   of   a   semiconductor  

generally  increases  when  the  sample  is  placed  in  a  magnetic  field.  The  measured  

resistance  between  point  1  and  point  2  (R12)  in  the  Hall  bar  is  not  the  actual  Hall  

resistance   (RH),   but   the   combination   of   magnetoresistance   (RM)   and   RH.   To  

extract   the   actual   RH   and   free   carrier   concentration   at   low   temperature,   the  

resistance  between  point  2  and  point  3  (R23)  was  also  recorded.  For  no  Hall  effect  

should  be  detected  between  point  2  and  point  3,  R23   is   consisted  of   its  original  

resistance  (RS)  and  RM.  As  an  example,  the  experimental  results  of  R23  and  R12  in  

the  function  of  magnetic  field  for  sample  #1  at  240  K  are  showed  in  Fig.  S7.  

Page 19: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

 

Figure   S7.   Hall   resistance   Vs   Magnetoresistance.   a)   Dependence   of   resistance  

between   point   1   and   point   2   (R12)   on   the   change   of   magnetic   field.   b)  

Dependence  of  resistance  between  point  2  and  point  3  (R23)  on  magnetic  field.    

 

The   two   curves   in   Fig.   S7   were   then   fitted   with   the   following   equations,  

respectively.  

R12 = RM12 + RH = r12 a+ cx2 + ex4 + gx6 +!( )+ rH (bx + dx3 + fx5 +!)        (3)  

R23 = RM 23 + RS = r23 a+ cx2 + ex4 + gx6 +!( )+ rS (bx + dx3 + fx5 +!)        (4)  

where   x  is  the  magnetic  flux  density  and  parameters  such  as  a,  b,  c,  d,  e,  f,  g  …  are  

the   same   ones   in   equation   (3)   as   in   equation   (4).   The   fitting   result   of  

rH (bx + dx3 + fx5 +!) is   the   actual   Hall   resistance.   Hence,   the   free   carrier  

concentration  nc  can  be  obtained  by  the  equation  (5).  

rHb = −1nce,nc = −

1rHbe

     (5)  

Page 20: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

Table   S1.   Free   carrier   concentration   (nc)   for   sample   #1   and   #2   at   different  

temperature.  

Sample  #1  (nitrogen  doping)  Sample  #2  (both  nitrogen  and  

phosphorus)  

Temperature  (K)   nc  (×  1011  cm-­‐2)   Temperature  (K)   nc  (×  1011  cm-­‐2)  

300   1.36   300   2.10  

280   1.12   290   1.85  

260   0.838   280   1.72  

240   0.562   270   1.61  

220   0.369   260   1.48  

200   0.202   250   1.34  

190   0.135   240   1.21  

180   0.0790   230   1.09  

170   0.0401   220   0.987  

160   0.0155   210   0.898  

    200   0.843  

 

Page 21: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

The  donor  energy  level  in  nitrogen  doped  silicon  

In   the  monolayer  doped   silicon   samples,   the  doping   is  highly  non-­‐uniform  and  

majority   of   the   dopants   are   located   within   a   thin   layer   below   the   surface.  

However,  the  electron  concentration  still  follows  the  universal  equation  below  to  

maintain  charge  neutral:  

nc = nD+ + pc  

where  nc  is  the  average  electron  concentration,  nD+  the  average  concentration  of  

ionized  dopants  and  pc  the  average  hole  concentration.  

 

For  our  case,  the  thin  doped  layer  is  always  n-­‐type,  and  hence  nc  >>  pc.  Then,  we  

have    

nc = nD+

 

→ nc = Nc expEF −Ec

kT#

$%

&

'(=

ND

1+ 2exp EF −Ed

kT#

$%

&

'(= nD

+      (a)  

exp EF −Ec

kT"

#$

%

&'=

ncNc

 

where  Nc  is  the  effective  density  of  states  function,  Nc ≈ 22πmn

*kTh2

"

#$

%

&'

3/2

= w kT( )3/2  

with  w  being  the  constant  related  to  the  band  structure  of  the  semiconductor,  ND  

the   concentration   of   donors,  EF   the   Fermi   level,  Ec   the   conductance   band   edge  

and  Ed  the  donor  energy  level.    

 

Further,  the  right  part  of  equation  (a)  can  be  written  as:  

Page 22: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

ND

1+ 2exp EF −Ed

kT"

#$

%

&'=

ND

1+ 2exp Ec −Ed

kT"

#$

%

&'exp

EF −Ec

kT"

#$

%

&'  

 

In  the  end,  equation  (a)  can  be  rewritten  as:  

nc = Nc expEF −Ec

kT"

#$

%

&'=

ND

1+ 2exp Ec −Ed

kT"

#$

%

&'ncNc

     (b)  

 

After  further  simplifying  equation  (b)  and  solving  the  2-­‐order  equation,  we  will  

find  

)exp(4

)exp(82

kTE

kTENNNN

nDccc

c Δ

Δ++−

=

(c)  

where  the  activation  energy  ΔE  =  Ec  -­‐  Ed.  

 

Page 23: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

Sheet  resistance  results  from  a  different  type  of  wafer  

Si  (111)  wafer  as  doping  substrate  was  also  employed  into  the  nitrogen  doping  

experiment.  The  sheet  resistance  results  of  the  Si  (111)  sample  before  and  after  

monolayer  doping  is  shown  in  Table  S2.  The  result   in   line  with  that  of  Si  (100)  

also   suggests   a   successfully   nitrogen   doping   in   Si   (111)   by   self-­‐assembled  

monolayer.  

 

Table  S2.  Sheet  resistances  measured  by  Van  der  Pauw  technique  

Si  (111)  Samples  with  an  original  resistivity  of  15  kΩ.cm,  

thickness  of  300  μm  

Sheet  Resistance  Rs  

(kΩ)  

Blank  sample:  intrinsic  silicon  sample  without  any  surface  

modification  457  ±  23  

Sample  #1:  modified  with  monolayer  containing  nitrogen  

 

165  ±  27  HN

O

O

Page 24: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

Additional  SIMS  profiles  for  monolayer  doped  samples  

 

Figure  S8.  The  duplicate  nitrogen  profiles  of  the  N  doped  sample  by  SIMS,  

showing  good  reproducibility  of  the  experimental  results.  

 

 

Figure  S9.  The  duplicate  phosphorus  profiles  of  the  N  and  P  doped  sample  by  

SIMS.  

 

Page 25: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

Table  S3.  Comparison  of  SiO2  formation  methods§  

Methods  for  SiO2  formation   ALD  Spin-­‐on-­‐glass  

(SOG)  

Thermal  

evaporation  

Silicon  sample  (measured  sheet  

resistance  of  143  ±  21  kΩ)  Sheet  Resistance  Rs  (kΩ)  

Control:  unmodified  wafer  coated  

with  SiO2  and  annealed  121±13   116±6   109*  

N-­‐doped:  modified  with  N  

monolayer,  coated  with  SiO2  and  

annealed  

35.6±0.2   41.2±6.8   N.A.  

§ Only   the  data  by  ALD-­‐SiO2   capping   are  displayed   in   the  manuscript,   because  

the  experimental  data  using  SOG  and  thermal  evaporation  to   form  SiO2  coating  

are  incomplete  due  to  non-­‐scientific  reasons.    

*  Only  one  repeat  was  done  for  thermally  evaporated  SiO2.  

 

References  

1   Webb,   L.   J.   &   Lewis,   N.   S.   Comparison   of   the   Electrical   Properties   and  

Chemical   Stability   of   Crystalline   Silicon(111)   Surfaces   Alkylated   Using  

Grignard  Reagents   or  Olefins  with   Lewis  Acid  Catalysts.   J.  Phys.  Chem.  B  

107,  5404-­‐5412,(2003).  

2   Sieval,   A.   B.,   Linke,   R.,   Zuilhof,  H.  &   Sudholter,   E.   J.   R.  High-­‐quality   alkyl  

monolayers  on  silicon  surfaces.  Adv.  Mater.  12,  1457-­‐1460,(2000).  

3   Cicero,   R.   L.,   Linford,   M.   R.   &   Chidsey,   C.   E.   D.   Photoreactivity   of  

Unsaturated   Compounds   with   Hydrogen-­‐Terminated   Silicon(111).  

Langmuir  16,  5688-­‐5695,(2000).  

Page 26: OPEN Nanoscale Nitrogen Doping in Silicon by Self ...yapingd.sjtu.edu.cn/upload/editor/file/20181024/... · SITII REPORTS 5121 I 1.138sre121 1 Nanoscale Nitrogen Doping in Silicon

4   Puniredd,  S.  R.,  Assad,  O.  &  Haick,  H.  Highly  stable  organic  monolayers  for  

reacting   silicon  with   further   functionalities:   the   effect   of   the   C−   C   Bond  

nearest  the  silicon  surface.  J.  Am.  Chem.  Soc.  130,  13727-­‐13734,(2008).  

5   Ciampi,  S.  et  al.  Functionalization  of  Acetylene-­‐Terminated  Monolayers  on  

Si(100)   Surfaces:   A   Click   Chemistry   Approach.   Langmuir   23,   9320-­‐

9329,(2007).  

6   Lavi,   A.,   Cohen,   H.,   Bendikov,   T.,   Vilan,   A.   &   Cahen,   D.   Si-­‐C-­‐bound   alkyl  

chains   on   oxide-­‐free   Si:   towards   versatile   solution   preparation   of  

electronic   transport   quality   monolayers.   Phys.   Chem.   Chem.   Phys.   13,  

1293-­‐1296,(2011).  

7   Peng,  W.,  DeBenedetti,  W.  J.  I.,  Kim,  S.,  Hines,  M.  A.  &  Chabal,  Y.  J.  Lowering  

the   density   of   electronic   defects   on   organic-­‐functionalized   Si(100)  

surfaces.  Appl.  Phys.  Lett.  104,  241601,(2014).  

8   Gouzman,   I.,   Dubey,   M.,   Carolus,   M.   D.,   Schwartz,   J.   &   Bernasek,   S.   L.  

Monolayer   vs.   multilayer   self-­‐assembled   alkylphosphonate   films:   X-­‐ray  

photoelectron  spectroscopy  studies.  Surf.  Sci.  600,  773-­‐781,(2006).  

9   Strother,   T.,   Hamers,   R.   J.   &   Smith,   L.   M.   Covalent   attachment   of  

oligodeoxyribonucleotides   to   amine-­‐modified   Si   (001)   surfaces.   Nucleic  

Acids  Res.  28,  3535-­‐3541,(2000).