Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Donor incomplete ionization and mobility enhancement in ultra-thin silicon-on-insulator films doped by phosphorus end-terminated polymers

Andrea Pulici ab, Stefano Kuschlan ac, Gabriele Seguini a, Marco De Michielis a, Riccardo Chiarcos c, Michele Laus c, Marco Fanciulli ab and Michele Perego *a
aCNR-IMM, Unit of Agrate Brianza, Via C. Olivetti 2, I-20864 Agrate Brianza, Italy. E-mail: michele.perego@cnr.it
bUniversità degli Studi di Milano-Bicocca, Via Roberto Cozzi 55, I-20125 Milano, Italy
cUniversità del Piemonte Orientale ‘‘A. Avogadro’’, Viale T. Michel 11, I-15121 Alessandria, Italy

Received 8th May 2024 , Accepted 13th October 2024

First published on 14th October 2024


Abstract

Ex situ doping of ultra-thin silicon-on-insulator (SOI) substrates is performed by using polymers terminated with a doping containing moiety. The injection of P impurity atoms is investigated confining the same P dose of ∼ 1 × 1013 cm−2 in a progressively thinner device layer, with thickness values (HSOI) from 6 to 70 nm. The dopant concentration is determined by Time-of-flight secondary ion mass spectroscopy (ToF-SIMS) measurements. Sample resistivity (ρ), carrier concentration (ne) and mobility (μ) are determined combining sheet resistance and Hall measurements in van der Pauw configuration. Almost complete activation and full ionization of the injected dopants is observed at room temperature in the samples with HSOI ≥ 30 nm. The ionization fraction progressively drops to 5% when reducing the thickness of the device layer. Dopant incomplete ionization is accompanied by an increase in electron mobility, with values significantly larger than those reported for bulk Si. In the SOI samples with HSOI > 20 nm, the fraction of ionized P atoms at room temperature is perfectly described by the 3D bulk model of Altermatt et al. For HSOI ≤ 20 nm, the bulk model must be corrected to account for the effect of interface states and dielectric mismatch between Si and surrounding SiO2.


Introduction

The microelectronics industry is heading into the age of ultra-scaled stacked 3D device architectures, such as gate-all-around field-effect transistors (GAAFETs).1 These transistors, in which the gate surrounds a Si nanosheet channel, are a direct evolution of FinFETs, designed to offer better channel control, be faster, and less power-consuming.2,3 This transition poses technological and fundamental problems regarding dopant incorporation, activation and ionization in Si-based nanostructures.4 In this respect, there have been extensive theoretical and experimental studies regarding dopants in 0D and 1D silicon nanostructures,5–7 but limited results are reported on 2D systems like doped ultra-thin Si films.8,9

Ex situ doping of ultra-scaled Si nanostructures by standard ion implantation is difficult because the release of energy by the impinging ions determines severe damages in the Si crystal that are difficult to restore. To overcome the limitations associated with conventional top-down doping techniques, the use of dopant end-terminated polymers was proposed as an alternative bottom-up approach to achieve semiconductor doping.10 Polymers are employed to create a δ-layer source of dopants at the interface between a SiO2 capping layer and the underlying Si substrate.11 The dopants are driven-in via high temperature annealing in a rapid thermal processing (RTP) system. This mild approach is compatible with mass production of advanced microelectronic components and was demonstrated not to introduce significant contaminations in the sample during the process.12 In particular, the incorporation of P dopants in a silicon-on-insulator (SOI) substrate with device layer thickness ∼30 nm using this doping strategy was already investigated.11 The experimental data demonstrated that drive-in of the dopants at 1000 °C for 100 s results in a uniform dopant concentration throughout the entire device layer with high activation rates (ηa > 90%) and electrical properties compatible with those reported for a P-doped bulk Si.13 This doping technology represents a simple tool to introduce dopant impurities in ultra-thin Si films and study their activation, providing useful information to support current transition toward the next generation of microelectronics devices.

In this work, the drive-in of P impurity atoms in ultra-thin Si films was investigated confining the same P dose in SOI samples with progressively thinner device layers. Time-of-flight secondary ion mass spectrometry (ToF-SIMS) data were combined with sheet resistance and Hall measurements at room temperature to investigate the effect of the reduction of the device layer thickness (HSOI) on the activation and ionization of P impurity atoms.

Experimental

1 × 1 cm2 SOI dies were cleaved from lightly doped p-type SOI wafers with (149.6 ± 0.1) nm thick buried oxide (BOX) and (71 ± 1) nm thick Si device layer. The device layer was thinned down following an oxidation procedure described in a previous paper,11 developed to produce ultra-thin SOI films with no evidence of sample degradation associated with high temperature oxidation.11 In particular, no significant differences in terms of oxygen diffusion into the device layer were observed between the pristine and the thinned SOI samples when performing the oxidation process in O2 atmosphere at T ≤ 1100 °C. Accordingly, to prevent any possible oxygen contamination and modification of the Si/SiO2 interfaces prior to the doping process, the thinning of the device layers was performed by an oxidation process at T = 1000 °C. A P dopant δ-layer source was formed at the interface between the Si device layer and a 10 nm thick SiO2 capping layer, using a poly(methylmetachrylate) end-terminated by a P-containing moiety (PMMA-P) grafted to the substrate. PMMA-P is characterized by an average molar mass Mn = 7.5 kg mol−1 and polydispersity index Đ = 1.14. More details are reported in previous publications.10,14–16 The samples underwent a high temperature treatment in a RTP system at temperatures ranging from 900 to 1100 °C in N2 atmosphere to promote the drive-in of P atoms into the Si device layer. Finally, after removal of the SiO2 capping layer, circular aluminum metal contacts were deposited by thermal evaporation in a square van der Pauw (VDP) geometry at the corners of each sample using a shadow mask. Spectroscopic ellipsometry (SE) was used to monitor HSOI at each step of the doping process, from oxidation to metal deposition. Sample preparation was optimized to produce SOI films with HSOI ranging from 6 to 70 nm. To guarantee accurate determination of the effective HSOI values and proper correlation with electrical characteristics, the thickness of each sample was determined at the end of the process, just before the electrical characterization of the sample. Sheet resistance and Hall measurements were performed in VDP configuration.17 ToF-SIMS measurements were performed in a IONTOF IV system using Cs+ ions for sputtering and Ga+ ions for analysis. Depth scale calibration was performed by individually measuring the sputter rate in each of the SOI samples. The Si/SiO2 interfaces were determined by the spikes in 30Si signals, while HSOI was accurately determined by SE. To obtain quantitative information about the P concentration, the counts measured by the ToF-SIMS were converted into atom concentrations (atoms cm−3) following a calibration protocol fully described in a previous publication.18 In particular, P secondary ion signal was normalized on the 30Si secondary ion signal that was selected as a reference. Subsequently, the normalized P signal was calibrated using a relative sensitivity factor that was determined considering a set of samples implanted with P at the same energy but different doses in order to verify the linearity of the calibration over a broad range of concentrations.

Results and discussion

The P concentration profiles in the different SOI samples upon drive-in at 1000 °C for 100 s are reported in Fig. 1. The ToF-SIMS depth profiles of the 50 and 70 nm thick SOI samples show a P concentration gradient throughout the device layer. The observed concentration profile is correctly predicted by Fick's law of diffusion for thick SOI films.19 Conversely, when HSOI ≤ 30 nm, P atoms reach the buried Si/SiO2 interface during diffusion, starting to segregate at the interface and rediffuse into the device layer. This results in a flattening of the concentration profile, as observed by ToF-SIMS analysis, which demonstrates that the P concentration remains constant within experimental variation throughout the entire film thickness in the case of the samples with HSOI ≤ 30 nm.
image file: d4tc01886a-f1.tif
Fig. 1 ToF-SIMS P depth profiles of SOI samples with HSOI ranging from 6 to 70 nm upon drive-in at 1000 °C. Grey dashed lines indicates the position of the BOX interface.

Fig. 2a shows the total P dose (ND) injected in the device layer, computed as the integral of the P concentration depth profile obtained by ToF-SIMS analysis. Data were obtained by averaging the P doses obtained across several samples annealed in the same RTP conditions with the same nominal HSOI. In the case where just one sample was prepared, the error bar was assumed to be around 10% of the dose determined from calibrated ToF-SIMS depth profiles. In the SOI samples with HSOI > 6 nm, ND is found to be constant within experimental error. The average P dose was determined to be ND ∼ (1.0 ± 0.2) × 1013 cm−2. The ND reduction that is observed in the 6 nm thick sample suggests that, in the case of ultra-thin SOI, a fraction of P atoms is lost because of P diffusion in the BOX, in agreement with previous studies about P diffusivity in Si and SiO2 under similar annealing conditions.11 Moreover, Fig. 1 shows an increasingly higher P signal at the device layer/BOX interface when reducing HSOI, suggesting a significant P accumulation at the interface in ultra-thin films. It is worth noting that, although important matrix effects are expected to occur at the buried Si/SiO2 interface, preventing a quantitative estimation of the effective amount of P segregated at this interface by the ToF-SIMS signals, a qualitative comparison of the P concentration profiles in this region is assumed to be reliable since were obtained from samples that are nominally identical since they were cut from the same wafer and experienced the same thermal treatment. The combination of these P segregation and diffusion phenomena accounts for the low P dose in the 6 nm thick device layer.


image file: d4tc01886a-f2.tif
Fig. 2 (a) The total P dose computed from analysis of ToF-SIMS depth profiles (black) and the dose of charge carriers obtained by Hall measurements (blue) versus HSOI upon drive-in at 1000 °C. Dashed lines correspond to different dopant activation rates. (b) Sheet resistance Rs (black) and carrier mobility μ (blue) versus HSOI. In the inset, μ versus the average carrier concentration ne. Blue solid line indicates electron mobility in bulk Si.13

Fig. 2a shows the dose of free electrons (Ne) in the device layer, obtained by Hall measurements at room temperature, as a function of HSOI. In the limit of full ionization, the free carrier dose is assumed as an indicator of the dose of electrically active P atoms. Accordingly, the activation rate ηa is computed as the ratio ηa = Ne/ND. Almost complete activation and full ionization is achieved when HSOI ≥ 30 nm, consistently with previous results.11 Apparently, the activation progressively drops when decreasing HSOI below this threshold value, achieving a minimum ηa ∼ 5% when HSOI ∼ 6 nm.

Fig. 2b reports the sheet resistance (Rs) and the carrier mobility (μ) values as a function of HSOI. The carrier mobility (μ) is computed by combining Rs and Hall measurements according to the equation

μ = (qNeRs)−1,
where q is the elemental charge. The RS value progressively increases when reducing of HSOI. Conversely, the mobility progressively decreases until HSOI ∼ 20 nm, consistent with the idea of increased Coulomb scattering that correlates with the progressive increase in carrier concentration, that is observed taking into account the progressive reduction of HSOI. These mobility values are perfectly compatible with those reported for bulk Si13 as shown in the inset of Fig. 2b. Interestingly, a strong increase in electron mobility is observed when HSOI < 20 nm, with values higher than those reported for bulk Si. This trend is consistent with data in the literature for highly doped SOI samples and can be explained assuming that HSOI reduction causes a transition from 3D to quasi-2D Coulomb scattering.20

Fig. 3a depicts a schematic representation of ionized impurities and the effect of Coulomb scattering on an electron moving in a thick or ultra-thin doped SOI. When HSOI is similar to the average distance between ionized impurities image file: d4tc01886a-t1.tif, the number of Coulomb scattering ions surrounding an electron decreases because no ions exist outside the Si device layer plane. The reduced number of neighbor ionized dopants results in reduced Coulomb scattering and leads to enhanced mobility.20 To effectively describe this effect, it is necessary to consider both HSOI and the concentration of ionized scattering centers. Following the procedure described by Kadotani et al.,20 we consider a new variable defined as HSOI/dave. Interestingly, the relative mobility, i.e. the mobility increase with respect to the mobility expected for bulk Si, is found to increase when decreasing HSOI/dave (Fig. 3b). When HSOI/dave ∼ 1, a significant rise of the relative mobility is observed, with values even double those expected for bulk Si. It is worth noting that data follow the same universal curve irrespective of the annealing temperature, indicating that the model provides a consistent picture of the system. Fig. 3b reports the electron mobility values of the SOI films annealed at 1000 °C for 100 s (red open circles). The results obtained from two other sets of samples are also shown in Fig. 3b. These SOI samples were prepared and characterized following the same sample preparation protocol but varying the temperature of the drive-in process of the dopants from 900 (blue open circles) to 1100 °C (green open circles). The duration of each annealing was properly adjusted to inject into the device layer the same ND ∼ 1.0 × 1013 cm−2. As shown in Fig. 3b, our experimental data, that were obtained from different sets of samples annealed in different RTP conditions, are fully compatible with those reported by Kadotani et al. (black closed symbols), particularly for the ultra-thin SOI samples, where the increased electron mobility is observed.20


image file: d4tc01886a-f3.tif
Fig. 3 (a) Schematic representation of ionized impurities and the effect of Coulomb scattering on an electron in a thick or ultra-thin SOI. (b) Relative mobility versus HSOI/dave, for SOI samples upon drive-in at different temperatures. Experimental data obtained in ultra-thin samples upon drive-in of the dopants at 900, 1000, and 1100 °C (open circles) are compared to those reported by Kadotani et al.20 (black closed circles).

Data in Fig. 2a demonstrated a significant ηa reduction for HSOI < 20 nm. In principle, this ηa reduction could be ascribed to incomplete ionization associated to the variation of P concentration in the conductive channel. Fig. 4 shows the average P concentration in the device layer as obtained by ToF-SIMS analysis. P concentration progressively increases as HSOI is reduced. The average electron carrier concentration ne reported in Fig. 4, is derived as the ratio between the Ne value, obtained by Hall measurements, and HSOI. The model of incomplete ionization proposed by Altermatt et al.21 predicts that the fraction of ionized P impurity atoms in bulk Si changes with the concentration of the dopants and exhibits a clear minimum at nD ∼ 2 × 1018 atoms cm−3. However, the model predicts that more than 80% of the dopants is expected to be ionized and, considering the P concentration range of this work, the percentage of ionized impurities is expected to increase as the P concentration increases, i.e., reducing HSOI. The P concentration expected to be ionized at room temperature computed using the parametrization of Abenante22 of the model proposed by Altermatt et al.21 is shown (red solid line) in Fig. 4. The model correctly predicts, within the experimental error, the values recorded for the SOI samples with HSOI ≥ 20 nm but does not account for the strong decrease in ηa observed in ultra-thin SOI. When HSOI < 20 nm, the model developed for bulk Si cannot be directly applied because interface and quantum confinement effects are expected to occur.


image file: d4tc01886a-f4.tif
Fig. 4 The average P concentration obtained by ToF-SIMS analysis (black) and the carrier concentration computed by Hall measurements (blue) versus HSOI upon drive-in at 1000 °C. The measured electron concentration corrected considering the effect of the DIT, the electron concentration predicted by the incomplete ionization model of Altermatt et al.21 and the range predicted by the 1D dielectric mismatch model of Bjork et al.23 corrected considering the reduction of EI, are shown for comparison.

The first interface effect to be considered is the presence of interface state traps between the Si device layer and the surrounding oxides. Fig. 5a shows a scheme elucidating the effect of interface states in n-type-doped SOI. Interface states trap negative charges at the Si/SiO2 interface, resulting in a depletion of mobile charges near the interface.24 As a result, the effective thickness of the conductive channel (Heff) is lower than the physical thickness of the device layer (HSOI). The dimensions of the depleted region can be computed by solving the 1D Poisson equation, assuming both interfaces are identical, considering the full depletion approximation and imposing the charge neutrality condition. Fig. 5b shows Heff plotted against HSOI for various density of interface states (DIT) values. As DIT increases, Heff is reduced considerably. It is worth noting that the graph was computed considering the average dopant concentration in the device layer obtained by ToF-SIMS analysis, which increases as HSOI is reduced. Higher nD leads to smaller depletion regions created by interface states, ultimately resulting in a small difference between HSOI and Heff. Fig. 5b suggests a maximum DIT value of ∼6 × 1012 cm−2 eV−1. Above this threshold value, the model predicts that the 6 nm thick SOI should be fully depleted, resulting in no conduction during the electrical characterization. The correct DIT value could vary when HSOI is reduced. In particular, we expect a higher DIT value as nD is increased due to the enhanced segregation of the P dopants at the interfaces, potentially causing additional traps and doping-induced defects.25 Additionally, according to the literature, DIT values between 1 and 6 × 1012 cm−2 eV−1 are appropriate for non-passivated Si/SiO2 interfaces.26,27


image file: d4tc01886a-f5.tif
Fig. 5 (a) Schematic representation of the effect of trapped charge carries at the interfaces, reducing the effective thickness of the conductive channel (Heff). (b) Heffversus HSOI calculated solving the Poisson equation considering the average P concentration obtained by ToF-SIMS analysis.

Accordingly, to correctly compute the average carrier concentration in the conductive layer, we must divide the Hall dose by Heff. The corrected carrier concentration value in the conductive channel varies in the blue range shown in Fig. 4, which was computed considering the reduction of Heff for DIT values ranging between 1 and 6 × 1012 cm−2 eV−1. Fig. 4 clearly indicates that there is no significant difference when comparing the lower limit to the one without any effect of interface states. However, increasing the DIT to 6 × 1012 cm−2 results in a significant increase in the average carrier concentration in the conductive channel. The reduced dimensions of the conductive channel could contribute to the significant ηa reduction observed in Fig. 2a. However, the correct DIT value may differ from the proposed upper limit, and even after considering the effects of interface states, the experimental data for ultra-thin SOI differ significantly from the values predicted by the model proposed by Altermatt et al.21 (Fig. 4).

The second interface effects to be considered is the dielectric mismatch between Si and SiO2. This effect was highlighted in Si nanowires by Bjork et al.23 Nanostructures usually exhibit sharp dielectric interfaces with their surroundings. These dielectric mismatches are responsible for significant self-energy corrections to the band structure.28 When an electron is injected into a solid, it repels nearby valence electrons.29 In nanostructures this charge accumulates around the dielectric interfaces in the vicinity of the impurity, leading to an extra term in the Coulomb potential.28 The interaction between the electron and these image charges is responsible for the additional self-energy correction to the band structure. The total charge seen far from the impurity as well as the potential is asymptotically unscreened.30 Diarra et al. theoretically studied dopant deionization due to dielectric mismatch between a nanowire and its surroundings. This additional contribution to the tight-band Hamiltonian leads to a strong increase in the ionization energy of the dopants.30 Using their model, an expression for the change in impurity ionization energy (EI) with respect to the ionization energy in bulk Si (EBULKI) is obtained:

 
image file: d4tc01886a-t2.tif(1)
where R is the effective radius of the nanowire, εin and εout are the dielectric constants of Si and surrounding respectively and F is a function of the dielectric ratio of the wire to the surroundings given by Niquet et al.28

The geometry of our system is planar rather than circular. However, following the 1D parametrization given by eqn (1), that we believe is accounting for the maximum contribution of the dielectric mismatch between Si and the surrounding, it is possible to determine the carrier concentration expected in the conductive channel, considering the contribution of dielectric mismatch. Moreover, in our system, nD increases when HSOI is reduced. As a consequence, EBULKI cannot be kept constant, as it decreases when nD is increased.21EBULKI was then evaluated using the incomplete ionization model of Altermatt et al.21 We assume that the two opposite contributions to EI, caused by the increase in nD and the effect of dielectric mismatch can be independently determined and that their overall effect can be accounted by a simple additive procedure. The R parameter in eqn (1) was assumed as Heff/2. The carrier concentration values expected considering the effect of the dielectric mismatch are reported in the bronze range in Fig. 4, for DIT values between 1 and 6 × 1012 cm−2 eV−1. When DIT increases, Heff decreases, resulting in a higher EI correction from eqn (1). Fig. 4 clearly demonstrates that interface effects start to be significant only when HSOI < 20 nm. Above this thickness value, any variation of the DIT produces no significant changes. Interestingly, the model significantly underestimates the fraction of ionized P impurities when HSOI ≥ 20 nm. The same effect was observed in the original paper of Bjork et al., in which less than 30% of dopants were assumed to be ionized at room temperature, even in the case of thick nanowires (R ≥ 30 nm) and high P concentrations (nD = 3 × 1019 cm−3). In this respect, it is worth noting that the 1D dielectric mismatch parametrization is based on the incomplete ionization model of Xiao et al.,31 which assumes a constant EBULKI and strongly overestimates the incomplete ionization in bulk Si when nD > 1018 cm−3.21 Correctly, when HSOI ≥ 20 nm, the experimental results were perfectly described by the incomplete ionization model for bulk Si of Altermatt et al. (red line in Fig. 4) further supporting the idea that above this threshold the system behaves like bulk silicon. Conversely, when HSOI < 20 nm, EI rapidly increases, and the proposed model predicts a significant dopant incomplete ionization. In particular, the model underestimates the ionization observed in the 6 nm thick SOI when DIT = 1 × 1012 cm−2 eV−1 but tremendously overestimates the ionization when DIT = 6 × 1012 cm−2 eV−1, with ne values dropping lower than 3 × 1016 cm−3. In this DIT range, ne values predicted by the 1D dielectric mismatch model overlap the ne values measured and corrected to account for the reduction of Heff. Consequently, the strong incomplete ionization observed in ultra-thin SOI samples could be tentatively described by the increase in EI, assuming a DIT value in the proposed range.

Conclusions

In conclusion, this work investigates the electrical properties of P doped ultra-thin SOI. The annealing treatment efficiently promotes diffusion and incorporation of P atoms in the SOI. If HSOI ≥ 20 nm, the percentage of active dopants at room temperature is perfectly described by the incomplete ionization model of Altermatt et al.21 and the samples behave as bulk-like, perfectly matching the electrical properties of a similarly doped bulk Si substrate.11 If HSOI < 20 nm, a significant increase in carrier mobility is observed, with values even greater than those reported for bulk Si.13 Moreover, a significant dopant incomplete ionization is observed for ultra-thin SOI, with ηa values dropping below 5%. Accordingly, the model developed for bulk Si21 must be corrected considering the contribution of interface effects due to depleted regions resulting from interface traps and dielectric mismatch between Si and SiO2. The proposed model was able to qualitatively describe the experimental data in the case of ultra-thin SOI, but it significantly overestimates the effect of incomplete ionization when HSOI ≥ 20 nm. A 2D parametrization to account for the dielectric mismatch contribution in planar geometry would be necessary to improve the accuracy of the proposed model and properly discriminate between interface effects and quantum confinement corrections that were not considered in this study, even though they are expected to play a role in SOI samples with HSOI < 10 nm, i.e. when HSOI is approximately 2 times the effective Bohr radius of P in Si.32

Author contributions

Andrea Pulici: writing – original draft, investigation, formal analysis, data curation. Stefano Kuschlan: writing – review & editing, investigation. Gabriele Seguini: writing – review & editing, investigation. Marco De Michielis: writing – review & editing, investigation, formal analysis, data curation. Riccardo Chiarcos: writing – review & editing. Michele Laus: writing – review & editing, supervision. Marco Fanciulli: writing – review & editing, supervision. Michele Perego: writing – original draft, supervision, formal analysis, data curation, conceptualization.

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of interest

The authors have no conflicts to disclose.

Acknowledgements

This work was financially supported by the Italian project DONORS (grant number 2022WBPHKF). The project received funding from the Italian programme for Research Projects of outstanding National Interest (PRIN) in the framework of the National Recovery and Resilience Plan (PNRR).

References

  1. H. Mertens, R. Ritzenthaler, A. Hikavyy, M. S. Kim, Z. Tao, K. Wostyn, S. A. Chew, A. De Keersgieter, G. Mannaert, E. Rosseel, T. Schram, K. Devriendt, D. Tsvetanova, H. Dekkers, S. Demuynck, A. Chasin, E. Van Besien, A. Dangol, S. Godny, B. Douhard, N. Bosman, O. Richard, J. Geypen, H. Bender, K. Barla, D. Mocuta, N. Horiguchi and A. V.-Y. Thean, 2016 IEEE Symposium on VLSI Technology, 2016, pp. 1–2.
  2. M. G. Bardon, Y. Sherazi, D. Jang, D. Yakimets, P. Schuddinck, R. Baert, H. Mertens, L. Mattii, B. Parvais, A. Mocuta and D. Verkest, 2018 IEEE Symposium on VLSI Technology, 2018, pp. 143–144.
  3. P. Ye, T. Ernst and M. V. Khare, IEEE Spectr., 2019, 56, 30–35 Search PubMed.
  4. R. Duffy, F. Meaney and E. Galluccio, ECS Meet. Abstr., 2020, MA2020-01, p. 1316.
  5. M. Perego, C. Bonafos and M. Fanciulli, Nanotechnology, 2009, 21, 025602 CrossRef.
  6. E. Arduca and M. Perego, Mater. Sci. Semicond. Process., 2017, 62, 156–170 CrossRef CAS.
  7. Z. Sun, O. Hazut, B.-C. Huang, Y.-P. Chiu, C.-S. Chang, R. Yerushalmi, L. J. Lauhon and D. N. Seidman, Nano Lett., 2016, 16, 4490–4500 CrossRef CAS PubMed.
  8. J. M. Yang, J. Lee, T.-E. Park, D. Seo, J. M. Park, S. Park, J. Na, J. Kwon, H.-J. Lee, J. Ryu and H.-J. Choi, Electron. Mater. Lett., 2019, 15, 208–215 CrossRef CAS.
  9. C. Barri, E. Mafakheri, L. Fagiani, G. Tavani, A. Barzaghi, D. Chrastina, A. Fedorov, J. Frigerio, M. Lodari, F. Scotognella, E. Arduca, M. Abbarchi, M. Perego and M. Bollani, Nanotechnology, 2020, 32, 025303 CrossRef.
  10. M. Perego, G. Seguini, E. Arduca, A. Nomellini, K. Sparnacci, D. Antonioli, V. Gianotti and M. Laus, ACS Nano, 2018, 12, 178–186 CrossRef CAS.
  11. A. Pulici, S. Kuschlan, G. Seguini, F. Taglietti, M. Fanciulli, R. Chiarcos, M. Laus and M. Perego, Mater. Sci. Semicond. Process., 2023, 163, 107548 CrossRef CAS.
  12. M. Perego, F. Caruso, G. Seguini, E. Arduca, R. Mantovan, K. Sparnacci and M. Laus, J. Mater. Chem. C, 2020, 8, 10229–10237 RSC.
  13. S. M. Sze and K. K. Ng, Physics of Semiconductor Devices, John Wiley & Sons, Inc., Hoboken, NJ, USA, 2006 Search PubMed.
  14. V. Gianotti, D. Antonioli, K. Sparnacci, M. Laus, C. Cassino, F. Marsano, G. Seguini and M. Perego, J. Anal. Appl. Pyrolysis, 2017, 128, 238–245 CrossRef CAS.
  15. M. Perego, S. Kuschlan, G. Seguini, R. Chiarcos, V. Gianotti, D. Antonioli, K. Sparnacci and M. Laus, ACS Appl. Polym. Mater., 2021, 3, 6383–6393 CrossRef CAS.
  16. R. Chiarcos, M. Laus and M. Perego, Eur. Polym. J., 2024, 208, 112849 CrossRef CAS.
  17. L. J. van der PAUW, Semiconductor Devices: Pioneering Papers, World Scientific, 1991, pp. 174–182 Search PubMed.
  18. M. Mastromatteo, E. Arduca, E. Napolitani, G. Nicotra, D. De Salvador, L. Bacci, J. Frascaroli, G. Seguini, M. Scuderi, G. Impellizzeri, C. Spinella, M. Perego and A. Carnera, Surf. Interface Anal., 2014, 46, 393–396 CrossRef CAS.
  19. A. Fick, Ann. Phys., 1855, 170, 59–86 CrossRef.
  20. N. Kadotani, T. Takahashi, T. Ohashi, S. Oda and K. Uchida, J. Appl. Phys., 2011, 110, 034502 CrossRef.
  21. P. P. Altermatt, A. Schenk and G. Heiser, J. Appl. Phys., 2006, 100, 113714 CrossRef.
  22. L. Abenante, AIP Adv., 2023, 13, 015109 CrossRef CAS.
  23. M. T. Björk, H. Schmid, J. Knoch, H. Riel and W. Riess, Nat. Nanotechnol., 2009, 4, 103–107 CrossRef.
  24. V. Schmidt, S. Senz and U. Gösele, Appl. Phys. A: Mater. Sci. Process., 2007, 86, 187–191 CrossRef CAS.
  25. J. Snel, Solid-State Electron., 1981, 24, 135–139 CrossRef CAS.
  26. H. Angermann, Th Dittrich and H. Flietner, Appl. Phys. A: Mater. Sci. Process., 1994, 59, 193–197 CrossRef.
  27. W. Lu, C. Leendertz, L. Korte, J. A. Töfflinger and H. Angermann, Energy Procedia, 2014, 55, 805–812 CrossRef CAS.
  28. Y. M. Niquet, A. Lherbier, N. H. Quang, M. V. Fernández-Serra, X. Blase and C. Delerue, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 165319 CrossRef.
  29. J. C. Inkson, Many-Body Theory of Solids, Springer US, Boston, MA, 1984 Search PubMed.
  30. M. Diarra, Y.-M. Niquet, C. Delerue and G. Allan, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 75, 045301 CrossRef.
  31. G. Xiao, J. Lee, J. J. Liou and A. Ortiz-Conde, Microelectron. Reliab., 1999, 39, 1299–1303 CrossRef.
  32. J. W. Orton, The Story of Semiconductors, OUP, Oxford, 2004 Search PubMed.

This journal is © The Royal Society of Chemistry 2024
Click here to see how this site uses Cookies. View our privacy policy here.