Temperature dependent Raman spectroscopy of electrochemically exfoliated few layer black phosphorus nanosheets

Amit S. Pawbakeab, Manisha B. Erandeb, Sandesh R. Jadkarb and Dattatray J. Late*a
aPhysical & Materials Chemistry Division, CSIR-National Chemical Laboratory, Dr. Homi Bhabha Road, Pune 411008, India. E-mail: dj.late@gmail.com; datta099@gmail.com
bDepartment of Physics, Savitribai Phule Pune University, Pune 411007, India

Received 20th June 2016 , Accepted 3rd August 2016

First published on 3rd August 2016


Abstract

The present investigation deals with temperature dependant Raman spectroscopy of electrochemically exfoliated few layer black phosphorus nanosheets. The temperature dependent study illustrates that softening of the A1g, B2g and A2g modes occurs as the temperature increases from 78 K to 573 K. The calculated temperature coefficients for the A1g, B2g and A2g modes were found to be −0.028 cm−1 K−1, −0.028 cm−1 K−1 and −0.018 cm−1 K−1 respectively. The observed phenomenon can be utilized for characterizing other emerging two-dimensional inorganic layered materials with atomic thickness.


Introduction

Graphene has attracted much attention due to its application in nanoelectronic and optoelectronic devices1 due to its huge mobility of ∼2 × 105 cm2 V−1 s−1 which is suitable to be used in fast communication and radio frequency devices.2 Graphene is a promising material due to its extraordinary and unique optical,3 electrical,4 mechanical5 and magnetic properties.6 The low Ion/Ioff ratio in field effect transistor devices only reinforces graphene’s potential use and commercialization in various optoelectronic devices. Graphene-like two dimensional (2D) transition metal dichalcogenide (TMDC) inorganic layered materials have been widely studied recently because they are semiconducting in nature and have a tuneable bandgap, a property which is necessary to switch on and switch off devices. Recently, 2D materials7–18 have been widely used for various applications including optoelectronics,8 gas sensors,9 environmental remediation,10 solid state lubricants,11 solar cells,12 light emitting diodes,13,14 field-effect transistors,15 photo-detectors16,17 and field emitters.18–20

Black phosphorus (P) has several crystal structures such as orthorhombic, rhombohedral, cubic etc.17 Due to its lone pair of electrons it is available in a puckered honeycomb structure instead of atomically thin flat sheets like graphene. The loan pair of electrons is one of the reasons for degradation under ambient conditions.17,21 Black phosphorus is an elemental p-type semiconductor produced by vertical stacking of individual atomic layers; each layer is bonded by three phosphorus atoms (shown in Fig. 1a and b). Overall, black phosphorus is a stable allotrope of phosphorus under ambient conditions. Orthorhombic black phosphorus forms wrinkled layers, which are held together by weak van der Waals forces.21 Similar to other semiconductor materials, the bandgap in bulk black phosphorus as well as in its monolayer form remains direct with a variation from 2.2 eV for monolayer to 0.3 eV for bulk which covers the entire range of the spectrum suitable for optoelectronic device applications.21–23 An exfoliated black phosphorus flake (∼10 nm thickness) based transistor has shown high mobility up to 1000 cm2 V−1 s.24,29 Black phosphorus/h-BN heterostructures have been used in solar cell applications with an efficiency of ∼0.05%,25 highly stable field emitters,26,27 field effect transistors,28 anode materials for lithium ion batteries,30 high performance sensors17,31,32 etc.


image file: c6ra15996f-f1.tif
Fig. 1 Single-layer black phosphorus nanosheet: (a) side view and (b) top view, (c) typical XRD pattern of a bulk black phosphorus crystal, and (d) room temperature Raman spectrum of few layer thick black phosphorus nanosheets.

A systematic study of the vibrational properties of electrochemically exfoliated black phosphorus nanosheets at a wide range of temperatures is still lacking in the literature. A detailed investigation of the vibrational properties of black phosphorus nanosheets is used to understand the electron–phonon interactions which can be a main limiting factor for charge carrier mobility. Thermal conductivity and heat dissipation might be some of the greatest challenges for fabricating existing silicon based electronic and optoelectronic devices. In device application, it is crucial to understand the effect of phonons due to self-heating of the device which can drastically affect the performance of the device. For manipulation of thermal conductivity and heat dissipation, it is important to investigate the electron–phonon interactions and thermodynamics of atomically thin nanosheets of black phosphorus materials synthesized using a chemical method, in order to use these nanosheets for various applications.

In this article, we report the vibrational properties of black phosphorus by means of Raman spectroscopy over a wide temperature range from 78 K to 573 K. The first order temperature coefficients (χ) associated with each Raman mode have been extracted from measured temperature dependence Raman spectroscopy. We observe the linear nature of the temperature dependence of the Raman shifts for the A1g, B2g and A2g modes respectively.

Experimental details

The few layer thick black phosphorus nanosheets have been synthesized using electrochemical exfoliation.26,28

Characterization

For temperature dependant Raman spectroscopy we have used a Renishaw inVia confocal Raman microscope with a backscattering configuration. All the experiments were carried out in ambient argon with 1 mW laser power and 532 nm laser sources. Temperature dependant Raman spectroscopy was carried out over a range of 78 K to 573 K. For the low temperature measurements, we have used a liquid N2 assembly and for high temperature we have used a heater kept below the substrate. Optical images of the electrochemically exfoliated black phosphorus nanosheet samples were captured using a Nikon Eclipse LV 150 NL microscope. For the optical images, the as-synthesized samples were dispersed in ethanol and then dropped on a Si substrate. Auto-exposure times were used during the image recording, and can be changed in the range of 10–500 ms depending upon the intensity. TEM images were captured using an FEI TECNAI G2-20 (TWIN, The Netherlands) instrument operating at 200 KV. X-Ray Diffraction (XRD) patterns for the bulk black phosphorus crystal were recorded using an X-ray diffractometer (Bruker D8 Advance, Germany) using a Cu Kα line (λ = 1.54056 Å).

Results and discussion

Fig. 1 shows a typical schematic presentation of the single-layer black phosphorus nanosheets from a (a) top view and (b) side view. Each ball represents a phosphorus atom in Fig. 1a and b. A typical XRD pattern of the bulk black phosphorus crystal shown in Fig. 1c confirms the orthorhombic crystal structure along the (040) direction. Orthorhombic black phosphorus is produced by repeated vertical stacking of individual atomic layers. The bonding of each atomic layer is oriented in multiple directions. A typical room temperature Raman spectrum of electrochemically exfoliated black phosphorus is as shown in Fig. 1d and shows high intensity A1g, B2g and A2g modes at ∼466 cm−1, ∼438.6 cm−1 and ∼362 cm−1 respectively. A typical optical image of electrochemically exfoliated black phosphorus nanosheets is shown in Fig. 2a. Further, a typical low magnification TEM image is shown in Fig. 2b, which shows a sheet-like morphology. A high resolution TEM image was also recorded which shows a d spacing of ∼0.25 nm as depicted in Fig. 2c which shows that its orientation is along the (040) plane. Fig. 2d shows a typical indexed selected area electron diffraction (SAED) pattern of a few layer black phosphorus nanosheet sample. The diffraction pattern confirms the highly crystalline nature of the black phosphorus nanosheets. Fig. 3 shows the Raman spectrum for a few layer black phosphorus nanosheet sample as a function of temperature. The Raman spectrum shows consecutive softening of all the Raman modes with increasing temperature. The temperature dependence of the Raman modes’ position is related to the electron–phonon, anharmonic phonon–phonon interactions, or pure thermal expansion. Further, an increase in the full width at half maximum (FWHM) for all the Raman modes was also observed with increasing temperature. Fig. 4a–c shows a typical plot for the Raman spectra’s peak positions as a function of temperature for the A1g, B2g and A2g modes respectively. The Raman modes of the electrochemically exfoliated black phosphorus nanosheets behave linearly over the temperature range from 78 K to 573 K. The changes in the Raman frequency (Δω) for the A2g, B2g and A1g Raman modes of the electrochemically exfoliated black phosphorous were found to be 13.44 cm−1, 13.66 cm−1 and 8.51 cm−1 respectively over the temperature range of 78 K to 573 K. The peak positions versus temperature were fitted using the following eqn (1),33
 
ω(T) = ω0 + χT (1)
where ω0 is the peak position of the A1g, B2g and A2g vibration modes at zero Kelvin, χ is the first order temperature coefficient of the A1g, B2g and A2g modes. The Raman modes A1g, B2g and A2g behave linearly with temperature and the slope of the fitted straight line gives the temperature coefficient (χ). The calculated values of the temperature coefficients for A1g, B2g and A2g are found to be −0.028 cm−1 K−1, −0.028 cm−1 K−1 and −0.018 cm−1 K−1 respectively. The observed values of the temperature coefficients are in good agreement with those reported for the TMDC;34–39 trichalcogenide40 and black phosphorus nanosheet44–47 materials which are listed in Table 1. The change in the Raman frequency is given by eqn (2),41,42
 
image file: c6ra15996f-t1.tif(2)
where χ = χT + χV in which χT is the self-energy shift due to the coupling of the phonon modes and χV is the shift due to the thermal expansion induced volume change. The reason these two terms arise is that all the measurements have been carried out at constant pressure rather than constant volume.40,41

image file: c6ra15996f-f2.tif
Fig. 2 Electrochemically exfoliated black phosphorus nanosheets: (a) optical image of the nanosheet sample deposited on a Si substrate, (b) typical low magnification TEM image, (c) high resolution TEM image, and (d) SAED pattern.

image file: c6ra15996f-f3.tif
Fig. 3 Raman shift as a function of temperature for an electrochemically exfoliated few layer black phosphorus nanosheet sample.

image file: c6ra15996f-f4.tif
Fig. 4 Raman spectra peak position as a function of temperature for the (a) A2g, (b) B2g and (c) A1g modes respectively. The FWHM as a function of temperature for the (d) A2g, (e) B2g and (f) A1g modes showing increasing FWHM with temperature.
Table 1 Comparison of the temperature coefficients reported in the literature for various 2D nanosheets
2D materials χ (cm−1 K−1) References
A1g B2g A1g
Black phosphorus −0.0164 −0.0271 −0.0283 44
−0.02175 −0.02877 −0.027 45
−0.008 −0.013 −0.014 46
−0.01895 −0.02434 −0.02316 45
−0.023 −0.018 −0.023 47
−0.028 −0.028 −0.018 Present

2D materials χ (cm−1 K−1) References
E12g A1g
MoS2 −0.016 −0.011 34
−0.0136 −0.0113 39
WS2 −0.008 −0.004 34
−0.0098 −0.014 39

2D materials χ (cm−1 K−1) References
A1g
MoSe2 −0.0054 36
−0.0096 39
WSe2 −0.0071 39
−0.0032 36


The changes in FWHM for the A1g, B2g and A2g modes are depicted in Fig. 4d–f respectively. The broadening in the Raman modes with temperature is based on the phonon dispersion and many body theoretical calculations. The change in the FWHM is mainly due to the contribution from the decay of the zone-centre optical phonon into one acoustic and one optical phonon and these two phonons were selected from the phonon density states. The change in the line width as a function of temperature is given by the equation,41–43

 
Γ(T) = Γ0 + A[1 + n(ω1, T) + n(ω2, T)] (3)
where Γ0 represents the background contribution, A is the anharmonic coefficient and n(ω, T) is the Bose–Einstein contribution function. The temperature dependent FWHM can be determined from the parameters such as Γ0, A, ω1 and ω2. The change in the FWHM, intensity and shift in the peak position as a function of temperature can be explained by using a double resonance phenomenon which is very active in atomically thin nanosheets.

Conclusion

In summary, we have systematically performed temperature dependent Raman spectroscopy of electrochemically exfoliated black phosphorus nanosheet samples. The TEM images, and XRD and room temperature Raman spectra show the highly crystalline nature of the sample. The temperature dependent Raman spectra of few layer black phosphorous nanosheets show softening of the A2g, B2g and A1g modes. The calculated temperature coefficients for the A1g, B2g and A2g modes of the few layer black phosphorus nanosheet sample were found to be −0.028 cm−1 K−1, −0.028 cm−1 K−1 and −0.018 cm−1 K−1 respectively. These results are explained on the basis of a double resonance phenomenon which is active in atomically thin nanosheets. Thus we realize that the A1g mode is more active via an electron–phonon interaction. These results will be helpful in understanding the vibrational properties of atomically thin black phosphorus nanosheet based integrated circuits and useful for their thermal manipulation in the near future.

Conflict of interest

The author declares no competing financial interest.

Acknowledgements

This research work was supported by the Department of Science and Technology (Government of India) under the Ramanujan Fellowship to Dr D. J. Late (Grant No. SR/S2/RJN-130/2012), NCL-MLP project grant 028626, DST-SERB Fast-track Young scientist project Grant No. SB/FT/CS-116/2013, Board of Research in Nuclear Sciences (BRNS Grant No. 34/14/20/2015), Government of India, and partially supported by the INUP IITB project sponsored by the DeitY, MCIT (Grant No. MT 114), Government of India.

Notes and references

  1. K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V. Khotkevich, S. V. Morozov and A. K. Geim, Proc. Natl. Acad. Sci. U. S. A., 2005, 102, 10451–10453 CrossRef CAS PubMed.
  2. Y.-M. Lin, C. Dimitrakopoulos, K. A. Jenkins, D. B. Farmer, H.-Y. Chiu, A. Grill and P. Avouris, Science, 2010, 327, 662 CrossRef CAS PubMed.
  3. A. N. Grigorenko, M. Polini and K. S. Novoselov, Nat. Photonics, 2012, 6, 749–758 CrossRef CAS.
  4. S. Stankovich, D. A. Dikin, G. H. B. Dommett, K. M. Kohlhaas, E. J. Zimney, E. A. Stach, R. D. Piner, S. T. Nguyen and R. S. Ruoff, Nature, 2006, 442, 282–286 CrossRef CAS PubMed.
  5. X. Zhao, Q. Zhang, D. Chen and P. Lu, Macromolecules, 2010, 43, 2357–2363 CrossRef CAS.
  6. H. S. S. R. Matte, K. S. Subrahmanyam and C. N. R. Rao, J. Phys. Chem. C, 2009, 113, 9982–9985 CAS.
  7. H. S. S. R. Matte, A. K. Manna, D. J. Late, R. Datta, S. K. Pati and C. N. R. Rao, Angew. Chem., Int. Ed., 2010, 122, 4153–4156 CrossRef.
  8. Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman and M. S. Strano, Nat. Nanotechnol., 2012, 7, 699–712 CrossRef CAS PubMed.
  9. D. J. Late, Y.-K. Huang, B. Liu, J. Acharya, S. N. Shirodkar, J. Luo, A. Yan, D. Charles, U. V. Waghmare, V. P. Dravid and C. N. R. Rao, ACS Nano, 2013, 7, 4879–4891 CrossRef CAS PubMed.
  10. P. F. Weck, E. Kim and K. R. Czerwinski, Dalton Trans., 2013, 42, 15288–15295 RSC.
  11. A. N. Enyashin, L. Yadgarov, L. Houben, I. Popov, M. Weidenbach, R. Tenne, M. Bar-Sadan and G. Seifert, J. Phys. Chem. C, 2011, 115, 24586–24591 CAS.
  12. J. Dai and X. C. Zeng, J. Phys. Chem. Lett., 2014, 5, 1289–1293 CrossRef CAS PubMed.
  13. G. L. Frey, K. J. Reynolds, R. H. Friend, H. Cohen and Y. Feldman, J. Am. Chem. Soc., 2003, 125, 5998–6007 CrossRef CAS PubMed.
  14. F. Xia, H. Wang, D. Xiao and M. Dubey, Nat. Photonics, 2014, 8, 899–907 CrossRef CAS.
  15. D. J. Late, B. Liu, H. S. S. R. Matte, V. P. Dravid and C. N. R. Rao, ACS Nano, 2012, 6, 5635–5641 CrossRef CAS PubMed.
  16. A. S. Pawbake, R. G. Waykar, D. J. Late and S. R. Jadkar, ACS Appl. Mater. Interfaces, 2016, 8, 3359–3365 CAS.
  17. M. Buscema, D. J. Groenendijk, S. I. Blanter, G. A. Steele, H. S. J. van der Zant and A. Castellanos-Gomez, Nano Lett., 2014, 14, 3347–3352 CrossRef CAS PubMed.
  18. C. S. Rout, P. D. Joshi, R. V. Kashid, D. S. Joag, M. A. More, A. J. Simbeck, M. Washington, S. K. Nayak and D. J. Late, Sci. Rep., 2013, 3, 3282 Search PubMed.
  19. S. R. Suryawanshi, A. S. Pawbake, M. S. Pawar, S. R. Jadkar, M. A. More and D. J. Late, Mater. Res. Express, 2016, 3, 035003 CrossRef.
  20. D. J. Late, P. A. Shaikh, R. Khare, R. V. Kashid, M. Chaudhary, M. A. More and S. B. Ogale, ACS Appl. Mater. Interfaces, 2014, 6, 15881–15888 CAS.
  21. A. Castellanos-Gomez, J. Phys. Chem. Lett., 2015, 6, 4280–4291 CrossRef CAS PubMed.
  22. A. Yuichi, S. Endo and S. Narita, J. Phys. Soc. Jpn., 1983, 52, 2148–2155 CrossRef.
  23. R. W. Keyes, Phys. Rev., 1953, 92, 580 CrossRef CAS.
  24. F. Xia, H. Wang and Y. Jia, Nat. Commun., 2014, 5, 4458 CAS.
  25. M. Buscema, D. J. Groenendijk, G. A. Steele, H. S. J. van der Zant and A. Castellanos-Gomez, Nat. Commun., 2014, 5, 4651 CrossRef CAS PubMed.
  26. M. B. Erande, S. R. Suryawanshi, M. A. More and D. J. Late, Eur. J. Inorg. Chem., 2015, 3102–3107 CrossRef CAS.
  27. S. R. Suryawanshi, M. A. More and D. J. Late, J. Vac. Sci. Technol., B, 2016, 34, 041803 Search PubMed.
  28. M. B. Erande, M. S. Pawar and D. J. Late, ACS Appl. Mater. Interfaces, 2016, 8, 11548–11556 CAS.
  29. L. Li, Y. Yu, G. J. Ye, Q. Ge, X. Ou, H. Wu, D. Feng, X. H. Chen and Y. Zhang, Nat. Nanotechnol., 2014, 9, 372–377 CrossRef CAS PubMed.
  30. C. M. Park and H. J. Sohn, Adv. Mater., 2007, 19, 2465–2468 CrossRef CAS.
  31. A. N. Abbas, B. Liu, L. Chen, Y. Ma, S. Cong, N. Aroonyadet, M. Kopf, T. Nilges and C. Zhou, ACS Nano, 2015, 9, 5618–5624 CrossRef CAS PubMed.
  32. D. J. Late, Microporous Mesoporous Mater., 2016, 225, 494–503 CrossRef CAS.
  33. E. S. Zouboulis and M. Grimsditch, Phys. Rev. B: Condens. Matter Mater. Phys., 1991, 43, 12490 CrossRef CAS.
  34. M. Thripuranthaka, R. V. Kashid, C. S. Rout and D. J. Late, Appl. Phys. Lett., 2014, 104, 081911 CrossRef.
  35. D. J. Late, U. Maitra, L. S. Panchakarla, U. V. Waghmare and C. N. R. Rao, J. Phys.: Condens. Matter, 2011, 23, 055303 CrossRef PubMed.
  36. D. J. Late, S. N. Shirodkar, U. V. Waghmare, V. P. Dravid and C. N. R. Rao, ChemPhysChem, 2014, 15, 1592–1598 CrossRef CAS PubMed.
  37. M. Thripuranthaka and D. J. Late, ACS Appl. Mater. Interfaces, 2014, 6, 1158–1163 Search PubMed.
  38. D. J. Late, ACS Appl. Mater. Interfaces, 2015, 7, 5857–5862 CAS.
  39. A. Pawbake, M. Pawar, S. R. Jadkar and D. J. Late, Nanoscale, 2016, 8, 3008–3018 RSC.
  40. A. S. Pawbake, J. O. Island, E. Flores, J. R. Ares, C. Sanchez, I. J. Ferrer, S. R. Jadkar, H. S. J. van der Zant, A. Castellanos-Gomez and D. J. Late, ACS Appl. Mater. Interfaces, 2015, 7, 24185–24190 CAS.
  41. P. H. Tan, Y. M. Deng and Q. Zhao, Phys. Rev. B: Condens. Matter Mater. Phys., 1998, 58, 5435 CrossRef CAS.
  42. P. H. Tan, Y. M. Deng, Q. Zhao and W. C. Cheng, Appl. Phys. Lett., 1999, 74, 1818 CrossRef CAS.
  43. J. Menendez and M. Cardona, Phys. Rev. B: Condens. Matter Mater. Phys., 1984, 29, 2051 CrossRef CAS.
  44. A. Lapinska, A. Taube, J. Judek and M. Zdrojek, J. Phys. Chem. C, 2016, 120, 5265–5270 CAS.
  45. Z. Luo, J. Maassen, Y. Deng, Y. Du, R. P. Garrelts, M. S. Lundstrom, P. D. Ye and X. Xu, Nat. Commun., 2015, 6, 8572 CrossRef CAS PubMed.
  46. D. J. Late, ACS Appl. Mater. Interfaces, 2015, 7, 5857–5862 CAS.
  47. S. Zhang, J. Yang, R. Xu, F. Wang, W. Li, M. Ghufran, Y. W. Zhang, Z. Yu, G. Zhang, Q. Qin and Y. Lu, ACS Nano, 2014, 8, 9590–9596 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.