Controllable Laser Reduction of Graphene Oxide Films for

Oct 5, 2016 - This article presents a new simple method of creating light-absorbing carbon material for optical devices such as bolometers. A simple m...
1 downloads 11 Views 6MB Size
Research Article www.acsami.org

Controllable Laser Reduction of Graphene Oxide Films for Photoelectronic Applications Stanislav Evlashin,*,†,∥ Pavel Dyakonov,† Roman Khmelnitsky,∥ Sarkis Dagesyan,‡ Andrey Klokov,∥ Andrey Sharkov,∥ Peter Timashev,⊥ Svetlana Minaeva,⊥ Konstantin Maslakov,§ Sergey Svyakhovskiy,‡ and Nikolay Suetin†,# †

D. V. Skobeltsyn Institute of Nuclear Physics, ‡Department of Physics, §Department of Chemistry, M. V. Lomonosov Moscow State University, 1(2) Leninskie Gory, 119991 Moscow, Russia ∥ P. N. Lebedev Physical Institute, Russian Academy of Sciences, 53 Leninsky Prospect, 119991 Moscow, Russia ⊥ Institute on Laser and Information Technologies RAS, 2 Pionerskaya,142092 Troitsk, Russia # Skolkovo Institute of Science and Technology, 3 Nobel, 143025 Skolkovo, Russia S Supporting Information *

ABSTRACT: This article presents a new simple method of creating lightabsorbing carbon material for optical devices such as bolometers. A simple method of laser microstructuring of graphene oxide is used in order to create such material. The absorption values of more than 98% in the visible and more than 90% in the infrared range are achieved. Moreover thermal properties of the films, such as temperature dependence and the thermal response of the samples, are studied. The change in resistance with temperature is 13 Ohm K−1, temperature coefficient of resistance (TCR) is 0.3% K−1, and the sensitivity is 0.17 V W−1 at 300 K. Thermal conductivity is rather high at ∼104 W m−1 K−1 at 300 K. The designed bolometer operates at room temperature using incandescent lamp as a light source. This technique suggests a new inexpensive way to create a selective absorption coating and/or active layer for optical devices. Developed GO and rGO films have a large surface area and high conductivity. These properties make carbon coatings a perfect candidate for creating a new type of optoelectronic devices (gas sensors, detectors of biological objects, etc.). KEYWORDS: graphene oxide, optical properties, thermal properties, bolometers, laser microstructuring

1. INTRODUCTION Optoelectronic devices are used in a wide range of applications: optical communications, night vision devices, gas sensors, motion detectors, and many others. All these applications require development of materials that on one hand have a minimum light reflectance and on the other hand can efficiently convert optical (or infrared) light into electrical signals. In order to solve the first problem, nonreflective coatings are used to transmit light in a wide range of wavelengths and at any angles of radiation incidence to the surface of the device. Conventional optical coatings can be divided into several types according to their morphology, including multilayer coatings, structured substrate, periodic grating, and chaotic structures.1,2 As it was shown in ref 1, in order to achieve high absorption of materials we need to control two main parameters of the films: reflection and absorption. Materials with low reflectance usually have a graded refractive index.3 Optical properties of different materials and various techniques of their preparation are discussed in numerous recent reviews.1,4 In order to solve the second problem, it is crucial to develop materials that can easily convert optical signals into electrical © 2016 American Chemical Society

signals. In traditional nonselective radiation detectors, such as bolometers, radiation leads to heating of the absorbing layer. As a result of the heating process, the resistance of the active layer changes what leads to the shift of the electric current in the circuit. In terms of increasing the sensitivity of a bolometer it is essential to maximize the absorption of radiation and minimize heat-sink from the active area. Also, in order to achieve fast response of the device it is significant to decrease heat capacity of absorber and active layers. Furthermore, the active layer should have high coefficient of conversion heat into electric current, i.e., maximum slope of of R(T) line, where R(T) is the dependence of active layer resistance on the temperature. Thus, bolometer sensitivity ≈ [dR/dT] × A/k, where [dR/dT] is the slope of R(T), A is the coefficient of radiation absorption, and k is the thermal conductance (value that is reciprocal of thermal resistance). Received: August 13, 2016 Accepted: October 5, 2016 Published: October 5, 2016 28880

DOI: 10.1021/acsami.6b10145 ACS Appl. Mater. Interfaces 2016, 8, 28880−28887

Research Article

ACS Applied Materials & Interfaces

Figure 1. (a) Process of sample preparation. We dropped graphene oxide onto a silicon (glass) substrate. When it got dry we processed it with laser. (b) Measurement scheme of complete and spectral reflection of the samples. (c) Process of installation and the contact diagram when measuring thermal properties.

thermal (bolometric) properties in a wide range of wavelengths and temperatures was carried out. We realized a bolometer prototype, which operated with conventional incandescent lamp. For the materials under study, we observed the maximum TCR value to be 0.3% K−1 at 300 K. Our structures display TCR values lower than some widely used bolometric materials, while absorption properties of our structures outperform all known analogues. The fact that rGO and GO have different optical, thermal, and electrical properties allows us to create a selective absorbing coating for optoelectronic devices.

A number of one- and two-dimensional materials, including carbon nanotubes and graphene, are suitable for this purpose.5,6 Due to the fact that graphene has zero band gap, the impact of any irradiation can generate charge carriers in a wide wavelength range from ultraviolet to terahertz. Bolometers,7−10 photodetectors,11 optical devices with nonlinear properties,12,13 and optical modulators14 have been implemented on the basis of carbon materials such as graphene, graphene oxide (GO), reduced graphene oxide (rGO), and carbon nanotubes. On the basis of carbon nanostructures, devices that have effective antireflective, absorbing, and converting properties can be created. Given the fact that carbon materials have high values of thermal conductivity along the layers, it can be concluded that we can obtain an efficient photoconverter for the infrared (IR) range by using carbon materials.15,16 Detailed properties of different bolometric materials are shown in Table S1. Values of metal temperature coefficient of resistance (TCR) were less than 1% K−1. The highest TCR value of 5% K−1 was reached for VOx and Si at room temperature. A lot of studies have already demonstrated that bolometers can be designed on basis of rGO and achieve a TCR value of 0.03% K−1, and the values of voltage responsivity vary from 4.3 to 1200 V W−1.8,17,18 Both TCR and absorption play crucial part in bolometer performance. The use of doubleprobe methods in some articles do not allow estimation of the contact phenomena and their contribution to TCR and volt− watt sensitivity.17,18 Moreover, there are no experimental studies on rGO bolometers response to the nanosecond laser pulses. Here we investigate the response of rGO bolometers to nanosecond laser pulses. Our previous papers have revealed that carbon nanowalls (CNW) also possess unique optical properties in visible and IR range.19,20 However, unlike nanotubes, CNW do not affect the polarization of electromagnetic radiation; they are resistant to moisture and have a remarkably small thickness (∼1−3 μm). At the same time, the creation of the materials and devices listed above requires expensive equipment and complex, multistage processes. That results in a substantial increase of the cost of the final products. Here we demonstrate a simple way to create an absorptive coating using laser microstructuring of GO. Almost any surface can be coated with GO by simply pouring solution on the surface for further drying, while laser microstructuring allows controllable change of material properties in certain spots. The precision of the structure creation is determined by resolution of the laser system and beam energy. In this work we study the effect of graphene oxide structuring modes on the obtained optical and structural properties of the samples. The absorption of more than 98 and 90% in the visible and IR wavelength ranges, correspondingly, was obtained. Moreover, a study of

2. EXPERIMENTAL SECTION 2.1. Synthesis of Graphene Oxide. GO was prepared from commercially available graphite powder Timcal (Timrex KS 15) by Hummers method.21 After centrifugation 18 mg mL−1 aqueous GO solution was obtained. Various surfaces such as glass and silicon were drop-cast with different GO solutions. The volume of suspension poured on the substrate surface remained constant and was 0.16 mL cm−2. The thickness of the films was estimated by a scanning electronic microscope (SEM). 2.2. Laser Microstructuring of the Samples. Laser microstructuring of the samples was carried out by the fiber laser by IREPolus company with wavelength of 1.064 μm. GO coatings were microstructured with laser beam and thus treated areas were reduced to rGO. Treated areas represented a grid of rGO. Laser fluence values were in range from 4 to 28 J cm−2. Most of the results were achieved at fluence of 4 J cm−2. Gaussian mode profile was focused on the sample surface in a 50 μm diameter spot, pulse duration time of 20 ns and frequency of 10 000 Hz was used. Therefore, total overlapping factor was about 3. 2.3. Analysis of Samples. Analysis of samples was carried out using scanning electron microscopy by Carl Zeiss (Supra 40), FTIR spectrometer Bruker (Vertex 70 V), and Raman spectrometer HORIBA Jobin Yvon (LabRAM HR800 UV−visible−NIR). X-ray photoelectron spectra (XPS) were recorded by a spectrometer Axis Ultra DLD (Kratos) in monochromatic Al Kα radiation with the analyzer energy transmission of 20 eV. 2.4. Optical Measurements. Total optical properties of the samples were studied by LOMO (SF 56) and Avesta 100 spectrophotometers. 2.5. Thermal Properties. To study thermal properties of obtained structures 200 nm thick Au contacts were deposited on microstructured rGO films. Magnetron sputtering was used to deposit Au contacts. Sputtering was carried out at constant current in argon atmosphere. Dependence of thermal response on laser impulse was studied in the same cryostat, where resistance−temperature characteristics were measured. Characteristics were measured using the doubleprobe method. To display output data we used LeCroy WaveRunner 62Xi-A oscillograph. Intercontact space was irradiated with spread-out laser beam (Tech 1053 laser, wavelength, 1.053 μm; pulse duration time, ∼4 ns). Contact region was covered with masking layer to avoid penetration of laser irradiation. Resistance of rGO film was measured using the 4-probe method at the temperature range from 80K to 370 28881

DOI: 10.1021/acsami.6b10145 ACS Appl. Mater. Interfaces 2016, 8, 28880−28887

Research Article

ACS Applied Materials & Interfaces K. For measurements at the temperature range from 100 to 250 K we used optical cryostat OptCry197. In order to determine the thermal parameters of rGO film, we compared an experimental response to laser pulse heating with the simulated one. It was obtained by solving the system of two heattransfer equations. It simulates the heat transmission in axially symmetric geometry in the two-layered structures. We took into account the thermal resistance of the interface between rGO film and substrate. The equations were solved using the method of separation of variables. The correct estimation of unknown thermal parameters requires knowledge of laser impulse shape and amplitude−frequency locus of the registration system. Using germanium photodiode FP-70 with pulse leading-edge time of 70 ps, we measured the shape of the laser impulse. The method of determining the thermal properties in details is similar to the one that was described in ref 22. Absorption of the film obeys Bouguer−Lambert law. Unknown thermal characteristics of film (such as volumetric heat capacity, thermal conductivity, and border heat resistance) came up as a result of comparison of experimental data and results of computer simulation.

3. RESULTS AND DISCUSSION 3.1. Structural Properties. Figure 1 represents schematics of creating GO film and its following reduction, also it shows the schematics of optical and thermal measurements which were carried out on obtained sample. Substrates were drop-cast with graphene oxide suspension of different concentrations. The increasing concentration from 3 to 18 mg mL−1 with a constant volume of 0.16 mL cm−2 resulted in a linear increase in film thickness from 5 to 30 μm. These measurements were carried out using SEM. Then, the samples were laser irradiated. Laser fluence was changed from 4 to 28 J cm−2; diameter of laser beam was 50 μm. Figures 2 and S1 show structures obtained by laser treatment. As can be seen, the structured surface represent a grid with lines of rGO. In order to study the structural properties of samples, we used XPS, Raman spectroscopy, and FTIR (Figure 3 and Figure S2). Figure 3a reveals the presence of oxygen-containing functionalities in the form of epoxy, hydroxyl, and carbonyl groups. Laser fluence 4 J cm−2 leads to significant changes. The number of C−O bonds drops by a factor of 2 while the range of

Figure 3. XPS C 1s peak spectra of graphite oxide before (a) and after (b) laser irradiation at laser fluence 4 J cm−2 and line densities 20 lines per mm. (c) Raman spectra of rGO at laser fluence varying from 4 to 28 J cm−2, line densities 20 line per mm.

C−C bonds simultaneously increases. These changes are shown in Figure 3a,b. Further increase of laser fluence up to 28 J cm−2 leads to a slight changes. More detailed data is presented in Table S2. Similar regularities in the processing of graphene with laser beam can be observed in.23,24 FTIR spectra well correlates with XPS data. FTIR spectra in Figure S2 shows common GO peaks: 3395 cm−1 (O−H stretching vibrations), 1730 cm−1 (CO stretching vibrations from carbonyl and carboxylic groups), 1595 cm−1 (skeletal vibrations of unoxidized graphitic domains), 1413 cm−1 (C− O), 1226 cm−1 (C−O (epoxy) stretching vibration), and 1050 cm−1 (and the C−O (alkoxy) stretching).25 Increase of laser beam energy leads to slight increase of C−C, C−O bands intensities, and decrease of O−H band intensity. Figure 3c shows a typical Raman spectra of GO and rGO. Raman spectrum has some characteristic D- and G-bands, located at 1351 and 1581 cm−1; the ratio of D-band intensity to intensity of G-band is 0.97. Because of the laser treatment Dband shifts significantly and now can be found at 1342 cm−1. Gband position changes slightly; its new location is 1586 cm−1. These changes can be described by structural modification of the films.26,27 The I(D)/I(G) ratio moves toward a value of about ∼0.82,28 which tells us sufficient decrease in number of domains upon reduction of the exfoliated GO and can be explained by creation of greater number of new graphite domains of smaller size compared to the size of ones that were present in GO before reduction.23,24,29 D-band dislocation can be explained by local reduction of GO after laser treatment of samples surface; the laser pulse overheats GO and thus leads to its partial reduction. Local destruction/reduction of GO also influences the morphology of sample.30 The 2D peak is not so well pronounced as it is shown in ref 24. It means that degree of graphitization is lower than ones shown in cited papers. 3.2. Optical Properties. Laser treatment of samples changes structure and color of the film. Sample surface that once was dark brown (GO) becomes black (rGO) after laser treatment. Optical properties of obtained samples were

Figure 2. SEM images of rGO structures produced from GO solution (concentration 4 mg mL−1) at fluence of 4 J cm−2 and line density of 20 lines per mm. (a) General view of structured surface. (b) Top view of structures. (c and d) Side views of the structures at different magnification. 28882

DOI: 10.1021/acsami.6b10145 ACS Appl. Mater. Interfaces 2016, 8, 28880−28887

Research Article

ACS Applied Materials & Interfaces

Figure 4. Total reflection of the sample. (a) Sample reflection for different concentrations of 3, 6, 12, and 18 mg mL−1 at the constant line density (20 lines per mm, laser fluence 4 J cm−2) (b) Sample reflection for different line densities of 0, 10, 15, 20, and 25 lines per mm at the constant concentration (12 mg mL−1, laser fluence 4 J cm−2). (c) Sample reflection for different laser fluence 4−10 J cm−2 at constant concentration and line density (20 lines per mm). (d) Dependences of films absorption for different line densities of 0, 10, 15, 20, and 25 lines per mm (12 mg mL−1, laser fluence 4 J cm−2).

be defined by reflection of the sample. Figure 4d shows absorption spectra of samples with different areas of structured surface (black line stands for a sample that remained untreated). The highest absorption coefficient for visible wavelength range reached a value of 98%. Figure S3 displays the reflection coefficient spectra for normal incident light and mirror-reflected light. The increase in angles of incidence and reflection leads to an increase of the reflection coefficient. The increase of wavelength also causes an increase in reflection coefficient. Observations did not show any correspondence between reflection coefficient and polarization direction Figure S4. Figure 5 shows the FTIR spectra of obtained samples; for comparison, each plot has a curve for GO samples which remained untreated. Figure 5a shows reflection spectra for samples of different film thickness. The lowest reflection was observed for 12 mg mL−1 GO solution. Figure 5b shows reflection spectra for samples of different line density. In comparison to the visible light range, the infrared range has the lowest reflection coefficient observed for line density of 20 lines mm−1. The lowest transmittance was observed for samples with line densities of 15 and 20 lines mm−1. The absorption of films was calculated in the same fashion as in ref 20. An absorption ratio of 90% was obtained for sample with maximum film thickness. Infrared absorption essentially is determined by film thickness, while surface morphology does not show any sufficient effects on it. These effects were well studied in our previous articles and also shown in Figure S5.20 Figure S5 represents the absorption of the samples with different numbers of lines per unit areas. It is shown that structures with different numbers of lines show no significant effect on IR absorption (in IR measurements, the beam size is larger than the distance

measured. Figure 4 displays a total reflection of samples for different conditions. Measurements scheme is shown on Figure 1b. Total reflection spectra of the samples with different film thickness are shown in Figure 4a. In order to obtain samples of different film thickness various concentrations aqueous GO solutions have been used, the volume of the solution poured on substrate kept constant. The diameter of laser beam was 50 μm. The density of structural lines was kept at a value of 20 lines mm−1. The lowest reflection value at optical wavelength range was obtained for samples with a GO concentration of 12 mg mL−1. The dependence between film thickness and reflection coefficient was observed. At first, the increase of film thickness leads to a slow decrease of reflection coefficient until it reaches its minimum value at 20 lines mm−1; further increase in film thickness results in slow increase of reflection coefficient. Reflection coefficients for both GO and rGO are shown on Figure 4b. The GO reflection coefficient appears to be higher than the one for rGO. The ratio of surface area covered with GO to surface area covered with rGO will defines the total reflection coefficient. Some samples were structured more then once; such a treatment causes destruction of GO suspended on a substrate, hence changing the reflection spectrum of the sample. Figure 4c shows correlation between reflection coefficient and laser beam energy. For 12 mg mL−1 GO solution, the lowest reflection coefficient was obtained for laser fluence of 4 J cm−2. Further increase of laser energy as well as induction of a number of patterning lines more than 20 leads to the destruction of material and changes the morphology of the structure. It has significant impact on reflection coefficient in visible wavelength range.19 Obtained samples are nontransparent for visible light range; therefore, absorption coefficient will 28883

DOI: 10.1021/acsami.6b10145 ACS Appl. Mater. Interfaces 2016, 8, 28880−28887

Research Article

ACS Applied Materials & Interfaces

Figure 5. Characteristics of the sample in IR wavelength range. (a) Sample reflection for different concentrations of 3, 6, 12, and 18 mg mL−1 at the constant line density and fluence of 20 lines per mm and 4 J cm−2, correspondingly. (b) Reflection of the sample structured at different line densities of 0, 15, 20, and 25 lines per mm (concentration 12 mg mL−1, laser fluence 4 J cm−2). (c) Transmission of the sample structured at different line densities of 0, 15, 20, and 25 lines per mm (concentration 12 mg mL−1, laser fluence 4 J cm−2). (d) Absorption of the sample at different concentrations of 3, 6, 12, and 18 mg mL−1 (20 lines per mm, fluence 4 J cm−2).

between the structured lines). The main difference can be observed in the presence of the reduced graphene oxide as there is a significant change in the energy band gap. That results in the increase of absorption.31 3.3. Thermal Properties. In order to measure thermal characteristics, several samples with low reflection coefficient and high absorption were selected. The concentration of original GO solution was 12 mg mL−1, and line density of 20 lines per mm. In order to measure thermal properties, Au contacts were deposited on structured film as shown on Figure 1c. Resistance versus temperature response characteristics were measured by four-probe method in a temperature ranges from 100 to 250 K and from 300 to 350 K. Measurements were carried out at bias current ranging from −150 to 150 μA with a step of 50 μA. Dependence of resistance on current was not observed. Figure 6a illustrates measurements for bias current of 100 μA and linear approximations of those measurements. Increase of temperature leads to a decrease of resistance. The value of dR/ dT ranges from −50 Ohm K−1 for 100 K to −13 Ohm K−1 for 350 K. Dependence was approximately linear within the temperature range from 120 to 220 K; slope was −27 Ohm K−1. The temperature coefficient of resistance nonlinearly depends on temperature and ranges from −6 × 10−3 K−1 at 100 K to −3 × 10−3 K−1 at 300 K. Such differences in slope can be explained by varying number of OH groups for rGO samples in vacuum and air. It has been observed and resulted in changes of ionic conductivity.32

Bolometric response of structure to laser pulse input was measured for two opposite bias current directions. The value measured for positive direction of bias current was subtracted from the value measured for negative direction with provision for the fact that nonbolometric current direction-independent component (for example, photoemf or thermal-emf) is not taken into consideration, and bolometric component comes into account twice. This is the reason why responses have positive polarity (negative value of dR/dT stands for negative response polarity). Figure 6b displays results for studied bolometric responses of rGO film to laser pulse illumination for two values of temperature (80 and 300 K). Response duration at half-height equals ∼80 μs. Duration of the rising edge for ∼300 ns is defined by a high value of film resistance of ∼10 kOhm and an input capacity of ∼20 pF for the registration system that gives an input time constant of ∼200 ns. Nonlinear correspondence between response amplitude and temperature is shown on Figure 6c. The decrease of temperature from 300 to 100 K leads to a doubling of response amplitude. At the temperature range from 120 to 220 K, dR/dT is temperature-independent, and the variation of response amplitude stems from changes in thermal properties of film/substrate interface. In order to obtain thermal characteristics of the film, mathematical simulation of bolometric response was performed. Film response simulation for two temperatures (80 and 300 K) are shown on Figure 6b.We note that simulation is in good agreement with experimental data. Figure 6c,d illustrates obtained thermal characteristics for bolometric structures. 28884

DOI: 10.1021/acsami.6b10145 ACS Appl. Mater. Interfaces 2016, 8, 28880−28887

Research Article

ACS Applied Materials & Interfaces

Figure 6. Bolometer produced from 12 mg mL−1 suspension at laser fluence 4 J cm−2 and line density of 20 lines per mm. (a) Resistance− temperature relationship for a sample. Slope of −27 Ω K−1 in temperature range 100−250 K and −27 Ω K−1 in temperature range 300−350 K. (b) Response of the sample at T = 80 and 300 K. The inset shows leading edge of responses. Thin black lines indicate the result of simulation. (c) Dependences of peak resistivity change (experimental data) and thermal conductivity of rGO film (simulation) on temperature. (d) Results of simulation for heat capacity and thermal boundary conductivity of the film/substrate interface.

Figure 7. (a) Schematic of the measurements; (b and c) characteristics of photoresponse at a voltage of 0.5 V, at different rates of switching on−off. Films produced from 12 mg mL−1 solution, laser fluence 4 J cm−2, and line density 20 lines per mm.

descends through the decrease of temperature faster than does graphite thermal conductivity (at the temperature of 80 K thermal conductivity of obtained sample reached the 15 W m−1

At room temperature thermal conductivity is rather high at ∼104 W m−1 K−1 (Figure 6c), close to the one for polycrystalline graphite15 or graphene laminate,33 but it 28885

DOI: 10.1021/acsami.6b10145 ACS Appl. Mater. Interfaces 2016, 8, 28880−28887

ACS Applied Materials & Interfaces K−1).15 High thermal conductivity of the rGO films may occur due to the large lateral size of graphene oxide sheets, ∼54 μm. The low density of the material can be the reason for rather low value of volumetric heat capacity. Boundary heat resistance stands for a heat resistance of 4 μm thick substrate layer. Figure 6d represents the heat capacity and thermal boundary conductivity. Figure 7 shows the realization of a 1 pixel bolometer. Au contacts were deposited on glass substrate; the distance between contacts was 1 mm. Contacts were covered with aqueous GO solution. GO solution then was dried for further laser treatment. A conventional incandescent lamp (50 W) located at a distance of 50 cm from sample was used as a light source. To focus the incandescent light, we used a cylindric lens with a focal length of 10 cm. Incident light intensity on substrate surface was at the value of 216 mW cm−2. Volt− ampere curves were measured with Keithley 6487 Picoammeter at constant voltages of 0.5 V. Incandescent light was irradiated on the structured surface of the substrate. Irradiation of incandescent lamp leads to a significant shift of current ∼0.4%. Studied materials show voltage responsivity of 0.17 V W−1 at 300 K. Using thermal parameters and TCR of rGO film, we estimated the volt−watt sensitivity of the designed bolometer structures at low frequencies. The results are shown in Figures S4. Here we represent the parameters for computer simulation on a film on substrate and a free-standing film. Maximum sensitivity that can be achieved is 42 V W−1 for free-standing films with a thickness of 4 μm at a current of 1 mA.

ACKNOWLEDGMENTS



REFERENCES

(1) Chattopadhyay, S.; Huang, Y. F.; Jen, Y. J.; Ganguly, A.; Chen, K. H.; Chen, L. C. Anti-Reflecting and Photonic Nanostructures. Mater. Sci. Eng., R 2010, 69 (1), 1−35. (2) Cui, Y.; He, Y.; Jin, Y.; Ding, F.; Yang, L.; Ye, Y.; Zhong, S.; Lin, Y.; He, S. Plasmonic and Metamaterial Structures as Electromagnetic Absorbers. Laser. Photonics. Rev. 2014, 8 (4), 495−520. (3) Poitras, D.; Dobrowolski, J. A. Toward Perfect Antireflection Coatings. Appl. Opt. 2004, 43 (6), 1286. (4) Cai, J.; Qi, L. Recent Advances in Antireflective Surfaces Based on Nanostructure Arrays. Mater. Horiz. 2015, 2 (1), 37−53. (5) Mizuno, K.; Ishii, J.; Kishida, H.; Hayamizu, Y.; Yasuda, S.; Futaba, D. N.; Yumura, M.; Hata, K. A Black Body Absorber From Vertically Aligned Single-Walled Carbon Nanotubes. Proc. Natl. Acad. Sci. U. S. A. 2009, 106 (15), 6044−6047. (6) Nair, R. R.; Blake, P.; Grigorenko, A. G.; Novoselov, K. S.; Booth, T. J.; Stauber, T.; Peres, N. M. R.; Geim, A. K. Fine Structure Constant Defines Visual Transparency of Graphene. Science 2008, 320 (5881), 1308−1308. (7) Yan, J.; Kim, M. H.; Elle, J. A.; Sushkov, A. B.; Jenkins, G. S.; Milchberg, H. W. M.; Fuhrer, M. S.; Drew, H. D. Dual-Gated Bilayer Graphene Hot-Electron Bolometer. Nat. Nanotechnol. 2012, 7, 472− 478. (8) Dickerson, W.; Hemsworth, N.; Gaskell, P.; Ledwosinska, E.; Szkopek, T. Bolometric Response of Free-Standing Reduced Graphene Oxide Films. Appl. Phys. Lett. 2015, 107 (24), 243103. (9) Itkis, M. E.; Borondics, F.; Yu, A.; Haddon, R. C. Bolometric Infrared Photoresponse of Suspended Single-Walled Carbon Nanotube Films. Science 2006, 312, 413−416. (10) Hazra, K. Sh.; Sion, N.; Yadav, A.; McLauhglin, J.; Misra, D. Sh. Vertically Aligned Graphene Based Non-Cryogenic Bolometer. ArXiv, 2013, arXiv:1301.1302v1. (11) Koppens, F. H. L.; Mueller, T.; Avouris, P.; Ferrari, A. C.; Vitiello, M. S.; Polini, M. Photodetectors Based on Graphene, Other Two-Dimensional Materials and Hybrid Systems. Nat. Nanotechnol. 2014, 9 (10), 780−793. (12) Shi, H.; Wang, C.; Sun, Z.; Zhou, Y.; Jin, K.; Redfern, S. A.; Yang, G. Tuning the Nonlinear Optical Absorption of Reduced Graphene Oxide by Chemical Reduction. Opt. Express 2014, 22 (16), 19375−19385. (13) Wang, J.; Chen, Y.; Blau, W. J. Carbon Nanotubes and Nanotube Composites for Nonlinear Optical Devices. J. Mater. Chem. 2009, 19 (40), 7425−7443. (14) Liu, M.; Yin, X.; Ulin-Avila, E.; Geng, B.; Zentgraf, T.; Ju, L.; Wang, F.; Zhang, X. A Graphene-Based Broadband Optical Modulator. Nature 2011, 474 (7349), 64−67. (15) Balandin, A. A.Thermal Properties of Graphene and Nanostructured Carbon Materials. Nat. Mater. 2011, 10 (8), 569−581. (16) Balandin, A. A.; Ghosh, S.; Bao, W.; Calizo, I.; Teweldebrhan, D.; Miao, F.; Lau, C. N. Superior Thermal Conductivity of SingleLayer Graphene. Nano Lett. 2008, 8, 902. (17) Liang, H. Mid-Infrared Response of Reduced Graphene Oxide and its High-Temperature Coefficient of Resistance. AIP Adv. 2014, 4 (10), 107131. (18) Bae, J.; Yoon, J.; Jeong, S.; Moon, B.; Han, J.; Jeong, H.; Lee, Y.; Hwang, H.; Lee, Y.; Jeong, Y.; Lim, S. Sensitive Photo-Thermal Response of Grapheme Oxide for Mid-Infrared Detection. Nanoscale 2015, 7 (38), 15695−15700. (19) Krivchenko, V. A.; Evlashin, S. A.; Mironovich, K. V.; Verbitskiy, N. I.; Nefedov, A.; Wöll, C.; Kozmenkova, A.; Suetin, N. V.; Svyakhovskiy, S. E.; Vyalikh, D. V.; Rakhimov, A. T.; Egorov, A. V.;

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.6b10145. Photo of the sample, Raman and FTIR spectra, mirror and angular dependencies, resistance change dependencies on frequency, comparison of various bolometric material (PDF)





This work is supported by the Russian Science Foundation, grant 14-13-01422. XPS measurements were supported by M. V. Lomonosov Moscow State University Program of Development.

4. CONCLUSIONS We have realized a bolometer prototype and investigated the properties of the derived material in wide temperature and wavelength ranges. The developed method allows us to carry out a controlled process of GO microstructuring, which in turn paves the way toward real-world applications. Due to the remarkably high values of rGO conductivity and GO resistance, various types of surfaces can be coated with the obtained structured films. The study of thermal and optical properties of rGO produced by laser microstructuring has been performed. For samples of several micrometer film thickness, we have obtained an absorption of more than 98% for visible and 90% for IR. The resistance of the samples, measured in a wide temperature range, followed the linear dependence with a slope value of −13 Ohm K−1, while TCR was 0.3% K−1 at 300 K.



Research Article

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest. 28886

DOI: 10.1021/acsami.6b10145 ACS Appl. Mater. Interfaces 2016, 8, 28880−28887

Research Article

ACS Applied Materials & Interfaces Yashina, L. V. Carbon Nanowalls: the Next Step for Physical Manifestation of the Black Body Coating. Sci. Rep. 2013, 3, 3328. (20) Evlashin, S.; Svyakhovskiy, S.; Suetin, N.; Pilevsky, A.; Murzina, T.; Novikova, N.; Stepanov, A.; Egorov, A.; Rakhimov, A. Optical and IR Absorption of Multilayer Carbon Nanowalls. Carbon 2014, 70, 111−118. (21) Hummers, W. S., Jr.; Offeman, R. E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80 (6), 1339−1339. (22) Klokov, A. Y.; Aminev, D. F.; Sharkov, A. I.; Ral’chenko, V. G.; Galkina, T. I. Thermal Parameters of Layers and Interfaces in Siliconon-Diamond Structures. Phys. Solid State 2008, 50 (12), 2263−2269. (23) Arul, R.; Oosterbeek, R. N.; Robertson, J.; Xu, G.; Jin, J.; Simpson, M. C. The Mechanism of Direct Laser Writing of Graphene Features into Graphene Oxide Films Involves Photoreduction and Thermally Assisted Structural Rearrangement. Carbon 2016, 99, 423− 431. (24) Sokolov, D. A.; Rouleau, C. M.; Geohegan, D. B.; Orlando, T. M. Excimer Laser Reduction and Patterning of Graphite Oxide. Carbon 2013, 53, 81−89. (25) Zhang, J.; Yang, H.; Shen, G.; Cheng, P.; Zhang, J.; Guo, S. Reduction of Graphene Oxide via L-Ascorbic Acid. Chem. Commun. 2010, 46 (7), 1112−1114. (26) Ferrari, A. C.; Robertson, J. Resonant Raman Spectroscopy of Disordered, Amorphous, and Diamondlike Carbon. Phys. Rev. B: Condens. Matter Mater. Phys. 2001, 64 (7), 075414. (27) Compagnini, G.; Giannazzo, F.; Sonde, S.; Raineri, V.; Rimini, E. Rimini, E. Ion Irradiation and Defect Formation in Single Layer Graphene. Carbon 2009, 47 (14), 3201−3207. (28) Cancado, L. G.; Takai, K.; Enoki, T.; Endo, M.; Kim, Y. A.; Mizusaki, H.; Jorio, A.; Coelho, L. N.; Magalhães-Paniago, R.; Pimenta, M. A. General Equation for the Determination of the Crystallite Size La of Nanographite by Raman Spectroscopy. Appl. Phys. Lett. 2006, 88, 163106. (29) Tuinstra, F.; Koenig, J. L. Raman Spectrum of Graphite. J. Chem. Phys. 1970, 53 (3), 1126−1130. (30) Krishnan, R. S. Temperature Variations of the Raman Frequencies in Diamond. Proc. - Indian Acad. Sci., Sect. A 1946, 24 (1), 45−57. (31) Shen, Y.; Yang, S.; Zhou, P.; Sun, Q.; Wang, P.; Wan, L.; Li, J.; Chen, L.; Wang, X.; Ding, Sh.; Zhang, D. W. Evolution of the BandGap and Optical Properties of Graphene Oxide with Controllable Reduction Level. Carbon 2013, 62, 157−164. (32) Gao, W.; Singh, N.; Song, L.; Liu, Z.; Reddy, A. L. M.; Ci, L.; Vajtai, R.; Zhang, Q.; Wei, B.; Ajayan, P. M. Direct Laser Writing of Micro-Supercapacitors on Hydrated Graphite Oxide Films. Nat. Nanotechnol. 2011, 6 (8), 496−500. (33) Malekpour, H.; Chang, K. H.; Chen, J. C.; Lu, C. Y.; Nika, D. L.; Novoselov, K. S.; Balandin, A. A. Thermal Conductivity of Graphene Laminate. Nano Lett. 2014, 14 (9), 5155−5161.

28887

DOI: 10.1021/acsami.6b10145 ACS Appl. Mater. Interfaces 2016, 8, 28880−28887