You are on page 1of 19

Version of Record: https://www.sciencedirect.

com/science/article/pii/S0925346719307608
Manuscript_60cb0f1a06d358c865a36e8e510d5ef1

Laser patterning and induced reduction of graphene


oxide functionalized silk fibroin
Kelly T. Paula1, Molíria V. Santos1, Murilo. H. M. Facure3,4, Marcelo B. Andrade1, Francineide L.
Araújo1, Daniel S. Correa3,4, Sidney J. L. Ribeiro2, *Cleber R. Mendonça1
1
São Carlos Institute of physics, University of São Paulo, 13560-970, São Carlos, SP, Brazil
2
Araraquara Institute of Chemistry, São Paulo State University,14800-060, Araraquara, SP, Brazil
3
Nanotechnology National Laboratory for Agriculture (LNNA), Embrapa Instrumentação 13560-970,
São Carlos, SP, Brazil
4
PPGQ, Department of Chemistry, Center for Exact Sciences and Technology, Federal University of São
Carlos, 13565-905, São Carlos, SP, Brazil

Abstract
This paper presents a one-step method for patterning and reducing graphene
oxide functionalized in silk fibroin using femtosecond laser induced forward transfer
(LIFT). Such approach renders fibroin, a natural biopolymer with excellent
biocompatibility, chemical stability, and mechanical properties, improved electrical
conductivity in a controlled and localized way. Composite films based on silk fibroin
and graphene oxide/reduced graphene oxide were fabricated by spin coating at room
temperature. The films composition was characterized by micro-Raman spectroscopy
demonstrating the presence of graphene oxide in the silk fibroin matrix. The influence
of pulse energy on the fs laser transfer process was investigated, yielding a threshold
energy of 24.0 nJ. Additionally, the laser-induced reduction of graphene oxide attained
by LIFT was determined by Raman spectroscopy and further confirmed by an increase
in the electrical conductivity. The approach presented here proved to be an efficient
route to simultaneously produce micrometric patterns of graphene oxide/silk fibroin
composite (with line widths on the order of 1 µm), and reduction of graphene oxide,
opening new opportunities for the development of green composites for electronics
devices, including for instance microfluidic sensors, capacitors and LEDs.

*Corresponding author. Tel: +55 16 3373-8085 ext. 236 1


E-mail: crmendon@ifsc.usp.br
© 2019 published by Elsevier. This manuscript is made available under the Elsevier user license
https://www.elsevier.com/open-access/userlicense/1.0/
1. Introduction
In recent years the interest in graphene has become the subject of intense
research due to its unique thermal[1,2], mechanical[3], chemical[4] and optical[5]
properties, being an excellent candidate for a variety of applications in nanotechnology.
However, its cost-effective mass production and difficulties with solubility and
processing are barriers for the development of further applications using graphene.
Graphene-based materials, such as graphene oxide (GO), produced by the oxidation of
graphite, and its reduced counterpart (rGO) have been extensively studied because of
their similarities to graphene, as well as due to the presence of functional groups on
their basal planes that allows solubility in a variety of solvents [6,7], and compatibility
with other materials [8], allowing their applications in biodevices[9–11], chemical
sensors[12–14] and photonics[15,16].
Patterning of graphene-based materials for microelectronic devices has also been
receiving considerable attention[17–19]. At the same time, novel transfer/printing
methods have been developed for fabricating high-resolution patterns, such as direct
laser writing (DLW)[20,21], as alternatives to standard lithographic processes. In this
direction, Laser Induced Forward Transfer (LIFT)[22], which is a versatile and non-
destructive method of patterning, have been attracting the attention of researchers in
different fields of sciences. LIFT employs laser radiation to transfer a donor material to
different types of substrates, and has been used for the microscale transfer of a variety
of materials, including metals[23], semiconductors[24], 2D materials[25,26],
polymers[27], metallic pastes[28,29] and biomaterials[30], with potential use as
chemical sensors, conductive structures, organic and optical devices[31–34]. More
specifically, the LIFT technique employs a pulsed laser to transfer a donor material
coated on a transparent substrate placed in close proximity or in contact with the
receiver substrate. A laser pulse with enough intensity is focused onto the donor,
heating and ejecting the material towards the receiver, printing pixels or droplets in it. In
this context, it is worth mentioning the LIFT carried out with femtosecond laser pulses
(fs-LIFT), which offers advantages in terms of deposition of smaller size features and
intact material, thanks to the reduced thermal effects and minimal collateral damage[35]
resulting from the nonlinear light-matter interaction.
In this work, we used fs-LIFT to fabricated graphene oxide functionalized silk
fibroin micrometric patterns on a glass substrate. In such structures, graphene oxide was

2
dispersed in silk fibroin, which is a natural biopolymer that presents biocompatibility,
chemical stability, and mechanical properties[36,37]. Scanning electron and atomic
force microscopy results revealed a localized, controllable and reproducible deposition,
with lines widths on the order of 1 µm, when the proper parameters are employed.
Furthermore, Raman spectroscopy reveals the reduction of graphene oxide upon fs-
LIFT, which is in agreement with the observed increase in the electrical conductivity of
the transferred material, opening perspectives on the use of graphene/biopolymer
composites for the development of electronic and biomedical devices.

2. Experimental

2.1. Synthesis of Silk Fibroin

Silk Fibroin (SF) solution was prepared using natural silk cocoons (Bombyx mori
silkworms) supplied by Bratac (Fiação de Seda S.A. - Bastos/SP, Brazil). Initially, 5 g
of silk cocoon pieces were diluted in 2 L of Na2CO3 solution (0.02 M) at 100 °C for 30
minutes and washed thoroughly with distilled water to remove sericin (degumming).
Degummed silk was then dissolved in 9.3 M LiBr, in a proportion of 1:4 (w/v) of
fibroin to LiBr, for 4 hours at 60 °C for dissolution. Each 25 mL of the resulting
solution was dialyzed against 2 L of milli-Q water for 48 hours to remove salts. An
8.5% (w/V) aqueous fibroin solution, free of impurities, was obtained after the
centrifugation (20,000 rpm) of the dialyzed solution at 4 oC for 20 min. The final
solution was stored at 4 oC before use.

2.2. Graphene Oxide and reduced Graphene Oxide

The Graphene Oxide (GO) was purchased from Sigma Aldrich (powder, 4-10%
edge-oxidized). The reduced Graphene Oxide (rGO) was obtained through the chemical
reduction of GO using ascorbic acid (AA). Ascorbic acid was employed because its use
as a reducing agent gives rise to a rGO with a high degree of reduction and
consequently to a material with desired properties (e.g. high electrical conductivity)
[38]. In the synthesis, 150 mg of AA was added to a 100 mL GO dispersion (0.5 mg
mL-1). The system was then maintained at 100 ºC under reflux for 24 h. The obtained
rGO was washed with double-distilled water to remove the excess of AA and
redispersed for further use.

2.3. Preparation of functionalized (GO or rGO) silk fibroin films

3
Thin films were prepared from the silk fibroin solution described before, which was
doped with GO or with rGO. First, aqueous dispersions of GO and rGO with
concentrations of 3 mg/ml was sonicated for 1 h to minimize the formation of clusters
and then added in the silk fibroin solution. The SF/GO and SF/rGO solutions were
mechanically stirred for 10 minutes. Each sample (donor film) was prepared by spin
coating the final solution onto glass substrates using 350 rpm for 10 minutes at room
temperature. This approach resulted in functionalized (GO or rGO) silk fibroin films of
approximately 3.0 µm of thickness. In addition, to optimize the GO concentration we
performed tests with higher and lower concentrations. GO concentrations higher than
approximately 4 mg/ml resulted in low-quality films due to aggregation, associated with
difficulties in the incorporation of GO into the SF solution. A lower concentration
sample (~ 1 mg/ml) was also prepared, but presented electrical properties similar to the
pure SF sample, considering the experimental error.

2.4. Femtosecond Laser Induced Forward Transfer experimental setup

LIFT experimental setup (illustrated in Fig. 1) employed a Ti: Sapphire laser


oscillator (50-fs pulses) centered at 800 nm with a repetition rate of 5 MHz. The laser
beam was focused at the interface of the donor material and its substrate using a
microscope objective with a numerical aperture (NA) of 0.65 (40×). Both donor and
receptor substrate were maintained in contact and placed on an x-y-z translation stage
that moved at a constant speed. The pulse energy and scanning speed were varied to
determine the optimal irradiation parameters. A CCD camera aligned to the microscope
objective allowed monitoring the entire process. The receiver glass substrate was
cleaned with NaOH (0.1 M) in an ultrasonic bath to enhance the adhesion of the donor
material.

2.5. Characterization

Raman spectroscopy of the samples was carried out using a confocal LabRAM
micro-Raman system, with a solid-state laser operating at 532 nm as the excitation
source and a 1800 g.mm-1 grating. The spectral resolution was 1 cm-1. The transferred
patterns were characterized by atomic force microscopy (AFM) using a Nanosurf’s
easyScan 2®, as well as by optical microscopy (Zeiss LSM 700) and scanning electron

4
microscopy (SEM) using a FEI’s Inspect F50 microscope (Eindhoven, The
Netherlands). Raman spectra of the LIFTed samples were carried out using Surface
Enhanced Raman Spectroscopy (SERS) [39].
For the electrical measurements, functionalized SF (modified with GO or rGO)
was transferred to glass substrates in which Gold (Au) electrodes were previously
deposited by thermal evaporation. A holder containing a metal shade mask was used to
define the areas of interest of the electrodes. At the end of the evaporation, each
substrate contained three Au electrodes (100 nm of thickness), formed by two contacts
separated by 1 mm gap. The electrical properties of the fs-LIFT patterns were measured
with a Keithley 6487 Picoammeter scanning using a DC voltage from -2.5 V to 2.5 V.

Figure 1 - LIFT experimental setup illustrating the main elements of the system.

3. Results and discussion

3.1. Morphology and characterization of the LIFTed lines

SEM images of lines (separated by 40 µm) transferred to the receptor substrate


by fs-LIFT using pulse energy of 36 nJ and scanning speed of 50 µm/s (whose
parameters were optimized in a set of subsidiary experiments), are shown in Fig. 2a
(SF), Fig. 2b (GO/SF) and Fig. 2c (rGO/SF), and reveal the high resolution achieved by
the process. Such images also show a controllable and reproducible deposition with line
widths on the order of 1 µm.

5
Figure 2 - Scanning electron microscopy images of fs-LIFT micromachined lines with 40 µm gap
obtained for Silk (a), GO-Silk (b) and rGO-Silk (c).

To investigate the morphology of the lines produced by fs-LIFT, a 65×65 µm2


area was inspected by AFM. In Fig 3a, as an example, it is presented an AFM
micrograph of two SF fs-LIFT lines (36 nJ and 50 µm/s) separated by 40 µm. The
profile reveals that such lines present a thickness of about 204 nm. For GO/SF and
rGO/SF, we obtained thicknesses of approximately 180 nm and 186 nm, respectively.
The average roughness of the SF fs-LIFTed lines was determined as 18 nm (Fig. 3b).
For GO/SF and rGO/SF, the average roughness values determined were 29 nm and 22
nm, respectively. From the AFM measurements, we noticed that the thickness of the
deposited lines increases with the pulse energy, varying from approximately 190 nm to
230 nm when the pulse energy increases from 30 to 50 nJ. The same goes for
transferred GO/SF and rGO/SF, with line thickness varying from approximately 140 nm
to 270 nm and 160 nm to 240 nm, respectively.

6
Figure 3 – AFM 3D image of the fs-LIFT lines of 40 µm gap obtained using laser pulse energy of 36 nJ
and scanning speed of 50 µm/s (a) and roughness profile of the obtained lines (b).

In order to determine the influence of pulse energy on the LIFT process, groups
of 500-µm lines separated by 40 µm were deposited on the receptor substrate using a
0.65 NA microscope objective, scanning speed of 50 µm/s and pulse energy varying
from 30 to 50 nJ. Such lines were analyzed by optical microscopy; their widths increase
with the pulse energy, varying from approximately 0.89 to 1.45 µm when the pulse
energy increases from 30 to 50 nJ.
Figure 4 presents the line width squared as a function of pulse energy (log scale).
By fitting the data presented in Figs. 4a and 4b, using the approach presented in
Ref.[40], we determined a threshold energy of 24.0 nJ for both, GO/SF and rGO/SF.

7
Figure 4 - Line width squares as a function of the pulse energy using scanning speed of 50 µm/s for
GO/SF (a) and rGO/SF (b) obtained from LIFT experiments.

3.2. Analysis and comparison of Raman spectra

Figure 5 shows the Raman spectra of donor film SF (black line), GO/SF (gray
line) and rGO/SF (light-gray line). The SF spectrum exhibits the characteristic peaks in
the region between 1250 and 1800 cm-1. The band at 1663 cm-1 corresponds to the
Amide I mode [41–43]. At 1615 cm-1, the peak observed is assigned to the tryptophan,
phenylalanine, and tyrosine vibrations [44,45]. The Raman bands at 1549 cm-1, 1448
cm-1 and 1330 cm-1 are assigned to tryptophan vibration [44,46], CH2 bending and CH3

8
anti-symmetric bending [47,48], and CH3 symmetric bending mode [47]. As shown in
Fig. 5, the GO/SF Raman spectrum exhibits the G peak at 1563 cm-1, corresponding to
the first order scattering of the E2g mode [49], while the D band is a breathing mode of
κ-point phonons of A1g symmetry. The Raman spectrum of the rGO/SF also contains
both G and D bands, at 1572 and 1339 cm-1, respectively, with an increased D/G
intensity ratio compared to that in GO/SF.

Figure 5 - Raman spectra of SF (black line), SF/GO (gray line) and SF/rGO (light-gray line).

As mentioned before, Raman spectra of LIFTed samples were carried out using
SERS using gold nanoparticles. Since the transferred materials were deposited onto
glass substrates, measurements on the glass slide were also carried out. The black lines
in Fig. 6a and 6b present the Raman spectrum of GO/SF and rGO/SF donor films,
respectively. From such data, it was obtained ID/IG of 0.61 and 0.95 for the GO/SF and
rGO/SF, respectively. The gray lines in Fig. 6a and 6b correspond to the spectrum for
the LIFTed samples that display the characteristic peaks of SF and GO. In this case, it
was found an ID/IG of 1.04 and 1.09 for the transferred GO/SF and rGO/SF,
respectively. In addition, there is no band shift in the Raman spectrum after the LIFT
process. Therefore, the ID/IG of transferred GO/SF and rGO/SF are very similar and

9
larger than that obtained for GO/SF donor, indicating the reduction of GO upon the
LIFT process.

Figure 6 - Raman spectrum of Donor (black line) and LIFTed (gray line) samples: a) SF/GO and b)
SF/rGO.

3.3. Electrical properties

In order to evaluate the changes in conductivity of GO/SF and rGO/SF donor


films aiming at applications in electronic devices, the electrical resistivity of samples

10
was measured using the four-point probe method. Considering rectangular samples
deposited on a non-conducting substrate of finite length a, finite width d of thickness w,
under the condition w < 4/10 s, the resistivity is given by [50]

= ( , , ) (1)

in which s is the distance between the probes, V is the voltage, I is the current and
( , , ) is a correction factor that depends on the dimension of the film and the
distance between the probes. After measuring the resistance, the resistivity of the pure
SF, GO/SF and rGO/SF donor films were calculated, providing values of 6.6 × 105
Ω.m, 6 × 105 Ω.m, and 2 × 104 Ω.m, respectively. As mentioned previously, higher
GO concentrations than the one described resulted in low-quality films. For the lower
GO content sample, the resistivity obtained is the same as the one observed for the pure
SF.
To investigate the resistivity of GO/SF and rGO/SF samples deposited by fs-
LIFT, a 1 mm area was transferred onto Gold (Au) electrodes, as illustrated in Fig. 7.
Figure 7a and 7b show SEM images of the transferred material in each case. The
thicknesses of the patterns were estimated by AFM as 180 nm for GO/SF and 186 nm
for rGO/SF. From the I-V curves shown in Fig. 8a and 8b for GO/SF and rGO/SF,
respectively, the electrical resistance (R) of the sample was obtained. The linear
dependence observed in Fig. 8 indicates an ohmic behavior, whose corresponding
electrical resistivities are ~ 300 Ω.m and ~ 40 Ω.m, for GO/SF and rGO/SF,
respectively. Such values are on the same order of those observed for other graphene
oxide-based composites and other carbon materials [8],[51],[52].

11
a b

Figure 7 - SEM image of a) GO/SF and b) rGO/SF transferred between gold electrodes. The scheme
located on the upper part of SEM images represents GO/SF and rGO/SF samples deposited by fs-LIFT
and transferred onto the Gold (Au) electrodes.

12
Figure 8 – Current – Voltage curve of transferred a) GO/SF and b) rGO/SF, displaying ohmic behavior
and electrical resistivities of 300 Ω.m and 40 Ω.m, respectively.

4. Conclusions
We demonstrate micro-patterning and reduction of graphene oxide
functionalized on silk fibroin films by fs-LIFT in an efficient, controlled and localized
way. The threshold energy for fs-LIFT for both, GO/SF and rGO/SF, were determined
as 24.0 nJ. Spectroscopy analyses of Raman confirmed the presence of graphene oxide
in the silk fibroin matrix and also indicated the reduction of GO upon the fs-LIFT
process. Scanning electron and atomic force microscopies revealed a controllable and

13
reproducible deposition, with line widths on the order of 1 µm. Additionally, an
increase on the electrical conductivity of the fs-LIFTed material was observed,
corroborating the reduction of the GO dispersed in SF. Our results indicate the potential
of fs-LIFT approach for designing green composites for electronics devices, including
microfluidic sensors and capacitors.

Acknowledgments

Authors are grateful to São Paulo Research Foundation (FAPESP, grants 2018/1283-7,
2016/11591-8, 2013/03487-8), Financial support from the Coordenação de
Aperfeiçoamento de Pessoal de Nível Superior (CAPES) - Finance Code 001, CNPq,
Army Research Laboratory W911NF-17-1-0123, Air Force Office of Scientifc Research
(FA9550-12-1-0028) and also Bruno B. M. Torres for the use of the four-point probe
station in the electrical characterization of the donor films.

References:
[1] D. Teweldebrhan, C.N. Lau, S. Ghosh, A.A. Balandin, W. Bao, I. Calizo, F.
Miao, Superior Thermal Conductivity of Single-Layer Graphene, Nano Lett. 8
(2008) 902–907. doi:10.1021/nl0731872.
[2] A.A. Balandin, Thermal properties of graphene and nanostructured carbon
materials, Nat. Mater. 10 (2011) 569–581. doi:10.1038/nmat3064.
[3] C. Lee, X. Wei, J.W. Kysar, J. Hone, Measurement of the Elastic Properties and
Intrinsic Strength of Monolayer Graphene, Science. 321 (2008) 385–388.
[4] F. Schedin, A.K. Geim, S. V. Morozov, E.W. Hill, P. Blake, M.I. Katsnelson,
K.S. Novoselov, Detection of individual gas molecules adsorbed on graphene,
Nat. Mater. 6 (2007) 652–655. doi:10.1038/nmat1967.
[5] F. Bonaccorso, Z. Sun, T. Hasan, A.C. Ferrari, Graphene Photonics and
Optoelectronics, Nat. Photonics. 4 (2010) 611–622.
doi:10.1038/nphoton.2010.186.
[6] S. Park, J. An, I. Jung, R.D. Piner, J. An, X. Li, A. Velamakanni, R.S. Ruoff, S.
Park, J. An, I. Jung, R.D. Piner, S.J. An, X. Li, A. Velamakanni, R.S. Ruoff,
Oxide in a Wide Variety of Organic Solvents Colloidal Suspensions of Highly
Reduced Graphene Oxide in a Wide Variety of Organic Solvents, (2009) 1–5.

14
doi:10.1021/nl803798y.
[7] J.I. Paredes, S.V. Rodil, A.M. Alonso, J.M.D. Tascón, Graphene Oxide
Dispersions in Organic Solvents Partners, 121 (2017) 1–2.
doi:10.1021/la801744a.
[8] S. Stankovich, D.A. Dikin, G.H.B. Dommett, K.M. Kohlhaas, E.J. Zimney, E.A.
Stach, R.D. Piner, S.B.T. Nguyen, R.S. Ruoff, Graphene-based composite
materials, Nature. 442 (2006) 282–286. doi:10.1038/nature04969.
[9] K.H. Liao, Y.S. Lin, C.W. MacOsko, C.L. Haynes, Cytotoxicity of graphene
oxide and graphene in human erythrocytes and skin fibroblasts, ACS Appl.
Mater. Interfaces. 3 (2011) 2607–2615. doi:10.1021/am200428v.
[10] S. Mishra, M. Ashaduzzaman, P. Mishra, H.C. Swart, A.P.F. Turner, A. Tiwari,
Stimuli-enabled zipper-like graphene interface for auto-switchable bioelectronics,
Biosens. Bioelectron. 89 (2017) 305–311. doi:10.1016/j.bios.2016.03.052.
[11] N. Mohanty, V. Berry, Resolution Biodevice and DNA Transistor : Interfacing
Graphene Derivatives with Nanoscale and Microscale Biocomponents, Am.
Chem. Soc. 8 (2008) 4469–4476. doi:10.1021/nl802412n.
[12] P.E. Sheehan, E.S. Snow, F.K. Perkins, J.T. Robinson, Z. Wei, Reduced
Graphene Oxide Molecular Sensors, Nano Lett. 8 (2008) 3137–3140.
doi:10.1021/nl8013007.
[13] R.B. Kaner, B.H. Weiller, Y. Yang, M.J. Allen, V.C. Tung, J.D. Fowler, Practical
Chemical Sensors from Chemically Derived Graphene, ACS Nano. 3 (2009)
301–306. doi:10.1021/nn800593m.
[14] X.C. Dong, H. Xu, X.W. Wang, Y.X. Huang, M.B. Chan-Park, H. Zhang, L.H.
Wang, W. Huang, P. Chen, 3D graphene-cobalt oxide electrode for high-
performance supercapacitor and enzymeless glucose detection, ACS Nano. 6
(2012) 3206–3213. doi:10.1021/nn300097q.
[15] G. Sobon, J. Sotor, J. Jagiello, R. Kozinski, M. Zdrojek, M. Holdynski, P.
Paletko, J. Boguslawski, L. Lipinska, K.M. Abramski, Graphene oxide vs.
reduced graphene oxide as saturable absorbers for Er-doped passively mode-
locked fiber laser., Opt. Express. 20 (2012) 19463–73.
doi:10.1364/OE.20.019463.
[16] H.X. Wang, Q. Wang, K.G. Zhou, H.L. Zhang, Graphene in light: Design,
synthesis and applications of photo-active graphene and graphene-like materials,
Small. 9 (2013) 1266–1283. doi:10.1002/smll.201203040.

15
[17] S.R. Das, Q. Nian, A.A. Cargill, J.A. Hondred, S. Ding, M. Saei, G.J. Cheng, J.C.
Claussen, 3D nanostructured inkjet printed graphene: Via UV-pulsed laser
irradiation enables paper-based electronics and electrochemical devices,
Nanoscale. 8 (2016) 15870–15879. doi:10.1039/c6nr04310k.
[18] D. Song, A. Mahajan, E.B. Secor, M.C. Hersam, L.F. Francis, C.D. Frisbie,
High-Resolution Transfer Printing of Graphene Lines for Fully Printed, Flexible
Electronics, ACS Nano. 11 (2017) 7431–7439. doi:10.1021/acsnano.7b03795.
[19] D. Song, E.B. Secor, Y. Wang, M.C. Hersam, C. Daniel Frisbie, Transfer
Printing of Sub-5 μm Graphene Electrodes for Flexible Microsupercapacitors,
ACS Appl. Mater. Interfaces. 10 (2018) 22303–22310.
doi:10.1021/acsami.8b06235.
[20] S. Papazoglou, V. Tsouti, S. Chatzandroulis, I. Zergioti, Direct laser printing of
graphene oxide for resistive chemosensors, Opt. Laser Technol. 82 (2016) 163–
169. doi:10.1016/j.optlastec.2016.03.009.
[21] E.C.P. Smits, A. Walter, D.M. De Leeuw, K. Asadi, Laser induced forward
transfer of graphene, Appl. Phys. Lett. 111 (2017) 1–5. doi:10.1063/1.5001712.
[22] J. Bohandy, B.F. Kim, F.J. Adrian, Metal deposition from a supported metal film
using an excimer laser, J. Appl. Phys. 60 (1986) 1538–1539.
doi:10.1063/1.337287.
[23] F.J. Adrian, A study of the mechanism of metal deposition by the laser-induced
forward transfer process, J. Vac. Sci. Technol. B Microelectron. Nanom. Struct. 5
(2002) 1490. doi:10.1116/1.583661.
[24] L. Rapp, C. Cibert, S. Nénon, A.P. Alloncle, M. Nagel, T. Lippert, C. Videlot-
Ackermann, F. Fages, P. Delaporte, Improvement in semiconductor laser printing
using a sacrificial protecting layer for organic thin-film transistors fabrication,
Appl. Surf. Sci. 257 (2011) 5245–5249. doi:10.1016/j.apsusc.2010.10.147.
[25] G. Li, X. Mo, W.C. Law, K.C. Chan, 3D printed graphene/nickel electrodes for
high areal capacitance electrochemical storage, J. Mater. Chem. A. 7 (2019)
4055–4062. doi:10.1039/c8ta11121a.
[26] N.T. Goodfriend, S.Y. Heng, O.A. Nerushev, A. V. Gromov, A. V. Bulgakov, M.
Okada, W. Xu, R. Kitaura, J. Warner, H. Shinohara, E.E.B. Campbell, Blister-
based-laser-induced-forward-transfer: A non-contact, dry laser-based transfer
method for nanomaterials, Nanotechnology. 29 (2018). doi:10.1088/1361-
6528/aaceda.

16
[27] J. Shaw-Stewart, T. Lippert, M. Nagel, F. Nüesch, A. Wokaun, Laser-induced
forward transfer of polymer light-emitting diode pixels with increased charge
injection, ACS Appl. Mater. Interfaces. 3 (2011) 309–316.
doi:10.1021/am100943f.
[28] J. Wang, R.C.Y. Auyeung, H. Kim, N.A. Charipar, A. Piqué, Three-dimensional
printing of interconnects by laser direct-write of silver nanopastes, Adv. Mater.
22 (2010) 4462–4466. doi:10.1002/adma.201001729.
[29] E. Turkoz, M. Morales, S.Y. Kang, A. Perazzo, H.A. Stone, C. Molpeceres, C.B.
Arnold, Laser-induced forward transfer from healing silver paste films, Appl.
Phys. Lett. 113 (2018) 0–5. doi:10.1063/1.5060717.
[30] J.M. Fernández-Pradas, M. Colina, P. Serra, J. Domínguez, J.L. Morenza, Laser-
induced forward transfer of biomolecules, Thin Solid Films. 453–454 (2004) 27–
30. doi:10.1016/j.tsf.2003.11.154.
[31] P. Sopeña, J. Arrese, S. González-Torres, J.M. Fernández-Pradas, A. Cirera, P.
Serra, Low-Cost Fabrication of Printed Electronics Devices through Continuous
Wave Laser-Induced Forward Transfer, ACS Appl. Mater. Interfaces. 9 (2017)
29412–29417. doi:10.1021/acsami.7b04409.
[32] D.M. Zhigunov, A.B. Evlyukhin, A.S. Shalin, U. Zywietz, B.N. Chichkov,
Femtosecond Laser Printing of Single Ge and SiGe Nanoparticles with Electric
and Magnetic Optical Resonances, ACS Photonics. 5 (2018) 977–983.
doi:10.1021/acsphotonics.7b01275.
[33] P. Sopeña, S. González-Torres, J.M. Fernández-Pradas, P. Serra, Spraying
dynamics in continuous wave laser printing of conductive inks, Sci. Rep. 8
(2018) 1–12. doi:10.1038/s41598-018-26304-9.
[34] A. Palla Papavlu, T. Mattle, S. Temmel, U. Lehmann, A. Hintennach, A. Grisel,
A. Wokaun, T. Lippert, Highly sensitive SnO 2 sensor via reactive laser-induced
transfer, Sci. Rep. 6 (2016) 1–9. doi:10.1038/srep25144.
[35] D. Banks, K. Kaur, C. Grivas, R. Fardel, Femtosecond laser induced forward
transfer for the deposition of nanoscale transparent and solid-phase materials,
Proc. LAMP. (2009) 1–9. http://eprints.soton.ac.uk/79005/.
[36] R.F.P. Pereira, M.M. Silva, V. De Zea Bermudez, Bombyx mori Silk Fibers: An
Outstanding Family of Materials, Macromol. Mater. Eng. 300 (2015) 1171–1198.
doi:10.1002/mame.201400276.
[37] L.D. Koh, Y. Cheng, C.P. Teng, Y.W. Khin, X.J. Loh, S.Y. Tee, M. Low, E. Ye,

17
H.D. Yu, Y.W. Zhang, M.Y. Han, Structures, mechanical properties and
applications of silk fibroin materials, Prog. Polym. Sci. 46 (2015) 86–110.
doi:10.1016/j.progpolymsci.2015.02.001.
[38] L. Guardia, J.I.L. Paredes, P. Solı, J.M.D. Tasco, M.J. Fernandez-Merino, L.
Guardia, J.I.L. Paredes, S. Villar-Rodil, P.Solis-Fernandez, A. Martinez-Alonso,
J.M.D. Tascon, Vitamin C Is an Ideal Substitute for Hydrazine in the Reduction
of Graphene Oxide Suspensions ´ n, J. Phys. Chem. C. 114 (2010) 6426–6432.
[39] B. Sharma, R.R. Frontiera, A.-I. Henry, E. Ringe, R.P. Van Duyne, SERS:
Materials, applications, and the future, Mater. Today (Oxford, U. K.). 15 (2012)
16–25. doi:10.1016/s1369-7021(12)70017-2.
[40] J.M. LIU, Simple technique for measurements of pulsed Gaussian-beam spot
sizes, Opt. Lett. 7 (1982) 196–198. doi:10.1364/OL.7.000196.
[41] M. Yao, D. Su, W. Wang, X. Chen, Z. Shao, Fabrication of Air-Stable and
Conductive Silk Fibroin Gels, ACS Appl. Mater. Interfaces. 10 (2018) 38466–
38475. doi:10.1021/acsami.8b14521.
[42] K.A. Trabbic, P. Yager, Comparative Structural Characterization of Naturally-
and Synthetically-Spun Fibers of Bombyx mori Fibroin, 9297 (1998) 462–471.
doi:10.1021/ma9708860.
[43] J. Ayutsede, M. Gandhi, S. Sukigara, M. Micklus, H. Chen, F. Ko, Regeneration
of Bombyx mori silk by electrospinning . Part 3 : characterization of electrospun
nonwoven mat, 46 (2005) 1625–1634. doi:10.1016/j.polymer.2004.11.029.
[44] P. Monti, P. Taddei, G. Freddi, T. Asakura, M. Tsukada, Raman spectroscopic
characterization of Bombyx mori silk fibroin: Raman spectrum of Silk I, J.
Raman Spectrosc. 32 (2001) 103–107. doi:10.1002/jrs.675.
[45] M.A. Koperska, D. Pawcenis, J. Bagniuk, M.M. Zaitz, M. Missori, Degradation
markers of fi broin in silk through infrared spectroscopy, 105 (2014) 185–196.
doi:10.1016/j.polymdegradstab.2014.04.008.
[46] P. Monti, G. Freddi, A. Bertoluzza, N. Kasai, M. Tsukada, Raman Spectroscopic
Studies of Silk Fibroin from Bombyx Mon, J. Raman Spectrosc. 29 (1998) 297–
304. doi:10.1002/(SICI)1097-4555(199804)29:4<297::AID-JRS240>3.0.CO;2-
G.
[47] L. Beaulieu, T. Asakura, M. Pe, Study of Protein Conformation and Orientation
in Silkworm and Spider Silk Fibers Using Raman Microspectroscopy, (2004)
2247–2257. doi:10.1021/bm049717v.

18
[48] J. Shao, J. Zheng, J. Liu, C.M. Carr, Fourier transform Raman and Fourier
transform infrared spectroscopy studies of silk fibroin, J. Appl. Polym. Sci. 96
(2005) 1999–2004. doi:10.1002/app.21346.
[49] D.A. Field, C.A. Ventrice, S. Park, A. Velamakanni, R.D. Piner, S. Stankovich, I.
Jung, R.S. Ruoff, D. Yang, M. Stoller, G. Bozoklu, Chemical analysis of
graphene oxide films after heat and chemical treatments by X-ray photoelectron
and Micro-Raman spectroscopy, Carbon N. Y. 47 (2008) 145–152.
doi:10.1016/j.carbon.2008.09.045.
[50] K. Ilse, T. Tänzer, C. Hagendorf, M. Turek, Geometrical correction factors for
finite-size probe tips in microscopic four-point-probe resistivity measurements, J.
Appl. Phys. 116 (2014) 1–8. doi:10.1063/1.4903964.
[51] S. Papazoglou, Y.S. Raptis, S. Chatzandroulis, I. Zergioti, A study on the pulsed
laser printing of liquid-phase exfoliated graphene for organic electronics, Appl.
Phys. A Mater. Sci. Process. 117 (2014) 301–306. doi:10.1007/s00339-014-
8260-3.
[52] D.D.L. Chung, Review: Electrical application of carbon materials, 9 (2004)
2645–2661.
http://wings.buffalo.edu/academic/department/eng/mae/cmrl/Electrical
applications of carbon materials.pdf.

19

You might also like