Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Stimulated Brillouin scattering in dispersed graphene

Open Access Open Access

Abstract

We explored the Stimulated Brillouin scattering (SBS) behavior of a transparent liquid containing a low concentration of strongly absorbing nanoparticles. We measured SBS energies in N-methyl-2-pyrrolidone (NMP) and water at 532 nm-wavelength. The previously unknown NMP Brillouin gain factor is gB = 18.6 ± 1.8 cm⋅GW−1. Graphene nanoflakes suspended in liquids strongly quench SBS. Linear dependence of the SBS-threshold on the graphene absorption coefficient (concentration) makes it suitable for the detection of small nanoparticles quantities in water, with a minimal detectable concentration of 5⋅10−8 g⋅cm−3. The effect is interpreted as an antagonism between electrostriction and thermal expansion, which is induced by the formation of carbon vapor bubbles. It is very sensitive to changes in density, refractive index and acoustic absorption coefficient, which parameters can be determined in the SBS method, including access to the bubbling nanosecond dynamics. SBS suppression may find applications in laser technologies and optical telecommunication networks.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Two dimensional (2D) nanomaterials (graphene, black phosphorus, transition atom chalcogenides, etc.) are an advanced trend in the science of nanostructures due to their peculiar topology and electronic features triggering unusual electron and heat transport properties. They also open a new manifold of possibilities for the design of nanostructured devices [1,2]. They show outstanding nonlinear optical properties like enhanced two-photon absorption and absorption saturation [3,4] providing new perspectives for application in laser technologies, optical computing and telecommunications [5].

Ongoing wave mixing studies in 2D materials are mainly focused on harmonic generation [6–8]. Four-wave mixing in near infrared was recently carried on a graphene monolayer [9], and revealed a third-order nonlinear susceptibility χ(3) value of 2.1⋅10−15 m2V−2 which is ca. 7 orders of magnitude larger than in bulk insulators like silica and BK7 glass, 3-5 orders of magnitude larger than in bulk semiconductors like silicon, germanium, cadmium and zinc chalcogenides, metal oxides, and ca. 10 times larger than in thin plasmonic gold films and nanoparticles [10]. Most recently, an even larger value of χ(3) = 6.3⋅10−14 m2V−2 was obtained in graphene nanoribbons at mid-infrared frequencies close to the transverse plasmon resonance [11]. Despite these not yet abundant but impressive advances of phase conjugation in graphene, the effect of 2D materials on stimulated Brillouin scattering (SBS) remains unknown. Meanwhile, this effect is fundamentally important in laser and fiber telecommunications, and currently attracts theoretical considerations [12–14] concerning bulk and composite semiconductor materials, including practical designs [15].

This motivates the investigation of SBS in 2D nanostructures. We report here on the peculiar character of SBS in liquid graphene suspensions. We observe a net SBS effect in water and N-methyl-2-pyrrolidone (NMP), which is a good solvent for nanocarbon, and SBS threshold energy changes upon graphene suspension addition into the solvents. The results show a strong quenching of SBS even using vanishingly small concentrations of graphene nanosheets, corresponding to an absorption coefficient in the order of 0.001 cm−1 in the case of water. We attribute the observed effect to the interference of gratings formed in the liquid both by electrostriction and thermal expansion. By means of computer simulations of measured concentration dependences of SBS threshold energies we show that the thermal expansion is determined by carbon vapor bubble formation which strongly influences on the refractive index through density changes and acoustic wave damping through bubble compressibility. This allows the evaluation of effective bubble size in the suspensions under laser propagation conditions.

We also observed this effect in an aqueous suspension of single wall carbon nanotubes (SWCNT) at the dilution having absorption coefficient value of 0.016 cm−1 at the excitation wavelength and found it lower than in the graphene suspension.

2. Experimental procedure

2.1 Graphene nanoparticles morphology

Graphene suspensions were prepared from graphite powder as described earlier [16]. Sodium cholate (NaC) surfactant was used to stabilize the aqueous suspension. Details of suspension preparation and absorption cross-section, σe, measurement are available in Appendix A. We evaluated σe = (0.80 ± 0.16)⋅104 cm2g−1 of graphene in the aqueous suspension using the density value measured by gravimetric method. Olympus BX53 optical microscope, FEI Tecnai G2 F20 TEM and Bruker Nano Inc. Dimension 3100 AFM were used to visualize graphene. Graphene flakes with transverse sizes from 0.2 to 2 μm were observed in a dried drop of aqueous suspensions (Grwater) on a glass plate using both optical [Fig. 1(a)] and atomic force [Fig. 1(b)] microscopes. We could not obtain the same pictures of graphene from NMP (GrNMP), where separate flakes probably strongly aggregate during drying. For aqueous suspension, it turned out apparently possible because the surfactant prevents graphene aggregation by forming a sort of coat [high spikes in Fig. 1(b) at the flake edges]. The surfactant can also improve the visibility of graphene sheets, and this can be the alternative reason why we can see the flakes dried from water and do not see them from NMP. However, transmission electron microscopy images show that the graphene flakes in NMP also have several hundred nanometers average transverse size. A typical flake is shown in Fig. 1(c).

 figure: Fig. 1

Fig. 1 (a) Optical and (b) AFM images with transverse profiles A and B of Grwater; (c) TEM image of GrNMP; Raman G′ band spectra of (d) Grwater and (e) GrNMP: curves A and B refer to the excitation wavelengths 488 nm and 514 nm respectively, dashed arrows show the corresponding maximum positions of this excitation-dependent band.

Download Full Size | PDF

Flakes thickness measured by AFM varies from 0.3 nm to 1.5 nm approximately, evidencing the few layers morphology with middle thickness around 1 nm corresponding to 3-layer graphene whose thickness is estimated as a sum of two 0.34 nm interlayer distances and two 0.17 nm van der Waals radii. The suspensions were further studied by Raman spectrum analysis. Raman studies were performed at the Research park of St. Petersburg State University, center for Geo-Environmental Research and Modeling “Geomodel”, using a Horiba Jobin-Yvon LabRAM HR800 spectrometer equipped with a microscope Olympus BX-41. Spectral resolution was better than 3 cm−1. Raman spectra were excited by an Ar+ laser at 514 nm and 488 nm wavelengths with maximal power of 5 mW at the sample. A double Raman resonance G′ mode band of graphene dried from the aqueous suspension [see Fig. 1(d)] shows the spectral position and shape (a complex asymmetric band with maximum at the Raman shift of ca. 2688 cm−1, and spectral width of ca. 60 cm−1) which also correspond to the 3-layer graphene [17]. Graphene dried from NMP suspension [Fig. 1(e)] shows an intermediate shape of its G′ mode band between 2-layer and 3-layer graphene band shapes. Based on these data we suppose the preponderance of 3-layer graphene flakes in our suspensions.

2.2 Nonlinear optical setup

SBS observations were performed at room temperature using the second harmonic (λ = 532 nm) of 4 ns pulsed Continuum Minilite II Nd:YAG laser operating at 1Hz pulse repetition frequency. The setup is schematically shown in Fig. 2. Crossed polarizers P1 and P2 were used to change the laser pulse energy, together maintaining the laser beam polarization unchanged. A beam expander telescope (BE) was used to decrease the beam divergence angle 5 times from its initial divergence of θo = 1.8 mrad (measured by a ruler at the 4 m base).

 figure: Fig. 2

Fig. 2 Schematic setup: Continuum Minilite II – laser source; P1, P2, P3 – polarizers; BE – beam expander; BS – beam splitter; L1, L2 – lenses; F1, F2, F3 – neutral filters sets; D1, D2, D3 – detectors; S – sample cuvette; DP – diaphragm; dashed rectangle shows a light protection mask. Insets: A – X-axe beam intensity distribution, measured at L1 focal position (solid) and its Gaussian fit (dash); B – image of SBS from pure NMP on the screen put on DP; C – time profiles of incident beam (solid) and backward scattered (dash); D – image of the interaction area from the cuvette with pure solvent in the SBS observation conditions; E – the same image from the graphene suspension when no SBS is observed.

Download Full Size | PDF

The beam was focused in a 3 cm-path quartz cuvette by 150 mm biconvex lens L1. The beam spatial distribution after the lens being close to Gaussian in the focal point (see the inset A in Fig. 2), is influenced by diffraction at other Z positions. Z-scan study of the beam spot radius (W) performed in the air and approximations of its results by the test function W(Z) = W0(1 + (Z/ZR)2)1/2 gave W0 = 41 ± 7 μm waist radius and ZR = 1.0 ± 0.2 mm Rayleigh-length.

The backward scattered beam was reflected by the beam splitter BS and observed on the screen placed after a 300 μm pinhole PH. Inset B in Fig. 2 shows a visual picture of the SBS effect which has specific threshold behavior. The backward beam polarization followed the polarization of the incident radiation. Time profiles measurements performed with Thorlabs DET 025AL/M 2 GHz silicon detector manifest a pulse-time reduction from τp = 3.7 ± 0.3 ns (FWHM) in the incident beam to τp = 2.6 ± 0.2 ns in the backward beam (inset C in Fig. 2) that evidences its nonlinear origin.

Energies of incident, back-scattered and transmitted beams were measured by silicon detectors Thorlabs PDA100A-EC calibrated both by Coherent J50MB-YAG-1561 and by Thorlabs ES111C pyroelectric heads to account for possible systematic errors related to the energy meters. The difference in the calibrations was within 5%. The maximal SBS beam energy was found to be ca. 0.9 mJ in organic liquids and ca. 0.05 mJ in water at the ultimate achievable input energy 4.5 mJ. No SBS was detected in pristine (not diluted) suspensions of graphene nanosheets in water and NMP. Then, graphene suspensions were added into pure solvents by small portions, evaluating its linear absorption coefficient αe [exponent index in Lambert-Beer law: αe = α10⋅ln(10), where α10 is instrumentally obtained absorption coefficient] by the dilution ratio.

3. Results of nonlinear optical measurements

3.1 Stimulated Brillouin scattering

SBS beam energy ESBS was measured simultaneously with energy of the beam transmitted through the sample, E (optical limiting), vs. pump laser energy, E0, for different linear absorption coefficients. The dependences obtained are shown in Fig. 3. In the studied diapason of E0, they look linear, therefore SBS threshold energies, Ethr, can be numerically found as X-interсepts of linear fittings as shown in Figs. 3(a) and 3(b). The threshold determines the exponential SBS gain coefficient, G, according to the condition [18]:

G=gBEthrLπW02τp=25,
which corresponds to the regime where spontaneous scattering is amplified into SBS. Here gB is the Brillouin gain factor independent on the incident beam intensity, which characterizes the SBS strength in a material. L – the interaction length which in our experiment can be considered as 2ZRn, n – refractive index.

 figure: Fig. 3

Fig. 3 Incident energy dependences of: (a) back scattered energy from NMP-based liquids, (b) back scattered energy from water-based liquids, (c) transmitted energy through NMP-based liquids, and (d) transmitted energy through water-based liquids. Data correspond to (⬤) pure NMP, (⬛) pure water, (⬤) aqueous solution of NaC, and suspensions of different graphene concentration with following absorption coefficients αe: () 0.013 cm−1, (▼) 0.023 cm−1, (▲) 0.032 cm−1, (◆) 0.048 cm−1, (◆) 0.001 cm−1, (⬣) 0.002 cm−1, (⬟) 0.004 cm−1.

Download Full Size | PDF

Small additions of graphene suspensions in pure solvents remarkably decrease the SBS beam energy and increase its threshold. This is opposite to a reduction of the SBS threshold calculated for a graphene-clad tapered silica fiber [14]. In such structure graphene modulates the hypersonic wave which enhance SBS due to an appropriate ratio of photo-acoustic and electro-optic parameters of the core and the clad (acoustic velocity, Brillouin frequency and refractive index). In our case, we see a manifestation of another effect, taking place on the background of acousto-optical modulation of the bulk material by dispersed graphene nanoparticles and apparently originating from light absorption.

The obtained concentration dependence of the threshold energy reveals a rather linear character. Figures 4(a) and 4(b) show the dependences of Ethr on αe, which look quite linear both in NMP and in water. Experimental errors in Fig. 4 are within the point’s size except the last point where error is large because of very small SBS energy values. The SBS threshold in water was shifted up even with the NaC addition, however, our measurements performed with higher concentrations of NaC (from 1 to 5 wt.%) did not reveal further threshold increase. Since the NaC concentration used in our suspension is in proximity of critical micellar concentration [19]: 0.68 wt.%, we consider the micellation as the reason of this SBS threshold change through an additional extinction both of optical and acoustic waves. It was extracted from the experimental points [see hollow squares in Fig. 4(b)] and only the effect connected with graphene was further considered.

 figure: Fig. 4

Fig. 4 SBS threshold energy against absorption coefficient of graphene suspensions in (a) NMP and in (b) water; SBS gain factor against absorption coefficient of graphene suspensions in (c) NMP and in (d) water. Filled points: experimental results; empty points in (b): results corrected for the shift due to NaC (blue circle); lines in (a) and in (b): linear fitting; curves in (c) and in (d): simulations with simultaneous variations of ρ and ΓB.

Download Full Size | PDF

We see from the Fig. 4 that SBS threshold is very strongly affected by graphene quantities which can hardly even be detected spectrometrically (α10 < 0.01), and which can be considered as impurities. Thus, the SBS threshold measurement can be considered as a method for such impurity detection in pure liquids. The minimal detectable graphene concentration, ρGr min, can be evaluated through the uncertainty of SBS energy at lower concentrations, ΔEthr, which is biggest in case of water: ΔEthr = 0.11 mJ. The minimal detectable graphene absorption is αe min = ΔEthr/b = 4⋅10−4 cm−1, where b is the slope taken from the linear fitting, and ρGr min = αe min /σe = 5⋅10−8 g⋅cm−3. The value for NMP can be even smaller because αe min consists only 2⋅10−5 cm−1, however, since the absorption cross-section for the NMP-based suspension is not determined we cannot quote it here.

The Brillouin gain factor value obtained from the intercept of Ethr(αe) fitting line by means of Eq. (1), gB(0) = 6.4 ± 0.1 cm⋅GW−1 for water, is in accordance with a theoretical one 6.4 cm⋅GW−1, and however by 35% greater than an experimental one 4.8 cm⋅GW−1 [20]. Although we could not find any published value of gB for NMP, the plausible result for water convinces us in the correctness of our value gB(0) = 18.6 ± 1.8 cm⋅GW−1. It should be noted that the uncertainties indicated for gB(0) are obtained from the spread of Ethr experimental points and do not include systematic errors. An analysis of the latter shows that the error of beam waist ΔW0 dominates among them in Eq. (1) and consists in 34%.

Since the Ethr(αe) dependence is linear, the gB(αe) one evaluated using Eq. (1) shows hyperbolic character [points in Figs. 4(c) and 4(d)].

3.2 Optical limiting

Figures 3(c) and 3(d) show the optical limiting curves manifesting moderate nonlinear behavior at energies below SBS thresholds (< 1 mJ for NMP and < 2 mJ for water) which has been considered as effective two-photon absorption (TPA) and fitted by parabolic functions shown in the figures as solid curves:

E(E0)=Tlin(E0βeffLπW02τpE02),
where Tlin – linear transmittance, βeff – TPA coefficient, or effective nonlinear absorption (NLA) coefficient in general case. Results of this approximation are given in Table 1.

Tables Icon

Table 1. Effective NLA coefficients of liquid systems with different content (absorption coefficient) of graphene obtained from fitting the experimental points with Eq. (2).

We plotted the obtained coefficients against the linear absorption ones and fitted by straight lines as shown in Fig. 5. The line slope determines the nonlinear-to-linear absorption cross-sections ratio:

 figure: Fig. 5

Fig. 5 Effective NLA coefficients of suspensions against graphene linear absorption (⬛) in NMP and (⬤) in water.

Download Full Size | PDF

βeffαe=σeffσe.

The NLA cross-section for the suspension in NMP is found to be σeff = (5.59 ± 1.21) cm2GW−1σe, and in water: σeff = (42.0 ± 5.6) cm2GW−1σe. The data on TPA coefficients of pure solvents in visible range are scarce. For water at λ = 532 nm we have just an upper bound of it: βeff < 0.004 cm GW−1, obtained with picosecond pulses [21]. Our value is 20 times greater. Hence, it comprises other processes, enhancing the effect from the picosecond to nanosecond time scales. In view of the absence of linear absorption in visible range, we can exclude reverse saturable absorption effects related to long-lived states and consider the light scattering on increasing density fluctuations as the most realistic reason. These fluctuations follow the intensity fluctuations (speckles) due to the electrostriction forces even prior to the visible appearance of SBS [22]. At higher energies, around SBS thresholds the limiting curves in Figs. 3(c) and 3(d) show pits and subsequent linear behavior signaling additional reduction of the transmitted energy by SBS whose energy linearly depends on E0 within our experiment accuracy.

On the other hand, for a graphene suspension in NMP with αe = 0.223 cm–1 we know a value of βeff = 0.29 cmGW−1 obtained from Z-scan measurements at λ = 532 nm, τp = 4 ns [23]. That should give us βeff ~0.03 cm⋅GW−1 for our suspensions with ten times smaller absorption. Our βeff value for the graphene suspension in NMP: βeff(αe = 0.023 cm–1) – βeff(αe = 0) = 0.14 cm GW−1 is about 5 times greater. Therefore, we think that the linear concentration dependence in Eq. (3) may saturate at higher graphene nanoparticle concentrations due to aggregation effects (flakes size redistribution).

4. Theory of the effect and simulations

4.1 Conceptual issues

Propagation of the SBS beam intensity IS through a weakly absorbing substance is described by the equation [24]:

(ddzαe)IS=gBISI0,
where I0 is the incident beam intensity. A graphene addition can change the absorption coefficient of the substance in the left side of the equation. It decreases the Brillouin gain in Eq. (1) by the term –αeL, whose greatest value in our case is achieved in the NMP-based suspension: αeL = 0.048 cm−1⋅0.294 cm = 0.014 and consists only 0.056% of the threshold gain G = 25, which has no influence on the effect observed. The effective NLA coefficient decreases the Brillouin gain factor explicitly: ΔgB = –βeff. This quantity is proportional to αe as we see from Eq. (3), but the ratio βeff/αe values we obtained is much less than the derivatives of the curves in Figs. 4(c) and 4(d): dgB(0)/e = –gB(0)⋅b/Ethr(0) ≈–670 cm2GW−1 (NMP) and dgB(0)/e ≈–1500 cm2GW−1 (water). Therefore, no linear losses in Eq. (4) can explain the effect of SBS quenching by graphene.

Somehow analogous effect was observed and described by McEwan and Madden [25] in experiments on degenerated four-wave mixing in carbon black suspensions. They showed that bubbles caused by local heating of carbon nanoparticles upon laser radiation, form a transient grating in the substance correlated to interference peaks of two incident laser beams. The fourth beam appears as a result of diffraction of the third beam on the grating. Graphene nanoparticles being good absorbers should form a similar grating from their thermally expanded microenvironment. The difference in our case is that this grating corresponds to the maxima of the incident beam intensity.

The sketch in Fig. 6 shows the processes taking place in the graphene suspensions under the conditions of SBS generation. First, electrostriction forces (red arrows) compress the matter and form the density peak (hence, the refractive index peak) grating which is responsible for the backward scattered wave amplification (the conventional SBS generation mechanism). At the same time, graphene nanoparticles form a “negative” grating of refractive index dips due to thermal expansion which is coherent with the first grating and destroys it.

 figure: Fig. 6

Fig. 6 Sketch of the electrostriction – thermal expansion antagonism occurring in a graphene suspension upon irradiation and giving impact on SBS quenching.

Download Full Size | PDF

This background requires a consideration of the thermal expansion process that leads to an additive to the Brillouin gain factor of the same form as for the thermal-induced scattering but opposite in sign [24]:

gB(αe)=gBegBT(αe),
where (with electric constant ε0 for S.I. units):
gBe=(γeω0)2ε02c3nρvacΓB
is the electrostrictive part of the gain factor, and
gBT(αe)=γeγT(αe)ω02ε02c3nρvacΓB
is the thermo-expansive part.

Notations are: γe = ε0ρ (∂ε/∂ρ)S – electrostrictive coefficient; γT(αe) = ε0c2βT vacαe/(Cpω0) – thermo-optic coupling coefficient; βT = (∂V/∂T)ρ/V – thermal expansion coefficient; Cp – specific heat capacity at constant pressure, ρ – solvent density; S – entropy; n – refractive index.

ΓB=(ω0ωB)2ηvac2ρ
is Brillouin line width (acoustic wave dumping coefficient); ω0, ωB, – incident laser and Brillouin waves circular frequencies correspondingly; vac – acoustic wave velocity; η = ηs(4/3 + ηb/ηs) – total viscosity of the solvent: ηs – shear viscosity, ηb – bulk viscosity.

4.2 Local heating of graphene flakes

A local temperature rise, ΔT, of absorbing graphene flakes drives the process, and the understanding of its scale is important. It can be upper estimated using the simple heat capacity definition relation: ΔT = ΔQabs/(mGrCp), where ΔQabs is the heat amount obtained by a graphene flake from the radiation absorbed, mGr is the flake mass, Cp = 0.7 J⋅K−1g−1 [26] for graphene.

Following the Lambert-Beer law at low absorption and neglecting the scattering: ΔQabs = E0F(t)σemGrL/V, where V is the irradiated suspension volume, F(t) = ∫0tI0(τ)/E0 is the time dependent radiation dose, which can be easily calculated from the laser beam time profile (see Fig. 2, inset C). Considering the irradiated volume is that inside the caustic surface: V = 8πW02ZR/3 for the Gaussian beam, we finally obtain that the temperature rise is not determined by the flake mass or size, but only by the absorption cross-section:

ΔT(t)=3E0F(t)σe4πW02Cp.
The calculated temporal dependence at E0 = 1 mJ which approximately corresponds to the minimal SBS threshold is shown in Fig. 7 by solid line.

 figure: Fig. 7

Fig. 7 Temperature rise of a graphene flake upon 1 mJ laser-pulse absorption (solid curve); laser beam temporal profile (dashed curve) in arbitrary units.

Download Full Size | PDF

For the first τ1 = 1.34 ns a temperature rise of 4830 K is achieved which approximately corresponds to the graphite sublimation condition reported by McEwan and Madden [25]. The subsequent sublimation process can be traced using the relation:

F(τ2)=F(τ1)+4πW02Λ3E0σe,
where Λ is heat of vaporization. From here the sublimation ending time τ2 = 4.65 ns was determined. It should be noted that the presented values are upper estimations. We used parameters known for graphite including the heat of vaporization value Λ = 169.7 kcal/mole from JANAF thermochemical Tables [27]. However, it is arguable that thin graphene flakes can be evaporated more easily. For example, a study of graphite filaments reported in [27] reveals their effective evaporation at remarkably lower temperatures (3200-3500 K) and with a lower heat of evaporation value: 102.7 kcal/mole. This tendency would reduce the evaporation time in our case, which would lead to an even stronger effect on the laser pulse propagation.

Carbon atom and its ions have no remarkable absorption lines at λ = 532 nm and 266 nm, therefore a further temperature growth after the sublimation can be determined either by carbyne forms in the carbon vapor or by plasma electrons. The plasma is apparently getting generated in the bubbles due to a partial oxidation of the carbon atoms. These processes are hardly evaluable and seem unimportant in the framework of the present study.

It is important to understand that at a laser pulse energy around the SBS threshold all dispersed graphite nanoparticles become evaporated after 4.65 ns. The characteristic time of SBS is the inverse acoustic wave dumping coefficient (2πB), and for most of the studied liquids it is equal to 3-4 ns, which means that the graphite evaporation begins to interfere with the SBS process. For greater pulse energies this time scale is shorter. Hence, in the presence of graphene flakes, SBS is developing in a medium which represents an emulsion of carbon vapor bubbles and has non-steady temperature fields. Thermodynamic characteristics of such medium can significantly differ from those of pure liquid. The increasing of graphene concentration will also increase the bubble concentration and the material characteristics will change drastically.

4.3 Simulation of gBe) dependence

Only thermo-expansive part of the Brillouin gain factor [see Eq. (7)] shows an explicit dependence on αe, which, however, does describe neither the derivative dgB(αe)/e observed, nor the hyperbolic form of the gB(αe) dependence. Thus, we must assume an influence of graphene flakes on SBS parameters in Eqs. (6) and (7).

We performed simulations of gB(αe) using one of these parameters, P, changed by small (less than 20%) variations δP:

P˜=P(1±δP)
at fixed values of all other parameters, and referring it to the observed scale of the absorption coefficient change Δαe:
δP=ΔαPΔαe,ΔαP=1PdPdαe.
Here P designates a varied parameter (P = n, ρ, vac, ΓB, and γT), ΔαP is an absorptive increment (or decrement) of the parameter. Having varied these parameters one by one, a comparative examination of the efficiency of each parameter on the Brillouin gain factor reduction was performed. This approach implies the priority of small variations of the substance characteristics in the observed changes of the net effect, and its reliability is based on numerous observations of the relatively small changes in refractive index [4,28], acoustic velocity [29,30], thermal expansion coefficient [31] and frequency shifts ω0ωB [29], occurring at high power laser action on nanocarbon species in suspensions, as well as on the moderate character of known temperature and pressure dependences of density, speed of sound, viscosity, refractive index, and thermal expansion coefficient of pure NMP [32–34] and water [32,35].

Some of the parameters depend on others, such as refractive index and acoustic velocity on density. In our simulations we considered the former using the Lorentz-Lorenz equation:

n=(1+2Rρ1Rρ)1/2,
where R is the mass refraction (see Appendix B for definition). Therefore, n was varied both by the direct variation δn and by the density variation δρ through Eq. (11). The dependences of other parameters on density including vac are not so explicit in view of an inhomogeneity of the substance and a thermodynamic unsteadiness of the process, so that they were varied only independently. The initial (unvaried) values of the parameters used in our simulation are collected in Table 2.

Tables Icon

Table 2. Physical properties of the solvents at T = 293 K and standard pressure.

Electrostrictive coefficients were also evaluated using Eq. (11), which gives a well-known relation: γe = ε0(n2–1)(n2 + 2)/3. The Brillouin line widths ΓB (αe = 0) were taking those to fit the observed values of gB(0) in calculating by Eq. (5) at αe = 0: ΓB = 2.545 GHz (NMP) and ΓB = 3.475 GHz (water). The heat capacity variation δCp in γT(αe) functionally gives the same behavior as δβT with the opposite sign, so that the latter can be considered as a variation of the thermo-optic coupling coefficient in total: δβT = δγT.

The simulated gB(αe) dependences due to different parameter variations are superimposed onto the experimental points in Figs. 8(a) and 8(b). The efficiency of each parameter towards the relative change of gB is illustrated by dependences in Figs. 8(c) and 8(d): more rapid curve at the (1,1) point indicates more influenced parameter. The efficiency curves make it clear that the thermal expansion gives a very small contribution to the Brillouin gain factor, which in case of water is only ca. 10−3 of gBe values in all the small variations range. In case of NMP, gBT still has nonzero impact, which has been considered mostly in simulations at other parameter variations. The simulated curves of gB(αe) due to δγT have too strong curvature [curve E in Fig. 8(a) and curve С in Fig. 8(b)] and cannot describe the effect observed.

 figure: Fig. 8

Fig. 8 SBS gain factor of (a) NMP and (b) water against linear absorption of graphene suspensions; points: experimental results; curves: simulations; varied parameters are: (a) n (curve A), ρ (curve B), vac (curve C), ΓB (curve D), γT (curve E), (b) n and ρ (curve A), vac and ΓB (curve B), γT (curve C). Relative changed SBS gain factor of (c) NMP and (d) water against corresponding varied parameters: (c) n (curve A), ρ (curve B), vac (curve C), ΓB (curve D), γT (curve E), (d) n (curve A), ρ (curve B), vac and ΓB (curve C), γT (curve D).

Download Full Size | PDF

We see that all the simulated curves in Figs. 8(a) and 8(b) do not follow the experimental points sufficiently well. However, the proximity of each curve to them indicates reliability of the parameter influence. Since the dependence is more nonlinear for NMP, Figs. 8(a) and 8(c) are more indicative. The changes of vac and ΓB, both positive (increase), give similar contributions to gB(αe), even equal for water where the gBT part independent on vac is negligible. But for NMP we can see the difference between them: the simulation with δvac [curve C in Fig. 8(a)] does not reproduce the curvature of the experimental points. On the other hand, the positive variation δΓB gives the best approaches to them. This makes ΓB more important parameter in the SBS quenching as compared to vac.

The negative variations of ρ and n (their decrease) give even stronger influence on the effect observed and reproduce the true sign of curvature [curves A, B in Fig. 8(a) and curve A in Fig. 8(b)]. The absorptive increment (decrements) providing the best fit of the experimental points are given in Table 3. The sensitivity of gB(αe) curves to these values defines their uncertainties which are within the last valid digits.

Tables Icon

Table 3. Relative absorptive changes (in cm) of refractive index, density, acoustic velocity, Brillouin line width and thermal optic coefficient providing the observed decrease of gB.

Lower ΔαP value shows stronger influence of the corresponding parameter. Despite n is the most influenced parameter, it is not independent, and its variation can be contributed from both density and mass refraction changes. An analysis of the mass refraction contribution given in Appendix II demonstrates the prevailing of density-determined nature of the refractive index changing. It can be calculated that values Δαρ = –11.1 cm (NMP) and –32 cm (water) is obtained when n(ρ) dependence is taken into account.

Therefore, the simulation results denoted a density decrease and an increase of Brillouin line width as the key changes leading to the SBS quenching, whereas the effect is very low sensitive to the thermo-optic coupling coefficient. Analyzing parameters that determine ΔαΓB in Eq. (8), we can see that its growth is unlikely determined by the Brillouin frequency shift (ω0ωB), whose change should be rather negative upon irradiation [29].

Thermodynamic studies show a remarkable and monotonous decrease of viscosity (both shear and bulk) of liquids [36] and steam [35] in a large range of temperature and pressure growth. Hence, if the Eq. (8) remains adequate for the inhomogeneous media, the growth of ΓB should be attributed to a drop of acoustic velocity. Such kind of decrease is anticipated in a bubbly liquid due to a high compressibility of bubbles [30]. The effect can be considered as an acoustic wave extinction in growing bubbles that increase the acoustic absorption coefficient αac = ΓB/vac of the liquid, which is most likely the main parameter in the SBS quenching along with density.

Apparently, both the density and acoustic absorption factors take place in the real process acting simultaneously. For this reason, we performed a fitting of the experimental gB(αe) points by a simulation with simultaneous variation of ρ and ΓB values. The curves obtained are shown in Figs. 4(c) and 4(d) and manifest a perfect coincidence with the measured dependence with correlation coefficients 0.992 (NMP) and 0.991 (water). Corresponding parameter changes are: Δαρ = –1.55 cm (NMP), –5.4 cm (water) and ΔαΓ = 50 cm (NMP), 90 cm (water).

4.4 Bubble size scaling

Since the density reduction is determined by the bubble formation: δρ = ρ (1 – fC), where fC is the bubble filling factor, the density decrement allows to estimate the average bubble radius:

rb=(9d2σeρsΔαρ4π)1/3
as it is shown in Appendix B. Here d denotes the average flake characteristic size, and ρs = 7.7⋅10−8 g⋅cm−2 [37] is the surface density of graphene. Estimating d = 600 nm from the AFM image scale, we can get the bubble radius: rb = 2.0 ± 0.5 μm. An error analysis is given in Appendix B. Since the SBS effect develops during few nanoseconds (due to the pulse shortening SBS intensity is negligible already after fourth nanosecond of the incident pulse) the radius obtained here corresponds to the starting point of bubble evolution: just after the termination of graphene sublimation. Thus, if to tune laser pulse duration and its energy, one can trace the bubble size evolution using the SBS method.

The bubble radius found is comparable with those obtained in SWCNT suspension in binary water-PEG solvent under nanosecond laser irradiation: 1.2 μm [30]. Therewith, the referenced study confirms a weak dependence of the initial bubble size on the laser pulse energy. Our one-point measurement of the SBS quenching in SWCNT suspension: ΔgBαe = –94 cm2GW−1 is less than our result for graphene that will also give a lower value for the bubble size in case of nanotubes.

Simple evaluations [see Eqs. (14) and (15) in Appendix B] using the same density decrement value give the ratio of bubble-to-flake volume Vb/Vf = 1⋅105 without data of flake surface size, but with the flake thickness. In our case it does not improve the uncertainty, but at some conditions the thickness can be made more uniform and precisely measured by AFM or Raman spectroscopy than the flake area. In this case the volume ratio can be more indicative.

In both cases the absorption cross-section of nanomaterial should be known. For graphene flakes numerous experiments give values which are several times different depending on the suspension preparation conditions [38–41]. Although similar σe values were demonstrated for different organic solvents [38], the value found in aqueous suspensions was twice less than that in NMP-based suspensions after the same treatment procedure [38,39], presumably, due to the surfactant influence. If, following this reason, we perform similar evaluations for NMP in assumption of the graphene absorption cross-section being twice lager than in water, we obtain values: rb = 1.7 μm and Vb/Vf = 8⋅104, which are comparable with the values obtained for aqueous suspension.

Despite the demonstrated possibility of the SBS method in obtaining the bubble size, the simultaneous fitting performed reveals a wide minimum which is crucial especially for Δαρ due to its small value and can increase the uncertainties of the characteristics found. Independent measurements of the acoustic absorption by photo-acoustic methods or Brillouin spectroscopy must be done to obtain the full image of the mechanism of light-matter interaction in conditions of the SBS type phase conjugation in liquid suspensions of strongly absorbing nanoparticles. Once the acoustic absorption coefficient is fixed, the dimension of vapor bubbles can be determined more precisely.

5. Conclusion

In summary, we have studied the stimulated Brillouin scattering in NMP and water in the presence of nanoscale graphene flakes and have measured its energetic characteristics. We found a strong SBS quenching effect in liquids by graphene with low concentrations of the nanoparticles that give no remarkable absorption in the material. Established linear dependences of SBS threshold on graphene absorption coefficient (i.e., concentration) can be used for the detection of small nanomaterial quantities in liquid media down to 5⋅10−8 g⋅cm−3 which was shown for the aqueous suspension.

Computer simulations of the Brillouin gain factor show the efficiency of different thermodynamic, electrooptic and photoacoustic parameters in the SBS quenching. The role of density and compressibility among them which change due to carbon vapor bubble formation is found to be decisive. Effective bubble size has been estimated, which can be specified in combination with acoustic wave absorption measurements providing more information on bubble properties in the nanosecond time scale. From this point of view, the SBS method can be used as a tool for the nanosecond-resolved bubble formation scaling in nanoparticle contained media.

The studied effect can be used for SBS suppression in situations where it is undesirable (both in laser technology and optical telecommunication networks). We expect an appearance of this effect in transparent solids containing traces of nanoparticles and believe that the method can be applied to other kinds of 2D-nanomaterials and maybe other nano-sized linear or nonlinear absorbers.

Appendix A Suspension, preparation, and absorption cross-section measurements

Graphene suspensions were prepared from graphite powder (Sigma-Aldrich 332461) by the way like that described earlier [16]. NMP (99.5%, Sigma-Aldrich 328634) and double deionized water were used as solvents. Graphite powders were added to the solvent with initial concentration of 5 mg/ml. Sodium cholate (NaC) hydrate > 99% (Sigma-Aldrich C6445) surfactant was added to the aqueous sample with initial concentration 0.03 mg/ml. Mixtures were then sonicated using Sonics Vibra-cell VCX-750 W ultrasonic processor at the amplitude 38% for 1 h. Then the suspensions were centrifugated in Rotofix 32A centrifuge at 4000 rpm for 1 h twice with 15 h interval. After that supernatant portions were taken, and their absorption spectra were measured.

The precipitate was dried in an evacuated drying box at ca. 1 Torr pressure and T = 90°C with a periodical control of weight using Mettler Toledo XS105 analytic balance. The drying ended when the weight stopped changing. The total weight of the dry precipitate was subtracted from the initial weight to evaluate the carbon concentration ρC in the suspension. For aqueous suspension we obtained ρC = 56.3 ± 11.2 μg cm−3. The uncertainty is mainly determined by the weight error ± 0.5 mg appearing from poorly controllable water vapor adsorption by dry carbon powder. No weight loss was found in carbon precipitate from NMP comparing to the initial carbon powder weight in several attempts of suspension preparation and drying during weeks. This apparently means that the procedure doesn’t allow to eject NMP molecules adsorbed by carbon. After spectral measurements 0.47 mg cm−3 of NaC was added to the aqueous suspension for better stability. The total concentration 0.5 wt.% of NaC in the suspension is close to its critical micelle concentration [19] so that we can expect that graphene flakes are totally surrounded by the surfactant. Both suspensions (NMP- and water-based) were stable, and no sedimentation was seen during more than 6 months.

Absorption spectra of the studied suspensions were measured in 1 cm path cell using PerkinElmer Lambda 750 dual-beam UV/VIS/NIR spectrometer, just after taking supernatant. The spectra were measured with reference to the solvent placed in the reference beam. The results are shown in Fig. 9.

 figure: Fig. 9

Fig. 9 Absorption spectra of the as-prepared graphene suspensions in water and NMP.

Download Full Size | PDF

The NMP suspension was approximately 5 times denser; its spectrum in the Fig. 9 is divided by 10 to compare with that of aqueous suspensions. Two main factors determine the absorption spectra shape. First, it is correlated with graphene density of states DOS(λ), which in case of electron-hole symmetry is determined by the complete elliptic integral of first kind [1]:

DOS(λ)=8π2hcλh2λZ01/20π/2dφZ0Z1sin2(φ),Z0=(1+λhλ)2+14[(λhλ)21]2,Z1=4λhλ.

Here λh is the peak position of DOS which corresponds to twofold nearest-neighbor hopping energy in graphene (Eh ≈ 2.8 eV), h is the Plank constant, c – the speed of light in vacuum. The calculated DOS(λ) function using Eq. (13) is shown in the Fig. 9 by a dotted curve with maximum at λh = 221.4 nm. The other dotted curve with maximum at λh = 270 nm is a Lorentz function which is a good approximation for a surface plasmon absorption cross-section [42]. The transverse plasmon resonance frequency in graphene takes place at ~5 eV [43].

It can be seen in the Fig. 9 on the example of aqueous suspension spectrum, that each of these dependences do not describe it separately. The Lorentzian can describe only its red part for wavelengths λ > 500 nm with a constant adduct (red dashed curve). Both DOS and Lorentzian together reproduce the experimental curve down to λ ≈ 300 nm (magenta dash-dotted curve). The constant adduct is also necessary to fit the spectrum, and it is apparently connected with the linear Dirac electronic spectrum of graphene, which leads to the constant absorption coefficient equal only απ per layer, where α = 1/137 is the fine structure constant, in all NIR and visible spectral region down to 450 nm [43].

The peak shape in Fig. 9 may be strongly influenced by light scattering which is small in visible for the nanoparticles concentrations studied, but becomes remarkable in UV not only due to diffraction enhancement, but also due to dramatic changes in dynamic dielectric permittivity through the factor [ε(ω) – εNaC(ω)].

These considerations assume a complicated character of the spectrum determined by different electron movements in graphene. In a real graphene flake electron-hole symmetry can be broken due to defects and a confinement effect that leads to a shift of the DOS peak position and distortion of its shape. The real spectrum of graphene suspension is even more complicated being a sum of different particles absorption, where the size and layer thickness distribution as well as defect structures are crucial.

Thus, the total absorption coefficient of graphene in water measured at λ = 532 nm is α10 = 0.1954 ± 0.0011 cm−1 that taking into account ρC value obtained gives the absorption cross-section σe = α10 ⋅ ln(10) /ρC = (0.80 ± 0.16)⋅104 cm2g−1. It is a little less than the value 1.39⋅104 cm2g−1 obtained for low-concentrated aqueous graphene suspensions [39], which we explain by difference in layer thickness and defect structures of our graphene flakes compared to those studied in the reference.

Appendix B Carbon filling factors and mass refraction

Fast carbon evaporation during 4 ns pulse leads to an increase of total volume of bubbles Vb. The bubble filling factor, which is the ratio of its volume to the caustic volume can be then evaluated: fC = Vb/V = rb3Nb/(2ZRW02), where rb is the bubble radius and Nb is the number of bubbles in the caustic region. The latter coincides with the number of graphite flakes in the same volume. We can calculate it from carbon density ρC, V, and the mass of one particle mGr, which can be obtained only by rough estimations. Assuming three-layer thickness in average, we can accept: mGr = 3d2ρs, where ρs = 7.7⋅10−8 g⋅cm−2 is the surface density of graphene [37]. Then, considering that αe = ρCσe, we obtain:

Nb=8πW02ZRαe9d2σeρs,
which gives the value Nb = 1.5⋅106αe(cm−1) with the assumption of the average characteristic flake size based on the AFM image scale to be around d = 600 nm. Its uncertainty comes mainly from the size distribution standard deviation SD(d), and apparently consists a few tens of percent. Thus, for the bubble filling factor we have:
fC=4πrb3αe9d2σeρs,
or fC = 5⋅1011⋅[rb(cm)]3αe(cm−1) in water with the said uncertainty.

At that, the initial graphene flakes filling factor f0 can be estimated more precisely because it does not require the particle size, but only its thickness h:

f0=hαe3σeρs.

Assuming the middle thickness of graphene flakes in our suspension h = 1 nm, we obtain f0 = 5.4⋅10−5αe(cm−1) for water. At the average number of layers equal to three, its standard deviation cannot be more than one that implies the 33% upper estimation for the Δh uncertainty. With the 20% uncertainty of density ρC it gives Δf0/f0 = 39%.

From the definition of filling factors also follows that their ratio is just a ratio of the average bubble volume to the average flake volume:

fCf0=VbVf.

Mass refraction is a ratio of the Lorentz-Lorenz function to the mass density:

Rn21n2+21ρ,
and it is a measure of polarizability a0 of a particle with the mass M0:

R=4π3a0M0.

In application to the substances under study, R was determined from Eq. (16) for NMP: R = 0.271 cm3g−1 and for water: R = 0.206 cm3g−1 using data presented in Table 2. For the graphene mass refraction, a refractive index value n3layer = 2.27 for the three-layered graphene [44], and a simple estimation of graphene density: ρGr = 3ρs/h = 2.31 g cm−3 were used: RGr = 0.251 cm3g−1.

Refractive index of the suspension is not practically affected by the mass refraction of graphene in view of its very low density:

nsusp=[1+2(Rsolvρsolv+RGrρC)1(Rsolvρsolv+RGrρC)]1/2,
nsusp = 1.472 (suspension in NMP) and 1.333 (suspension in water) at maximal concentrations used.

The mass refraction of carbon in bubbles is evaluated from Eq. (17) and literature data on carbon atom polarizability [45]: RC = 0.374 cm3g−1 and found even greater than all other refractions. Thus, carbon vaporization could increase the refractive index in changing the term RGrρC, which is negligible, by RCρC/fC = RC/(5⋅1011rb3σe) which can be larger than solvent’s ones at small bubbles. Since we know from the experiment and related computer simulations that the refractive index of the suspension heated by laser pulse is decreasing, we can establish the conditions on the bubble radius value: rb >> 10 nm. In that case RCρC/fC << 0.1, and the decreasing of the refraction is assured only by the solvent density reduction due to bubbles formation:

δ(Rρ)=Rsolvρsolv(1fC).

From here we can see that fC = Δαρ⋅Δαe, where Δαρ the absorptive decrement of density introduced by Eq. (10). The substitution of Eq. (14) into Eq. (15) and then into Eq. (18) gives Eq. (12) for the bubble size.

The expression is exact, i.e., obtained from the geometric consideration with the only assumption of the square shape of the flake. In case of other shape types d2 should be replaced by the corresponding area value. Therefore, the bubble radius depends on the flake surface area, absorption cross-section and absorptive decrement of density.

Roughly supposing the relative standard deviation of the flake size distribution to be SD(d)/d = 30%, the fractional error δdrb due to flake size uncertainty is 23%. The 20% uncertainty of the absorption cross-section gives the fractional error δσrb = 7%. The error in Δαρ obtained from the experimental data fitting is not large and consists only few percent, however, other minima in the fitting related to other ΔαΓ values are possible. In total, we evaluate δrb error by 25%. Therefore, simultaneous measurement of acoustic absorption will help to obtain more reliable data of average bubble size.

Funding

National Natural Science Foundation of China (NSFC) (61675217, 61522510); the Strategic Priority Research Program of Chinese Academy of Sciences (CAS) (XDB16030700); the Key Research Program of Frontier Science of CAS (QYZDB-SSW-JSC041); the Program of Shanghai Academic Research Leader (17XD1403900); President’s International Fellowship Initiative of CAS (2017VTA0010, 2017VTB0006, 2018VTB0007).

References

1. A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim, “The electronic properties of graphene,” Rev. Mod. Phys. 81(1), 109–162 (2009). [CrossRef]  

2. Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, “Electronics and optoelectronics of two-dimensional transition metal dichalcogenides,” Nat. Nanotechnol. 7(11), 699–712 (2012). [CrossRef]   [PubMed]  

3. Y. Li, N. Dong, S. Zhang, X. Zhang, Y. Feng, K. Wang, L. Zhang, and J. Wang, “Giant two-photon absorption in monolayer MoS2,” Laser Photonics Rev. 9(4), 427–434 (2015). [CrossRef]  

4. Y. Chen, T. Bai, N. Dong, F. Fan, S. Zhang, X. Zhuang, J. Sun, B. Zhang, X. Zhang, J. Wang, and W. J. Blau, “Graphene and its derivatives for laser protection,” Prog. Mater. Sci. 84, 118–157 (2016). [CrossRef]  

5. X. Liu, Q. Guo, and J. Qiu, “Emerging low-dimensional materials for nonlinear optics and ultrafast photonics,” Adv. Mater. 29(14), 1605886 (2017). [CrossRef]   [PubMed]  

6. L. M. Malard, T. V. Alencar, A. P. M. Barboza, K. F. Mak, and A. M. de Paula, “Observation of intense second harmonic generation from MoS2 atomic crystals,” Phys. Rev. B 87(20), 201401 (2013) [CrossRef]   [PubMed]  

7. N. Yoshikawa, T. Tamaya, and K. Tanaka, “High-harmonic generation in graphene enhanced by elliptically polarized light excitation,” Science 356(6339), 736–738 (2017). [CrossRef]   [PubMed]  

8. H. G. Rosa, Y. W. Ho, I. Verzhbitskiy, M. J. F. L. Rodrigues, T. Taniguchi, K. Watanabe, G. Eda, V. M. Pereira, and J. C. V. Gomes, “Characterization of the second- and third-harmonic optical susceptibilities of atomically thin tungsten diselenide,” Sci. Rep. 8(1), 10035 (2018). [CrossRef]   [PubMed]  

9. E. Hendry, P. J. Hale, J. Moger, A. K. Savchenko, and S. A. Mikhailov, “Coherent nonlinear optical response of graphene,” Phys. Rev. Lett. 105(9), 097401 (2010). [CrossRef]   [PubMed]  

10. Y. Wang, C.-Y. Lin, A. Nikolaenko, V. Raghunathan, and E. O. Potma, “Four-wave mixing microscopy of nanostructures,” Adv. Opt. Photonics 3(1), 1–52 (2011). [CrossRef]  

11. D. Kundys, B. Van Duppen, O. P. Marshall, F. Rodriguez, I. Torre, A. Tomadin, M. Polini, and A. N. Grigorenko, “Nonlinear light mixing by graphene plasmons,” Nano Lett. 18(1), 282–287 (2018). [CrossRef]   [PubMed]  

12. C. Wolff, P. Gutsche, M. J. Steel, B. J. Eggleton, and C. G. Poulton, “Impact of nonlinear loss on stimulated Brillouin scattering,” J. Opt. Soc. Am. B 32(9), 1968–1978 (2015). [CrossRef]  

13. M. J. A. Smith, B. T. Kuhlmey, C. M. de Sterke, C. Wolff, M. Lapine, and C. G. Poulton, “Metamaterial control of stimulated Brillouin scattering,” Opt. Lett. 41(10), 2338–2341 (2016). [CrossRef]   [PubMed]  

14. H. J. Lee, F. Abdullah, and A. Ismail, “Numerical modelling on stimulated Brillouin scattering characterization for graphene-clad tapered silica fiber,” EPJ Web Conf. 162, 01029 (2017). [CrossRef]  

15. B. Morrison, A. Casas-Bedoya, G. Ren, K. Vu, Y. Liu, A. Zarifi, T. G. Nguyen, D.-Y. Choi, D. Marpaung, S. J. Madden, A. Mitchell, and B. J. Eggleton, “Compact Brillouin devices through hybrid integration on silicon,” Optica 4(8), 847–854 (2017). [CrossRef]  

16. Y. Feng, N. Dong, Y. Li, X. Zhang, C. Chang, S. Zhang, and J. Wang, “Host matrix effect on the near infrared saturation performance of graphene absorbers,” Opt. Mater. Express 5(4), 802–808 (2015). [CrossRef]  

17. L. M. Malard, M. A. Pimenta, G. Dresselhaus, and M. S. Dresselhaus, “Raman spectroscopy in graphene,” Phys. Rep. 473(5–6), 51–87 (2009). [CrossRef]  

18. R. W. Boyd, K. Rzaewski, and P. Narum, “Noise initiation of stimulated Brillouin scattering,” Phys. Rev. A 42(9), 5514–5521 (1990). [CrossRef]   [PubMed]  

19. G. González-Gaitano, A. Compostizo, L. Sanchez-Martin, and G. Tardajos, “Speed of sound, density, and molecular modeling studies on the inclusion complex between sodium cholate and β-cyclodextrin,” Langmuir 13(8), 2235–2241 (1997). [CrossRef]  

20. M. Denariez and G. Bret, “Investigation of Rayleigh wings and Brillouin-stimulated scattering in liquids,” Phys. Rev. 171(1), 160–171 (1968). [CrossRef]  

21. B. A. Rockwell, C. P. Cain, G. D. Noojin, W. P. Roach, M. E. Rogers, M. W. Mayo, and C. A. Toth, “Nonlinear refraction in vitreous humor,” Opt. Lett. 18(21), 1792–1794 (1993). [CrossRef]   [PubMed]  

22. B. Ya. Zel’dovich, N. F. Pilipetsky, and V. V. Shkunov, Principles of phase conjugation (Springer-Verlag Berlin Heidelberg, 1985), Chap. 3.

23. N. Dong, Y. Li, Y. Feng, S. Zhang, X. Zhang, C. Chang, J. Fan, L. Zhang, and J. Wang, “Optical limiting and theoretical modeling of layered transition metal dichalcogenide nanosheets,” Sci. Rep. 5(1), 14646 (2015). [CrossRef]   [PubMed]  

24. D. Pohl and W. Kaiser, “Time-resolved investigations of stimulated Brillouin scattering in transparent and absorbing media: determination of phonon lifetimes,” Phys. Rev. B 1(1), 31–43 (1970). [CrossRef]  

25. K. J. McEwan and P. A. Madden, “Transient grating effects in absorbing colloidal suspensions,” J. Chem. Phys. 97(11), 8748–8759 (1992). [CrossRef]  

26. F. Ma, H. B. Zheng, Y. J. Sun, D. Yang, K. W. Xu, and P. K. Chu, “Strain effect on lattice vibration, heat capacity, and thermal conductivity of graphene,” Appl. Phys. Lett. 101(11), 111904 (2012). [CrossRef]  

27. J. T. Clarke and B. R. Fox, “Rate and heat of vaporization of graphite above 3000 °K,” J. Chem. Phys. 51(8), 3231–3240 (1969). [CrossRef]  

28. V. S. Nair, A. Pusala, M. Hatamimoslehabadi, and C. S. Yelleswarapu, “Impact of carbon nanotube geometrical volume on nonlinear absorption and scattering properties,” Opt. Mater. 73, 306–311 (2017). [CrossRef]  

29. W. Zhou, R. P. Tiwari, R. Annamalai, R. Sooryakumar, V. Subramaniam, and D. Stroud, “Sound propagation in light-modulated carbon nanosponge suspensions,” Phys. Rev. B Condens. Matter Mater. Phys. 79(10), 104204 (2009). [CrossRef]  

30. I.-Y. S. Lee, Y. Hayama, H. Suzuki, and T. Osawa, “Photoacoustic sensitization and laser-induced cavitation in polymer solutions by carbon nanotubes,” J. Phys. Chem. C 114(51), 22392–22397 (2010). [CrossRef]  

31. A. C. Beveridge, T. E. McGrath, G. J. Diebold, and A. A. Karabutov, “Photoacoustic shock generation in carbon suspensions,” Appl. Phys. Lett. 75(26), 4204–4206 (1999). [CrossRef]  

32. A. García-Abuín, D. Gomez-Diaz, M. D. La Rubia, and J. M. Navaza, “Density, speed of sound, viscosity, refractive index, and excess volume of n-methyl-2-pyrrolidone + ethanol (or water or ethanolamine) from T = (293.15 to 323.15) K,” J. Chem. Eng. Data 56(3), 646–651 (2011). [CrossRef]  

33. M. J. Dávila and J. P. Martin Trusler, “Thermodynamic properties of mixtures of N-methyl-2-pyrrolidinone and methanol at temperatures between 298.15 K and 343.15 K and pressures up to 60 MPa,” J. Chem. Thermodyn. 41(1), 35–45 (2009). [CrossRef]  

34. B. García, S. Aparicio, R. Alcalde, M. J. Davila, and J. M. Leal, “Modeling the PVTx behavior of the N-Methylpyrrolidinone/Water mixed solvent,” Ind. Eng. Chem. Res. 43(12), 3205–3215 (2004). [CrossRef]  

35. NIST Chemistry WebBook, “Thermophysical Properties of Fluid Systems” (National Institute of Standards and Technology, U. S. Department of Commerce, 2017) as recalculated in https://en.wikipedia.org/wiki/Water_(data_page).

36. N. Hirai and H. Eyring, “Bulk viscosity of liquids,” J. Appl. Phys. 29(5), 810–816 (1958). [CrossRef]  

37. Scientific background on the Nobel Prize in physics 2010: Graphene (Royal Swedish Academy of Sciences, 2010).

38. Y. Hernandez, V. Nicolosi, M. Lotya, F. M. Blighe, Z. Sun, S. De, I. T. McGovern, B. Holland, M. Byrne, Y. K. Gun’Ko, J. J. Boland, P. Niraj, G. Duesberg, S. Krishnamurthy, R. Goodhue, J. Hutchison, V. Scardaci, A. C. Ferrari, and J. N. Coleman, “High-yield production of graphene by liquid-phase exfoliation of graphite,” Nat. Nanotechnol. 3(9), 563–568 (2008). [CrossRef]   [PubMed]  

39. M. Lotya, Y. Hernandez, P. J. King, R. J. Smith, V. Nicolosi, L. S. Karlsson, F. M. Blighe, S. De, Z. Wang, I. T. McGovern, G. S. Duesberg, and J. N. Coleman, “Liquid phase production of graphene by exfoliation of graphite in surfactant/water solutions,” J. Am. Chem. Soc. 131(10), 3611–3620 (2009). [CrossRef]   [PubMed]  

40. U. Khan, A. O’Neill, M. Lotya, S. De, and J. N. Coleman, “High-concentration solvent exfoliation of graphene,” Small 6(7), 864–871 (2010). [CrossRef]   [PubMed]  

41. M. Lotya, P. J. King, U. Khan, S. De, and J. N. Coleman, “High-concentration, surfactant-stabilized graphene dispersions,” ACS Nano 4(6), 3155–3162 (2010). [CrossRef]   [PubMed]  

42. T. A. Vartanyan, N. B. Leonov, and S. G. Przhibelskii, “Application of localized surface plasmons to study morphological changes in metal nanoparticles,” in Plasmons: Theory and Applications, K. N. Helsey, ed. (Nova Science Publishers, 2011).

43. R. R. Nair, P. Blake, A. N. Grigorenko, K. S. Novoselov, T. J. Booth, T. Stauber, N. M. R. Peres, and A. K. Geim, “Fine structure constant defines visual transparency of graphene,” Science 320(5881), 1308 (2008). [CrossRef]   [PubMed]  

44. B. G. Ghamsari, J. Tosado, M. Yamamoto, M. S. Fuhrer, and S. M. Anlage, “Measuring the complex optical conductivity of graphene by FabryPérot reflectance spectroscopy,” Sci. Rep. 6(1), 34166 (2016). [CrossRef]   [PubMed]  

45. W. Wang, M. Rong, A. B. Murphy, Y. Wu, J. W. Spencer, J. D. Yan, and M. T. C. Fang, “Thermophysical properties of carbon-argon and carbon-helium plasmas,” J. Phys. D 44(35), 355207 (2011). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (9)

Fig. 1
Fig. 1 (a) Optical and (b) AFM images with transverse profiles A and B of Grwater; (c) TEM image of GrNMP; Raman G′ band spectra of (d) Grwater and (e) GrNMP: curves A and B refer to the excitation wavelengths 488 nm and 514 nm respectively, dashed arrows show the corresponding maximum positions of this excitation-dependent band.
Fig. 2
Fig. 2 Schematic setup: Continuum Minilite II – laser source; P1, P2, P3 – polarizers; BE – beam expander; BS – beam splitter; L1, L2 – lenses; F1, F2, F3 – neutral filters sets; D1, D2, D3 – detectors; S – sample cuvette; DP – diaphragm; dashed rectangle shows a light protection mask. Insets: A – X-axe beam intensity distribution, measured at L1 focal position (solid) and its Gaussian fit (dash); B – image of SBS from pure NMP on the screen put on DP; C – time profiles of incident beam (solid) and backward scattered (dash); D – image of the interaction area from the cuvette with pure solvent in the SBS observation conditions; E – the same image from the graphene suspension when no SBS is observed.
Fig. 3
Fig. 3 Incident energy dependences of: (a) back scattered energy from NMP-based liquids, (b) back scattered energy from water-based liquids, (c) transmitted energy through NMP-based liquids, and (d) transmitted energy through water-based liquids. Data correspond to (⬤) pure NMP, (⬛) pure water, (⬤) aqueous solution of NaC, and suspensions of different graphene concentration with following absorption coefficients αe: () 0.013 cm−1, (▼) 0.023 cm−1, (▲) 0.032 cm−1, (◆) 0.048 cm−1, (◆) 0.001 cm−1, (⬣) 0.002 cm−1, (⬟) 0.004 cm−1.
Fig. 4
Fig. 4 SBS threshold energy against absorption coefficient of graphene suspensions in (a) NMP and in (b) water; SBS gain factor against absorption coefficient of graphene suspensions in (c) NMP and in (d) water. Filled points: experimental results; empty points in (b): results corrected for the shift due to NaC (blue circle); lines in (a) and in (b): linear fitting; curves in (c) and in (d): simulations with simultaneous variations of ρ and ΓB.
Fig. 5
Fig. 5 Effective NLA coefficients of suspensions against graphene linear absorption (⬛) in NMP and (⬤) in water.
Fig. 6
Fig. 6 Sketch of the electrostriction – thermal expansion antagonism occurring in a graphene suspension upon irradiation and giving impact on SBS quenching.
Fig. 7
Fig. 7 Temperature rise of a graphene flake upon 1 mJ laser-pulse absorption (solid curve); laser beam temporal profile (dashed curve) in arbitrary units.
Fig. 8
Fig. 8 SBS gain factor of (a) NMP and (b) water against linear absorption of graphene suspensions; points: experimental results; curves: simulations; varied parameters are: (a) n (curve A), ρ (curve B), vac (curve C), ΓB (curve D), γT (curve E), (b) n and ρ (curve A), vac and ΓB (curve B), γT (curve C). Relative changed SBS gain factor of (c) NMP and (d) water against corresponding varied parameters: (c) n (curve A), ρ (curve B), vac (curve C), ΓB (curve D), γT (curve E), (d) n (curve A), ρ (curve B), vac and ΓB (curve C), γT (curve D).
Fig. 9
Fig. 9 Absorption spectra of the as-prepared graphene suspensions in water and NMP.

Tables (3)

Tables Icon

Table 1 Effective NLA coefficients of liquid systems with different content (absorption coefficient) of graphene obtained from fitting the experimental points with Eq. (2).

Tables Icon

Table 2 Physical properties of the solvents at T = 293 K and standard pressure.

Tables Icon

Table 3 Relative absorptive changes (in cm) of refractive index, density, acoustic velocity, Brillouin line width and thermal optic coefficient providing the observed decrease of gB.

Equations (23)

Equations on this page are rendered with MathJax. Learn more.

G = g B E thr L π W 0 2 τ p = 25 ,
E ( E 0 ) = T lin ( E 0 β eff L π W 0 2 τ p E 0 2 ) ,
β eff α e = σ eff σ e .
( d d z α e ) I S = g B I S I 0 ,
g B ( α e ) = g B e g B T ( α e ) ,
g B e = ( γ e ω 0 ) 2 ε 0 2 c 3 n ρ v ac Γ B
g B T ( α e ) = γ e γ T ( α e ) ω 0 2 ε 0 2 c 3 n ρ v ac Γ B
Γ B = ( ω 0 ω B ) 2 η v ac 2 ρ
Δ T ( t ) = 3 E 0 F ( t ) σ e 4 π W 0 2 C p .
F ( τ 2 ) = F ( τ 1 ) + 4 π W 0 2 Λ 3 E 0 σ e ,
P ˜ = P ( 1 ± δ P )
δ P = Δ α P Δ α e , Δ α P = 1 P d P d α e .
n = ( 1 + 2 R ρ 1 R ρ ) 1 / 2 ,
r b = ( 9 d 2 σ e ρ s Δ α ρ 4 π ) 1 / 3
D O S ( λ ) = 8 π 2 h c λ h 2 λ Z 0 1 / 2 0 π / 2 d φ Z 0 Z 1 sin 2 ( φ ) , Z 0 = ( 1 + λ h λ ) 2 + 1 4 [ ( λ h λ ) 2 1 ] 2 , Z 1 = 4 λ h λ .
N b = 8 π W 0 2 Z R α e 9 d 2 σ e ρ s ,
f C = 4 π r b 3 α e 9 d 2 σ e ρ s ,
f 0 = h α e 3 σ e ρ s .
f C f 0 = V b V f .
R n 2 1 n 2 + 2 1 ρ ,
R = 4 π 3 a 0 M 0 .
n susp = [ 1 + 2 ( R solv ρ solv + R Gr ρ C ) 1 ( R solv ρ solv + R Gr ρ C ) ] 1 / 2 ,
δ ( R ρ ) = R solv ρ solv ( 1 f C ) .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.