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

Analysis of radiation pressure induced nonlinear optofluidics

Open Access Open Access

Abstract

We analyze two nonlinear optofluidic processes where nonlinearity is induced by the interplay between optical field and liquid interface. Specifically, guided optical waves generate radiation pressure on the liquid interface, which can in turn distort the liquid interface and modify the properties of the optical field. In the first example, we discuss the feasibility of nonlinear optofluidic solitons, where optical field is governed by the nonlinear Schrödinger equation and nonlinearity is effectively determined by liquid properties. Then, we analyze a nonlinear optofluidic process associated with a high quality (Q) factor whispering gallery mode (WGM) in a liquid droplet. Similar to Kerr effects, the WGM can produce a frequency shift proportional to the WGM power. Using liquid properties that are experimentally attainable, we find that it may only take a few photons to generate measurable WGM resonance shift. Such a possibility may eventually lead to nonlinear optics at single photon energy level.

© 2014 Optical Society of America

1. Introduction

Optofluidics [1–6] and optomechanics [7–9] have recently emerged as two important areas of research. Optofluidics aims to synergize microfluidics and optics to achieve novel functionalities including reconfigurable optical systems [1, 2], integrated optics [3, 4], lasers [5], and sensing [6]. Optomechanics, on the other hand, involves the dynamic interplay between optical field and mechanical motion [7–9]. To date, most optomechanics-related research utilizes solid resonators [10–12]. Only a limited number of studies investigated the mechanical interaction between optical field and fluids [13–18]. The earliest example is perhaps the classic work in [13], where the authors used a focused high power laser beam to create a bulge over a flat air-liquid interface. Similarly, in [15, 16], the authors used a high power laser beam focused onto a liquid droplet to distort its interface. More recently, optical radiation pressure was used in [17] to distort a flat liquid interface and form a tunable lens. By using a liquid mixture with extremely small interfacial tension, which can occur near the critical temperature of phase transition, it is possible to significantly reduce the optical power required for large interface distortion [17–19]. Additionally, stimulated Brillouin scattering in a hollow capillary tube filled with liquid has been reported recently [20]. The existence of high-Q whispering gallery mode (WGM) in an all-liquid droplet has also been experimentally confirmed [21, 22]. The focus of this work, however, is distinct from existing studies. Specifically, our goal is to demonstrate that under appropriate conditions, the dynamic interplay between optical force and liquid interface can lead to processes that are very similar to classical third-order nonlinear processes such as Kerr effects and optical solitons. As will be made clear in this paper, a defining feature of these nonlinear optofluidic processes is that the nonlinearity arises from the distortion of liquid interfaced induced by optical radiation pressure.

To analyze nonlinear optofluidic processes, it is necessary to solve the optical equations (e.g., Maxwell’s equations) and the fluidic equations (e.g., Navier-Stokes equations) in a self-consistent manner, where optical radiation pressure must be balanced by fluidic forces such as surface tension and buoyancy. Consequently, by tuning fluid parameters such as surface tension and density, one should be able to modulate and control the effective strength of optofluidic nonlinearity. This is in sharp contrast with traditional nonlinear optics, where nonlinear susceptibility (e.g., χ(2) or χ(3)) is an intrinsic material property and cannot be easily tuned. Furthermore, by using liquids with low surface tension, it should be possible to achieve large nonlinearity with low optical power. In fact, in this paper, we find that by reducing surface tension to a low but experimentally attainable level, the effective strength of nonlinear optofluidics can be several orders of magnitude stronger than the traditional Kerr effects. To the best of our knowledge, this possibility has never been discussed in existing literature.

To quantitatively analyze nonlinear optofluidic effects, we consider two examples in this paper. Figure 1(a) shows a liquid structure which contains a high index liquid as the waveguiding core, air as the top cladding, and a lower index liquid as the bottom cladding. Assuming low interfacial tension for the bottom liquid-liquid interface, the presence of optical field should deform the interface and create a bulge. At the same time, this bulge can enhance optical confinement along the transverse direction. Under certain conditions, we find that the propagation of optical field within the liquid bulge can be described by the nonlinear Schrödinger equation, the solutions to which are in the familiar form of optical solitons. This example illustrates a key feature of nonlinear optofluidics, namely that the presence of the optical field distorts the liquid interface; simultaneously, the change in the liquid interface shape also modifies the optical field. As a result, analyses of nonlinear optofluidics often require solving the coupled system of optical waves and fluids. Specifically, the optical field must satisfy the Maxwell’s equations plus the boundary conditions imposed by the deformable liquid interface. Meanwhile, the liquid system is governed by the relevant fluid equations, where the impact of optical radiation pressure must also be included. Clearly, the most important feature of nonlinear optofluidics is the coupling between the optical field and the liquids, where the coupling strength is effectively determined by the radiation pressure and the interfacial tension of the liquid interface.

 figure: Fig. 1

Fig. 1 (a) An optofluidic soliton formed from a self-guided optical wave that is confined within a liquid bulge formed through radiation pressure. (b) A liquid droplet that contains a high-Q WGM circulating along the equator. The radiation pressure of the WGM forms the bulge, which in turn shifts the WGM resonance frequency.

Download Full Size | PDF

The optofluidic soliton depicted in Fig. 1(a) may not be easy to implement experimentally. A more practical example is the system shown in Fig. 1(b), which is based on a liquid droplet that supports high-Q WGMs. In this example, the optical field of the WGM exerts radiation pressure on the droplet surface. Since the direction of radiation pressure always points from the high index core to the low index cladding [16], a bulge should form along the droplet equator. As a result, given sufficiently high optical power, both the effective optical path of the resonator and the corresponding resonance frequency should change. This phenomenon is very similar to the Kerr effect, where high optical power also shifts the effective cavity length by changing the refractive index of the liquid. In this paper, we derive a closed-form formula that can provide an order of magnitude estimate for this nonlinear optofluidic process. Using common liquid parameters, we find that this nonlinear optofluidic effect can be significantly stronger than the Kerr effect. In fact, for liquids with low but experimentally achievable surface tension, it may even be possible to produce measurable change in WGM resonance frequency at single photon energy level. Such a possibility may ultimately enable us to demonstrate nonlinear interaction between two single photons, which is obviously important for quantum information technology.

2. Optofluidic soliton

The optofluidic solitons can exist in the asymmetric liquid waveguide shown in Fig. 2(a). The refractive indices of the three layers are represented by n1, n2, and n3, while their fluid densities are given by  ρ1, ρ2, and  ρ3, respectively. For simplicity, we assume layer 3 is air and n3=1. In the absence of optical signals, the waveguide core is the planar middle layer with thickness h0 and has the highest refractive index (i.e., n2>n1 and n2>n3). By using the water-in-oil microemulsion system described in [17], it is possible to ensure that the core liquid layer (layer 2) possesses a higher refractive index but a lower mass density than the cladding liquid layer (layer 1), i.e., n2>n1 and ρ2<ρ1. Thus the configuration in Fig. 2 should be hydrodynamically stable for small deformation.

 figure: Fig. 2

Fig. 2 (a) Illustration of an optofluidic soliton. The structure contains two liquids (refractive indicesn1 and n2) and air (n3=1). The thickness of the waveguide is h0 in the absence of optical field. The radiation pressure of the guided optical signal produces the bulge shown in the figure. The thickness of the bulge that serves as the waveguide core is denoted as h(x)=h0+Δh(x). (b) The effective index of the asymmetric dielectric waveguide defined in (a) as a function of the core layer thickness h(x). The waveguide parameters are n1=1.5, n2=1.5043, and n3=1. The operation wavelength is 1μm. The exact solution (blue line) is calculated using standard waveguide theory. The dashed black line represents linear fitting of the exact solution in the range of 2.5 μm<h<4 μm. The slope gives n/h=1×103μm1.

Download Full Size | PDF

After coupling light into layer 2, the presence of the optical field generates radiation pressure on the dielectric boundaries. We assume that the interfacial tension between layer 1 and 2 is much smaller than that of layer 2 and 3. As a result, we can ignore the deformation of the air-liquid interface (the interface between layer 2 and 3) and focus on the liquid-liquid interface instead. We note that the direction of radiation pressure always points from the high refractive index medium towards the low refractive index medium [16]. Consequently, the presence of optical field should generate a liquid bulge as illustrated in Fig. 2(a). In turn, the geometrical deformation can lead to better optical confinement within the bulge. Qualitatively, such a nonlinear optofluidic process is very similar to the self-focusing of spatial solitons. In the following analysis, we establish an analytical framework and confirm that under certain conditions, the aforementioned optofluidic process can indeed be described by the nonlinear Schrödinger equation.

Our first step is to apply the effective index theory [23], a widely used theory that allows us to reduce the original three-dimensional (3D) problem into a two-dimensional (2D) one. In our case, similar to the treatment of structures such as ridge waveguides, we assume that the optical field depends only on x and z and use an effective index neff(x,z) to account for the field’s variation along the y direction. For the specific example in Fig. 2(a), we assume that the optical field’s distribution along the y axis is simply given by the fundamental y-polarized mode in a planar waveguide. Then, we approximate the x and z dependence of the optical field as:

E=A(x,z)ei(ωtkx)e^y,
where A(x,z) represents field amplitude, ω is angular frequency, k=n0ω/c is the wave vector in the effective medium, with n0 being the effective mode index of the fundamental y-polarized mode in the slab waveguide with no surface deformation (i.e., h=h0), and c is the speed of light. With optical field in the liquid core, the waveguide thickness h is no longer a constant. Correspondingly, the effective index, determined by the effective mode index of a planar waveguide with thickness h, also depends on x and z. Figure 2(b) shows the relationship between neff and h, with waveguide parameters given in the figure caption. We note that in the thickness range of 2.5μm<h<4 μm, neff increases almost linearly as a function of h. Therefore, assuming small interface deformation, we can linearize the relationship between neff and h:

neff(x,z)=n0+(nh)Δh(x,z).

Given the effective index in Eq. (2), the electric field in Eq. (1) should satisfy:

2E+neff2(x,z)ω2c2E=0.
Note that Eq. (3) no longer depends on y. Then, assuming a slow varying envelop A(x,z), i.e., kA/x>>2A/x2, we can simplify Eq. (3) into:
2ikAx+2Az2=2n0ω2c2(nh)Δh(x,z)A(x,z).
Equation (4) shows that the shape of the liquid bulge, i.e., Δh(x,z), can significantly change the properties of the guided optical wave.

In order to determine Δh(x,z), we must first derive its governing equation. For the electric field polarized along the y axis, the optical radiation pressure at the layer 1 and 2 interface is given by popt=ε0(n22n12)n22|A(x,z)|2/(2n12) [16]. Since we look for a solution that is self-guided and propagates along the x direction, its intensity, i.e., |A(x,z)|2, should not depend on x. Therefore, the shape of the bulge Δh(x,z) should not depend on x either. Then, in equilibrium, interfacial tension on a deformed interface is balanced with buoyancy and radiation pressure, according to the Young-Laplace equation. When the bulge has a small slope (Δh/z<<1), this balance leads to:

σd2Δh(z)dz2+ΔρgΔh(z)=12ε0(n22n12)n22n12|A(x,z)|2,
where Δρ=ρ1ρ2, and σ is the interfacial tension between layer 1 and 2. In Eq. (5), we ignore the deformation of the air-liquid interface, since the air-liquid surface tension is assumed to be much higher than the liquid-liquid interfacial tension. To obtain a closed form solution, we further assume that the interfacial tension term in Eq. (5) is even smaller than gravity and is therefore neglected. Under this scenario, the height of the liquid bulge is proportional to optical intensity, i.e., Δh(z)=ε0(n22n12)n22|A(x,z)|2/(2n12Δρg). Substituting this result into Eq. (4), we obtain:
2ikAx+2Az2=χeffω2c2|A|2A,
whereχeff=(nh)ε0n0(n22n12)n22Δρgn12.
Equation (6a) takes the familiar form of the nonlinear Schrödinger equation [24]. Its fundamental solution is:

A(x,z)=A0sech(zz0)eiγx,
wherez0=n0k|A0|2χeff,γ=kχeff|A0|24n02.

Our assumptions in deriving the form of optofluidic solitons are either the standard approach or can be justified using carefully controlled experimental conditions. For example, the slow varying envelope assumption and the effective index theory are commonly used in nonlinear optics and integrated optics. Our assumptions on liquid properties such as density, refractive index, and surface tension, however, warrant additional discussion. For example, two key requirements are: 1) in comparison with the cladding medium (layer 3), the core medium (layer 2) should possess lower density and higher refractive index; 2) the surface tension between layer 1 and 2 should be small. Both requirements can be satisfied by using a water-in-oil microemulsion system that contains a mixture of water, sodium dodecyl sulfate, toluene, and n-butanol-1 [17]. At a temperature slightly higher than the critical temperature (Tc = 35 °C) for phase transition, the mixture phase separates into two different micellar phases, with interfacial tension given by σ=σ0(T/Tc1)2v, where σ0=1.04×104 N/m, and v=0.63. The two micellar phases possess different densities and refractive indices, with the density difference given by Δρ=(Δρ)0(T/Tc1)β, and the index difference given by Δn=(n/ρ)T(ρ1ρ2), where (Δρ)0=285 kg/m3,β=0.325 and (n/ρ)T=1.22×104 m3/kg. Given these conditions, at the temperature of T=35.5 , the interfacial tension is σ=3.18×108N /m, the density and refractive index differences are Δρ=35kg/m3, and Δn=0.0043. The relative magnitude of buoyancy to interfacial tension force is defined as the Bond number (B=ΔρgΔh/(σΔh/Δz2)). Assuming a soliton solution with a bulge size of Δz=20 μm and height Δh=1 μm, the Bond number is in the order of O(10). Such a large Bond number indicates that the interfacial tension effect is negligible compared to buoyancy effect. When the Bond number is small, the interfacial tension effect becomes comparable or even stronger than gravity, then we must solve the two coupled equations (Eqs. (4) and (5)) without any approximations. For such cases, analytical solutions would unfortunately be unattainable and numerical methods must be applied.

An attractive feature of the optofluidic soliton is that its nonlinear behaviors can be tuned by varying the parameters of the liquid systems. In particular, an interface with weak interfacial tension should possess high nonlinear efficiency. As an example, consider the waveguide structure analyzed in Fig. 2(b). From linear fitting of effective index for 2.5 μm<h<4 μm, we find n/h=1×103μm1. Using this value and liquid parameters defined in the previous paragraph, the effective nonlinear coefficient defined in Eq. (6b) is approximately χeff=7.44×1013 m2/V2. Comparing Eq. (6a) to the equation that describes spatial self-focusing, e.g., Eq. (7).1.18) of [24], it is clear that χeff essentially represents the third order nonlinear susceptibility of the optofluidic system. For typical optical materials, the third order nonlinear susceptibility χ(3) varies in the range of 1022 m2/V2 to 1018 m2/V2 [24]. Therefore, by using liquid systems with weak interfacial tension, we can effectively increase the third order nonlinearity by several orders of magnitude.

We can use Eq. (7) to estimate the optical power required to form the aforementioned optofluidic soliton. For this calculation, we assume z0=10 µm, λ=1 μm, n01.5, and use χeff=7.44×1013 m2/V2 given in the paragraph above. Then, based on Eq. (7b), the required electric field strength is A0=2.61×104 V/m. If we assume that the optical beam is confined in a cross-section of width  W=10 µm (i.e., soliton width) and height H=3 µm (i.e., waveguide thickness), the optical power of such a beam can be simply estimated as n0cε0|A0|2WH/2. Using parameters given above, we find that the optical power of this optofluidic soliton is only  41 µW. We can also use the value for A0 to estimate the bulge height. Applying the discussion immediately above Eq. (6), we can relate the magnitude of bulge height to the optical field amplitude as Δhε0(n22n12)n22|A0|2/(2n12Δρg). Using waveguide parameters given in the caption of Fig. 2, g=9.8m/s2, and assuming Δρ=35kg/m3 (refer to the discussion on the water-in-oil microemulsion system), we find the bulge height is approximately Δh114 nm. The relative deformation (Δh/h) is approximately 4%.

The discussion in this section is mainly theoretical, where our primary aim is to show that radiation pressure induced interface deformation can lead to nonlinear processes very similar to traditional third order nonlinear processes such as optical solitons. In terms of experiment, it is perhaps easier to investigate nonlinear optofluidic processes using a high-Q optical resonator based on a liquid droplet, as will be discussed in the next section.

3. WGM induced droplet deformation

Figure 3(a) illustrates the second example of radiation pressure induced nonlinear optofluidic processes, where the key element is the high-Q WGM circulating along the equator of a liquid droplet in air. Since the direction of the radiation pressure always points from the high refractive index material to the low index material, the presence of a high-Q WGM would push the spherical droplet surface outwards and consequently enlarge the circumference of the droplet’s equator. Since radiation pressure is proportional to the optical power carried by the WGM, under the limit of small droplet deformation, we expect that the equator circumference should increase linearly as a function of WGM power and shift the WGM resonance frequency as a direct result. This phenomenon is very similar to the Kerr effect. In the following analysis, we establish an analytical framework and provide an order of magnitude estimate for the frequency shift associated with this nonlinear optofluidic process. Our analysis suggests that the effective strength of the aforementioned nonlinear optofluidic process can be several orders of magnitude larger than the traditional Kerr effect.

 figure: Fig. 3

Fig. 3 (a) A liquid droplet with a high-Q WGM circulating near its equator. The radiation pressure of the WGM deforms the original spherical droplet (the blue circle) and generates the bulge (the solid black line), which is approximated as an oblate spheroid (the dashed black line). The normalized equator radius xe is defined as the ratio of the spheroid radius at the equator (a+ΔR) and the radius of the original sphere a. (b) The integral F(xe)=0πaκ¯(xe,θ)Y20(θ)sinθdθ (blue circles) as a function of the normalized equator radius xe. κ¯(xe,θ) is given in Eq. (12). The dimensionless constant Γσ is extracted using Eq. (14) and least square fitting (dashed red line).

Download Full Size | PDF

The starting point of our analysis is the Young-Laplace equation. We assume a small Bond number, i.e., buoyant force is much smaller than surface-tension force and is ignored. As a result, we find:

ΔP+Popt=2σκ¯,
where Δp represents the pressure difference inside and outside of the droplet, Popt is the radiation pressure generated by the WGM, σ represents the surface tension of the droplet, and κ¯ is the mean curvature of the droplet. Since our main purpose is to provide an order of magnitude estimate, we take the following steps to simplify our analysis. First, we restrict our considerations to the fundamental transverse electric (TE) mode. Mathematically, this means that the optical field of the WGM can be written as [25]:
ElmTE=gl(kqr)Xlm(θ,ϕ)eiωt,
HlmTE=ikqZq×[gl(kqr)Xlm(θ,ϕ)]eiωt,
where the angular mode number l and m are integers and satisfy l>0 and lml, ω is angular frequency. The subscript “q” is either “co” if it represents core parameters of the WGM resonator or “cl” if it denotes cladding parameters. For example, kco refers to the optical wave vector within the droplet, and kcl is the cladding wave vector. Similarly, Zco and Zcl represent the impedance of the droplet core and cladding material, respectively. In Eq. (9), Xlm is the vector spherical harmonic function [25] and represents the angular variation of the WGM. As to the radial profile of the WGM, the function gl(kqr) is either the l-th order spherical Bessel function (for core) or the spherical Hankel function of the first kind (for cladding), i.e.,:
gl(kqr)={Acojl(kcor),   r<aAclhl(1)(kclr),  r>a
where a is the radius of the un-deformed sphere. For the fundamental WGM, the m=l mode and the m=l mode are identical except for the direction of light circulation. For convenience, we only consider the m=l mode and denote it as |ll. Given Eq. (9a), the WGM radiation pressure Popt on the droplet interface is given by [16]:
popt=12ε0(nco2ncl2)|Esurf|2,
where ε0 is free space permittivity, nco and ncl represent core and cladding refractive indices, and Esurf is the electric field on the droplet surface. Note that the expression for the optical force in Eq. (11) is slightly different from the one we used in Eq. (5), which is due to the difference in electric field direction [16]. From the expression in Eq. (9a), we can verify that |Esurf|2, and therefore Popt, depends on θ but not on ϕ. Additionally, for the |ll mode, |Esurf|2 is symmetric with respect to the equator plane (i.e., θ=π/2).

Numerically solving Eq. (9) to (11) can give us the form of droplet deformation. However, the mathematical process is complex and can obscure the physics. In this paper, we emphasize the physical reality and utilize several assumptions to obtain an intuitive estimate of droplet deformation. First, we assume that the droplet deformation can be approximated by an oblate spheroid, as shown in Fig. 3(a). This assumption simplifies our analysis considerably. For example, assuming a spheroid shape, the mean curvature of the droplet interface is simply:

κ¯(xe,θ)=12axpxe2xe2+(xp2xe2)sin2θ[xe2+(xp2xe2)sin2θ]3/2,
where xe and xp are defined as R(θ=π/2)/a and R(θ=0)/a, which correspond to the normalized droplet radius at the equator (θ=π/2) and at the north pole (θ=0), respectively. (Volume conservation requires xe2xp=1.) Furthermore, under the spheroid assumption, the WGM resonance shift can be readily estimated using the perturbation theory in [26].

The next step is to expand both sides of Eq. (8) using spherical harmonic function YLM(θ,ϕ). From symmetry arguments, we can conclude that most of the spherical harmonic expansion coefficients are zero. First, Δp is constant over the droplet interface, whereas both popt and κ¯ are independent of ϕ. Therefore, any M0 term must be zero. Additionally, both popt and κ¯ are symmetric with respect to θ=π/2. Hence, any term with odd L number must also be zero. The first two non-zero terms of spherical harmonic expansion involve overlap integrals with Y00 and Y20, and are given by:

2Δp+0πpopt(θ)sinθdθ=2σ0πκ¯(xe,θ)sinθdθ,
0πpopt(θ)Y20(θ)sinθdθ=2σ0πκ¯(xe,θ)Y20(θ)sinθdθ,
where Y20=5/16π(3cos2θ1). We can physically interpret Eq. (13) as follows. Given the presence of radiation pressure, the difference in fluid pressure Δp should be changed by the amount given in Eq. (13a). In other words, the reduction in fluid pressure is a constant term that balances the average magnitude of the radiation force (integrated over the entire droplet surface). After subtracting this constant component (i.e., Δp), it is the θ dependence of the radiation pressure that determines droplet deformation. In applying Eq. (13b), we essentially use the first non-zero alternating component to estimate the magnitude of droplet deformation. Obviously, to satisfy Eq. (8) exactly, the mean curvature must contain higher “frequency” components. However, using the curvature to calculate droplet shape requires double integration, which can be regarded as a low-pass filter that significantly suppresses these higher “frequency” components. Therefore, it is reasonable to discard the higher order spherical harmonic terms (L4), and use Eq. (13) to estimate the magnitude of droplet deformation.

To further simplify Eq. (13b), we note that the right hand side of Eq. (13b) is a function of xe only. Multiplying the integral by droplet radius a, we can define a dimensionless parameter F(xe)=0πaκ¯(xe,θ)Y20(θ)sinθdθ and numerically evaluate it as a function of xe. The result, which is shown in Fig. 3(b), suggests that this integral is almost a perfect linear function of xe. Thus we introduce a dimensionless parameter Γσ as:

F(xe)=0πaκ¯(xe,θ)Y20(θ)sinθdθΓσ(xe1).
Through least square fitting, we determine the value of Γσ to be 1.01.

To obtain a closed form formula, we express |Esurf|2 in Eq. (11) as:

|Esurf|2=|Esurfpeak|2flm(θ),
where |Esurfpeak|2 is the maximum value of |E|2 over the droplet surface. Hence flm(θ) represents its angular variation and is normalized to 1. Then, the left-hand side of Eq. (13b) becomes:
0πpopt(θ)Y20(θ)sinθdθ=12ε0(nco2ncl2)|Esurfpeak|20πflm(θ)Y20(θ)sinθdθ.
This leads us to introduce another dimensionless parameter Γθlm as:
Γθlm=ncoaλ0πflm(θ)Y20(θ)sinθdθ.
The coefficient ncoa/λ is introduced to ensure that Γθlm is dimensionless and remains reasonably close to 1. Combining the results in Eqs. (13b), (14), (16) and (17), we find that the normalized droplet deformation (ΔR/a) near the spheroid equator is given by:

ΔRa=xe1=Γθlmε0λ4Γσσnco(nco2ncl2)|Esurfpeak|2.

Equation (18) relates droplet deformation to the peak electric field intensity on the droplet interface and is the key result of this paper. In the next section, we consider several examples of liquid droplets and calculate their deformation as a function of optical power associated with the circulating WGM.

4. Estimate of nonlinear optofluidic effects based on high-Q WGMs

Using Eq. (18), we can calculate the radiation pressure induced WGM frequency shift and quantitatively compare nonlinear optofluidics with the traditional Kerr effect. We choose the following parameters in our calculations. The core index is nco=1.33, the cladding index is ncl=1.0. The droplet radius varies from 10μm to 400μm. All WGMs are the fundamental TE mode |ll, and their angular momenta l are chosen to ensure that the WGM wavelengths λ are in the vicinity of 1.56 μm. Table 1 lists the l and λ of the TE |ll mode in such droplets with different radii a. The resonance frequencies are obtained by matching transverse field components across the droplet interface [27], which is assumed to be a perfect sphere.

Tables Icon

Table 1. The Angular Mode Number l and Resonance Wavelength λ of WGMs in Liquid Droplets.

Equation (18) links droplet deformation with peak surface field intensity, which is unfortunately not directly measurable. Therefore, we need to further relate the peak field intensity in Eq. (18) with the optical power or energy of the circulating WGM. To accomplish this, we take advantage of the fact that the deformation is linearly proportional to |E¯surfpeak|2. Specifically, we choose an electric field amplitude such that |E¯surfpeak|=1 and calculate the corresponding field distribution, WGM power flux and total energy. Figure 4 shows the field distribution and the ϕ component of the Poynting vector of a WGM (l=257) in a droplet with radius a=50μm. By integrating the Poynting vector across the ϕ=0 plane, we can calculate the total optical power carried by the WGM. Similarly, through volume integration of energy density, we can obtain the total energy stored within the WGM. Then, we multiply the electric field intensity with an appropriate normalization factor such that the total power or the energy of the WGM becomes the desired value. This normalization factor is then used to calculate |E¯surfpeak|2 associated with the desired WGM power or energy. Once |E¯surfpeak|2 is obtained, the only unknown parameter in Eq. (18) is the dimensionless factor Γθlm, which can be obtained through numerical integration of Eq. (17) and is shown in Fig. 4(d).

 figure: Fig. 4

Fig. 4 (a) The radial dependence of |E|2 of a fundamental TE mode (l=257) in a spherical droplet with radius a=50μm. We assume |Esurfpeak|=1 and take θ=π/2 and ϕ=0. (b) The angular dependence of |E|2 for the WGM in (a). The value of |E|2 is evaluated over the droplet surface, with ϕ=0. Due to our normalization scheme, the curve is also flm(θ). (c) The power flux of the WGM in (a) within the ϕ=0 plane. Only the eϕ component of the Poynting vector is shown. The white circle represents the droplet surface. (d) The value of Γθlm for the fundamental TE mode |ll in droplets with different radius. The sphere radii and WGM parameters are listed in Table 1. The Γθlm values (blue circles) are obtained numerically using Eq. (17) and simply connected together using the dashed line.

Download Full Size | PDF

Using the procedure described above, we set the power associated with the circulating WGM to be 1 W and use Eq. (18) to calculate the radiation induced interface deformation for different droplet radius. The results are shown in Fig. 5(a). The red circles give the normalized droplet deformation (ΔR/a) induced by the WGM radiation pressure, where we assume the surface tension is that of water (σ=72mN/m). The blue and black diamonds represent the estimated Kerr effect in water (blue) and CS2 (black), produced by the same peak surface field |Esurfpeak|2. These estimates are simply obtained by using Δnχ(3)|Esurfpeak|2, where χ(3) is the third order nonlinear susceptibility of water (2.5×1022m2/V2) or CS2 (3.1×1020m2/V2). Since both ΔR/a and Δn represent the same physical effects, i.e., the relative change in the optical path of the high-Q resonator, they can be shown in the same figure for direct comparison. Clearly, the nonlinear optofluidic effect is three to five orders of magnitude stronger than the Kerr effect.

 figure: Fig. 5

Fig. 5 (a) Radiation pressure induced droplet deformation (ΔR/a) in droplets with different radii (red circles). The total power of the WGM that circulates along the droplet equator is fixed at 1 W. For comparison, the changes in refractive index (Δn) due to the Kerr effect are shown in the same figure, which are estimated using Δnχ(3)|Esurfpeak|2. The Kerr effects in water and in CS2 are represented as the blue and black diamonds, respectively. All points are connected by dashed lines. (b) The relative shift in WGM frequency induced by the radiation pressure of a single photon. Two different values are used for surface tension: σ=72mN/m (blue circles), and σ=1mN/m (red crosses). All points are connected by dashed lines.

Download Full Size | PDF

Since the nonlinear optofluidic process can be significantly stronger than the Kerr effect, it is worth asking the question: Is it possible to use only a few photons to generate measurable WGM frequency shift? To answer this question, we first calculate the value of |Esurfpeak|2 such that the total WGM energy equals to single photon energy ω. Then, we substitute this value into Eq. (18) and calculate the normalized radius change ΔR/a induced by the presence of a single photon. Finally, using the perturbation theory results in [26], the WGM frequency shift associated with the deformed spheroid is:

Δωω=2l12(l+1)ΔRa.

Combining Eqs. (18) and (19), we can now calculate the radiation pressure induced frequency shift. The results are shown in Fig. 5(b). One set of data (blue circles) is obtained using σ=72mN/m, which is a typical value for air-liquid interface. The other set of data (red crosses) assumes a low but experimentally achievable surface tension σ=1mN/m. Given the fact that WGMs with Q factors as high as 2.3 × 106 have been observed using a liquid resonator [22], Fig. 5(b) suggests that one should be able to use only a few photons to produce experimentally observable resonance shift. If one can reduce the surface tension further down to the level of σ=0.1mN/m, then the presence of even a single photon can significantly change the characteristics of the liquid droplet resonator.

It is worth remarking that liquid systems with ultralow surface tensions have been experimentally attained using surfactants [28–30]. For example, by introducing a bimolecular layer of preformed ferric stearate, one can reduce the surface tension of water down to 1mN/m [28]. In emulsion or microemulsion systems containing water (a polar fluid) and oil (a nonpolar fluid), surfactants can lower surface tension as small as 1μN/m at an optimal concentration [29]. Therefore, for these liquid systems with ultralow surface tension, nonlinear optics at single photon energy level should be experimentally feasible.

In our WGM analyses, the effect of gravity is ignored. To justify this choice, we can estimate the Bond number of the liquid droplet. For a sphere with radius a, the gravitational and the surface tension effects can be estimated as ΔρgΔR and σΔR/a2, respectively. Consequently, the Bond number is simply B=ΔρgΔR/(σΔR/a2). For an order of magnitude estimate, we can assume Δρ~103kg/m3, g~10m/s2, and a=100 μm. With these values, even if the surface tension is as low as σ=1mN/m, the Bond number remains small (B~0.1) and we can safely neglect the gravitational effect in Eq. (8).

It is also of interest to briefly consider other types of liquid resonators such as an oil droplet (nco=1.45) immersed in water (nco=1.33). Such a resonator is similar to the one investigated in [21]. As a specific example, we estimate the deformation induced by 1W WGM power circulating within an oil-in-water droplet with radius a=70 μm and interfacial tension σ=72 mN/m. This resonator supports an TE-polarized WGM with l=397 and resonance wavelength λ=1.5610 μm. Using the procedure described above, we find that 1 W WGM power should induce a relative radius change of ΔR/a=2.94×104. In comparison with the results shown in Fig. 5(a), for an water-in-air droplet with identical radius, interfacial tension and WGM power, the deformation is ΔR/a=2.70×104. This result suggests that the conclusions we obtained using water-in-air droplets are broadly applicable to other types of liquid resonators.

Equation (18) suggests that droplet deformation depends on interfacial tension but not on liquid viscosity. This is to be expected, since our discussion is based on static analysis, i.e., the Young-Laplace equation in Eq. (8), which does not involve viscosity. We do, however, expect that viscosity should play a role in the dynamics of droplet deformation.

Finally, we point out that both the spherical WGM system analyzed here and the planar soliton case discussed in section 2 can be connected to liquid-based spheroid resonators. On the one hand, the spherical droplet analyzed in section 3 is a special case of the liquid spheroid resonators. On the other hand, a prolate spheroid resonator with an extremely large major axis can essentially be regarded as an infinitely long dielectric cylinder, which can also support WGM that circulates along the perimeter of the cylinder. And if we “unwrap” the infinitely long cylinder, the WGM traveling along the cylinder perimeter becomes similar to the self-guided soliton waves analyzed in section 2.

5. Summary

In this paper we analyze the possibility of using all liquid systems to achieve nonlinear optical effects with extremely low power threshold. The defining feature of the proposed nonlinear optofluidic processes is the interface deformation due to optical radiation pressure. In particular, we find that through the formation of a bulged interface, optical waves could self-focus within a liquid slab waveguide. Under appropriate conditions, these self-guided optical waves are governed by the nonlinear Schrödinger equation and can be described by the familiar soliton solutions. We then consider a spherical liquid droplet that supports high-Q WGMs. With sufficiently high power, the optical force associated with the circulating WGM could deform the liquid droplet and induce a frequency shift of the WGM resonance. By applying spherical harmonic expansion, we estimate that the radiation pressure induced nonlinearity is several orders of magnitude stronger than the traditional Kerr effect. In fact, we find that by using a liquid system with low but experimentally achievable surface tension, it is possible to produce measurable frequency shift at the energy level of a few photons. Such effect may ultimately lead to nonlinear interactions between two single photons.

Acknowledgments

We gratefully acknowledge Dr. William Olbricht and the support from the National Science Foundation (CBET 1438112).

References and links

1. D. Psaltis, S. R. Quake, and C. Yang, “Developing optofluidic technology through the fusion of microfluidics and optics,” Nature 442(7101), 381–386 (2006). [CrossRef]   [PubMed]  

2. U. Levy and R. Shamai, “Tunable optofluidic devices,” Microfluid Nanofluidics 4(1-2), 97–105 (2008). [CrossRef]  

3. C. Monat, P. Domachuk, and B. J. Eggleton, “Integrated optofluidics: A new river of light,” Nat. Photonics 1(2), 106–114 (2007). [CrossRef]  

4. S.-H. Kim, J.-H. Choi, S.-K. Lee, S.-H. Kim, S.-M. Yang, Y.-H. Lee, C. Seassal, P. Regrency, and P. Viktorovitch, “Optofluidic integration of a photonic crystal nanolaser,” Opt. Express 16(9), 6515–6527 (2008). [CrossRef]   [PubMed]  

5. Z. Li and D. Psaltis, “Optofluidic dye lasers,” Microfluid Nanofluidics 4(1-2), 145–158 (2007). [CrossRef]  

6. H. Zhu, I. M. White, J. D. Suter, P. S. Dale, and X. Fan, “Analysis of biomolecule detection with optofluidic ring resonator sensors,” Opt. Express 15(15), 9139–9146 (2008). [CrossRef]   [PubMed]  

7. H. Rokhsari, T. J. Kippenberg, T. Carmon, and K. J. Vahala, “Theoretical and experimental study of radiation pressure-induced mechanical oscillations (parametric instability) in optical microcavities,” IEEE J. Sel. Top. Quantum Electron. 12(1), 96–107 (2006). [CrossRef]  

8. T. J. Kippenberg and K. J. Vahala, “Cavity opto-mechanics,” Opt. Express 15(25), 17172–17205 (2007). [CrossRef]   [PubMed]  

9. T. J. Kippenberg and K. J. Vahala, “Cavity optomechanics: Back-action at the mesoscale,” Science 321(5893), 1172–1176 (2008). [CrossRef]   [PubMed]  

10. J. Hofer, A. Schliesser, and T. J. Kippenberg, “Cavity optomechanics with ultrahigh-Q crystalline microresonators,” Phys. Rev. A 82(3), 031804 (2010). [CrossRef]  

11. S. Tallur, S. Sridaran, and S. A. Bhave, “A monolithic radiation-pressure driven, low phase noise silicon nitride opto-mechanical oscillator,” Opt. Express 19(24), 24522–24529 (2011). [CrossRef]   [PubMed]  

12. A. Cho, “Putting light’s light touch to work as optics meets mechanics,” Science 328(5980), 812–813 (2010). [CrossRef]   [PubMed]  

13. A. Ashkin and J. M. Dziedzic, “Radiation pressure on a free liquid surface,” Phys. Rev. Lett. 30(4), 139–142 (1973). [CrossRef]  

14. I. I. Komissarovak, G. V. Ostrovskaya, and E. N. Shedova, “Light pressure induced deformations of a free liquid surface,” Opt. Commun. 66(1), 15–20 (1988). [CrossRef]  

15. J.-Z. Zhang and R. K. Chang, “Shape distortion of a single water droplet by laser-induced electrostriction,” Opt. Lett. 13(10), 916–918 (1988). [CrossRef]   [PubMed]  

16. H. M. Lai, P. T. Leung, K. L. Poon, and K. Young, “Electrostrictive distortion of a micrometer-sized droplet by a laser pulse,” J. Opt. Soc. Am. B 6(12), 2430–2437 (1989). [CrossRef]  

17. A. Casner and J.-P. Delville, “Adaptative lensing driven by the radiation pressure of a continuous-wave laser wave upon a near-critical liquid liquid interface,” Opt. Lett. 26(18), 1418–1420 (2001). [CrossRef]   [PubMed]  

18. A. Casner, J.-P. Delville, and I. Brevik, “Asymmetric optical radiation pressure effects on liquid interfaces under intense illumination,” J. Opt. Soc. Am. B 20(11), 2355–2362 (2003). [CrossRef]  

19. J.-P. Delville, M. Robert de Saint Vincent, R. D. Schroll, H. Chraïbi, B. Issenmann, R. Wunenburger, D. Lasseux, W. W. Zhang, and E. Brasselet, “Laser microfluidics: fluid actuation by light,” J. Opt. A, Pure Appl. Opt. 11(3), 034015 (2009). [CrossRef]  

20. G. Bahl, K. H. Kim, W. Lee, J. Liu, X. Fan, and T. Carmon, “Brillouin cavity optomechanics with microfluidic devices,” Nat Commun 4, 1994 (2013). [CrossRef]   [PubMed]  

21. M. Hossein-Zadeh and K. J. Vahala, “Fiber-taper coupling to Whispering-Gallery modes of fluidic resonators embedded in a liquid medium,” Opt. Express 14(22), 10800–10810 (2006). [CrossRef]   [PubMed]  

22. A. Jonáš, Y. Karadag, M. Mestre, and A. Kiraz, “Probing of ultrahigh optical Q-factors of individual liquid microdroplets on superhydrophobic surfaces using tapered optical fiber waveguides,” J. Opt. Soc. Am. 29(12), 3240–3247 (2012). [CrossRef]  

23. A. Yariv and P. Yeh, Photonics: Optical Electronics in Modern Communications (Oxford University, 2007).

24. R. W. Boyd, Nonlinear Optics, 3rd ed. (Academic, Amsterdam, 2008).

25. J. D. Jackson, Classical Electrodynamics, 3rd ed. (John Wiley & Sons, 1998).

26. H. M. Lai, P. T. Leung, K. Young, P. W. Barber, and S. C. Hill, “Time-independent perturbation for leaking electromagnetic modes in open systems with application to resonances in microdroplets,” Phys. Rev. A 41(9), 5187–5198 (1990). [CrossRef]   [PubMed]  

27. J. R. Buck and H. J. Kimble, “Optimal sizes of dielectric microspheres for cavity QED with strong coupling,” Phys. Rev. A 67(3), 033806 (2003). [CrossRef]  

28. A. Datta, S. Kundu, M. K. Sanyal, J. Daillant, D. Luzet, C. Blot, and B. Struth, “Dramatic enhancement of capillary wave fluctuations of a decorated water surface,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 71(4), 041604 (2005). [CrossRef]   [PubMed]  

29. H. Leitão, A. M. Somoza, M. M. Telo da Gama, T. Sottmann, and R. Strey, “Scaling of the interfacial tension of microemulsions: A phenomenological description,” J. Chem. Phys. 105(7), 2875 (1996). [CrossRef]  

30. H. Chraibi, D. Lasseux, R. Wunenburger, E. Arquis, and J.-P. Delville, “Optohydrodynamics of soft fluid interfaces: optical and viscous nonlinear effects,” Eur Phys J E Soft Matter 32(1), 43–52 (2010). [CrossRef]   [PubMed]  

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 (5)

Fig. 1
Fig. 1 (a) An optofluidic soliton formed from a self-guided optical wave that is confined within a liquid bulge formed through radiation pressure. (b) A liquid droplet that contains a high-Q WGM circulating along the equator. The radiation pressure of the WGM forms the bulge, which in turn shifts the WGM resonance frequency.
Fig. 2
Fig. 2 (a) Illustration of an optofluidic soliton. The structure contains two liquids (refractive indices n 1 and n 2 ) and air ( n 3 = 1 ). The thickness of the waveguide is h 0 in the absence of optical field. The radiation pressure of the guided optical signal produces the bulge shown in the figure. The thickness of the bulge that serves as the waveguide core is denoted as h ( x ) = h 0 + Δ h ( x ) . (b) The effective index of the asymmetric dielectric waveguide defined in (a) as a function of the core layer thickness h ( x ) . The waveguide parameters are n 1 = 1.5 , n 2 = 1.5043 , and n 3 = 1 . The operation wavelength is 1 μ m . The exact solution (blue line) is calculated using standard waveguide theory. The dashed black line represents linear fitting of the exact solution in the range of 2.5   μ m < h < 4   μ m . The slope gives n / h = 1 × 10 3 μ m 1 .
Fig. 3
Fig. 3 (a) A liquid droplet with a high-Q WGM circulating near its equator. The radiation pressure of the WGM deforms the original spherical droplet (the blue circle) and generates the bulge (the solid black line), which is approximated as an oblate spheroid (the dashed black line). The normalized equator radius x e is defined as the ratio of the spheroid radius at the equator ( a + Δ R ) and the radius of the original sphere a . (b) The integral F ( x e ) = 0 π a κ ¯ ( x e , θ ) Y 20 ( θ ) sin θ d θ (blue circles) as a function of the normalized equator radius x e . κ ¯ ( x e , θ ) is given in Eq. (12). The dimensionless constant Γ σ is extracted using Eq. (14) and least square fitting (dashed red line).
Fig. 4
Fig. 4 (a) The radial dependence of | E | 2 of a fundamental TE mode ( l = 257) in a spherical droplet with radius a = 50 μ m . We assume | E s u r f p e a k | = 1 and take θ = π / 2 and ϕ = 0 . (b) The angular dependence of | E | 2 for the WGM in (a). The value of | E | 2 is evaluated over the droplet surface, with ϕ = 0 . Due to our normalization scheme, the curve is also f l m ( θ ) . (c) The power flux of the WGM in (a) within the ϕ = 0 plane. Only the e ϕ component of the Poynting vector is shown. The white circle represents the droplet surface. (d) The value of Γ θ l m for the fundamental TE mode | l l in droplets with different radius. The sphere radii and WGM parameters are listed in Table 1. The Γ θ l m values (blue circles) are obtained numerically using Eq. (17) and simply connected together using the dashed line.
Fig. 5
Fig. 5 (a) Radiation pressure induced droplet deformation ( Δ R / a ) in droplets with different radii (red circles). The total power of the WGM that circulates along the droplet equator is fixed at 1 W. For comparison, the changes in refractive index ( Δ n ) due to the Kerr effect are shown in the same figure, which are estimated using Δ n χ ( 3 ) | E s u r f p e a k | 2 . The Kerr effects in water and in CS2 are represented as the blue and black diamonds, respectively. All points are connected by dashed lines. (b) The relative shift in WGM frequency induced by the radiation pressure of a single photon. Two different values are used for surface tension: σ = 72 m N / m (blue circles), and σ = 1 m N / m (red crosses). All points are connected by dashed lines.

Tables (1)

Tables Icon

Table 1 The Angular Mode Number l and Resonance Wavelength λ of WGMs in Liquid Droplets.

Equations (23)

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

E = A ( x , z ) e i ( ω t k x ) e ^ y ,
n e f f ( x , z ) = n 0 + ( n h ) Δ h ( x , z ) .
2 E + n e f f 2 ( x , z ) ω 2 c 2 E = 0.
2 i k A x + 2 A z 2 = 2 n 0 ω 2 c 2 ( n h ) Δ h ( x , z ) A ( x , z ) .
σ d 2 Δ h ( z ) d z 2 + Δ ρ g Δ h ( z ) = 1 2 ε 0 ( n 2 2 n 1 2 ) n 2 2 n 1 2 | A ( x , z ) | 2 ,
2 i k A x + 2 A z 2 = χ e f f ω 2 c 2 | A | 2 A ,
w h e r e χ e f f = ( n h ) ε 0 n 0 ( n 2 2 n 1 2 ) n 2 2 Δ ρ g n 1 2 .
A ( x , z ) = A 0 sech ( z z 0 ) e i γ x ,
w h e r e z 0 = n 0 k | A 0 | 2 χ e f f , γ = k χ e f f | A 0 | 2 4 n 0 2 .
Δ P + P o p t = 2 σ κ ¯ ,
E l m T E = g l ( k q r ) X l m ( θ , ϕ ) e i ω t ,
H l m T E = i k q Z q × [ g l ( k q r ) X l m ( θ , ϕ ) ] e i ω t ,
g l ( k q r ) = { A c o j l ( k c o r ) ,       r < a A c l h l ( 1 ) ( k c l r ) ,     r > a
p o p t = 1 2 ε 0 ( n c o 2 n c l 2 ) | E s u r f | 2 ,
κ ¯ ( x e , θ ) = 1 2 a x p x e 2 x e 2 + ( x p 2 x e 2 ) sin 2 θ [ x e 2 + ( x p 2 x e 2 ) sin 2 θ ] 3 / 2 ,
2 Δ p + 0 π p o p t ( θ ) sin θ d θ = 2 σ 0 π κ ¯ ( x e , θ ) sin θ d θ ,
0 π p o p t ( θ ) Y 20 ( θ ) sin θ d θ = 2 σ 0 π κ ¯ ( x e , θ ) Y 20 ( θ ) sin θ d θ ,
F ( x e ) = 0 π a κ ¯ ( x e , θ ) Y 20 ( θ ) sin θ d θ Γ σ ( x e 1 ) .
| E s u r f | 2 = | E s u r f p e a k | 2 f l m ( θ ) ,
0 π p o p t ( θ ) Y 20 ( θ ) sin θ d θ = 1 2 ε 0 ( n c o 2 n c l 2 ) | E s u r f p e a k | 2 0 π f l m ( θ ) Y 20 ( θ ) sin θ d θ .
Γ θ l m = n c o a λ 0 π f l m ( θ ) Y 20 ( θ ) sin θ d θ .
Δ R a = x e 1 = Γ θ l m ε 0 λ 4 Γ σ σ n c o ( n c o 2 n c l 2 ) | E s u r f p e a k | 2 .
Δ ω ω = 2 l 1 2 ( l + 1 ) Δ R a .
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.