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

Opto-mechanical analysis of nonlinear elastomer membrane deformation under hydraulic pressure for variable-focus liquid-filled microlenses

Open Access Open Access

Abstract

Nonlinear large deformation of a transparent elastomer membrane under hydraulic pressure was analyzed to investigate its optical performance for a variable-focus liquid-filled membrane microlens. In most membrane microlenses, actuators control the hydraulic pressure of optical fluid so that the elastomer membrane together with the internal optical fluid changes its shape, which alters the light path of the microlens to adapt its optical power. A fluid-structure interaction simulation was performed to estimate the transient behavior of the microlens under the operation of electroactive polymer actuators, demonstrating that the viscosity of the optical fluid successfully stabilizes the fluctuations within a fairly short period of time during dynamic operations. Axisymmetric nonlinear plate theory was used to calculate the deformation profile of the membrane under hydrostatic pressure, with which optical characteristics of the membrane microlens were estimated. The effects of gravitation and viscoelastic behavior of the elastomer membrane on the optical performance of the membrane microlens were also evaluated with finite element analysis.

© 2014 Optical Society of America

1. Introduction

Miniaturized adaptive optical systems have attracted a great deal of attention due to their potential applications in consumer electronics and biomedical diagnostic systems. Variable-focus liquid-filled membrane microlenses have been developed by several researchers as promising candidates for miniaturized adaptive optical systems [113]. The application of hydraulic pressure on optical fluid makes a transparent elastomer membrane deform, which tunes the focal length of a membrane microlens. Unlike conventional solid-state focus-tunable lens systems, variable-focus liquid-filled membrane microlenses cause several design issues including response time, precision control and spherical aberration. Furthermore, for hand-held electronics applications, variable-focus microlenses are expected to be small and to have low driving voltage, low power consumption, high thermal stability and low fabrication cost. Therefore, actuators to produce hydraulic pressure in liquid-filled microlenses are a crucial aspect of opto-mechanical performance. Several actuators such as electromagnetic actuators [2,10,11], piezo-bending actuators [36], relaxor ferroelectric polymer actuators [7,14], thermal actuators [8], and dielectric elastomer actuators [13] were developed and integrated into microfluidic membrane lenses.

In our previous work [7], we designed and fabricated a varifocal liquid-filled microlens operated by four poly(vinylidene fluoride-trifluoroethylene-chlorotrifluoroethylene) [P(VDF-TrFE-CTFE)] polymer actuators [14]. The design and operation concept of the varifocal microlens developed by Choi et al. [7] is shown in Fig. 1. The main frame of the varifocal microlens is composed of three parts: a glass frame and two micromachined silicon frames. The lower silicon frame micromachined to have a lens hole and four actuator chambers was bonded to the glass frame. The top surface of the lower silicon frame was covered with transparent elastomer membrane so that the internal volume confined by the elastomer membrane, the lower silicon frame, and the glass frame is filled with optical fluid. Multilayered P(VDF-TrFE-CTFE) polymer actuators were fabricated on the upper silicon frame, and directly bonded onto the elastomer membrane on the lower silicon frame. Therefore, polymer actuators produce bending deformation under applied electric field, which pushes optical fluid in fluidic chambers into the lens part, which in turn changes the shape of the transparent elastomer membrane. Therefore, the shape change of the membrane tunes the focal length of the varifocal microlens.

 figure: Fig. 1

Fig. 1 The varifocal microlens developed by Choi et al. [7] as a typical example of variable-focus liquid-filled membrane microlenses, whose focal length varies with the deformation of a transparent elastomer membrane under hydraulic pressure tailored by four P(VDF-TrFE-CTFE) polymer actuators [14].

Download Full Size | PDF

For hand-held electronics applications, Choi et al. [7] and Pouydebasque et al. [9] designed and fabricated an array of varifocal liquid-filled microlenses on a wafer using microfabrication. This method is considered to be cost-effective for mass production and the fabricated varifocal microlenses are quite thin (less than 1 mm thick), which is desirable for mobile phone applications in particular. However, as the thickness of optical liquid confined in the microlens decreases, the response time of the liquid-filled microlens increases due to the viscosity of the optical fluid. Therefore, design of varifocal microlenses should be optimized to provide a fast response time with a small size.

Silicone elastomers such as polydimethylsiloxane (PDMS) are usually selected for optical membranes, since they have high optical transmittance and low Young’s modulus, and are capable of withstanding large deformation without permanent deformation. On the other hand, mechanical deformation of transparent optical membranes specifies the lens shape of liquid-filled microlenses, which results in several design issues with regard to optical performance such as effective aperture, dioptric power, optical transmittance, aberration, etc. First of all, the dynamic response of the membrane microlenses depends highly on the viscosity of the optical fluid and the viscoelastic properties of the optical membrane, in which the deformed shape of the membrane under hydraulic pressure is critical to the optical performance of the microlens.

In this paper, nonlinear large deformation of a transparent elastomer membrane under hydraulic pressure is analyzed and its effect on the optical performance of a variable-focus liquid-filled membrane microlens is investigated. The dynamic fluid-structure interaction (FSI) among the P(VDF-TrFE-CTFE) polymer actuators, optical fluid, and elastomer membrane in the liquid-filled varifocal microlens developed by Choi et al. [7] is analyzed in Sec. 2. The deformation of an elastomer membrane under hydrostatic pressure is analyzed by geometrically-nonlinear plate theory and its effects on the optical performance of the varifocal microlens are evaluated in Sec. 3. Using finite element analysis (FEA), Sec. 4 assesses the effect of gravity on spherical aberration of the varifocal microlens. Optical drift originating from the viscoelasticity of the elastomer membrane is analyzed with FEA in Sec. 5. Finally, Sec. 6 concludes this study.

2. Response time of microlenses: fluid-structure interaction analysis

The profile of a transparent elastomer membrane in liquid-filled membrane microlenses is controlled by hydraulic pressure generated by electroactive polymer actuators, and the dynamic deformation behavior of the membrane is crucial for opto-mechanical performance. For example, the shorter response time assures faster autofocus time of the microlenses. In this section, as a model problem, the dynamic fluid-structure interaction (FSI) among electroactive polymer actuators, optical fluid, and elastomer membrane in the liquid-filled varifocal microlens developed by Choi et al. [7] is analyzed with commercial software, COMSOL 3.3 [15]. As illustrated in Fig. 1, four electroactive polymer actuators produce bending deformation to push optical fluid through the microfluidic channel into the center portion. The hydraulic pressure of optical fluid in turn causes the shape of the transparent elastomer membrane to change. The analysis model consists of a transparent elastomer membrane, optical fluid contained in a volume of 6 × 6 × 0.3 mm3, and four poly(vinylidene fluoride-trifluoroethylene-chlorotrifluoroethylene) [P(VDF-TrFE-CTFE)] polymer actuators as the electroactive polymer actuators as shown in Fig. 2(a). The height of the microfluidic channel, the diameter of the membrane, and the thickness of the membrane are thus considered to be 0.3 mm, 2.4 mm, and 50 μm, respectively. Young’s modulus, Poisson’s ratio, density, and thickness of the P(VDF-TrFE-CTFE) polymer actuators are set to be 400 MPa, 0.48, and 1.879 g/cc, and 50 μm, respectively. In FSI analysis, piezoelectric deformation of the P(VDF-TrFE-CTFE) polymer actuators was simulated by applying uniform pressure on the outer surface of the actuator, given as a function of time as

pa=5π[tan1(2t6)+2](kPa),
which is also plotted in Fig. 2(b). The actuator pressure given in Eq. (1) was chosen to take into account the time delay of electromechanical response of the polymer actuator itself. A transparent elastomer membrane, polydimethylsiloxane (PDMS) is used, which has density, Young’s modulus, and Poisson’s ratio of 0.965 g/cc, 1.12 MPa, and 0.48, respectively [16]. The density of the optical fluid was assumed to be 1.80 g/cc, while the viscosity of the optical fluid varied from 1 cP to 1000 cP. Therefore, the FSI simulation was performed for various values of the viscosity of optical fluid. The transient response of the deflection at the center of the membrane is plotted in Fig. 2(b), showing that the varifocal microlens is a dynamically over-damped system due to the viscosity of the optical fluid, and thus, no fluctuation was observed. The steady-state deflection of the membrane was determined to be 47.8 μm. The response time of the microlens was defined to be the elapsed time for the membrane deflection to reach 98% of its steady-state value, 47.8 μm. The response time was plotted as a function of the viscosity of optical fluid in Fig. 2(c), showing that the response time is approximately linear with the viscosity of optical fluid. Figure 2(c) also shows that when the viscosity of optical fluid is 100 cP, the response time is 17 ms. Choi et al. [7] used Fomblin PFPE M03 manufactured by Solvay S. A. as the optical fluid, which has viscosity and refractive index of 54 cP at 20°C and 1.29, respectively. Therefore, we expect that the response time of the varifocal microlens is less than 17 ms, which agrees well with the experimental results reported by Choi et al. [7]. This level of response time is considered to be fast enough for autofocus function in mobile applications. It is also worth noting that if the height of the microfluidic channel increases, the response time decreases, which may be desirable for optical fluid with high viscosity.

 figure: Fig. 2

Fig. 2 Fluid-structure interaction analysis of a varifocal liquid-filled membrane microlens with commercial software COMSOL 3.3. (a) Simulation model, (b) transient response of deflection at the center of the varifocal microlens membrane, and (c) response time as a function of the viscosity of optical fluid.

Download Full Size | PDF

3. Optical performance of nonlinear elastomer membrane

In the previous section, it was observed that the varifocal liquid-filled microlens is a dynamically over-damped system such that no fluctuation of the elastomer membrane takes place during dynamic operation. We therefore infer that static plate theory can be used to analyze the deformation profile of the elastomer membrane, which is essential for its optical design. In this section, the elastomer membrane under hydrostatic pressure is analyzed by geometrically-nonlinear plate theory, which provides excellent estimation of its deformation profile unless any transient response within 20 ms is under consideration.

3.1 Deformation profile of an optical membrane under hydrostatic pressure

Let us assume that a uniformly loaded circular membrane with a clamped edge undergoes axisymmetric large deformation as depicted in Fig. 3(a). No surface tension, residual stress, or dynamic effect is considered here. Then, the governing equation becomes a system of nonlinear differential equations as follows [17]:

d2udr2+1rdudrur2=1ν2r(dwdr)2dwdrd2wdr2,
d3wdr3+1rd2wdr21r2dwdr=12t2dwdr[dudr+νur+12(dwdr)2]+pr2D,
where u=u(r) and w=w(r) are displacements in the r and z directions, respectively, t is the thickness of the membrane, and p is the uniform hydrostatic pressure acting on the membrane. The flexural rigidity of the axisymmetric membrane is expressed as D=Et3/12(1ν2), where E and ν are Young’s modulus and Poisson’s ratio, respectively. The boundary conditions of this problem can be expressed as

 figure: Fig. 3

Fig. 3 (a) Model geometry of an axisymmetric elastomer membrane with clamped edge loaded by uniform hydrostatic pressure. (b) Deformation profile of the elastomer membrane (Young’s modulus = 1.12 MPa, Poisson’s ratio = 0.48, thickness = 50 μm). (c) Maximum deflection at the center of the elastomer membrane as a function of the normalized pressure.

Download Full Size | PDF

u=dwdr=0,atr=0,
u=w=dwdr=0,atr=ra.

Details of the solution technique of governing Eqs. (2) and (3) with the boundary conditions given in Eqs. (4) and (5) are given by Timoshenko and Woinowsky-Krieger [17], and thus we only bring their results here for compactness. Equations (2) and (3) can be solved by assuming

w(r)=8t[C12(rt)2+C34(rt)4+C56(rt)6+],
Nr(r)=tE[B0+B2(rt)2+B4(rt)4+],
where Nr(r) is the force per unit length in the r direction. Governing Eqs. (2) and (3) together with boundary conditions given in Eqs. (4) and (5), are automatically satisfied if the unknown coefficients Bk (k=0,2,4,)and Ck (k=1,3,5,) satisfy the following recursive relations:
Bk=4k(k+2)m=1,3,5,k1CmCkm,k=2,4,6,
Ck=12(1ν2)k21m=0,2,4,k3BmCk2m,k=5,7,9,
C3=32(1ν2)(p42E+B0C1).
When the two constants B0 and C1 are assigned, all the other constants are determined by the above recursive relations. For given values of ν and p/E, B0 and C1 can be numerically searched with the conjugate gradient method.

Numerical results for the deformation profile of an elastomer membrane are shown in Fig. 3(b), for which a Young’s modulus, Poisson’s ratio, thickness, and radius of PDMS membrane of 1.12 MPa, 0.48, 50 μm, and 1,200 μm, respectively, were used. Near the clamped edge, the deformation profile largely deviates from a spherical shape due to the boundary constriction, which restricts a change of curvature in the membrane, and thus, forces the edge region to be discarded as an optical lens. Figure 3(c) displays the maximum deflection at the center of the optical membrane as a function of normalized pressure for several thicknesses of the optical membrane. As the thickness of the optical membrane increases, the maximum deflection at the center of the optical membrane becomes linearly proportional to the normalized pressure.

3.2 Optical performance of aspherically-deformed elastomer membrane

For optical analysis, the deformation profile of an elastomer membrane given in Eq. (6) can be converted into the following function:

w(r)=cr21+1(1+K)c2r2+k=2A2kr2k,
where c and K are the vertex curvature and conic constant of a profile, respectively. The types of standard symmetric conic surfaces, i.e., optical conicoids, can be determined by the conic constant, K: hyperboloid (K<1), paraboloid (K=1), sphere (K=0), and ellipsoid (1<K<0 or K>0). However, it is worth noting that when the surface profile of a conicoid, given as the first term in Eq. (11), is expanded in a power series, the convergence of the power series is not guaranteed [18]. The deformation profile given in Eq. (11) with vertex curvature c, conic constant K, and higher aspheric coefficients A2k determined in terms of Ck (k=1,3,) turns out to be valid only near the center, and not near the edge of the membrane. Therefore, without loss of generality, the conic constant is assumed to be K=1 (paraboloid), and thus, the vertex curvature and the higher aspheric coefficients are determined in terms of Ck (k=1,3,) as
c=8tC1,A2k=2C2k1kt2k1,(k=2,3,)
The vertex curvature of the deformation profile of an elastomer membrane as given in Eq. (12) is plotted in terms of normalized pressure p/E for various membrane thicknesses in Fig. 4. When t=50μm, the vertex curvature dramatically changes at the low pressure level (p/E< 1). As the thickness of the optical membrane increases, the vertex curvature of the optical membrane becomes linearly proportional to the normalized pressure. This tendency may suggest that the focal length of the varifocal microlens can be easily controlled for a thicker membrane.

 figure: Fig. 4

Fig. 4 Vertex curvature of the deformation profile of an elastomer membrane under hydrostatic pressure.

Download Full Size | PDF

The ray tracing simulation of the variable-focus liquid-filled microlens as shown in Fig. 1 was performed by using commercial software, Code V, in order to calculate the focal length and wave front error (WFE). The profile given in Eq. (11) with c and A2k (k=2,3,) given in Eq. (12) and with K=1 serves as the axisymmetric profile of the optical membrane made of PDMS (nPDMS = 1.38) in the optical analysis. Fomblin PFPE M03 manufactured by Solvay S. A. was used as the optical fluid, whose initial thickness and refractive index are tfluid = 300 μm and nfluid = 1.29, respectively [7]. The thickness and refractive index of the glass frame were assumed to be tglass = 300 μm and nglass = 1.52, respectively. The aperture stop of the microlens was assumed to be 2.0 or 1.6 mm in the optical simulation. The ray tracing simulation showed that the membrane deflection, 47.8 μm, obtained from the FSI analysis in Sec. 2 corresponds to the change of focal length from infinity to 38.3 mm (aperture stop = 2.0 mm) and 34.5 mm (aperture stop = 1.6 mm), representing the change of optical power by more than 25 diopters. The root-mean-square (RMS) WFE (WFERMS) was plotted as a function of focal length for various values of the membrane thickness in Fig. 5. As the focal length increases, the WFERMS gradually decreases. For a thicker optical membrane, the WFERMS is slightly greater. It is worth noting that the WFERMS for the 1.6 mm aperture stop is much smaller than that for the 2.0 mm aperture stop. This difference implies that the edge effect due to the bending stiffness tends to increase spherical aberration.

 figure: Fig. 5

Fig. 5 Wave front error of the varifocal liquid-filled microlens obtained by ray tracing simulation with commercial software, Code V.

Download Full Size | PDF

4. Spherical aberration due to gravity

The membrane shape of variable-focus liquid-filled microlenses is influenced by gravitational force, since the membrane surface may be positioned parallel to the gravitational force and the hydrostatic pressure varies along the direction of the gravitation force. The gravitational effect on the membrane deformation was simulated with a commercial finite element analysis (FEA) program, ABAQUS Version 6.12 [19], in which the optical membrane is located in the xy plane, and the gravitational force is assumed to act along the y direction. Therefore, the hydrostatic pressure is given as

p(x,y)=ρgy+p0,
where ρ=1.80g/cc and g=9.81m/s2 are the density of the optical fluid and acceleration of gravity, respectively. p0 is the uniform pressure applied by the P(VDF-TrFE-CTFE)] polymer actuators. Figure 6(a) shows a contour plot of the deformed shape of the elastomer membrane with a thickness of 50 μm under a hydraulic pressure p0 = 2 Pa. It can be inferred from Fig. 6(a) that the linear distribution of the hydrostatic pressure caused undulation of the elastomer membrane in the y direction. Figure 6(b) represents the deformation profile along the vertical center line (y axis) of the elastomer membrane for thicknesses of 50 and 70 μm for various values of applied pressure p0.

 figure: Fig. 6

Fig. 6 Finite element analysis of gravity effects on membrane deformation: (a) Contour plot of the deformed shape of the elastomer membrane under hydraulic pressure in the vertical position, and (b) the deformation profile along the vertical center line of the elastomer membrane.

Download Full Size | PDF

The deformed surface profiles of the elastomer membrane obtained above by FEA were used in the optical simulation with commercial software, Code V, with which the effect of gravity on the optical performance (focal length and WFE) of the variable-focus liquid-filled microlens as shown in Fig. 1 was investigated. The root-mean-square (RMS) WFE (WFERMS) was plotted as a function of focal length for various values of the membrane thickness in Fig. 7. It is worth noting that since a thinner optical membrane produces larger deflection under a given hydrostatic pressure as shown in Fig. 6(b), the WFERMS due to gravity becomes greater for a thinner optical membrane as can be observed in Fig. 7 for large focal length (greater than 50 mm). On the other hand, as the focal length decreases (in other words, as the deformation of the optical membrane increases), the WFERMS due to gravity decreases due to the geometrical nonlinearity of the optical membrane, while the WFERMS due to the aspherically-deformed membrane increases as shown in Figs. 5 and 7. Therefore, compared to the WFERMS due to gravity, the WFERMS due to the aspherically-deformed membrane becomes dominant for small focal length (smaller than 50 mm). When the WFERMS in Fig. 7 is compared with that in Fig. 5, the WFERMS due to gravity seems to be negligible for the optical membrane of thickness 90, 110, and 130 μm. However, the optical membrane of thickness 50 μm shows dramatic increase of the WFERMS in the range of 50-100 mm focal length, meaning that the optical membrane of thickness 50 μm should not be used without applied pressure p0 to prevent excessive WFE due to gravity. Therefore, it can be inferred that a thicker elastomer membrane reduces WFE due to the gravity effect. It is worth recalling that in order to reduce the WFE due to an aspherically-deformed membrane, the thickness of the elastomer membrane needs to be decreased. Therefore, the thickness of the elastomer membrane must be optimized to minimize the total WFE due to both the aspherically-deformed membrane and gravity.

 figure: Fig. 7

Fig. 7 Wave front error of the varifocal liquid-filled microlens due to gravity obtained by ray tracing simulation with Code V.

Download Full Size | PDF

In order to reduce spherical aberrations or WFE of a liquid-filled microlens, it is important to find the ranges of design parameters for which deformation of an elastomer membrane under hydrostatic pressure does not exceed a given limit. The hydrostatic pressure by the gravitational force can be given as ρgy and the coordinate y has the order of the membrane diameter, d. Therefore, the hydrostatic pressure due to gravitational force can be assumed to be pg=ρgd for a parametric study. Let us consider the deformation of a linear elastic circular membrane under a uniform pressure pg=ρgd, for which the maximum deflection at the center of the membrane can be given as [17]

δmax=3(1ν2)pgd416Et3=3(1ν2)ρgd516Et3.
Here, E, ν, and t are the Young’s modulus, Poisson’s ratio, and thickness of the membrane, respectively. Therefore, the shape deviation of a membrane due to gravitational force can be expressed as
δgravity=AP(1ν2)ρd5Et3,
where AP is a coefficient depending on the distribution of hydrostatic pressure and thus can be accurately evaluated by FEA. Equation (15) provides us a design guideline for liquid-filled microlenses: All the parameters, ρ,E,ν,t,andd in Eq. (15) must be selected to make a shape deviation δgravity less than a predetermined limit. It is worth noting that the diameter and thickness of the membrane are key parameters controlling the effects of gravity, since the shape deviation δgravity is proportional to the fifth power of the membrane diameter and inversely proportional to the third power of the membrane thickness.

5. Optical drift due to viscoelasticity of the elastomer membrane

Silicone elastomers like PDMS, typical optical membrane materials for liquid-filled microlenses, exhibit viscoelastic behavior. The viscoelasticity of silicone elastomers makes the opto-mechanical behavior of the membrane microlenses time-dependent. In particular, creep behavior (gradual change of deformation over a relatively long period of time under applied stresses) of the elastomer membrane may cause optical drift of liquid-filled membrane microlenses, for example, gradual change of focal length as a function of time. The viscoelastic behavior of an elastomer membrane was analyzed by a finite element method with a commercial program, ABAQUS Version 6.12 [19]. Let us assume that the uniformly loaded circular elastomer membrane with a clamped edge undergoes axisymmetric deformation as depicted in Fig. 3(a). However, the elastomer membrane is considered to be a viscoelastic PDMS, for which the relaxation modulus is given as [16]

μPDMS(t)=188.59+19.78et3.58+37.10et44.55+32.47et311.23+99.97et4908.22(kPa).
On the other hand, the Poisson’s ratio of the PDMS is assumed to be constant, that is, νPDMS = 0.48 [16], which is a typical assumption in viscoelasticity [20]. The uniform pressure was applied at t = 0 s, that is, p(t)=p0H(t), where H(t) is a Heaviside step function. Figure 8(a) represents the deformation profile of the PDMS membrane under the uniform pressure obtained by FEA. As time increases, the deformation of the PDMS membrane gradually increases. The maximum deflection at the center of the viscoelastic PDMS membrane is denoted by δmax(t), and the relative deflection at the center of the membrane is defined to be Δδmax(t)=δmax(t)δmax(0). If the normalized relative deflection, Δδmax(t)/Δδmax(10), at the center of the membrane is plotted as a function of time, the creep behavior of the membrane can be unified into a single curve as shown in Fig. 8(b).

 figure: Fig. 8

Fig. 8 (a) Spontaneous deformation profiles and creep behavior of a viscoelastic PDMS membrane obtained by FEA, and (b) normalized relative maximum deflection at the center of the viscoelastic PDMS membrane as a function of time.

Download Full Size | PDF

The creep behavior illustrated in Figs. 8(a) and 8(b) may cause optical drift of liquid-filled membrane microlenses. For example, the focal length of a microlens may gradually change due to the viscoelastic behavior of the viscoelastic optical membrane as a function of time, which must be compensated to provide continuous quality of imaging. Figure 9 shows the percent change of the focal length of the liquid-filled membrane microlens for 10 seconds due to the viscoelastic behavior of the PDMS membrane. For focal lengths greater than 150 mm, the percent change in focal length over 10 seconds is fairly uniform, while for focal length less than 150 mm, the percent change in focal length over 10 seconds dramatically decreases as the hydrostatic pressure increases and thus the focal length decreases. This observation implies that when relatively high pressure is applied, creep behavior of the optical membrane and thus the optical drift of the liquid-filled membrane microlens diminish due to the geometrical nonlinearity of the optical membrane.

 figure: Fig. 9

Fig. 9 Percent change of the focal length of the liquid-filled membrane microlens for 10 seconds due to the viscoelastic behavior of the PDMS membrane.

Download Full Size | PDF

6. Conclusion

Nonlinear large deformation of a transparent elastomer membrane in a variable-focus liquid-filled membrane microlens was analyzed to study its optical performance. In most membrane microlenses, actuators control hydraulic pressure of the optical fluid so that the elastomer membrane together with the internal optical fluid changes their shape, which alters the light path of the microlens to adapt its optical power. Fluid-structure interaction simulation demonstrated that the viscosity of the fluid successfully stabilizes the fluctuations during dynamic operations, and the lens changes the optical power by more than 25 diopters within 20 ms. Geometrically-nonlinear plate theory was used to calculate the deformation profile of the elastomer membrane under hydrostatic pressure, with which optical characteristics of the membrane microlens were estimated. In order to estimate the effect of gravity on the optical performance (focal length and WFE) of the variable-focus liquid-filled microlens, FE analysis and optical simulation were performed. It is worth noting that the diameter and thickness of the membrane are the key parameters to control the effect of gravity, since the shape deviation is proportional to the fifth power of the membrane diameter and inversely proportional to the third power of the membrane thickness. The effect of the viscoelastic behavior of the elastomer membrane on the optical performance of the membrane microlens was also evaluated with FEA. The percent change in focal length over 10 seconds was approximately 7%.

Acknowledgment

This work was supported by the 2012 Research Fund of the University of Ulsan.

References and links

1. N. Chronis, G. Liu, K. H. Jeong, and L. Lee, “Tunable liquid-filled microlens array integrated with microfluidic network,” Opt. Express 11(19), 2370–2378 (2003). [CrossRef]   [PubMed]  

2. S. W. Lee and S. S. Lee, “Focal tunable liquid lens integrated with an electromagnetic actuator,” Appl. Phys. Lett. 90(12), 121129 (2007). [CrossRef]  

3. F. Schneider, C. Müller, and U. Wallrabe, “A low cost adaptive silicone membrane lens,” J. Opt. A, Pure Appl. Opt. 10(4), 044002 (2008). [CrossRef]  

4. F. Schneider, J. Draheim, C. Müller, and U. Wallrabe, “Optimization of an adaptive PDMS-membrane lens with an integrated actuator,” Sens. Actuator A Phys. 154(2), 316–321 (2009). [CrossRef]  

5. J. Draheim, F. Schneider, R. Kamberger, C. Mueller, and U. Wallrabe, “Fabrication of a fluidic membrane lens system,” J. Micromech. Microeng. 19(9), 095013 (2009). [CrossRef]  

6. F. Schneider, J. Draheim, R. Kamberger, P. Waibel, and U. Wallrabe, “Optical characterization of adaptive fluidic silicone-membrane lenses,” Opt. Express 17(14), 11813–11821 (2009). [CrossRef]   [PubMed]  

7. S. T. Choi, J. Y. Lee, J. O. Kwon, S. Lee, and W. Kim, “Varifocal liquid-filled microlens operated by an electroactive polymer actuator,” Opt. Lett. 36(10), 1920–1922 (2011). [CrossRef]   [PubMed]  

8. W. Zhang, K. Aljasem, H. Zappe, and A. Seifert, “Completely integrated, thermo-pneumatically tunable microlens,” Opt. Express 19(3), 2347–2362 (2011). [CrossRef]   [PubMed]  

9. A. Pouydebasque, C. Bridoux, F. Jacquet, S. Moreau, E. Sage, D. Saint-Patrice, C. Bouvier, C. Kopp, G. Marchand, S. Bolis, N. Sillon, and E. Vigier-Blanc, “Varifocal liquid lenses with integrated actuator, high focusing power and low operating voltage fabricated on 200 mm wafers,” Sens. Actuator A Phys. 172(1), 280–286 (2011). [CrossRef]  

10. H. Choi, D. S. Han, and Y. H. Won, “Adaptive double-sided fluidic lens of polydimethylsiloxane membranes of matching thickness,” Opt. Lett. 36(23), 4701–4703 (2011). [CrossRef]   [PubMed]  

11. H. Choi, D. S. Han, and Y. H. Won, “Fluidic lens of PDMS membrane driven by voice-coil and magnet,” IEEE Photonics Technol. Lett. 24(19), 1683–1685 (2012). [CrossRef]  

12. J. K. Lee, K. Park, J. C. Choi, H. Kim, and S. H. Kong, “Design and fabrication of PMMA-micromachined fluid lens based on electromagnetic actuation on PMMA-PDMS bonded membrane,” J. Micromech. Microeng. 22(11), 115028 (2012). [CrossRef]  

13. S. Shian, R. M. Diebold, and D. R. Clarke, “Tunable lenses using transparent dielectric elastomer actuators,” Opt. Express 21(7), 8669–8676 (2013). [CrossRef]   [PubMed]  

14. S. T. Choi, J. O. Kwon, and F. Bauer, “Multilayered relaxor ferroelectric polymer actuators for low-voltage operation fabricated with an adhesion-mediated film transfer technique,” Sens. Actuators A Phys. 203, 282–290 (2013). [CrossRef]  

15. COMSOL, “COMSOL Multiphysics, Version 3.3” (2006).

16. S. T. Choi, S. J. Jeong, and Y. Y. Earmme, “Modified-creep experiment of an elastomer film on a rigid substrate using nanoindentation with a flat-ended cylindrical tip,” Scr. Mater. 58(3), 199–202 (2008). [CrossRef]  

17. S. Timoshenko, S. Woinowsky-Krieger, and S. Woinowsky, Theory of Plates and Shells (McGraw-Hill, 1959).

18. D. Malacara and Z. Malacara, Handbook of Optical Design (Marcel Dekker, 2004).

19. Dassault Systèmes Simulia Corp., “ABAQUS Version 6.12” (2013).

20. R. Christensen, Theory of Viscoelasticity: An Introduction (Elsevier, 1982).

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 The varifocal microlens developed by Choi et al. [7] as a typical example of variable-focus liquid-filled membrane microlenses, whose focal length varies with the deformation of a transparent elastomer membrane under hydraulic pressure tailored by four P(VDF-TrFE-CTFE) polymer actuators [14].
Fig. 2
Fig. 2 Fluid-structure interaction analysis of a varifocal liquid-filled membrane microlens with commercial software COMSOL 3.3. (a) Simulation model, (b) transient response of deflection at the center of the varifocal microlens membrane, and (c) response time as a function of the viscosity of optical fluid.
Fig. 3
Fig. 3 (a) Model geometry of an axisymmetric elastomer membrane with clamped edge loaded by uniform hydrostatic pressure. (b) Deformation profile of the elastomer membrane (Young’s modulus = 1.12 MPa, Poisson’s ratio = 0.48, thickness = 50 μm). (c) Maximum deflection at the center of the elastomer membrane as a function of the normalized pressure.
Fig. 4
Fig. 4 Vertex curvature of the deformation profile of an elastomer membrane under hydrostatic pressure.
Fig. 5
Fig. 5 Wave front error of the varifocal liquid-filled microlens obtained by ray tracing simulation with commercial software, Code V.
Fig. 6
Fig. 6 Finite element analysis of gravity effects on membrane deformation: (a) Contour plot of the deformed shape of the elastomer membrane under hydraulic pressure in the vertical position, and (b) the deformation profile along the vertical center line of the elastomer membrane.
Fig. 7
Fig. 7 Wave front error of the varifocal liquid-filled microlens due to gravity obtained by ray tracing simulation with Code V.
Fig. 8
Fig. 8 (a) Spontaneous deformation profiles and creep behavior of a viscoelastic PDMS membrane obtained by FEA, and (b) normalized relative maximum deflection at the center of the viscoelastic PDMS membrane as a function of time.
Fig. 9
Fig. 9 Percent change of the focal length of the liquid-filled membrane microlens for 10 seconds due to the viscoelastic behavior of the PDMS membrane.

Equations (16)

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

p a = 5 π [ tan 1 ( 2 t 6 ) + 2 ] ( kPa ) ,
d 2 u d r 2 + 1 r d u d r u r 2 = 1 ν 2 r ( d w d r ) 2 d w d r d 2 w d r 2 ,
d 3 w d r 3 + 1 r d 2 w d r 2 1 r 2 d w d r = 12 t 2 d w d r [ d u d r + ν u r + 1 2 ( d w d r ) 2 ] + p r 2 D ,
u = d w d r = 0 , at r = 0 ,
u = w = d w d r = 0 , at r = r a .
w ( r ) = 8 t [ C 1 2 ( r t ) 2 + C 3 4 ( r t ) 4 + C 5 6 ( r t ) 6 + ] ,
N r ( r ) = t E [ B 0 + B 2 ( r t ) 2 + B 4 ( r t ) 4 + ] ,
B k = 4 k ( k + 2 ) m = 1 , 3 , 5 , k 1 C m C k m , k = 2 , 4 , 6 ,
C k = 12 ( 1 ν 2 ) k 2 1 m = 0 , 2 , 4 , k 3 B m C k 2 m , k = 5 , 7 , 9 ,
C 3 = 3 2 ( 1 ν 2 ) ( p 4 2 E + B 0 C 1 ) .
w ( r ) = c r 2 1 + 1 ( 1 + K ) c 2 r 2 + k = 2 A 2 k r 2 k ,
c = 8 t C 1 , A 2 k = 2 C 2 k 1 k t 2 k 1 , ( k = 2 , 3 , )
p ( x , y ) = ρ g y + p 0 ,
δ max = 3 ( 1 ν 2 ) p g d 4 16 E t 3 = 3 ( 1 ν 2 ) ρ g d 5 16 E t 3 .
δ gravity = A P ( 1 ν 2 ) ρ d 5 E t 3 ,
μ PDMS ( t ) = 188.59 + 19.78 e t 3.58 + 37.10 e t 44.55 + 32.47 e t 311.23 + 99.97 e t 4908.22 (kPa) .
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.