We analytically consider the effect of meniscus curvature on heat transfer to laminar flow across structured surfaces. The surfaces considered are composed of ridges. Curvature of the menisci, which separates liquid in the Cassie state and gas trapped in cavities between the ridges, results from the pressure difference between the liquid and the gas. A boundary perturbation approach is used to develop expressions that account for the change in the temperature field in the limit of small curvature of a meniscus. The meniscus is considered adiabatic and a constant heat flux boundary condition is prescribed at the tips of the ridges in a semi-infinite and periodic domain. A solution for a constant temperature ridge is also presented using existing results from a mathematically equivalent hydrodynamic problem. We provide approximate expressions for the apparent thermal slip length as function of solid fraction over a range of small meniscus protrusion angles. Numerical results show good agreement with the perturbation results for protrusion angles up to ± 20 deg.

Introduction

Micro- and nanostructured surfaces are engineered to reduce drag by using a combination of roughness and hydrophobic coatings. They are being considered for microchannel flows where surface friction forces dominate in such applications as electronics cooling [1] and lab-on-a-chip technologies [2]. A meniscus separates liquid in the Cassie (unwetted) state on a micro- or nanostructured surface from gas trapped in the cavities. Meniscus curvature is often considered a secondary effect and neglected in the analysis of flow and heat transfer along these surfaces. Recent work has shown that the effect of curvature on hydrodynamic slip length is not necessarily small and depends on the geometry of the structures, solid fraction,2 and the magnitude and direction of the protrusion angle. We use a boundary perturbation approach to assess the change to apparent thermal slip length across ridge-type structures due to meniscus curvature.

A composite interface, comprised of solid–liquid and liquid–gas regions, exists along the structured surface which results in a mixed-boundary condition on the flow. The no-slip boundary condition applies at the solid–liquid interface where the primary heat transfer occurs. At the liquid–gas interface, i.e., the meniscus, the flow boundary condition is low shear and is often assumed shear-free. The thermal boundary condition is low heat flux and is often assumed adiabatic.

Slip length expressions may be used to model the velocity and temperature discontinuities at the surface. When considering channel flow where the height of the channel is large compared to the length scale of the structures, the hydrodynamic and thermal boundary conditions on structured surfaces are treated similarly to those in the molecular slip regime of rarefied gas flows using Navier's [3] and Poisson's3 slip conditions, respectively. These provide the means to solve the outer flow and temperature problems to determine key engineering parameters, Poiseuille and Nusselt numbers, as a function of slip lengths [57]. For the case of a flat boundary, the thermal boundary condition is
(1)

where T is the temperature in the liquid, T¯sl is the mean temperature of the solid–liquid interface, T¯ĉ is the mean temperature at the composite interface, bt is the apparent thermal slip length, henceforth referred to as thermal slip length, and n is the direction normal to the composite interface pointing into the liquid flow. The left-hand side of Eq. (1) is the temperature jump (or discontinuity) which results from the added thermal resistance at the composite boundary relative to that of an unstructured surface. Thermal slip length expressions are available in the literature. Enright et al. [6] analytically developed them for three structure geometries: parallel ridges, perpendicular ridges, and pillar arrays. They assumed a flat, adiabatic meniscus and considered isoflux and isothermal boundary conditions at the solid–liquid interface. Ng and Wang [8] semi-analytically developed slip length expressions for ridges which account for heat transfer across a flat meniscus. They considered an isothermal boundary condition. They found that conduction through the gas has a negligible impact on slip length when the cavity depth is on the order of the spacing of the structures and larger, i.e., the adiabatic condition at the meniscus applies. They also considered pillar and hole arrays. The case of evaporation and condensation across the meniscus has been considered by Hodes et al. [9], and a slip length expression is provided for isoflux ridges as a function of solid fraction and interfacial heat transfer coefficient.

The thermal slip length is found from the solution to the inner problem, near the structures. Heat transfer is governed by the Péclet number, Pea, based on the length scale of the structures, a. When it is sufficiently small, the transport is considered purely diffusive. When the domain of the inner problem can be treated as semi-infinite, the temperature field becomes uniform at a vertical distance that is on the order of the pitch of the structures. For the case of the isothermal ridge, the temperature jump is mathematically equivalent to a hydrodynamic slip velocity and produces a change in thermal energy added to the channel flow. For the isoflux case, the temperature jump is captured in the change of the mean temperature at the solid–liquid interface relative to the no-slip case. In both cases, the thermal slip length captures not only the spreading resistance but also the change to the one-dimensional resistance that results from the presence of the irregular boundary.

Previous Work

The effect of hydrodynamic slip on drag reduction across structured surfaces is well documented in the literature [1013]. Heat transfer to such flows has only recently received attention [1,6,14,15].

A number of studies have shown that the shape of the meniscus and the degree of protrusion significantly affect drag. Ybert et al. [16] developed scaling laws for hydrodynamic slip. They derive expressions for secondary effects such as a finite shear force at the meniscus and pressure-induced curvature of the meniscus. They assumed a meniscus with a positive protrusion angle, i.e., the pressure in the gas is greater than that in the liquid. In the limit of large solid fraction, the correction to slip length is negatively proportional to the cavity fraction and the height of the meniscus. In the limit of small solid fraction, the correction to slip length varies inversely with curvature. They note that geometries with larger slip lengths are more sensitive to curvature effects. Bocquet and Barrat [17] discuss the analogy between thermal slip and hydrodynamic slip. We note that these scaling laws are also applicable to thermal slip in the presence of curvature.

Sbragaglia and Prosperetti [18] used a perturbation approach to study the effect of small curvature of the meniscus on hydrodynamic slip for parallel ridges. They identify two significant effects on the flow that result. The first is the change in the cross-sectional area of the flow due to the deformation of the meniscus, and the second is the change in the velocity field. They provided hydrodynamic slip length expressions for parallel ridges for a pressure-driven flow with a finite channel height and for a shear flow with an infinite channel height.

Several analytical studies have addressed the effect of meniscus curvature in the limit of large solid fraction for domains with semi-infinite heights. Crowdy [19] used a series of conformal maps to model a meniscus between parallel ridges. He derived an expression for slip length as a function of solid fraction and protrusion angle which is applicable for a periodic distribution of ridges in the limit of large solid fractions. He extended his work in Ref. [20] developing additional terms for his slip length expression such that it now applies for solid fractions down to ϕ=0.1. Davis and Lauga [21] used a two-dimensional model of shear flow past an array of bubbles trapped between transverse ridges and provided an expression to calculate the critical protrusion angle at which the slip length becomes negative. Their results also show the asymmetry between the effects of negative and positive protrusion angles. Recently, Crowdy [22] used a new transform scheme to model fluid flow over bubbles with partial slip at the meniscus. He provides a semi-analytical slip length expression for a meniscus with a positive protrusion angle of 90 deg. For solid fractions down to ϕ=0.1, results are within 10% of numerical data.

Teo and Khoo performed numerical studies of the effect of meniscus curvature on drag for menisci formed between parallel ridges for both Poiseuille and Couette flows [23] and in Poiseuille flow over transverse ridges [24]. For transverse ridges, they report a critical protrusion angle θc6265deg at which the slip length becomes zero that is independent of shear-free fraction, geometry, and flow type but which decreases as channel height is reduced for a given shear-free fraction. Above θc, friction was enhanced. For parallel ridges, slip length displayed asymmetry with respect to positive and negative protrusion angles; however, there was no critical angle at which drag reduction became zero. Instead, the slip length exhibited a monotonic increase with protrusion angle even for large angle values up to 90deg. They also found that the ratio of parallel ridge slip length to transverse ridge slip length did not remain constant as protrusion angle was varied.

To the authors' knowledge, the effect of meniscus curvature on thermal slip has not been studied. This motivates the present work.

Problem Formulation

The geometry under consideration is that of a concave or convex meniscus formed between two ridges of a structured surface as shown in Fig. 1. The center-to-center spacing of the ridges is 2d, the width of the ridge is 2a, the width of the cavity is 2c, and c/d=1ϕ is the cavity fraction. The solid fraction, ϕ, is the ratio of the solid–liquid area to the projected total area of the composite interface in the horizontal plane. The temperature field is symmetric about x = 0 and x = d. Typical length scales for structure width and spacing are on the order of microns [2527]. The protrusion angle, α, is measured from the horizontal to the line tangent to the meniscus at the triple contact line per Fig. 1. A positive protrusion angle corresponds to the meniscus projecting upward into the liquid and is limited by the maximum receding contact angle, θR, where α=πθR. A negative protrusion angle corresponds to the meniscus extending downward into the cavity and is limited by the maximum advancing contact angle, θA, where α=θAπ/2. Thermophysical properties are assumed constant. We address the inner problem when the Péclet number, based on the ridge width, is very small and transport is diffusive. Evaporation and condensation [9] and thermocapillary stress along the meniscus [28] are not considered here.

Fig. 1
Schematic of a curved meniscus between the ridges. The protrusion angle, α, is that between the line tangent to the meniscus at the triple contact line and the horizontal. The pitch of the structures is 2d. The width of the cavity is 2c.
Fig. 1
Schematic of a curved meniscus between the ridges. The protrusion angle, α, is that between the line tangent to the meniscus at the triple contact line and the horizontal. The pitch of the structures is 2d. The width of the cavity is 2c.
Close modal
We consider a unit cell of width d and infinite height. The governing equation is the Laplace equation
(2)
Symmetry boundary conditions exist at the midpoint of the meniscus and the midpoint of the ridge such that
(3)
Constant heat flux leaves the domain as y
(4)
In Sec. 3.2, we consider a constant heat flux boundary condition at the solid–liquid interface
(5)
where qsl=aq/d. In Sec. 3.3, we present a solution from the literature for the case when a constant temperature boundary condition applies at the solid–liquid interface such that
(6)
The adiabatic condition along the meniscus is
(7)

where n̂ is the vector normal to the meniscus.

Identification of the Perturbation Parameter.

We follow the approach of Sbragaglia and Prosperetti [18] by approximating the shape of the meniscus. It is an arc of radius, R, which results from a pressure difference across the meniscus, Δplg, between the liquid above the composite interface and the gas in the cavities per the Laplace–Young equation
(8)
where σlg is the liquid–gas surface tension. We note that, for a pressure-driven flow, the pressure gradient in the channel causes R to change in the streamwise direction, which is not addressed here. A circular arc pinned at the corners of the ridges is of the form
(9)
Using the binomial approximation of the square roots results in
(10)
In the limit of a small protrusion angle, as would potentially be the case with water, the curvature of the meniscus is taken to be small and its shape is defined as
(11)
where ε is a small quantity [18]. It follows that
(12)
and
(13)
Per the Young–Laplace equation
(14)
Notably, ε is a small parameter which varies inversely with R and represents the degree of curvature of the meniscus. Using a force balance and geometry, Δplg can be expressed in terms of surface tension, solid fraction, and meniscus protrusion angle, α, as discussed in Ref. [1]; ε becomes
(15)

When α = 0, ε=0 corresponds to a flat meniscus. +ε and ε correspond to α and +α, respectively. For small values of α, ε is a small parameter which is used here to consider the change in the temperature field resulting from small deflections of the meniscus via a boundary perturbation approach. For example, water in the Cassie state with maximum contact angle of θA=110deg on a surface with solid fraction, ϕ=.01, ε normalized to the pattern width is 0.17.

Isoflux Ridge.

In this section, we solve for the temperature field with constant heat flux, qsl, supplied at the solid–liquid interface. The problem is rendered dimensionless by using T̃=Tk/(qsld),x̃=x/d,ỹ=y/d,c̃=c/d,η̃=η/d2, and ε̃=εd. The width of the analytical domain is unity per Fig. 2.

Fig. 2
Schematic of isoflux domain, a half-ridge and half-cavity of a parallel ridge structured surface with meniscus between the ridges. Dashed vertical lines indicate symmetry.
Fig. 2
Schematic of isoflux domain, a half-ridge and half-cavity of a parallel ridge structured surface with meniscus between the ridges. Dashed vertical lines indicate symmetry.
Close modal

Boundary Condition at the Meniscus.

Expressing the dimensionless isoflux temperature field as a Taylor series in ε̃, it takes the form
(16)

where T̃0 is the temperature profile assuming a flat adiabatic meniscus between the structures, and T̃1 is the change to the first order in ε̃ that would result from the deflection of the meniscus. For small ε̃, the effect of meniscus curvature on the temperature field is captured by T̃1.

Taking the normal vector as n̂=ε̃(dη̃/dx̃)î+ĵ, the adiabatic condition along the meniscus is
(17)
where η̃=c̃2x̃2. Then, using Taylor expansions of T̃/x̃ and T̃/ỹ about ỹ=0
(18)
(19)
Using Eq. (16) and the above, Eq. (17) becomes
(20)
Matching powers of ε̃, the meniscus boundary conditions for T̃0 and T̃1 become
(21)
and
(22)

respectively. Further discussion of the perturbation expansion is provided in Appendix  A.

Zeroth-Order Problem.

The temperature profile takes the following form:
(23)
The first term on the RHS is the temperature profile in the absence of a meniscus, and ũ is the change to the temperature field that results from the presence of an adiabatic, flat meniscus. In the context of spreading resistance, Mikic and Rohsenow [29] obtained the temperature profile for T̃0 to within an inconsequential constant as
(24)
Enright et al. [6] showed that the thermal slip length is analogous to a thermal spreading resistance and adapted the above to provide an expression for thermal slip length for a flat meniscus. The thermal slip length associated with T̃0 is
(25)

where thermal slip length is nondimensionalized by the pitch of the structures, i.e., b̃t=bt/(2d).

First-Order Problem.

The correction to the temperature field in the presence of deflection of the meniscus is ε̃T̃1. The governing equation for T̃1 is the Laplace equation. It is
(26)
subjected to Eq. (22) and the following
(27a)
(27b)
Symmetry boundary conditions apply at x̃=0 and x̃=1. The method of separation of variables is used to solve the system. Applying the homogeneous boundary conditions leads to
(28)
At y = 0, the boundary condition is composed of the boundary condition at the meniscus (Eq. (22)) and the boundary condition at the ridge (Eq. (27a)). These are applied to Eq. (28) and yield a set of dual series equations
(29)
(30)
where
(31)
Using orthogonality and combining Eqs. (29) and (30) results in
(32)

The coefficients, gm, are found from the above, and the expression is provided in Appendix  B.

The total thermal slip associated with an adiabatic meniscus with small curvature in a semi-infinite domain for the isoflux ridge is
(33)
The contribution to thermal slip length associated with the deflection of the meniscus is found from Eq. (1). Noting that T̃¯1,ĉ=0
(34)
We note that the perturbation domain, T̃1, is adiabatically connected to the unperturbed domain such that the denominator of Eq. (34) is two times the heat flux in the unperturbed domain, namely, 2ϕ. It follows that
(35)
which yields, to a first order in ε̃
(36)

We compute b̃t,1 for prescribed values of ϕ by setting n = 200 and m = 200 and increasing first n by 200 until no change is observed in the result to five digits, and then increasing m by 200 until no change is observed in the result to five digits. For example, ϕ=0.01 convergence was achieved when gm was truncated at m = 3200 and n = 2600 terms. Further details of the application of the boundary perturbation method to compute the thermal slip length are discussed in Appendix  C. The salient point developed therein is that when neglecting terms of o(ε̃), we may evaluate all quantities in the expression for thermal slip length at ỹ=0. The same applies to the case of an isothermal ridge discussed next.

Isothermal Ridge.

The dimensionless thermal slip length expression for a flat meniscus and isothermal ridge is provided by Enright et al. [6]
(37)

Sbragaglia and Prosperetti [18] developed an expression which can be numerically integrated to calculate the hydrodynamic slip length associated with small deflections of the meniscus in a diffusive domain with infinite height. We adapt it here to provide a thermal slip length expression accounting for meniscus curvature for the case of an isothermal ridge.

The contribution to thermal slip length associated with the deflection of the meniscus is found using Eq. (B5) from Ref. [18]
(38)
It follows that the total thermal slip length accounting for meniscus curvature is
(39)

Results

Thermal slip lengths versus solid fraction is plotted for flat menisci and for a range of positive and negative protrusion angles. The cases of an isoflux ridge and isothermal ridge are shown in Figs. 3 and 4, respectively.

Fig. 3
Dimensionless thermal slip length versus solid fraction, ϕ, for selected protrusion angles, α, for the case of an isoflux ridge
Fig. 3
Dimensionless thermal slip length versus solid fraction, ϕ, for selected protrusion angles, α, for the case of an isoflux ridge
Close modal
Fig. 4
Dimensionless thermal slip length versus solid fraction, ϕ, for selected protrusion angles, α, for the case of an isothermal ridge
Fig. 4
Dimensionless thermal slip length versus solid fraction, ϕ, for selected protrusion angles, α, for the case of an isothermal ridge
Close modal

The presence of a flat meniscus imposes a restriction on the heat flow between the solid and the liquid domain. A thermal spreading (or constriction) resistance results, which is captured in the slip length for the flat meniscus. As solid fraction is decreased, the constriction increases and the thermal resistance of the inner liquid domain increases resulting in increased thermal slip. In the limit as ϕ0, the resistance grows to resemble that of an adiabatic boundary at ỹ=0. In the limit as ϕ1, the resistance reverts to that of a domain with a solid surface at ỹ=0. As the shape of the meniscus is perturbed from the flat state, it produces additional changes in the thermal resistance of the liquid domain. The shape of the irregular boundary affects the one-dimensional thermal resistance in addition to the spreading resistance in the liquid domain. The meniscus protruding downward decreases the thermal resistance resulting in lower slip lengths and enhancement to heat transfer. Conversely, as the meniscus protrudes upward, the area for heat flow becomes more constricted and the thermal resistance is increased.

Figure 5 displays the isothermal and isoflux thermal slip lengths as a function of protrusion angle for selected solid fractions. Notably, for lower solid fractions, the meniscus protrusion angle has a strong effect on slip length, and thus on heat transfer for both cases. This highlights the importance of accounting for the effect of curvature on heat transfer at low solid fractions which are desirable for the achievement of drag reduction.

Fig. 5
Dimensionless thermal slip length versus protrusion angle, α, for selected solid fractions, ϕ. Solid lines correspond to an isoflux ridge. Dashed lines correspond to an isothermal ridge.
Fig. 5
Dimensionless thermal slip length versus protrusion angle, α, for selected solid fractions, ϕ. Solid lines correspond to an isoflux ridge. Dashed lines correspond to an isothermal ridge.
Close modal

Model Validation.

The partial differential equation (PDE) toolbox in matlab [30] was used to compute the temperature field in the liquid domain with the meniscus, TPDE. The change in temperature in the solution domain was compared to that of a domain with no meniscus, T1D, having the same height, width, thermal conductivity, and heat flux through it per
(40)

where w is the width of the domain. The quantity β is equivalent to a dimensionless thermal spreading/constriction resistance [31].

The Laplace equation with relevant boundary conditions was solved with the finite element solver in PDE toolbox. The flow domain was discretized on average with 440,000 elements. Slip length results from this model were first compared to the analytical slip length for the flat meniscus and to comparable fluent [32] simulations. The values from the PDE toolbox were found to match the known flat meniscus slip length to five decimal places; therefore, it was used to evaluate the analytical values in order to determine the range of angles and solid fractions over which the analytical results are valid. The mesh at each case was adapted and refined multiple times for both the matlab and the fluent simulations to ensure mesh independence.

Tables 1 and 2 compare the slip length values obtained from the perturbation method and Refs. [6] and [18] to β obtained from the PDE toolbox. We note that, insofar as the perturbation analysis to o(ε̃), the mean temperatures along ỹ=0 and ỹ are unchanged as per Eqs. (24) and (26). Thus, in Eq. (C8), the corresponding terms in the numerator, i.e., T̃¯0(x̃,0) and T̃1¯(x̃,0), may be replaced by those as ỹ. Thus, the slip may also be computed from Eq. (40) for β. However, once the magnitude of ϵ̃ is large enough that terms of o(ε̃) become relevant, T̃¯0(x̃,0) and T̃¯1(x̃,0) may not be replaced by those as ỹ, and thus, we do not expect b̃t and β to agree as borne out by our results. The same argument applies for the case of an isothermal ridge. In summary, β=b̃t+o(ε̃).

Table 1

Comparison of perturbation method values, b̃t, with numerical values, β, for the isoflux ridge case

Solid fractionProtrusion angle (deg)ε̃b̃tβPercent difference
0.01−200.1731.116771.19293−6.38
0.01−100.0881.235681.26721−2.49
0.0110−0.0881.480941.471710.63
0.0120−0.1731.599851.61557−0.97
0.05−200.1800.728940.76476−4.68
0.05−100.0910.786730.80128−1.82
0.0510−0.0910.905920.902230.41
0.0520−0.1800.963710.97311−0.97
0.1−200.1900.553350.57584−3.91
0.1−100.0960.589240.59840−1.53
0.1−50.0480.607680.61157−0.64
0.15−0.0480.644840.642650.34
0.110−0.0960.663280.661040.34
0.120−0.1900.699170.70507−0.84
Solid fractionProtrusion angle (deg)ε̃b̃tβPercent difference
0.01−200.1731.116771.19293−6.38
0.01−100.0881.235681.26721−2.49
0.0110−0.0881.480941.471710.63
0.0120−0.1731.599851.61557−0.97
0.05−200.1800.728940.76476−4.68
0.05−100.0910.786730.80128−1.82
0.0510−0.0910.905920.902230.41
0.0520−0.1800.963710.97311−0.97
0.1−200.1900.553350.57584−3.91
0.1−100.0960.589240.59840−1.53
0.1−50.0480.607680.61157−0.64
0.15−0.0480.644840.642650.34
0.110−0.0960.663280.661040.34
0.120−0.1900.699170.70507−0.84
Table 2

Comparison of perturbation method values, b̃t, with numerical values, β, for the isothermal ridge case

Solid fractionProtrusion angle (deg)ε̃b̃tβPercent difference
0.01−200.1731.113931.14665−2.85
0.01−100.0881.216431.22594−0.78
0.0110−0.0881.427851.44062−0.89
0.0120−0.1731.530341.58957−3.73
0.05−200.1800.706990.71899−1.67
0.05−100.0910.757820.7603−0.33
0.0510−0.0910.862670.87112−0.97
0.0520−0.1800.913490.94698−3.54
0.1−200.1900.520590.53100−1.96
0.1−100.0960.555010.55802−0.54
0.1−50.0480.572690.57349−0.14
0.15−0.0480.608320.60924−0.15
0.110−0.0960.626000.63000−0.63
0.120−0.1900.660410.67884−2.71
Solid fractionProtrusion angle (deg)ε̃b̃tβPercent difference
0.01−200.1731.113931.14665−2.85
0.01−100.0881.216431.22594−0.78
0.0110−0.0881.427851.44062−0.89
0.0120−0.1731.530341.58957−3.73
0.05−200.1800.706990.71899−1.67
0.05−100.0910.757820.7603−0.33
0.0510−0.0910.862670.87112−0.97
0.0520−0.1800.913490.94698−3.54
0.1−200.1900.520590.53100−1.96
0.1−100.0960.555010.55802−0.54
0.1−50.0480.572690.57349−0.14
0.15−0.0480.608320.60924−0.15
0.110−0.0960.626000.63000−0.63
0.120−0.1900.660410.67884−2.71

For the isothermal case, the perturbation results for negative angles are closer to β than those with corresponding positive angles. For the isoflux case, the perturbation results for positive angles show better agreement with the numerical values. As to be expected, the difference between b̃t and β increases with the absolute value of α commensurate with the degree of curvature of the meniscus. It is significant to note that the effect of meniscus curvature on heat transfer is not symmetrical for positive and negative protrusion angles. For a given solid fraction, positive angles produce somewhat larger changes in thermal resistance relative to their negative counterparts. As shown in Figs. 6 and 7, as |α| increases, ε̃ increases, and ε̃ is an indication of the magnitude of error to be expected.

Fig. 6
Dimensionless slip length versus solid fraction for the case of an isoflux ridge. Solid lines are perturbation method results, and Δ's are values for β from the numerical results from matlab.
Fig. 6
Dimensionless slip length versus solid fraction for the case of an isoflux ridge. Solid lines are perturbation method results, and Δ's are values for β from the numerical results from matlab.
Close modal
Fig. 7
Dimensionless slip length versus solid fraction for the case of an isothermal ridge. Solid lines are perturbation method results from Ref. [18], and Δ's are values for β from the numerical results from matlab.
Fig. 7
Dimensionless slip length versus solid fraction for the case of an isothermal ridge. Solid lines are perturbation method results from Ref. [18], and Δ's are values for β from the numerical results from matlab.
Close modal

For the isoflux case, the difference between the perturbation method, b̃t and β, is less than 6.5% for protrusion angles between ±20deg, and solid fractions below 0.1. For the isothermal case, for angles between ±20deg, and solid fractions below 0.1, values that are typical for water-based cooling, the difference is less than 3%.

Conclusions

Thermal slip length expressions have been developed for flow across ridges which account for the change in thermal resistance associated with the presence of a curved adiabatic meniscus. They are applicable for a range of small protrusion angles and solid fractions which are typical of water-based cooling in a microchannel. When liquid pressure is higher than that of the gas, a negative protrusion angle exists and heat transfer is enhanced. Conversely, the presence of menisci protruding into the liquid phase because the pressure in the gas is higher than that of the liquid will reduce heat transfer to the liquid at low protrusion angles.

Two effects are important to note. First, at lower solid fractions, the meniscus protrusion angle has a significant effect on slip length, and thus on heat transfer for both cases. Second, the effect of meniscus protrusion on heat transfer is not symmetrical for positive and negative protrusion angles.

Future work will develop thermal slip lengths over a wider range of protrusion angles and solid fractions. The slip lengths presented here are applicable for use in Nusselt number expressions for channels where curvature of the meniscus is constant along the channel such as in Ref. [7]; however, further work must be undertaken to incorporate them into Nusselt number expressions for channels where curvature changes along the channel as in a Poiseuille flow.

Acknowledgment

The authors gratefully acknowledge Durwood Marshall and Shawn Doughty of Tufts Technology Services for their assistance and access to Tufts UIT Research Computing facilities. Funding for L.L., M.H., and G.K. was provided in part by the NSF CBET Award No. 1402783. T.K. was supported by an EPSRC doctoral scholarship. The work of S.M. was partially supported by an NSERC Discovery Grant

Appendix A: Full Perturbation Expansion

The Taylor expansion of Eq. (17) (previously used in Ref. [18]) is problematic when ỹ=0 falls outside of the physical domain (for ε̃<0). A more rigorous treatment of the perturbation expansion in Eq. (16) is possible through a domain transformation, taking Ỹ=ỹ+ε̃η̃(x̃) and writing
(A1)

where we extend η̃(x̃) to the solid–liquid interface by taking its value to be 0 for c̃x̃1. Note that, with this definition, the domains of T̃0 and T̃1 are naturally for 0<x̃<1 and 0<Ỹ, since points (x̃,ε̃η̃(x̃)) on the meniscus are mapped to points on the x̃ -axis by this definition. Note that derivatives of T̃(x̃,Ỹ) with respect to x̃ must now be evaluated via the chain rule, but that the boundary condition along the meniscus can be imposed directly as a boundary condition along Ỹ=0 for T̃0 and T̃1.

Expanding all terms, we recover the same zeroth-order problem as discussed in Sec. 3.2.2, so T̃0(x̃,Ỹ)=T̃0(x̃,Ỹ) as defined in Eq. (24); however, there are two key changes to the first-order problem. First, the boundary condition along the meniscus simplifies to
(A2)
Second, the governing PDE no longer satisfies the homogeneous Laplace's equation in Eq. (26), but a nonhomogeneous form, given by
(A3)
Writing T̃1(x̃,Ỹ)=T̃1(x̃,Ỹ)η̃(x̃)(T̃0/ỹ)(x̃,Ỹ), we find that T̃1(x̃,Ỹ) satisfies both Laplace's equation and the boundary condition given in Eq. (22), so that its solution is, indeed, T̃1 as given in Eq. (28), with coefficients as given in Eq. (32). Considering, then, the perturbation in the thermal slip length defined in Eq. (34), the only changes from that given in Eq. (35) come from the difference
(A4)

Note that this difference is, in fact, zero along the boundary, since (T̃0/ỹ)(x̃,0)=0 for 0<x̃<c̃, by Eq. (21), and η̃(x̃)=0 for c̃x̃1 as defined above. Thus, there is no change to the numerator in Eq. (34), while the denominator remains the same due to the boundary condition as ỹ. In summary, the full perturbation expansion as given by Eq. (A1) generates a distinct first-order problem from when the perturbation is applied only to the meniscus boundary condition, but this does not change the expressions presented above for the thermal slip length.

Appendix B: Coefficients gm

(B1)

Appendix C: Boundary Perturbation to Compute Thermal Slip Length

The dimensionless form of Eq. (1) is
(C1)
where ĉ denotes the composite interface. Defining s̃ as a coordinate (normalized to d̃) along the composite interface from the center of the (curved) meniscus (s̃=0) to the center of the solid–liquid interface, it becomes
(C2)
A general expression for η̃(x̃) that applies along the entire length of the composite interface is η̃(x̃)=H(1ϕx̃)(c̃2x̃2), where H denotes the Heaviside step function. Utilizing the formula for arc length to integrate along x̃ rather than s̃ and noting that ĉds̃=1+o(ε̃), Eq. (C1) becomes
(C3)
Expanding the terms in brackets as per the binomial theorem and noting that, since the meniscus is adiabatic, we need only evaluate the integral in the denominator from 1ϕ to 1
(C4)
Expanding the terms evaluated at (x̃,ε̃η̃) in the numerator such that we may evaluate them at (x̃,0) and recalling that η̃=0 along the solid–liquid interface when evaluating the integral in the denominator, we find that
(C5)
Too, as per the expansion T̃=T̃0+ε̃T̃1+o(ε̃), Eq. (C5) becomes
(C6)
We note that the term ε̃η̃(T̃0/ỹ)|x̃,0 in the numerator equals zero because the partial derivative equals 0 for 0x̃1ϕ and η̃=0 over the solid–liquid interface. Since the perturbation domain adds no net heat flux to the problem, 1ϕ1(T̃1/ỹ)(x̃,0)dx̃=0 and can be eliminated. Thus, ignoring terms of o(ε̃)
(C7)
or using overbars to denote mean quantities
(C8)

Thus, when neglecting terms of o(ε̃), we have shown that we may evaluate all quantities in the expression for slip length at ỹ=0. We note this analysis applies to both isoflux and isothermal solid–liquid interfaces.

2

Ratio of ridge tip area to total projected area of the surface.

3

According to Kennard [4].

References

1.
Steigerwalt Lam
,
L.
,
Hodes
,
M.
, and
Enright
,
R.
,
2015
, “
Analysis of Galinstan-Based Microgap Cooling Enhancement Using Structured Surfaces
,”
ASME J. Heat Transfer
,
137
(
9
), p.
091003
.
2.
Zhao
,
Y.
,
Liu
,
Y.
,
Xu
,
Q.
,
Barahman
,
M.
, and
Lyons
,
A. M.
,
2015
, “
Catalytic, Self-Cleaning Surface With Stable Superhydrophobic Properties: Printed Polydimethylsiloxane (PDMS) Arrays Embedded With TiO2 Nanoparticles
,”
ACS Appl. Mater. Interfaces
,
7
(
4
), pp.
2632
2640
.
3.
Navier
,
C.
,
1823
, “
Mémoire sur les du mouvement des fluids
,”
Mém. l'Académie R. Sci. l'Institut Fr.
,
6
, pp.
389
440
.
4.
Kennard
,
E. H.
,
1938
,
Kinetic Theory of Gases
, 1st ed.,
McGraw-Hill
,
New York
.
5.
Inman
,
R. M.
,
1964
, “
Laminar Slip Flow Heat Transfer in a Parallel-Plate Channel or a Round Tube With Uniform Wall Heating
,” NASA, Washington, DC, Technical Note D-2393.
6.
Enright
,
R.
,
Hodes
,
M.
,
Salamon
,
T.
, and
Muzychka
,
Y.
,
2014
, “
Isoflux Nusselt Number and Slip Length Formulae for Superhydrophobic Microchannels
,”
ASME J. Heat Transfer
,
136
(
1
), p.
012402
.
7.
Steigerwalt Lam
,
L.
,
Melnick
,
C.
,
Hodes
,
M.
,
Ziskind
,
G.
, and
Enright
,
R.
,
2014
, “
Nusselt Numbers for Thermally Developing Couette Flow With Hydrodynamic and Thermal Slip
,”
ASME J. Heat Transfer
,
136
(
5
), p.
051703
.
8.
Ng
,
C.-O.
, and
Wang
,
C. Y.
,
2014
, “
Temperature Jump Coefficient for Superhydrophobic Surfaces
,”
ASME J. Heat Transfer
,
136
(
6
), p.
064501
.
9.
Hodes
,
M.
,
Steigerwalt Lam
,
L.
,
Cowley
,
A.
,
Enright
,
R.
, and
MacLachlan
,
S.
,
2015
, “
Effect of Evaporation and Condensation at Menisci on Apparent Thermal Slip
,”
ASME J. Heat Transfer
,
137
(
7
), p.
071502
.
10.
Davies
,
J.
,
Maynes
,
D.
,
Webb
,
B. W.
, and
Woolford
,
B.
,
2006
, “
Laminar Flow in a Microchannel With Superhydrophobic Walls Exhibiting Transverse Ribs
,”
Phys. Fluids
,
18
(
8
), p.
087110
.
11.
Maynes
,
D.
,
Jeffs
,
K.
,
Woolford
,
B.
, and
Webb
,
B. W.
,
2007
, “
Laminar Flow in a Microchannel With Hydrophobic Surface Patterned Microribs Oriented Parallel to the Flow Direction
,”
Phys. Fluids
,
19
(
9
), p.
093603
.
12.
Teo
,
C. J.
, and
Khoo
,
B. C.
,
2008
, “
Analysis of Stokes Flow in Microchannels With Superhydrophobic Surfaces Containing a Periodic Array of Micro-Grooves
,”
Microfluid. Nanofluid.
,
7
(
3
), pp.
353
382
.
13.
Cheng
,
Y.
,
Teo
,
C.
, and
Khoo
,
B.
,
2009
, “
Microchannel Flows With Superhydrophobic Surfaces: Effects of Reynolds Number and Pattern Width to Channel Height Ratio
,”
Phys. Fluids
,
21
(
12
), p.
122004
.
14.
Maynes
,
D.
, and
Crockett
,
J.
,
2014
, “
Apparent Temperature Jump and Thermal Transport in Channels With Streamwise Rib and Cavity Featured Superhydrophobic Walls at Constant Heat Flux
,”
ASME J. Heat Transfer
,
136
(
1
), p.
011701
.
15.
Cheng
,
Y.
,
Xu
,
J.
, and
Sui
,
Y.
,
2014
, “
Numerical Study on Drag Reduction and Heat Transfer Enhancement in Microchannels With Superhydrophobic Surfaces for Electronic Cooling
,”
Appl. Therm. Eng.
,
88
, pp.
1
11
.
16.
Ybert
,
C.
,
Barentin
,
C.
,
Cottin-Bizonne
,
C.
,
Joseph
,
P.
, and
Bocquet
,
L.
,
2007
, “
Achieving Large Slip With Superhydrophobic Surfaces: Scaling Laws for Generic Geometries
,”
Phys. Fluids
,
19
(
12
), p.
123601
.
17.
Bocquet
,
L.
, and
Barrat
,
J.-L.
,
2007
, “
Flow Boundary Conditions From Nano- to Micro-Scales
,”
Soft Matter
,
3
(
6
), pp.
685
693
.
18.
Sbragaglia
,
M.
, and
Prosperetti
,
A.
,
2007
, “
A Note on the Effective Slip Properties for Microchannel Flows With Ultrahydrophobic Surfaces
,”
Phys. Fluids
,
19
(
4
), p.
043603
.
19.
Crowdy
,
D.
,
2010
, “
Slip Length for Longitudinal Shear Flow Over a Dilute Periodic Mattress of Protruding Bubbles
,”
Phys. Fluids
,
22
(
12
), p.
121703
.
20.
Crowdy
,
D.
,
2016
, “
Analytical Formulae for Longitudinal Slip Lengths Over Unidirectional Superhydrophobic Surfaces With Curved Menisci
,”
J. Fluid Mech.
,
791
, p.
065507
.
21.
Davis
,
A. M. J.
, and
Lauga
,
E.
,
2009
, “
Geometric Transition in Friction for Flow Over a Bubble Mattress
,”
Phys. Fluids
,
21
(
1
), p.
011701
.
22.
Crowdy
,
D.
,
2016
, “
Effective Slip Lengths for Longitudinal Shear Flow Over Partial-Slip Circular Bubble Mattresses
,”
Fluid Dyn. Res.
,
47
(
6
), p.
065507
.
23.
Teo
,
C. J.
, and
Khoo
,
B. C.
,
2010
, “
Flow Past Superhydrophobic Surfaces Containing Longitudinal Grooves: Effects of Interface Curvature
,”
Microfluid. Nanofluid.
,
9
(
2–3
), pp.
499
511
.
24.
Teo
,
C. J.
, and
Khoo
,
B. C.
,
2014
, “
Effects of Interface Curvature on Poiseuille Flow Through Microchannels and Microtubes Containing Superhydrophobic Surfaces With Transverse Grooves and Ribs
,”
Microfluid. Nanofluid.
,
17
(
5
), pp.
891
905
.
25.
Ou
,
J.
, and
Rothstein
,
J.
,
2005
, “
Direct Velocity Measurements of the Flow Past Drag-Reducing Ultrahydrophobic Surfaces
,”
Phys. Fluids
,
17
(
10
), p.
103606
.
26.
Woolford
,
B.
,
Prince
,
J.
,
Maynes
,
D.
, and
Webb
,
B. W.
,
2009
, “
Particle Image Velocimetry Characterization of Turbulent Channel Flow With Rib Patterned Superhydrophobic Walls
,”
Phys. Fluids
,
21
(
8
), p.
085106
.
27.
Schäffel
,
D.
,
Koynov
,
K.
,
Vollmer
,
D.
,
Butt
,
H.-J.
, and
Schönecker
,
C.
,
2016
, “
Local Flow Field and Slip Length of Superhydrophobic Surfaces
,”
Phys. Rev. Lett.
,
116
(
13
), p.
134501
.
28.
Hodes
,
M.
,
Karamanis
,
G.
,
Steigerwalt Lam
,
L.
,
Maclachlan
,
S.
,
Kirk
,
T.
, and
Papageorgiou
,
D.
,
2015
, “
Conformal Map and Asymptotic Solutions for Apparent Slip Lengths in the Presence of Thermocapillary Stress
,”
23rd National Heat and Mass Transfer Conference and 1st International ISHMT-ASTFE Heat and Mass Transfer Conference
, Thiruvananthapuram, India, Dec. 17–20, Paper No. IHMTC2015-1254.
29.
Mikic
,
B. B.
, and
Rohsenow
,
W. M.
,
1966
, “
Thermal Contact Resistance
,” Department of Mechanical Engineering, MIT, Cambridge, MA, Technical Report No. DSR-74542-41.
30.
MathWorks
,
2004
, “
Partial Differential Equation ToolboxTM User's Guide
,” MathWorks, Inc., Natick, MA.
31.
Yovanovich
,
M.
, and
Marotta
,
E.
,
2003
, “
Thermal Spreading and Contact Resistances
,”
Handbook of Heat Transfer
,
A.
Bejan
and
A. D.
Kraus
, eds.,
Wiley
,
Hoboken, NJ
, pp.
261
394
.
32.
ANSYS
,
2013
, “
ANSYS Fluent Theory Guide
,” ANSYS, Inc., Canonsburg, PA.