throbber
J. Fluid Mech. (2008), vol. 595, pp. 141–161.
`doi:10.1017/S002211200700910X Printed in the United Kingdom
`
`c(cid:1) 2008 Cambridge University Press
`
`141
`
`Transition from squeezing to dripping in a
`microfluidic T-shaped junction
`
`M. D E M E N E C H1, P. G A R S T E C K I2, F. J O U S S E3
`A N D H. A. S T O N E4
`1Max–Planck Institute for the Physics of Complex Systems, N¨othnitzer Str. 38,
`01187, Dresden, Germany
`2Institute of Physical Chemistry, Polish Academy of Sciences, Kasprzaka 44/52,
`01-224, Warsaw, Poland
`3Unilever Corporate Research, Colworth House, Sharnbrook, Bedfordshire, MK44 1LQ, UK
`4School of Engineering and Applied Sciences, Harvard University, Cambridge, MA 02138, USA
`
`(Received 12 May 2006 and in revised form 5 September 2007)
`
`We describe the results of a numerical investigation of the dynamics of breakup of
`streams of immiscible fluids in the confined geometry of a microfluidic T-junction.
`We identify three distinct regimes of formation of droplets: squeezing, dripping and
`jetting, providing a unifying picture of emulsification processes typical for microfluidic
`systems. The squeezing mechanism of breakup is particular to microfluidic systems,
`since the physical confinement of the fluids has pronounced effects on the interfacial
`dynamics. In this regime, the breakup process is driven chiefly by the buildup of
`pressure upstream of an emerging droplet and both the dynamics of breakup and
`the scaling of the sizes of droplets are influenced only very weakly by the value
`of the capillary number. The dripping regime, while apparently homologous to the
`unbounded case, is also significantly influenced by the constrained geometry; these
`effects modify the scaling law for the size of the droplets derived from the balance
`of interfacial and viscous stresses. Finally, the jetting regime sets in only at very high
`flow rates, or with low interfacial tension, i.e. higher values of the capillary number,
`similar to the unbounded case.
`
`1. Introduction
`Because of the small size of the microchannels (widths of the order of 10 to 100 µm)
`−1), flows in microfluidic systems are generally dominated
`and typical flow rates (1 µl s
`by viscous effects. Two characteristics of microflows, i.e. the laminar flow and the
`typically large values of the P´eclet number (measuring the ratio of the convective
`to diffusive transport), allow for an extensive control both in space and time over
`the transport of chemical substances (Kenis, Ismagilov & Whitesides 1999; Stone,
`Strook & Ajdari 2004; Squires & Quake 2005). This control, in conjunction with the
`ease of fabrication of the microfluidic devices (Duffy et al. 1998; McDonald et al.
`2000), is one of the main features driving the interest in microfluidic systems for
`engineering and research applications.
`The laminar flow of a single Newtonian fluid at low Reynolds numbers, which is
`described by linear equations and boundary conditions, is to be contrasted with a wide
`class of nonlinear phenomena which have been uncovered with the first experiments
`on two-phase flows in microfluidic systems (Thorsen et al. 2001; Gan ´an-Calvo &
`
`Fluidigm Exhibit 2022
`Page 1 of 21
`
`1
`
`

`
`142
`
`M. De Menech, P. Garstecki, F. Jousse and H. A. Stone
`
`Application
`
`Reference
`
`Chemical processing
`Micromixing inside drops
`Drops and slugs as mixing elements
`High-throughput screening
`Kinetic analyses
`Organic chemistry
`Bioanalysis
`Diagnostic assays
`Handling and/or analysis of living cells
`Molecular evolution
`Material science
`Anisotropic particles
`
`Song et al. (2003)
`Gunther et al. (2004); Garstecki et al. (2005a)
`Zheng et al. (2003); Zheng & Ismagilov (2005)
`Song & Ismagilov (2003)
`Cygan et al. (2005)
`
`Sia et al. (2004)
`He et al. (2005)
`Cornish (2006)
`
`Jeong et al. (2004, 2005); Nisisako et al. (2004);
`Dendukuri et al. (2005); Xu et al. (2005)
`Microcapsules
`Takeuchi et al. (2005)
`Colloidal shells
`Subramaniam et al. (2005)
`Gunther & Jensen (2006); Song et al. (2006)
`Review articles
`Table 1. Examples of applications of multiphase flow in microfluidic devices.
`
`Gordillo 2001; Anna, Bontoux & Stone 2003; Dreyfus, Tabeling & Willaime 2003).
`The existence of an interface and the influence of interfacial tension introduce strong
`nonlinearities in the flow, which are responsible for the appearance of a range of
`novel effects, some of which are particular to microfluidic systems when the interfacial
`dynamics is strongly influenced by the confinement of the fluids by the walls of the
`channel (Garstecki et al. 2005b,c,d , 2006; Guillot & Colin 2005).
`The interest in detailed understanding of the emulsification processes in microfluidic
`systems is additionally motivated by the wide range of work on applications of
`microfluidic multiphase flows (see table 1). For example, microfluidics offers new
`routes and better control to chemistry, including chemistry inside small containers
`(e.g. Li, Zheng & Harris 2000; Song, Chen & Ismagilov 2006), control of dispersion
`(e.g. Pedersen & Horvath 1981), etc., approaches that have a long history in chemical
`engineering. Most of the applications require precise control over the process of
`formation of droplets or bubbles (e.g. Basaran 2002), and characterization or,
`preferably, understanding of the scaling laws that describe the volume of the bubbles
`or droplets formed in the devices as a function of the material (e.g. viscosities,
`interfacial tension) and flow parameters (e.g. pressures or rates-of-flow applied to the
`system). Understanding the flows is closely related to more classical studies of drop
`breakup and emulsification in sheared unbounded fluid systems (Rallison 1984; Stone
`1994). Nevertheless, as we shall discuss, the confinement that naturally accompanies
`flow in small devices has significant qualitative and quantitative effects on the drop
`dynamics and breakup (Garstecki et al. 2005d , 2006; Guillot & Colin 2005).
`Several methods of formation of both bubbles and droplets have already been
`described. Table 2 provides references to several recent experimental reports.
`Numerical simulations of breakup in microfluidic geometries have also been
`conducted, see for example studies of axisymmetric geometries by Jensen, Stone &
`Bruus (2006), Suryo & Basaran (2006) and Zhou, Yue & Feng (2006). For a thorough
`up-to-date review of drop formation in microfluidic devices see Christopher & Anna
`(2007). Here we report fully three-dimensional numerical simulations to understand
`and characterize drop formation in a microfluidic T-junction geometry (figure 1)
`which was first introduced for the controlled formation of water-in-oil dispersions by
`
`Fluidigm Exhibit 2022
`Page 2 of 21
`
`2
`
`

`
`Planar geometries
`Review article
`Planar geometries
`Flow-focusing
`
`Crossflow
`
`Diverging flow T-junction
`Axisymmetric geometries
`
`Transition from squeezing to dripping in a microfluidic T-shaped junction
`
`143
`
`Subject
`
`Reference
`
`Christopher & Anna (2007)
`
`Anna et al. (2003); Dreyfus et al. (2003); Cubaud & Ho (2004);
`Garstecki et al. (2004); Xu & Nakajima (2004);
`Ward et al. (2005)
`Blackmore et al. (2001); Thorsen et al. (2001); Song & Ismagilov
`(2003); Okushima et al. (2004); Zheng et al. (2003); Gerdts et al.
`(2004); Tice et al. (2004); Dendukuri et al. (2005);
`Guillot & Colin (2005); Garstecki et al. (2006)
`Link et al. (2004); Engl et al. (2005)
`Gan ´an-Calvo & Gordillo (2001); Jeong et al. (2004, 2005);
`Takeuchi et al. (2005); Utada et al. (2005)
`Sugiura et al. (2001, 2005)
`Nonplanar geometries
`Table 2. Examples of droplet and bubble formation in microfluidic devices. Given the rapid
`growth in the number of such two-phase-flow studies, the above references are representative
`of the kinds of studies that have been performed.
`
`Vd
`

`d
`
`Vc
`

`c
`
`L
`
`Figure 1. Diagram of a T-junction with crossflow. The channels have square cross-section
`with side L. vc and vd are the mean flow velocities of the continuous and dispersed phases,
`while µc and µd are the corresponding shear viscosities.
`
`Thorsen et al. (2001). The authors made the reasonable suggestion that the dynamics
`of droplet formation is dominated by the balance of tangential shear stresses and
`interfacial tension (i.e. the capillary number) as expected in unbounded shear flows, via
`an analogy to breakup processes in shear and extensional flows (Taylor 1934; Rallison
`1984; Stone 1994). A detailed experimental study of the T-junction configuration
`(Garstecki et al. 2006) identified a different (squeezing) mechanism that is directly
`connected to the confined geometry in which the drop is formed (see also Guillot &
`Colin 2005). It was proposed that when the capillary number is sufficiently small,
`the dominant contribution to the dynamics of breakup arises from the buildup of
`pressure upstream of the emerging droplet (Garstecki et al. 2005d , 2006). This model
`
`Fluidigm Exhibit 2022
`Page 3 of 21
`
`3
`
`

`
`144
`
`M. De Menech, P. Garstecki, F. Jousse and H. A. Stone
`
`results in a scaling law for the size of the droplets that is independent of the value
`of the capillary number and includes only the ratio of the rates of flow of the two
`immiscible fluids.
`The numerical results that we present here provide a unifying picture of the
`dynamics of formation of droplets in microfluidic T-junction geometries that includes
`both of the aforementioned squeezing (Garstecki et al. 2005d , 2006) and shear-driven
`(Thorsen et al. 2001) types of breakup; for another three-dimensional simulation
`of drop formation at a T-junction, see van der Graaf et al. (2006). We confirm
`the existence of the ‘rate-of-flow controlled’ or ‘squeezing’ breakup mechanism at
`low values of the capillary number Ca. We provide the details of the dynamics,
`which includes fluctuations of pressure upstream of the immiscible tip postulated
`by Garstecki et al. (2006). We identify a critical value of the capillary number
`at which the system transits into a shear-dominated or dripping mechanism of
`droplet formation. Also, we indicate the differences in drop formation in confined and
`unbounded systems. Finally, similarly to breakup into an unbounded fluid, we observe
`a transition from dripping to jetting at larger values of the capillary number. We
`note that although jetting refers to the formation of long threads prior to formation
`of a drop, which is usually associated with inertial effects of the internal phase, here
`we have a case where shear in the external phase drives a jetting transition at low to
`modest Reynolds numbers (e.g. Utada et al. 2007).
`In the following section, we define the geometry and the parameters of the system
`
`that we study and introduce the important dimensionless quantities. In § 3, we describe
`the numerical methods employed in our work. In § 4, we detail the results, both for the
`low- and high-capillary-number regimes, and we summarize our observations in § 5.
`
`2. Description of the system
`For a planar geometry, the characteristic dimensions of the T-shaped junction are
`the height h, and the widths of the main and side channels. We will consider the
`simplest case in which the widths of both ducts equal L, and the channels have a
`square cross-section h = L (figure 1). The dispersed phase is injected into the main
`channel from the side inlet. For simplicity, we set the densities of both phases equal;
`we expect that this choice has negligible influence on the results, since in most
`microfluidic configurations buoyancy-driven speeds are much smaller than the actual
`flow speeds. Besides the width L, which is constant, the problem is fully described
`by six parameters characterizing the flow and material properties of the fluids. These
`parameters are the mean speeds of the continuous and dispersed phases, vc and vd
`respectively, the viscosities of the two fluids µc and µd , the interfacial tension γ , and
`the density ρ. We will assume perfect wetting for the continuous phase, while the
`dispersed fluid does not wet the walls. The rescaled volume V = Vd /L3 of the droplets
`formed in the device is the eighth physical quantity, and following the Buckingham-Π
`theorem, it can be described as a function of four dimensionless parameters. We
`chose the following groups: the capillary number calculated for the continuous phase,
`Ca = µcvc/γ , the Reynolds number Re = ρvcL/µc, the viscosity ratio λ = µd /µc, and
`the flow rate ratio Q = vd /vc = Qd /Qc, where Qd = vd L2 and Qc = vcL2 are the flow
`rates at the two inlets. The pressure will be rescaled by the typical viscous shear stress
`µcvc/L, while the time unit is L/vc.
`For the flow regimes under consideration, the Reynolds number is small (Re < 1),
`and does not influence the droplet size, which leaves us with the three governing
`parameters: Ca, λ and Q. Our main focus is the discussion of the transition
`
`Fluidigm Exhibit 2022
`Page 4 of 21
`
`4
`
`

`
`Transition from squeezing to dripping in a microfluidic T-shaped junction
`
`145
`
`from the interfacial-tension dominated to the shear-dominated regime, which is best
`characterized by considering the effects of the capillary number on the droplet size.
`Within this framework, we will also consider the influence of λ and Q.
`
`3. The numerical model
`We use a phase-field model (De Menech 2006) to simulate numerically the flow of
`the two immiscible fluids at the T-junction. In common experimental configurations,
`low capillary numbers characterize the microfluidic flows, hence surface tension
`stresses are large in comparison to viscous stresses. These flows can be effectively
`tackled numerically with the diffuse interface method that we employ here, which
`models the phase boundary separating the two fluids as a diffuse region. The
`equilibrium properties of the mixture, including phase behaviour, wetting properties
`and the concentration profile in the interface region, are derived from a generalized
`free-energy functional, which also determines the diffusive and capillary forces in the
`transport equations. The transport equations are solved on a three-dimensional grid,
`and the method has been tested successfully in the case of droplet breakup in a
`microfluidic T-junction with diverging flow (De Menech 2006).
`
`3.1. Modelling multiphase flow
`The key issue in the numerical modelling of droplet formation and breakup in
`microfluidic devices is the requirement of a consistent and robust description of
`the effects related to the large interfacial tensions, which dominate over inertial
`and viscous stresses. As for the broader context of multiphase flow simulation,
`we can choose from the two main approaches: interface tracking and interface
`capturing methods (for an introductory review focused on microfluidics, see for
`example Cristini & Tan 2004).
`In interface tracking methods, the displacement of the boundary surface is followed
`explicitly by an adapting mesh, which represents the discrete approximation of the
`classical sharp interface limit. The main advantage of this approach is the accurate
`description of the interface, which comes at the expense of a greater complexity of the
`algorithms that are required to manage the motion and addition of mesh nodes. The
`handling of the singularities associated with surface merging and breakup represents
`a difficult technical problem, since the remeshing procedure has to cope with the
`changes in topology (see, e.g. Cristini, Balwzdziewicz & Loewenberg 1998; Wilkes,
`Phillips & Basaran 1999). In this framework, a few examples are available for the
`description of the dynamics of droplets pinned at or sliding on flat surfaces, both
`in two- and three-dimensional studies (Feng & Basaran 1994; Li & Pozrikidis 1996;
`Schleizer & Bonnecaze 1999).
`In interface capturing methods, the computational mesh remains fixed and the
`boundary discontinuities are smeared out over the finite width ξ of a diffuse interface,
`whose location is reconstructed from the gradients of an auxiliary scalar field. Surface
`stresses are included in the Navier–Stokes equation as volume forces that depend on
`the spatial derivatives of the auxiliary scalar field, such that the momentum transport
`equation remains consistent with the single-phase equation in the bulk regions. The
`lower bound to the manageable interface thickness ξ is determined by the mesh
`spacing, and the discontinuity of the sharp interface limit behaviour is recovered
`asymptotically as ξ becomes negligible with respect to the characteristic length of the
`flow L. Broadly speaking, we can distinguish two mainstream approaches in the class
`of interface capturing methods, based on the physical interpretation of the auxiliary
`
`Fluidigm Exhibit 2022
`Page 5 of 21
`
`5
`
`

`
`146
`
`M. De Menech, P. Garstecki, F. Jousse and H. A. Stone
`
`scalar field. In phase-field models (also called diffuse interface models; see Anderson,
`McFadden & Wheeler 1998), the behaviour of the scalar field in the transition region
`strictly follows from a variational principle applied to a generalized free-energy
`functional, which determines the equilibrium and non-equilibrium properties of the
`system. On the other hand, in the case of popular techniques such as the volume of
`fluid (VOF) method (Scardovelli & Zaleski 1999), or the level-set method (Osher &
`Fedkiw 2001), the scalar field (also known as a colour function in VOF methods) is
`merely an indicator of the different phases; its profile across the transition region has
`no meaning, other than serving to define the position of the phase boundary.
`Phase-field methods were developed to investigate phase transition phenomena, such
`as nucleation, evaporation and coarsening (Cahn 1965), where the diffuse interface
`provides a mean field description of the concentration or density profile. The phase-
`field version of the Navier–Stokes equations converges to the classical sharp interface
`behaviour as the interface thickness is reduced to zero along with the diffusivity
`(Jacqmin 1999). In the same manner, contact line dynamics can be related to that of
`immiscible fluids (Jacqmin 2000).
`In this paper, we will be using the phase-field model in which the capillary stresses
`at the interfaces are represented as volume forces, and are distributed over the
`characteristic thickness ξ of the diffuse phase boundary. This important feature of
`the model allows the treatment of multiphase flows with relatively coarse grids,
`since typically only a few mesh points are required in order to resolve the smooth
`variation of the order parameter across the interface. The interface thickness should
`be compared to a characteristic length of the system, which could be the radius of
`a droplet or the size of the domain (here L) in the case of confined flow, leading to
`the definition of the Cahn number C = ξ /L. The droplet dynamics of two immiscible
`
`fluids, described classically in the sharp interface limit ξ → 0, is exactly recovered as
`
`C and the diffusivity goes to zero (Jacqmin 1999).
`There are two possible ways to interpret the results of the phase-field model.
`The first one is to relateξ
`to the physical width of the interface for real fluids,
`which is of the order of 1 nm or larger. In this approach, our simulations should,
`within the continuum model, reflect the actual flow of fluids at the nano-scale,
`with typical channel widths of the order of 10 nm. Another interpretation is to
`consider the diffuse interface model as an approximation of the flow of immiscible
`fluids at the microscale, which is justified when the interface thickness ξ is small
`compared with the characteristic length of the flow L. Despite ξ not being the interface
`thickness separating real bulk fluids (e.g. since we typically choose C = 1/20 for
`
`numerical convenience, then ξ ∼ 5 µm when simulating drops in 100 µm channels), the
`
`experimentally observed droplet dynamics in microfluidic devices can be reproduced
`effectively (De Menech 2006).
`De Menech (2006) showed that the numerical model, despite the limitations related
`to the coarse description of the interface description, is capable of capturing the
`quantitative details of the dynamics of droplet breakup in a T-junction, in agreement
`with the experimental results for water droplets in a continuous oil phase. The results
`that we describe in this report, which are expressed in terms of the non-dimensional
`groups, also match closely the experimental observation of flow of droplets and
`bubble formation in microfluidic T-junctions (Garstecki et al. 2006).
`
`3.2. The phase-field model
`Phase-field models and a range of related numerical methods based on the generalized
`free-energy functional approach (lattice Boltzmann) have already been applied to
`
`Fluidigm Exhibit 2022
`Page 6 of 21
`
`6
`
`

`
`Transition from squeezing to dripping in a microfluidic T-shaped junction
`
`147
`
`the modelling of droplet dynamics in unbounded flow or in the presence of solid
`boundaries, including droplet deformation and breakup (Wagner & Yeomans 1997;
`Badalassi, Ceniceros & Banerjee 2003; Wagner, Wilson & Cates 2003; Kim 2005),
`droplet formation (Kuksenok et al. 2003), spreading on patterned surfaces (Dupuis &
`Yeomans 2004), and droplet formation in axially symmetric flow-focusing devices
`(Zhou et al. 2006). In this section, we summarize the main features of the phase-field
`model we used. A more complete description is given in De Menech (2006).
`We consider a mixture of two fluids, A and B, with the Cahn–Hilliard–van der
`Waals form for the free energy
`
`(cid:1)
`
`(cid:2)
`
`(cid:3)
`|∇ϕ|2 + nW (ϕ) + f (n)
`of component B, κ determines the interface thickness, ξ ∝ √
`tension, γ ∝ κ/ξ (De Menech 2006). In (3.1), f is the sum of the free-energy densities
`
`F [n, ϕ] =
`
`dr
`
`κn
`2
`
`,
`
`(3.1)
`
`where n = nA + nB is the total particle number density, ϕ = nB /n is the molar fraction
`κ, and the interfacial
`
`of the pure components, which for simplicity is assumed to depend only on the total
`density. Also, W is the free energy of mixing, having the characteristic double-well
`shape below the critical temperature, which therefore determines the coexistence of
`the A-rich and B-rich separate phases, with corresponding concentrations ϕA and ϕB.
`We will consider the incompressible limit, n = const, such that the equilibrium
`concentration satisfies the Euler–Lagrange equation
`(cid:7)
`
`µch (r) =− κ∇2ϕ(r) +W
`where µch (r) =δF /δϕ (r) is the chemical potential difference, and µcoex is the Lagrange
`multiplier which fixes the total molar fraction. In the framework of the Cahn–Hilliard
`theory of diffuse interfaces, the interaction of the fluid components with a wall is
`introduced by adding to the functional (3.1) a surface energy term (Cahn 1977), which
`fixes the relative affinity of the two components, A or B, for the solid boundaries. The
`surface energy therefore determines the wetting properties of the two phases, which
`can be expressed in terms of the equilibrium wetting angle. We choose the surface free
`energy such that the A component wets the walls, while the B component is repelled.
`In non-equilibrium conditions, local imbalances of the chemical potential µch (r)
`will generate diffusion currents which tend to restore the configuration satisfying (3.2).
`The advection–diffusion equation for the phase field is therefore
`
`(ϕ(r)) = µcoex ,
`
`(3.2)
`
`=−∇·(ϕ v) +Λ∇ 2[−κ∇2ϕ + W
`
`(cid:7)
`
`(ϕ)]
`
`∂ϕ
`
`∂t
`
`(3.3)
`
`where Λ is the mobility coefficient. Since the flow is considered incompressible,
`
`∇·v = 0. The expression for the pressure tensor, which includes the capillary forces
`
`generated by the presence of the interface, follows directly from the free-energy
`functional (3.1), based on the Gibbs–Duhem equation or on Noethers theorem
`(Anderson et al. 1998; De Menech 2006). The fluid momentum equation, besides the
`nonlinear advection term, contains both reactive and dissipative forces, depending,
`respectively, on the pressure tensor and the shear viscosity µ, which depends on ϕ.
`The Navier–Stokes equation is
`

`
`∂v
`∂t
`
`=−ρ∇· (vv) − ∇p +∇· (µ[∇v + (∇v)T ]) +
`
`∇· (∇ϕ∇ϕ),
`
`γ ξ
`
`Ω
`
`(3.4)
`
`Fluidigm Exhibit 2022
`Page 7 of 21
`
`7
`
`

`
`148
`
`M. De Menech, P. Garstecki, F. Jousse and H. A. Stone
`
`where the last term comes from the non-isotropic part of the pressure tensor, as
`first introduced by Korteweg, and Ω is a dimensionless number that depends on the
`choice of the mixing potential W (De Menech 2006).
`Finally, we remark that in the case of contact line dynamics, the results of a phase-
`field model match those of immiscible fluids in the far-field region, i.e. at a distance
`from the contact line which is larger than the characteristic interface thickness ξ
`(Jacqmin 2000). Similarly to the case of droplet breakup and interface merging, the
`dynamics of diffuse interface models in the near-field region have the significant
`advantage of removing the singularities that cause the breakdown of the sharp
`interface limit description of these phenomena. In our three-dimensional simulations,
`it is not strictly possible to identify a far-field region, since the pipe diameter is
`comparable with the interface width for the values of the Cahn number we have
`used. In other words, the adoption of the phase-field model for the description of two
`immiscible fluids is not justified a priori. On the other hand, as for the case of droplet
`breakup in a T-junction (De Menech 2006), we will show that the agreement with
`the experiments corroborates the validity of our approach, which in essence aims at
`classifying the variety of patterns observed in multiphase flows in microfluidic devices.
`
`3.3. The numerical method
`The transport equations (3.3) and (3.4) are discretized on a uniform three-dimensional
`Cartesian grid with staggered velocities; the molar fraction and pressure fields are
`collocated at the centres of the cubic cells, while the velocity components are placed
`on the faces, and the boundaries of the simulated domain always coincide with a
`face of a grid cell, be it a wall, an inlet or a pressure outlet. The time evolution is
`implemented with a fully implicit Euler scheme, which is first order in time (see De
`Menech 2006). The boundary condition (BC) for the molar fraction ϕ is of Neumann
`
`type at the outlet, n·∇ϕ = 0, where n is the vector normal to the outlet. A Dirichlet BC
`
`is set at the inlets, ϕ = ϕA at the main channel inlet and ϕ = ϕB at the side inlet. The
`BCs for the fluid velocity v are of Dirichlet and Neumann type at the inlets and outlet,
`respectively, with the inlet velocity profile set equal to the fully developed Poiseuille
`flow in a square pipe. The BCs for the pressure field p are of Neumann and Dirichlet
`type at the inlet and outlet, respectively. Finally, on all boundaries it is assumed that
`the chemical potential gradient is zero, so that locally there are no contributions to a
`flux of mole fraction. With respect to the actual discretization the typical grid spacing
` x/L was equal to C , which implies 20 grid points across the tube diameter. For the
`smallest domain constructed, with a tube length of 6L and a side inlet of length L,
`we have therefore 56 000 grid points in the three-dimensional domain.
`Given the stress tensor T in the Navier–Stokes equation (3.4), which is the sum
`
`of the viscous stress tensor µ[∇v + (∇v)T ] and the negative of the pressure tensor,
`P = pI + (γ ξ /Ω)∇ϕ∇ϕ, then at the outlet we haveT zz = Pzz = p = p0 (where z is normal
`
`to the boundary); at the outlet, the pressure is constant and equal to the isotropic
`part of the pressure tensor. This set of boundary conditions at the outlet is equivalent
`to the assumption that the flow is fully developed at the end of the tube, and that
`there is no net flux of the two components A and B (note that at the inlet there is
`still a net influx because of the Dirichlet BC).
`
`The constant pressure condition and the condition n·∇ϕ = 0 at the outlet deforms
`
`the shape of droplets which touch the outlet, since the Laplace pressure difference
`here is forced to be zero and the curved interface is rapidly stretched and stabilized to
`become perpendicular to the boundary. Very long jets which touch the outlet are, in
`fact, stabilized, with their shape remaining stationary. These effects cannot be avoided
`
`Fluidigm Exhibit 2022
`Page 8 of 21
`
`8
`
`

`
`Transition from squeezing to dripping in a microfluidic T-shaped junction
`
`149
`
`in a finite simulation domain. On the other hand, such effects play a role only when
`the droplet or the thread is very close to the boundary (a distance comparable to the
`interface thickness ξ ), and the pressure fluctuations associated with these events have
`very high frequency, and do not affect the slow dynamics of droplet formation inside
`the junction. These two facts still enable us to trust reliably the breakup dynamics
`observed far from the outlet, as for the squeezing and dripping regimes. We also
`performed a few tests to check that the size of droplets formed in the channel is not
`affected by changing the length of the pipe downstream from the T-junction.
`
`4. Results
`Before we present our results, we sketch the characteristics of the two main
`dynamical models for breakup in a microfluidic T-junction: shear-driven breakup
`(Thorsen et al. 2001), which is derived from the balance of shear and interfacial
`stresses, and the rate-of-flow-controlled breakup (Garstecki et al. 2005d , 2006), which
`was proposed for systems operating at low values of the capillary number and is
`based on the evolution of pressure in the continuous phase upstream of the tip of the
`dispersed phase.
`In shear-driven breakup, the volume of the droplet can be estimated from a balance
`of the viscous drag that the continuous fluid exerts on the emerging droplet and the
`interfacial force that opposes the elongation of the neck, which connects the reservoir
`of the dispersed fluid with the droplet (e.g. Umbanhowar, Prasad & Weitz 2000;
`Thorsen et al. 2001). This model leads to a relation in which the diameter of the
`droplet is inversely proportional to the capillary number calculated for the flow of the
`continuous liquid. Because the drag force exerted on the droplet depends only very
`weakly on the viscosity of the droplet, within the shear-driven regime, the viscosity
`of the dispersed phase does not influence the size of droplets appreciably (e.g. see
`discussion of flow past a drop in Batchelor 1967). This effect has been confirmed
`experimentally by Cramer, Fischer & Windhab (2004).
`Qualitatively, the rate-of-flow-controlled breakup (Garstecki et al. 2005d , 2006) can
`be described as follows (see figure 1): the tip of the stream of the immiscible fluid
`enters the main channel, and because interfacial stresses dominate the shear stresses,
`the tip blocks almost the entirety of the cross-section of the main channel; the shear
`stresses exerted on the tip by the continuous fluid are not strong enough to deform
`the tip significantly away from an area minimizing shape. As a result, the continuous
`phase fluid is confined to thin films between the tip of the other immiscible fluid and
`the walls of the device. Flow in these thin films is subject to an increased viscous
`resistance, which leads to a build-up of pressure in the continuous phase upstream of
`the tip (Stone 2005). This pressure is larger than the pressure in the immiscible tip,
`and the continuous fluid displaces the interface or squeezes the neck of the inner fluid,
`which leads to breakup and detachment of a droplet. Notably, within this regime the
`breakup is not driven by interfacial stresses, at least not until the very last stage; the
`speed at which the neck of the immiscible thread collapses is proportional to the rate
`of flow of the continuous fluid, and does not depend significantly on the value of
`the interfacial tension, or on the values of the viscosities of either of the two fluids
`(Garstecki et al. 2005d ). Within this regime, we can expect that the volume of the
`droplet will be a function of the ratio of the rates of flow of the two fluids, and will
`not depend strongly on the value of the capillary number (Garstecki et al. 2006).
`The qualitatively different predictions of each of the two models of breakup make
`it possible to distinguish between them by inspecting how the volume of the droplets
`
`Fluidigm Exhibit 2022
`Page 9 of 21
`
`9
`
`

`
`150
`
`M. De Menech, P. Garstecki, F. Jousse and H. A. Stone
`
`depends on the material (viscosities and interfacial tension) and flow (rates of flow)
`parameters. In this case, an advantage of numerical simulations over the experiments
`is the control over the different parameters. In the experiments, for example, it is
`difficult to change the interfacial tension between the two fluids over a wide range of
`values without introducing dynamic interfacial tension effects (at low concentrations
`of surfactant), or changing the wetting properties of the two fluids. In contrast,
`in simulations, it is straightforward to change the value of any parameter without
`affecting any others. Here we present the results of three-dimensional simulations for
`a range of the values of Ca that is broad enough to observe all three mechanisms of
`breakup in the confined geometry of a T-junction: squeezing, dripping and jetting.
`4.1. Overview of the dynamics of the system
`As the dispersed phase ent

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket