11.07.2015 Views

Removing the Stiffness from Interfacial Flows with Surface Tension

Removing the Stiffness from Interfacial Flows with Surface Tension

Removing the Stiffness from Interfacial Flows with Surface Tension

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

FLOWS WITH SURFACE TENSION 315lower order terms in <strong>the</strong> evolution. There is also an equationanalogous to Eq. (8) for L. Such a reformulation of fluidinterface evolution, as in Eq. (11 ), is a "small scale decomposition."It is <strong>with</strong>in <strong>the</strong> small scale decomposition thatimplicit integration or linear propagator methods can beapplied to a high order linear term. With periodic boundaryconditions, <strong>the</strong>se methods are explicit in Fourier space andhave no high-order time step constraint. The time-steppingmethods are coupled to spectrally accurate spatial discretizations.Small scale decompositions are given also formore general Hele-Shaw flows and for interface flows under<strong>the</strong> Euler equations.An example of what is possible <strong>with</strong> <strong>the</strong>se methods is seenin Fig. 1, which shows <strong>the</strong> simulation of a gas bubbleexpanding into a Hele-Shaw fluid (see [ 15, 16]) over longtimes. From <strong>the</strong> competition of surface tension <strong>with</strong> <strong>the</strong> fluidpumping, this simulation shows <strong>the</strong> development oframification through successive tip-splitting events and <strong>the</strong>competition between adjacent fingers. This simulation isalso spectrally accurate in space and uses a second-order intime linear propagator method for integrating <strong>the</strong> smallscaledecomposition. There are no high-order time stepconstraints. The fluid velocity is calculated using <strong>the</strong> fastmultipole method [23], and GMRES [42] is used to solve<strong>the</strong> integral equation that arises <strong>from</strong> having a viscosity con-" trast [22]. The operation count is O(N) at each time-step,where N is <strong>the</strong> number of points describing <strong>the</strong> boundary.Here N= 4096, S = 0.001, and At = 0.001. The time step is103 times larger than that used by Dai and Shelley [ 16] incomputations of a similar flow using an explicit method15<strong>with</strong> a lesser number of points, and <strong>the</strong> interface here hasdeveloped far more structure.The organization of <strong>the</strong> paper is as follows. In Section 2,boundary integral formulations are given for <strong>the</strong> motion offluid interfaces under surface tension in both Hele-Shawand two-dimensional Euler flows. The interface is interpretedas a vortex sheet whose normal velocity is determinedby <strong>the</strong> Birkhoff-Rott integral. Its vortex sheetstrength is determined by <strong>the</strong> equations of motion andboundary conditions. In Section 3, <strong>the</strong> source of stiffness in<strong>the</strong>se problems is discussed and motivated by <strong>the</strong>generalized linear stability analysis of Beale, Hou, andLowengrub [10-12]. In Section4, <strong>the</strong> Birkhoff-Rottintegral is carefully examined. It is shown that this integral,asymptotically at small scales, becomes a Hilbert transformof <strong>the</strong> sheet strength over a flat interface, <strong>with</strong> a variableprefactor. In Section 5, <strong>the</strong> description of <strong>the</strong> interface isreformulated in terms of 0 and L. This step makes <strong>the</strong>variable coefficient depend only upon time, and at smallscales, <strong>the</strong> leading order terms are only nonlocal through aHilbert transform. This leads to small scale decompositionsfor <strong>the</strong> evolution problems. In Section 6, numerical methodsare discussed. This includes second-order integrationmethods that exploit <strong>the</strong> small scale decomposition toremove <strong>the</strong> high-order stiffness, generalizations to higherorder time discretizations, as well as o<strong>the</strong>r related numericalissues such as spectrally accurate spatial discretizations. InSection 7 <strong>the</strong> results of numerical simulations using <strong>the</strong>semethods are presented. These results include <strong>the</strong> motionof Hele-Shaw interfaces moving under <strong>the</strong> competinginfluences of gravity and surface tension, in addition to <strong>the</strong>expanding gas bubble. The roll-up and collision of vortexsheets <strong>with</strong> surface tension in an Euler flow is also given.Concluding remarks are given in Section 8.10-5-10-15-1I I I I I-10 -5 0 5 10 15Time =0 to 45FIG. 1.An expanding Hele-Shaw bubble.2. THE FORMULATIONIn this section, boundary integral formulations are givenfor two illustrative incompressible flows. The first describes<strong>the</strong> motion of an interface separating two Hele-Shaw fluidsof differing densities and viscosities. The second describes<strong>the</strong> motion of a vortex sheet in a two-dimensional, inviscidfluid. As <strong>the</strong> concept of a vortex sheet arises also in <strong>the</strong>Hele-Shaw case, this second case is referred to as an inertialvortex sheet.Consider an incompressible and irrotational velocity fieldin two dimensions given in terms of a velocity potential:(u,v)=V~. Suppose that ~b has a jump across aparametrized interface F= (x(~), y(0~)), but that its normalderivative is continuous (see Fig. 2). This implies that <strong>the</strong>velocity has a tangential discontinuity across F while <strong>the</strong>component normal to F is continuous (i.e., <strong>the</strong> kinematicboundary condition is satisfied). Such an interface is called


316 HOU, LOWENGRUB, AND SHELLEYP$-~pa,/z2FIG. 2. A schematic showing F, <strong>the</strong> interface separating twoHele-Shaw fluids.a vortex sheet (see [43 ] ). The velocity away <strong>from</strong> <strong>the</strong> interfacehas <strong>the</strong> integral representation(u(x, y), v(x, Y))1 f (--(y-y(og)),x-x(og)) , ,J ~(~') (x - x(¢)) ~ + (y- y(¢))= ao~+ V(x, y, t), (12)where (x,y)¢(x(oQ, y(ct)). The velocity V accounts foro<strong>the</strong>r contributions to <strong>the</strong> motion not given by <strong>the</strong> integralterm. V is assumed smooth, at least across F, and can arisefor many reasons, such as to satisfy far-field boundary conditionsor to account for o<strong>the</strong>r interfaces. 7 is called <strong>the</strong>(unnormalized) vortex sheet strength and measures <strong>the</strong>velocity difference across F. It is given by~(~)=s,((u,, v~)-(u2, v2))l~.s.This representation is well known; see [6 and 51, 27] forsome applications to inertial and Hele-Shaw flows, respectively.While <strong>the</strong>re is a discontinuity in <strong>the</strong> tangential componentof <strong>the</strong> velocity at F, <strong>the</strong> normal component, U(00, iscontinuous and is given by (12) aswhere1w(~) =~ P.V. f ~(~')U(~)=W(~).nviscosities, and densities. For simplicity, F is assumedperiodic in <strong>the</strong> x-direction. The fluid below F is labeledfluid 1 and that above is labeled fluid 2, and likewise for<strong>the</strong>ir respective viscosities, etc. The density and viscosity areassumed to be constant above and below F, but <strong>the</strong>y candiffer across F. The velocity in each fluid is given by Darcy'slaw, toge<strong>the</strong>r <strong>with</strong> <strong>the</strong> incompressibility constraint:b 2oj= (uj, vj) = 12ktjV(ps-psgy), V.uj= O. (16)Here b is <strong>the</strong> gap width of <strong>the</strong> Hele-Shaw cell, pj is <strong>the</strong>viscosity, pj is <strong>the</strong> pressure, Ps is <strong>the</strong> density, and gy is <strong>the</strong>gravitational potential. The boundary conditions we takeare(i) [U]r'n =0, <strong>the</strong> kinematicboundary condition (17)(ii) [ P ] r = zK, <strong>the</strong> Laplace-Youngcondition (18)(iii) uj(x,y)-~O as lyl---' ~, (19)where [ f ] r = fl - f2 and fl, f2 are <strong>the</strong> limiting values <strong>from</strong>(13) below and above <strong>the</strong> interface, respectively. In addition, x isdefined as in Eq. (1) so that a circle has positive curvature.Condition (i) requires that F moves <strong>with</strong> <strong>the</strong> fluids on ei<strong>the</strong>rside. Condition (ii) relates <strong>the</strong> pressure jump across F to <strong>the</strong>interfacial curvature x, where z is <strong>the</strong> surface tension.Condition (iii) specifies that <strong>the</strong> fluid is at rest far <strong>from</strong> <strong>the</strong>interface.That <strong>the</strong> velocity field has <strong>the</strong> form given in (12) <strong>with</strong>V = 0 follows <strong>from</strong> Darcy's law (16), which implies that <strong>the</strong>(14) flow is irrotational, <strong>the</strong> incompressibility constraint, and<strong>from</strong> <strong>the</strong> boundary conditions (i) and (iii). An equation for7 follows <strong>from</strong> <strong>the</strong>se, toge<strong>the</strong>r <strong>with</strong> <strong>the</strong> Laplace-Young condition(ii); see [51 or 16] for details. In nondimensionalvariables, 7 satisfies(- (y(0Q- y(0()), x(0Q - x(ct'))x (x(~) -x(~')) 2 + (y(~) - y(~,)12d~'+V (15)and P.V. denotes <strong>the</strong> principal value integral. This integralis called <strong>the</strong> Birkhoff-Rott integral. This representation canbe given for closed or open interfaces and for situations <strong>with</strong>multiple fluids and interfaces. For Hele-Shaw and Eulerflows, our two main examples, <strong>the</strong> full formulations nowfollow.2.1. Hele-Shaw <strong>Flows</strong>Consider an interface F, as shown in Fig. 2, whichseparates two Hele-Shaw fluids of differing, but uniform,7 = -2Aus~W. ~ + Sx~- Rye. (20)is <strong>the</strong> Atwood ratio of <strong>the</strong>viscosities, S is <strong>the</strong> nondimensional surface tension, and R isa signed measure of density stratification (Pl


FLOWS WITH SURFACE TENSION 317Again, <strong>the</strong> shape of <strong>the</strong> interface is determined solely by <strong>the</strong>normal velocity, and a choice of tangential velocity onlymodifies <strong>the</strong> frame of <strong>the</strong> parametrization. Accordingly, <strong>the</strong>motion of F is given byX,(~, t) = (x,, y,) = Un + Ts. (21)Once T is specified, Eqs. (20) and (21) determine <strong>the</strong> entireflow through <strong>the</strong> motion of F. The usual choice of frame,T = W. s, is called <strong>the</strong> Lagrangian frame and corresponds tochoosing T to be <strong>the</strong> average of <strong>the</strong> limiting tangential fluidvelocities <strong>from</strong> above and below F.Remarks. (1) Hele-Shaw flow can also be interpretedas two-dimensional porous media flow. There is also acorrespondence of <strong>the</strong> Hele-Shaw flow to <strong>the</strong> Ostwaldripening problem, a quasi-stationary appoximation to diffusion.See [52], for example.(2) Only <strong>the</strong> simplest, classical dynamic boundary conditionfor Hele-Shaw flows have been considered here.More physically realistic boundary conditions have beenderived in [ 40 ].2.2. Inertial Vortex SheetsThe formulation of <strong>the</strong> motion of an interface, F,separating inviscid, incompressible, and irrotational fluids,is similiar to that for <strong>the</strong> Hele-Shaw case. As before, <strong>the</strong>density is assumed to be constant on each side of F. Here,<strong>the</strong> velocity on ei<strong>the</strong>r side of Fis evolved by Euler's equationujt+(uj.V)uj=-lv(&+pjgy), V. uj= O. (22)&where Ap=(pl--P2)/(pl+P2 ) is <strong>the</strong> Atwood ratio ofdensities and S= 2z/(pl + P2) is a rescaled surface tensionparameter (see [6, 45]). In contrast to Hele-Shaw flows,<strong>the</strong> vortex sheet strength y is an independent dynamicvariable. This is because <strong>the</strong> motion is governed by inertialforces (Euler's equation) ra<strong>the</strong>r than by viscous forces(Darcy's law). And now, Eq. (27) is a Fredholm integral of<strong>the</strong> second kind for y, due to <strong>the</strong> presence of ~)t in W,. Thekernel of <strong>the</strong> equation is <strong>the</strong> same as that in <strong>the</strong> Hele-Shawintegral equation <strong>with</strong> A, replaced by Ap. Finally, <strong>the</strong> meanof 7 is preserved by Eq. (27) and must be chosen to be 2 Vo,initially, to guarantee that condition (iii) is satisfied.3. STIFFNESS AND LINEAR THEORYNumerical stiffness arises through <strong>the</strong> presence of highorderterms (i.e., many spatial derivatives) in <strong>the</strong> evolution.The role of surface tension in producing numerical stiffnessis illustrated by considering first <strong>the</strong> linearized equations ofmotion about <strong>the</strong> flat equilibrium. That is, considerx(ct, t) = ~ + e~(~, t) and y(ct, t) = eq(ct, t), <strong>with</strong> e ~ 1. This issufficient for Hele-Shaw flows as ~ is a dependent variable.For inertial vortex sheets, this is suppliemented by 7(~, t) =So + ee~(0~, t), where 7o is a constant. For definiteness, <strong>the</strong>Lagrangian frame T= W. s is taken, and for simplicity,A, = A s = 0 for both <strong>the</strong> Hele-Shaw and inertial flows.The linearized Hele-Shaw equations of motion arewhere ~ is <strong>the</strong> Hilbert transform,~t=0 (28)qt=½Jf[Sq=~-Rr/~], (29)There are now <strong>the</strong> boundary conditions:(i) [U]r-n=0 (23)(ii) [P]r=rX (24)(iii) nj(x,y)--*(+_Vo, O) as y~ _+~. (25)Since <strong>the</strong> fluid is irrotational and incompressible, condition(i) again guarantees a vortex sheet representation of <strong>the</strong>solution, again <strong>with</strong> V = 0. Using <strong>the</strong> representation (12) of<strong>the</strong> velocity, Euler's equation at <strong>the</strong> interface, and <strong>the</strong>Laplace-Young condition, <strong>the</strong> equations of motion for <strong>the</strong>interface areX t = Un + Ts7,- O~((T- W. s) 7/s~)= - 2AAs,Wt. s + 1~,(~/s~)2 + gy,- (T-W-s) W~ .s/s~) + Sx~,W[f](ct)1 f+~ f(a') do~'. (30)n -oo 0~--0~'is a skew-symmetric linear operator, is diagonalized by<strong>the</strong> Fourier transform, and satisfies~[e ikx] = - i sgn(k) e ikx, af[1] =0. (31)Consequently, <strong>the</strong> linear growth rate for <strong>the</strong> amplitudeperturbation t/isak= --½(S Ikl3 +R Ikt). (32)Therefore, if R < 0, <strong>the</strong>re is a band of unstable modes near(26) k = 0. This is a Saffman-Tayl0r type instability, driven by<strong>the</strong> unstable density stratification. At higher wavenumbers,this instability is cut off by <strong>the</strong> surface tension term whichacts as third-order diffusion. Thus, in linearized Hele-Shaw(27) flows, surface tension is a dissipative regularization.


318 HOU, LOWENGRUB, AND SHELLEYThe source of stiffness is made clear by <strong>the</strong> linear motion.The stability constraint for an explicit time integrationmethod applied to Eq. (29) has <strong>the</strong> form~It ff[rl~] + S1 )ffEq=~]. (34)S~ S aA "frozen coefficient" analysis of Eq. (34) reveals <strong>the</strong> stricterstability constraint~It < C. (g=h)3/S, (35)where g,=min, s,. Therefore, <strong>the</strong> stability constraint isdetermined by <strong>the</strong> minimum grid spacing in arclength,which is strongly time dependent. Our experience is that <strong>the</strong>Lagrangian motion of <strong>the</strong> points can lead to "pointclustering" and hence to very stiff systems, even for flows inwhich <strong>the</strong> interface is smooth and <strong>the</strong> surface tension issmall.For inertial vortex sheets, <strong>the</strong> growth rate for perturbationsabout <strong>the</strong> flat equilibrium is given byS~2 = V~k2_~ ikl 3. (36)Again, <strong>the</strong> surface tension controls a high wave-numberinstability. The instability here is <strong>the</strong> Kelvin-Helmholtzinstability and it is due to <strong>the</strong> shearing motion across <strong>the</strong>interface. As can be seen <strong>from</strong> <strong>the</strong> growth rate, <strong>the</strong> surfacetension is a dispersive regularization in contrast to <strong>the</strong>Hele-Shaw case where it is dissipative. Again, by linearizingaround <strong>the</strong> time dependent inertial vortex sheet F=(x(0~, t), y(~, t)) <strong>with</strong> strength 7(~, t), Beale et al. [ 10, 12]find <strong>the</strong> dominant behavior for r/, again <strong>the</strong> normal componentof <strong>the</strong> perturbation, to be),2 Stltt= --~S J~O[t/~ct~t]'a'=4 ?/~t"3I- ~-~-3 2s= (37)The perturbation in y is eliminated, to leading order, byusing two time derivatives on ~/. A frozen coefficient analysisleads to <strong>the</strong> dynamic stability constraintWhile <strong>the</strong> time step constraint (38) is less restrictive thanthat for Hele-Shaw flows, our experience is that pointclustering, through <strong>the</strong> Lagrangian point motion, still leadsto prohibitively stiff systems.4. THE SMALL SCALE BEHAVIOR OF UThe stiffness of an explicit method occurs because <strong>the</strong>evolution at small length scales is controlled by a high orderterm introduced by <strong>the</strong> curvature. The normal velocity U<strong>from</strong> <strong>the</strong> Birkhoff-Rott integral contains <strong>the</strong> physically relevantpart of <strong>the</strong> velocity field, and <strong>the</strong> curvature appears init through <strong>the</strong> vortex sheet strength. In this section, <strong>the</strong>small scale behavior of U is analyzed and precisely determinedin terms of <strong>the</strong> vortex sheet strength.First, <strong>the</strong> normal velocity U is given in a convenient form.Let <strong>the</strong> complex position of <strong>the</strong> interface be given byz(e, t) = x(e, t) + iy(~, t). Then, <strong>the</strong> normal velocity is givenbyU(o~,t)=_llm{Z ~ +oo y(o~',t) }P.v. f tl .(39)This quantity is clearly related to <strong>the</strong> Hilbert transform of<strong>the</strong> vortex sheet strength 7 over <strong>the</strong> curved interface F. Ouroriginal intuition was that <strong>the</strong> small-scale behavior of thisexpression could be found by simply retaining one term in<strong>the</strong> expansion of <strong>the</strong> denominator in <strong>the</strong> Birkhoff-Rottintegral to yield a Hilbert transform over a flat interface,z = ~t, <strong>with</strong> a variable coefficient prefactor. This intuition isset rigorously in <strong>the</strong> following way. The kernel in <strong>the</strong>Birkhoff-Rott integral is rewritten as1 1mz(o~,t)-z(og, t) z,.(o~-o~')[ 1+ z(oq t)- z(og, t)(40)Note that <strong>the</strong> bracketed term has a removable singularity at= e', provided that z is a smooth function ofcq <strong>the</strong> interfacedoes not self-intersect, and s~ > 0. The Birkhoff-Rottintegral can <strong>the</strong>n be rewritten asf+~ 7(0~', t)P.V. -~ z(cq t) -z(~', t) d0(n ~ff[7]=_ +f+~Z~ oo7(~',t)g(ot, o~',t) d o~',(41)At < C. (gah)3/2/S. (38) where g is <strong>the</strong> term in <strong>the</strong> brackets of Eq. (40). Thus, <strong>the</strong>


FLOWS WITH SURFACE TENSION 319integral term containing g is a smoothing operator on y.This proves <strong>the</strong> following result.TrmOREM 1 (Small scale behavior of U). Suppose that<strong>the</strong> coordinates ofF, (x(oc, t), y(ct, t)), are real analytic functionsof oc for t O. ThenU(0c, t)=~s ~[y](0c , t)+E[y](oq t) (42)for t


320 HOU, LOWENGRUB, AND SHELLEY8~s=xn and 8~n= -xs, toge<strong>the</strong>r <strong>with</strong> O~=x and O/8s=(1/s~)(8/OoO, it is easy to see that s~ and 0 satisfys~t=T~--O~U (51)O, =--1 u~+T o~. (52)S~ S~Given s~ and 0, <strong>the</strong> position (x(a, t), y(0c, t)) is reconstructedup to a translation by direct integration of Eq. (49). In manyproblems, <strong>the</strong> motion is translation invariant and thus thislost constant is irrelevant. However, a single point on <strong>the</strong>interface can be evolved to provide <strong>the</strong> constant of integration.The velocities U and T can <strong>the</strong>n be constructed and s~and 0 updated through Eqs. (51) and (52).5.1. The Equations of Motion ReposedUsing this formulation, <strong>the</strong> general small scale decompositionis given for Hele-Shaw and inertial flows. Thedominant small scale behavior is assumed to come <strong>from</strong> <strong>the</strong>U~ term in <strong>the</strong> 0 equation and <strong>the</strong> curvature term in <strong>the</strong> yequation. U~ explicitly contains <strong>the</strong> curvature in <strong>the</strong>Hele-Shaw case and <strong>the</strong> vortex sheet strength in <strong>the</strong> inertialvortex sheet case. The evolution of s~ will be recast in termsof L in <strong>the</strong> next section.(A)Small scale decomposition: Hele-Shaw flow. For <strong>the</strong>Hele-Shaw case, using that tc = O~/s~ and recalling Eqs. (45)and (47), Eq. (52) becomes=s2(±0, 2s ,s+ ~U~ 2s~\s~* ~ ~ ~ +lO~Ts= (53)O~-SI (I~I~],),+N(cc't"2s, (54,where N is defined as <strong>the</strong> remaining terms, on <strong>the</strong> right-handside of Eq. (53), not included in <strong>the</strong> first term. By extracting<strong>the</strong> dominant term, <strong>the</strong> Hele-Shaw evolution is now in aform that reveals clearly <strong>the</strong> dominating behavior at smallscales. But most importantly, if s, is considered to be given,<strong>the</strong> dominant small scale term is linear in <strong>the</strong> tangentangle 0, but nonlocal by virtue of <strong>the</strong> Hilbert transform andvariable coefficient by <strong>the</strong> presence of s~.(B) Small scale decomposition: Inertial vortex sheets.Analogously, Eqs. (45) and (48) are used to give <strong>the</strong> 0 andy evolution for an inertial vortex sheet, in a way that displays<strong>the</strong> small scale dominating terms,ll(1 ) +P, (55)X~ = \s~l~+Q.ct(56)Again, P and Q represent <strong>the</strong> remainder terms. Assuming,as before, that s, is given, <strong>the</strong> dominant small scale terms arelinear in 0 and 7, nonlocal, and also variable coefficient.As was <strong>the</strong> case for curve shortening ( U = x), <strong>the</strong> leadingorder terms in both cases (A) and (B) simplify considerablyif s~ is independent of co. Again, this is enforced by a choicefor T, <strong>the</strong> tangential velocity. We do remark, though, that insome situations it may be preferable to use o<strong>the</strong>r choicesof T. O<strong>the</strong>r choices may prove to be useful to resolve regionsof high curvature or to naturally handle anisotropic surfaceenergies. If that is <strong>the</strong> case, <strong>the</strong> general small scale decomposition,given above, may still be useful for implicitintegration methods since <strong>the</strong> leading order terms are linearin0 (and X). Additional reference frames are given inAppendix 2 and are currently under study.5.2. The 0 - L FormulationThe general expression for T is now given so that s~ isindependent of ~ in its evolution; s= will <strong>the</strong>n depend onlyupon t, and <strong>the</strong> PDE (51) for s~ reduces to an ODE.Moreover, <strong>the</strong> dominant terms at small scales in Eq. (54) forHele-Shaw and Eqs. (55) and (56) for inertial vortex sheets,become constant coefficient in space.As in <strong>the</strong> Introduction, s~ is required to be everywhereequal to its mean, that iss~(oq t) = sw(~', t) d~' = L(t), (57)where L is <strong>the</strong> length of <strong>the</strong> interface. By differentiatingEq. (57) <strong>with</strong> respect to t and using Eq. (51), Tis found tobeT(c~,t)=T(O,t)+ O~,Vdoc-~ O~,Vdoc', (58)fowhich expresses T entirely in terms of 0 and U. The spatialconstant T(0, t) just gives an overall temporal shift in frame.For simplicity, it will be taken to be 0, although in principle,it could be evolved as well. Thus, if s, is initially uniformin ~, <strong>the</strong>n <strong>the</strong> choice for T in Eq. (58) will maintain thatconstraint in time. An expression for L, is found directly byusing Eqs. (51), (57), and (58). The evolution of <strong>the</strong> interfaceis now given in terms of L and 0 by/, 2~L t = -| 0~, UdoC (59)ao0,= T (60)Given U, Eqs. (58), (59), and (60) are a complete formulationof <strong>the</strong> evolution problem. We note that this choice offrame through T is actually a special case of o<strong>the</strong>r, moregeneral choices of frames. See Appendix 2.


FLOWS WITH SURFACE TENSION 321The small scale decompositions for Hele-Shaw and inertialvortex sheet flows in <strong>the</strong> 0-L formulation are nowgiven:(A) Small scale decomposition: Hele-Shaw flow. In <strong>the</strong>0- L formulation, Eq. (55) simplifies tos(2"/3O'=2\Lj+ut t(61)Here, N is that inferred <strong>from</strong> Eq. (54) <strong>with</strong> <strong>the</strong> choice of Tin Eq. (58). Equation (61 ) is posed, toge<strong>the</strong>r <strong>with</strong> Eq. (59),<strong>the</strong> ODE for evolving L, and is a complete specification of<strong>the</strong> interfacial problem, <strong>with</strong> <strong>the</strong> highest order, linearbehavior prominently displayed. This term is nowdiagonalizable by <strong>the</strong> Fourier transform, and soO,(k) = - ~ Ikl 30(k) + ~(k). (62)Implicit time integration methods, such as Crank-Nicholson differencing, or linear propagator methods, cannow be easily applied.(B) Small scale decomposition: lnertial vortex sheets.Analogously, for <strong>the</strong> inertial vortex sheet, <strong>the</strong> equations ink'-space areOt(k)= ~ lkl (~)2 ~(k) + P(k), (63)~,(k) = -Sk 2 ~ O(k) + O(k). (64)LHere, P and Q are those inferred by in Eqs. (55) and (56),<strong>with</strong> <strong>the</strong> choice of T in Eq. (58). The diagonalization of <strong>the</strong>leading order, linear system is trivial. And again, <strong>the</strong>se equationsmust be posed toge<strong>the</strong>r <strong>with</strong> Eq. (59).6. NUMERICAL METHODSIn this section, discrete methods are discussed for computing<strong>the</strong> motion of inertial vortex sheets and Hele-Shawinterfaces. We begin by considering two types of timeintegration methods. Both methods use <strong>the</strong> 0-L smallscale decomposition in a crucial way. The first uses anintegrating factor to remove <strong>the</strong> leading order term andgives an explicit time discretization under <strong>the</strong> Fourier transform.This is <strong>the</strong> linear propagator method. The second is animplicit Crank-Nicholson discretization of <strong>the</strong> leadingorder term, which also gives an explicit method under <strong>the</strong>Fourier transform. Spatial discretizations are chosen to bespectrally accurate.Assume for now that space is continuous. We begin bydiscussing <strong>the</strong> time integration of <strong>the</strong> length L. Its ordinarydifferential equation (59) can be discretized <strong>with</strong> an explicitmethod. In <strong>the</strong> calculations presented here, <strong>the</strong> secondorderAdams-Bashforth method is used,L .+1 =L.+_~(3M._M.-1), (65)where <strong>the</strong> superscript denotes <strong>the</strong> time level and M is givenbyM= -- f0 TM 0~, Ud~'. (66)As this is an explicit method, L can be updated before ei<strong>the</strong>r0 or T. The 0 (and V) discretizations will require thisupdated L.6.1. Hele--Shaw Time DiscretizationsThe Linear Propagator MethodLinear propagator methods factor out <strong>the</strong> leading orderlinear term prior to discretization. They usually providestable, even high-order, methods for integrating diffusiveproblems. Linear propagator methods apply also to dispersiveproblems, but must be used more carefully. Thesemethods have <strong>the</strong> property that in <strong>the</strong> absence of nonlinearityand variable coefficients, <strong>the</strong> discrete solution gives<strong>the</strong> exact solution to <strong>the</strong> constant coefficient linear problem.That is, <strong>the</strong> linear modes are propagated exactly. The firstuse of such a method (of which <strong>the</strong> authors are aware) isRogallo [41] in simulations of <strong>the</strong> Navier-Stokes equations,although it has been rediscovered and used by severalresearchers in different contexts.For Hele-Shaw, we rewrite Eq. (62) aswhereOt ~b(k' t)=exp (2zclkl) 3 ~t dt' ) ~(k, t), (67)Jo L3(t')]S ' dt' ) O(k, t)~k(k, t)= exp (2x Ikl)3 ~0 Z3(t')J (68)Equation (67) follows <strong>from</strong> Eq. (62) by finding an integratingfactor to incorporate <strong>the</strong> linear term into <strong>the</strong> timederivative. It is now Eq. (67) that is discretized using <strong>the</strong>second-order Adams-Bashforth method. In terms of 0, <strong>the</strong>result iszttOn* l(k)=ek(tn, tn+ l) i~n(k) +-~ (3ek( tn, t~+l) ~n(k)--ek(t~_~, tn+ 1)/~"-l(k)), (69)


322 HOU, LOWENGRUB, AND SHELLEYwhere t. = n At, andand soS t2 dt' "]ek(tl, tz) = exp -- ~ (2zr ]kl) 3 ft, za(t')J" (70)The use of "linear propagator" is now clear; 0 at <strong>the</strong> n thtime step is propagated forward to <strong>the</strong> (n + 1 )th timestep at<strong>the</strong> exact exponential rate associated <strong>with</strong> <strong>the</strong> linear term. IfN-0, this yields <strong>the</strong> exact solution to <strong>the</strong> linear problem.Of course, <strong>the</strong> factor e(tl, t2) still has a continuous timedependence. We retain second order by replacing <strong>the</strong>integrals <strong>with</strong> <strong>the</strong>ir trapezoidal rule approximations, i.e.,1ek(t., tn+l)=exp (-S~t(21r lkl)3[(~ ~ (Ln~- 1)3])SAtek(tn_l, tn+l)=exp -----~-(2/l:lkl) 3 (71)- 1X _2( Ln _ 1 ) 3Thus, in <strong>the</strong> numerical scheme, 0 is propagated by a secondorderapproximation of <strong>the</strong> exact exponential rate. Recallthat L" +1 is computed explicitly by Eq. (65).Remarks. 1. Near equilibrium, <strong>the</strong> second-order linearpropagator method is unconditionally stable. That is, werequire only d t ~< C(S) independently of <strong>the</strong> wavenumber k.More generally, however, <strong>the</strong>re may be a CFL conditionarising <strong>from</strong> <strong>the</strong> transport term hidden in N (see Eqs. (53)and (54)). This term is not seen in <strong>the</strong> near equilibriumanalysis. The long time simulations are always performed<strong>with</strong> a CFL condition in mind. The exponential dampingfactors also tend to suppress instabilities such as thosearising <strong>from</strong> aliasing errors or underresolution. It has alsobeen observed that, in <strong>the</strong> presence of singular behavior of<strong>the</strong> interface, this method tends to oversmooth <strong>the</strong> solution[ 19]. This difficulty is overcome by reducing <strong>the</strong> timestep. We believe that it may also be ameliorated by usinghigher order time discretizations. If <strong>the</strong> solution is smooth,<strong>the</strong>n this difficulty is not noticeable.2. An advantage of a linear propagator method is thatit is easy to derive higher order time discretizations. This isclear if <strong>the</strong>re is no time-dependence in <strong>the</strong> coefficient of <strong>the</strong>linear term. On <strong>the</strong> o<strong>the</strong>r hand, if <strong>the</strong>re is time dependence(as is <strong>the</strong> case here), <strong>the</strong>n a time integral appears in <strong>the</strong>integrating factor (see Eq. (68)). An explicit quadrature ofthis integral is avoided by introducing a new independentvariable corresponding exactly to <strong>the</strong> integral. LetI= I[L](t) be given byI(t)= fo1dt'L3(t,) (72)1I t = L--- 3, I(0) = 0. (73)Then, we supplement <strong>the</strong> original system (59) and (67) andsolveL, = M[L, I, ~k](t) (74)1I, =~--~ (75)~k, = eZ(S/2)~z~lkl)3b~[ L, L ~b](k, t). (76)It is now straightforward to discretize this system to highorder. Fourth-order Runge-Kutta is currently beingimplemented.Crank-Nicholson DiscretizationAno<strong>the</strong>r stable scheme is obtained by using <strong>the</strong>Crank-Nicholson discretization on <strong>the</strong> leading order stiffterm and leapfrog on <strong>the</strong> nonlinearity. This gives <strong>the</strong>second-order integrationO°+l(k)--O° l(k)2At+ A?"(k). (77)Since L n+l, L n, and L n-1 are known, Eq. (77) gives anexplicit expression for O n + l(k).Remarks. I. A near equilibrium analysis shows thatthis method is unconditionally stable. Again, <strong>the</strong>re may bea CFL condition <strong>from</strong> <strong>the</strong> transport term. Our computationsconfirm stability for short times, but <strong>the</strong>y also indicatethat for long times and, in <strong>the</strong> fully nonlinear regime, that<strong>the</strong> Crank-Nicholson method is susceptible to an aliasinginstability. The time of onset of this instability increases as<strong>the</strong> spatial resolution is increased. However, we find that wecan control this instability and retain accuracy, even overlong times, by <strong>the</strong> application of a high-order Fourier filter(25 th order). We will discuss this fur<strong>the</strong>r when we give <strong>the</strong>spatial discretizations and numerical results.2. The Crank-Nicholson discretization may also be usefulto compute in o<strong>the</strong>r frames of reference (i.e., using o<strong>the</strong>rchoices of T). Of course, s~ is no longer constant in ~, but<strong>the</strong> general small scale decompositions (54) and (55), (56)still apply. The PDE for s, is not expected to be stiff and <strong>the</strong>leading order terms in <strong>the</strong> 0 (and y) equations are still linearin 0 (and y), albeit <strong>with</strong> variable coefficients. Thus by


FLOWS WITH SURFACE TENSION 323updating s~ explicitly, a linear, but not diagonal, systemcan be obtained fer <strong>the</strong> updates of 0 (and 7). This systemcould <strong>the</strong>n be solved by an iterative method such asGMRES [42].O__ (e --i x/(S/2)(2~ Ik[) 3 IO (dt'/L3/2(t'))l) 1 )0t= e-i'/(s/2)(2" Ikl) 3 ~ (dt'/L3/2(t'))~, 1(8o)6.2. Inertial Vortex Sheet Time DiscretizationsThe Crank-Nicholson DiscretizationA stable integration can be obtained by discretizing<strong>the</strong> leading order stiff terms implicitly using <strong>the</strong> Crank-Nicholson discretization. There are two possible ways ofdoing this. In one, <strong>the</strong> Eqs. (63) and (64) are diagonalizedbefore <strong>the</strong> Crank-Nicholson discretization is used.However, we find it sufficient (and simpler) to use <strong>the</strong>Crank-Nicholson discretization directly on Eqs. (63) and(64), to give <strong>the</strong> method# n + 1 -- ~n 12At=[klf(4 \\Ln+lj2n ~2 n \{ 2~ ~2 )+l..~/Ln_l / ~,-1 +P"(k) (78)2At=1Sk: ( 2z: On+l+L~_lOn_l)+On(k)" (79)Now O"+~(k) and )3n+l(k) can be found explicitly byinverting a 2 x 2 matrix.Remarks. 1. A near equilibrium analysis shows thatthis method is unconditionally stable. Again <strong>the</strong>re is <strong>the</strong>possibility that a CFL condition must be satisfied. As in <strong>the</strong>Hele-Shaw case, this method suffers <strong>from</strong> an aliasinginstability at long times in <strong>the</strong> fully nonlinear regime. Asbefore, <strong>the</strong> onset time increases <strong>with</strong> increasing resolution.And again, we find that we can control this instability andretain accuracy, even over long times, by <strong>the</strong> application ofa high-order Fourier filter (25 th order).2. The scheme resulting <strong>from</strong> diagonalizing Eqs. (63)and (64) first, and <strong>the</strong>n applying Crank-Nicholson is alsounconditionally stable near equilibrium. It has not beenimplemented.3. As in <strong>the</strong> Hele-Shaw case, <strong>the</strong> Crank-Nicholson discretizationis adaptable to different choices of referenceframes.The Linear Propagator MethodA linear propagator method can also be constructed. Thisinvolves diagonalizing Eqs. (63) and (64) so as to writewhereTt(e i ~/(s/2)(2n Ikl )3 j~ (dt,/L3/2(t,))U2= e i x/(S/2)(2zc Ikl) 3 ~ (dt'/L3/2(t'))~2 '(81)V = (Vl, /)2) = ~r~--I W, W= (0, ~) (82)F--(F1,F2)--a-IF-~-IQt v, F=(/5,0) (83)<strong>with</strong> <strong>the</strong> basis matrix 0 given byg2= 2 \L/ 2 \LJ |F(2 Ikl73 s7' 2/-iL\-Z-; JThese equations can now be straightforwardly discretizedin time in an analogous manner as was done for <strong>the</strong>Hele-Shaw case (see <strong>the</strong> previous section). The secondorderAdams-Bashforth method is implemented for <strong>the</strong>seequations. While it is difficult to identify analytically <strong>the</strong>stability constraint near equilibrium, it is clear that thisdiscretization is not unconditionally stable. While <strong>the</strong>re iscertainly a restriction of <strong>the</strong> form At


324 HOU, LOWENGRUB, AND SHELLEYtime-stepping instability. This instability is observed in timecontinuous and space discrete schemes. It is due to <strong>the</strong>fact that lower order schemes unphysically suppress <strong>the</strong>stabilizing effects of surface tension at <strong>the</strong> highest modes.Beale et al. [ 12] show how lower order accurate schemescan be modified to be stable and convergent. However, noneof <strong>the</strong>se works address <strong>the</strong> issue of <strong>the</strong> temporal stabilityconstraints, which is a central focus of <strong>the</strong> work here.We use spectrally accurate spatial discretizations. Anydifferentiation, partial integration, or Hilbert transform isfound at <strong>the</strong> mesh points by using <strong>the</strong> discrete Fourier transform(DFT). To compute <strong>the</strong> complex Lagrangian velocityof <strong>the</strong> interface (84), we use <strong>the</strong> spectrally accurate alternatepoint discretization-' --- ~ ykcot , (85)uj tVJ -- 4r~i j + k oddwhere zj-=-xj + iyj denotes <strong>the</strong> approximation to <strong>the</strong> positionof <strong>the</strong> interface at <strong>the</strong> grid point jh <strong>with</strong> h = 2zr/N, andN is <strong>the</strong> number of grid points (see [ 46, 44 ]). Similar spectrallyaccurate quadratures employing prior removal of <strong>the</strong>singularity are given by Baker [4]. O<strong>the</strong>r integrals over <strong>the</strong>period, such as that in Eq. (66), are evaluated to spectralaccuracy using <strong>the</strong> trapezoidal rule.Care must also be taken <strong>with</strong> spectrally accuratemethods. Although a near equilibrium, spatially discreteanalysis indicates that <strong>the</strong>y are stable [7, 10, 12], thisanalysis involves only constant coefficients of perturbedquantities. Beale et al. [ 12] performed also a spatially discretelinear analysis far <strong>from</strong> equilibrium and noted that <strong>the</strong>spatial variation of coefficients leads to aliasing errors.These errors can be destabilizing if left uncontrolled, andBeale et al. describe several types of Fourier filtering thatguarantee stability and convergence. Such errors arise as aresult of <strong>the</strong> imposed periodicity on <strong>the</strong> discrete solutions inboth space and wavenumber, and <strong>the</strong> resultant instabilitieshave been well studied in <strong>the</strong> context of hyperbolic equations.See [21, 25, 32, 49], for example. Fourier filtering isoften used to remove <strong>the</strong>m.Typically, if <strong>the</strong> time discrete method is not smoothing at<strong>the</strong> highest modes, <strong>the</strong>n aliasing instabilities can occur overlong times. The linear propagator method for Hele-Shawflows is inherently smoothing at <strong>the</strong> highest modes, due to<strong>the</strong> exponential damping factors and does not suffer<strong>from</strong> aliasing instabilities. The o<strong>the</strong>r methods we havedescribed--Crank-Nicholson for Hele-Shaw flows andboth methods for inertial vortex sheets-~lo suffer <strong>from</strong>aliasing instabilities since <strong>the</strong>y are not naturally dampingat <strong>the</strong> highest modes. As is expected <strong>with</strong> an aliasinginstability, its onset time increases as <strong>the</strong> spatial resolutionis increased. Thus, in principle, <strong>the</strong> instability can becontrolled by simply increasing <strong>the</strong> spatial resolution. Thisis expensive. Instead, <strong>the</strong> instability is controlled by usingFourier filtering to damp <strong>the</strong> highest modes. In particular,<strong>the</strong> 25 th order Fourier filter/~[f](k) -- e-1°(Ikr/u)25f(k) (86)is employed. The filtering determines <strong>the</strong> overall accuracy of<strong>the</strong> method, and so <strong>the</strong> formal accuracy is O(h25). Certainlyan infinite-order filter could have been used, but we did notdo so. In addition, as high derivatives are computed in <strong>the</strong>code, Krasny filtering [31 ] (setting to 0 all Fourier modesbelow a tolerance level e) is employed along <strong>with</strong>/7 to controlamplification of noise through taking derivatives. It isapplied at <strong>the</strong> same time as //<strong>with</strong> e= 10 -~3, near <strong>the</strong>round-off level. This combination of both is now referred toas filtering. Again, filtering is used only in long-time computationsinvolving <strong>the</strong> nonsmoothing methods. The use ofKrasny filtering, however, is strictly a precautionarymeasure to keep <strong>the</strong> spectrum "clean" and is not necessaryfor stability. This is demonstrated in Section 7. Moreover,Krasny filtering alone would not control <strong>the</strong> aliasinginstability as <strong>the</strong> highest modes are above <strong>the</strong> tolerance leveland, hence, are not removed when <strong>the</strong> instability occurs.As a final note, we add that <strong>the</strong> role of filtering here issomewhat different <strong>from</strong> its use in zero surface tensioncalculations [31, 44]. There, Krasny filtering is used tocheck an anomolous growth of noise induced by <strong>the</strong>Kelvin-Helmholtz instability. Uncontrolled, this growthdestroys <strong>the</strong> accuracy of computations in very short times,well before any singular behavior occurs, and increasing <strong>the</strong>spatial resolution only enhances <strong>the</strong> instability. Here, <strong>the</strong>ma<strong>the</strong>matical problems are at least linearly well posed[ 10, 12] and so for stable schemes, it should not be crucialthat <strong>the</strong> computations are free of spurious erorrs. This isborne out for our methods as simply increasing <strong>the</strong> spatialresolution delays <strong>the</strong> onset of instability. The purpose of <strong>the</strong>filtering here is to provide an efficient way to preserve <strong>the</strong>overall accuracy in time. Finally, <strong>the</strong> techniques of Beale etal. [ 12] can be used to show <strong>the</strong> convergence of our numericalmethods.6.4. O<strong>the</strong>r ConsiderationsIt is useful to discuss <strong>the</strong> construction of <strong>the</strong> initial equalarclength parametrization, as well as <strong>the</strong> construction of <strong>the</strong>mapping <strong>from</strong> <strong>the</strong> (x,y) description of <strong>the</strong> interfaceto (L, 0),Equal Arclength InitializationThe procedure for obtaining <strong>the</strong> initial equal arclengthparametrization is presented in Appendix B of [9]. Theidea is just to solve <strong>the</strong> nonlinear equation for <strong>the</strong> equal


FLOWS WITH SURFACE TENSION 325arclength grid points using Newton's method and Fourierinterpolation. The equation that must be solved isfor flj as j=0(1)NLs,,d~'=~-~ s~,d~'=jh~-~n, (87)<strong>with</strong> h=2n/N. Here ~ is a givenparametrization, not necessarily arclength, and flj gives <strong>the</strong>location of points in <strong>the</strong> ~ parametrization that are equallyspaced in arclength. L is obtained by trapezoidal integrationofs~ over its period.Moreover, in <strong>the</strong> case of inertial vortex sheets, <strong>the</strong> unnormalizedvortex sheet strength in <strong>the</strong> arbitrary parametrization~, must be rescaled appropriately for use in <strong>the</strong> equalarclength frame. This is because in <strong>the</strong> inertial vortex sheetcase, <strong>the</strong> unnormalized vortex sheet strength appears in aframe-dependent way in <strong>the</strong> equations of motion, see (15)and (27). In particular, suppose that ~(~, 0) is given. Then,<strong>the</strong> initial data used in <strong>the</strong> equal arclength frame isy(flj, 0). fl~(jh), where fl~ is computed <strong>from</strong> <strong>the</strong> solution ofEq. (87) using <strong>the</strong> DFT. The true vortex sheet strength (i.e.,<strong>the</strong> tangential jump in velocity: 7/s~), on <strong>the</strong> o<strong>the</strong>r hand, isframe-independent. Of course, it is not necessary to rescale<strong>the</strong> unnormalized vortex sheet strength in <strong>the</strong> Hele-Shawcase as it is not an independent variable and does, in fact,appear in a frame-independent way.The Forward Mapping (x, y) ~ (L, O)Recall that <strong>the</strong> tangent angle is given byO= tan-~(y~/x~). However, this is not a good formula touse numerically as it is tricky to ensure that <strong>the</strong> variation ofy~/x~ over <strong>the</strong> interface does not result in jumps in 0 due tobranching in <strong>the</strong> inverse tangent. It is better to construct 0<strong>from</strong> <strong>the</strong> curvature by integrating <strong>the</strong> formulaO~=s~K, where x=(x~y~-y~x~)/s]. (88)The constant of integration may be chosen by using <strong>the</strong>inverse tangent formula at <strong>the</strong> initial point of integration.The Inverse Mapping (L, 0) ~ (x, y)The difficulties in recovering (x, y) <strong>from</strong> (L, 0) are purelyimplementational; (x, y) are obtained by integrating <strong>the</strong>formulaeLLx, = cos(0), y~ = ~ sin(0). (89)For <strong>the</strong> flows that are periodic in x, <strong>the</strong> formulation requiresthat x = cc + p(0c, t) and y = q(~, t), where p, q are periodicin 0~. In particular, <strong>the</strong> coefficient of <strong>the</strong> linear term inmust be exactly 1 in <strong>the</strong> x coordinate and 0 in <strong>the</strong> ycoordinate. Unfortunately, integrating (89) using <strong>the</strong> DFTperturbs <strong>the</strong>se coefficients slightly due to numerical error.Unchecked, this has a devastating effect on <strong>the</strong> code as <strong>the</strong>assumed periodicity of <strong>the</strong> solution is altered. The alternatingpoint quadrature role, for example, loses spectralaccuracy and <strong>the</strong> code becomes unstable. This difficulty iseasily fixed by forcing <strong>the</strong> coefficients to be exactly 1 and 0,respectively, after <strong>the</strong> reconstruction process. For example,we reconstruct x by using <strong>the</strong> discretization ofx(~,t)=x(O,t)+~I 1 L ~2~ doCl- Jo+ ~ cos(0(~')) d~'. (90)Of course, in <strong>the</strong> absence of numerical error, <strong>the</strong> explicitcoefficient of~ in (90) vanishes. We have considered o<strong>the</strong>rways to enforce this condition, but we have found that justexplicitly forcing <strong>the</strong>se coefficients to be 1 and 0, respectively,performs <strong>the</strong> best numerically. We remark that Strainalso noted this difficulty in [48], but he did not seem toemploy a correction of it computationally.7. SOME NUMERICAL RESULTSIn this section, <strong>the</strong> results of numerical simulations arepresented for several fluid interface problems. The first of<strong>the</strong>se is <strong>the</strong> motion of Hele-Shaw interfaces, which evolveby a competition of surface tension and unstable densitystratification. The second is <strong>the</strong> expansion of a gas bubbleinto a Hele-Shaw fluid. The third is <strong>the</strong> motion of inertialvortex sheets, which evolve by a competition of surface tensionand <strong>the</strong> Kelvin-Helmholtz instability. And <strong>the</strong> fourth is<strong>the</strong> motion of an interface <strong>with</strong> surface tension in anunstably stratified Boussinesq fluid. All of <strong>the</strong>se simulationsuse <strong>the</strong> appropriate small scale decomposition, toge<strong>the</strong>r<strong>with</strong> <strong>the</strong> associated numerical methods discussed in <strong>the</strong>previous section.The computations in <strong>the</strong> open geometry (i.e., all cases but<strong>the</strong> expanding bubble) assume one-periodic interfacesra<strong>the</strong>r than 2n-periodic as was assumed in <strong>the</strong> previous sections.The two are equivalent by a rescaling of time, space,and <strong>the</strong> physical parameters.7.1. Numerical Results: Hele-Shaw7.1.1. Numerical <strong>Stiffness</strong>This section begins <strong>with</strong> a comparison of <strong>the</strong> stabilityconstraints of three methods for Hele-Shaw flow--<strong>the</strong>second-order linear propagator, <strong>the</strong> Crank-Nicholson/leapfrog, and <strong>the</strong> explicit second-order Adams-Bashforthmethod applied to small scale decomposition (62). Todemonstrate <strong>the</strong> constraints on <strong>the</strong>se methods (especially


326 HOU, LOWENGRUB, AND SHELLEYfor <strong>the</strong> explicit scheme), it is sufficient to consider shorttimes and <strong>the</strong> simple initial conditionx(a, O) =~, y(~, O) = -0.01 sin 2m~. (91)Here, R =- 1 and S = 0.01 so that <strong>the</strong> flow is unstablystratified although k = + 1 are <strong>the</strong> only linearly growingmodes. A~ = 0 is taken for simplicity.For each method, Fig. 3 shows plots of <strong>the</strong> Fourier transformlOglo IP(k)l versus k at t = 0.1 for <strong>the</strong> two resolutionsN= 64 (top) and N= 128 (bottom). The results <strong>from</strong> <strong>the</strong>integrating factor and Crank-Nicholson methods use <strong>the</strong>timestep At = 0.01 at both spatial resolutions. Instability ina time-integration method is usually manifested as rapidand unphysical growth in <strong>the</strong> amplitudes of <strong>the</strong> highwavenumbermodes. As is clear <strong>from</strong> <strong>the</strong> spectrum, no timestep reduction is necessary for stability of ei<strong>the</strong>r method as<strong>the</strong> resolution is increased, at least at <strong>the</strong>se resolutions. Forlarger values of N, i.e., N up to 2048, we do see <strong>the</strong>emergence of a first-order CFL constraint imposed by <strong>the</strong>transport term. Although this constraint is not seen in<strong>the</strong> near equilibrium linear analysis, as <strong>the</strong> transport term(see Eq. (53)) is neglected, it is not surprising that it is seenin <strong>the</strong> full computation for large N.The results for <strong>the</strong> explicit scheme are shown in <strong>the</strong> lastcolumn. The solid curve corresponds to t = 0.10 <strong>with</strong> At =2.0x10 -5 for N=64, and At=O.25xlO -5 for N=128.This reduction in time step is exactly <strong>the</strong> factor of 8 that is0-5logz0l:9(k)t-10-15-20 00-5loSxol~'(k)l-lOLin. Prop., (a)t=.10d~.01, N=6401-51-15t=.10C-N. (b)dt=.01, N=64w20 40 60 80 20 40 60 80kkt=.10(d)dt=.01, N=128C-5t=.10(e)dr=.01. N=128Explicit A-B, (c)0,._ t=.002, dt-=4.0d-5- t=. 10. dt=2.0d-5-5~ N=64-lOlli ,,"iii't i-15 ~-2C0 20•40 60 80k(00 - t=.0002, dt=0.5d-.*- t=.10, dt=O.25d-5-5 N=128-1C -10 t,II-15 -15 -15 r JJ i-20 ~0 20 40 60 80-2C0 20 40 60 80-200 20 40 60 80k k kFIG. 3. Hele-Shaw numerical stiffness: acomparison oflogl0 ]~(k)[ vsk at t=0.10 and R=-I, S=0.01: (a) linear propagator, N=64,At = 0.01; (b) Crank-Nicholson N= 64, At = 0.01; (c) explicit secondorderAdams-Bashforth, N= 64, At = 2.0 x l0 -5 and at t = 2.0 x 10 -3 <strong>with</strong>At=4.0xl0-5; (d) lin. prop., N=128, At=0.01; (e) C-N, N=128,At=0.01; (f) explicit A-B, N= 128, At=0.25 x 10 -5 and t=2.0 x 10 -4<strong>with</strong> At = 0.5 x 10-st*stipulated by <strong>the</strong> near equilibrium stability constraint/it~ Ch 3. To demonstrate <strong>the</strong> proximity of <strong>the</strong> stabilitythreshold, <strong>the</strong> results are given also for computations using<strong>the</strong> intermediate time steps At = 4.0 × 10-5, at t = 0.002 forN= 64, and At = 0.5 x 10 -5, at t = 0.0002 forN= 128 (givenby <strong>the</strong> dashed lines). That <strong>the</strong> stability criterion is violatedis shown by <strong>the</strong> unphysical growth of <strong>the</strong> spectrum at highwavenumbers. The latter two computations cannot be continuedmuch beyond <strong>the</strong>se times as <strong>the</strong> numerical solutionblows up.At <strong>the</strong>se early times, <strong>the</strong> flow has not yet developed spatialcomplexity. Therefore, <strong>the</strong>re are no significant aliasingerrors and hence no filtering is used in <strong>the</strong>se calculations.Aliasing errors become especially important when <strong>the</strong> activeportion of <strong>the</strong> spectrum approaches and exceeds <strong>the</strong>Nyquist frequency k = N/2. Finally, <strong>the</strong> explicit computation<strong>with</strong> N = 128 and/It = 0.25 × 10 -5 takes approximately175 min on an IRIS Indigo workstation while <strong>the</strong> linearpropagator and <strong>the</strong> Crank-Nicholson computations eachtake less than 30 s (due to <strong>the</strong> fact that <strong>the</strong>ir time steps are/It = 0.01 each).7.1.2. Longer Time ComputationsConsider now <strong>the</strong> evolution <strong>from</strong> <strong>the</strong> multimodal intialconditionx(~, O) = a, y(~, O) = 0.01 cos 2zr~ - 0.01 sin 6try, (92)<strong>with</strong> S= 0.1, R = -50, and A, = 0. Now, <strong>the</strong> modes [k[ ~< 3are linearly unstable, and <strong>the</strong> competition between <strong>the</strong> sur-20-1-2(a)-0.5 0 0.5 1 1.5t=0(d)1110-1(b)-0.5 0 0.5 I 1.5t=0.04(e)-0.5 0 0.5 1 1.5 -0.5 0 0.5 1 1.5t=0.08t=0.10o{-2}17, )/1"65 I(c)-0.5 0 0.5 1 1.5t=0.061.6 ]1.55I(f)1.5 I0.75 0.8 0.85t=0.10FIG. 4. Long-time evolution of a Hele-Shaw interface: S=0.1,R = -50, N= 2048, At= 3.125 x 10-5: (a) t= 0; (b) t= 0.04; (c) t= 0.06;(d) t = 0.08; ( e ) t = 0.10; (f) close-up of topmost pinching region, t = 0.10.


FLOWS WITH SURFACE TENSION 327face tension and <strong>the</strong> unstable stratification causes <strong>the</strong> interfaceto rapidly develop a ramified spatial structure.A time sequence of inteface positions is shown in Fig. 4using N=2048 and At= 3.125 × 10 -5. The second-orderlinear propagator method is used, and <strong>the</strong> time step ischosen small enough so as to effectively eliminate time-steppingerrors <strong>from</strong> <strong>the</strong> computation. At early times, threerising and falling fingers of fluid form. As time progresses,<strong>the</strong> tips of <strong>the</strong>se fingers thicken and begin to resemblebubbles. The necks of <strong>the</strong> fastest moving fingers begin tonarrow and <strong>the</strong>ir sides approach tangentially. The o<strong>the</strong>rnecks appear to follow suit. These necks become localizedjets, fluxing fluid <strong>from</strong> <strong>the</strong> bulk into <strong>the</strong> bubbles. At latertimes, <strong>the</strong> sides of <strong>the</strong> necks seem to self-intersect or pinch.However, a close-up of <strong>the</strong> neck region (box (f)) of <strong>the</strong>topmost bubble reveals that <strong>the</strong> pinching has not yet takenplace by t = 0.10 and that <strong>the</strong> neck still has a nonzero width.As <strong>the</strong> necks narrow, however, it becomes more and moredifficult to maintain resolution. This is because <strong>the</strong> alternatepoint discretization of <strong>the</strong> Birkhoff-Rott integral requires aminimum width of six or seven grid lengths in <strong>the</strong> neckregion for accuracy [ 8 ]. This effect is illustrated in Fig. 5,where <strong>the</strong> interface positions are compared at t = 0.10 <strong>with</strong>N= 1024 and N---2048. Although pinching has alreadytaken place in <strong>the</strong> N = 1024 calculation, it is unphysical. Therrrinimum width of <strong>the</strong> neck as a function of time is shownin Fig. 6 for several spatial resolutions: N= 512, N= 1024and N=2048 (again using <strong>the</strong> same time step Lit=3.125 x 10-5). The width is computed by minimizing <strong>the</strong>distance function between <strong>the</strong> bounding curves, which arerepresented by Fourier polynomials. The width of <strong>the</strong> neckdecreases rapidly at early times. By t = 0.065 <strong>with</strong> N= 512,<strong>the</strong> neck is but one grid length wide, and <strong>the</strong> calculationbecomes inaccurate. The higher resolution calculations21.510.50-0.5-1-1.5-2 1024i0LiI 2FIG. 5. Comparison of Hele-Shaw interfaces at t = 0.10 for differentspatial resolutions: <strong>the</strong> left corresponds to N= 1024 and <strong>the</strong> right toN=2048; S=0.1, R= -50, At= 3.125 x l0 -5.)L32O480.12-- 512-.- 10240.1 - 20480.08~ 0.0~0.0,t0.020 5 6 7 8 :, 10T x 10 .2FIG. 6. Minimum distance to pinching for N= 512, N= 1024, andN= 2048; S = 0.1, R = -50, At = 3.125 x 10 -5.indicate that <strong>the</strong> narrowing slows shortly <strong>the</strong>reafter.However, by t = 0.08, <strong>the</strong> N = 1024 calculation has a neckwidth of only five grid lengths, and that calculation becomesinaccurate. But by using N = 2048, <strong>the</strong> width is seen tosaturate after t= 0.08 and shows only a slight fur<strong>the</strong>rdecrease by t = 0.10. The neck region is 11 of its grid lengthswide. Presumably, <strong>the</strong> width of <strong>the</strong> neck region scales <strong>with</strong><strong>the</strong> surface tension, although this has not been studied indetail.This lack of self-intersection in <strong>the</strong> neck region is consistent<strong>with</strong> behavior found by Goldstein, Pesci, and Shelley[ 18 ] in asymptotic models of jets in Hele-Shaw flows. Theirmodelling suggests also that <strong>the</strong> minimum neck widthdecreases by a factor of two <strong>with</strong> a fourfold decrease in <strong>the</strong>surface tension. It is only in <strong>the</strong> limit of zero surface tensionthat <strong>the</strong>ir model equation predicts <strong>the</strong> breaking of <strong>the</strong> neck.They do consider o<strong>the</strong>r cases, however, where <strong>the</strong> surfacetension does not prevent <strong>the</strong> pinching of material interfaces.Figure 7 shows <strong>the</strong> evolution of <strong>the</strong> curvature. The topleft plot shows <strong>the</strong> inverse of <strong>the</strong> maximum absolute curvature.Vanishing in this plot would correspond to adivergence of <strong>the</strong> curvature. At early times, <strong>the</strong>re is a rapidincrease in <strong>the</strong> curvature (a decrease in <strong>the</strong> plot). It peaksaround t= 0.015 and <strong>the</strong>n begins to decrease. This lastsuntil t -- 0.05. The curvature <strong>the</strong>n refocuses and <strong>the</strong> processrepeats itself several times. By <strong>the</strong> end of <strong>the</strong> computation,<strong>the</strong> curvature has nearly reached again its peak at t = 0.015.The o<strong>the</strong>r plots in Fig. 7 show x as a function of ~ at severaltimes; <strong>the</strong>se graphs indicate that in fact, <strong>the</strong> curvaturesaturates at one part of <strong>the</strong> interface and refocuses inano<strong>the</strong>r, leading to a very complicated overall structure.The boundaries of <strong>the</strong> narrow neck regions are indicated bya pairs of closely spaced peaks of like signed curvature. Thephenomenon of saturation and refocussing will be seenagain in <strong>the</strong> context of inertial vortex sheets.


328 HOU, LOWENGRUB, AND SHELLEY0.250.20.150A0.05Inverse Curvature, (a)20110101Curvature, (b)0 0.05 O. 1 0.5TT=0.02Curvature, (d)20 20110-10101O~ilCurvature, (e)Curvature, (c)-200 0.5 0.5-200.5T--0.06 T--0.08 T--0.10200-10-20 0 0.5T---0.04200-10Curvature, (f)FIG. 7. Time evolution of <strong>the</strong> curvature; S = 0.1, R = -50, N = 2048,At= 3.125 x 10-5: (a) plot of inverse maximum of curvature (absolutevalue); (b) curvature at t=0.02; (c) t=0.04; (d) t=0.06; (e) t=0.08;(f) t = 0.10.Error Analysis. In Fig. 8, spatial convergence isdemonstrated by comparing computations using N= 256,512, and 1024 <strong>with</strong> those <strong>from</strong> N= 2048. The time step isAt = 3.125 × 10-5 for all resolutions. The error is measuredas eN(t) = maxj Ixj(t; N) - xj(t; 2048)1, where Nis <strong>the</strong> numberof points in <strong>the</strong> lower resolution calculations. The erroris plotted on a negative logarithm (base 10) vertical scale.The error is consistently around 10- lO, for each of <strong>the</strong> lowerresolutions, until <strong>the</strong> interfaces nearly pinch. Then, rapidlosses of accuracy are seen <strong>from</strong> about t = 0.03 for N = 256,~ 6"712 ~ ~ . _ _113842Ci0.01 0.02 0.(13-- 256-~ -.- 512i~ - 1024' ~ .,\~,, ,.\,i .ii i i i i . . . . . i'- ....0.04 0.05 0.06 0.07 0.08 0.09 0.1TFIG. 8. Error in x-coordinate (-loglo(error)); S=0.1, R=-50,Jt = 3.125 x 10-5, exact solution approximated by N= 2048.t=0.05 for N=512, and t=0.07 for N= 1024. It is clearthat more resolution will be required for fur<strong>the</strong>r computationof this flow. Fur<strong>the</strong>r, as <strong>the</strong> quadrature we use is <strong>the</strong>chief culprit in loss of accuracy, it would be useful to considero<strong>the</strong>r types of quadrature (i.e., product integrationmethods) that do not lose accuracy so catastrophicallythrough <strong>the</strong> close approach of interfaces. The second-ordertemporal convergence can be shown similarly, but it is notpresented here.7.1.3. The Expanding BubbleThe calculation of a gas bubble expanding into aHele-Shaw fluid, shown in Fig. 1, is now briefly discussed.The dynamics of expanding bubbles in <strong>the</strong> radial geometryhave attracted a great deal of attention due to <strong>the</strong> formationof striking patterns observed in experiments (see <strong>the</strong> manyreferences and contributions in [47 ], for example). In thisflow, an expanding, unstable interface is produced byplacing a mass source at <strong>the</strong> center of <strong>the</strong> bubble. Linearizationabout an expanding circular bubble, <strong>with</strong> radius R(t)and pumping rate dA/dt, gives <strong>the</strong> instantaneous growthrate [ 16 ]1(dA/dtak(t) =R--~ (k- 1) --2reS\-- (k- 1)k(k+ 1)).R(t)(93)The pumping term replaces <strong>the</strong> gravitational term in <strong>the</strong>previous example as <strong>the</strong> source of instability. For a constantpumping rate (here dA/dt=2g) we obtain a Mullins-Sekerka type instability, which shows <strong>the</strong> competitionbetween <strong>the</strong> destabilization effect due to pumping and <strong>the</strong>stabilizing effect due to surface tension.Unlike <strong>the</strong> previous calculation, <strong>the</strong>re is now a viscositycontrast (A, = 1,/t = 0 inside <strong>the</strong> bubble). Consequently, anintegral equation analagous to Eq. (20) must be solved for<strong>the</strong> vortex sheet strength y (see Dai and Shelley [ 16 ] ). Here,<strong>the</strong> integral equation is solved in its dipole form, for<strong>the</strong> dipole strength v, which is related to y by ? = v~ (seeGreenbaum, Greengard, and McFadden [22]). Theintegral equation is solved iteratively using <strong>the</strong> GMRESmethod [42]. Birkhoff-Rott type integrals over a closedinterface (although <strong>with</strong> smooth kernels) must be evaluatedat each step. The fast multipole method [23] is used toevaluate <strong>the</strong>m. Thus, at each iteration, <strong>the</strong> operation countis O(N) as opposed to O(N 2) for direct summation. Theconvergence tolerance for <strong>the</strong> GMRES iteration is set to10-x2. The rate of convergence is improved considerably bya simple diagonal preconditioning, as used in [22], and bysupplying a good first guess through an extrapolation ofsolutions at previous time steps. Once <strong>the</strong> solution to <strong>the</strong>integral equation is obtained, <strong>the</strong> Dirichlet-Neumann mapis used to determine <strong>the</strong> normal velocity of <strong>the</strong> interface.


FLOWS WITH SURFACE TENSION 329This results again in a Birkhoff-Rott type integral (now<strong>with</strong> a singular kernel) that must be evaluated and alternate 3point quadrature using <strong>the</strong> fast multipole methods is2.5employed to this end. See [ 16, 22] for details. Time integra-2~ a5tion is accomplished using <strong>the</strong> second-order linearpropagator method on <strong>the</strong> small scale decomposition; i.e., -~<strong>the</strong> method is given by Eq. (69) where <strong>the</strong> term A ~ is 0.5modified to account for a pumping term (see [ 16 ]). The 0method is spectrally accurate in space.The initial condition is given by2.42.2Log-Log plot of A(t) vs. G(t), area vs. radius of gyration, (a).J0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9log10 G(t)Plot of Slope, (b).(Xo(~), yo(~)) = r(~)(cos 0t, sin a)<strong>with</strong> r(0~) = 1 +0.1 sin 20t+0.1 cos 30t (94)and is shown as <strong>the</strong> innermost curve of Fig. 1; it imposes noparticular symmetry on <strong>the</strong> ensuing motion. The value of<strong>the</strong> surface tension is S = 0.001. At t = 45, <strong>the</strong> time step isAt = 0.0005 and N= 8192.Figure 1 shows <strong>the</strong> expansion of this bubble <strong>from</strong> t = 0 upto 45, at unit intervals of time. This simulation displaysmuch of <strong>the</strong> behavior that has stimulated interest in patternformation in Hele-Shaw flows. An early times, three main"l]ords" form in <strong>the</strong> interface. These t]ords separate threeexp"anding fronts. The number of fjords arises <strong>from</strong> <strong>the</strong> k = 3component in <strong>the</strong> initial data. The expanding fronts rapidlydevelop oscillations, particularly along <strong>the</strong>ir outer edgesnear <strong>the</strong> t]ords, which <strong>the</strong>mselves form "fingers" and"petals." The petals expand outwards and eventually tipsplitinto two petals. Although this tip splitting temporarilyrestabilizes <strong>the</strong>m, <strong>the</strong> process repeats itself. One petal in <strong>the</strong>second quadrant has already tip-split four times and isapproaching a fifth tip-splitting. There is also abundantevidence of competition between <strong>the</strong>se various structures.Of <strong>the</strong> approximately 25 protuberances that develop atearly times (say between t = 5 and 8), only about 15 of <strong>the</strong>mare still actively growing outwards as t = 45. The remainderhave ei<strong>the</strong>r stopped growing outwards, or have receded andbeen absorbed back towards <strong>the</strong> main bulk of <strong>the</strong> bubble.Although fur<strong>the</strong>r details of <strong>the</strong>se calculations will appearelsewhere, we present one interesting diagnostic. Figure 9shows <strong>the</strong> bubble area A(t) versus G(t), <strong>the</strong> "radius of gyration,"on a log-log scale (upper graph). G(t) is defined as <strong>the</strong>maximum distance of a point on <strong>the</strong> interface <strong>from</strong> <strong>the</strong> injectionpoint. It is computed numerically by maximizing <strong>the</strong>Fourier interpolant of <strong>the</strong> distance function <strong>from</strong> <strong>the</strong> origin(<strong>the</strong> injection point) to <strong>the</strong> interface. The radius of gyrationhas been used as a measurement of complexity, as <strong>the</strong> slopeof log G versus log A, shown in <strong>the</strong> middle and lowergraphs, gives roughly a dimension d; where d is <strong>the</strong> dimensionof <strong>the</strong> bubble (i.e., A ~ Gd). For example, <strong>the</strong> slope istwo for a circular bubble. This slope is, of course, notsmooth, as <strong>the</strong> point of maximum radius jumps between dif-2o 1.81.61.41.2 0 5 10 15 20 25 30 35 40 45TimeFIG. 9. (a) The bubble area A(t) vs <strong>the</strong> radius of gyration G(t) on alog-log scale. (b) The points are <strong>the</strong> slope (dimension) dmeasured <strong>from</strong> (a)as a function of time. The dashed lines are at d= 1.66 and 1.71 (DLAestimates).ferent sites during <strong>the</strong> evolution. The jumps <strong>the</strong>mselves areassociated typically <strong>with</strong> tip-splitting events, as <strong>the</strong> splittingallows a newly formed finger to move outwards morequickly. The dashed lines have <strong>the</strong> values 1.71 and 1.66 (see[53, 26], respectively), which are estimates for <strong>the</strong> fractaldimension of a branched object grown via diffusion limitedaggregation (DLA), a prototypic model for <strong>the</strong> growthof branched structures [53]. There is a short period(15


330 HOU, LOWENGRUB, AND SHELLEYspectrum had risen to 10 -8 (<strong>from</strong> 10-16). The N= 8192computation was stopped at t = 45, even though its highmodes had risen to only 10-11. We note that <strong>the</strong> computedarea of <strong>the</strong> bubble at t = 45 still agrees to six digits <strong>with</strong> itsexact value.The calculation to t = 25.0 took about 9 days on anIBM R6000 workstation. Had <strong>the</strong>re been no accompayingintegral equation to solve, it would have taken but a fewhours. The best approach to reducing this time appears tobe finding better preconditioners for <strong>the</strong> iteration. Previouscalculations of similar flows have been performed [ 15, 16].The time step here is 103 times larger than that used by Daiand Shelley [ 16 ] in computations of a similar flow using anexplicit method <strong>with</strong> a lesser number of points, and <strong>the</strong>interface here has developed far more structure. Fur<strong>the</strong>r, nosymmetries have been imposed. Again, fur<strong>the</strong>r details willappear elsewhere.7.2. Numerical Results: Inertial Vortex SheetsIn <strong>the</strong>se last two subsections, we examine <strong>the</strong> long-timeevolution of inertial vortex sheets <strong>with</strong> surface tension, bothin a homogeneous fluid and in a fluid <strong>with</strong> slight densitystratification. Leaving aside issues such as dimensionalityand viscous effects, <strong>the</strong> first case is not a completely artificialsituation. Two immiscible fluids can be density matched andstill produce a surface tension. Fur<strong>the</strong>r, in <strong>the</strong> absence ofsurface tension, <strong>the</strong> singularities of a vortex sheet embedded<strong>with</strong>in a homogeneous fluid give <strong>the</strong> basic form of <strong>the</strong>singularities observed on an interface evolving by <strong>the</strong>Rayleigh-Taylor instability [ 5 ]. Our results in <strong>the</strong> last subsectionsuggest that this still holds true <strong>with</strong> <strong>the</strong> addition ofsurface tension.Computationally, such problems have also been consideredby Pullin [ 38 ], by Rangel and Sirignano [ 39], byBaker and Nachbin [7], and by Beale et al. [ 10, 12]. Pullinused a method based on cubic splines and observed apersistent numerical instability that was only slightlyameliorated by additional smoothing of <strong>the</strong> interface position.His calculations are restricted to short times. Rangel etaL used a similar method, combined <strong>with</strong> a remeshing, usinglinear interpolation of <strong>the</strong> interface positions and sheetstrength, at every time step. This is strongly smoothing.Indeed, <strong>the</strong>y calculate <strong>the</strong> smooth roll-up of <strong>the</strong> sheet, even<strong>with</strong>out surface tension, for which it is known that asingularity interrupts <strong>the</strong> motion well before roll-up occurs[33, 31, 44]. Baker and Nachbin and Beale et al. investigate<strong>the</strong> stability of spatial discretizations. In <strong>the</strong>se works, <strong>the</strong>issue of stiffness is not addressed, and <strong>the</strong> calculations areagain restricted to short times.7.2.1. Numerical <strong>Stiffness</strong>In this section, <strong>the</strong> stability constraints are compared for<strong>the</strong> second-order linear propagator, Crank-Nicholson/leapfrog, and explicit second-order Adams-Bashforthmethods for small scale decomposition of <strong>the</strong> 0 - L formulationof inertial vortex sheets, Eqs. (55) and (56). For <strong>the</strong>explicit scheme, <strong>the</strong> Adams-Bashforth discretization is usedin all three equations. The initial condition isx(~, 0) = ~ + 0.01 sin 2zt~, y(ct, 0) -- -0.01 sin 2~,7(~, 0)= 1.0.(95)This initial data was used by Krasny [31] in <strong>the</strong> absence ofsurface tension. In that case, a curvature singularity forms att ~ 0.375. The calculations here have S = 0.005, which gives16 linearly growing modes in <strong>the</strong> period.Figure 10 shows plots of log10 IP(k)F for <strong>the</strong> threemethods <strong>with</strong> N = 64 and N= 256 at t = 0.12. This is a factorof 4 increase in N, and so <strong>the</strong> near equilibrium constraintAt < Ch 3/2 requires a factor of 8 decrease in time step. Theresults <strong>from</strong> <strong>the</strong> Crank-Nicholson/leapfrog scheme aregiven in <strong>the</strong> first column. The time step At = 0.01 is used forboth resolutions and this scheme is clearly stable. As <strong>the</strong>reis little spatial complexity (i.e., no large aliasing errors),filtering has not been used. The results <strong>from</strong> <strong>the</strong> explicitAdams-Bashforth scheme are given in <strong>the</strong> middle column.The +-curves correspond to At = 3.1 x 10-4 for N = 64 andAt = 3.9 x 10-5 for N= 256. This is again <strong>the</strong> factor of 8 aspredicted by <strong>the</strong> stability constraint. The solid lines showthat <strong>the</strong> time step constraint is violated <strong>with</strong> At = 1.2 x 10 -3<strong>with</strong> N= 64 and At = 7.8 x 10-5 <strong>with</strong> N = 256. In <strong>the</strong> thirdcolumn of <strong>the</strong> figure <strong>the</strong> results are <strong>from</strong> <strong>the</strong> integratingl°glol$'(k)!l 0C-N, (a)° i , t=.12, N=64I-15 ~I]- dt=.01o-5-IO-15Expficit A-B. (b)t=. 12, N=64- dt=l.2d-3+ dt~3.1d-4Lin. Prop., (c)-2c-20 I50 1000 50 100kkk-2C-2-2050 100 50 I00 0 50 100k k k(d)(t=.12, N=256(g)t=.12, N=256(Dt=.12, N=256-5 - dt=.01- dt=7.Sd-5- dt=9.Sd-6-5+ dt=3.9d-5logaol~'(k)l-K-15-10:-10-15-10-15t=.12, N=64- dt=6.3d-4+ dt=l.6d-4FIG. 10. Inertial vortex sheet numerical stiffness: a comparison oflog10 I~(k)l vs k <strong>with</strong> S= 0.005 at t = 0.12: (a) Crank-Nicholson, N= 64,At = 0.01; (b) explicit second-order Adams-Bashforth, N= 64, At =1.2 × 10-3, 3.1 x 10-4; (c) linear propagator, N= 64, At = 6.3 x 10-4,1.6x 10-4; (d) C-N, N=256, At=0.01; (e) explicit A-B, N=256, At=7.8 x 10-5, 3.9 × 10 -5; (f) lin. prop., N= 256, At, At = 9.8 x 10-6.


FLOWS WITH SURFACE TENSION 331factor method. The stability region of this scheme does notdiffer significantly <strong>from</strong> that of <strong>the</strong> explicit scheme. This issomewhat surprising but confirmed by near equilibriumanalysis of <strong>the</strong> second-order linear propagator scheme.O<strong>the</strong>r time discretizations, such as fourth-order Runge-Kutta, for <strong>the</strong> linear propagator method are currently beingstudied.By virtue of its stability, <strong>the</strong> Crank-Nicholson method isabout 250 times faster than both <strong>the</strong> explicit and linearpropagator methods <strong>with</strong> N = 256. Calculations <strong>with</strong> up to8192 points will be presented for <strong>the</strong> Crank-Nicholsonmethod. Following this scaling, <strong>the</strong> explicit Adams-Bashforth and <strong>the</strong> linear propagator Would require a timestep 4 x 105 times smaller.7.2.2. Long-Time ComputationThe long-time evolution of <strong>the</strong> initial data (95) <strong>with</strong>S=0.005 is now studied, using <strong>the</strong> Crank-Nicholsonscheme, toge<strong>the</strong>r <strong>with</strong> filtering. The effect of <strong>the</strong> filtering isdiscussed in detail below.In Fig. 11, a time sequence of interface positions is given,starting <strong>from</strong> <strong>the</strong> initial condition. The calculation usesN= 1024 and/It = 1.25 x 10 -4. At early times, 0 ~< t < 0.40,<strong>the</strong> interface steepens and behaves similarly to <strong>the</strong> zero surfatetension case. It passes smoothly through <strong>the</strong> S--0Kelvin-Helmholtz singularity time (t ~0.375). At t=0.45,<strong>the</strong> interface becomes vertical at <strong>the</strong> center. It subsequentlyrolls over and produces two fingers (t ~ 0.50). These grow inlength in <strong>the</strong> sheet-wise direction (box (b)). The tips of <strong>the</strong>fingers broaden and roll <strong>with</strong> <strong>the</strong> sheet. This is clearly seenat t = 0.80 and <strong>the</strong>re is evidence of capillary waves, seen aswiggles along <strong>the</strong> sheet. By t = 1.20, <strong>the</strong> sheet producesano<strong>the</strong>r turn and <strong>the</strong> fingers have become broader andlarger. Additional capillary waves also have been producedand <strong>the</strong>y traverse <strong>the</strong> interface outwards <strong>from</strong> <strong>the</strong> centerregion. This is a dispersive effect of <strong>the</strong> surface tension thatis seen more clearly in plots of <strong>the</strong> curvature and vortexsheet strength. Note that <strong>the</strong> part of <strong>the</strong> interface (on <strong>the</strong>inner turn) closest to <strong>the</strong> fingers has become quite flat andbends very slightly towards <strong>the</strong> fingers. At later times, thispart bends even more towards <strong>the</strong> fingers, <strong>the</strong> tips of <strong>the</strong>fingers narrow and both pieces of <strong>the</strong> interface approacheach o<strong>the</strong>r. At t--1.40, <strong>the</strong> interface appears to self-intersect,but a close-up of <strong>the</strong> region at this time indicates that<strong>the</strong>re is still a finite distance between <strong>the</strong> upper finger and<strong>the</strong> inner roll. The same is true for <strong>the</strong> lower finger by symmetry,although that symmetry is not imposed in <strong>the</strong>simulation. At this time, <strong>the</strong> gap is but five grid lengths wide,and <strong>the</strong> calculation is stopped here. The length of <strong>the</strong> interfacehas more than doubled (a factor of 2.6).This type of self-approach is quite different <strong>from</strong> that seenin <strong>the</strong> Hele-Shaw flow. Here, each part of <strong>the</strong> interfaceappears to form a finite angle during <strong>the</strong> approach, whereasfor Hele-Shaw, <strong>the</strong> approach is tangential. The mechanismdriving <strong>the</strong> interface to self-approach is not yet well understoodand is under active study. Ra<strong>the</strong>r than ultimatelypreventing <strong>the</strong> pinching, as is apparently <strong>the</strong> case in<strong>the</strong> Hele-Shaw jet flow, surface tension appears to push <strong>the</strong>flow fur<strong>the</strong>r towards it by producing a jet through <strong>the</strong>narrow gaps between <strong>the</strong> fingers and <strong>the</strong> inner turn of<strong>the</strong> spiral. Fluid streams into each of <strong>the</strong> fingers and <strong>the</strong>°i[-0.2 t-0.2(a)0 0.2 0.4 0.6 0.8T=0(c)(b)0 0.2 0.4 0.6 0.8T=0.60(d)0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8T=0.80T=I.20(e)(f)0.I(a)4-!f-4 ' '0 0.2 0.4 0.6 0.8T=0.0(c)4 ' ' '0 0.2 0.4 0.6 0.8T=0.80(¢)44(b)-4 , ,0 0.2 0.4 0.6 0.8T~.~(d)4-i0 0.2 0.4 0.6 0.8T=1.20(t)0 0.2 0.4 0.6 0.8 1 0.4 0.45 0.5 0.55T=l.40T=l.40FIG. 11. Inertial vortex sheet; sequence of interface positions:S= 0.005, N= 1024, z/t= 1.25 x 10-4: (a) t=0; (b) t =0.60; (c) t=0.80;(d) t = 1.20; (e) t = 1.40; (f) close-up of top pinching region, t = 1.40.- top-40 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8T= 1.40 T= 1.40FIG. 12. Inertial vortex sheet: sequence ofy vs. ~t; S = 0.005, N= 1024,dt=1.25×10-4: (a) t=0; (b) t=0.60; (c) t=0.80; (d) t=l.20;(e) t = 1.40; (f) strength on opposite pieces of <strong>the</strong> upper pinching region,t = 1.40.


332 HOU, LOWENGRUB, AND SHELLEYtwo pieces of <strong>the</strong> interface are drawn toge<strong>the</strong>rl This jet is 0.2recognized in that <strong>the</strong> approaching parts of <strong>the</strong> interfacecome <strong>with</strong> oppositely signed vortex sheet strengths, and<strong>the</strong>ir signs are such that fluid flows into <strong>the</strong> fingers. For thisinitial data (single-signed sheet strength), such a sign 0change in <strong>the</strong> vortex sheet strength can only occur in <strong>the</strong>presence of surface tension.Figure 12 shows <strong>the</strong> unnormalized vortex sheet strength,~, versus 0~. Recall that y, initially, is not identically equal to1 as it has been rescaled appropriately for <strong>the</strong> 0- L frame.In this frame, <strong>the</strong> unnormalized vortex sheet strength differs<strong>from</strong> <strong>the</strong> true vortex sheet strength (i.e., <strong>the</strong> tangential jump 0in velocity) by only <strong>the</strong> time-dependent scale factor L/2zc. Atearly times, two positive peaks form in <strong>the</strong> vortex sheet 20ostrength (see box (b)). These are <strong>the</strong> remnants of <strong>the</strong> S = 0,Kelvin-Helmholtz instability, which concentrates vorticity 0into <strong>the</strong> center. In <strong>the</strong> S = 0 case, <strong>the</strong>re is only one peak[31] but <strong>the</strong> addition of a small surface tension splits it in -20%two. There are small waves that form at <strong>the</strong> outer edges of<strong>the</strong>se peaks and are presumably dispersive waves producedby <strong>the</strong> surface tension saturation of <strong>the</strong> Kelvin-Helmholtzinstability. At t = 0.80 (box (e)), <strong>the</strong> peaks have saturatedand more waves have been produced. These disperseoutward along <strong>the</strong> interface. The strength ? has also formeddownward peaks at <strong>the</strong> edge of <strong>the</strong> wave packet. The saturationand dispersion continues through t--1.20 (box (f)).However, when <strong>the</strong> interface begins to self-approach, <strong>the</strong>vortex sheet strength refocusses, forming <strong>the</strong> jet. By <strong>the</strong> finaltime t = 1.40 (box (e)), two pairs of positive and negativepeaks of vortex sheet strength have formed in each pinchingregion. These peaks have been isolated for <strong>the</strong> top pinchingregion in box (f). The top of <strong>the</strong> pinching region (inner turnof <strong>the</strong> spiral) comes <strong>with</strong> a negative signed vortex sheetstrength and <strong>the</strong> bottom comes <strong>with</strong> a positive sign. Thisimplies that fluid is streaming through <strong>the</strong> gap towards <strong>the</strong>center and into <strong>the</strong> downward pointing finger.Saturation and refocussing is also observed in <strong>the</strong> curvature.Its evolution is plotted in Fig. 13. The first graphshows <strong>the</strong> inverse maximum of <strong>the</strong> curvature (in absolutevalue) as a function of time. There is an initial region ofrapid growth in <strong>the</strong> curvature (decrease in <strong>the</strong> plot) due to<strong>the</strong> S = 0 Kelvin-Helmholtz instability. Recall that for <strong>the</strong>Kelvin-Helmholtz singularity, <strong>the</strong> curvature diverges. Here,<strong>the</strong> curvature saturates and eventually refocusses. By <strong>the</strong>final time t = 1.40, <strong>the</strong> maximum of <strong>the</strong> curvature nearly ~reaches that attained during <strong>the</strong> initial period of growth. 4The refocussing, however, occurs at different places along<strong>the</strong> interface as seen in <strong>the</strong> o<strong>the</strong>r plots of t¢ versus ~. Disper- 2sive waves are also generated and move outwards along <strong>the</strong>interface. At t = 1.40, <strong>the</strong> pinching points are indicated bypairs of like-signed peaks in <strong>the</strong> curvature.Evidence is now presented to show that <strong>the</strong> pinchingtakes place in a finite time. Maintaining resolution is criticaland so <strong>the</strong> neck widths obtained <strong>from</strong> different spatial and0.I ~ , ~i1c8Inverse Curvanu~, (a)0.5 ITCurvature, (c)-200 . . . .0.2 0.4 0.6 0.8T=0.60Curvature, (e)J ~0.2 0.4 0.6 0.8T=l.201.5Curvature, (b)-200~ 0.2 0.4 0.6 0.8T=0.40Curvature, (d)-2000 0.2 0.4 0.6 0.8T=0.800Curvature, (O2o00 o~ o-q" o-g o-~T=I.40FIG. 13. Inertial vortex sheet: sequence ofx vs. a; S = 0.005, N= 1024,A t = 1.25 x 10 - 4: (a) plot of inverse maximum curvature (absolute value ):(b) plot of curvature, t=0.40; (c) t=0.60; (d) t=0.80; (e) t= 1.20;(f) t= 1.40.temporal resolutions are compared. This is shown in <strong>the</strong> topgraph of Fig. 14, which shows minimum width in <strong>the</strong> toppinching region, graphed as a function of time starting att = 1.40. As in <strong>the</strong> Hele-Shaw case, <strong>the</strong> minimum width iscomputed by minimizing <strong>the</strong> distance function betweeneach piece of <strong>the</strong> interface and using <strong>the</strong> Fourier interpolant1.4x 10 -3Minimum Distance to Pinching, (a).. 1024-- 2048-.- 4096- 81921.405 1.41 1.415 1.42 1.425 1.43 1.435TNumber of Accurate Digits in Energy, (b)- 2048 ............................... - ._ '.~.. 1024-- 5120.2 0.4 0.6 0.8 I 1.2 1.4TFIG. 14. Pinching in, and accuracy in <strong>the</strong> energy of, <strong>the</strong> inertial vortexsheet: (a) minimum width of <strong>the</strong> top pinching region: S= 0.005; N= 1024<strong>with</strong> At = 2.5 x I0-4; N = 2048 <strong>with</strong> At = 1.25 × I0-4; N= 4096 <strong>with</strong> At =6.25 x 10-s; N=8192 <strong>with</strong> At=3.125 x 10 -5. (b) Number of accuratedigits in <strong>the</strong> energy: S=0.005; N=512, N= 1024, N=2048 <strong>with</strong> At=1.25 × 10-4, and N= 1024 <strong>with</strong> At = 3.125 × 10 -5 (1024ext).


FLOWS WITH SURFACE TENSION 333of <strong>the</strong> curve. The spatial resolutions shown are N= 1024<strong>with</strong> At=2.5xlO-4, N=2048 <strong>with</strong> At=1.25xlO-4,N=4096 <strong>with</strong> At=6.25 x 10 -5 , and N=8192 <strong>with</strong> At=3.125 x 10-5. The computations <strong>with</strong> N= 2048 and aboveare computed by restarting Fourier interpolants of lowerresolutions. In particular, <strong>the</strong> N= 2048 computation isrestarted <strong>from</strong> <strong>the</strong> N= 1024 computation at t = 1.25. TheN=4096 and N= 8192 computations are restarted <strong>from</strong>N = 2048 at t = 1.40. These are times at which <strong>the</strong> active spatialspectrum of <strong>the</strong> lower resolution calculations begin toapproach <strong>the</strong> Nyquist frequency. Additional details of <strong>the</strong>secalculations will appear in a separate study [ 28 ].In Fig. 14a, it is seen that <strong>the</strong> time of pinching is adecreasing function of resolution and strongly suggests thatpinching occurs at a finite time. Moreover, <strong>the</strong> width seemsto vanish <strong>with</strong> an infinite slope, implying that <strong>the</strong> pinchingrate intensifies as <strong>the</strong> width narrows. Again, fur<strong>the</strong>r detailswill appear in [ 28 ].This apparent singularity is of a completely different typethan that of <strong>the</strong> S = 0 Kelvin-Helmholtz singularity. Here,<strong>the</strong> singularity is a topological singularity and may signal animminent change in <strong>the</strong> topology of <strong>the</strong> flow. The curvature,vortex sheet strength, and fluid velocity all apparentlydiverge at <strong>the</strong> pinching time [28]. In <strong>the</strong> S = 0 case, it is <strong>the</strong>velocity gradient (strain rate) and <strong>the</strong> curvature thatdiverge, but not 7- More fundamentally, <strong>with</strong> S = 0 <strong>the</strong>singularity occurs through a rapid compression of vorticityalong <strong>the</strong> sheet at a single isolated point [33, 31, 9, 44].Here, <strong>the</strong> divergence occurs through a rapid production ofvorticity that is associated <strong>with</strong> <strong>the</strong> pinching. In a separatestudy, we will show evidence this collapse is a localphenomenon in <strong>the</strong> flow and satisfies a scalingbehavior [28]. We have also derived localized equationsand <strong>the</strong>y are currently under investigation.Error Analysis and <strong>the</strong> Role of Filtering. We now turn toan analysis of <strong>the</strong> errors in our scheme, first by considering<strong>the</strong> energy E(t), which is a conserved quantity not conservedby <strong>the</strong> discrete methods. The energy E is <strong>the</strong> sum of <strong>the</strong> perturbationkinetic and <strong>the</strong> interfacial energies, given bywhereis <strong>the</strong> interfacial energyE( t) = Es( t) + Ek( t), (96)Es(t) = S(L - 1 ) (97)1Ek(t) = ½ fo ~,(o~', t) ~(od, t) d~' (98)is <strong>the</strong> perturbation kinetic energy and ~k is <strong>the</strong> stream function,ff(~, t) = - 2-~ Re {f2 Y(c(, t)× log[sin rc(z(oq t)- z(~', t))] do(}. (99)The formula for ff can be rewritten, by explicitly subtractingoff <strong>the</strong> logarithmic singularity and integrating by parts, toyield an expression that can be computed <strong>with</strong> spectralaccuracy numerically. See Pullin [38] and Baker andNachbin [ 7 ] for details.The lower graph of Fig. 14 shows -lOgl0 [(E(t)-E(O) )/E(O ) I for several spatial resolutions. This is essentiallynumber of accurate digits in <strong>the</strong> energy. The time step isAt= 1.25 x 10 -4 for all <strong>the</strong> resolutions except <strong>the</strong> onemarked 1024ext. This uses At= 3.125 x 10 -5 and uses asinitial data <strong>the</strong> o<strong>the</strong>r N= 1024 computation at t = 0.40.Prior to <strong>the</strong> time t = 0.40, <strong>the</strong>re is essentially no differencebetween <strong>the</strong> different resolutions and <strong>the</strong> errors aredominated by <strong>the</strong> time stepping. Over nine digits ofaccuracy are obtained. Around t = 0.42, <strong>the</strong>re is a suddenoverall decrease in accuracy. This is an effect <strong>from</strong> <strong>the</strong> S = 0Kelvin-Helmholtz singularity. This decrease is primarilydue to time stepping since <strong>the</strong> N= 512, 1024, and 2048 curvesall lie on top of one ano<strong>the</strong>r immediately afterwards.This is reinforced by <strong>the</strong> 1024ext curve which has a fourtimes smaller time step and lies well above <strong>the</strong> o<strong>the</strong>rs. Fromt = 0.43 until nearly <strong>the</strong> end of <strong>the</strong> computation, <strong>the</strong> energychanges only slowly, and <strong>the</strong> N = 1024 and 2048 computationsare slightly more accurate than <strong>the</strong> N= 512. Aroundt = 1.40, when <strong>the</strong> pinching begins to take place, <strong>the</strong>re isano<strong>the</strong>r sudden loss of accuracy and now N = 2048 is moreaccurate than N= 1024 as spatial resolution becomes veryimportant.In view of <strong>the</strong>se results, our code would certainly benefit<strong>from</strong> adaptive time stepping. A small time step is really onlyneeded near <strong>the</strong> Kelvin-Helmholtz singularity time andnear <strong>the</strong> time of pinching. Larger time steps could likely beused safely for <strong>the</strong> remainder of <strong>the</strong> computation. Finally,we remark that <strong>the</strong> energy of <strong>the</strong> N= 8192 computationused to compute <strong>the</strong> pinching distance still has over sixdigits of accuracy at <strong>the</strong> pinching time [ 28 ].Since <strong>the</strong> energy is an integral quantity, using it tomeasure error is smoo<strong>the</strong>r than using a pointwise error. Asin <strong>the</strong> Hele-Shaw case, we also consider <strong>the</strong> pointwise errorin <strong>the</strong> x-component of <strong>the</strong> interface position. In Fig. 15,-lOgl0(e~t)) is plotted (see <strong>the</strong> Hele-Shaw results for adefinition of eN), <strong>with</strong> <strong>the</strong> N = 2048 calculation again usedas an approximation of <strong>the</strong> exact solution. Several resolutionsare graphed and all use <strong>the</strong> time step At = 1.25 x 10-4.This graph is qualitatively similar to that of <strong>the</strong> energy anddemonstrates <strong>the</strong> spatial convergence of our scheme. Thetemporal convergence can be similarly shown, but is notpresented.


334 HOU, LOWENGRUB, AND SHELLEYle',io L,~8• -r64 ,-.',;iI-1024-.-512--256t-J . . . . . . . . . . . . . . . . . . . _.,ItI2 012 0'.4 0'.6 018 ; 1;~ 1.4TFIG. 15. Error in x-coordinate; S=0.005, At= 1.25 x 10 -4, exactsolution approximated by N = 2048.The role of Fourier filtering (H in Eq. (86)) is nowdiscussed. As <strong>the</strong> higher Fourier modes rise above <strong>the</strong>round-off level (due to nonlinear interaction), an aliasinginstability appears. This instability can be controlled byei<strong>the</strong>r increasing <strong>the</strong> resolution or by using some kind offiltering. Both have been tried, and we conclude that it ismore practical to use a high order Fourier filtering thatsmooths <strong>the</strong> highest modes. By this, it is possible to achieve<strong>the</strong> same level of accuracy, at late times, using many fewerpoints than would be required <strong>with</strong>out it. This is seen mostdirectly by comparing <strong>the</strong>ir energies.Figure 16 compares <strong>the</strong> energies of calculations both <strong>with</strong>and <strong>with</strong>out filtering on <strong>the</strong> time interval 0.4 to 0.6. Recall-'xthat <strong>the</strong> filtering combines <strong>the</strong> Fourier and Krasny filters.The effect of Krasny filtering is discussed below. This is avery good test, as <strong>the</strong> spectrum is broadly populated due to<strong>the</strong> effect of <strong>the</strong> S= 0 singularity. The time step is At =1.25 x 10-4 for all <strong>the</strong> calculations. The filtered calculationsshow a generally slower loss of accuracy and are moreaccurate than <strong>the</strong>ir unfiltered counterparts by t = 0.46. Infact, at t = 0.46, <strong>the</strong> N= 256 filtered calculation is moreaccurate than <strong>the</strong> unfiltered N= 1024 computation! Theunfiltered computations both break down before t = 0.6.The aliasing instability can be observed directly in spectralplots of <strong>the</strong> numerical solution. In Figs. 17a, b areshown <strong>the</strong> y-spectra of <strong>the</strong> filtered and unfiltered computations,respectively. In Fig. 17b, <strong>the</strong> aliasing instability isclearly seen through <strong>the</strong> artificial rise in <strong>the</strong> high wavenumberspectrum at t = 0.46. The unfiltered calculation cannotbe continued long after this time, as this rise causes <strong>the</strong> codeto blow up. The filtered computation, on <strong>the</strong> o<strong>the</strong>r hand,does not exhibit such a rise. Moreover, it is really only <strong>the</strong>last few modes that are damped by <strong>the</strong> filtering (modes450-512). Since Krasny filtering is also employed in <strong>the</strong>filtering method, <strong>the</strong> modes at <strong>the</strong> roundoff level 10 -]6 in<strong>the</strong> unsmoo<strong>the</strong>d calculation are set to zero in <strong>the</strong> smoo<strong>the</strong>done. They are seen at <strong>the</strong> 10 -19 level in <strong>the</strong> plot (this kept<strong>the</strong> logarithm function in <strong>the</strong> graphics package <strong>from</strong>diverging). Moreover, it is clear <strong>from</strong> <strong>the</strong> behavior of <strong>the</strong>spectrum of <strong>the</strong> unfiltered computation that roundoff errorsgrow ra<strong>the</strong>r slowly and, <strong>the</strong>refore, <strong>the</strong> main effect of Krasnyfiltering is to keep <strong>the</strong> spectrum clean and is not necessaryfor stability.The spectrum also indicates a very interesting behavior.At <strong>the</strong> early times 0 ~< t ~< 0.46, energy is rapidly moved to<strong>the</strong> high wavenumbers due to <strong>the</strong> near Kelvin-Helrnholtz108:~ 64Number of Aecttrate Digits in Energy, (a)256-- unfiltered20.4 0.42 0,44 0.46 0.48 0.5 0.52T(b)108:g 64"" 10245121024256-- unfiltered- filtered0.54 0.56 0.5820.4 0.42 0.44 0.46 0.48 0.5 0.52 0.54 0.56 0.58TFIG. 16. Accuracy in energy: filtered vs unfiltered, S= 0.005, At =1.25x 10-4: (a) filtered vs unfiltered, N=256, 512, 1024; (b) filtered(N= 512) vs unfiltered (N= 1024).0Filtered C-N, (a)loga01#(k)l- I C 1 ~ - -t=0,42 ~ t=0.46"2G0 100 200 300 400 500 600kUnfiltered C-N, (b)0log10lP(k)l- 1 0 ~ ~---------~ t=0.46 tlogloJ~'(k)l-20 0 100 200 300 400 500 600kFiltered C-N, Later Times, (c)p=1.20 r-20 0 100 200 300 400 500 600kFIG. 17. Log plot of power spectra (log [~(k)l ) at different times,S = 0.005, At = 1.25 x 10 -4, N= 1024: (a) filtered Crank-Nicholson, shorttime; (b) unfiltered C-N, short time; (c) filtered C-N, long time.


FLOWS WITH SURFACE TENSION 335singularity, and at t--0.46 all <strong>the</strong> modes lie above <strong>the</strong>round-off. Then, as seen in Fig. 17c, energy begins to move<strong>from</strong> high wavenumbes back to low wavenumbers. Forexample, <strong>the</strong> highest nonzero mode at t--0.60 is aboutk = 450 and by time t = 1.20, it moves to about k = 360. Thisbehavior reflects <strong>the</strong> saturation observed in 7 and x.Between t = 1.20 and t = 1.40, <strong>the</strong> energy resumes its generationat high wavenumbers (refocussing) and by t = 1.40, allmodes again lie above <strong>the</strong> round-off.7.3. Boussinesq InterfacesIn this section, a preliminary long time computation ofan interface in an unstably stratified Boussinesq flow ispresented. In Boussinesq flow, <strong>the</strong> fluids are only weaklystratified across <strong>the</strong> interface and <strong>the</strong> only contribution tovorticity production is through <strong>the</strong> buoyancy term. Weremark that Boussinesq interface flows in <strong>the</strong> absence ofsurface tension develop finite time singularities of a formwell described by those for inertial vortex sheets <strong>with</strong>outsurface tension [ 37, 5 ]. The 7 evolution equation isV,-~3~((T-W.s)7/s~)=Sx~-2Ry~. (100)The term Q in <strong>the</strong> small scale decomposition (56) is replacedb'y Q-2Rye. The Crank-Nicholson scheme is used <strong>with</strong>filtering for <strong>the</strong> same initial data in x and y as in <strong>the</strong> inertialvortex sheet calculation. Initially, 7 = 0. Fur<strong>the</strong>r, R = -- 5 sothat <strong>the</strong> flow is unstably stratfied, and <strong>the</strong>re are 19 linearlyunstable modes.Figure 18 shows <strong>the</strong> time evolution of two periods of <strong>the</strong>0.5.0.05 - -(a)0 0.5 1t=01.5 2(c)0.5(b)0 . 5 ~ 0-0.50 0.5 1 1.5T=0.60(d)0.501-0.5 0 0.5 1 1.5 2 0 0.5 1 1.5T=0.70T=0.80(e)(f)0.5 0.10 0.5 1 1.5 2 0.15 0.2 0.25 0.3T--0.89T=0.89FIG. 18. Evolution ofa Boussinesq interface; S = 0.005, R = - 5, At =3.125x 10 -5 , N= 1024: (a) t=0; (b) t=0.60; (c) t=0.70; (d) t=0.80;(e) t = 0.89; (f) close-up of a left-most spiral, t = 0.89.0.35interface. This computation uses N=1024 and At=3.125× 10 5. The classical Rayleigh-Taylor structure isseen as two spirals form on ei<strong>the</strong>r side of <strong>the</strong> fluid plumes.The spirals are smaller than those for <strong>the</strong> inertial vortexsheet, but an examination of <strong>the</strong>ir inner structure shows<strong>the</strong>m to be very similar. In particular, note <strong>the</strong> inward bulgein <strong>the</strong> second turn of <strong>the</strong> interface (t = 0.89, close-up). Itseems very plausible that <strong>the</strong>re will again be finite timepinching. One consequence of <strong>the</strong>se smaller spirals is thatthis flow is more difficult to resolve than <strong>the</strong> previous caseof inertial vortex sheets. More highly resolved computationswill be presented in a separate study [28].An intriguing question is whe<strong>the</strong>r such "bubble" formation,as is apparently predicted by <strong>the</strong>se calculations, isobservable in an experimental setting. Of course viscosity(and dimensionality) are crucial effects as <strong>the</strong> singularity isapproached. But in his experimental study of <strong>the</strong> developmentof instability in sharply stratified shear flow betweentwo immiscible fluids, Thorpe [ 50 ] remarks that <strong>the</strong> interfacebetween <strong>the</strong> two fluids "became very irregular, sometimesbeing broken and drops of one fluid being producedin <strong>the</strong> o<strong>the</strong>r,...".8. CONCLUSIONSIn this paper, we present new numerical methods forcomputing <strong>the</strong> motion of a fluid interface <strong>with</strong> surface tensionin a two-dimensional, irrotational, and incompressiblefluid. These methods have no high order stability constraint,are effectively explicit, and have high spatial accuracy. Ourmethods are applied to computing <strong>the</strong> motion, in 2D, ofinterfaces in Hele-Shaw flows (viscously dominated) and to<strong>the</strong> motion of free surfaces in inviscid, inertially dominatedflows governed by <strong>the</strong> Euler equations.In <strong>the</strong> former case, we compute <strong>the</strong> long-time developmentof an expanding gas bubble and a complicated"mixing zone" by <strong>the</strong> interpenetration of a heavy fluid <strong>with</strong>a lighter one. We show <strong>the</strong> robustness and accuracy of <strong>the</strong>method, even in <strong>the</strong> presence of ramified spatial structure. In<strong>the</strong> latter case, we compute <strong>the</strong> long-time evolution of avortex sheet: The simulations suggest that in <strong>the</strong> process offorming a spiral, <strong>the</strong> turns of <strong>the</strong> spiral collide <strong>with</strong> oneano<strong>the</strong>r, creating trapped bubbles of fluid. The results alsoindicate that this behavior is strongly tied to <strong>the</strong> presence of<strong>the</strong> surface tension, which allows <strong>the</strong> creation of localcirculation along <strong>the</strong> sheet. To our knowledge, such asingularity has not been previously observed in such a flowin 2D.These methods are finding application in problemsbeyond those discussed in this paper. To study <strong>the</strong> "quasistatic"limit of dendritic crystal growth, Almgren, Dai,and Hakim [ 1 ] have recently applied our methods tocomputing <strong>the</strong> expansion of a Hele-Shaw bubble <strong>with</strong>anisotropic surface tension. By being able to compute for


336 HOU, LOWENGRUB, AND SHELLEYlong-times <strong>with</strong> good resolution and moderate time steps,<strong>the</strong>y have discovered new scaling behaviors in <strong>the</strong> expansionand propagation of <strong>the</strong> large scale fingers• In collaboration<strong>with</strong> A. Greenbaum and D. Meiron, we have impleted <strong>the</strong>second-order linear propagator method, toge<strong>the</strong>r <strong>with</strong> fastsummation methods, to compute <strong>the</strong> many inclusionOstwald ripening problem. Preliminary results indicate that<strong>the</strong> linear propagator method is substantially faster than anearlier implementation that used a standard ODE solver forstiff systems.APPENDIX 1: SMALL SCALE DECOMPOSITION OFINTEGRAL EQUATIONIn this appendix, it shown that at small scales, <strong>the</strong> integralequations for Hele-Shaw interfaces and inertial vortexsheets yield explicit relations for y and Yt, respectively. Bothintegral equations can be written as~(a, t)+ K[~](ot, t)=f(a, t), (101)where ~ is <strong>the</strong> funcion to be determined (i.e., 7 or 7t) andfcontains all <strong>the</strong> terms not involving ~. K is <strong>the</strong> integraloperator and is given byf+ct)K[~](ot, t)= -2A ~(a', t)--ooxRe 2hi z(~,t)-z(ot',t) do~', (102)where A=Au (Hele-Shaw) or Ap (inertial sheets) and[A[ ~< 1 in ei<strong>the</strong>r case. The integral equation (101) is aFredholm integral equation of <strong>the</strong> second kind and it is wellknown that it is invertible and has a bounded inverse on anopen and periodic geometry if z is smooth in 0~ (see [ 6 ] ).Moreover, if <strong>the</strong> components of <strong>the</strong> complex interfaceposition z satisfy <strong>the</strong> assumptions of Theorem 1 (Section 4),<strong>the</strong>n it is easy to see that <strong>the</strong> integral kernelz( Zz(ot,)j(103)is a smooth function. Thus, K is a smoothing operator. Asin Theorem 1, this means that for any ~k e L 2 <strong>the</strong> highwavenumber behavior of K is given by/~[ ~k ] (k) = O(e -P Ikt~(k)), (104)where p is <strong>the</strong> assumed strip width of analyticity of z in ot.Consider <strong>the</strong> integral equation (101). It can be invertedformally to give~ = (1+ K)-lf, (105)where I is <strong>the</strong> identity operator. Moreover, <strong>the</strong> boundednessof <strong>the</strong> inverse implies that if f eL 2 <strong>the</strong>n so is (I+K)-lfFur<strong>the</strong>r, Eq. (101) can be rewritten as~ =f--K[~] =f-K[(I+K)-~f] (106)by using (105) only on <strong>the</strong> right-hand side. Finally, using(104) <strong>with</strong> ~, = (I + K) - if, gives~f (107)in <strong>the</strong> notation of section 4 and thus, an explicit expressionat small scales.Remark. Actually, invertibility of <strong>the</strong> integral equationis not strictly necessary. It is only important that on itsrange, <strong>the</strong> inverse is bounded. In that case, if <strong>the</strong> Fredholmalternative is satisfied by <strong>the</strong> right-hand side, f, <strong>the</strong> aboveargument can <strong>the</strong>n be repeated. Such singular integral equationsappear in harmonic problems on multiply connecteddomains.APPENDIX 2: OTHER CHOICES OF TThe equal arclength frame, given in Section 5.2., isactually a special case of two o<strong>the</strong>r more general choices offrame:(1) Scaled initial parametrization frame. In <strong>the</strong> first,Eq. (57) is replaced <strong>with</strong> <strong>the</strong> requirement that1s~(ot, t) = ~n L(t) R(ct),t~ 2n<strong>with</strong> Jo R(o~')dot'=2n. (108)- R(ot) 2n f~- s~, dot'. (109)This constraint is dynamically satisfied by choosing T to beT(o~, t) = T(O, t) +f;0~, Udot'2ng- . R(~') dot' • O~,Udod. (llO)ZT~ JO fOThis is easily seen by differentiating Eq. (109) <strong>with</strong> respectto time and using Eq. (51 ). This choice of frame implies thatan initial parametrization of <strong>the</strong> curve is preserved, up to atime-dependent scaling. If it is known where complexity of<strong>the</strong> curve will develop, <strong>the</strong>n a finer grid could be placedinitially in this area. The equal arclength frame follows <strong>from</strong>Eq. (108) by setting R = 1.


FLOWS WITH SURFACE TENSION 337(2) Complexity Measuring Frame. In <strong>the</strong> second choice,one defines a strictly positive, smooth functionf(a, t), whichmeasures <strong>the</strong> complexity of <strong>the</strong> curve. For example,[= 1 + x 2. Now, <strong>the</strong> product s~f is required to be equaleverywhere to its mean. That is,1 f2n Is~(o~, t) f(oq t) =~-~ j0 s~,(o~, t) f(og, t) do(1 L= 2-~ fo f(s, t)ds. (111)Thus, where f is large, s~ (<strong>the</strong> infinitesimal grid spacing) issmall, and vice versa. For general f, it is not possible toobtain an explicit expression for <strong>the</strong> tangential velocityneeded to dynamically enforce condition (111 ). However, itturns out that if f= C(K), <strong>the</strong>n an explicit expression isobtained. Previously, curvature constraints have been usedsuccessfully to statically remesh interfaces (see [9, 27], forexample)• However, <strong>the</strong> choice of T to dynamically enforce(111 ) is given byT(ct, t)= Cl~=0 • T(0, t)C(112)and <strong>the</strong> equal arclength frame again follows by settingC=I.More general reference frames (i.e., more general choicesfor f) may be useful, however. An important example where<strong>the</strong>y may be needed is in <strong>the</strong> case of anisotropic surfacetension [24, 1]. There, <strong>the</strong> surface tension coefficient isdirectionally dependent and so <strong>the</strong>re are preferred directionsof growth. It is possible to choose a reference framethat dynamically equidistributes grid points along <strong>the</strong> interfaceaccording to a measure of <strong>the</strong> anisotropic surfaceenergy• In this case, <strong>the</strong> appropriate tangential velocity isfound by solving a linear integro-differential equation• Thisis currently under study.ACKNOWLEDGMENTSWe thank Russ Caflisch, Anne Greenbaum, Leslie Greengard, JoeKeller, Mike Ward, and especially Dan Meiron and Ray Goldstein, forimportant and stimulating conservations. All three of <strong>the</strong> authors thank <strong>the</strong>hospitality and support of <strong>the</strong> Institute for Advanced Study where <strong>the</strong>yparticipated in <strong>the</strong> special year in fluid mechanics during <strong>the</strong> 1991 academicyear, where some of <strong>the</strong> initial work of this paper was done. There, <strong>the</strong>ywere supported by NSF Grant DMS-9100383 and by Department of <strong>the</strong>Air Force Grant F-49620-92-J-0023F. T.Y.H. acknowledges partial support<strong>from</strong> NSF Grant DMS-9003202 and Air Force Office of ScientificResearch Grant AFOSR-90-0090. J.S.L. acknowledges <strong>the</strong> support of <strong>the</strong>National Science Foundation through a Ma<strong>the</strong>matical Sciences PostdoctoralFellowship, <strong>the</strong> University of Minnesota Army High PerformanceComputing Research Center, <strong>the</strong> Minnesota Supercomputer Center, wheresome of <strong>the</strong> data analysis was performed, and <strong>the</strong> University of MinnesotaGeometry Center, where a movie was made of <strong>the</strong> pinching inertial vortexsheet. M.J.S. acknowledges partial support <strong>from</strong> DOE Grant DE-FG02-88ER25053 and <strong>the</strong> National Science Foundation through a PresidentialYoung Investigator Grant DMS-9396403.REFERENCES1. R. Almgren, W.-S. Dai, and V. Hakim, Phys. Rev. Lett. 71, 3461 (1993).2. W. F. Ames, Numerical Methods for Partial Differential Equations(Academic press, New York, 1977).3. C. Anderson, J. Comput. Phys. 62, 111 (1986).4. G. Baker, "Generalized Vortex Methods for Free-<strong>Surface</strong> <strong>Flows</strong>," inWaves on Fluid Interfaces, edited by R. E. Meyer (Academic Press,New York/London, 1983).5. G. Baker, R. Caflisch, and M. Siegel, J. FluidMech. 252, 51 (1993).6. G. Baker, D. Meiron, and S. Orszag, J. FluidMech. 123, 477 (1982).7. G. Baker and A. Nachbin, preprint.8. G. R. Baker and M. J. Shelley, J. Comput. Phys. 64, 112 (1986).9. G. R. Baker and M. J. Shelley, J. FluidMech. 215, 161 (1990).10. J. T. Beale, T. Y. Hou, and J. S. Lowengrub, "On <strong>the</strong> Well-Posednessof Two Fluid Interracial <strong>Flows</strong> <strong>with</strong> <strong>Surface</strong> <strong>Tension</strong>," in Singularitiesin Fluids, Plasmas and Optics, edited by R. C. Caflisch and G. C.Papanicolaou (Kluwer Academic, London, 1993).11. J. T. Beale, T. Y. Hou, and J. S. Lowengrub, Commun. Pure AppkMath., to appear.12. J. T. Beale, T. Y. Hou, and J. S. Lowengrub, in preparation.13. G. Birkhoff, Proc. Syrup. Appl. Math. 13, 55 (1962).14. G. Carrier, M. Krook, and C. Pearson, Functions of a ComplexVariable (McGraw-Hill, New York, 1966).15. W. Dai, L. Kadanoff, and S. Zhou, Phys. Rev. A 43, 6672 (1991).16. W.-S. Dai and M. J. Shelley, Phys. Fluids A 5, 2131 (1993).17. A. J. DeGregoria and L. W. Schwartz, J. FluidMech. 164, 383 (1986).18. R. Goldstein, A. Pesci, and M. Shelley, Phys. Rev. Lett. 70, 3043 (1993).19. R. Goldstein, A. Pesci, and M. Shelley, in preparation.20. R. Goldstein and D. M. Petrich, Phys. Rev. Lett. 67, 3203 (1991).21. J. Goodman, T. Y. Hou, and E. Tadmor, Numer. Math., to appear.22. A. Greenbaum, L. Greengard, and G. B. McFadden, J. Comput. Phys.105, 267 (1993).23. L. Greengard and V. Rokhlin, J. Comput. Phys. 73, 325 (1987).24. M. Gurtin, Thermomechanics of Evolving Phase Boundaries in <strong>the</strong> Plane(Clarendon Press, Oxford, 1993).25. D. Gottlieb and S. A. Orszag, Numerical Analysis of Spectral Methods:Theory and Aplications, CBMS-NSF Regional Conference Series inApplied Ma<strong>the</strong>matics, Vol. 26 (SIAM, Philadelphia, 1977).26. T. Halsey and M. Leibig, Phys. Rev. A 46, 7793 (1992).27. E. Meiburg and G. Homsy, Phys. Fluids 31,429 (1988).28. T. Hou, J. Lowengrub, and M. Shelley, in preparation.29. J. M. Hyman and M. J. Naughton, in Proc. SIAM-AMS Conf. onLarge Computation in Fluid Mechanics (SIAM, Philadelphia, 1984).30. D. Kellogg, Foundations of Potential Theory (Dover, New York, 1929).


338 HOU, LOWENGRUB, AND SHELLEY31. R. Krasny, J. FluidMech. 167, 65 (1986).32. H.-O. Kreiss and J. Ofiger, SIAM J. Numer. Anal, 16, 421 (1979).33. D. Moore, Proc. R. Soc. London A 365, 105 (1979).34. W. Mullins, J. Appl. Phys. 27, 900 (1956).35. W. Mullins and R. Sekerka, J. Appl. Phys. 34, 323 (1963).36. A. Prosperetti, L. A. Crum, and H. C. Pumphrey, J. Geophys. Res. 94,3255 (1989).37. D. A. Pugh, Ph.D. <strong>the</strong>sis, Imperial College, London, 1989.38. D. I. Pullin, J. FluidMech. 119, 507 (1982).39. R. Rangel and W. Sirignano, Phys. Fluids 31, 1845 (1988).40. D. Reinelt, J. FluidMech. 183, 219 (1987).41. R. Rogallo, NASA TM-73203 (unpublished).42. Y. Saad and M. Schultz, SIAMJ. ScL Stat. Comput. 7, 856 (1986).43. P. G. Saffman and G. R. Baker, Annu. Rev. FluidMech. 11, 95 (1979).44. M. Shelley, J. FluidMeeh. 244, 493 (1992).45. M. J. Shelley and M. Vinson, Nonlinearity 5, 323 (1992).46. A. Sidi and M. Israeli, J. Sei. Comput. 3; 201 (1988).47. H. E. Stanley and N. Ostrowsky (Eds.), Random Fluctuations andPattern Growth: Experiments and Models (Kluwer Academic, London,1988).48. J. Strain, J. Comput. Phys. 85, 342 (1989).49. E. Tadmor, SIAM Rev. 29, 525 (1987).50. S. Thorpe, J. FluidMeeh. 39, 25 (1969).51. G. Tryggvason and H. Aref, J. FluidMech. 136, 1 (1983).52. P. W. Voorhees, G. B. McFadden, R. F. Boisvert, and D. I. Meiron,Acta Metall. 36, 207 (1988).53. T. A. Witten and L. M. Sander, Phys. Rev. B 27, 5686 (1983).

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!