1 i ' I Geophys. J. Int. (1992) 109, 553-564 Diapirism and topography A. ~oliakovland Y. Podladchikov2 I I Hans Ramberg Tectonic Laboratory, Institute of Geology, Uppsala University, Box 555. 751 22 Uppsala, Sweden Institute of Experimental Mineralogy, Chernogolovka, Moscow District, 142 432, USSR Accepted 1991 December 4. Received 1991 December 4; in original form 1991 April 13 SUMMARY A new numerical technique using markers and a deformable Lagrangian mesh is used t o study the deformation of the surface above a rising diapir. This method allows modelling of a free-surface boundary in a self-consistent way. The method makes it possible to investigate many different problems with non-regular geometry. The codes were verified through comparison with analytical solutions for the initial stages and with other numerical codes for mature stages. Our simulations led to the following conclusions. (1) The growth rate of the diapir is more strongly influenced by the viscosity contrast than the wavelengths. This is due to the strong growth rate difference during the initial stage. (2) The maximum elevation above the diapir linearly increases with increasing wavelength and is approximately the same for different viscosity contrasts. (3) The elevation will be considerably higher for a low-viscosity diapir at a given depth than for an isoviscous diapir . Comparisons for different thicknesses of the buoyant layer showed that the highest topography is produced when the two layers have equal thickness. We showed the dependence of the topographic behaviour on the parameter R (the ratio of the density difference between two layers to the density of the upper layer). It was found that topography behaves linearly up to R = 0.15. It indicates that a posteriori estimation of topography for traditional calculations using the free-slip boundary condition works well within this limit. Key words: diapirism, finite element method, topography. 1 INTRODUCTION Salt and granite diapirs, and mantle plumes can exert a strong influence on the overlying surface topography. Surface topography can then have a feedback effect on the development of diapirs. These interrelated processes have not been explored numerically in a comprehensive manner. A systematic simulation is required, that will yield quantitative information about the temporal and spatial evolution of height, shape and relief above a buoyant body. The results of these simulations, combined with geophysical and geological studies of topography, can then be used to estimate various parameters for natural diapirs such as their viscosity, size, and depth of origin. A new, more sophisticated numerical model is therefore required, in order to directly investigate the interrelated phenomena of diapirism and topography. It is first necessary to briefly review previous studies of topography related to diapirism. Rambeg (1968a, b) derhed analytical formulae for the Rayleigh-Taylor instability for the linear stage of diapiric growth and for the rate of isostatic compensation. He suggested that linear solutions are valid up to amplitudes of 10 times the initial perturbation. An analytical solution for the motion of the surface above a deep rising cylinder or sphere has been obtained by Morgan (1965) using the method of images. Talbot & Jarvis (1984) studied the shape of a diapir extruding onto the surface by using an analytical solution for flow from a point source. One disadvantage of these methods is that it is impossible to consider complicated non-linear interaction between the surface and the rising diapir. Laboratory investigation of the surface dynamics of viscous flows has been conducted using centrifuge techniques by Ramberg (1968a, b) and Jackson et al. (1988). Olson & Nam (1986) and Griffiths et al. (1989) studied these phenomena without a centrifuge using much lower viscosity materials such as oils. Visualization of the surface motion was quite difficult in the centrifuge experiments; the experiments with oils were complicated by surface tension effects. Similar problems have been studied numerically. In most 554 A. Poliakov and Y. Podladchikov cases the surface topography is calculated with a following procedure (e.g. see McKenzie 1977; Fleitout, Froidevaux & Yuen 1986; Lliboutry 1987). The subsurface flow is calculated for a fixed upper boundary (vertical velocity is zero). Topography is then calculated a posteriori from the normal stress along this boundary. The relationship between the surface topography, Ah, and the dynamical normal stress a,,, to first order, is given by Coupling between surface topography and subsurface flow was studied by Myasnikov & Savushkin (1978), Fadeev (1988) and Pak (1989). If the surface topography is small, the problem can be treated as two coupled problems and can be solved iteratively. The technique is based on an asymptotic decomposition in which the ratio of the thickness of the hydrodynamic boundary layer to the characteristic size of the problem is a small parameter. Thus, it is difficult to use this method for problems with large dynamically induced surface topography. Zaleski & Julien (1990) used a top layer with a very low viscosity and density to represent air or water above the surface. This allows a simple representation of the free surface. However, due to the very high viscosity and density contrast and diffusion between the top layer and the underlying layers, calculations sometimes become unstable and give significant errors. In the present study an alternative method is proposed for modelling these problems numerically using the Finite Element Method (FEM). Raleigh-Taylor (RT) instability is simulated with a moving, free-surface boundary. The technique is explained below and compared with other approaches. Several examples are presented. Surface topography is studied for both the free-surface boundary condition, using the new technique, and for the free-slip condition using traditional linearization methods. Also, the dynamics of surface topography were studied for different viscosity contrasts and geometries. 2 PHYSICAL A N D NUMERICAL M O D E L 2.1 Raleigh-Taylor instability The Raleigh-Taylor instability for two layers of viscous fluid occurs when the density of the upper layer is higher than that of the bottom layer so that the system is gravitationally unstable. Any small perturbation on the interface between the layers grows until finally the lower material rises up and displaces the upper layer. During the rise of a diapir, the non-hydrostatical part of pressure (the dynamical pressure) must be balanced by a rise in the surface above the diapir producing topography. The geometry of the problem modelled is shown in Fig. 1. The two layers are described by their density, viscosity and thickness; p, p and h respectively. The botton of the mesh was given a non-slip boundary condition. Along the vertical sides the free-slip condition was chosen which implies mirror symmetry. The upper boundary was flexible and had a free-surface (free stress) boundary condition. c=o (free surface) or vz -0 , 0 = 0 (free slip) XZ height of diapir Figure 1. Model geometry, boundary conditions and physical variables used in the calculations. An initial cosinusoidal perturbation with a wavelength equal to the characteristic wavelength and with an amplitude of 4 per cent of the box height is imposed on the buoyant layer. 2.2 Mathematical formulation In the present study the following forms of Stokes equations for slow incompressible fluids was simulated: where aij is the stress tensor, p is the viscosity, p is the density, ui the ith component of velocity and aij is the Kronecker delta. The change in the viscosity and density fields during each time step were calculated using the continuity equation: where A is any characteristic material property (e.g. density, Newtonian viscosity, etc.), 2.3 Comparison of methods Numerical modelling of the slow flows governed by Stokes equation is well established. Temam (1977) studied the application of FEM technique for the solution of this equation from a mathematical point of view. The main difficulty in numerical simulation of Stokes equation is to satisfy the incompressibility condition. There is a class of methods based on the introduction of a stream function which satisfies this condition automatically and gives good results. However, the stress-free boundary condition causes significant difficulties because it requires the calculation of third-order derivatives of the stream function. These algorithms are also not applicable for irregular meshes. Thus, it is difficult to solve problems in regions with complicated geometry, and to refine the mesh in areas of special interest. A further disadvantage is that the approximation functions are not expressed explicitly in terms of the nodal variables and an additional system of linear equations must be solved. Diapirism and topography For these reasons we use the pressure-velocity formulation for Stokes equations. This allows us to impose arbitrary velocity or stress boundary conditions and to work with arbitrary geometry. However, the two general strategies (mixed and penalty methods) which are used for satisfying the incompressibility condition have their own drawbacks. This is especially true for problems with a free surface, non-linear rheology or buoyancy forces. Pelletier et al. (1989) showed that in this case the FEM formulation reduces to an ill-conditioned system of linear equations. Erroneous solutions can be obtained for density-driven flow problems due to a poor choice of pressure discretization. The main advantage of the integrated method is that velocity and pressure are computed directly, without numerical differentiations. However, the system of equations becomes very large and the matrix contains zeros on the main diagonal, which makes partial pivoting necessary. For the mixed method Pelletier et al. (1989) proposed pressure scaling for the stiffness matrix in order to overcome this ill-conditioning. There are several different types of penalty methods including continuous, discrete and iterative methods as well as the Uzawa algorithm. Cuvelier, Segal & Steenhoven (1986) have shown that the discrete penalty method (where the pressure is eliminated after discretization of the continuity equation and the momentum equation) is superior to other penalty function methods. The main advantage of the penalty method is the large reduction of the system of equations as well as the fact that partial pivoting is not necessary. 2.4 Choice of elements Because the system includes both buoyancy forces and free-surface boundary conditions, basis functions for the approximation of the pressure and velocity fields must be chosen carefully. Many studies have compared different elements for the solution of Stokes equations (Cuvelier et al. 1986; Crochet, Davies & Walters 1984; Hughes 1987). Quadrilateral elements are preferable (Cuvelier et al. 1986), but triangular elements are more convenient for complex geometries. Suitable basis functions for velocity and pressure must satisfy the following requirements. (1) Pressure must be approximated by interpolation polynomials which are of order at least one degree less than the polynomials for the velocity. (2) The number of equations for incompressibility in the global matrix should not exceed the number of degrees of freedom for the velocity field. Otherwise, the system will be overconstrained and the velocity will be completely determined by the condition of incompressibility (Crochet et al. 1984). ( 3 ) The constraint ratio, r, introduced by Hughes (1987) should be approximately equal to 2. r is the ratio of the number of displacements to the number of incompressibility conditions per element. If r is less than 2, the element will be overconstrained and mesh locking will occur. In other words, if the volume of the element is kept fixed (incompressibility), the element is not flexible enough to simulate the velocity field. Following these recommendations and from our own experience, we have chosen triangular elements with enriched continuous quadratic basis functions for velocity and discontinuous linear basis 555 Velocity: enriched quadratic (continuous) (7 nodal points: x) & Pressure:(discontinuous) linear (3 nodal points: 0 ) Figure 2. The Crouzeix-Raviart element (after Cuvelier et a!. 1986). functions for pressure (Fig. 2). Discontinuous functions were used because continuous functions were found to produce a 'chessboard' effect (the incompressibility condition was satisfied for the domain as a whole but not for the individual elements). This element satisfies all three conditions listed above. We use this element for both mixed and penalty techniques depending on the problem. 2.5 Solution procedure Problems are solved with a time-stepping procedure. At each time step the velocity field is obtained by solving steady-state Stokes equations (2)-(4) for a given geometry. The velocity field is then used to move the material according to the continuity equation (5). 2.5.1 Velocity calculations At each time step, our program first calculates the velocity field based on the current density and viscosity distribution. The Galerkin method was used to reduce Stokes equations to a symmetric, non-positive definite system of linear equations in the case of mixed formulation and a positive definite system in the case of the discontinuous penalty method. This system was assembled and solved using the frontal method (Hood 1976) with double precision (the relative accuracy of the solution was approximately and the divergence of velocity in each element was about for the mixed method and for the penalty method. We have used typically about 21 nodal points in the z-direction and a proportional number of nodes in x-direction depending on the goemetry. 2.5.2 Method of markers Moving a density or viscosity field which contains sharp discontinuities through a discrete mesh is difficult due t o problems with numerical diffusion (Hirt & Nichols 1981). These diffusion problems can be overcome by using a method of characteristics based on marker points (particles). Weinberg & Schmeling (1991) developed a marker technique which is very effective for multiphase flow where each phase has different rheological properties. In their method, each marker contains information on all material phases present. This allows the effective material properties at the nodal points to be estimated correctly. Also, uncertainties which arise when more than two phases are present are avoided. Marker points, containing material property information are distributed throughout the numerical mesh. The markers are moved with the numerical mesh according to the velocity field. In places where the density of markers becomes too low (as in extentional zones) new, 556 A. Poliakov and Y. Podladchikov additional markers are added with material properties interpolated from neighbouring markers. This procedure is easy to program, computationally cheap and produces a minimum of diffusion compared with techniques which redistribute all the markets in order to maintain the same number of marker particles. Also, sedimentation can be simulated by introducing new markers on the upper surface. Erosion can be simulated by removing markers on the surface. A high-order, 13 point integration formula (Hughes 1987) for calculation of the stiffness matrix was used in order to conserve detailed information from the marker field in the coarse FEM mesh. Direct projection from non-regulary distributed markers to non-regulary located integration points is a time consuming operation. It is necessary to search through all markers for each integration point, looking for the closest markers. Instead of doing this, we use a two-step algorithm. We introduce a very fine regular mesh and first interpolate properties from the markers to this mesh. We then project from this mesh to the integration points. We tested our codes for different grid sizes, numbers of markers and resolution of the fine interpolation grid. In the cases presented in this work we found it optimal to have 100 markers and a 5 x 5 interpolation fine mesh per quadrilateral (which contains two triangles). Thus, for a typical problem with 21 x 21 FE mesh we will have lo4 markers and 2.5 x lo3 fine mesh nodes that use only about 200-300 K of memory. 2.5.3 Technique for moving mesh, boundaries and markers Markers are usually moved through a fixed, Eulerian mesh. Unfortunately, this method is very expensive, because it is necessary to update the locations of markers and material properties at every time step (or even several times per time step for Runge-Kutta, or predictor-corrector methods). These calculations involve finding the element to which each marker belongs and interpolating the velocity field of the element to the marker location. Also, the updated material properties must be projected onto the integration points. The Langrangian method, where the mesh moves with the material, is faster. Values of density and viscosity are conserved in the integration points. Therefore, the searching and interpolation required for the Eulerian method is not necessary for every time step. Of course, this method is not good for extremely large strains because the mesh becomes too distorted. Following a suggestion of P. Cundall (personal communication 1991), we combined both the Lagrangian technique and markers as shown on Fig. 3. At the initial stage (Fig. 3a), material properties of the different layers are projected to both the integration points of each element and to the markers (schematically represented by 'stars' and 'crosses' indicating the different materials). For each marker, the number of the element which contains the marker is stored. Also, within each element the Cartesian coordinates of the markers are converted to local coordinates (area coordinates for triangular elements). The mesh is then updated according to the Lagrangian technique. At each time step new positions for the mesh grid nodes are calculated from the current velocity field using an explicit two stage Runge-Kutta method (Fig. 3b). Figure 3. Explanation of the method for moving the boundaries and mesh used in the present work. (a) Initial stage; 'crosses' and 'stars' indicate different materials (b) Moving mesh with markers 'chilled in element' until it is necessary to remesh. (3) Mesh and markers after remeshing. This Lagrangian movement is very fast because it is onhy necessary to move the mesh nodes according to the calculated velocity field. No interpolation is necessary. The moving mesh automatically tracks evolution of the upper surface. When the mesh becomes too distorted (Fig. 3b), it is necessary to remesh. Since the local coordinates of the markers remain unchanged during the Lagrangian movement of the mesh, the Cartesian coordinates of the markers can be obtained by simple interpolation from the nodes of the elements. Only at this stage is it necessary to interpolate from the markers to the integration points as is described in Section 2.4.2. This is in contrast to the Eulerian method where this interpolation must be performed at every time step. This procedure is demonstrated in Fig. 3(c), which show the same stage as Fig. 3(b) but with a new regular I 557 Dinpirism and topography mesh. The new mesh is adjusted to fit the deformed upper boundary. We found that this Lagrangian approach combined with markers was much faster then the Eulerian approach and was much easier to program. This method is probably more precise, because interpolation is required only for remeshing so that the total number of interpolations is only about 5-10 for overturn. We feel that this method would also be useful for heat transfer problems. The moving elements would eliminate convective terms from the heat transfer equation leaving only the diffusion term. h, = 0.3, R = 0.1 (a) 0.025--'; ' , " , ' 1 i I i i i/ - 0.020 - i l ' ' ' 1 2 - 0.010 'b ~ I ' " ' I " ' ' I - -.-.-. analytical,(Ramberg1988b) II - = 1.1 analytical, Ramberg,l968b) - - - . p1/p2= 1 0 , X = 3.0 , - '1 !I !I /: 1; - I 0.005 - f 0.000 I 0 3 ' -...- ' 0 . 0 1 5 ~!,I Ez4 " - pI/~2 = 1, h 50 . . 100 . , I . , , 150 , 200 TIME Numerical results 3.1 Non-dimensionalization and verification of the method h, = 0.3, R = 0.1 (b) All variables were non-dimensionalized on the basis of characteristic velocity: where L is the sum of the thickness of both layers, p , is the dynamic viscosity of the upper layer, Ap = (p, - p,) is the density contrast between the two layers, g is the acceleration of gravity, t is time and x and z are coordinates. For calculations with the free-surface condition, the solution depends not only on the density Ap as in the free-slip caie but also on the additional parameter p, and can be neglected. We will show where pair<< (Section 4) that for geophysical applications (R < 0.15), the results are similar and topography depends linearly on R. The initial shape of the simulated region was rectangular with unit height and a length equal to one half of the wavelength, A. The amplitude of the initial cosinusoidal perturbation was 0.04 times the height of the region. This makes possible a comparison with the linear analytic theory during the initial stages of flow. We verified our code for the initial linear stages by simulating Raleigh-Taylor (RT) instability of a viscous buoyant layer covered by another viscous layer with a free surface. Non-dimensional parameters for this problem are the ratio of the densities (parameter R), the viscosities of the two layers and the thickness of the buoyant layer h, (see Fig. 4). The results of our numerical simulations were compared t o the analytical solutions of Ramberg (1968b) in Fig. 4. The height of the diapir (Fig. 4a) and the surface elevation (Fig. 4b) are plotted against non-dimensional time. For each viscosity ratio a characteristic wavelength, A, was also calculated from the formula of Ramberg (1968b). We can see that there is satisfactory agreement between theory and numerical experiments for the height of the diapir up to a non-dimensional time of 60 for the isoviscous case and up to 10 for a viscosity of contrast lo3. This is in agreement with the suggestion of Schmeling (1987) that the analytical solution is valid up to amplitudes of 10 per cent of the wavelength or up t o 30 per cent of the height of the thinnest layer. However, surface elevation above the diapir E g - 0 , v , PI/& 6 , - - -. -0.6 0 50 , = , 1, h = , 1.1 , , , , analytical, Ramberg,l968b) pl/p2 = 10 , h = 3.0 analytical,(Ramber ,1968b) 100 150 200 TIME - pl/p2 = --'-, -0.8 0 lo3, X = 3.0 Peter van Keken, (pers. comm.,1991) 20 40 TIME Figure 4, Comparison of present codes with analytial solution of Ramberg (1968b) for linear stage (a, b) for free-surface boundary condition and (c) non-linear stages for free-slip upper boundary condition. from the numerical solution starts to deviate from the analytical solutions two times earlier (Fig. 4). For the non-linear stage, with a free-slip upper boundary, we verified our codes by comparison with calculations performed by U. Christensen for the isoviscous case. We also compared our codes for a viscosity contrast of lo3 with the code of Peter van Keken (personal communication 1991) (see Fig. 4c). 3.2 Dynamics of the upper surface under different conditions We have carried out a brief parametric study in order to determine which factors influence the evolution of surface topography. The geologic setting in which actual diapirs develop is, of course, much more complicated than the simple, two-layer isothermal model we study. 558 A. Poliakov and Y.Podladchikov 3.2.1 Effectof z~iscosity The viscosity contrast between a diapir and the overburden significantly influencesthe evolution of the system, including the surface topography and the height, shape and growth rate of the diapir. In various geological settings the viscosity contrast can range over several orders of magnitude due to compositional, rheological or temperature differences. Therefore, we will first study the influence of viscosity. The growth of diapirs is also strongly influenced by the initial conditions such as the shape, amplitude and wavelength of the perturbation on the interface between two materials. It is usually assumed that a wide spectrum of perturbations is present on the interface between the buoyant layer and the overburden in natural systems. Since perturbations with different wavelengths grow at different rates, the perturbation growing at the fastest rate will suppress dominate the others. Thus, the diapirs will be spaced at this fastest growing interval (characteristic wavelength). Under certain geologic conditions, such as faulting, sedimentation or erosion, growth of diapirs at noncharacteristic wavelengths can be induced. Schrneling (1987) explored, for the isoviscous case, the type of perturbations necessary to excite diapir growth at non-characteristic wavelengths. We will consider here cases where the initial perturbation is equal to, shorter than or longer than the characteristic wavelength. The initial prescribed perturbation must be strong enough to inhibit the natural tendency of the system to revert to the characteristic wavelength. In all cases we set the amplitude of the initial perturbation equal to 0.04 which was adequate to ensure that the desired wavelength survived. Figs 5(c), (d) and (e) show calculations for viscosity contrasts of 1, 10 and lo3 respectively. The calculations were given initial perturbations so that each grew at the characteristic wavelength for that viscosity contrast. The wavelengths, A, were 1.1, 1.68 and 3.0 respectively. The velocity field (left column) and the passive strain markers (right column) are shown at several stages, The values of time and the maximum velocity (for scaling of the arrows) are also shown. Strain markers are shown for visualization of the deformations. Markers were not attached to the boundaries due to the high strain on the boundaries. When distortion is too large, some markers were excluded (white fields). Figs 5(a) and (b) show the elevation directly above the diapir and the height of the diapir versus nondimensional time (see Fig. 1 for terminology). It is difficult to compare results for different wavelengths because of the extra parameter, A. However, the difference in topography can be explained as follows. For longer wavelengths, the size of the diapir increases and if the buoyant body is larger in the vertical direction the buoyant forces will be larger. Therefore topography will be higher for larger diapirs in order to compensate this force. U. Christensen (personal communication 1991) suggested a comparison between the size of a diapir in the vertical direction and the difference in topography. We can see that for the case pl/p, = lo3 that the vertical size of the diapir is two times higher than for the isoviscous case and that the topography is also two times higher. This is in good agreement with the results shown in Fig. 5(a). We now compare calculations for the same three viscosity contrasts for A = 3.0, the characteristic wavelength for a viscosity contrast of lo3 ((~ig. 6) and for h = 1.1, the characteristic wavelength for the isoviscous case (Fig. 7). To see the temporal evolution of the surface topography, exaggerated surface profiles the shape of the diapir are shown on the same plot for the three different cases (Fig. 6c, d, e). Each line type corresponds to different time steps for the same diapir (bottom) and topography (top). The growth rate of the diapir is strongly dependent on the viscosity of the bottom, buoyant layer. For a given wavelength (A = 1.1 or 3.0) the growth rate of the diapir increases significantly with decreasing viscosity of the buoyant layer. Also, the wavelength has very little effect on the diapir growth rate. Topography, in constrast to diapir growth rate, is very sensitive to the wavelength. For a fixed wavelength the maximum value of elevation depends very little on the viscosity contrast. However, as the wavelength becomes longer, maximum surface elevation increases. The dependence of the growth rate on the viscosity contrast is more pronounced at the earlier stages of diapir growth. This can be seen from Figs 6(b) and 7(b), where the slope of the elevation versus time curves are quite different for early times but become similar at later stages. For example, at the initial stage the growth rate is five times faster for a viscosity contrast, p1/p2= 10 and 10 times faster for the viscosity contrast, pI/p2 = lo3 when compared to the isoviscous case. By the time the top of the diapir has (a) h, = 0.3, R = 0.1 7 0 . 0 2 5 L , ," " ' " " ' 1 " " ' TIME (b) - 0 . 8 1 . . 0 h, = 0.3, R = 0.1 . , I 50 . . . . I 100 . . . . , . . 150 , , I 200 TIME Figure 5. Elevation (a) and height of diapir (b) for different viscosity contrasts at the characteristic wavelengths for each viscosity contrast. (c, d, e) Passive strain marker field (left columns) and velocity (right columns). For each part the value of the maximum velocity and time are also given. Diapirisrn and topography velmax= 3.e-02 time35. Figure %(Continued) 559 560 A. Poliakou and Y. Podladchikou 2I I i Diapirism and topography I (a) h = 1.1, h, = 0.3,R = 0.1 TIME (b) -0.81 0 . . A = 1.1, . . I 50 561 h, = 0.3, R = 0.1 . . . . , . . . . , . . . 100 150 200 TIME Figure 7. As Fig. 6 with A = 1.1. reached an elevation of -0.4, the growth rate for all three viscosity contrast are very close. This observation can be explained by dividing diapir growth into two stages. During the first stage, the 'bulb' of the diapir forms from the buoyant layer. During the second stage the mature diapir rises toward the surface. The first stage, in which the diapir is formed is most strongly influenced by the viscosity contrast. After the mature diapir is formed the time of ascent is very similar for the various viscosity contrasts. The total time for the ascent of the diapir is the sum of the two stages. Therefore, in cases where the contribution of the second stage is dominant, the total time of ascent will be only slightly influenced by the viscosity contrast. The second stage will be dominant when the thickness of the buoyant layer is very thin or the perturbation is very strong. In order to isolate the influence of wavelength we present Fig. 8, where the viscosity contrast was fixed (PI = 10)) but the wavelength was varied to 1.1, 1.68 and 3.0 (1.68 was the characteristic wavelength). Fig. 8 shows that the time in which the diapir rises to the depth -0.05 is very close for all A. However, the surface topography varies considerably and is two times greater for A = 3.0 compared to A = 1.1. Although the maximum elevation of surface topography is approximately the same for a given h, the evolution of the growth rate is very different. We now compare surface topography for different viscosity contrasts for a fixed wavelength, A = 3. We examine the rate of growth when the diapir has reached a given depth. For example, when the diapirs rise to a depth of -0.4, the maximum elevation above the isoviscous diapir is only 0.005 while the elevation is 0.016 for the lo3 viscosity contrast case (Figs 6c-e). This difference can be explained by examining the shapes of the diapirs at this stage. It can be seen that the diapir with viscosity one (Fig. 6c, solid line) is more narrow and there is a layer of the buoyant material at the bottom. The diapir with a viscosity of (Fig. 6e, solid line) has coalesced into one 'bulb'. The buoyant force (or dynamic pressure) is higher for this configuration. The difference in pressure is higher for the high-viscosity case because there is no layer of buoyant material. The same effect could be seen for h = 1.1. 3.2 2 Ef~ec. of the thickness of .he buoyant layer We now examine the effect of the relative thickness of the buoyant layer on the surface topography. The combined thickness of the two layers is held constant, while the thickness of the lower layer is set to h, = 0.1, 0.3, 0.5, and 0.7 (Figs 9c-f, respectively). The wavelength of the initial perturbation is the same for all cases. 562 A. Poliakov and Y. Podladchikov 20 (d) g g 40 80 60 100 TIME TIME h = 1.68, 0.010 0.000 -0.010 ..... -0.020 -0.030 0.0 0.2 0.4 0.6 0.8 1.0 Figure 8. Elevation (a) and height of diapir (b) for fixed viscosity contrast and at the variable wavelengths. (c, d, e) Contours of diapirs (bottom) and exaggerated topography (top) for different moments in time. The surface elevation directly above the diapir is plotted versus non-dimensional time for these three cases in Fig. 9(a). The height of the diapir is plotted versus time in Fig. 9(b). The maximum height of the topography is highest when the two layers have equal thickness. This is because the total potential energy available to drive the flow is greatest for this configuration and causes the most intensive flow. The maximum height of the topography differs for h, = 0.3 and h, = 0.7 due to the asymmetry of boundary conditions on the surface and bottom. Nonetheless, potential energies of these systems are similar. The time for topographic growth increases with decreasing h, as one would expect. interior processes. These assumptions are usually met for large-scale geophysical problems. For free-slip calculations, topography is linearly dependent on R. In order to compare this prediction with the results of free surface simulations we made calculations for various values of R. Fig. 10 shows the elevation of the surface directly above the diapir for two viscosity contrasts for the same parameters used in Section 3.1. Horizontal profiles of topography (scaled by R ) do not differ considerably with increases of this parameter. Thus, the relation between topography and R is indeed linear from R = 0.03 to 0.15. This indicates that the linear theory is valid for geologically reasonable density contrasts. 4 INFLUENCE OF THE PARAMETER R ON THE TOPOGRAPHIC B E H A V I O U R 5 Raleigh-Taylor instability is usually modelled with a free-slip boundary condition on the upper surface. The topography is then estimated a posteriori from equation (1). This procedure is valid as long as the topography is small and adjustment of the surface occurs much faster than GENERAL CONCLUSIONS We have presented a new numerical technique using markers and a deformable Lagrangian mesh to study the deformation of the surface above a rising diapir. This method allowed us to model a free-surface boundary in a self-consistent way. The method makes it possible to investigate many different problems with non-regular - - I - Diapirisrn and topography (a) 0.020 A = 1.1, P,/P, = I , 563 R = 0.1 -1 I -1.0 0 200 TIME I 200 TIME 400 b, ... I"'.. I :*. , / ! I \ '. \;* \ I ,' : '.. ................ ----.a, \ Figure 9. Elevation (a) and height of diapir (b) for different thicknesses of buoyant layer at the fixed wavelength A = 1.1. (c, d, e) Contours of diapirs (bottom) and exaggerated topography (top) for different moments in time. geometry. Due to lagrangian updating of the elements, codes become much faster and are easier to program than when using the Eulerian approach. Markers are combined with moving elements and are used only during a few remeshing procedures. Our results led to the following conclusions. (1) The growth rate of the diapir is more strongly influenced by the viscosity contrast than the wavelengths. This is due to the strong growth rate difference during the initial stage. In our cases the growth times is 3-5 times less for a viscosity contrast of lo3 than for a viscosity contrast of 1 (for a given A). However, for different A the time of ascent varies by only 30 per cent (for a given viscosity contrast). (2) The maximum elevation above the diapir linearly increases with increasing wavelength and is approximately the same for different viscosity contrasts. (3) The elevation will be considerably higher for a low-viscosity diapir at a given depth than for an isoviscous diapir. This is because the shape of the diapir depends on the viscosity contrast. It was found that topography behaves linearly up to R = 0.15 which indicates that a posteriori estimation of topography for the free-slip calculations are valid within this limit. ACKNOWLEDGMENTS DENSITY RATIO, R = P ,-P,/P, Figure 10. The maximum elevation dependence on the parameter R, for different viscosity contrasts. We are grateful to Christopher Talbot for suggesting the free-surface problem. David Yuen is thanked both for suggesting comparison with traditional modelling methods 564 A. Poliakov and Y. Podladchikov and for his critical reading of our manuscript. We thank Harro Schmeling for introducing us to the marker technique, and Peter Cundall for suggestion the Lagrangian approach. Andrei Malevsky and Harro Schmeling gave many helpful explanations of physical processes. Peter van Keken and Ulrich Christensen kindly run their codes for a benchmark comparison of our program. V. Novikov and A. Lyakhovsky significantly contributed in the beginning of the present codes. We thank Ethan Dawson for patiently revising the English of our manuscript. Laura Weyer is thanked for drawing the pictures. Debbie Downs and Sara Green established a proper research atmosphere. U. Christensen and H. Schmeling provided excellent reviews and gave many ideas on how to improve the paper. We thank the Minnesota Supercomputer Institute for providing facilities during the course of this work. A study grant and computer facilities from the University of Uppsala is also acknowledged. Y. Podladchikov is grateful to the Th. Nordstroms fond, Royal Swedish Academy of Sciences for research grant 'Numerical modelling of Geological structures'. REFERENCES Crochet, M. J.., Davies, A. R. & Walters, K., 1984. Numerical Simulation of Non-Newtonian Flow, Elsevier, Amsterdam. Cuvelier, C., Segal, A. & Steenhoven, A. A., 1986. Finite Element Methods and Navier-Stokes Equations, Reidel, Dordrecht. Fadeev, V. E., 1988. Geodynamical models in regional geophysics, Full doctor thesis, Moscow State University. Fleitout, L., Froidevaux, C. & Yuen, D., Active lithospheric thinning, Tectonophysics, 132, 271-278. Griffiths, R. W., Gurnis, M. & Eitelbeg, G., 1989. Holographic measurements of surface topography in laboratory models of mantle hotspots, Geophys. J., 96,477-495. Hirt, C. W. & Nichols, B. D., 1981. Volume of fluid (VOF) method for the dynamics of free boundaries problem, J. Comp. Phys., 39,201. Hood, P., 1976. Frontal solution program for unsymmetric matrices, Znt. J. Num. Meth. Eng., 10, 379-399. Hughes, T. J. R., 1987. The Finite Element Method, Prentice-Hall, Englewood Cliffs, NJ. Jackson, M. P. A., Talbot, C. J. & Cornelius, R. R., 1988. Centrifuge modelling of the effect of aggradation and progradation on syndepositional salt structures, Report 1973, Bur. Econ. Geol., University Texas at Austin. Lliboutry, L. A., 1987. Very Slow Flows of Solidr, Martinus Nijhoff Publishers, Dordrecht. McKenzie, D. P., 1977. Surface deformation, gravity anomalies and convection, Geophys. J. R. astr. Soc., 48, 21 1-238. Morgan, W. J., 1965. Gravity anomalies and convection currents. 1. A. sphere and cylinder sinking beneath the surface of a viscous fluid, J. geophys. Res., 70, 6175-6187. Myasnikov, V. P. & Savushkin, V. D., 1978. Structure of the surface movement in the hydrodynamic model of the Earth evolution, Dokl. Acad. nauk SSSR., 239, 200. Olson, P. & Nam, I. S., 1986. Formation of seaflow swells by mantle plumes, J. geophys. Res., 91, 7181-7191. Pak, V. V., 1989. Numerical modelling of the process of the mantle heterogeneous rising, PhD thesis, Moscow State University. Pelletier, D., Fortin, A. & Camarero, R., 1989. Are FEM solutions of incompressible flows really incompressible? (Or how simple flows can cause headaches!), Int. J. Numer. Meth. Fluids, 9, 99-112. Ramberg, H., 1968a. Fluid dynamics of layered systems in the field of gravity, a theoretical basis for certain global structures and isostatic adjustment, Phys. Earth planet. Inter., 1, 63-87. Ramberg, H., 1968b. Instability of layered systems in the field of gravity I an 11, Phys. Earth planet. Inter., 1, 427-474. Schmeling, H., 1987. On the relation between initial conditions and late stages of Rayleigh-Taylor instabilities, Tectonophysics, 133, 65-80. Talbot, C. J. & Jarvis, R. J., 1984. Age, budget and dynamics of an active salt extrusion salt in Iran, J. Struct. Geol., 6, 521-524. Temam, R., 1977. Navier-Stokes Equations, North-Holland, Amsterdam. Weinberg, R. B. & Schmeling, H., 1992. Polydiapirs: Multiwavelength gravity structures, J. Struct Geol., in press. Zaleski, S. & Julien, P., 1990. Numerical simulation of salt diapir formation in small and extended geometries, Report, Total-Compagnie Francaise des Petroles.