CS计算机代考程序代写 flex data structure algorithm Abstract

Abstract
The bottle-neck in most cloth simulation systems is that time steps must be small to avoid numerical instability. This paper describes a cloth simulation system that can stably take large time steps. The simulation system couples a new technique for enforcing constraints on individual cloth particles with an implicit integration method. The simulator models cloth as a triangular mesh, with internal cloth forces derived using a simple continuum formulation that supports modeling operations such as local anisotropic stretch or compres- sion; a unified treatment of damping forces is included as well. The implicit integration method generates a large, unbanded sparse lin- ear system at each time step which is solved using a modified con- jugate gradient method that simultaneously enforces particles’ con- straints. The constraints are always maintained exactly, independent of the number of conjugate gradient iterations, which is typically small. The resulting simulation system is significantly faster than previous accounts of cloth simulation systems in the literature.
Keywords—Cloth, simulation, constraints, implicit integration, physically-based modeling.
1 Introduction
Physically-based cloth animation has been a problem of interest to the graphics community for more than a decade. Early work by Ter- zopoulos et al. [17] and Terzopoulos and Fleischer [15, 16] on de- formable models correctly characterized cloth simulation as a prob- lem in deformable surfaces, and applied techniques from the me- chanical engineering and finite element communities to the prob- lem. Since then, other research groups (notably Carignan et al. [4] and Volino et al. [20, 21]; Breen et al. [3]; and Eberhardt et al. [5]) have taken up the challenge of cloth.
Although specific details vary (underlying representations, nu- merical solution methods, collision detection and constraint meth- ods, etc.), there is a deep commonality amongst all the approaches: physically-based cloth simulation is formulated as a time-varying partial differential equation which, after discretization, is numeri- cally solved as an ordinary differential equation
x ̈=M−1−∂E+F. (1) ∂x
In this equation the vector x and diagonal matrix M represent the geometric state and mass distribution of the cloth, E—a scalar func-
tion of x—yields the cloth’s internal energy, and F (a function of x and x ̇ ) describes other forces (air-drag, contact and constraint forces, internal damping, etc.) acting on the cloth.
In this paper, we describe a cloth simulation system that is much faster than previously reported simulation systems. Our system’s faster performance begins with the choice of an implicit numerical integration method to solve equation (1). The reader should note that the use of implicit integration methods in cloth simulation is far from novel: initial work by Terzopoulos et al. [15, 16, 17] applied such methods to the problem.1 Since this time though, research on cloth simulation has generally relied on explicit numerical integra- tion (such as Euler’s method or Runge-Kutta methods) to advance the simulation, or, in the case of of energy minimization, analogous methods such as steepest-descent [3, 10].
This is unfortunate. Cloth strongly resists stretching motions while being comparatively permissive in allowing bending or shear- ing motions. This results in a “stiff” underlying differential equa- tion of motion [12]. Explicit methods are ill-suited to solving stiff equations because they require many small steps to stably advance thesimulationforwardintime.2 Inpractice,thecomputationalcost of an explicit method greatly limits the realizable resolution of the cloth. For some applications, the required spatial resolution—that is, the dimension n of the state vector x—can be quite low: a reso- lution of only a few hundred particles (or nodal points, depending on your formulation/terminology) can be sufficient when it comes to modeling flags or tablecloths. To animate clothing, which is our main concern, requires much higher spatial resolution to ad- equately represent realistic (or even semi-realistic) wrinkling and folding configurations.
In this paper, we demonstrate that implicit methods for cloth overcome the performance limits inherent in explicit simulation methods. We describe a simulation system that uses a triangular mesh for cloth surfaces, eliminating topological restrictions of rect- angular meshes, and a simple but versatile formulation of the inter- nal cloth energy forces. (Unlike previous metric-tensor-based for- mulations [15, 16, 17, 4] which model some deformation energies as quartic functions of positions, we model deformation energies only as quadratic functions with suitably large scaling. Quadratic energy models mesh well with implicit integration’s numerical properties.) We also introduce a simple, unified treatment of damping forces, a subject which has been largely ignored thus far. A key step in our simulation process is the solution of an O(n) × O(n) sparse lin- ear system, which arises from the implicit integration method. In this respect, our implementation differs greatly from the implemen- tation by Terzopoulos et al. [15, 17], which for large simulations
1 Additional use of implicit methods in animation and dynamics work in- cludes Kass and Miller [8], Terzopoulos and Qin [18], and Tu [19].
2Even worse, the number of time steps per frame tends to increase along with the problem size, for an explicit method. Cloth simulations of size n— meaning x ∈ IRO(n)—generally require O(n) explicit steps per unit simu- lated time. Because the cost of an explicit step is also O(n) (setting aside complications such as collision detection for now) explicit methods for cloth require time O(n2)—or worse.
􏰀􏰁
Large Steps in Cloth Simulation
David Baraff Andrew Witkin
Robotics Institute Carnegie Mellon University
SIGGRAPH ’98 1
CD-ROM version

used an “alternating-direction” implicit (ADI) method [12]. An ADI method generates a series of tightly banded (and thus quickly solved) linear systems rather than one large sparse system. (The price, however, is that some of the forces in the system—notably between diagonally-adjacent and non-adjacent nodes involved in self-collisions—are treated explicitly, not implicitly.) The speed (and ease) with which our sparse linear systems can be robustly solved—even for systems involving 25,000 variables or more—has convinced us that there is no benefit to be gained from using an ADI method instead (even if ADI methods could be applied to irregular triangular meshes). Thus, regardless of simulation size, we treat all forces as part of the implicit formulation. Even for extremely stiff systems, numerical stability has not been an issue for our simulator.
1.1 Specific Contributions
Much of the performance of our system stems from the development of an implicit integration formulation that handles contact and ge- ometric constraints in a direct fashion. Specifically, our simulator enforces constraints without introducing additional penalty terms in the energy function E or adding Lagrange-multiplier forces into the force F. (This sort of direct constraint treatment is trivial if equa- tion (1) is integrated using explicit techniques, but is problematic for implicit methods.) Our formulation for directly imposing and maintaining constraints is harmonious with the use of an extremely fast iterative solution algorithm—a modified version of the conju- gate gradient (CG) method—to solve the O(n) × O(n) linear sys- tem generated by the implicit integrator. Iterative methods do not in general solve linear systems exactly—they are run until the solu- tion error drops below some tolerance threshold. A property of our approach, however, is that the constraints are maintained exactly, re- gardless of the number of iterations taken by the linear solver. Addi- tionally, we introduce a simple method, tailored to cloth simulation, for dynamically adapting the size of time steps over the course of a simulation.
The combination of implicit integration and direct constraint sat- isfaction is very powerful, because this approach almost always al- lows us to take large steps forward. In general, most of our simu- lations require on average from two to three time steps per frame of 30 Hz animation, even for (relatively) fast moving cloth. The large step sizes complement the fact that the CG solver requires rel- atively few iterations to converge. For example, in simulating a 6, 000 node system, the solver takes only 50–100 iterations to solve the 18, 000 × 18, 000 linear system formed at each step. Addition- ally, the running time of our simulator is remarkably insensitive to the cloth’s material properties (quite the opposite behavior of ex- plicit methods). All of the above advantages translate directly into a fast running time. For example, we demonstrate results similar to those in Breen et al. [3] and Eberhardt et al. [5] (draping of a 2,600 node cloth) with a running time just over 2 seconds per frame on an SGI Octane R10000 195 Mhz processor. Similarly, we show gar- ments (shirts, pants, skirts) exhibiting complex wrinkling and fold- ing behavior on both key-framed and motion-captured characters. Representative running times include a long skirt with 4,530 nodes (8,844 triangles) on a dancing character at a cost of 10 seconds per frame, and a shirt with 6,450 nodes (12,654 triangles) with a cost varying between 8 to 14 seconds per frame, depending on the un- derlying character’s motion.
1.2 Previous Work
Terzopoulos et al. [15, 17] discretized cloth as a rectangular mesh. Energy functions were derived using a continuum formulation. This work recognized the need for damping forces; however, only a sim- ple viscous drag force −kx ̇ was used. The linear systems result-
ing from the use of implicit integration techniques were solved, for
small systems, by direct methods such as Choleski factorization, or
using iterative techniques such as Gauss-Seidel relaxation or conju-
gate gradients. (For a square system of n nodes, the resulting linear
SIGGRAPH ’98 2
CD-ROM version

n. In this case, banded Choleski factoriza- tion [6] requires time O(n2 ).) As previously discussed, Terzopoulos
et al. made use of an ADI method for larger cloth simulations. Following Terzopoulos et al.’s treatment of deformable surfaces, work by Carignan et al. [4] described a cloth simulation system using rectangular discretization and the same formulation as Ter- zopoulos et al. Explicit integration was used. Carignan et al. recog- nized the need for damping functions which do not penalize rigid- body motions of the cloth (as simple viscous damping does) and they added a force which damps cloth stretch and shear (but not bend). Later work by the same group includes Volino et al. [20], which focuses mainly on collision detection/response and uses a tri- angular mesh; no mention is made of damping forces. The system uses the midpoint method (an explicit method) to advance the simu- lation. Thus far, the accumulated work by this group (see Volino et al. [21] for an overview) gives the only published results we know of for simulated garments on moving characters. Reported resolutions of the garments are approximately two thousand triangles per gar- ment (roughly 1,000 nodal points) [21] with running times of sev- eral minutes per frame for each garment on an SGI R4400 150 Mhz
processor.
Breen et al. [3] depart completely from continuum formulations
of the energy function, and describe what they call a “particle- based” approach to the problem. By making use of real-world cloth material properties (the Kawabata measuring system) they produced highly realistic static images of draped rectangular cloth meshes with reported resolutions of up to 51 × 51 nodes. The focus of this work is on static poses for cloth, as opposed to animation: thus, their simulation process is best described as energy minimization, although methods analogous to explicit methods are used. Speed was of secondary concern in this work. Refinements by Eberhardt et al. [5]—notably, the use of higher-order explicit integration meth- ods and Maple-optimized code, as well as a dynamic, not static treat- ment of the problem—obtain similarly realistic results, while drop- ping the computational cost to approximately 20–30 minutes per frame on an SGI R8000 processor. No mention is made of damp- ing terms. Provot [13] focuses on improving the performance of ex- plicit methods by a post-step modification of nodal positions. He it- eratively adjusts nodal positions to eliminate unwanted stretch; the convergence properties of this method are unclear. A more compre- hensive discussion on cloth research can be found in the survey pa- per by Ng and Grimsdale [9].
2 Simulation Overview
In this section, we give a brief overview of our simulator’s architec- ture and introduce some notation. The next section derives the linear system used to step the simulator forward implicitly while section 4 describes the specifics of the internal forces and their derivatives that form the linear system. Section 5 describes how constraints are maintained (once established), with a discussion in section 6 on col- lision detection and constraint initialization. Section 7 describes our adaptive step-size control, and we conclude in section 8 with some simulation results.
2.1 Notation and Geometry
Our simulator models cloth as a triangular mesh of particles. Given
a mesh of n particles, the position in world-space of the ith particle
is x ∈ IR3 . The geometric state of all the particles is simply x ∈ IR3n . i
system has bandwidth

The same component notation applies to forces: a force f ∈ IR3n act-
ing on the cloth exerts a force fi on the ith particle. Real-world cloth
is cut from flat sheets of material and tends to resist deformations
away from this initial flat state (creases and pleats not withstanding).
We capture the rest state of cloth by assigning each particle an un-
changing coordinate (u , v ) in the plane.3 Section 4 makes use of ii
these planar coordinates.
Collisions between cloth and solid objects are handled by pre-
venting cloth particles from interpenetrating solid objects. Our cur- rent implementation models solid objects as triangularly faced poly- hedra. Each face has an associated thickness and an orientation; par- ticles found to be sufficiently near a face, and on the wrong side, are deemed to have collided with that face, and become subject to a con- tact constraint. (If relative velocities are extremely high, this simple test may miss some collisions. In this case, analytically checking for intersection between previous and current positions can guarantee that no collisions are missed.) For cloth/cloth collisions, we detect both face-vertex collisions between cloth particles and triangles, as well as edge/edge collisions between portions of the cloth. As in the case of solids, close proximity or actual intersection of cloth with it- self initiates contact handling.
2.2 Energy and Forces
The most critical forces in the system are the internal cloth forces which impart much of the cloth’s characteristic behavior. Breen et al. [3] describes the use of the Kawabata system of measurement for realistic determination of the in-plane shearing and out-of-plane bending forces in cloth. We call these two forces the shear and bend forces. We formulate the shear force on a per triangle basis, while the bend force is formulated on a per edge basis—between pairs of adjacent triangles.
The strongest internal force—which we call the stretch force— resists in-plane stretching or compression, and is also formulated per triangle. Under normal conditions, cloth does not stretch apprecia- bly under its own weight. This requires the stretch force to have a high coefficient of stiffness, and in fact, it is the stretch force that is most responsible for the stiffness of equation (1). A common prac- tice in explicitly integrated cloth systems is to improve running time by decreasing the strength of the stretch force; however, this leads to “rubbery” or “bouncy” cloth. Our system uses a very stiff stretch force to combat this problem, without any detrimental effects on the run-time performance. While the shear and bend force stiffness co- efficients depend on the material being simulated, the stretch coef- ficient is essentially the same (large) value for all simulations. (Of course, if stretchy cloth is specifically called for, the stretch coeffi- cient can be made smaller.)
Complementing the above three internal forces are three damp- ing forces. In section 5, we formulate damping forces that subdue any oscillations having to do with, respectively, stretching, shear- ing, and bending motions of the cloth. The damping forces do not dissipate energy due to other modes of motion. Additional forces in- clude air-drag, gravity, and user-generated generated mouse-forces (for interactive simulations). Cloth/cloth contacts generate strong repulsive linear-spring forces between cloth particles.
Combining all forces into a net force vector f, the acceleration x ̈i of the ith particle is simply x ̈ i = fi / mi , where mi is the ith particle’s mass. The mass mi is determined by summing one third the mass
3In general, each particle has a unique (u,v) coordinate; however, to accommodate pieces of cloth that have been topologically seamed together (such as a sleeve), particles lying on the seam must have multiple (u, v) co- ordinates. For these particles, we let the (u, v) coordinate depend on which triangle we are currently examining. The (u, v) coordinates are useful for texturing.
of all triangles containing the ith particle. (A triangle’s mass is the product of the cloth’s density and the triangle’s fixed area in the uv coordinate system.) Defining the diagonal mass matrix M ∈ IR3n×3n by diag(M) = (m1,m1,m1,m2,m2,m2,… ,mn,mn,mn), we can write simply that
x ̈ = M−1f(x,x ̇). (2) 2.3 Sparse Matrices
The use of an implicit integration method, described in the next section, generates large unbanded sparse linear systems. We solve these systems through a modified conjugate gradient (CG) itera- tive method, described in section 5. CG methods exploit sparsity quite easily, since they are based solely on matrix-vector multiplies, and require only rudimentary sparse storage techniques. The spar- sity of the matrix generated by the implicit integrator is best repre- sented in block-fashion: for a system with n particles, we deal with an n × n matrix, whose non-zero entries are represented as dense 3 × 3 matrices of scalars. The matrix is represented as an array of n rows; each row is a linked list of the non-zero elements of that row, to accommodate possible run-time changes in the sparsity pattern, due to cloth/cloth contact. The (dense) vectors that are multiplied against this matrix are stored simply as n element arrays of three- component vectors. The overall implementation of sparsity is com- pletely straightforward.
2.4 Constraints
An individual particle’s position and velocity can be completely controlled in either one, two, or three dimensions. Particles can thus be attached to a fixed or moving point in space, or constrained to a fixed or moving surface or curve. Constraints are either user-defined (the time period that a constraint is active is user-controlled) or auto- matically generated, in the case of contact constraints between cloth and solids. During cloth/solid contacts, the particle may be attached to the surface, depending on the magnitudes of the frictional forces required; otherwise, the particle is constrained to remain on the sur- face, with sliding allowed. The mechanism for releasing a contact constraint, or switching between sliding or not sliding, is described in section 5.
The constraint techniques we use on individual particles work just as well for collections of particles; thus, we could handle cloth/cloth intersections using the technique described in section 5, but the cost is potentially large. For that reason, we have chosen to deal with cloth/cloth contacts using penalty forces: whenever a par- ticle is near a cloth triangle or is detected to have passed through a cloth triangle, we add a stiff spring with damping to pull the parti- cle back to the correct side of the triangle. The implicit solver easily tolerates these stiff forces.
3 Implicit Integration
Given the known position x(t0 ) and velocity x ̇ (t0 ) of the system at time t0, our goal is to determine a new position x(t0 + h) and veloc- ity x ̇(t0 + h) at time t0 + h. To compute the new state and veloc- ity using an implicit technique, we must first transform equation (2) into a first-order differential equation. This is accomplished simply by defining the system’s velocity v as v = x ̇ and then writing
􏰀􏰁􏰀􏰁􏰀􏰁
dx=dx= v . (3) d t x ̇ d t v M − 1 f ( x , v )
To simplify notation, we will define x0 = x(t0 ) and v0 = v(t0 ). We alsodefine􏰤x=x(t0 +h)−x(t0)and􏰤v=v(t0 +h)−v(t0).
SIGGRAPH ’98 3
CD-ROM version

The explicit forward Euler method applied to equation (3) ap- proximates 􏰤x and 􏰤v as
􏰀􏰁􏰀􏰁
􏰤x =h v0 􏰤v M−1 f0
where the force f0 is defined by f0 = f(x0,v0). As previously dis- cussed, the step size h must be quite small to ensure stability when using this method. The implicit backward Euler method appears similar at first: 􏰤x and 􏰤v are approximated by
􏰀􏰁􏰀 􏰁
􏰤x v0 +􏰤v
􏰤v =h M−1f(x0+􏰤x,v0+􏰤v) . (4)
The difference in the two methods is that the forward method’s step is based solely on conditions at time t0 while the backward method’s step is written in terms of conditions at the terminus of the step itself.4
The forward method requires only an evaluation of the function f but the backward method requires that we solve for values of 􏰤x and 􏰤v that satisfy equation (4). Equation (4) is a nonlinear equa- tion: rather than solve this equation exactly (which would require iteration) we apply a Taylor series expansion to f and make the first- order approximation
f(x0 +􏰤x,v0 +􏰤v)=f0 + ∂f􏰤x+ ∂f􏰤v. ∂x ∂v
In this equation, the derivative ∂f/∂x is evaluated for the state (x0 , v0 ) and similarly for ∂f/∂v. Substituting this approximation into equation (4) yields the linear system
􏰀􏰁􏰂 v+􏰤v 􏰃 􏰤x0
4 Forces
Cloth’s material behavior is customarily described in terms of a
scalar potential energy function E(x); the force f arising from this
energy is f = −∂E/∂x. Equation (6) requires both the vector f
and the matrix ∂f/∂x. Expressing the energy E as a single mono-
lithic function—encompassing all aspects of the cloth’s internal
behavior—and then taking derivatives is impractical, from a book-
keeping point of view. A better approach is decompose E into a sum
sible.
However, even decomposing E into sparse energy functions is
not enough. Energy functions are an undesirable starting point be-
cause sensible damping functions cannot be derived from energy
functions. Instead, we define internal behavior by formulating a
vector condition C(x) which we want to be zero, and then defining
the associated energy as k C(x)T C(x) where k is a stiffness constant. 2
In section 4.5, we show how sensible damping functions can be con- structed based on this formulation. An added bonus is that starting from this vector-based energy description tends to result in a sim- pler, more compact, and more easily coded formulation for ∂f/∂x than proceeding from an energy function in which the structure of C has been lost.
4.1 Forces and Force Derivatives
Given a condition C(x) which we want to be zero, we associate an energy function E with C by writing E (x) = k C(x)T C(x) where
􏰀􏰁
−1 ∂f ∂f
f0 + h(v0 + 􏰤v) + 􏰤v .
α Eα(x)where each Eα depends on as few elements of x—as few particles—as pos-
ofsparseenergyfunctions;thatis,towriteE(x)=
􏰢
􏰤v =h M−1(f0 + ∂f􏰤x+ ∂f􏰤v) ∂x ∂v
on,
. (5) Taking the bottom row of equation (5) and substituting 􏰤x =
fi =−
∂ EC ∂xi
∂C(x)
=−k C(x); (7)
CC2
k is a stiffness constant of our choice. Assuming that C depends on only a few particle, C gives rise to a sparse force vector f. Re- call from section 2.1 that we view the vector f in block form; each element fi is a vector in IR3. For each particle i that C depends
h(v0 + 􏰤v) yields 􏰤v = hM
∂xi
Similarly, the derivative of f is also sparse. Defining the deriva-
tive matrix K = ∂f/∂x, the nonzero entries of K are Kij for all pairs of particles i and j that C depends on. Again, we treat K in block fashion: K ∈ IR3n×3n, so an element Kij is a 3 × 3 matrix. From equation (7), we have
􏰁
C(x) . (8)
Additionally, since Kij is a second derivative—that is, Kij = ∂fi/∂xj = ∂2 E/∂xi∂xj—we have Kij = KTji so K is symmetric. Note that since C does not depend on v, the matrix ∂f/∂v is zero.
We can now easily describe the internal forces acting on the cloth, by just writing condition functions. Forces and their derivatives are easily derived using equations (7) and (8).
4.2 Stretch Forces
Recall that every cloth particle has a changing position xi in world space, and a fixed plane coordinate (ui , vi ). Even though our cloth is modeled as a discrete set of points, grouped into triangles, it will be convenient to pretend momentarily that we have a single continuous function w(u, v) that maps from plane coordinates to world space. Stretch can be measured at any point in the cloth surface by examin- ing the derivatives wu = ∂w/∂u and wv = ∂w/∂v at that point. The magnitude of wu describes the stretch or compression in the u direc- tion; the material is unstretched wherever ∥wu ∥ = 1. Stretch in the
all the other elements of f are zero.
∂x ∂v
Letting I denote the identity matrix, and regrouping, we ob-
tain􏰀 􏰁􏰀􏰁 ∂f∂f ∂f
∂f 􏰀∂C(x) ∂C(x)T ∂2C(x) K = i =−k +
I−hM−1 −h2M−1 􏰤v=hM−1 f0+h v0 (6) ∂v∂x ∂x
ij∂x ∂x∂x∂x∂x jijij
which we then solve for 􏰤v. Given 􏰤v, we trivially compute 􏰤x = h(v0 +􏰤v).
Thus, the backward Euler step consists of evaluating f0, ∂f/∂x and ∂f/∂v; forming the system in equation (6); solving the system for 􏰤v; and then updating x and v. We use the sparse data structures described in section 2.3 to store the linear system. The sparsity pat- tern of equation (6) is described in the next section, while solution techniques are deferred to section 5.
4 The method is called “backward” Euler because starting from the output state (x0 + 􏰤x, v0 + 􏰤v) and using a forward Euler step to run the system backward in time (i.e. taking the step −h(v(t0 + h), f(x(t0 + h), v(t0 + h))) brings you back to (x0 , v0 ). What is the value in this? Forward Euler takes no notice of wildly changing derivatives, and proceeds forward quite blindly. Backward Euler, however, forces one to find an output state whose deriva- tive at least points back to where you came from, imparting, essentially, an additional layer of consistency (or sanity-checking, if you will).
SIGGRAPH ’98
4
CD-ROM version

v direction is measured by ∥wv∥. (Some previous continuum for-
mulations have modeled stretch energy along an axis as essentially
(wT w − 1)2, which is a quartic function of position [15, 16, 17, 4]. uu
We find this to be needlessly stiff; worse, near the rest state, the force gradient—a quadratic function of position—is quite small, which partially negates the advantage implicit integration has in exploit- ing knowledge of the force gradient. A quadratic model for energy is, numerically, a better choice.)
We apply this stretch/compression measure to a triangle as fol-
lows. Let us consider a triangle whose vertices are particles i, j and
k. Define􏰤x1 =xj −xi and􏰤x2 =xk −xi. Also,let􏰤u1 =uj −ui,
while 􏰤u = u − u and similarly for 􏰤v and 􏰤v . We approxi- 2ki12
mate w(u, v) as a linear function over each triangle; this is equiva- lent to saying that wu and wv are constant over each triangle. This lets us write 􏰤x1 = wu􏰤u1 + wv􏰤v1 and 􏰤x2 = wu􏰤u2 + wv􏰤v2. Solving for wu and wv yields
􏰀 􏰁−1
(wu wv)=(􏰤x1 􏰤x2) 􏰤u1 􏰤u2 . (9)
n2 and e as having a constant length at each step of the simulation. This makes differentiating θ with respect to x a manageable task.
Rectangular meshes make it simple to treat bending anisotropi- cally. The uv coordinates associated with particles make this possi- ble for triangular meshes as well. Given material for which bending in the u and v directions are weighted by stiffnesses ku and kv, we can emulate this anisotropy as follows. Let the edge between the triangles be between particles i and j, and define 􏰤u = ui − u j and 􏰤v = vi − v j . The stiffness weighting for this edge should simply be
Note that x1 and x2 vary during the simulation but the matrix in the above equation does not.
We can treat wu and wv as functions of x, realizing that they de- pend only on xi, xj and xk and using equation (9) to obtain deriva- tives. The condition we use for the stretch energy is
􏰀􏰁
C(x) = a ∥wu (x)∥ − bu (10) ∥wv(x)∥−bv
where a is the triangle’s area in uv coordinates. Usually, we set bu = bv = 1, though we need not always do so. In particular, if we want to slightly lengthen a garment (for example, a sleeve) in the u direction, we can increase bu , which causes wu to seek a larger value, and tends to induce wrinkles across the u direction. Likewise, we might decrease bv near the end of a sleeve, inducing a tight cuff, as on a sweatshirt. We have found the ability to control shrink/stretch anisotropically to be an indispensable modeling tool.
4.3 Shear and Bend Forces
Cloth likewise resists shearing in the plane. We can measure the ex-
tent to which cloth has sheared in a triangle by considering the inner
product wuT wv . In its rest state, this product is zero. Since the stretch
term prevents the magnitudes of wu and wv from changing overly
much, we need not normalize. By the small angle approximation,
the product wT w is a reasonable approximation to the shear angle. uv
The condition for shearing is simply
C(x) = awu (x)T wv (x)
with a the triangle’s area in the uv plane.
We measure bend between pairs of adjacent triangles. The con-
dition we write for the bend energy depends upon the four particles defining the two adjoining triangles. If we let n1 and n2 denote the unit normals of the two triangles and let e be a unit vector parallel to the common edge, the angle θ between the two faces is defined by the relations sinθ = (n1 ×n2)·e and cosθ = n1 ·n2. We define a condition for bending by writing simply C(x) = θ which results in a force that counters bending.5 The assumption that the stretch en- ergy will keep the cloth from stretching much allows us to treat n1,
5For reasonably equilateral triangles, as edge lengths decrease, the cur- vature represented by a particular angle θ between triangles increases. Since the square of the curvature—a a good measure of the bend energy in cloth— increases at the same rate that the triangle’s area decreases, the condition C should not be scaled by the triangles’ areas. See Breen et al. [3] for a further discussion of relating curvature to bend angle.
∂C(x) ̇
􏰤v1 􏰤v2
To the above forces we also add easily implemented forces such as gravity and air-drag (which is formulated on a per-triangle basis, and opposes velocities along the triangle’s normal direction). When the simulation is fast enough to interact with, we add user-controlled “mouse” forces. These forces and their gradients are easily derived.
4.5 Damping
The energies we have just described are functions of position only. Robust dynamic cloth simulation, however, is critically dependent on well-chosen damping forces that are a function of both position and velocity. For example, the strong stretch force must be ac- companied by a suitably strong damping force if we are to prevent anomalous in-plane oscillations from arising between connected particles. However, this strong damping force must confine itself solely to damping in-plane stretching/compressing motions: stretch damping should not arise due to motions that are not causing stretch or compression. Terzopoulos et al.’s [16, 17] treatment of cloth used a simple viscous damping function which dissipated kinetic energy, independent of the type of motion. Carignan et al. [4] im- proved upon this somewhat, borrowing a formulation due to Platt and Barr [11]; however, their damping function—a linear function of velocity—does not match the quartic energy functions of their continuum formulation. In this section we describe a general treat- ment for damping that is independent of the specific energy function being damped.
It is tempting to formulate a damping function for an energy func- tion E(x) by measuring the velocity of the energy, E ̇ = d E(x).
dt
This is an easy trap to fall into, but it gives nonsensical results. At an equilibrium point of E, the gradient ∂E/∂x vanishes. Since E ̇ = (∂E/∂x)Tx ̇, we find that E ̇ is zero when E is at its ̇minimum, regardless of the system’s velocity x ̇ = v. In general, E is always too small near the system’s rest state. Clearly, basing the damping force on E ̇ is not what we want to do.
We believe that the damping function should be defined not in terms of the energy E, but in terms of the condition C(x) we have been using to define energies. The force f arising from the energy acts only in the direction ∂C(x)/∂x, and so should the damping force. Additionally, the damping force should depend on the com- ponent of the system’s velocity in the ∂C(x)/∂x direction; in other words, the damping strength should depend on (∂C(x)/∂x)Tx ̇ = C ̇ (x). Putting this together, we propose that the damping force d associated with a condition C have the form
(11)
4.4
ku (􏰤u)2 + kv (􏰤v)2 (􏰤u)2 + (􏰤v)2
Additional Forces
.
d = −kd
This neatly parallels the fact that f = −ks
∂x
C(x).
∂C(x) ∂x
C(x).
SIGGRAPH ’98 5
CD-ROM version

Given the condition functions C we have defined in this sec- tion for stretch, bend and shear forces, we can now add accompa- nying damping forces by applying equation (11). As before, di is nonzero only for those particles that C depends on, and ∂d/∂x has the same sparsity pattern as ∂f/∂x. Differentiating equation (11), we obtain
no coordinates, while a particle with one dimension of constraint would have two coordinates. This is possible—but it complicates the system immensely. If we change the number of coordinates per particle, we alter the size of the derivative matrices in equa- tion (6), as well as the sparsity pattern (this happens when a particle changes from having no coordinates to some coordinates, or vice versa). Given the transient nature of contact constraints between cloth and solids, this is most unappealing. The computation of the derivative matrices’ entries is also greatly complicated, because we must now introduce extra Jacobian matrices that relate a particle’s reduced coordinates to its motion in world-space. Finally, correct constraint-release behavior between cloth and solid objects is diffi- cult to achieve using a reduced coordinate formulation. Considering all of this, we immediately rejected this method of constraints.
Penalty Methods
We could constrain particles through the use of strong energy functions—essentially, stiff springs that attempt to prevent illegal particle motions. Since our entire formulation is geared to han- dle stiffness, the usual objections to enforcing constraints with springs—very stiff equations—do not carry as much weight. We tried this for a time, and found it to be a not unreasonable con- straint enforcement mechanism. However, penalty methods do not enforce constraints exactly, and they do add some additional stiff- ness to the system. Since the mechanism we describe enforces con- straints exactly, and adds no extra stiffness, we turned away from penalty methods except in the case of cloth/cloth interactions.
Lagrange Multipliers
We could introduce additional constraint forces—that is, Lagrange multipliers—into our system to satisfy the constraints. This in- volves augmenting the linear system of equation (6) with extra vari- ables (the multipliers) and extra equations (the constraint condi- tions). Unfortunately, this turns a positive definite system into an indefinite system, which means that iterative methods such as CG will need to square the system first, thereby doubling the running time and degrading the numerical conditionining of the linear sys- tem. Additionally, an iterative method will generally not enforce the constraints exactly without a large number of iterations. (A direct method for solving the augmented system would, however, avoid this problem.) Again, the constraint method we describe steps past these difficulties, so we turned away from using Lagrange multipli- ers.
5.1 Mass Modification
The idea behind our constraint enforcement mechanism is described
quite simply, although the actual implementation is somewhat more
complicated, to maximize performance. A dynamic simulation usu-
ally requires knowledge of the inverse mass of objects; for example,
note the appearance of M−1, and not M in equation (6). In the case
of a single particle, we write x ̈i = 1 fi to describe a particle’s ac- mi
celeration. When inverse mass is used, it becomes trivial to enforce constraints by altering the mass.
Suppose for example that we want to keep particle i’s velocity from changing. If we take 1/mi to be zero, we give the particle an infinite mass, making it ignore all forces exerted on it. Complete control over a particle’s acceleration is thus taken care of by storing a value of zero for the particle’s inverse mass. What if we wish to constrain the particle’s acceleration in only one or two dimensions?
not always do so. Suppose we write x ̈i =
∂ d i ∂xj
=−kd
∂ C ( x ) ∂ C ̇ ( x ) T ∂xi ∂xj
+
∂ 2 C ( x ) ̇
􏰂􏰃
Note that ∂d/∂x is not a second derivative of some function as was the case in equation (8) so we cannot expect ∂d/∂x to be symmetri- cal. In equation (12), it is the term (∂C(x)/∂xi )(∂C ̇ (x)/∂x j )T which breaks the symmetry. Anticipating section 5.2, we find it expedi- ent simply to leave this term out, thereby restoring symmetry. This simplification is clearly not physically justifiable, but we have not observed any ill effects from this omission. (Omitting all of equa- tion (12), however, causes serious problems.)
Finally, equation (6) requires the derivative ∂d/∂v. Since C ̇ (x) = (∂C(x)/∂x)T v, we have
∂ C ̇ ( x ) ∂ 􏰀 ∂ C ( x ) T 􏰁 ∂ C ( x ) =v=.
∂xi∂xj
C(x) . (12)
∂v ∂v ∂x ∂x Using this fact, we can write
∂C(x) ∂C ̇ (x) T ∂C(x) ∂C(x) T =−kd .
∂d
∂vj ∂xi ∂vj ∂xi ∂xj
i =−kd
In this case, the result is symmetrical without dropping any terms.
5 Constraints
In this section, we describe how constraints are imposed on indi- vidual cloth particles. The constraints we discuss in this section are either automatically determined by the user (such as geometric at- tachment constraints on a particle) or are contact constraints (gener- ated by the system) between a solid object and a particle. The tech- niques we describe in this section could be used for multi-particle constraints; however, constraints that share particle would need to be merged. Thus, a set of four-particle constraints (such as ver- tex/triangle or edge/edge contacts in the cloth) might merge to form a single constraint on arbitrarily many particles, which would be ex- pensive to maintain. Because of this, we handle cloth/cloth contacts with strong springs (easily dealt with, given the simulator’s underly- ing implicit integration base) and “position alteration,” a technique described in section 6.
At any given step of the simulation, a cloth particle is either com- pletely unconstrained (though subject to forces), or the particle may be constrained in either one, two or three dimensions. Given the differential nature of our formulation, it is the particle’s accelera- tion, or equivalently, the change in the particle’s velocity, that is con- strained. If the particle is constrained in all three dimensions, then we are explicitly setting the particle’s velocity (at the next step). If the constraint is in two or one dimensions, we are constraining the particle’s velocity along either two or one mutually orthogonal axes. Before describing our constraint method, we discuss several other possible enforcement mechanisms and explain why we chose not to use them.
Reduced Coordinates
An obvious and quite exact method for constraining a particle is to reduce the number of coordinates describing the particle’s po- sition and velocity. A completely constrained particle would have
Although we normally think of a particle’s mass as a scalar, we need
SIGGRAPH ’98 6
CD-ROM version
􏰄1/mi 0 0􏰅
0 1/mi 0 fi. Now x ̈i 000

must lie in the xy plane; no acceleration in the z direction is possible.
Note than an unconstrained particle can be considered to have the
3 × 3 inverse mass matrix 1 I, with I the identity matrix. mi
Of course, we are not restricted to coordinate-aligned constraints.
More generally, given a unit vector p ∈ IR3, a particle is prevented
from accelerating along p by using an inverse mass matrix 1 (I − mi
ppT ); this follows from the fact that (I − ppT )p = 0. Similarly, given two mutually orthogonal unit vectors p and q, we prevent a particle from accelerating in either the p or q direction by using the inverse mass matrix 1 (I − ppT − qqT ).
mi
By allowing constrained particles to have these sorts of inverse
masses, we can build constraints directly into equation (6). We will
create a modified version W of M−1; W will be a block-diagonal
matrix, with off-diagonal blocks being zero, and diagonal blocks
defined as follows: let ndof(i) indicate the number of degrees of
freedom particle i has, and let particle i’s prohibited directions be
pi (ifndof(i)=2)orpi andqi (ifndof(i)=1)withpi andqi mu- 1
a symmetric (and positive definite) system by left-multiplying the entire equation by M: the system
􏰀 ∂f ∂f􏰁 􏰀 ∂f􏰁
M−h −h2 􏰤v=h f0+h v0 (15)
is symmetric and has the same solution 􏰤v as equation (6). Unfor- tunately, we cannot apply the same transformation to equation (14), because W is singular—the filtering blocks in equation (13) are rank deficient—so we cannot multiply through by W−1.
The solution to the problem of asymmetry is to modify the CG method so that it can operate on equation (15), while procedurally applying the constraints inherent in the matrix W at each iteration. The modified method will need to know about the particles’ con- straints and the vector z. Let us define the symmetric positive defi- nite matrix A by
∂v∂x ∂x
tually orthogonal unit vectors. W’s diagonal blocks are Wii = m Si i
∂f
A= M−h −h2
(16)
􏰀􏰁
∂f ∂x
r=A􏰤v−b.
Given A, b, constraints on the particles, and z, our modified CG
• For each particle i, the component of ri in the particle’s uncon- strained direction(s) will be made equal to zero (assuming the method is run for sufficiently many iterations).
• For each particle i, the component of 􏰤vi in the particle’s con- strained direction(s) will be exactly zi (no matter how many iterations are taken).
Note that these two conditions imply that unconstrained particles
have ri close to zero, while completely constrained particles have
where
∂v and the vector b and residual vector r as
S = i
(13)
b=h
method will try to find 􏰤v that satisfies two conditions:
   I
(I−pipiT)
(I − pipiT − qiqiT )
0
if ndof(i) = 3 ifndof(i)=2 if ndof(i) = 1 if ndof(i) = 0.
􏰀􏰁
􏰤v =z. Thusinthecasewhennoparticlesareconstrained,our ∂v∂x ∂x ii
f0 +h∂fv0 and ∂x
We are not limited to constraining particles to have zero accel- erations in certain directions; rather, we control exactly what the change in velocity is along the constrained directions. For every particle i, let zi be the change in velocity we wish to enforce in the particle’s constrained direction(s). (This implies we can choose any value of zi for a completely constrained particle, since all directions are constrained; an unconstrained particle must have zi = 0 since it has no constrained directions.) Using W and z, we rewrite equa- tion (6) to directly enforce constraints. If we solve
∂f 2 ∂f ∂f
I−hW −hW 􏰤v=hW f0+h v0 +z (14)
􏰀 􏰁􏰀􏰁
for 􏰤v, we will obtain a 􏰤v which is consistent with our constraints. Completelyconstrainedparticleswillhave􏰤vi =zi,whilepartially constrained particles will have a 􏰤vi whose component in the con- strained direction(s) is equal to zi.
5.2 Implementation
We initially implemented constraints using equation (14) and found that it worked exactly as advertised. For very small test systems, we solved equation (14) using a direct method (Gaussian elimina- tion) without any problems. For larger systems, we planned to use the iterative, sparsity-exploiting CG method, which immediately presents us with a problem: equation (14) is not a symmetric lin- ear system. (For that matter, neither is equation (6) unless all par- ticles have the same mass.) CG methods, however, require sym- metric matrices.6 We could apply a CG method to the unsymmetric matrix of equation (14) by use of the “normal equations”; but this involves multiplying the matrix of equation (14) with its transpose which doubles the cost of each iteration while squaring the condition number of the system [14]—a less than desirable plan. We decided that using a CG method to solve the unsymmetric problem was not acceptable.
Note that without constraints, applying a CG method to equa- tion (6) is not difficult, because we can transform this equation to
6 In fact, they work best on positive definite symmetric matrices. The ma- trices we ultimately hand to our CG method are positive definite.
modified CG method should produce the same result as the regular CG method.
5.3 The Modified Conjugate Gradient Method
The CG method (technically, the preconditioned CG method) takes a symmetric positive semi-definite matrix A, a symmetric positive definite preconditioning matrix P of the same dimension as A, a vector b and iteratively solves A􏰤v = b. The iteration stops when ∥b − A􏰤v∥ is less than ε∥b∥ where ε is a user-defined tolerance value. The preconditioning matrix P, which must be easily invert- ible, speeds convergence to the extent that P−1 approximates A. We wholeheartedly refer the reader to Shewchuk [14] for information on the CG method.
We derive our modified conjugate gradient method by observing that the effect of the matrix W in equation (14) is to filter out veloc- ity changes in the constrained directions. Our idea then is to define an invariant— for all i, the component of 􏰤vi in the constrained di- rection(s) of particle i is equal to zi—and then establish and main- tain the invariant at each iteration, by defining a filtering procedure filter. The role of filter is to take a vector a and perform the same filtering operation (see equation (13)) as multiplying by W, but leav- ing out the scaling by 1/ mi :
procedure filter(a) for i = 1 to n
aˆ i = S i a i return aˆ
SIGGRAPH ’98 7
CD-ROM version

Using filter, we define the modified CG method modified-pcg as follows:
1 procedure modified-pcg
2 􏰤v=z
The particles’ accelerations are inherently dependent on one an- other through the matrix A of equation (16). This means that the cor- rect approach to determing constraint release is combinatoric, as in Baraff [2]. We reject this approach as impractical given the dimen- sion of A. Instead, we allow contacts to release when the constraint force between a particle and a solid switches from a repulsive force to an attractive one. In practice, this has proven to work well.
Friction presents a similar problem. When cloth contacts a solid, we lock the particle onto the surface, if the relative tangential ve- locity is low. We monitor the constraint force, and if the tangential force exceeds some fraction of the normal force, we allow the par- ticle to slide on the surface. For high sliding velocities, we apply a dissipative tangential force, opposite the relative sliding direction, proportional to the normal force.
6 Collisions
Much has been written about collision detection for cloth; we have nothing substantial to add to the subject of collision detection per se. Cloth/cloth collisions are detected by checking pairs ( p, t) and (e1 , e2 ) for intersections, where p and t are a cloth particle and a cloth triangle respectively, and e1 and e2 are edges of cloth trian- gles. Given a previous known legal state of the cloth, we postulate a linear motion for the cloth particles to the current (possibly illegal) state and check for either particle/triangle or edge/edge crossings. To avoid O(n2) comparisons, we use a coherency-based bounding- box approach [1] to cull out the majority of pairs.
When collisions between a cloth vertex and triangle, or two cloth edges are detected, we insert a strong damped spring force to push the cloth apart. A dissipative force tangent to the contact is also applied, countering any sliding motion. The force is not, strictly speaking, a frictional force: rather it is proportional to the slip ve- locity, so it is in actuality a damping force, although it reasonably emulates dynamic friction. Applying static friction forces to cloth contacts is far more difficult, and is a problem we have not solved yet. The forces, and their derivatives with respect to position and velocity, are of course included in equation (15).
Our system detects collisions between cloth particles and solid objects by testing each individual cloth particle against the faces of each solid object. A solid object’s faces are grouped in a hierarchical bounding box tree, with the leaves of the tree being individual faces of the solid. The tree is created by a simple recursive splitting along coordinate axes. The maintenance of contacts and the application of friction forces was described in the previous section.
6.1 Constraint Initiation
Both cloth/cloth and cloth/solid collisions give rise to the same problem whenever two contacts form. For both types of collisions, our detection algorithm reports an intersection, and then takes action to remedy the situation: either by enforcing a constraint (cloth/solid collisions) or by adding a penalty force (cloth/cloth) collisions. However, since our simulator proceeds in discrete steps, collisions resulting in a reasonably substantial interpenetration depth can oc- cur between one step and the next. Clearly, this situation needs to be remedied.
For cloth/cloth collisions, this would not appear to be a problem: the spring forces that are added work to counter the colliding veloc- ities and then push the cloth apart. For cloth/solid collisions, how- ever, the situation is more complicated. If we simply enforce a con- straint which causes the colliding cloth particle to have a velocity consistent with the solid object’s velocity, and continue to enforce that constraint, the cloth particle will continue to remain embedded somewhere below the solid object’s surface. This is unacceptable.
3 δ
0
= filter(b)T P filter(b)
4 r = filter(b − A􏰤v)
5 c = filter(P−1r)
6 δ =rTc
new
7 while δ
> ε2δ
8 q = filter(Ac)
new
0
9 α=δ /(cTq) new
10 􏰤v=􏰤v+αc
11 r=r−αq
12 s=P−1r
13 δold = δnew
14 δnew=rTs
15 c = filter(s + δnew c)
δold
Line 2 of the procedure establishes our invariant. Lines 5 and 15 maintain the invariant by filtering c before adding it to 􏰤v. The un- modified conjugate gradient method establishes a stopping criterion based on bT Pb. Since our constrained formulation ignores certain components of b, our stopping criterion should as well, so we add filtering to line 3. The vector r measures the solution error b − A􏰤v, and should not include error due to the constraints; hence we add filtering at lines 4 and 8. (Note that removing the calls to filter and changing line 2 to 􏰤v = 0 yields the standard preconditioned con- jugate gradient method.)
We use a simple preconditioner P by making P be a diagonal ma- trix with Pii = 1/Aii so products involving P−1 are trivially com- puted. More elaborate preconditioners could be used, though we doubt there is a large speedup to be gained. Matrix-vector products with A are of course implemented in sparse matrix-vector fashion, using the data structures defined in section 2.3.
Given modified-pcg, obvious questions are “does it work?” fol- lowed by “how does it compare with the unmodified CG method?” Proofs about CG methods are difficult in general; in practice, our method always converges, which answers the first question. Prior to implementing modified-pcg, we used a penalty method and applied the standard CG method to equation (15). When we began using procedure modified-pcg, we did not notice any substantial change in the number of iterations required by the method. Empirically, we conclude that the two methods have similar convergence behavior. Result in section 8 indicate that the running time is close to O(n1.5 ), which is what unmodified CG would be expected to deliver on this sort of problem [14].
5.4 Determining the Constraint Forces
For contact constraints (between cloth and solid objects) we need to know what the actual force of constraint is, in order to deter- mine when to terminate a constraint. Additionally, we need to know the constraint force actually exerted in order to model fric- tional forces properly. Fortunately, it is easy to add one more step to modified-pcg to determine the constraint force. When modified-pcg terminates, the residual error e = A􏰤v − b has the property that ei need not be close to zero if particle i is constrained. In fact, ei is exactly the extra constraint force that must have been supplied to enforce the constraint. Thus, we can compute constraint forces at the end of modified-pcg by performing one last matrix- vector product to compute A􏰤v − b. (The vector r in modified-pcg is equal to filter(A􏰤v − b), so the extra matrix-vector product to compute e really is necessary.)
SIGGRAPH ’98 8
CD-ROM version

One solution is to use Baumgarte stabilization [18], which sched- ules the particle’s acceleration so that the position and velocity er- ror of the particle with respect to the surface decay asymptotically to zero. We experimented with this technique, but found it lacking. In particular, a fast rise to the surface was prone to noise and “jumpi- ness”; this could be eliminated, but at the cost of decreasing the step size. A slower rise to the surface caused visual artifacts.
We tried a simpler solution: when intersections occurred, rather than wait for a scheduled constraint or a penalty force to eliminate the intersection, we simply altered the positions of the cloth parti- cles, effecting an instantaneous (and discontinuous) change in posi- tion. While this would be problematic when using a multi-step dif- ferential equation solver which expects continuity (such as a Runge- Kutta method), it should not interfere with a one-step solver such as the backward Euler method. Unfortunately, simply changing parti- cle positions produced disastrous results. The stretch energy term in a cloth system is extremely strong, and altering particle positions ar- bitrarily introduced excessively large deformation energies in an al- tered particle’s neighborhood. This resulted in visibly “jumpy” be- havior of the cloth in localized regions.
6.2 Position Alteration
Despite its initial failure, the ability to make arbitrary small changes in a particle’s position continued to attract our attention. The en- tire process of implicit integration can be considered to be a filtering process [7], and we postulated that a mechanism for filtering energy changes caused by displacing particles might make position alter- ation a viable technique. We considered that perhaps some sort of extra implicit step could be used as a filter, but forming and solv- ing an additional linear system at each step seemed too expensive. Happily, we can make use of the filtering effect of implicit integra- tion without any extra work.
Consider a particle that has collided with a solid object. The par- ticle’s change in velocity at each step is under our control, using the constraint techniques described in section 5. Meanwhile, the parti- cle’s position at the next step follows from equation (4):
􏰤xi =h(v0i +􏰤vi)
(recall that v0i is the particle’s current velocity). The reason that changing positions after a step has been taken doesn’t work is be- cause the particle’s neighbors receive no advance notification of the change in position: they are confronted with the alteration at the be- ginning of the next step. This presents an obvious solution: we sim- ply modify the top row of equation (4) to
􏰤xi=h(v0i+􏰤vi)+yi (17)
where yi is an arbitrary correction term of our choice, introduced solely to move a particle to a desired location during the backward Euler step. Having modified the top row of equation (4), we must follow this change through: using equation (17) and repeating the derivation of section 3 and the symmetric transform from section 5 yields the modified symmetric system
7 Adaptive Time Stepping
The methods introduced in all of the previous sections usually allow us to take sizeable steps forward, without loss of stability. Even so, there are still times when the step size must be reduced to avoid di- vergence. There are a large number of methods for altering the size of a time step, for both explicit and implicit integrators, but these methods tend to concentrate on the accuracy of the simulation, and not the stability. Our goal is animation, not engineering; thus visu- ally pleasing results, meaning a numerically stable solution, rather than overall accuracy, is the deciding voice. The trick is to recognize instability before you see it on your screen—by then it’s too late.
Stiffness, and thus any potential instability, arises almost com- pletely from the strong stretch forces in the cloth. After each im- plicit step, we treat the resulting 􏰤x as a proposed change in the cloth’s state, and examine the stretch terms (section 4.2) for each tri- angle in the newly proposed state. If any triangle undergoes a drastic change in its stretch (in either the u or v direction) we discard the proposed state, reduce the step size, and try again. Subtlety is not required: we find that an unstable step invariably results in stretch changes that are quite large, and are thus easily detected.
Our simulation is run with a parameter that indicates the maxi- mum allowable step size: this parameter is set by the user, and is always less than or equal to one frame. (Most of our simulations in- volving human motions use a step size of 0.02 seconds.) Whenever the simulator reduces the step size, after two successes with the re- duced step size the simulator tries to increase the step size. If the simulator fails at the larger step size, it reduces the size again, and waits for a longer period of time before retrying to increase the step size. At its limit, the simulator will try increasing the step size ev- ery 40 steps; thus, if the user chooses too large a step, the simulator settles down to wasting only one out of every 40 steps in attempting too large a step. This method, though simple, has served us well.
8 Results
Table 1 gives a performance summary of assorted animations, shown in figures 1–6. Unaccounted overhead of the simulation (typ- ically about 5%) includes tasks such as geometry transformations, memory allocation, etc. The clothes in figures 3–6 were modeled as discrete planar panels, and then topologically seamed. The simula- tor was used to relax the clothing from an initial deformed state, that got the clothes around the characters, to a well-fitting state on the characters. The bu and bv parameters (see equation (10)) were then made smaller in certain regions to produce cuffs and waistbands, or strategically increased to induce wrinkling behavior in other re- gions.
We also ran the simulation in figure 1 with a range of stiffnesses for the bend term. Using the stiffness parameters in figure 1 as a ref- erence, we ran the simulation with those bend stiffnesses multiplied by 0.1, 1.0, 10, 100 and 1,000 (for a total range of 10,000 in the stiff- ness). The variance in the running times was under 5%. We doubt that simulators based on explicit integration methods could make a similar claim.
Finally, we tried to estimate our simulator’s performance as a function of n, the number of cloth particles. We ran the simulation in figure 1 with cloth resolutions of 500, 899, 2,602 (shown in figure 1) and 7,359 particles. The running times were, respectively, 0.23 seconds/frame, 0.46 seconds/frame, 2.23 seconds/frame, and 10.3 seconds/frame. This is slightly better than O(n1.5) performance, which is in line with the convergence rates of the conjugate gradient method [14] for systems such as equation (18).
􏰀 ∂f ∂f􏰁 􏰀 ∂f ∂f􏰁
M−h −h2 􏰤v=h f0+h v0+ y . (18)
∂v ∂x ∂x ∂x
This modification gives us complete control over both the posi- tion and velocity of a constrained particle in just one step, without any extra computational cost. We use this technique to bring parti- cles quickly and stably to the surface of solid objects without creat- ing visual artifacts or limiting the allowable step size. We can also add correction terms to particles involved in cloth/cloth collisions. Without a constraint on those particles’ velocities there is no guar- antee that they will go exactly where we want in one step, but the ability to induce sizeable jumps in position without excessively stiff spring forces adds greatly to the stability of the simulation.
SIGGRAPH ’98 9
CD-ROM version

figure
no. vertices/no. triangles cloth solid
time/frame (CPU sec.)
step size min/max (ms)
total frames/ total steps
task breakdown percentage EVAL CG C/C C/S
1
2,602/4,9442 322/640
2.23
16.5/33
75/80
25.7 50.4 18.3 1.4
2
2,602/4,9442 322/640
3.06
16.5/33
75/80
17.9 63.6 15.3 0.2
3
6,450/12,654 9,941/18,110
7.32
16.5/33
50/52
18.9 37.9 30.9 2.6
4 (shirt) (pants)
6,450/12,654 9,941/18,110 8,757/17,352 9,941/18,110
14.5 38.5
2.5/20 0.625/20
430/748 430/1214
16.7 29.9 46.1 2.2 16.4 35.7 42.5 1.7
5 (skirt) (blouse)
2,153/4,020 7,630/14,008 5,108/10,016 7,630/14,008
3.68 16.7
5/20 5/20
393/715 393/701
18.1 30.0 44.5 1.5 11.2 26.0 57.7 1.3
6 (skirt) (blouse)
4,530/8,844 7,630/14,008 5,188/10,194 7,630/14,008
10.2 16.6
10/20 1.25/20
393/670 393/753
20.1 36.8 29.7 2.6 13.2 30.9 50.2 1.4
Table 1: System performance for simulations in figures 1–6. Minimum and maximum time steps are in milliseconds of simulation time. Time/frame indicates actual CPU time for each frame, averaged over the simulation. Percentages of total running time are given for four tasks: EVAL— forming the linear system of equation (18); CG—solving equation (18); C/C—cloth/cloth collision detection; and C/S—cloth/solid collision detection.
9 Acknowledgments
This research was supported in part by an ONR Young Investiga- tor award, an NSF CAREER award, and grants from the Intel Cor- poration. We thank Alias|Wavefront for supplying the models and motion capture data used in figures 5 and 6.
References
[1] D.Baraff.DynamicSimulationofNon-penetratingRigidBod- ies. PhD thesis, Cornell University, May 1992.
[2] D. Baraff. Fast contact force computation for nonpenetrating rigid bodies. Computer Graphics (Proc. SIGGRAPH), 28:23– 34, 1994.
[3] D.E. Breen, D.H. House, and M.J. Wozny. Predicting the drape of woven cloth using interacting particles. Computer Graphics (Proc. SIGGRAPH), pages 365–372, 1994.
[4] M.Carignan,Y.Yang,N.Magenenat-Thalmann,andD.Thal- mann. Dressing animated synthetic actors with complex de- formable clothes. Computer Graphics (Proc. SIGGRAPH), pages 99–104, 1992.
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
J.C. Platt and A.H. Barr. Constraint methods for flexible mod- els. In Computer Graphics (Proc. SIGGRAPH), volume 22, pages 279–288. ACM, July 1988.
W.H. Press, B.P. Flannery, S.A. Teukolsky, and W.T. Vetter- ling. Numerical Recipes. Cambridge University Press, 1986.
X. Provot. Deformation constraints in a mass-spring model to describe rigid cloth behavior. In Graphics Interface, pages 147–155, 1995.
J. Shewchuk. An introduction to the conjugate gradient method without the agonizing pain. Technical Report CMU- CS-TR-94-125, Carnegie Mellon University, 1994. (See also http://www.cs.cmu.edu/~quake-papers/ painless-conjugate-gradient.ps.).
D. Terzopoulos and K. Fleischer. Deformable models. Visual Computer, 4:306–331, 1988.
D.TerzopoulosandK.Fleischer.Modelinginelasticdeforma- tion: Viscoelasticity, plasticity, fracture. In Computer Graph- ics (Proc. SIGGRAPH), volume 22, pages 269–278. ACM, August 1988.
D. Terzopoulos, J.C. Platt, and A.H. Barr. Elastically de- formable models. Computer Graphics (Proc. SIGGRAPH), 21:205–214, 1987.
D. Terzopoulos and H. Qin. Dynamics nurbs with geomet- ric constraints for interactive sculpting. ACM Transactions on Graphics, 13:103–136, 1994.
X. Tu. Artificial Animals for Computer Animation: Biome- chanics, Locomotion, Perception and Behavior. PhD thesis, University of Toronto, May 1996.
P. Volino, M. Courchesne, and N. Magnenat Thalmann. Ver- satile and efficient techniques for simulating cloth and other deformable objects. Computer Graphics (Proc. SIGGRAPH), pages 137–144, 1995.
P. Volino, N. Magnenat Thalmann, S. Jianhua, and D. Thal- mann. An evolving system for simulating clothes on virtual actors. IEEE Computer Graphics and Applications, 16:42–51, 1996.
[5] B. Eberhardt, A. Weber, and W. Strasser. particle-system model for cloth draping. Graphics and Applications, 16:52–59, 1996.
A fast, flexible,
IEEE Computer
[6] G. Golub and C. Van Loan. Matrix Computations. John Hop- kins University Press, 1983.
[7] M. Kass. An Introduction To Physically Based Modeling, chapter Introduction to Continuum Dynamics for Computer
Graphics. SIGGRAPH Course Notes, ACM SIGGRAPH, 1995.
[8] M. Kass and G. Miller. Rapid, stable fluid dynamics for computer graphics. Computer Graphics (Proc. SIGGRAPH), pages 49–58, 1990.
[9] H.N. Ng and R.L. Grimsdale. Computer graphics techniques for modeling cloth. IEEE Computer Graphics and Applica- tions, 16:28–41, 1996.
[10] H. Okabe, H. Imaoka, T. Tomiha, and H. Niwaya. Three di- mensional apparel cad system. Computer Graphics (Proc. SIGGRAPH), pages 105–110, 1992.
SIGGRAPH ’98 10
CD-ROM version

Figure 1 (top row): Cloth draping on cylinder; frames 8, 13 and 35. Figure 2 (second row): Sheet with two fixed particles; frames 10, 29 and 67. Figure 3 (third row): Shirt on twisting figure; frames 1, 24 and 46. Figure 4 (bottom row): Walking man; frames 30, 45 and 58.
SIGGRAPH ’98 11 CD-ROM version

Figure 5 (top row): Dancer with short skirt; frames 110, 136 and 155. Figure 6 (middle row): Dancer with long skirt; frames 185, 215 and 236. Figure 7 (bottom row): Closeups from figures 4 and 6.
SIGGRAPH ’98 12 CD-ROM version

Leave a Reply

Your email address will not be published. Required fields are marked *