Srb 1

  • Uploaded by: Vincent S Ryan
  • 0
  • 0
  • October 2019
  • PDF

This document was uploaded by user and they confirmed that they have the permission to share it. If you are author or own the copyright of this book, please report to us by using this DMCA report form. Report DMCA


Overview

Download & View Srb 1 as PDF for free.

More details

  • Words: 6,775
  • Pages: 12
ASCI

VIRTUAL PROTOTYPING OF SOLID PROPELLANT ROCKETS Researchers seek a detailed, whole-system simulation of solid propellant rockets under normal and abnormal operating conditions. A virtual prototyping tool for solid propellant rocket motors based on first principles models of rocket components and their dynamic interactions meets this goal.

S

afety and reliability are paramount concerns in rocket motor design because of the enormous cost of typical payloads and, in the case of the Space Shuttle and other manned vehicles, for the crew’s safety. In the spring of 1999, for example, a series of three consecutive launch failures collectively cost more than US$3.5 billion. The most notorious launch failure, of course, was the tragic loss of the Space Shuttle Challenger and its seven crew members. Thus, there is ample motivation for improving our understanding of solid rocket motors (SRMs) and the materials and processes on which they are based, as well as the methodology for designing and manufacturing them. The use of detailed computational simulation in the virtual prototyping of products and devices has heavily influenced some industries— for example, in automobile and aircraft design— but to date, it hasn’t made significant inroads in rocket motor design. Reasons for this include the market’s relatively small size and the lack of sufficient computational capacity. Traditional design practices in the rocket industry primarily

1521-9615/00/$10.00 © 2000 IEEE

MICHAEL T. HEATH AND WILLIAM A. DICK Center for Simulation of Advanced Rockets, UIUC

21

use top-down, often one-dimensional modeling of components and systems based on gross thermomechanical and chemical properties, combined with engineering judgement based on many years of experience, rather than detailed, bottom-up modeling from first principles. Moreover, there has been a tendency to study individual components in isolation, with relatively little emphasis on the often intimate coupling between the various components. For example, SPP1—an industry-standard code for analyzing solid propulsion systems—includes a fairly detailed model of the propellant thermochemistry, but no structural analysis and no detailed model of internal flow. One of our primary goals at the Center for Simulation of Advanced Rockets (CSAR) is to develop a virtual prototyping tool for SRMs based on detailed modeling and simulation of their principal components and the dynamic interactions among them. Given a design specification—geometry, materials, and so on—we hope to be able to predict the entire system’s resulting collective behavior with sufficient fidelity to determine both nominal performance characteristics and potential weaknesses or failures. Such a “response tool” could explore the space of design parameters much more quickly, cheaply, and safely than traditional build-andtest methods. Of course, we must validate such a

COMPUTING IN SCIENCE & ENGINEERING

Figure 1. Idealized solid propellant rocket. Major components are indicated in red, our initial simulation models in green.

Solid propellant Igniter

Combustion-injection boundary layer model Compressible turbulence LES model Thermo-visco-elastic model

Interior cavity Combustion interface

Nozzle

Exhaust plume

Elasticity/ LES ablation model model Thermo-mechanical foam model Thermo-elastic model

capability through rigorous and extensive comparison with data for known situations to have confidence in its predictions for unknown situations. Although it is unlikely that simulation will ever totally replace empirical methods, it can potentially dramatically reduce the cost of those methods by identifying the most promising approaches before building actual hardware. Challenges in rocket simulation Solid propellant boosters are the “heavy lifters” of the space launch industry. Most of the world’s large, multistage launch vehicles—including the Ariane, Delta, Titan, and Space Shuttle—employ two or more SRBs in the initial stage to provide 80% or more of the immense thrust needed to lift a payload in excess of 10,000 pounds off the launch pad and propel it the first few tens of miles above Earth. Beyond this point, subsequent stages—typically liquidfueled—take over into orbit and beyond. SRMs are notably simpler than liquid rocket engines.2 The latter have far more moving parts (pumps, valves, and so on) and require storing and handling of liquids that might be cryogenic or potentially hazardous. SRMs, though, have almost no moving parts (often only a gimballed nozzle for thrust vector control), and the composite solid propellant (containing both fuel and oxidizer) forms the combustion chamber. The main disadvantage of SRMs is that once ignited, combustion is essentially uncontrollable: the propellant burns at maximum rate until exhausted. Thus, solid motors are ideal for the initial stages of flight, when raw power is more important than finesse, and then liquid-propellant rockets take over for the portions of flight requiring more delicate maneuvering. Despite their relative simplicity, SRMs are still fiendishly complex in terms of the chemical and thermomechanical processes that take place during fir-

22

Case Insulation

ing, as well as the design and manufacturing processes required to make them reliable, safe, and effective. Figure 1 shows a schematic drawing of a typical SRM—the major parts are indicated along with the types of mathematical models that might be used to represent them. Reality, of course, is considerably more complex than this 2D picture, and we note the following major challenges in achieving our goals. • The complex behavior of SRMs requires fully 3D modeling to capture the essential physics adequately. Examples include the combustion of composite energetic materials; the turbulent, reactive, multiphase fluid flows in the core and nozzle; the global structural response of the propellant, case, liner, and nozzle; and potential accident scenarios such as pressurized crack propagation, slag ejection, and propellant detonation. • The coupling between components is strong and nonlinear. For example, the loading due to fluid pressure deforms the solid propellant, which changes the geometry of the fluid flow, which in turn affects pressure, and so on. Similarly, the burn rate increases with pressure and vice versa. • The geometry is complex and changes dynamically as the rocket consumes propellant. The inclusion of slots and fins, which forms a star-shaped cross-section, enhances the amount of burning surface area. Whatever its initial shape, the propellant surface regresses at a pressure-dependent rate as the propellant burns, and discrete representations of the solid and fluid components, as well as the interface between them, must adapt accordingly. • The spatial and temporal scales are extremely diverse. For example, processes

COMPUTING IN SCIENCE & ENGINEERING







In September 1997, CSAR embarked on an ambitious plan to tackle these daunting challenges and produce a virtual prototyping tool for SRMs.3 This article is a progress report almost two years into our five-year plan. Our initial plans seemed audacious, but the substantial resources our sponsor (the US Department of Energy’s Accelerated Strategic

MARCH/APRIL 2000

Accidents Geometrical complexity



such as combustion and crack propagation occur on micron length scales and microsecond time scales, or less, which are entirely infeasible to treat a two-minute burn of a 125-foot-long rocket. Manufacturing and transportation constraints necessitate the use of numerous joints, including field joints where motor segments are assembled at the launch site. This significantly complicates the geometry and structural response of the motor and introduces potential points of failure. Modeling and simulating each component is challenging both methodologically and computationally. Although there is considerable experience in the field in modeling the various rocket motor components, a more fundamental understanding of the constitutive and energetic properties of materials and of the processes they undergo requires much greater detail along with terascale computational capacity. Modeling and simulating component coupling is even more demanding because it requires not only still greater computational capacity, but it also demands that the corresponding software modules interact in a manner that is physically, mathematically, and numerically correct and consistent. When data are transferred between components, they must honor physical conservation laws, mutually satisfy mathematical boundary conditions, and preserve numerical accuracy, even though the corresponding meshes might differ in structure, resolution, and discretization methodology. Integrated, whole-system SRM simulation requires enormous computational capacity, currently available only through massively parallel systems that have thousands of processors. Thus, the software integration framework, mesh generation, numerical algorithms, input/output, and visualization tools necessary to support such simulations must be scalable to thousands of processors.

GEN2 family

Joints Star grain 3D

GEN1 family

2D 1D

GEN0 Weakly Fully coupled coupled Physical complexity

Figure 2. Our code development follows this staged approach with increasing complexity in component models and coupling.

Detailed

Computing Initiative program) provided let us assemble a team of over 100 researchers, including roughly 40 faculty, 40 graduate students, and 20 staff (research scientists, programmers, and postdoctoral associates), that represented 10 departments across our university. This diverse group provided the broad expertise needed in combustion, fluid dynamics, structural mechanics, and computer science, but it also presented the additional challenge of coordinating a large collaborative project that cuts across traditional departmental boundaries. We organized our effort along basic disciplinary lines without regard to the academic departments of the participants. This has had the salutary effect of inducing collaboration among faculty and students in a given discipline, such as fluid dynamics, regardless of which department they might occupy. Cross-cutting teams—such as System Integration and Validation and Specification—draw members from all four disciplinary groups and require an additional level of collaboration. Staged approach We realized from the outset that a project of this complexity would require a staged approach: we would need to learn to walk before we could run (much less fly). The primary axes of complexity in our problem are physical and geometric (see Figure 2). Physical complexity refers to the detail and sophistication of physical models employed and the degree of coupling among them. Geometric complexity refers to the dimension of the problem and the degree of detail and fidelity in representing a real SRM. In essence, we wish to move along the diagonal of this diagram over time. In this spirit, we defined three successive generations of integrated rocket simulation codes:

23

Space Shuttle Reusable Solid Rocket Motor We chose the Space Shuttle Reusable Solid Rocket Motor as our primary simulation target for a variety of reasons, including its national importance, its public visibility, its fairly typical design, and the availability of detailed specifications and extensive test data. We outline here the basic technical facts about the Space Shuttle RSRM.1,2 Figure A shows a composite drawing of the RSRM. Height: 126.11 ft Diameter: 12.16 ft Weight: 149,275 lb empty; 1,255,415 lb full Case: High-strength D6AC steel alloy, 0.479 in. to 0.506 in. thick Nozzle: Aluminum nose-inlet housing and steel exit cone, with carbon-cloth phenolic ablative liners and glass-cloth phenolic insulators. Nozzle is partially submerged and is movable for thrust vector control. Insulation: Asbestos-silica-filled nitrile butadiene rubber Propellant (material percent by weight): Ammonium perchlorate oxidizer: 70 Powdered aluminum fuel: 16 Polybutadiene polymer (PBAN) binder: 12 Epoxy curative agent: 2 Ferric oxide burn rate catalyst: trace Propellant grain: 11-point star-shaped perforation in head end of forward segment, aft-tapered cylindrical perforation in remaining segments. Liquid and solid ingredients are first thoroughly mixed into a thick paste, then curative agent is added before mixture is vacuum-cast into a mold and then cured in a “slow” oven for several days. Consistency of

resulting composite solid is similar to that of a pencil eraser. Igniter: Solid rocket pyrogen igniter mounted in forward end, 47.5-in. long and containing 134 lb of TP-H1178 propellant Total launch weight: 4.5 million lb (including two SRBs, external tank, orbiter, and payload) Maximum thrust: 3,320,000 lb. force (each SRB) Acceleration: Lift-off 1.6 g (maximum 3.0 g) Launch timeline: Liquid engines fire: –6.0 sec SRB igniter initiated: 0.0 sec Lift-off pressure: 564 psia reached at 0.23 sec All exposed propellant ignited: 0.3 sec Maximum operating pressure: 914 psia reached at 0.6 sec Roll program begins: 10 sec Star grain burnout: 21 sec Liquid engines throttled down: 30 sec Mach 1 reached: 40 sec Solid propellant burnout: 111 sec SRB separation: 126 sec Velocity at separation: 3,100 mph Altitude at separation: 25 nmi References 1.

Design Data Book for Space Shuttle Reusable Solid Rocket Motor, Thiokol Space Operations, Publication No. 930480, TRW-16881, Revision A, Brigham City, Utah, 1997.

2.

A.J. McDonald, “Return to Flight with the Redesigned Solid Rocket Motor,” Proc. AIAA/ASME/SAE/ASEE 25th Joint Propulsion Conf., AIAA Paper No. 892404, AIAA Press, Reston, Va., 1989, pp. 1–15.

• GEN0: 2D ideal rocket with steady-state burning at chamber pressure, power law for propellant regression, Euler equations for fluid flow, a rigid case, linearly elastic propellant, and one-way coupling from fluid to solid. We based its physical parameters on the Space Shuttle reusable solid rocket motor (see sidebar). GEN0 was intended primarily as a warm-up exercise. • GEN1: Fully 3D whole-system simulation code using relatively simple component models, two-way coupling, and reasonably realistic geometry approximating that of the Space Shuttle RSRM. Star grain of Shuttle RSRM is included, but not joints, inhibitors, or cracks. Solid components include viscoelastic propellant and linearly elastic case. Fluid component is an unsteady, viscous, compressible flow, with a large-eddy simulation turbulence model but with no particles, radiation, or chemical reactions in the

24

flow. Combustion model assumes homogeneous surface burning and pressure-dependent regression rate. There is full, two-way aeroelastic coupling between fluid and solid components. Development of GEN1 was expected to span the first three years of the five-year project. • GEN2: Fully capable rocket simulation tool with detailed component models, complex component interactions, and support for subscale simulations of accident scenarios such as pressurized crack propagation, slag accumulation and ejection, and potential propellant detonation. GEN2 includes more detailed geometric features, such as joints and inhibitors, and also includes more detailed and accurate models for materials and processes based on separate subscale simulations. GEN2 was expected to span the last three years of the five-year project, overlapping with the final year of GEN1.

COMPUTING IN SCIENCE & ENGINEERING

Figure A. The Reusable Solid Rocket Motor (RSRM) is a primary booster for NASA’s Space Transportation System (STS). Section A-A shows 11-point slot and fin star grain structure in the RSRM’s forward segment. Propellant in forward-center and aft-center segments form straightwalled cylinders; aft-segment propellant tapers outward to submerged nozzle. Inhibitors between segments are asbestos-filled carboxylterminated polybutadiene used to tailor burning surface to meet the motor’s thrust requirements.

Progress to date We assembled the integrated GEN0 code from existing in-house modules for fluids, solids, and combustion components, and we completed it in May 1998. We ran it with modest levels of parallelism on a shared-memory SGI Power Challenge. Computed results agreed reasonably well with predictions of classical 1D theory, but we didn’t extensively validate it because we never intended GEN0 as a realistic, high-fidelity simulation; we saw it simply as a start-up system integration exercise to get our team accustomed to working together. Visit www.csar.uiuc.edu to view animations of our results. We based the subsequent GEN1 code on newly written or substantially modified in-house codes for the various modules. We completed a simplified, serial, but fully 3D version of it in October 1998. Its principal simplifications included our use of a strictly cylindrical geometry (no star grain, which is a slot-and-fin geometry used to

MARCH/APRIL 2000

increase the propellant’s initial burning surface area), no support for interface regression due to burning (not a significant factor for short burn times, but obviously necessary for longer runs), the requirement that the solid and fluid meshes must match at the interface to simplify data transfer, and our use of a linearly elastic (rather than viscoelastic) model for the propellant. Computed results without the star grain in the propellant gave a head-end pressure at the onset of steady burning of roughly half the empirically measured value for the Space Shuttle RSRM. This was not surprising, as the whole point of the star grain is to increase the exposed surface area of propellant early in the burn, which increases pressure and thrust accordingly. Thus, implementing the star grain became high priority. Another high priority was a fully parallel implementation, not only because this was an important general goal, but also for the practical reason that we could not otherwise make

25

Figure 3. Solids and fluids codes have different approaches to multicomponent simulation.

Rocsolid • Finite element • Linear elastodynamics • Unstructured hexahedral meshes • ALE treatment of interface regression • Implicit time integration • Multigrade equation solver • F90, MPI parallelism

Scaled speedup

1,000

100

10

T3E O2K SP2

1 1

(a)

10

100

1,000

Processors

Scaled speedup

1,000

100

T3E O2K CLU

10

1 1

(b)

10

100

1,000

Processors

Scaled speedup

1,000

100

10

T3E O2K

1

(c)

1

10

100

1,000

Processors

Figure 4. Graphs show scaled speedup of separate (a) Rocsolid and (b) Rocflo modules and of (c) integrated GEN1 code for Space Shuttle RSRM. Problem size is fixed at 14,625 fluid grid points and 8,192 solid elements per processor as number of processors grows. Computers used include Cray T3E, SGI Origin 2000 (02K), IBM SP2, and Intel Pentium cluster (CLU).

26

Rocflo • Finite volume • Unsteady, viscous, compressible flow • Block-structured meshes • ALE moving boundaries • Explicit time integration • 2nd order upwind total variation • F90, MPI parallelism

runs of significant duration at a reasonable resolution. By May 1999, we had completed a fully parallel implementation of GEN1.4 With the star grain geometry, computed head-end pressure was now within a few percentage points of the Space Shuttle RSRM’s measured value. Recent efforts have focused on implementing interface regression in both fluid and solid modules and on allowing for nonmatching meshes at the interface. Both fluid and solid modules now support interface regression using an ALE (Arbitrary Lagrangian-Eulerian) approach in which the mesh moves to allow for the dynamically changing geometry. We also devised generalmesh-association and conservative data-interpolation schemes that permit accurate data transfer between components with nonmatching meshes at the interface. The main features of the solids codes (Rocsolid) include finite elements, unstructured hexahedral meshes, linear elastodynamics, ALE treatment of regression, implicit time integrations, multigrid equation solvers, and Fortran 90 with MPI parallelism. The main features of the fluids codes (Rocflo) include finite volume; block-structured meshes; unsteady, viscous, compressible flow; ALE treatment of moving boundaries; second-order upwind total variation diminishing schemes; explicit time integrations; and Fortran 90 with MPI parallelism. Figure 3 shows how the two compare. Parallel scalability of Rocsolid, Rocflo, and the GEN1 code that integrates them has been excellent. We’ve run Rocflo on up to 2,048 processors, and we’ve run Rocsolid and GEN1 on up to 512 processors on a variety of platforms, including the SGI Origin at NCSA, the Cray T3E at Pittsburgh Supercomputer Center, and all three ASCI platforms at the US Department of Energy laboratories. Figure 4 shows “scaled speedup,” meaning that the problem size per processor is constant. The largest mesh sizes have about four million elements for the solid and seven million zones for the fluid. Visualizations of some computational results are shown in Figures 5 and 6. The features we still have to implement to

COMPUTING IN SCIENCE & ENGINEERING

Figure 5. Fully coupled 3D simulation of star grain in Space Shuttle RSRM showing stress in propellant and gas pressure isosurfaces in slots and core region. Executed on 256-processor SGI Origin 2000, visualized with Rocketeer, our in-house visualization tool.

complete the full GEN1 code include viscoelasticity, large deformations, and thermal effects in the solid; large-eddy simulation turbulence model in the fluid; more detailed model of combustion and interface regression; and a flamespreading model to capture ignition transients. System integration issues A number of technical issues arise in building an integrated, multicomponent code such as GEN1. First is the overall integration strategy, where the fundamental choice is between modular and monolithic approaches. In building the GEN1 code, we chose a modular or partioned approach in that we used separately developed component modules and created an interface to tie them together. In such an approach, separately computed component solutions might require subiterations back and forth between components to attain self-consistency. This approach contrasts with a more monolithic strategy in which all the physical components are incorporated into a single system of equations and all the relevant variables are updated at the same time, thereby obviating the need to iterate to self-consistency. Although it has some theoretical advantages, a monolithic approach impedes separate development and maintenance of individual

MARCH/APRIL 2000

Figure 6. Gas temperature computed by Rocflo in star grain region of Space Shuttle RSRM near onset of steady burning, visualized by Rocketeer. Values range from 3,364 K (magenta) to 3,392 K (red). Temperature is represented as (a) tint on interior surface of propellant and as (b) a series of translucent colored isosurfaces in interior at slightly later time. Rocket is cut in half along lateral axis to improve visibility.

component modules by specialists in the respective areas. The modular approach not only expedites separate development and maintenance, it also allows swapping of individual modules without replacing the entire code or even affecting the other modules. The modular strategy seemed to offer clear practical advantages in our somewhat dispersed organizational setting, as well as potentially letting users include commercial modules when appropriate. However, even less coupled approaches are commonly used in practice, in which entirely in-

27

dependent codes interact only offline (often with human intervention), perhaps through exchange of input and output data files. By contrast, in our modular GEN1 code, the component modules are compiled into a single executable code and they exchange data throughout a run with subroutine calls and interprocessor communication. Another important issue is the physical, mathematical, and geometric description of the interface between components, which in our case includes the jump conditions combustion induces. A careful formulation of the interface boundary conditions is necessary to satisfy the relevant conservation laws for mass and linear momentum, as well as the laws of thermodynamics. Time-stepping procedures are another significant issue in component integration. Here, the time steps for the fluid are significantly smaller than those for the solid. Thus, we employ a predictor-corrector approach in which the fluid is explicitly stepped forward by several (say, 10) time steps, based on the current geometry the solid determines. The resulting estimate of fluid pressure at the future time is then available for taking an implicit time step for the solid. However, the resulting deformation and velocity of the solid change the fluid’s geometry, so the time-stepping of the fluid repeats and so on, until we attain convergence, which usually requires only a few subiterations. Unless iterated until convergence, this scheme is only first-order accurate, and it is “serial” in that only one component computes at a time. Parallel time-stepping schemes that are second-order accurate without subiterations are possible,6 and we plan to investigate these. But because we map both fluid and solid components onto each processor, our current scheme does not prevent us from utilizing all processors. As we mentioned briefly earlier, data transfer between disparate meshes of different components is another highly nontrivial issue in component integration. In our approach, we let the meshes differ at the interface in structure, resolution, and discretization methodology, and indeed this is the case in GEN1, because the fluid mesh is block-structured, relatively fine, and based on cell-centered finite volumes, whereas the solid mesh is unstructured, relatively coarse, and based on node-centered finite elements. Although in principle the two interface meshes should abut because they discretize the same surface, in practice we can’t assume this because of discretization or rounding errors. Thus, we have developed general mesh association algorithms that efficiently determine which fluid points are

28

associated with each element (facet) of the solid interface mesh;7 we then use the local coordinates of the associated element to interpolate relevant field values in a physically conservative manner. Yet another thorny issue in component integration is partitioning the component meshes for parallel implementation in distributed memory. The block-structured fluid mesh is relatively easy to partition in a highly regular manner, but the unstructured solid mesh is partitioned by a heuristic approach, currently using Metis, which often yields irregular partitions (visit wwwusers.cs.umn.edu/~karypis/metis for further information). Moreover, because we partition the two component meshes separately, there is no way to maintain locality at the interface—adjacent partitions across the interface may not be placed on the same or nearby processors. In our current approach, this effect complicates the communication pattern and might increase communication overhead, but it has not been a serious drag on parallel efficiency so far. Nevertheless, we plan to explore more global, coordinated partitioning strategies that will preserve locality and perhaps simplify communication patterns. Software integration framework Our overarching goal in CSAR is not only to develop a virtual prototyping tool for SRMs, but also to develop a general software framework and infrastructure to make such integrated, multicomponent simulations much easier. Toward this end, we have initiated research and development efforts in several relevant areas of computer science, including parallel programming environments, performance monitoring and evaluation, parallel input/output, linear solvers, mesh generation and adaptation, and visualization. Our work in parallel programming environments has focused on creating an adaptive software framework for component integration based on decomposition and encapsulation through objects. This environment provides automatic adaptive load balancing in response to dynamic change or refinement, as well as highlevel control of parallel components. It is purposely designed to maintain compatibility with component modules in conventional languages such as Fortran 90, and it provides an automated migration path for the existing parallel MPI code base. This work is in part an extension of the previously developed Charm++ system, which has successfully built a large parallel code for molecular dynamics, NAMD.8 Although this frame-

COMPUTING IN SCIENCE & ENGINEERING

Rocketeer by Robert A. Fiedler and John Norris We need a powerful scientific visualization tool to analyze the large, complex 3D data sets our whole system and subscale rocket simulations generate. After an extensive review of existing packages, we decided to develop our own tool, which we call Rocketeer. (Visit www.csar.uiuc.edu/F_ software/rocketeer to download a user guide and software.) The tool has a number of features that make it ideal for visualizing data from multicomponent simulations, including its support for both structured and unstructured grids, cell-centered and node-centered data, ghost cells, seamless merging of multiple data files, automated animation, and a smart reader for HDF (hierarchical data format). A particularly useful feature for visualizing field data in the interior of an SRM is Rocketeer’s ability to depict translucent isosurfaces, which lets us view isosurfaces of temperature or pressure, for example, without blocking the view of other isosurfaces deeper inside (see Figure B). Although our need to visualize data from SRM simulations specifically motivated many of these features, Rocketeer is a broadly useful, general-purpose visualization tool. Rocketeer is based on the Visualization Toolkit (VTK),1 which is in turn based on OpenGL to take advantage of graphics hardware acceleration. It currently runs on Microsoft Windows and Unix/Linux. Planned enhancements include implementing a client-server model so that we can perform most compute-intensive operations (in parallel) on a remote supercomputer while interactive control and rendering are performed locally on a graphics workstation.

work is not yet mature enough to serve as the basis for the current GEN1 code, results for pilot implementations of the GEN1 component modules using the framework show great promise. In performance evaluation, we are leveraging ongoing development of the Pablo performance environment,9 which provides capabilities for dynamic, multilevel monitoring and measurement, real-time adaptive resource control, and intelligent performance visualization and steering of distributed, possibly geographically dispersed, applications. We used these tools to instrument the GEN1 code, view the resulting performance data, and relate it back to the application code’s call graph to identify performance bottlenecks. We also updated the popular ParaGraph performance visualization tool10 to display the behavior and performance of parallel programs using MPI, and we used it to analyze the GEN1 component modules. Parallel I/O is often a bottleneck in large-scale simulations, both in terms of performance and

MARCH/APRIL 2000

Figure B. Rocketeer visualization of temperature profile computed by Rocflo in the star grain region of RSRM at the onset of steady burning. Values range from 3,364 K (magenta) to 3,392 K (red). Temperature is represented by tint on the propellant fin’s surface and by a series of translucent colored isosurfaces inside the slot. The image is cut in half along the rocket’s axis to enhance visibility. The hottest point in this imaged flow is in the stagnation region in the middle of the rocket’s head end; the coolest point is in the middle of the star grain region’s open end.

Reference 1.

W.Schroeder, K. Martin, and W.E. Lorensen, The Visualization Toolkit: An Object-Oriented Approach to 3D Graphics, Prentice Hall, New York, 1997.

of programming effort required to manage I/O explicitly. For large-scale rocket simulations, we need good performance for collective I/O across many processors for purposes such as periodically taking snapshots of data arrays for visualization or checkpoint/restart. Because we use geographically dispersed machines, we also need efficient and painless data migration between platforms and back to our home site. Panda11 provides all these services—it runs on top of MPI and uses the standard file systems provided with the platforms we use. Its library offers server-driven collective I/O, support for various types of data distributions, self-tuning of performance, and integrated support for data migration that exploits internal parallelism. We’ve already incorporated Panda into Rocflo, and are doing the same in Rocsolid. Again, early results are promising, and we plan to use Panda to handle parallel I/O within our main visualization tool, Rocketeer (see sidebar).

29

Figure 7. The propellant surface model employs two sizes of AP particles imbedded in a fuel/binder matrix. Mixture models for this matrix account for smaller AP particles. (a) 3D flames supported by AP and binder decomposition product gases for configuration appear at midline region. (b) Primary flame results from burning binder/AP mixture.

Validation Comparison of simulation results with known test cases and laboratory data is essential to establish this approach’s validity for science and engineering. With our GEN1 integrated code rapidly maturing, we have begun an aggressive series of computational experiments to verify and validate its efficacy and fidelity. Fluid-solid interaction problems we use for validation include flow over an elastic panel, flow over a wing (Agard Wing 445.6), and a model of inhibitor deformation in an SRM. We hope also to be able to make comparisons with test data for small laboratory-scale rockets through collaboration with various government laboratories. A larger-scale test we are pursuing is to try to predict the propellant “slumping” that led to failure for the Titan IV SRB’s original design. Our ultimate test will be comparison with the immense amount of test data for the Space Shuttle RSRM taken during static firing tests after its redesign. These data include literally thousands of readings from strain gauges, pressure curves, and so on for a liberally instrumented test version of the Shuttle RSRM. New research directions Our second-generation rocket simulation code, GEN2, will require significantly more detailed and sophisticated component models than those in GEN1. Moreover, GEN2 will also support accident scenarios that will require even greater detail and finer resolution. To have these new models ready by the time we need them in GEN2, we have already begun extensive re-

30

search into these modeling issues, some of which we outline here. Heterogeneous propellant flames

The propellant in the Shuttle SRM consists of a high-density packing of ammonium perchlorate (AP) oxidizer particles embedded in a matrix of powdered aluminum fuel and polymeric binder. The aluminum particles burn in the gasphase products of AP-binder combustion. The propellant uses an initial bimodal distribution of AP particle sizes of 20 and 200 microns, and the primary combustion field is located within a few tens of microns of the propellant surface. Flow disturbances originating in the chamber due to acoustic-wave or mean-flow interactions and turbulence can affect this field, leading to what is known as erosive burning. Some of the heat generated in the combustion field is conducted to the propellant surface. This heat is responsible for the surface regression and the conversion of solid propellant to combustible gases. The resulting burning surface is not flat because the instantaneous regression rates of AP and binder differ, and if cracks form in the propellant, the increase in propellant surface can lead to a sharp increase in combustion intensity. To describe the surface regression and to generate boundary conditions for chamber flow, we must resolve the 3D combustion field and couple it with physical processes such as heat conduction in a thin surface layer in the propellant and allow for pressure and thermal feedback from the chamber flow (see Figure 7).

COMPUTING IN SCIENCE & ENGINEERING

Crack propagation in solid propellant

The initiation and propagation of one or more cracks in the solid propellant or along the graincase interface can dramatically affect rocket performance. Cracks create additional burning surfaces in the SP, so the propagation of cracks can greatly affect the pressure history in the rocket chamber, leading in some cases to the rocket’s destruction. Modeling crack propagation in burning SP is quite complex, because the problem is characterized by a tight coupling between structural dynamics, combustion, and fluid mechanics, along with a rapidly evolving geometry. Using a fully coupled explicit aeroelastic finiteelement, finite-volume code, we are investigating potential accident scenarios associated with the presence of pre-existing cracks at various locations in the solid booster, with special emphasis on propellant-liner interfacial failures. We’re using a novel cohesive-volumetric finite-element scheme to capture the spontaneous motion of the crack, allowing for the possibility of crack arrest or branching. As the crack propagates and the reacting gas pressurizes newly created crack faces, the fluid domain undergoes complex changes that an adaptive unstructured finite-volume scheme can capture. As illustrated in Figure 8, the reactive gas flow emanating from a pre-existing radial crack in the SP interacts with the core flow in the rocket motor. This generates a region of high pressure on the leeward face of the crack and a region of low pressure in the downstream vicinity of the crack that leads to substantial deformation of the propellant grain. Aluminum particle combustion

The solid propellants in modern rockets use aluminum (Al) particles as fuel. As combustion proceeds, these particles in the propellant melt and agglomerate on the combustion interface. A complex process follows as Al droplets detach from the surface and are injected into the core flow. The droplets, whose initial size varies from 20 to 300 microns, are injected into a strong cross-flow, and they provide a significant source of heat as they burn to form Al oxides. Near-wall turbulence plays an important role in the dispersion of Al droplets, and as a result heat release is volumetrically distributed, although dominant mainly in the near-wall region. Al2O3 is the primary product of combustion, and it appears either as fine powder of micron size or as larger residual particles. The deposition of Al2O3 in the form of slag in the submerged nozzle can adversely affect motor performance.

MARCH/APRIL 2000

Figure 8. Effect of pre-existing radial crack on motor’s core flow: Colored contours denote pressure in core flow and hydrostatic pressure in solid propellant. Note region of high pressure (red) in deformed crack and of low pressure (blue) downstream from crack.

The combustion of Al droplets strongly couples the dispersed phase (droplets and Al2O3 particles) with the continuous phase (the surrounding core flow). We expect the GEN2 version of Rocflo to include a Eulerian implementation of the core flow and a Lagrangian implementation of the Al droplets and the larger oxide particles. The simulation will introduce tens of millions of Al droplets into the flow at the combustion interface according to a specified probability distribution and local mass injection rate. The position, velocity, temperature, and species concentration of each droplet will be tracked in the simulation over time by solving a set of ordinary differential equations. The effect of the surrounding flow will be parameterized in terms of lift and drag coefficients, heat and mass transfer coefficients, and droplet burn rate. So far we have developed a detailed time-dependent 3D subsystem simulation of the flow, evaporation, and burning of an isolated Al droplet to obtain accurate state-of-the-art parameterizations (see Figure 9).

T

he individual problems just discussed are themselves examples of fluid− solid interactions, albeit on finer scales, for which we will be able to use the same software integration framework as for the global simulation of the SRM. In this way, we hope to leverage much of our current software development effort in component integration, and to be able to spin off these subscale simulations from the larger-scale system simulation in a highly compatible manner.

31

Advanced Solid-Rocket Flow Simulation Program ROCFLO,” Proc. 38th AIAA Aerospace Sciences Meeting and Exhibit, AIAA Press, Reston, Va., 2000. 6. C. Farhat and M. Lesoinne, “Two Efficient Staggered Procedures for the Serial and Parallel Solution of Three-Dimensional Nonlinear Transient Aeroelastic Problems,” Computer Methods in Applied Mechanics and Eng., Vol. 182, Nos. 3 and 4, 2000. 7. X. Jiao, H. Edelsbrunner, and M.T. Heath, “Mesh Association: Formulation and Algorithms,” Proc. Eighth Int’l Meshing Roundtable, Tech. Report 99-2288, Sandia Nat’l Labs., Albuquerque, New Mexico, 1999, pp. 75–82. 8. L. Kale et al., “NAMD2: Greater Scalability for Parallel Molecular Dynamics,” J. Computational Physics, Vol. 151, No. 1, May 1999, pp. 283–312. 9. L. DeRose et al., “An Approach to Immersive Performance Visualization of Parallel and Wide-Area Distributed Applications,” Proc. Eighth IEEE Symp. High-Performance Distributed Computing, IEEE Computer Soc. Press, Los Alamitos, Calif., 1999, pp. 247–254. 10. M.T. Heath and J.A. Etheridge, “Visualizing the Performance of Parallel Programs,” IEEE Software, Vol. 8, No. 5, Sept. 1991, pp. 29–39.

Figure 9. Results obtained from a time-dependent 3D simulation of flow and heat transfer from a spherical droplet in cross-flow. The simulation employs a resolution of 81 × 96 × 32 points along the radial, circumferential, and azimuthal directions at Reynolds number Re = U∞D/ν = 350, based on free stream velocity (U∞) and droplet diameter (D). At this Reynolds number, flow is unsteady with timeperiodic vortex shedding. (a) Azimuthal velocity contours at an instant in time, where we can see an imprint of vortex shedding. (b) Temperature contours, where approaching flow is hotter (red) than the droplet (blue), which is considered isothermal.

Acknowledgments We thank our many colleagues at CSAR for their research contributions to this article. This program is truly a collaborative effort based on the technical strengths of many people. We thank Amit Acharya, Prosenjit Bagchi, S. Balachandar, Dinshaw Balsara, John Buckmaster, Philippe Geubelle, Changyu Hwang, Thomas L. Jackson, and Biing-Horng Liou for their contributions to the “New research directions” section. The CSAR research program is supported by the US Department of Energy through the University of California under subcontract B341494.

References 1. G.R. Nickerson et al., The Solid Propellant Rocket Motor Performance Prediction Computer Program (SPP), Version 6.0, Tech. Report AFAL-TR-87-078, US Air Force Materials Lab., Edwards Air Force Base, Calif., 1987. 2. G.P. Sutton, Rocket Propulsion Elements, 6th ed., John Wiley & Sons, New York, 1992. 3. M.T. Heath and W.A. Dick, “Virtual Rocketry: Rocket Science Meets Computer Science,” IEEE Computational Science & Eng., Vol. 5, No. 1, Jan.–Mar. 1998, pp. 16–26. 4. I.D. Parsons et al., “Coupled Multi-Physics Simulations of Solid Rocket Motors,” Parallel and Distributed Processing Techniques and Applications Conf., Vol. VI, CSREA Press, 1999. 5. P.V.S. Alavilli, D. Tafti, and F. Najjar, “The Development of an

32

11. Y. Cho et al., “Parallel I/O for Scientific Applications on Heterogeneous Clusters: A Resource-Utilization Approach,” Proc. 13th ACM Int’l Conf. Supercomputing, ACM Press, New York, 1999, pp. 253–259.

Michael T. Heath is the director of the Center for Simulation of Advanced Rockets at the University of Illinois, Urbana-Champaign. He is also a professor in the Department of Computer Science, the director of the Computational Science and Engineering Program, and a senior research scientist at the National Center for Supercomputing Applications at the university. His research interests are in numerical analysis—particularly numerical linear algebra and optimization—and in parallel computing. He wrote Scientific Computing: An Introductory Survey (McGraw-Hill, 1997), and has served as editor of several journals in scientific and high-performance computing. He received a BA in mathematics from the University of Kentucky, an MS in mathematics from the University of Tennessee, and a PhD in computer science from Stanford University. Contact him at CSAR, 2262 Digital Computer Lab., 1304 West Springfield Ave., Urbana, IL 61801; [email protected]; www.csar.uiuc.edu. William A. Dick is the managing director of the Center for Simulation of Advanced Rockets at the University of Illinois, Urbana-Champaign. He received a BS in mechanical engineering from the University of Delaware and an MBA from the University of Illinois. Prior to coming to the University of Illinois, he was a deputy director and composites engineer in the National Science Foundation’s Engineering Research Center for Composites Manufacturing Science and Engineering. Contact him at CSAR, 2266 Digital Computer Lab., 1304 West Springfield Ave., Urbana, IL 61801; [email protected]; www.csar.uiuc.edu.

COMPUTING IN SCIENCE & ENGINEERING

Related Documents

Srb 1
October 2019 22
Srb
October 2019 26
Srb Test
November 2019 11
Nasa 146685main Srb-et
October 2019 17
19.esp-srb
May 2020 9

More Documents from ""

The Cataclysm Of Supernova
October 2019 54
Supersonic Flight
October 2019 63
Silent Aircraft
October 2019 59
A Cosmological Model
October 2019 71
Relativity
October 2019 68