06.08.2013 Views

structural geology, propagation mechanics and - Stanford School of ...

structural geology, propagation mechanics and - Stanford School of ...

structural geology, propagation mechanics and - Stanford School of ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

STRUCTURAL GEOLOGY, PROPAGATION MECHANICS AND<br />

HYDRAULIC EFFECTS OF COMPACTION BANDS IN SANDSTONE<br />

A DISSERTATION<br />

SUBMITTED TO THE DEPARTMENT OF<br />

GEOLOGICIAL AND ENVIRONMENTAL SCIENCES<br />

AND THE COMMITTEE ON GRADUATE STUDIES<br />

OF STANFORD UNIVERSITY<br />

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS<br />

FOR THE DEGREE OF<br />

DOCTOR OF PHILOSOPHY<br />

Kurt Richard Sternl<strong>of</strong><br />

March 2006


© Copyright by Kurt Richard Sternl<strong>of</strong> 2006<br />

All Rights Reserved<br />

ii


Abstract<br />

Low-porosity, low-permeability compaction b<strong>and</strong>s (CBs) form a pervasive planar<br />

fabric throughout the upper 600 meters <strong>of</strong> the 1,400-meter-thick æolian Aztec s<strong>and</strong>stone<br />

<strong>of</strong> southeastern Nevada, an exhumed analog for aquifers <strong>and</strong> reservoirs <strong>of</strong> similar<br />

lithology. The existence <strong>of</strong> CBs, only recently recognized, raises issues <strong>of</strong> practical<br />

significance. How do they impact local permeability? Can they influence fluid flow <strong>and</strong><br />

transport at production scales? When, how <strong>and</strong> why do they form? Can their presence,<br />

geometry <strong>and</strong> effects be forecast in the subsurface from sparse data?<br />

To address these questions, we applied a range <strong>of</strong> techniques, including: outcrop <strong>and</strong><br />

thin-section observations <strong>and</strong> measurements; <strong>structural</strong> <strong>and</strong> tectonic analysis; scanning<br />

electron microscopy; detailed field mapping <strong>and</strong> aerial photograph interpretation;<br />

mechanical analysis <strong>and</strong> modeling; computational permeability estimation <strong>and</strong> fluid-flow<br />

simulation. Although derived entirely from the Aztec s<strong>and</strong>stone, our results <strong>of</strong>fer a more<br />

general phenomenological underst<strong>and</strong>ing <strong>of</strong> CBs <strong>and</strong> their effects.<br />

Up to centimeters thick <strong>and</strong> hundreds <strong>of</strong> meters long, CBs comprise thin discs <strong>of</strong><br />

uniaxial porosity-loss compaction that form anastomosing arrays generally symmetric to<br />

the maximum compression. Idealized as anticracks, CB <strong>propagation</strong> <strong>and</strong> pattern<br />

development can be modeled using a linear-elastic boundary element method.<br />

Mechanical interaction between CBs is inversely proportional to differential stress, such<br />

that connectivity within a fabric, <strong>and</strong> its directional impact on fluid flow, can be<br />

estimated from the stress state in which it formed, <strong>and</strong> vice versa. Computational<br />

estimation corroborates measured permeability reductions <strong>of</strong> 10 -3 inside CBs. At outcrop<br />

scales, this yields bulk permeability reductions up to 10 -2 <strong>and</strong> anisotropy up to 10 1 . Flow<br />

simulations performed on a detailed, 40-acre CB map reveal significant practical impacts:<br />

pumping <strong>and</strong> injection pressures increase three-fold; reservoir production efficiency<br />

depends on proper alignment <strong>of</strong> the well array to the CB fabric, <strong>and</strong> contaminant<br />

transport channels along the CBs regardless <strong>of</strong> the regional pressure gradient.<br />

We interpret CBs in the Aztec as resulting from the tectonic compression <strong>of</strong> a high-<br />

porosity, low-cohesion s<strong>and</strong>stone at modest confining pressures (


Acknowledgements<br />

Dissertations can be ephemeral beasts—mythic, elusive <strong>and</strong> notoriously difficult to<br />

capture on paper. Coaxing a good dissertation into publication has to be one <strong>of</strong> the more<br />

arduous, solitary journeys a person can make while so entirely swaddled in support. This<br />

volume represents the culmination <strong>of</strong> one such saga, <strong>and</strong> there are many people to thank<br />

for seeing me through the <strong>of</strong>ten dense fog to that distant, shining shore <strong>of</strong> “doneness.”<br />

First, I acknowledge the U.S. Department <strong>of</strong> Energy, Office <strong>of</strong> Basic Energy Sciences<br />

for its generous funding <strong>of</strong> my research <strong>and</strong> tenure at <strong>Stanford</strong>. Additional support was<br />

provided by the <strong>Stanford</strong> Rock Fracture Project. The Valley <strong>of</strong> Fire State Park in Nevada<br />

is a beautiful <strong>and</strong> convenient place to work—thanks to all the helpful, friendly staff there.<br />

My sincere thanks to the faculty who served on my committees <strong>and</strong> provided valuable<br />

guidance: Mark Zoback, Steve Graham, Amos Nur <strong>and</strong>, particularly, Lou Durl<strong>of</strong>sky <strong>and</strong><br />

Atilla Aydin. Lou is great to work with—enthusiastic, incisive <strong>and</strong> human—<strong>and</strong> Atilla<br />

imbues every effort with dedication <strong>and</strong> principle. Thanks also to the administrative staff<br />

for their logistical support, especially Elaine Anderson—font <strong>of</strong> knowledge, c<strong>and</strong>y <strong>and</strong><br />

conversation. My gratitude also goes to Manika Prasad <strong>and</strong> Bob Jones for their tutelage<br />

in the rock physics lab <strong>and</strong> on the SEM, respectively.<br />

This truly has been a multidisciplinary effort, <strong>and</strong> I gratefully acknowledge all my co-<br />

authors <strong>and</strong> collaborators: John Rudnicki (Northwestern University), Lou Durl<strong>of</strong>sky,<br />

Mohammad Karimi-Fard, Jeffrey Chapin, Youngseuk Keehm, Tapan Mukerji, Gaurav<br />

Chopra <strong>and</strong> Ovunc Mutlu. Thanks also to David Lockner (U.S.G.S.), Bill Olsson <strong>and</strong><br />

David Holcomb (S<strong>and</strong>ia National Labs) <strong>and</strong> Bezalel Haimson (Univ. <strong>of</strong> Wisconsin) for<br />

opening their minds <strong>and</strong> their laboratories to my experimental aspirations.<br />

No group deserves more credit than my fellow students <strong>and</strong> researchers, past <strong>and</strong><br />

present. In particular, I acknowledge Eric Flodin, Phil Resor, Nick Davatzes, Jordan<br />

Muller, Juan Mauricio Florez, Frantz Maerten, Peter Eichhubl, Xavier Du Bernard,<br />

Fabrizio Agosta, Ian Mynatt, Ole Kaven, Ashley Griffith, Ghislain de Joussineau <strong>and</strong><br />

Ramil Ahmadov. Special thanks to Brita Graham <strong>and</strong> Tricia Fiore for being “awesome”<br />

<strong>of</strong>fice mates, <strong>and</strong> to everyone for their humor, insights <strong>and</strong> precious computer skills.<br />

My good friend from Yale daze, John Childs, deserves an enormous thank you for<br />

donating his time <strong>and</strong> assistance in the field on three occasions over three years—all<br />

v


delivered with fortitude, humor, surprisingly good penmanship, enlightened conversation<br />

<strong>and</strong> a car. What more could one ask from a history major <strong>and</strong> wooden objects conservator<br />

who could have spent the extra time on his sailboat headed to Catalina? Thanks John.<br />

This brings me to David Pollard, my principal advisor, collaborator <strong>and</strong> patron. I<br />

arrived at <strong>Stanford</strong> 15 years after turning down his initial <strong>of</strong>fer <strong>of</strong> admission, a story Dave<br />

tells with relish. No matter my reasons for that youthful decision, Dave had made an<br />

impression on me. When I finally decided to hunker down on that Ph.D., I approached<br />

only him. Incredibly, he remembered <strong>and</strong> welcomed me, <strong>of</strong>fering full funding to plug into<br />

a project in some valley full <strong>of</strong> “b<strong>and</strong>s.” The details mattered less to me than the pure<br />

educational opportunity <strong>of</strong> it, the exposure to a mind <strong>and</strong> philosophy at the cutting edge<br />

<strong>of</strong> <strong>structural</strong> <strong>geology</strong> <strong>and</strong> scientific inquiry itself. I have not been disappointed, <strong>and</strong> along<br />

the way managed to carve out my own niche in a fascinating field. Dave enabled this,<br />

with measured patience, open-minded rigor <strong>and</strong> humanity. No matter how dark the day, I<br />

never left his <strong>of</strong>fice without feeling better—more calm, clear <strong>and</strong> confident—than when I<br />

had entered. Anyone who has experienced graduate school knows what a compliment<br />

that is. And yes Dave, <strong>Stanford</strong> is an incredible place, golf carts notwithst<strong>and</strong>ing.<br />

Of course family plays a huge, albeit at times less visible, role, <strong>and</strong> my journey is no<br />

exception. Thanks to my mother <strong>and</strong> father, Eleanor <strong>and</strong> Paul, for instilling me with<br />

determination, <strong>and</strong> to my siblings, Karl, Mark <strong>and</strong> Erika, for helping with humor <strong>and</strong><br />

perspective maintenance. To Sarah, the best mother in law a guy could have, thanks for<br />

all the baked goods, barbeques <strong>and</strong> babying. And to deep-thinker Rye the dog, who so<br />

<strong>of</strong>ten kept my feet warm <strong>and</strong> pinned to the floor beneath my desk..........he’s a good boy.<br />

Finally, I owe my greatest debt <strong>of</strong> gratitude <strong>and</strong> all my love to Beth, who encouraged<br />

<strong>and</strong> nurtured me throughout the entire Ph.D. saga—the ups <strong>and</strong> downs from conception<br />

to completion <strong>and</strong> everything in between. She even exhibited the unmitigated faith <strong>and</strong><br />

optimism to marry me smack dab in the middle <strong>of</strong> it all. Although I can never hope to<br />

compensate her for the years <strong>of</strong> late nights, lost weekends, fits <strong>of</strong> distraction <strong>and</strong> bouts <strong>of</strong><br />

insomnia, I promise with all my heart to try—lavishing her with all the extra attention<br />

until so recently squ<strong>and</strong>ered on my laptop. Beth is my inspiration, my strength, my love<br />

<strong>and</strong> my best friend. I dedicate this dissertation to her.<br />

vi<br />

KRS, March 21 st , 2006


Table <strong>of</strong> Contents<br />

Abstract..............................................................................................................................iv<br />

Acknowledgements............................................................................................................v<br />

Table <strong>of</strong> Contents.............................................................................................................vii<br />

List <strong>of</strong> Tables.....................................................................................................................ix<br />

List <strong>of</strong> Illustrations.............................................................................................................x<br />

Introduction........................................................................................................................1<br />

Chapter 1—Structural <strong>geology</strong> <strong>and</strong> tectonic interpretation <strong>of</strong> compaction b<strong>and</strong>s<br />

in the Aztec s<strong>and</strong>stone <strong>of</strong> southeastern Nevada............................................9<br />

1. Abstract......................................................................................................................9<br />

2. Introduction..............................................................................................................10<br />

3. Setting <strong>and</strong> history <strong>of</strong> study area..............................................................................11<br />

3.1. Deposition.......................................................................................................11<br />

3.2. Diagenesis.......................................................................................................15<br />

3.3. Deformation....................................................................................................16<br />

4. Structural <strong>geology</strong> <strong>of</strong> compaction b<strong>and</strong>s..................................................................19<br />

4.1. Thin-section to outcrop scale..........................................................................19<br />

4.2. Anticrack model..............................................................................................22<br />

4.3. Outcrop to regional scale................................................................................22<br />

5. Compaction b<strong>and</strong> orientations.................................................................................24<br />

6. Tectonic interpretation.............................................................................................28<br />

6.1. Temporal, spatial <strong>and</strong> material constraints.....................................................28<br />

6.2. Paleostress analysis.........................................................................................31<br />

6.3. Geomechanical implications...........................................................................33<br />

7. Concluding observations..........................................................................................37<br />

9. Acknowledgements..................................................................................................38<br />

Chapter 2—Anticrack-inclusion model for compaction b<strong>and</strong>s in s<strong>and</strong>stone................39<br />

1. Abstract....................................................................................................................39<br />

2. Introduction..............................................................................................................39<br />

3. Methods....................................................................................................................42<br />

4. Field <strong>and</strong> petrographic analysis...............................................................................44<br />

4.1. Geological setting <strong>and</strong> paleostress state..........................................................44<br />

4.2. Outcrop analysis..............................................................................................46<br />

4.3. Petrographic analysis......................................................................................50<br />

5. Anticrack-inclusion conceptual model....................................................................57<br />

6. Elastic properties......................................................................................................60<br />

7. Mechanical analysis.................................................................................................61<br />

7.1. Embedded layer model...................................................................................61<br />

7.2. Eshelby inclusion model.................................................................................64<br />

7.3. Anticrack model..............................................................................................67<br />

8. Discussion................................................................................................................71<br />

9. Acknowledgements..................................................................................................74<br />

vii


Chapter 3—Energy-release model <strong>of</strong> compaction b<strong>and</strong> <strong>propagation</strong>............................75<br />

1. Abstract....................................................................................................................75<br />

2. Introduction..............................................................................................................75<br />

3. Formulation..............................................................................................................78<br />

4. Special cases............................................................................................................81<br />

5. Discussion................................................................................................................82<br />

6. Acknowledgements..................................................................................................84<br />

Chapter 4—Propagation <strong>of</strong> compaction b<strong>and</strong>s in s<strong>and</strong>stone as anticracks:<br />

Observational evidence, mechanical theory <strong>and</strong> numerical simulation.....85<br />

1. Abstract....................................................................................................................85<br />

2. Introduction..............................................................................................................85<br />

3. Field evidence <strong>and</strong> interpretation.............................................................................88<br />

3.1. Outcrop observations......................................................................................89<br />

3.2. Petrographic observations...............................................................................92<br />

3.3. Anticrack-inclusion interpretation..................................................................96<br />

4. Mechanical theory....................................................................................................98<br />

5. Numerical model....................................................................................................104<br />

6. Propagation simulation..........................................................................................107<br />

6.1. Model calibration <strong>and</strong> stability testing..........................................................107<br />

6.2. Symmetric <strong>propagation</strong>.................................................................................112<br />

6.3. Approaching tip interactions.........................................................................114<br />

7. Discussion <strong>and</strong> conclusions...................................................................................119<br />

8. Acknowledgements................................................................................................123<br />

Chapter 5—Computational estimation <strong>of</strong> compaction b<strong>and</strong> permeability:<br />

From thin-section estimations to reservoir implications...........................125<br />

1. Abstract..................................................................................................................125<br />

2. Introduction............................................................................................................125<br />

3. Computational method...........................................................................................130<br />

3.1. Image processing..........................................................................................130<br />

3.2. Pore structure realization..............................................................................130<br />

3.3. Permeability estimation................................................................................131<br />

4. Application to the Aztec s<strong>and</strong>stone........................................................................134<br />

4.1. Simulation results..........................................................................................134<br />

4.2. Comparison to measured values...................................................................137<br />

5. Conclusions............................................................................................................139<br />

6. Acknowledgements................................................................................................140<br />

Chapter 6—Permeability effects <strong>of</strong> deformation b<strong>and</strong> arrays in s<strong>and</strong>stone................141<br />

1. Abstract..................................................................................................................141<br />

2. Introduction............................................................................................................141<br />

3. Deformation b<strong>and</strong>s.................................................................................................143<br />

4. Characteristic DB patterns in the Aztec s<strong>and</strong>stone................................................147<br />

4.1. Parallel..........................................................................................................148<br />

viii


4.2. Cross-hatch...................................................................................................148<br />

4.3. Anastomosing...............................................................................................150<br />

5. Numerical modeling methods................................................................................150<br />

5.1. Governing equations.....................................................................................152<br />

5.2. Boundary conditions <strong>and</strong> solution................................................................154<br />

5.3. Finite difference/finite element method........................................................157<br />

6. Effective permeability results................................................................................157<br />

6.1. Parallel..........................................................................................................158<br />

6.2. Cross-hatch...................................................................................................160<br />

6.3. Anastomosing...............................................................................................163<br />

7. Summary................................................................................................................165<br />

8. Discussion <strong>and</strong> conclusions...................................................................................166<br />

9. Acknowledgements................................................................................................167<br />

Chapter 7—Flow <strong>and</strong> transport effects <strong>of</strong> compaction b<strong>and</strong>s in s<strong>and</strong>stone<br />

at scales relevant to aquifer <strong>and</strong> reservoir management...........................169<br />

1. Abstract..................................................................................................................169<br />

2. Introduction............................................................................................................169<br />

3. Mapping.................................................................................................................173<br />

4. Flow simulation model..........................................................................................175<br />

5. Simulations............................................................................................................182<br />

5.1. Well-to-well flow..........................................................................................182<br />

5.1.1. Results..................................................................................................182<br />

5.1.2. Discussion............................................................................................185<br />

5.2 Reservoir production......................................................................................187<br />

5.2.1. Results..................................................................................................187<br />

5.2.2. Discussion............................................................................................190<br />

5.3. Contaminant transport...................................................................................190<br />

5.3.1. Results..................................................................................................192<br />

5.3.2. Discussion............................................................................................192<br />

6. General discussion.................................................................................................196<br />

7. Conclusions............................................................................................................198<br />

8. Acknowledgements................................................................................................199<br />

References.......................................................................................................................201<br />

List <strong>of</strong> Tables<br />

Table 1.1. Summary <strong>of</strong> compaction b<strong>and</strong> orientation data...............................................29<br />

Table 1.2. Summary <strong>of</strong> restored compaction b<strong>and</strong> orientation data..................................32<br />

Table 5.1. Summary <strong>of</strong> permeability estimates...............................................................136<br />

Table 7.1. Relative permeability data for reservoir production simulations...................189<br />

ix


List <strong>of</strong> Illustrations<br />

Figure A. Cover photo from AAPG Bulletin (publication <strong>of</strong> Chapter 6)..........................xii<br />

Figure 1.1. Location <strong>and</strong> general <strong>geology</strong> <strong>of</strong> study area...................................................12<br />

Figure 1.2. Generalized stratigraphic column...................................................................14<br />

Figure 1.3. Schematized fault map for Lake Mead region................................................17<br />

Figure 1.4. Photo collage <strong>of</strong> compaction b<strong>and</strong>s in the Aztec s<strong>and</strong>stone...........................20<br />

Figure 1.5. Schematic <strong>of</strong> anticrack conceptual model......................................................23<br />

Figure 1.6. Collection <strong>of</strong> orientation data in the field.......................................................25<br />

Figure 1.7. Located orientation data from Valley <strong>of</strong> Fire.................................................26<br />

Figure 1.8. Orientation data for outcrops with multiple compaction b<strong>and</strong> sets................27<br />

Figure 1.9. Stereographs for combined orientation data...................................................29<br />

Figure 1.10. Stereographs for restored orientation data....................................................32<br />

Figure 1.11. Paleostress orientations <strong>and</strong> interpretation...................................................36<br />

Figure 2.1. Location <strong>of</strong> the Valley <strong>of</strong> Fire State Park, Nevada.........................................41<br />

Figure 2.2. Compaction b<strong>and</strong>s (CBs) in outcrop near Silica Dome..................................43<br />

Figure 2.3. Typical CB fin in outcrop...............................................................................48<br />

Figure 2.4. Compaction b<strong>and</strong> thickness pr<strong>of</strong>ile data.........................................................49<br />

Figure 2.5. Composite photomicrograph <strong>of</strong> a CB.............................................................51<br />

Figure 2.6. Porosity pr<strong>of</strong>iles across a CB..........................................................................52<br />

Figure 2.7. Backscatter electron images <strong>of</strong> s<strong>and</strong>stone <strong>and</strong> a CB......................................53<br />

Figure 2.8. Distribution <strong>of</strong> porosity inside a CB...............................................................55<br />

Figure 2.9. Backscatter electron images <strong>of</strong> grain damage in a CB...................................56<br />

Figure 2.10. Backscatter electron image <strong>of</strong> incipient CB.................................................55<br />

Figure 2.11. Schematic representations <strong>of</strong> idealized CB model.......................................58<br />

Figure 2.12. Near-tip stress components for the embedded layer model..........................65<br />

Figure 2.13. Stress distributions for the Eshelby inclusion model....................................66<br />

Figure 2.14. Stress comparison <strong>of</strong> Eshelby <strong>and</strong> anticrack models....................................69<br />

Figure 2.15. Near-tip stress comparison <strong>of</strong> Eshelby <strong>and</strong> anticrack models......................70<br />

Figure 2.16. Mean normal stress distribution around an anticrack tip..............................70<br />

Figure 3.1. Anatomy <strong>of</strong> a compaction b<strong>and</strong> in the Aztec s<strong>and</strong>stone.................................76<br />

Figure 3.2. Model <strong>of</strong> a semi-infinite CB embedded in an infinite layer...........................79<br />

Figure 3.3. Outcrop basis for the embedded layer CB model...........................................79<br />

Figure 3.4. Hypothetical 1-D distribution <strong>of</strong> stress near a CB tip....................................84<br />

Figure 4.1. Location <strong>of</strong> the Valley <strong>of</strong> Fire State Park, Nevada.........................................87<br />

Figure 4.2. Array <strong>of</strong> subparallel CBs in the Aztec s<strong>and</strong>stone...........................................87<br />

Figure 4.3. Typical compaction b<strong>and</strong> patterns exposed in outcrop..................................90<br />

Figure 4.4. Mechanical interaction between CBs <strong>and</strong> æolian bedding.............................91<br />

Figure 4.5. Tip-to-tip thickness pr<strong>of</strong>ile <strong>of</strong> a 25-m-long CB..............................................91<br />

Figure 4.6. Various CB <strong>propagation</strong> behaviors observed in outcrop................................93<br />

Figure 4.7. Examples <strong>of</strong> tip-to-tip interactions observed in outcrop................................94<br />

Figure 4.8. Composite photomicrograph <strong>of</strong> CB anatomy.................................................95<br />

Figure 4.9. Schematic representations <strong>of</strong> the idealized CB model...................................97<br />

Figure 4.10. Schematic <strong>of</strong> the global <strong>and</strong> anticrack tip coordinate systems.....................99<br />

Figure 4.11. Contour plots <strong>of</strong> near tip stress around an anticrack..................................100<br />

Figure 4.12. Polar near-tip stress as a function <strong>of</strong> θ for a fixed radius...........................102<br />

Figure 4.13. Effect <strong>of</strong> differential stress on <strong>propagation</strong> path stability...........................102<br />

x


Figure 4.14. Mechanical interactions between anticrack tips.........................................105<br />

Figure 4.15. Essential schematic <strong>of</strong> displacement discontinuity BEM...........................105<br />

Figure 4.16. Stiffness dependence <strong>of</strong> closing-mode displacement discontinuity...........109<br />

Figure 4.17. Near-tip stress comparison <strong>of</strong> Eshelby <strong>and</strong> anticrack models....................109<br />

Figure 4.18. Propagation path stability as a function <strong>of</strong> scale <strong>and</strong> remote stress............111<br />

Figure 4.19. Scale <strong>and</strong> remote stress dependence <strong>of</strong> <strong>propagation</strong> path stability.............111<br />

Figure 4.20. Element length <strong>and</strong> stress dependence <strong>of</strong> <strong>propagation</strong> path stability.........113<br />

Figure 4.21a. Remote stress <strong>and</strong> spacing effects on CB tip interactions........................116<br />

Figure 4.21b. Remote stress <strong>and</strong> spacing effects on CB tip interactions........................117<br />

Figure 4.21c. Remote stress <strong>and</strong> spacing effects on CB tip interactions........................118<br />

Figure 4.22. Influence <strong>of</strong> a CB length on approaching tip interactions..........................119<br />

Figure 4.23. Conceptual model <strong>of</strong> strain <strong>and</strong> grain damage around a CB......................122<br />

Figure 4.24. Relationship <strong>of</strong> a CB array to permeability <strong>and</strong> paleostress.......................124<br />

Figure 5.1. Location <strong>of</strong> the Valley <strong>of</strong> Fire State Park, Nevada.......................................126<br />

Figure 5.2. Typical CB in outcrop <strong>and</strong> its effect on fluid flow.......................................128<br />

Figure 5.3. Representative outcrop array <strong>of</strong> subparallel, anastomosing CBs.................129<br />

Figure 5.4. Composite photomicrograph <strong>of</strong> a representative CB...................................132<br />

Figure 5.5. Permeability estimation methodology workflow.........................................133<br />

Figure 5.6. Backscatter electron images <strong>of</strong> s<strong>and</strong>stone <strong>and</strong> CBs......................................135<br />

Figure 5.7. Porosity-permeability estimates versus measured values.............................136<br />

Figure 5.8. Computational estimation <strong>of</strong> permeability anisotropy.................................138<br />

Figure 6.1. Location <strong>of</strong> the Valley <strong>of</strong> Fire State Park, Nevada.......................................144<br />

Figure 6.2. Compactive deformation b<strong>and</strong>s in the Aztec s<strong>and</strong>stone...............................145<br />

Figure 6.3. Typical outcrop pattern <strong>of</strong> subparallel compactive DBs..............................149<br />

Figure 6.4. Typical cross-hatch pattern <strong>of</strong> compactive DBs...........................................149<br />

Figure 6.5. Typical anastomosing pattern <strong>of</strong> compactive DBs.......................................151<br />

Figure 6.6. Idealized b<strong>and</strong> pattern <strong>and</strong> model grid representation..................................153<br />

Figure 6.7. Schematic representations <strong>of</strong> b<strong>and</strong> pattern upscaling...................................153<br />

Figure 6.8. Key parameters for computing effective block permeability.......................156<br />

Figure 6.9. Effective permeability for perfectly parallel b<strong>and</strong> patterns..........................159<br />

Figure 6.10. Effective permeability for orthogonal cross-hatch b<strong>and</strong> patterns...............161<br />

Figure 6.11. Effective permeability for acute cross-hatch b<strong>and</strong> patterns........................162<br />

Figure 6.12. Effective permeability for a real anastomosing b<strong>and</strong> pattern.....................164<br />

Figure 7.1. Location <strong>and</strong> air photo <strong>of</strong> the Aztec s<strong>and</strong>stone, Valley <strong>of</strong> Fire, NV............170<br />

Figure 7.2. Compaction b<strong>and</strong>s in outcrop <strong>and</strong> thin section.............................................172<br />

Figure 7.3. Compaction b<strong>and</strong> trace map.........................................................................174<br />

Figure 7.4. Schematic <strong>of</strong> transmissibility calculation parameters..................................177<br />

Figure 7.5. Triangular discretization <strong>of</strong> compaction b<strong>and</strong> trace map.............................181<br />

Figure 7.6. Well locations for single-phase flow simulations.........................................183<br />

Figure 7.7. Pressure drop results for single-phase flow simulations..............................184<br />

Figure 7.8. Well configurations for reservoir production simulations............................186<br />

Figure 7.9. Saturation-map snapshots for reservoir production simulations..................188<br />

Figure 7.10. Production efficiency for the two reservoir production scenarios..............189<br />

Figure 7.11. Contaminant leak simulation configurations..............................................191<br />

Figure 7.12. Contaminant plume with b<strong>and</strong>s parallel to regional gradient.....................193<br />

Figure 7.13. Contaminant plume with b<strong>and</strong>s oblique to regional gradient.....................194<br />

Figure 7.14. Contaminant plume with b<strong>and</strong>s normal to regional gradient.....................195<br />

xi


Figure A. Cover photo that accompanied publication <strong>of</strong> Chapter 6 in AAPG Bulletin<br />

(Sternl<strong>of</strong> et al., 2004). This view shows steeply east-dipping compaction b<strong>and</strong>s in the<br />

bleached middle Aztec s<strong>and</strong>stone, just above the red, iron oxide-stained lower Aztec.<br />

Dark Paleozoic rocks <strong>of</strong> the over-riding Cretaceous Muddy Mountain Thrust loom in the<br />

background. These compaction b<strong>and</strong>s are approximately 1 cm thick <strong>and</strong> 100 m long.<br />

xii


Introduction<br />

Some brittle structures in rock act to enhance <strong>and</strong> channel fluid flow, others to<br />

impede <strong>and</strong> compartmentalize it. Commonly, structures <strong>of</strong> both types interact with <strong>and</strong><br />

overprint each other to form complex fabrics that exert a dominating influence on flow in<br />

subsurface aquifers <strong>and</strong> reservoirs, even those endowed with high intrinsic permeability<br />

such as porous s<strong>and</strong>stone. Underst<strong>and</strong>ing the distribution, geometry <strong>and</strong> flow properties<br />

<strong>of</strong> these <strong>structural</strong> fabrics is therefore vital to optimal resource production <strong>and</strong><br />

management. Unfortunately, data on the attributes <strong>of</strong> these fabrics—whether from 3-D<br />

seismic surveys, borehole geophysical logs, pump tests <strong>and</strong> production data, or core<br />

analysis—are generally limited <strong>and</strong> may be virtually nonexistent for small-scale<br />

structures. One approach to addressing this problem is to study exhumed analog aquifers<br />

<strong>and</strong> reservoirs, where structures <strong>and</strong> fabrics can easily be observed (at least in 2-D),<br />

sampled <strong>and</strong> analyzed. The ultimate goal <strong>of</strong> such studies is to develop insights <strong>and</strong> tools<br />

for forecasting the presence <strong>and</strong> influence <strong>of</strong> structures in the subsurface based on limited<br />

hard data.<br />

That has been the theme <strong>of</strong> an ongoing, decade-long research project undertaken by<br />

David Pollard <strong>and</strong> Atilla Aydin with funding from the U.S. Department <strong>of</strong> Energy, Office<br />

<strong>of</strong> Basic Energy Sciences—Structural Heterogeneities <strong>and</strong> Paleo Fluid Flow in an<br />

Analog S<strong>and</strong>stone Reservoir. My thesis work fits into this larger research initiative, just<br />

as my topic—deformation b<strong>and</strong>s <strong>and</strong>, more specifically, compaction b<strong>and</strong>s—fits into the<br />

framework <strong>of</strong> structures that did demonstrably impact paleo fluid flow in the Aztec<br />

s<strong>and</strong>stone <strong>and</strong> are now breathtakingly exposed in the aptly named Valley <strong>of</strong> Fire <strong>of</strong><br />

southeastern Nevada. My effort follows three earlier <strong>Stanford</strong> dissertations that also took<br />

shape from the colorful outcrops <strong>of</strong> the Valley <strong>of</strong> Fire: Fluid flow <strong>and</strong> chemical alteration<br />

in fractured s<strong>and</strong>stone (W.L. Taylor, 1999); Structure <strong>and</strong> hydraulics <strong>of</strong> brittle faults in<br />

s<strong>and</strong>stone (R.D. Myers, 1999); <strong>and</strong> Structural evolution, petrophysics <strong>and</strong> large-scale<br />

permeability <strong>of</strong> faults in s<strong>and</strong>stone, Valley <strong>of</strong> Fire, Nevada (E.A. Flodin, 2003).<br />

This dissertation focuses on the oldest <strong>structural</strong> fabric components present in the<br />

Aztec s<strong>and</strong>stone—compaction b<strong>and</strong>s—<strong>and</strong> endeavors to comprehend what they are, why<br />

they formed, <strong>and</strong> how they influence bulk permeability <strong>and</strong> fluid flow. In essence, I take<br />

a broad look at a specific <strong>structural</strong> phenomenon, one which physical logic dictates is<br />

1


likely present in subsurface s<strong>and</strong>stone equivalents <strong>of</strong> the Aztec. Coming at this research<br />

as a traditional <strong>structural</strong> geologist, I have therefore collaborated widely—in reservoir<br />

engineering, petrophysics, <strong>and</strong> theoretical, experimental <strong>and</strong> applied <strong>mechanics</strong>—to bring<br />

the requisite tools to bear. I could not have achieved the breadth <strong>of</strong> this thesis working in<br />

isolation, <strong>and</strong> consider these collaborations to be a particular strength <strong>of</strong> the effort.<br />

The thesis is comprised <strong>of</strong> seven chapters, each written as a st<strong>and</strong>-alone manuscript<br />

intended for peer-reviewed publication. As a result, some repetition, particularly with<br />

introductory material, was inevitable. Also, although arranged thematically, the chapters<br />

do not necessarily flow seamlessly one into the next, as they might in a thesis written to<br />

be a coherent book. Indeed, each paper was written with an eye toward the editorial<br />

requirements <strong>of</strong> a different journal, <strong>and</strong> they were finished in the order: 6, 2, 3, 7, 5, 4, 1.<br />

A discerning reader might detect evidence for the evolution <strong>of</strong> my thinking over the past<br />

few years preserved within these pages. For example, Chapter 6 does not differentiate<br />

compaction b<strong>and</strong>s (CBs), which are prevalent in the Aztec s<strong>and</strong>stone, from the more<br />

general category <strong>structural</strong> category <strong>of</strong> deformation b<strong>and</strong>s (DBs). A brief overview <strong>of</strong><br />

each chapter follows.<br />

Chapter 1 describes the occurrence <strong>and</strong> <strong>structural</strong> <strong>geology</strong> <strong>of</strong> CBs <strong>and</strong> CB arrays, <strong>and</strong><br />

considers their significance within the context <strong>of</strong> the depositional, diagenetic <strong>and</strong> tectonic<br />

history <strong>of</strong> the Aztec s<strong>and</strong>stone through Cretaceous time. As the final paper written, this<br />

draft manuscript represents a preliminary synthesis <strong>of</strong> my work in the Aztec <strong>and</strong> its<br />

implications for predicting CB occurrence in s<strong>and</strong>stone based on material <strong>and</strong> stress<br />

history or, conversely, interpreting material <strong>and</strong> stress history based on the occurrence <strong>of</strong><br />

CBs. The paper is based on my detailed observations conducted at thin-section, outcrop<br />

<strong>and</strong> regional scales, as well as on my review <strong>and</strong> interpretation <strong>of</strong> the work <strong>of</strong> others.<br />

David Pollard deserves substantial credit for his tutelage, editing <strong>and</strong> ability to foment<br />

ideas. Submission <strong>of</strong> a final manuscript based on Chapter 1 to Tectonics is anticipated.<br />

Chapter 2 presents the evidence <strong>and</strong> arguments for a mechanical interpretation <strong>of</strong> CBs<br />

as contractile Eshelby inclusions, <strong>and</strong> for considering them as anticracks for the purposes<br />

<strong>of</strong> <strong>propagation</strong> modeling using linear elastic theory <strong>and</strong> the displacement discontinuity<br />

boundary element method. All <strong>of</strong> the data collection, both in the field <strong>and</strong> the lab, all <strong>of</strong><br />

the analytical <strong>and</strong> numerical modeling, <strong>and</strong> all <strong>of</strong> the original writing <strong>and</strong> figure drafting<br />

2


was performed by me. The ideas expressed were developed in collaboration with my co-<br />

authors, principally David Pollard <strong>and</strong> secondarily John Rudnicki <strong>of</strong> Northwestern<br />

University. The embedded-layer model analysis was originated by Rudnicki. The Eshelby<br />

code was generously provided by Pradeep Sharma <strong>of</strong> the University <strong>of</strong> Houston, <strong>and</strong><br />

adapted for my use by fellow student Ole Kaven. The boundary element code was written<br />

to my specifications by another fellow student, Gaurav Chopra, in close consultation with<br />

Pollard <strong>and</strong> with substantial input from me. This paper was published in the November<br />

2005 issue <strong>of</strong> Journal <strong>of</strong> Geophysical Research (v. 110, n. B11).<br />

Chapter 3 represents an additional fruit <strong>of</strong> my collaboration with John Rudnicki, who<br />

had conceived <strong>of</strong> an energy-release analysis for compaction-b<strong>and</strong> <strong>propagation</strong>, but was<br />

lacking a convincing physical context in which to frame it. As a result <strong>of</strong> my work with<br />

him on Chapter 2, he came around to the view that his theoretical model could <strong>and</strong> should<br />

be grounded primarily in field observations rather than in experimental results, which I<br />

contend produce a phenomenology largely distinct from CB formation in nature. This<br />

subtle feat <strong>of</strong> scientific diplomacy helped to pierce a somewhat insular <strong>and</strong><br />

counterproductive feedback loop between theorists <strong>and</strong> experimentalists working on<br />

compaction localization. I am the second author after Rudnicki on this paper, <strong>and</strong><br />

contributed extensive rewrites <strong>and</strong> editing (both technical <strong>and</strong> grammatical) to<br />

fundamentally recast the paper as outlined above. I also conducted targeted fieldwork in<br />

the Valley <strong>of</strong> Fire to produce the second figure <strong>of</strong> two that appears in the manuscript. As<br />

always, substantial credit is due David Pollard for his guidance <strong>and</strong> editing. This succinct<br />

contribution, an important step toward establishing a stress/strain/strength activation limit<br />

for compaction <strong>propagation</strong> in s<strong>and</strong>stone, was published in the second August 2005 issue<br />

<strong>of</strong> Geophysical Research Letters (v. 32, n. 16).<br />

Chapter 4 takes the anticrack model for CBs developed in Chapter 2 <strong>and</strong> applies it<br />

through the boundary element method (BEM) code (also introduced in Chapter 2) to<br />

investigate the <strong>mechanics</strong> <strong>of</strong> CB <strong>propagation</strong>, interaction <strong>and</strong> pattern development in the<br />

Aztec s<strong>and</strong>stone. The scientific premise behind this preliminary draft manuscript is that<br />

all the various 2-D patterns <strong>of</strong> CB interaction revealed in outcrop express the same<br />

fundamental mechanism <strong>of</strong> <strong>propagation</strong>. I attempt to recreate some simple diagnostic<br />

patterns—in-plane <strong>propagation</strong>, “zig-zag” <strong>and</strong> anastamosing <strong>propagation</strong> instability <strong>and</strong><br />

3


hooking-tip interactions—using the BEM code. The degree <strong>of</strong> success attained suggests<br />

that the linear elastic anticrack mechanical model provides a valid first approximation for<br />

conceptualizing <strong>and</strong> simulating CB <strong>propagation</strong>. Perhaps the most interesting <strong>and</strong><br />

potentially useful result <strong>of</strong> this paper is the model observation that anticrack CBs respond<br />

to high compressive normal stress (σ2 ≈ σ1) oriented parallel to their trend with<br />

<strong>propagation</strong>-path instability. This is opposite to the instability effect observed for<br />

opening-mode cracks (Olson <strong>and</strong> Pollard, 1989; Thomas <strong>and</strong> Pollard, 1993), <strong>and</strong> could be<br />

used to help forecast the degree <strong>of</strong> anastomosis in subsurface CB arrays as a function <strong>of</strong><br />

stress history. By the same token, the configuration <strong>of</strong> an exhumed CB array could be<br />

used to constrain the ambient principal paleostress state (orientation <strong>and</strong> magnitude) in<br />

which it formed. For this paper, I performed all the fieldwork, modeling, writing <strong>and</strong><br />

figure drafting, with guidance <strong>and</strong> editing provided by second-author David Pollard. For<br />

the modeling, I used the BEM code written by third-author Gaurav Chopra (Chapter 2).<br />

Submission <strong>of</strong> a final manuscript based on Chapter 4 to Journal <strong>of</strong> Structural Geology is<br />

anticipated.<br />

Chapter 5 represents another outgrowth <strong>of</strong> my collaborative efforts, in this case with<br />

Tapan Mukerji <strong>and</strong> Youngseuk Keehm <strong>of</strong> the <strong>Stanford</strong> Rock Physics <strong>and</strong> Borehole<br />

Geophysics (SRB) Group. I had gone to Mukerji for help in using MATLAB® to<br />

perform automated image-analysis measurements <strong>of</strong> porosity from my thin-sections, the<br />

results <strong>of</strong> which contributed to every other chapter in this thesis. For the purposes <strong>of</strong> the<br />

fluid-flow modeling presented in Chapter 7, I had also contemplated collecting<br />

permeability measurements, but was dissuaded by the difficulty <strong>of</strong> obtaining reliable<br />

results from thin CBs at reasonable expense. Through Mukerji, however, I learned <strong>of</strong> the<br />

computational permeability estimation algorithm that had been developed in the SRB,<br />

principally by Keehm, <strong>and</strong> could be used to generate virtual permeability measurements<br />

from my existing thin sections. I recognized the opportunity to perform a practical test <strong>of</strong><br />

this new tool within the context <strong>of</strong> the analog reservoir research concept. Specifically, we<br />

applied the method to my prize thin section <strong>and</strong> compared the estimation results to<br />

available permeability measurement data for both CBs <strong>and</strong> host rock from the Aztec <strong>and</strong><br />

Navajo s<strong>and</strong>stones. Correspondence was excellent, suggesting that, for a subsurface<br />

equivalent <strong>of</strong> the CB-rich Aztec from which only scarce, potentially unconsolidated<br />

4


samples were available, representative flow properties can be gleaned from even a single<br />

thin section. In addition to conceiving <strong>of</strong> the project, I conducted all the fieldwork <strong>and</strong><br />

microscopy involved, <strong>and</strong> produced all the figures <strong>and</strong> text (except for the Methods<br />

section, which I edited, <strong>and</strong> Figure 5). Keehm <strong>and</strong> Mukerji performed the permeability<br />

estimations. David Pollard provided editing <strong>and</strong> valuable guidance. This manuscript has<br />

yet to be placed for publication.<br />

Chapter 6 began as a project undertaken in 1999 by Jeffrey Chapin, a visiting<br />

researcher working with David Pollard. Chapin ab<strong>and</strong>oned the work while in draft form<br />

<strong>and</strong> switched disciplines from <strong>geology</strong> to mechanical engineering. I inherited the project<br />

<strong>and</strong> brought it to fruition. Initially I performed a thorough editing <strong>and</strong> culling job on the<br />

rough draft, including moderate rewriting, <strong>and</strong> submitted it to American Association <strong>of</strong><br />

Petroleum Geologists Bulletin in September 2002 as second author after Chapin, with co-<br />

authors Louis Durl<strong>of</strong>sky <strong>and</strong> Pollard. That version <strong>of</strong> the manuscript was rejected. In<br />

assuming first authorship at Pollard’s behest, I then performed a complete rewrite,<br />

including a strategic reconceptualization, additional culling <strong>and</strong> the redrafting <strong>of</strong> all<br />

figures. This reformulated version <strong>of</strong> the paper, which presents an analytical <strong>and</strong><br />

numerical framework for underst<strong>and</strong>ing how characteristic patterns <strong>of</strong> deformation b<strong>and</strong>s<br />

exposed in the Aztec s<strong>and</strong>stone would transform bulk effective permeability, was<br />

published in the September 2004 issue <strong>of</strong> American Association <strong>of</strong> Petroleum Geologists<br />

Bulletin (v. 88, n. 9).<br />

Chapter 7 grew out <strong>of</strong> reviewer feedback that came with shepherding Chapter 6<br />

through to publication, <strong>and</strong> my desire to assume more complete ownership <strong>of</strong> the earlier<br />

work by exp<strong>and</strong>ing the effort. Comments by the reviewers, both cogent <strong>and</strong> <strong>of</strong>f-target,<br />

suggested that, in order convincingly to assess the potential flow effects <strong>of</strong> compaction<br />

b<strong>and</strong>s, I needed to consider a realistic b<strong>and</strong> pattern at a scale <strong>of</strong> clear relevance to aquifer<br />

<strong>and</strong> reservoir modeling, rather than appealing to arguments based on up-scaling<br />

techniques as in Chapter 6. I saw the opportunity to achieve this goal in low-altitude<br />

aerial photographs taken <strong>of</strong> the Valley <strong>of</strong> Fire for another purpose. Viewed at full<br />

resolution, the patterns <strong>of</strong> CBs cropping out in positive relief are clearly visible, which<br />

enabled me to construct a detailed <strong>and</strong> representative map <strong>of</strong> cm-thick CBs over an area<br />

<strong>of</strong> 150,000 m 2 . Using this map in conjunction with realistic flow properties for the b<strong>and</strong>s<br />

5


<strong>and</strong> s<strong>and</strong>stone, my co-authors—Mohammad Karimi-Fard, David Pollard <strong>and</strong> Louis<br />

Durl<strong>of</strong>sky—<strong>and</strong> I were able to model the impact <strong>of</strong> the b<strong>and</strong>s for a variety <strong>of</strong> practical<br />

scenarios: well-pair pump testing, five-spot reservoir production <strong>and</strong> contaminant plume<br />

migration. Our results confirm that the anastomosing b<strong>and</strong> pattern as mapped would exert<br />

significant pressure drop <strong>and</strong> directional fluid transport effects over hundreds <strong>of</strong> meters.<br />

In addition to conceiving <strong>of</strong> the project, I conducted all <strong>of</strong> the fieldwork <strong>and</strong> photo<br />

mapping (including arranging for additional air photos to be taken), designed the five-<br />

spot <strong>and</strong> plume scenarios modeled, <strong>and</strong> produced all the text (except for the Flow Model<br />

section) <strong>and</strong> final figures. Karimi-Fard ran the flow simulations—an arduous task—using<br />

a model code previously developed, primarily by him. He also collaborated closely in<br />

designing the simulation scenarios <strong>and</strong> figures, wrote the Flow Model section, <strong>and</strong><br />

participated in every sense as an equal research partner. Pollard <strong>and</strong> Durl<strong>of</strong>sky provided<br />

guidance <strong>and</strong> thorough editing. This paper has been accepted for publication in Water<br />

Resources Research <strong>and</strong> is currently in press.<br />

This introduction would not be complete without touching on my experimental efforts<br />

<strong>and</strong> collaborations, which failed to bear publishable fruit in time for inclusion in this<br />

dissertation, but did add about a year to my tenure at <strong>Stanford</strong>. The desire to test the<br />

anticrack CB mechanical interpretation in the lab, <strong>and</strong> by so doing shed light on the<br />

enigmatic nature <strong>of</strong> CB tip-zone processes, was irresistible to me—practical timing<br />

concerns voiced by my committee notwithst<strong>and</strong>ing. Besides, as mentioned above in the<br />

introduction to Chapter 3, it seemed that experimental compaction localization research,<br />

which had already become a hot area, was headed in a direction at odds with observable<br />

natural reality. In forging ahead despite my committee’s accurate admonitions, I did<br />

manage to glean valuable “negative result” insights into the balance <strong>of</strong> material <strong>and</strong><br />

loading conditions conducive to CB formation. Specifically, I suggest that natural<br />

compaction b<strong>and</strong>s, such as observed in the Aztec, form in unconsolidated to barely<br />

lithified porous s<strong>and</strong>s subjected to tectonic <strong>and</strong> burial stress states typical <strong>of</strong> the shallow<br />

crust. Despite using weakly lithified Aztec specimens in my experiments, which I seeded<br />

with stress-concentrating flaws <strong>and</strong> loaded uniaxially over a range <strong>of</strong> geologically<br />

plausible confining pressures, failure invariably occurred by dilational shear. On the other<br />

h<strong>and</strong>, experiments that force compaction to occur in lithified s<strong>and</strong>stone at mid-crustal<br />

6


confining pressures produce, in my opinion, a failure phenomenon distinct from<br />

compaction b<strong>and</strong>ing as observed in outcrop.<br />

The greater value <strong>of</strong> my fledgling laboratory efforts probably lies in the collaborative<br />

relationships developed within the experimental community pursuing compaction<br />

localization research. I visited <strong>and</strong> worked in the labs <strong>of</strong> Bezalel Haimson (University <strong>of</strong><br />

Wisconsin-Madison), William Olsson <strong>and</strong> David Holcomb (S<strong>and</strong>ia National Labs) <strong>and</strong><br />

David Lockner (USGS-Menlo Park), <strong>and</strong> corresponded extensively with the laboratory <strong>of</strong><br />

Teng-fong Wong (Stony Brook University). Through these interactions <strong>and</strong> a DOE-<br />

sponsored workshop (October 2004), which led directly to my association with co-author<br />

John Rudnicki (Chapters 2 <strong>and</strong> 3), David Pollard <strong>and</strong> I have introduced a much-needed<br />

field-based reality check into compaction localization experimentation. Whether this<br />

influence will become more tangible through my own future experimental efforts remains<br />

to be seen.<br />

Finally, I would be remiss in not observing that, although a strong process-based<br />

<strong>mechanics</strong> perspective <strong>and</strong> approach pervade this thesis, it is based primarily on<br />

observations made from outcrops <strong>and</strong> samples <strong>of</strong> the Aztec s<strong>and</strong>stone. While this fact<br />

does not detract from my results <strong>and</strong> interpretations, the issue <strong>of</strong> ubiquity is relevant from<br />

a practical applications st<strong>and</strong>point. That is, are CB arrays such as observed in the Aztec<br />

common in s<strong>and</strong>stone aquifers <strong>and</strong> reservoirs, or are they a relative freak <strong>of</strong> nature? From<br />

a fundamental physical <strong>and</strong> mechanical point <strong>of</strong> view, it is unlikely that CBs in the Aztec<br />

represent an isolated phenomenological fluke. However, shear deformation b<strong>and</strong>s have<br />

been observed in a wide variety <strong>of</strong> clastic deposits exposed around the world, <strong>and</strong> similar<br />

observations <strong>of</strong> CBs are needed to confirm them as a common structure in porous<br />

s<strong>and</strong>stone.<br />

Beyond this obvious direction for future analog reservoir-based fieldwork, a<br />

concerted effort also is needed to develop borehole geophysical methods for detecting<br />

<strong>and</strong> characterizing deformation b<strong>and</strong>s <strong>of</strong> all types in the subsurface. The general<br />

petrophysical attributes <strong>of</strong> b<strong>and</strong>s should, in theory at least, render them visible to a<br />

variety <strong>of</strong> imaging technologies, ranging from acoustic to electromagnetic. In conjunction<br />

with a solid mechanical underst<strong>and</strong>ing <strong>of</strong> how CBs <strong>and</strong> CB arrays form—the foundations<br />

for which are laid here in Chapters 1 through 4—such imaging techniques would provide<br />

7


the hard data from which accurate projections <strong>of</strong> b<strong>and</strong> configurations within subsurface<br />

s<strong>and</strong>stones could be made. With reliable configuration projections <strong>and</strong> flow-property<br />

estimations (Chapter 5) in h<strong>and</strong>, the fluid flow impacts <strong>of</strong> any CB array could then<br />

accurately be modeled <strong>and</strong> generalized for the purposes <strong>of</strong> aquifer/reservoir-scale<br />

simulation—as demonstrated in Chapters 6 <strong>and</strong> 7.<br />

8


Chapter 1<br />

Structural <strong>geology</strong> <strong>and</strong> tectonic interpretation <strong>of</strong> compaction b<strong>and</strong>s<br />

in the Aztec s<strong>and</strong>stone <strong>of</strong> southeastern Nevada<br />

1. Abstract<br />

The Aztec s<strong>and</strong>stone in the Valley <strong>of</strong> Fire <strong>of</strong> southeastern Nevada comprises an<br />

exhumed analog for active æolian aquifers <strong>and</strong> reservoirs subjected to tectonic<br />

compression <strong>and</strong> shortening. A planar fabric <strong>of</strong> anastomosing compaction b<strong>and</strong>s (CBs)<br />

pervades the upper 600 m <strong>of</strong> the 1,400-m-thick Aztec. Individual CBs are observed to be<br />

thin, tabular localizations <strong>of</strong> uniaxial porosity-loss compaction that can be considered<br />

kinematically <strong>and</strong> mechanically as anticracks. In outcrop, CBs form sometimes complex<br />

patterns that can, nonetheless, be generally understood as resulting from mechanical<br />

interactions between anticracks. Compaction b<strong>and</strong>s are the oldest structures in the Aztec,<br />

<strong>and</strong> abundant geological evidence indicates that they formed during the earliest period <strong>of</strong><br />

regional compression associated with Cretaceous Sevier tectonism in a weakly cemented<br />

material buried less than a few hundred meters.<br />

Viewed at larger scales (> 100 m), a dominant north-south trend <strong>and</strong> steep east dip are<br />

apparent, as well as localized pockets <strong>of</strong> a coexisting, coeval CB set trending generally<br />

northeast <strong>and</strong> dipping steeply northwest. The mean trend <strong>of</strong> the dominant CB set is<br />

generally orthogonal to the east-vergent direction <strong>of</strong> Sevier thrusting, while the secondary<br />

set is, on average, orthogonal to the dominant set. Based on this geometry, <strong>and</strong> our thin-<br />

section to outcrop-based underst<strong>and</strong>ing <strong>of</strong> CBs as anticracks, we interpret the dominant<br />

set as having formed orthogonal to the maximum compressive paleostress (σ1), with the<br />

secondary set orthogonal to σ2. This completely constrains the orientations <strong>and</strong> relative<br />

magnitudes <strong>of</strong> the principal paleostresses during CB formation—σ1 → 270°/38°, σ2 →<br />

138°/39°, σ3 → 24°/27°—which, however, depart radically from the classic view <strong>of</strong> σ3 as<br />

subvertical in thrust-fault environments. We suggest that a diffuse stress perturbation<br />

caused by as yet unrecognized regional structures—e.g. a detachment beneath the Valley<br />

<strong>of</strong> Fire along which it moved relatively westward—could explain the inferred state <strong>of</strong><br />

paleostress.<br />

9


Although the Aztec s<strong>and</strong>stone experienced less than 1% total shortening as the result<br />

<strong>of</strong> CB formation, the low-porosity, low-permeability b<strong>and</strong> fabrics developed have been<br />

shown capable <strong>of</strong> exerting pr<strong>of</strong>ound fluid flow effects at scales <strong>of</strong> practical interest. The<br />

analysis presented here <strong>of</strong> CBs as indicators <strong>of</strong> paleostress <strong>and</strong> tectonic structure in this<br />

exhumed analog aquifer/reservoir also applies in reverse. That is, given knowledge <strong>of</strong> the<br />

stress <strong>and</strong> material history <strong>of</strong> a subsurface s<strong>and</strong>stone, predictions regarding the possible<br />

presence <strong>and</strong> geometry <strong>of</strong> CBs can be made.<br />

2. Introduction<br />

Compaction b<strong>and</strong>ing as a recognized mode <strong>of</strong> localized brittle failure in porous<br />

s<strong>and</strong>stone has a brief history, having first been described <strong>and</strong> named by Mollema <strong>and</strong><br />

Antonellini (1996) based on their work in the Navajo s<strong>and</strong>stone <strong>of</strong> south central Utah.<br />

Compaction b<strong>and</strong>s (CBs) represent one kinematic end member <strong>of</strong> the suite <strong>of</strong> structures<br />

known collectively as deformation b<strong>and</strong>s, which includes shear <strong>and</strong> dilation b<strong>and</strong>s (Du<br />

Bernard et al., 2002; Borja <strong>and</strong> Aydin, 2004). Mollema <strong>and</strong> Antonellini (1996) described<br />

CBs as long (meters), thin (centimeters) tabular zones <strong>of</strong> mechanical porosity loss formed<br />

within the compressional quadrants <strong>of</strong> shear b<strong>and</strong> style faults (Aydin, 1978), interpreting<br />

them as “a <strong>structural</strong> analog for anti-mode I cracks in æolian s<strong>and</strong>stone...that<br />

accommodate pure compaction,” following the conceptual anticrack model for pressure<br />

solution surfaces <strong>of</strong> Fletcher <strong>and</strong> Pollard (1981).<br />

The recognition <strong>of</strong> CBs, <strong>and</strong> their obvious potential to act as low-porosity, low-<br />

permeability impediments to fluid flow in otherwise highly transmissive s<strong>and</strong>stones, has<br />

excited an ongoing flurry <strong>of</strong> work both experimental (e.g. Olsson <strong>and</strong> Holcomb, 2000;<br />

Klein et al., 2001; Vajdova <strong>and</strong> Wong, 2003; Haimson, 2003; Baud et al., 2004; Tembe et<br />

al., 2006) <strong>and</strong> theoretical (e.g. Olsson, 1999; Issen <strong>and</strong> Rudnicki, 2000, 2001; Detournay<br />

et al., 2003; Bésuelle <strong>and</strong> Rudnicki, 2004; Rudnicki, 2002, 2003, 2004; Katsman et al.,<br />

2005; Rudnicki <strong>and</strong> Sternl<strong>of</strong>, 2005; Sternl<strong>of</strong> et al., 2005; Katsman <strong>and</strong> Aharonov, 2006)<br />

all <strong>of</strong> it aimed at comprehending the <strong>mechanics</strong> <strong>of</strong> compaction localization in porous,<br />

granular materials. To date, however, the only other field-based study published on CB<br />

occurrence <strong>and</strong> <strong>geology</strong> also focused on an æolian Jurassic s<strong>and</strong>stone <strong>of</strong> the southwestern<br />

U.S.—the Navajo-equivalent Aztec s<strong>and</strong>stone <strong>of</strong> southeastern Nevada (Sternl<strong>of</strong> et al.,<br />

2005). That paper presents an anticrack-inclusion mechanical model for CBs that is based<br />

10


on detailed outcrop <strong>and</strong> thin-section observations <strong>of</strong> the Aztec as extensively exposed in<br />

the Valley <strong>of</strong> Fire State Park (Figure 1.1).<br />

Stepping back from the wide variety <strong>of</strong> patterns visible in individual outcrops, the two<br />

most striking aspects <strong>of</strong> CBs in the Aztec as a whole are their ubiquity <strong>and</strong> the general<br />

consistency <strong>of</strong> their planar orientation: north-northwest trending, steeply east dipping.<br />

Considering the Aztec as an exhumed analog for active s<strong>and</strong>stone aquifers <strong>and</strong> reservoirs,<br />

the potential for such an anastomosing CB fabric to restrict <strong>and</strong> channel fluid flow is<br />

substantial (Sternl<strong>of</strong> et al., 2004; Sternl<strong>of</strong> et al., 2006). The practical goal behind all CB<br />

investigations is to develop the capability to predict, or at least forecast their occurrence<br />

in the subsurface such that their hydraulic effects can be mitigated. Achieving this goal<br />

will require underst<strong>and</strong>ing how, in what kinds <strong>of</strong> materials <strong>and</strong> under what loading<br />

conditions CBs form. As a contribution to this endeavor, we present a broad look at the<br />

thin-section to regional scale <strong>structural</strong> <strong>geology</strong> <strong>of</strong> CBs in the Aztec s<strong>and</strong>stone at the<br />

Valley <strong>of</strong> Fire, <strong>and</strong> consider their implications for the tectonic framework <strong>and</strong> history <strong>of</strong><br />

southern Nevada.<br />

3. Setting <strong>and</strong> history <strong>of</strong> study area<br />

The following synopsis <strong>of</strong> the geological setting <strong>and</strong> history <strong>of</strong> the Aztec s<strong>and</strong>stone<br />

was synthesized from data <strong>and</strong> interpretations published in a wide variety <strong>of</strong> sources over<br />

the past 85 years as cited below. Given that the methods <strong>and</strong> perspectives <strong>of</strong> <strong>structural</strong><br />

<strong>geology</strong> have evolved a great deal since the first research on the area was published<br />

during the Wilson administration (Longwell, 1921), long before any modern notion <strong>of</strong><br />

tectonics, we emphasize here more recent work <strong>and</strong> interpretations.<br />

3.1. Deposition<br />

The Aztec s<strong>and</strong>stone was deposited during latest Triassic to middle Jurassic time in a<br />

back-arc basin associated with uplift <strong>of</strong> the ancestral Sierra Nevadan magmatic arc<br />

(Marzolf, 1986; Burchfiel et al., 1992). It comprises the southwestern edge <strong>of</strong> a much<br />

larger æolian erg system, which includes the Navajo <strong>and</strong> Nugget s<strong>and</strong>stones <strong>and</strong><br />

blanketed much <strong>of</strong> the intermountain southwest (Blakey, 1989; Marzolf, 1986).<br />

Prevailing winds during deposition <strong>of</strong> the Aztec are interpreted to have blown out <strong>of</strong> the<br />

northeast <strong>and</strong> over the shallow sea occupying what is now the continental interior<br />

(Marzolf, 1982, 1986).<br />

11


Waterpocket Fault<br />

36<br />

o<br />

26‘<br />

N<br />

0 1<br />

km<br />

114<br />

o<br />

32‘<br />

Willow Tank<br />

Thrust<br />

Park Road<br />

Overton Syncline<br />

Lower Aztec<br />

12<br />

Oregon<br />

Washingto n<br />

California<br />

30<br />

Nevada<br />

Canada<br />

Las Vegas<br />

Idaho<br />

Air photo detail<br />

Valley <strong>of</strong> Fire<br />

Cretaceous units<br />

25<br />

Upper Aztec<br />

Park Office<br />

25<br />

Montana<br />

Utah<br />

Wyoming<br />

Arizona<br />

Baseline Fault<br />

Route 169


Figure 1.1 (opposite page) Location <strong>and</strong> general <strong>geology</strong> <strong>of</strong> the study area in the Aztec<br />

s<strong>and</strong>stone at the Valley <strong>of</strong> Fire State Park, southeastern Nevada. The Aztec crops out as a<br />

1,400-m-thick sequence <strong>of</strong> gently folded, northeasterly dipping (~25°) æolian deposits<br />

consisting <strong>of</strong> ~800 m <strong>of</strong> red-stained s<strong>and</strong>stone (lower Aztec) <strong>and</strong> ~600 m <strong>of</strong> bleached <strong>and</strong><br />

variegated s<strong>and</strong>stone (upper Aztec). The Aztec is immediately, though unconformably<br />

overlain by up to 1,300 m <strong>of</strong> Cretaceous deposits (principally the quartz-rich Baseline<br />

s<strong>and</strong>stone in the air photo). The Aztec has been extensively affected by Basin <strong>and</strong> Range<br />

extension, principally by north-northeasterly trending left-lateral strike slip-faults <strong>and</strong><br />

subsidiary northwesterly trending right-lateral strike-slip faults (apparent in the many<br />

<strong>of</strong>fsets <strong>of</strong> the red/buff alteration front). The major left-lateral bounding faults for the<br />

immediate study area (Waterpocket fault <strong>and</strong> Baseline fault) are indicated (after<br />

Bohannon 1983), as are the Overton Syncline (after Carpenter <strong>and</strong> Carpenter, 1994) <strong>and</strong><br />

two representative bedding orientations in the lower Cretaceous units (after Bohannon,<br />

1983). Less obvious are remnants <strong>of</strong> the east-vergent Willow Tank thrust (upper right<br />

corner, teeth denote hanging wall), marking the easternmost extent <strong>of</strong> compressional<br />

tectonism related to the Cretaceous Sevier ororogeny. The aerial photo base is a mosaic<br />

constructed <strong>of</strong> images downloaded from Google Earth.<br />

Within the general vicinity <strong>of</strong> the Valley <strong>of</strong> Fire (Figure 1.1), the 1,400-m-thick Aztec<br />

unconformably overlies terrigenous clastic, evaporitic <strong>and</strong> limestone deposits <strong>of</strong> the<br />

Triassic <strong>and</strong> Permian, which themselves unconformably overlie miogeoclinal carbonates<br />

<strong>of</strong> Pennsylvanian age <strong>and</strong> older. The Aztec in turn is unconformably overlain by up to<br />

1,300 m <strong>of</strong> upward-coarsening Cretaceous clastics <strong>and</strong>, following a depositional hiatus<br />

through the Paleogene, another 300 m <strong>of</strong> Neogene limestones <strong>and</strong> clastics (Bohannon,<br />

1983; Bohannon et al., 1993; Eichhubl et al., 2004). Figure 1.2 presents a generalized<br />

stratigraphic section <strong>of</strong> the region from latest Paleozoic time through the present day, as<br />

modified from Bohannon (1977).<br />

The Aztec s<strong>and</strong>stone itself is a subarkose containing up to 8% orthoclase <strong>and</strong> minor<br />

amounts <strong>of</strong> lithic fragments (Eichhubl et al., 2004). It is composed <strong>of</strong> moderately rounded,<br />

fine to medium-sized detrital grains (0.1 to 0.5 mm, 0.25 mm average) arranged in a<br />

large-scale wedge <strong>and</strong> tabular-planar cross stratified sedimentary architecture<br />

characteristic <strong>of</strong> æolian deposition (Marzolf, 1983). The Aztec is generally quite friable,<br />

with porosity averaging about 20-25% (Sternl<strong>of</strong> et al., 2005) <strong>and</strong> permeability ranging<br />

from about 100 to 2,500 mD (Flodin et al., 2005; Chapter 5 this thesis). Clay, chiefly<br />

kaolinite, comprises up to 6% <strong>of</strong> the rock by volume <strong>and</strong> forms the predominant grain-<br />

bridging, pore-filling cement.<br />

13


Cenozoic<br />

Mesozoic<br />

Paleozoic<br />

Quaternary<br />

Permian Triassic Triassic?<br />

Jurassic Cretaceous Tertiary<br />

Miocene &<br />

Pliocene<br />

Miocene<br />

Thickness (m)<br />

5,400<br />

4,800<br />

4,200<br />

3,600<br />

3,000<br />

2,400<br />

1,800<br />

1,200<br />

600<br />

0<br />

Quaternary Deposits<br />

Muddy Creek Formation<br />

Horse Spring Formation<br />

Baseline S<strong>and</strong>stone<br />

Willow Tank Formatio n<br />

Aztec S<strong>and</strong>stone<br />

Moenave Formation<br />

Chinle Formation<br />

Moenkopi Formation<br />

Kaibab Limestone<br />

Toroweap Formatio n<br />

Hermit Shale<br />

Lithologic Descriptions<br />

S<strong>and</strong>stone<br />

Shale<br />

Siltstone<br />

Calcareous<br />

shale<br />

Limestone<br />

Dolomite<br />

Magnesite<br />

Conglomerate<br />

Breccia<br />

Evaporites<br />

(gypsum)<br />

Calcareous<br />

conglomerate<br />

Figure 1.2. Generalized stratigraphic column for the Valley <strong>of</strong> Fire area from the<br />

Permian Hermit shale through Quaternary alluvium (adapted from Bohannon, 1977).<br />

14<br />

Chert


Abundant deeply etched <strong>and</strong>/or kaolin-replaced orthoclase grains indicate that much <strong>of</strong><br />

the clay present is internally derived (Eichhubl et al., 2004; Sternl<strong>of</strong> et al., 2005). Small<br />

amounts <strong>of</strong> inter-layered smectite <strong>and</strong> illite are also present (Eichhubl et al., 2004).<br />

Abundant mutual indentation <strong>of</strong> quartz grains via pressure solution has been reported<br />

(Flodin et al., 2003; Eichhubl et al., 2004), although much <strong>of</strong> this texture may in fact be<br />

due to mechanical indentation accommodated by pervasive intra-grain micr<strong>of</strong>racturing<br />

(Sternl<strong>of</strong> et al., 2005).<br />

3.2. Diagenesis<br />

Syndepositional precipitation <strong>of</strong> hematite grain coats resulted in the Aztec being<br />

stained a uniform red color during initial burial (Eichhubl et al., 2004). The color from<br />

this earliest diagenetic alteration still dominates the lower half <strong>of</strong> the s<strong>and</strong>stone. The<br />

upper Aztec, however, has been subjected to at least two stages <strong>of</strong> bleaching <strong>and</strong> iron<br />

oxide redistribution apparently associated with upward <strong>and</strong> eastward expulsion <strong>of</strong><br />

reducing basinal brines from beneath Sevier-related thrust sheets advancing from the<br />

west (Eichhubl et al., 2004). These fluid-flow events are responsible for the vividly<br />

colorful patterns <strong>of</strong> iron mineralization from which the Valley <strong>of</strong> Fire derives its name<br />

(Figure 1.1).<br />

Despite its long alteration history, however, the presence <strong>of</strong> only limited quartz<br />

overgrowth cementation <strong>and</strong> diagenetic illite, both restricted to the very bottom <strong>of</strong> the<br />

pile (<strong>and</strong> within the damage zones <strong>of</strong> some major faults), indicate that the base <strong>of</strong> the<br />

Aztec in the Valley <strong>of</strong> Fire has likely never been buried deeper than about 3 km (~ 80° C<br />

for a normal geothermal gradient). Interestingly, this equals the sum <strong>of</strong> its own thickness<br />

<strong>and</strong> <strong>of</strong> all overlying sedimentary deposits through (Eichhubl et al., 2004; Sternl<strong>of</strong> et al.,<br />

2005). In any case, the Aztec is today at best moderately well cemented toward the<br />

bottom, where it is still red, <strong>and</strong> poorly cemented toward the middle <strong>and</strong> top (Flodin et al.,<br />

2003), where it has been extensively bleached <strong>and</strong> now exhibits an unconfined uniaxial<br />

compressive strength <strong>of</strong> only 2-3 MPa (Haimson <strong>and</strong> Lee, 2004). Neither is there<br />

evidence to suggest that the state <strong>of</strong> lithification anywhere in the Aztec was ever<br />

appreciably greater than it is today. Our own limited triaxial testing, however, does<br />

indicate that the resistance <strong>of</strong> present-day Aztec to uniform compaction with increasing<br />

pressure approaches that <strong>of</strong> quartzite.<br />

15


3.3. Deformation<br />

The Aztec also has been subjected to a varied <strong>and</strong> punctuated history <strong>of</strong> deformation<br />

driven by tectonic activity, starting with late Mesozoic compression <strong>of</strong> the Sevier<br />

Orogeny (e.g. Armstrong, 1968; Fleck, 1970; Brock <strong>and</strong> Engelder, 1977; Bohannon,<br />

1983) <strong>and</strong> jumping to Basin <strong>and</strong> Range extension beginning in Miocene time <strong>and</strong><br />

continuing to the present (e.g. Bohannon, 1983; Bohannon et al., 1993; Campagna <strong>and</strong><br />

Aydin, 1994; Flodin <strong>and</strong> Aydin, 2004; Myers <strong>and</strong> Aydin, 2004)). Figure 1.3 presents a<br />

generalized structure <strong>and</strong> tectonic map for the region, as compiled from multiple sources.<br />

Thin-skinned, east-vergent thrust faulting <strong>of</strong> the Sevier Orogeny may have initiated<br />

along the flanks <strong>of</strong> the exp<strong>and</strong>ing magmatic arc to the west <strong>of</strong> the study area during<br />

deposition <strong>of</strong> the Aztec, <strong>and</strong> progressively encroached on the Valley <strong>of</strong> Fire into latest<br />

Cretaceous time (Burchfiel et al., 1992). In general, Paleozoic carbonate rocks were<br />

placed atop Mesozoic clastic rocks along a piggy-backed sequence <strong>of</strong> low angle thrust<br />

ramps, each one accommodating up to tens <strong>of</strong> kilometers <strong>of</strong> shortening (Armstrong,<br />

1968). The easternmost extent <strong>of</strong> Sevier tectonism, at least ins<strong>of</strong>ar as expressed by<br />

exposed thrusts, reached just up to the Valley <strong>of</strong> Fire (Figure 1.3), first with the thin,<br />

short-traveled (kilometers) Summit-Willow Tank thrust, which placed lower Aztec atop<br />

upper Aztec <strong>and</strong> as much as 600 m <strong>of</strong> Cretaceous deposits, then with the kilometers thick,<br />

far-traveled (tens <strong>of</strong> kilometers) Muddy Mountain thrust, which placed Paleozoic<br />

carbonates over Aztec riding atop the Summit-Willow Tank thrust (Armstrong, 1968;<br />

Bohannon, 1983; Brock <strong>and</strong> Engelder, 1979; Carpenter <strong>and</strong> Carpenter, 1994; Longwell,<br />

1949). The preponderance <strong>of</strong> stratigraphic, <strong>structural</strong> <strong>and</strong> diagenetic evidence indicates<br />

that the Aztec s<strong>and</strong>stone now exposed in the main tourist part <strong>of</strong> the Valley <strong>of</strong> Fire State<br />

Park (the primary study area) was never buried by either thrust sheet, only by the upward-<br />

coarsening sequence <strong>of</strong> Cretaceous clastic deposits shed from them (Bohannon, 1984;<br />

Taylor, 1999; Eichhubl et al., 2004; Sternl<strong>of</strong> et al., 2005), which consist largely <strong>of</strong><br />

reworked Aztec s<strong>and</strong>stone (Bohannon, 1983).<br />

Structures formed in the Aztec s<strong>and</strong>stone that are associated with Sevier tectonism<br />

include the north-northwest trending, steeply east-dipping CBs that are the subject <strong>of</strong> this<br />

paper, <strong>and</strong> low-angle shear b<strong>and</strong>s that tend to parallel depositional bedding <strong>and</strong> exhibit a<br />

top-to-the-east sense <strong>of</strong> shear which mimics the large-scale thrust faults.<br />

16


Clark County Nevada<br />

LVVSZ<br />

40 km<br />

N<br />

Sevier Thrust Front<br />

Las<br />

Vegas<br />

N<br />

20 km<br />

LVVSZ<br />

114 W<br />

36<br />

o<br />

N<br />

o<br />

AHF<br />

MMT<br />

Study Area<br />

WTT<br />

LMFS<br />

CWF<br />

AHF<br />

X<br />

114<br />

o<br />

30’ N<br />

STUDY<br />

AREA<br />

Mead<br />

Lake<br />

LMFS<br />

Figure 1.3. Schematized fault map for the Lake Mead region <strong>of</strong> southeastern Nevada.<br />

The main map shows major strike-slip <strong>and</strong> normal faults associated with Basin <strong>and</strong><br />

Range extension (LVVSZ = Las Vegas Valley Shear Zone; LMFS = Lake Mead Fault<br />

System; AHF = Arrowhead Fault; CWF = California Wash Fault). Study area box<br />

indicates approximate location <strong>of</strong> Figure 1.1, while the X (circled in white) marks the<br />

location <strong>of</strong> measurements made in Aztec s<strong>and</strong>stone exposed at Buffington Pockets (see<br />

Section 4.3). The inset map indicates the approximate southeastern most extent <strong>of</strong><br />

Cretaceous Sevier thrusting, as dissected by the LVVSZ, LMFS <strong>and</strong> AHF. The study area<br />

just avoided being overridden (MMT = Muddy Mountain Thrust; WTT = Willow Tank<br />

Thrust [also the Summit Thrust]; thrust faults south <strong>of</strong> the LVVSZ are undifferentiated).<br />

The color base image was downloaded from Google Earth. The <strong>structural</strong> data was<br />

compiled from Stewart <strong>and</strong> Carlson (1978); Bohannon (1983b); Campagna <strong>and</strong> Aydin<br />

(1994) <strong>and</strong> Flodin (2003).<br />

17<br />

36<br />

o<br />

30’ N


There also are relatively high-angle tabular zones composed <strong>of</strong> multiple shear b<strong>and</strong>s<br />

(Flodin <strong>and</strong> Aydin, 2004). Abundant cross-cutting relationships consistently indicate that<br />

shear b<strong>and</strong>ing represents a distinct phase <strong>of</strong> deformation subsequent to CB formation,<br />

leaving CBs as the oldest <strong>structural</strong> fabric present (Hill, 1989; Taylor <strong>and</strong> Pollard, 2000;<br />

Myers <strong>and</strong> Aydin, 2004; Flodin <strong>and</strong> Aydin, 2004; Eichhubl et al., 2004; Sternl<strong>of</strong> et al.,<br />

2005). The highest density <strong>of</strong> low-angle shear b<strong>and</strong>ing occurs in close proximity to the<br />

Summit-Willow Tank thrust <strong>and</strong> extends into the overlying Cretaceous units exposed<br />

beneath it, suggesting a direct genetic relationship with emplacement <strong>of</strong> the thrust sheet<br />

(Hill, 1989; Eichhubl et al., 2004; Sternl<strong>of</strong> et al., 2005). Compaction b<strong>and</strong>s, on the other<br />

h<strong>and</strong>, are restricted to the Aztec s<strong>and</strong>stone itself (Sternl<strong>of</strong> et al., 2005).<br />

It appears that the Valley <strong>of</strong> Fire region was not affected by compression associated<br />

with the early Tertiary Laramide orogeny (Bohannon, 1983), although it is possible that a<br />

distributed tectonic fabric <strong>of</strong> joints may have formed during this time, when the study<br />

area is interpreted to have been situated on the northeastern flank <strong>of</strong> a broad, gently<br />

north-plunging arch (Bohannon, 1984). In any case, following some 40 million years <strong>of</strong><br />

relative tectonic quiescence, the next major event to impact the area was Basin <strong>and</strong> Range<br />

extension beginning 15 to 20 Ma. This largely expressed itself as nested systems <strong>of</strong> both<br />

left <strong>and</strong> right lateral strike-slip faults (generally with some dip slip) <strong>and</strong> associated<br />

normal faults (Bohannon 1983, 1984; Burchfiel et al., 1992; Duebendorfer et al., 1998).<br />

The major regional-scale features are the northeast-trending str<strong>and</strong>s <strong>of</strong> the left-lateral<br />

Lake Mead fault zone <strong>and</strong> the northwest-trending str<strong>and</strong>s <strong>of</strong> the right-lateral Las Vegas<br />

Valley shear zone (Figure 1.3). Associated structures developed within the Aztec<br />

s<strong>and</strong>stone at the Valley <strong>of</strong> Fire include joints, sheared joints <strong>and</strong> a hierarchical network <strong>of</strong><br />

left-lateral <strong>and</strong> right-lateral joint-based strike-slip faults (Taylor et al., 1999; Myers <strong>and</strong><br />

Aydin, 2004; Flodin <strong>and</strong> Aydin, 2004). These Basin <strong>and</strong> Range related structures dissect<br />

<strong>and</strong> to some degree obscure the CBs <strong>of</strong> interest, but the effects are generally local <strong>and</strong> not<br />

difficult to account for.<br />

Finally, the Aztec s<strong>and</strong>stone <strong>and</strong> overlying Cretaceous deposits in the central Valley<br />

<strong>of</strong> Fire study area exhibit a very gentle, northeast-plunging synclinal fold (Figure 1.1)<br />

referred to as the Overton syncline by Carpenter <strong>and</strong> Carpenter (1994) <strong>and</strong> attributed to<br />

drag along the east-west trending, left-lateral oblique-slip Arrowhead fault, which bounds<br />

18


the south side <strong>of</strong> the valley (Figure 1.3). The fold axis currently plunges 20-30° to the<br />

northeast, matching the general dip <strong>of</strong> the surrounding Cretaceous strata (Bohannon,<br />

1983; Flodin, 2003; Eichhubl et al., 2004). This indicates that the folding largely predates<br />

regional tilting.<br />

4. Structural <strong>geology</strong> <strong>of</strong> compaction b<strong>and</strong>s<br />

Extensive, anastomosing arrays <strong>of</strong> subparallel CBs pervade the upper 600 m <strong>of</strong> the<br />

Aztec s<strong>and</strong>stone with a dominant north-northwest trend <strong>and</strong> steep (~70°) dip to the east.<br />

There also are common, but generally isolated outcrops exhibiting more complicated<br />

patterns <strong>of</strong> b<strong>and</strong>s, including multiple sets <strong>of</strong> CBs <strong>and</strong> abundant shear b<strong>and</strong>s dissecting<br />

them (Sternl<strong>of</strong> et al., 2004). Phenomenal exposure <strong>of</strong> the Aztec throughout the main part<br />

<strong>of</strong> the Valley <strong>of</strong> Fire State Park permits an integrated view <strong>of</strong> the <strong>structural</strong> <strong>geology</strong> <strong>of</strong><br />

CBs from thin-section to regional scale (Figure 1.4).<br />

4.1. Thin-section to outcrop scale<br />

Viewed in thin section (Figure 1.4b), the boundaries <strong>of</strong> tabular CBs are clearly<br />

delineated as abrupt drops in porosity directly attributable to inelastic mechanical<br />

compaction associated with grain crushing—specifically, micro-fracture accommodated<br />

plasticity <strong>of</strong> the quartz grains (Sternl<strong>of</strong> et al., 2005) apparent under backscatter electron<br />

imaging (Figure 1.4c <strong>and</strong> d). The total volume change realized, as measured by the<br />

relative volume fraction occupied by detrital grains inside versus outside the b<strong>and</strong>, is<br />

about 10% <strong>and</strong> appears to be consistent along the b<strong>and</strong>s from tip to middle. Judging from<br />

the overall continuity <strong>of</strong> depositional bedding cutting across the b<strong>and</strong>s (Figure 1.4b), <strong>and</strong><br />

the almost complete absence <strong>of</strong> granular disaggregation within them despite intense grain<br />

fracturing, the volume change appears to have resulted from uniaxial compaction directed<br />

normal to the plane <strong>of</strong> the b<strong>and</strong> in the absence <strong>of</strong> appreciable shear accommodated across<br />

it (Sternl<strong>of</strong> et al., 2005).<br />

In the otherwise weakly lithified Aztec s<strong>and</strong>stone, CBs tend to weather out in positive<br />

relief as distinctive tabular fins, rendering them readily visible in outcrop <strong>and</strong><br />

accentuating the absence <strong>of</strong> appreciable shear accommodated (Figure 1.4e). The relative<br />

resistance <strong>of</strong> CBs is due primarily to the preferential precipitation <strong>of</strong> kaolinite as a pore-<br />

filling, grain-bridging cement (Sternl<strong>of</strong> et al., 2005). This post-compaction clay<br />

accumulation further exacerbates porosity loss <strong>and</strong> consequent permeability<br />

19


(e) (f)<br />

(c)<br />

500µm<br />

compaction b<strong>and</strong> trend<br />

depositional bedding<br />

N<br />

0 1<br />

km<br />

(a)<br />

(b)<br />

~ 10 mm<br />

compaction b<strong>and</strong><br />

20<br />

compaction b<strong>and</strong> trend<br />

500µm<br />

(g)<br />

(d)


Figure 1.4. (opposite page) Photo collage depicting compaction b<strong>and</strong>s in the Aztec<br />

s<strong>and</strong>stone. (a) Study area air photo from Figure 1.1 with arrows indicating approximate<br />

locations where other photos were taken. (b) Photomicrograph mosaic <strong>of</strong> a compaction<br />

b<strong>and</strong> at high angle to depositional bedding, which passes through the b<strong>and</strong> with no<br />

apparent shear <strong>of</strong>fset, change in direction or change in thickness (white grains are quartz,<br />

dark grains are mostly stained orthoclase <strong>and</strong> some hematite, blue is epoxy-filled pore<br />

space. (c) Backscatter electron image (BEI)from just outside the compaction b<strong>and</strong> (light<br />

gray is quartz, <strong>of</strong>f-white is orthoclase, dark gray is kaolinite, stark white is hematite,<br />

black is pore space). (d) BEI from inside the compaction b<strong>and</strong>. Note the abrupt drop in<br />

porosity, pore size <strong>and</strong> pore connectivity accommodated by intense damage to some<br />

quartz grains. This damage, <strong>and</strong> attendant grain interpenetration, is predominantly due to<br />

pervasive micr<strong>of</strong>racturing. (e) Close-up view <strong>of</strong> compaction b<strong>and</strong> fin capturing its<br />

essential tabular-planar aspect, cm thickness <strong>and</strong> lack <strong>of</strong> obvious shear displacement. (f)<br />

Typical view <strong>of</strong> subparallel, anastomosing compaction b<strong>and</strong> array (resistant fins) in<br />

outcrop. View is north-northwest along the dominant trend). (g) Cross-hatch pattern<br />

formed by two distinct b<strong>and</strong> orientations at high angle to each other. B<strong>and</strong>s belonging to<br />

the dominant trend set (view is northward) pass through both the upper <strong>and</strong> lower crossbed<br />

boundaries, while the secondary set <strong>of</strong> b<strong>and</strong>s is constrained within the middle<br />

bedding package.<br />

reduction—from ~25% <strong>and</strong> ~1,500mD in the s<strong>and</strong>stone to ~10% <strong>and</strong> 1.5 mD in the b<strong>and</strong>s<br />

(Sternl<strong>of</strong> et al., 2004; Sternl<strong>of</strong> et al., 2006; Chapter 5 this thesis).<br />

In outcrop, individual CBs appear to be grossly penny-shaped <strong>and</strong> elliptical in pr<strong>of</strong>ile,<br />

ranging in maximum thickness up to a few cm (averaging ~ 1 cm) <strong>and</strong> in length to more<br />

than 100 m. They <strong>of</strong>ten exhibit well-defined, elliptically tapered tips, <strong>and</strong> a variety <strong>of</strong><br />

meter-scale patterns <strong>of</strong> near-tip interactions that are diagnostic <strong>of</strong> in-plane <strong>propagation</strong> as<br />

anticracks (Sternl<strong>of</strong> et al., 2005; Chapter 4 this thesis). By far the dominant pattern at the<br />

outcrop-scale, however, is an anastomosing planar fabric trending north-northwest <strong>and</strong><br />

dipping steeply eastward (Figure 1.4f). Spacing between adjacent b<strong>and</strong>s within this fabric<br />

can vary from 1 mm to several meters, averaging ~0.7 m (based on 286 m <strong>of</strong> orthogonal<br />

scan lines intersecting 423 b<strong>and</strong>s at 23 locations—see Section 5 below). Also distinctly<br />

notable, although relatively rare <strong>and</strong> spatially isolated, are patterns <strong>of</strong> CBs that cross <strong>and</strong><br />

sometimes turn into each other at high angle (Figure 1.4g). These “cross-hatch” patterns<br />

(Sternl<strong>of</strong> et al., 2004) are comprised <strong>of</strong> through-going b<strong>and</strong>s belonging to the dominant<br />

orientation set, <strong>and</strong> a subsidiary local set trending generally southwest, dipping steeply<br />

northwest <strong>and</strong> spaced ~1.2 m on average (based on 73 m <strong>of</strong> scan lines intersecting 61<br />

b<strong>and</strong>s at 5 locations). B<strong>and</strong>s from these two sets generally cross-cut each other at a<br />

21


dihedral angle <strong>of</strong> 80° or more, <strong>and</strong> so appear to have formed contemporaneously (Sternl<strong>of</strong><br />

et al., 2004).<br />

4.2. Anticrack model<br />

The highly eccentric elliptical tip-to-tip pr<strong>of</strong>iles <strong>and</strong> uniform uniaxial internal<br />

compaction observed for individual CBs corresponds to a pure closing-mode sense <strong>of</strong><br />

relative boundary displacement distributed across a very thin b<strong>and</strong>. This geometry <strong>and</strong><br />

anti-mode I sense <strong>of</strong> displacement discontinuity defines CBs kinematically <strong>and</strong><br />

mechanically as anticracks (Mollema <strong>and</strong> Antonellini, 1996; Sternl<strong>of</strong> et al., 2005). As<br />

such, they would form symmetric with (orthogonal to) the direction <strong>of</strong> maximum<br />

compressive stress in an otherwise homogeneous, isotropic material (Figure 1.5), except<br />

when mechanically interacting with each other (Sternl<strong>of</strong> et al., 2005; Chapter 4 this<br />

thesis). The anticrack interpretation suggests that the normal to the dominant planar<br />

orientation <strong>of</strong> CBs in the Valley <strong>of</strong> Fire coincides with the direction <strong>of</strong> maximum<br />

compressive paleostress (σ1) acting regionally when they formed. In locations where a<br />

subsidiary set <strong>of</strong> cross-hatch CBs also formed, we suggest these reflect the orientation <strong>of</strong><br />

the intermediate principal paleostress (σ2), speculating that σ1 ≈ σ2 locally <strong>and</strong> that the<br />

preferred CB orientation flip-flopped between the two (see Chapter 4, this thesis).<br />

4.3. Outcrop to regional scale<br />

As described in the section on deposition, the Aztec s<strong>and</strong>stone is neither truly<br />

homogeneous nor isotropic. Rather, it is a granular material deposited in variable bedding<br />

orientations arranged as packages within a complex æolian sedimentary architecture. We<br />

suggest that this heterogeneous reality, coupled with apparently strong mechanical<br />

interactions between adjacent CBs, leads to the highly variable b<strong>and</strong> patterns commonly<br />

observed at outcrop scales. As the scale <strong>of</strong> observation increases, however, the influence<br />

<strong>of</strong> these local complications drops away. At a scale <strong>of</strong> hundreds <strong>of</strong> meters (Sternl<strong>of</strong> et al.,<br />

2006) to kilometers, the approximately planar fabric defined by north-northwest trending,<br />

steeply east-dipping b<strong>and</strong>s dominates, suggesting distributed formation in response to,<br />

<strong>and</strong> generally orthogonal with a paleo direction <strong>of</strong> regional compression.<br />

Finally, while CBs can be found in upper Aztec s<strong>and</strong>stone outcrops throughout<br />

southeastern Nevada, it is important to note that large exposures essentially devoid <strong>of</strong><br />

22


<strong>and</strong>s are also fairly common. Generally up to hundreds <strong>of</strong> meters in lateral extent,<br />

although some may be larger, we speculate that these holes in an otherwise pervasive CB<br />

fabric represent combinations <strong>of</strong> loading <strong>and</strong> mechanical properties not conducive to<br />

b<strong>and</strong> formation. Whether they are due primarily to small differences in porosity,<br />

cementation <strong>and</strong> cohesion, disadvantageous bedding orientations, pore-pressure<br />

variations or location within larger thrust structures remains to be deciphered. However,<br />

abundant evidence that sedimentary architecture can strongly influence CB <strong>propagation</strong><br />

(Hill, 1989; Sternl<strong>of</strong> et al., 2004; Flodin <strong>and</strong> Aydin, 2004) suggests that mechanical<br />

anisotropy related to depositional bedding, <strong>and</strong> perhaps also relative movement between<br />

adjacent cross-bed packages that alters the local stress state, might play decisive roles.<br />

(a)<br />

(b)<br />

σ 3<br />

σ 1<br />

x 3<br />

x 1<br />

x 1<br />

Figure 1.5. Schematic <strong>of</strong> anticrack-inclusion conceptual model. (a) Axisymmetric<br />

geometry <strong>of</strong> eccentric ellipsoidal b<strong>and</strong> aligned with the principal remote stresses (σ1 > σ2<br />

> σ3, compression positive). (b) Cross-sectional area <strong>of</strong> the b<strong>and</strong> (solid ellipse) relative to<br />

the pre-compacted area originally occupied by the same detrital grains (dashed ellipse).<br />

Net inward movement <strong>of</strong> the boundary (u1) during uniaxial compaction corresponds to a<br />

closing-mode (anti-mode I) sense <strong>of</strong> displacement. Their extreme eccentricity (~ 1)<br />

allows compaction b<strong>and</strong>s to be modeled as anticracks (adapted from Sternl<strong>of</strong> et al., 2005).<br />

23<br />

u 1<br />

x 2<br />

x 2<br />

σ 2


5. Compaction b<strong>and</strong> orientations<br />

Orientations were collected for 484 CBs intersected along 28 orthogonal scan lines<br />

totaling 359 m in length at 23 locations around the Valley <strong>of</strong> Fire. One additional set <strong>of</strong><br />

data (33 orientations) was collected as a control from outcrops <strong>of</strong> Aztec s<strong>and</strong>stone located<br />

~15 km southwest <strong>of</strong> the Valley <strong>of</strong> Fire in Buffington Pockets, a window through<br />

Paleozoics <strong>of</strong> the overriding Muddy Mountain thrust plate (Figure 1.3). All orientation<br />

data were collected as dip azimuth/dip angle, <strong>and</strong> only deformation b<strong>and</strong>s identifiable as<br />

CBs were measured (Figure 1.6). A b<strong>and</strong> was judged to be a CB if it was relatively thick<br />

(> 5 mm toward the middle), planar <strong>and</strong> continuous, <strong>and</strong> did not show discernable shear<br />

<strong>of</strong>fset. These simple criteria were applied to avoid counting low-angle shear b<strong>and</strong>s, as<br />

well as subsidiary b<strong>and</strong>s (e.g. linking structures) judged to be expressions <strong>of</strong> more<br />

localized stress regimes.<br />

As a point <strong>of</strong> interest, we note that b<strong>and</strong>s clearly belonging to the dominant CB set do<br />

occasionally show right-lateral <strong>of</strong>fsets on the order <strong>of</strong> centimeters. These b<strong>and</strong>s are<br />

generally encountered in close proximity to much younger, but similarly north-northwest<br />

trending dextral-slip faults (Myers <strong>and</strong> Aydin, 2004; Flodin <strong>and</strong> Aydin, 2004) <strong>and</strong> display<br />

an unusually high degree <strong>of</strong> induration. We suggest that these b<strong>and</strong>s represent Cretaceous<br />

CBs reactivated in shear during Miocene-Pliocene faulting, possibly due both to their<br />

favorable orientation <strong>and</strong> their potential for low shear strength as a result <strong>of</strong> the grain<br />

damage, comminution <strong>and</strong> preferential clay accumulation accommodated. Alternatively,<br />

one could argue that decreased porosity within these b<strong>and</strong>s resulted in increased shear<br />

modulus, causing them to concentrate subsequent fault-related stresses <strong>and</strong>, ultimately, to<br />

fail in synthetic shear themselves. In either case, the greater induration <strong>of</strong> these b<strong>and</strong>s<br />

today is likely due to frictional heating <strong>and</strong> healing <strong>of</strong> the shattered quartz grains,<br />

although this has yet to be confirmed with photomicroscopy. During our collection <strong>of</strong><br />

orientation data, these b<strong>and</strong>s were included as CBs.<br />

Figure 1.7 displays all the CB orientation data as planar poles in equal area, lower<br />

hemispheric projection stereograms for each location. These data clearly demonstrate the<br />

pervasive dominance <strong>of</strong> the steeply east-dipping CB set, although two other distinct<br />

orientation groups are apparent—steeply northwest dipping <strong>and</strong> very steeply southwest<br />

dipping. Figure 1.8 shows the stereograms for the five locations where multiple CB sets<br />

24


Figure 1.6. Collection <strong>of</strong> orientation data in the Valley <strong>of</strong> Fire (view is to the south). The<br />

compaction b<strong>and</strong>s shown are typical <strong>of</strong> those belonging to the dominant, through-going<br />

set in terms <strong>of</strong> their cm thickness, sub-meter spacing <strong>and</strong> planar yet anastomosing pattern.<br />

Acquisition <strong>of</strong> quality data is made generally easy by the prominent b<strong>and</strong> fins formed in<br />

outcrop. All data were collected as dip azimuth/dip angle, with spot checks <strong>of</strong><br />

reproducibility indicating about ± 5° for both.<br />

25


n = 20<br />

M<br />

n = 20<br />

P<br />

n = 22<br />

B<br />

n = 20<br />

R<br />

n = 22<br />

Q<br />

n = 20<br />

S<br />

n = 29<br />

C<br />

n = 20<br />

J<br />

n = 11<br />

N<br />

n = 20<br />

K<br />

n = 30<br />

O<br />

n = 23<br />

L<br />

26<br />

30<br />

n = 20<br />

A<br />

n = 13<br />

U<br />

n = 14<br />

D<br />

N<br />

0 1<br />

km<br />

25<br />

OS<br />

25<br />

n = 20<br />

V<br />

BF<br />

n = 32<br />

X<br />

n = 20<br />

I<br />

n = 20<br />

H<br />

n = 20<br />

W<br />

n = 18<br />

F<br />

n = 20<br />

G<br />

n = 17<br />

T<br />

n = 19<br />

E


Figure 1.7. (opposite page) Compaction b<strong>and</strong> orientation data from the Valley <strong>of</strong> Fire.<br />

Stereograms are lower hemispheric equal area projections <strong>of</strong> normals to the CB planes;<br />

lines to the central air photo (same as in Figure 1.1) indicate the approximate location<br />

where each set <strong>of</strong> data was collected; capital letter designations for each plot reflect the<br />

order <strong>of</strong> collection. Three distinct orientation groupings are apparent, although only five<br />

locations (C, B, F, L <strong>and</strong> O) exhibit more than one (see Figure 1.8). Relevant <strong>structural</strong><br />

data from Figure 1.1 (Overton syncline [OS], northern tip <strong>of</strong> Baseline fault [BF] <strong>and</strong><br />

Cretaceous bedding orientations) are as indicated. Stereogram X is from the location<br />

marked with a circled X in Figure 1.4, an area <strong>of</strong> Aztec outcrops called Buffington<br />

Pockets that comprise a window through the Paleozoics <strong>of</strong> the overriding Muddy<br />

Mountain thrust.<br />

P<br />

n = 22<br />

B<br />

S<br />

dihedral angle<br />

P to S = 80.5<br />

P<br />

n = 29<br />

C<br />

S<br />

T<br />

dihedral angles<br />

o<br />

P to S = 80.0<br />

o<br />

P to T = 32.7<br />

o<br />

S to T = 80.0<br />

P?<br />

n = 18<br />

F<br />

S<br />

T<br />

dihedral angles<br />

o<br />

P to S = 83.1<br />

o<br />

P to T = 7.5<br />

o<br />

S to T = 83.6<br />

P<br />

n = 23<br />

L<br />

S<br />

dihedral angle<br />

o<br />

P to S = 75.9<br />

P<br />

n = 30<br />

O<br />

S<br />

T<br />

dihedral angle<br />

P to S = 87.7<br />

o o<br />

Figure 1.8. Angular relationships between compaction b<strong>and</strong> sets at the five outcrop<br />

locations exhibiting multiple orientations (see Figure 1.7). P = primary set in terms <strong>of</strong><br />

park-wide abundance, S = secondary set, T = tertiary set. Dihedral angles were computed<br />

between the mean orientations for each b<strong>and</strong> set. Sets P <strong>and</strong> S are at high dihedral angle<br />

to each other at all locations, as are sets S <strong>and</strong> T. Sets P <strong>and</strong> T, however, are at variably<br />

low dihedral angle.<br />

27


were encountered, giving the dihedral angles between the mean orientations for each<br />

set—labeled P (primary), S (secondary) <strong>and</strong> T (tertiary), based on relative abundance. P<br />

<strong>and</strong> S share a dihedral angle <strong>of</strong> about 80° at all five locations, as do S <strong>and</strong> T at the two<br />

locations where they coexist. P <strong>and</strong> T, however share a variably low dihedral angle at<br />

these same two locations, having similar trends, but opposite dip directions.<br />

Figure 1.9 displays the combined data for all locations (n = 484) as poles, density<br />

contours, <strong>and</strong> Rose diagrams <strong>of</strong> both strike <strong>and</strong> dip orientations. Table 1.1 provides a<br />

brief statistical summary <strong>of</strong> the numbers, including mean <strong>and</strong> median dihedral angles, <strong>and</strong><br />

relative abundance. Two basic conclusions are immediately apparent. Firstly, wherever it<br />

occurs, the secondary CB set forms essentially orthogonal to the primary set. Secondly,<br />

the tertiary CB group is a steeply dipping subset <strong>of</strong> the primary orientation. Although the<br />

data were not collected to provide a statistically valid measure <strong>of</strong> relative abundance<br />

between the CB sets, they do also reasonably illustrate the dominance <strong>of</strong> the primary set<br />

at ~83% <strong>of</strong> all b<strong>and</strong>s.<br />

6. Tectonic interpretation<br />

Hill (1989) first suggested that high-angle deformation b<strong>and</strong>s in the Aztec s<strong>and</strong>stone<br />

that do not exhibit macroscopic shear (CBs in current usage) resulted from tectonic<br />

compression related to the Sevier orogeny. His primary arguments were that the b<strong>and</strong>s<br />

trend generally parallel to the encroaching thrust front <strong>and</strong> orthogonal to the east-vergent<br />

tectonic transport direction, <strong>and</strong> that they comprise the oldest structures present based on<br />

cross-cutting relationships. Subsequent workers (Taylor et al., 1999; Taylor <strong>and</strong> Pollard,<br />

2000; Myers <strong>and</strong> Aydin, 2004; Flodin <strong>and</strong> Aydin, 2004; Eichhubl et al., 2004; Sternl<strong>of</strong> et<br />

al., 2004, 2005, 2006) have all come to the same conclusion <strong>and</strong> we concur, <strong>of</strong>fering a<br />

more in-depth examination <strong>of</strong> the evidence <strong>and</strong> implications below.<br />

6.1. Timing, spatial <strong>and</strong> material constraints<br />

The analysis <strong>of</strong> Taylor <strong>and</strong> Pollard (2000) demonstrates that CBs were already<br />

present in the Aztec to influence the initial upward <strong>and</strong> eastward expulsion <strong>of</strong> reducing<br />

basinal brines from beneath the advancing Sevier thrust front, which bleached the middle<br />

<strong>and</strong> upper parts <strong>of</strong> the s<strong>and</strong>stone as observed today (Eichhubl et al., 2004) (Figure 1.1).<br />

Thus, CBs formed in the s<strong>and</strong>stone while it was stained uniformly red with hematite<br />

grain coatings, suggesting that this trace (~1% by volume) cement (Flodin et al., 2003;<br />

28


(a)<br />

(c)<br />

P<br />

P<br />

S<br />

T<br />

S<br />

T<br />

Figure 1.9. Stereograms for combined orientation data (equal area, lower hemispheric<br />

projections). P = primary set, S = secondary set, T = tertiary set, n = 484. (a) Normals to<br />

the b<strong>and</strong> planes. (b) Density contours <strong>of</strong> planar normals (2.5% contour interval). (c) Rose<br />

diagram <strong>of</strong> strike azimuth data. (d) Rose diagram <strong>of</strong> dip azimuth data.<br />

Table 1.1. Summary <strong>of</strong> compaction b<strong>and</strong> orientation data<br />

Primary Set (P)<br />

(n = 401)<br />

P<br />

Secondary Set (S)<br />

(n = 61)<br />

S<br />

T<br />

S<br />

T<br />

Tertiary Set (T)<br />

(n = 22)<br />

(b)<br />

P<br />

(d)<br />

Control Set<br />

(n = 33)<br />

Azimuth<br />

mean 81.7° 337.8° 241.2° 116.9°<br />

median 82.0° 338.0° 245.0° 116.0°<br />

st<strong>and</strong>ard dev. 13.3° 6.7° 11.5° 6.7°<br />

Dip angle<br />

mean 68.7° 56.9° 80.7° 61.6°<br />

median 69.0° 62.0° 80.5° 62.0°<br />

st<strong>and</strong>ard dev. 6.7° 13.6° 6.2° 4.9°<br />

Dihedral angles mean data median data<br />

P to S 89.4° 88.2°<br />

P to T 36.6° 34.8°<br />

S to T 89.6° 88.2°<br />

P to Control 32.6° 31.6°<br />

Rel. abundance 82.9% 12.6% 4.5% na<br />

29


Eichhubl et al., 2004) did not lend significant cohesion to the detrital grains (Sternl<strong>of</strong> et<br />

al., 2005) or, at least that it did not impede CB formation. Why the lower boundary <strong>of</strong> the<br />

alteration front is located stratigraphically where it is, <strong>and</strong> why this horizon roughly<br />

coincides with the lower extent <strong>of</strong> compaction b<strong>and</strong> formation remains a significant open<br />

question. Given that the b<strong>and</strong>s predate alteration <strong>and</strong> pure coincidence is unlikely, it is<br />

interesting to speculate, however, that the relatively newly formed CB arrays in the<br />

Cretaceous Aztec s<strong>and</strong>stone played an active role in attracting <strong>and</strong> concentrating flow <strong>of</strong><br />

the expulsing brines.<br />

As described above in Section 3, a preponderance <strong>of</strong> evidence indicates that the Aztec<br />

s<strong>and</strong>stone in the immediate study area just avoided being overridden by the Willow Tank<br />

<strong>and</strong> Muddy Mountain thrust sheets <strong>and</strong> that, based on diagenetic evidence, has never<br />

been buried by much more than the combined thickness <strong>of</strong> overlying Cretaceous <strong>and</strong><br />

Neogene sediments (~1,600 m). Directly overlying the Aztec is the basal conglomerate <strong>of</strong><br />

the Willow Tank formation, interpreted to be a gravel pediment indicative <strong>of</strong><br />

erosional/depositional stasis. An upward coarsening, 1,300-m-thick sequence <strong>of</strong><br />

Cretaceous deposition followed (Figure 1.2), dominated by more than 1,100 m <strong>of</strong> the<br />

Baseline s<strong>and</strong>stone representing aqueous redeposition <strong>of</strong> the Aztec from western upl<strong>and</strong><br />

sources into a forel<strong>and</strong> basin (Longwell, 1949; Bohannon, 1983). Up to 600 m <strong>of</strong> these<br />

lower Cretaceous sediments are overridden by the Willow Tank thrust, which is<br />

associated with shear b<strong>and</strong>ing in the Aztec that post-dates CB formation.<br />

Sternl<strong>of</strong> et al. (2005) conclude that no more than 600 m <strong>of</strong> fine-grained forel<strong>and</strong> basin<br />

sediments buried the Aztec s<strong>and</strong>stone during CB formation, <strong>and</strong> likely less. We further<br />

suggest that the evolution <strong>of</strong> this flexural forel<strong>and</strong> basin was in its earliest stages during<br />

CB formation <strong>and</strong> that no more than 200 m <strong>of</strong> clay-rich swampl<strong>and</strong> deposits <strong>of</strong> the<br />

Willow Tank formation covered the Aztec. The end-member scenario is that the Aztec<br />

s<strong>and</strong>stone was essentially unburied, covered only by the gravel pediment, which is<br />

suggestive <strong>of</strong> mild uplift related to broad regional warping <strong>of</strong> the crust at the onset <strong>of</strong> the<br />

Sevier compression (Fleck, 1970; Brock <strong>and</strong> Engelder, 1977). In any case, it appears<br />

certain that CB formation in the Aztec occurred well before the thrust front reached the<br />

area. This interpretation is further corroborated by the observation that CBs exposed at<br />

Buffington Pockets (Figure 1.3) are substantially similar in <strong>geology</strong> <strong>and</strong> orientation to<br />

30


those at the Valley <strong>of</strong> Fire, despite having been overridden by up to 5 km <strong>of</strong> the Muddy<br />

Mountain thrust sheet (Brock <strong>and</strong> Engelder, 1977) <strong>and</strong> probably rotated in a clockwise<br />

sense during later movement along the Las Vegas Valley shear zone <strong>and</strong> Lake Mead fault<br />

system (Figure 1.3).<br />

To summarize, we interpret CBs to have formed in the Aztec s<strong>and</strong>stone when it was<br />

shallowly buried (< 600 m) <strong>and</strong> well before active east-vergent thrust faulting affected<br />

the area. There is no evidence to suggest that the Aztec at this time was anything more<br />

than a weakly cemented granular aggregate (albeit one with a complex depositional<br />

architecture) topped with relatively thin (


P<br />

(a)<br />

(c)<br />

S<br />

P<br />

S<br />

Figure 1.10. Stereograms for restored orientation data (equal area, lower hemispheric<br />

projections0. P = primary set, S = secondary set, n=484. All orientation data were rotated<br />

25° counterclockwise about an axis with azimuth 315° (NW) <strong>and</strong> plunge <strong>of</strong> 0°, in order to<br />

restore them to pre-tilting attitudes. (a) Normals to the b<strong>and</strong> planes. (b) Density contours<br />

<strong>of</strong> planar normals (2.5% contour interval). (c) Rose diagram <strong>of</strong> strike azimuth data. (d)<br />

Rose diagram <strong>of</strong> dip azimuth data.<br />

Table 1.2. Summary <strong>of</strong> restored compaction b<strong>and</strong> orientation data<br />

Primary Set (P)<br />

(n = 423)<br />

P<br />

S<br />

Secondary Set (S)<br />

(n = 61)<br />

S<br />

(b)<br />

(d)<br />

P<br />

Control Set<br />

(n = 33)<br />

Azimuth<br />

mean 89.9° 317.9° 116.9°<br />

median 91.9° 320.4° 116.0°<br />

st<strong>and</strong>ard dev. 16.9° 12.0° 6.7°<br />

Dip angle<br />

mean 51.7° 51.2° 61.6°<br />

median 51.2° 55.2° 62.0°<br />

st<strong>and</strong>ard dev. 9.1° 12.1° 4.9°<br />

Dihedral angles mean data median data<br />

P to S 88.8° 86.2°<br />

P to Control 24.5° 22.7°<br />

Rel. abundance 87.4% 12.6% na<br />

32


otated all <strong>of</strong> the CB orientation data 25° counterclockwise about an axis trending 315°<br />

(northwest) <strong>and</strong> plunging 0°, in keeping with available bedding orientation data for the<br />

Cretaceous units (Bohannon 1977, 1983). The restored data are presented graphically in<br />

Figure 1.10 <strong>and</strong> summarized in Table 1.2.<br />

The most obvious result <strong>of</strong> the restoration is that the 22 measurements originally<br />

attributed to a distinct third set <strong>of</strong> CBs trending northwest <strong>and</strong> dipping steeply southwest<br />

(also previously identified by Hill, 1989 <strong>and</strong> others) can reasonably be categorized as<br />

falling within the periphery <strong>of</strong> the primary (dominant) orientation group. This leaves just<br />

two sets <strong>of</strong> CBs with mean orientations that are essentially orthogonal (dihedral angle <strong>of</strong><br />

88.8°). On average, the dominant set trends due north <strong>and</strong> dips 52° east, while the<br />

secondary set trends almost due northeast <strong>and</strong> dips 51°northwest. The axis <strong>of</strong> intersection<br />

between these mean CB orientations—azimuth <strong>of</strong> 24° <strong>and</strong> plunge <strong>of</strong> 27°—represents the<br />

approximate line <strong>of</strong> σ3, which turns out to be more nearly horizontal than vertical. The<br />

inferred azimuth/plunge directions <strong>of</strong> all three paleostresses are: 270°/38° (σ1), 138°/39°<br />

(σ2) <strong>and</strong> 24°/27° (σ3) (Figure 1.11a). These inferred paleostress directions represent<br />

approximate estimates, both because σ1 <strong>and</strong> σ2 (the poles to the dominant <strong>and</strong> secondary<br />

mean CB orientations) are not quite orthogonal (88.8°), <strong>and</strong> as a result <strong>of</strong> the scatter in<br />

the raw orientation data (± 17° in azimuth <strong>and</strong> 12° in dip) (Table 1.2).<br />

6.3. Geomechanical implications<br />

The radical departure <strong>of</strong> our inferred paleostress state from the classic Andersonian<br />

expectation that one <strong>of</strong> the principal stress components is always subvertical, σ3 in thrust-<br />

faulting environments (Anderson, 1951) means one <strong>of</strong> three things: the reconstruction <strong>of</strong><br />

the paleo CB orientations is faulty <strong>and</strong> the two sets were actually subvertical as well as<br />

orthogonal; the underlying interpretation <strong>of</strong> CBs as anticracks is incorrect <strong>and</strong> they do not<br />

reflect the orientations <strong>of</strong> the principal paleostresses; or the analysis is correct <strong>and</strong> points<br />

to an as yet unrecognized regional tectonic explanation. We address the two former<br />

possibilities first.<br />

While our first-order approach to the reconstruction is certainly approximate, it is<br />

fundamentally sound <strong>and</strong> anything more complex is currently unwarranted given the<br />

available data. The best way to refine the reconstruction would be to collect a high<br />

density <strong>of</strong> bedding orientation data as stratigraphically low in the Cretaceous deposits as<br />

33


possible, <strong>and</strong> use these to better constrain the azimuth <strong>and</strong> plunge <strong>of</strong> the tilting axis, as<br />

well as the amount <strong>of</strong> tilting. It would also be reasonable to avoid the southeastern limb<br />

<strong>of</strong> the Overton syncline as adding an unnecessary complication <strong>and</strong> source <strong>of</strong> potential<br />

error.<br />

These refinements, however, are unlikely to alter substantially the existing paleostress<br />

assessment. Varying the tilting axis azimuth <strong>and</strong> plunge by ± 15° around the 315°/0°<br />

orientation used produces a mild range <strong>of</strong> variation, while the amount <strong>of</strong> the tilt (~25°) is<br />

reasonably well constrained by field measurements. In any event, no geologically<br />

reasonable restoration can bring both sets <strong>of</strong> CBs to paleo vertical, as this would also<br />

result in near vertical Cretaceous strata during their deposition.<br />

Could the underlying interpretation <strong>of</strong> CBs as anticracks be flawed? The specific<br />

<strong>mechanics</strong> <strong>of</strong> CB <strong>propagation</strong> remains an area <strong>of</strong> active research—observationally,<br />

theoretically <strong>and</strong> experimentally—<strong>and</strong> it is possible that the b<strong>and</strong>s will turn out to<br />

propagate obliquely to the principal stresses. This could be the case so long as anti-mode<br />

I deformation dominates such that shear displacements are on the order <strong>of</strong> the mean grain<br />

size (0.25 mm) <strong>and</strong> thus nearly impossible to distinguish. Such a result would, <strong>of</strong> course,<br />

alter our principal paleostress interpretation based on the CB orientations. In this regard,<br />

it is interesting to note that, in individual outcrops containing both CB sets, the mean<br />

dihedral angle is ~80°, <strong>and</strong> can be less (Figure 1.8). This departure from the more nearly<br />

orthogonal result when looking at the mean b<strong>and</strong> orientations could be a clue that some<br />

degree <strong>of</strong> resolved shear is involved in <strong>propagation</strong> at the micromechanical scale. On the<br />

other h<strong>and</strong>, 80° is closer to 90° than to 60° (the typical dihedral angle between conjugate<br />

shear planes) <strong>and</strong> the preponderance <strong>of</strong> available data points strongly to the dominance <strong>of</strong><br />

uniaxial compaction (Sternl<strong>of</strong> et al., 2005; Chapter 4, this thesis). We suggest therefore<br />

that a significantly different interpretation <strong>of</strong> paleostress direction relative to CB<br />

orientation—such as σ1 bisecting the dihedral angle between the primary <strong>and</strong> secondary<br />

b<strong>and</strong> sets, <strong>and</strong> σ2 corresponding to the axis <strong>of</strong> intersection between them—is highly<br />

unlikely. In any case, no plausible reinterpretation would yield principal paleostress<br />

directions in agreement with Andersonian expectations.<br />

This leaves the third possible interpretation—that the current analysis is substantially<br />

correct <strong>and</strong> points to an as yet unrecognized tectonic explanation. Any such explanation<br />

34


almost certainly involves movement along regional bounding faults, such that tractions<br />

applied to a tectonic block containing the Aztec s<strong>and</strong>stone resulted in the orientation <strong>of</strong><br />

paleostress inferred from the CBs. Such a configuration is difficult to visualize in 3-D.<br />

A highly schematized 2-D treatment that ignores the secondary CB set, however, serves<br />

to illustrate the essential concept (Figure 1.11b).<br />

For this simple model, assuming homogeneous, isotropic, linear elasticity <strong>and</strong> using<br />

the finite element method s<strong>of</strong>tware Abaqus®, we consider a vertical, east-west cross<br />

section through the Mesozoic units <strong>of</strong> the study area as a block sliced orthogonally to the<br />

mean orientation <strong>of</strong> the dominant CB set <strong>and</strong> sitting atop a horizontal detachment at the<br />

top <strong>of</strong> the Hermit shale. If this 3.6-km-thick block is moving stably westward along the<br />

detachment (i.e. being pushed quasi-statically), the trajectories <strong>of</strong> σ1 induced a few km in<br />

from its eastern edge are as shown by the tick marks in Figure 1.11b. Within the zone<br />

indicated, 200 to 800 m deep <strong>and</strong> 5 to 9 km from the eastern edge <strong>of</strong> the block, which<br />

coincides with a lateral section <strong>of</strong> the CB-rich upper Aztec, these trajectories roughly<br />

coincide with our inferred orientation <strong>of</strong> σ1 that produced east dipping CBs. This<br />

rudimentary treatment is not well constrained, but it does demonstrate that the inferred<br />

orientation <strong>of</strong> σ1 can be produced over a broad, km-scale region in a geologically realistic<br />

way. In fact, any variety <strong>of</strong> geologically reasonable configurations <strong>of</strong> mechanical<br />

stratigraphy, <strong>and</strong> detachment <strong>and</strong> lateral bounding faults can be expected to produce the<br />

inferred paleostress orientations. Nonetheless, even the simple 2-D model presented here<br />

provides two relatively robust predictions testable by observation. Does the dip <strong>of</strong> the<br />

dominant b<strong>and</strong> set systematically decrease with increasing depth <strong>and</strong> increase from west<br />

to east? And is there any independent evidence for westward motion <strong>of</strong> the study area<br />

along a relatively shallow detachment?<br />

On this last point, we note the interpretation <strong>of</strong> Carpenter <strong>and</strong> Carpenter (1994) that<br />

west-vergent reverse faulting along what they called the North Buffington back-thrust<br />

system comprised the earliest expression <strong>of</strong> Sevier tectonism in southeastern Nevada,<br />

predating the formation <strong>of</strong> the dominant east-vergent thrusts (Muddy Mountain, Willow<br />

Tank, etc.) now exposed in the study area.<br />

35


1.5(ρgz)<br />

(b)<br />

WEST<br />

Willow Tank<br />

Upper Aztec<br />

(zone <strong>of</strong> interest)<br />

Lower Aztec<br />

Triassic strata<br />

(a)<br />

σ1<br />

~ 270/38<br />

traction-free ground surface<br />

detachment<br />

0 m<br />

200 m<br />

800 m<br />

3,600 m<br />

36 km 9 km<br />

5 km 0 km<br />

ρgz<br />

σ3<br />

~ 024/27<br />

σ2<br />

~ 138/39<br />

z<br />

EAST<br />

2(ρgz)<br />

Figure 1.11. Paleostress orientations <strong>and</strong> tectonic interpretation for the Aztec s<strong>and</strong>stone<br />

during compaction b<strong>and</strong> formation. (a) Equal area, lower hemisphere stereographic<br />

projection <strong>of</strong> the approximate inferred principal paleostress orientations (σ1 > σ2 > σ3,<br />

compression positive). (b) Simplified 2-D linear elastic, finite element model (using<br />

Abaqus®) <strong>of</strong> one tectonic scenario that could produce the inferred orientation <strong>of</strong> the<br />

maximum compressive paleostress (σ1). In this formulation, we consider a homogeneous,<br />

isotropic, linear elastic block representing an east-west cross section through the<br />

Mesozoic strata <strong>of</strong> the study area, which is orthogonal to the mean orientation <strong>of</strong> the<br />

dominant CB set. This block, the hanging wall <strong>of</strong> a horizontal thrust, is subjected to<br />

gravitational forces <strong>and</strong> unbalanced lateral forces that push it quasi-statically westward<br />

along a rigid detachment, which resists the motion by generating opposing shear tractions.<br />

The resulting trajectories <strong>of</strong> σ1 induced a few km from the eastern edge <strong>of</strong> the block <strong>and</strong><br />

within the CB-genic zone <strong>of</strong> the upper Aztec (tick marks) roughly coincide with the<br />

inferred orientation (z is depth, g is gravity, ρ is density, each grid block is 200 m by<br />

1,000 m).<br />

36


7. Concluding observations<br />

The data <strong>and</strong> analyses presented above indicate that CBs formed as anticracks in the<br />

Aztec while it comprised a shallowly buried, weakly lithified s<strong>and</strong>stone aquifer subjected<br />

to regional compression associated with the earliest phases <strong>of</strong> Sevier tectonism in the area.<br />

Interpreted as such, the presence <strong>of</strong> two orthogonal, coeval <strong>and</strong> coexisting sets <strong>of</strong> CBs<br />

can be used to constrain the 3-D paleostress state in which they formed, with the<br />

dominant set orthogonal to the maximum compressive stress (σ1) <strong>and</strong> the secondary set<br />

orthogonal to intermediate compressive stress (σ2). Furthermore, the mutually cross-<br />

cutting presence <strong>of</strong> both sets in the same outcrop suggests that σ1 ≈ σ2. That the inferred<br />

paleostress orientations depart radically from classical Andersonian expectations can<br />

reasonably be interpreted as evidence <strong>of</strong> as yet unrecognized regional structures,<br />

specifically a relatively shallow detachment along which a coherent <strong>structural</strong> block<br />

containing the Aztec now exposed in the Valley <strong>of</strong> Fire moved relatively westward.<br />

Based on an average spacing between CBs <strong>of</strong> the dominant, north-south trending set<br />

<strong>of</strong> 0.7 m, <strong>and</strong> an average b<strong>and</strong> thickness <strong>of</strong> 1 cm representing 10% uniaxial compaction<br />

(~1.1 mm), the Aztec s<strong>and</strong>stone experienced less than 0.2% total shortening as the result<br />

<strong>of</strong> CB formation. Even assuming an average b<strong>and</strong> spacing <strong>of</strong> only 10 cm hardly bumps<br />

the shortening past 1%. If homogeneously distributed, dropping overall porosity from<br />

25% to 24%, such a mild deformation event would be <strong>of</strong> no consequence to fluid flow.<br />

The highly localized, yet pervasively interconnected nature <strong>of</strong> the CB fabric formed,<br />

however, constitutes a considerable network <strong>of</strong> low-permeability baffles to fluid flow<br />

(CB permeability ~0.1% <strong>of</strong> the inherent s<strong>and</strong>stone permeability) that has been shown<br />

capable <strong>of</strong> exerting pr<strong>of</strong>ound effects at scales <strong>of</strong> practical interest (Sternl<strong>of</strong> et al., 2004;<br />

Sternl<strong>of</strong> et al., 2006).<br />

The analysis presented here <strong>of</strong> compaction b<strong>and</strong>s as indicators <strong>of</strong> paleostress in <strong>and</strong><br />

tectonic structure around the exhumed analog aquifer/reservoir that is the Aztec<br />

s<strong>and</strong>stone can also be applied in reverse. That is, given independent knowledge <strong>of</strong> the<br />

tectonic stress <strong>and</strong> material history <strong>of</strong> a s<strong>and</strong>stone similar to the Aztec, reasonable<br />

forecasts <strong>of</strong> the possible presence <strong>and</strong> gross geometry <strong>of</strong> CBs in the subsurface can be<br />

made. Such insights could prove highly valuable in the efficient development aquifers<br />

<strong>and</strong> reservoirs.<br />

37


9. Acknowledgements<br />

My sincere thanks go to John Childs for his able <strong>and</strong> entertaining assistance in the<br />

field, <strong>and</strong> to David Pollard for many thought-provoking discussions. Thanks also to the<br />

Valley <strong>of</strong> Fire State Park staff for their indulgence. This work was funded by the U.S.<br />

Department <strong>of</strong> Energy, Office <strong>of</strong> Basic Energy Science, Geosciences Research Program<br />

under grant DE-FG03-94ER14462, awarded to Pollard <strong>and</strong> Atilla Aydin as principal<br />

investigators. Additional support was provided by <strong>Stanford</strong> Rock Fracture Project.<br />

38


1. Abstract<br />

Chapter 2<br />

Anticrack-inclusion model for compaction b<strong>and</strong>s in s<strong>and</strong>stone<br />

Detailed observations <strong>of</strong> compaction b<strong>and</strong>s exposed in the Aztec s<strong>and</strong>stone <strong>of</strong><br />

southeastern Nevada indicate that these thin, tabular, bounded features <strong>of</strong> localized<br />

porosity loss initiated at pervasive grain-scale flaws, which collapsed in response to<br />

compressive tectonic loading. From many <strong>of</strong> these Griffith-type flaws, an apparently self-<br />

sustaining progression <strong>of</strong> collapse propagated outward to form b<strong>and</strong>s <strong>of</strong> compacted grains<br />

a few centimeters thick <strong>and</strong> tens <strong>of</strong> meters in planar extent. These compaction b<strong>and</strong>s can<br />

be idealized as highly eccentric ellipsoidal bodies that have accommodated uniform<br />

uniaxial plastic strain parallel to their short dimension within a surrounding elastic<br />

material. They thus can be represented mechanically as contractile Eshelby inclusions,<br />

which generate near-tip compressive stress concentrations consistent with self-sustaining,<br />

in-plane <strong>propagation</strong>. The combination <strong>of</strong> extreme aspect ratio (>10 -4 ) <strong>and</strong> significant<br />

uniaxial plastic strain (~10%) also justifies an approximation <strong>of</strong> the b<strong>and</strong>s as anticracks—<br />

sharp boundaries across which a continuous distribution <strong>of</strong> closing-mode displacement<br />

discontinuity has been accommodated. This anticrack interpretation <strong>of</strong> compaction b<strong>and</strong>s<br />

is analogous to that <strong>of</strong> pressure solution surfaces, except that porosity loss takes the place<br />

<strong>of</strong> material dissolution. We find that displacement discontinuity boundary element<br />

method modeling <strong>of</strong> compaction b<strong>and</strong>s as anticracks within a two-dimensional linear<br />

elastic continuum can accurately represent the perturbed external stress fields they<br />

induce.<br />

2. Introduction<br />

Compaction b<strong>and</strong>s represent one kinematic end member <strong>of</strong> the suite <strong>of</strong> structures<br />

known collectively as deformation b<strong>and</strong>s, which also includes shear <strong>and</strong> dilation b<strong>and</strong>s<br />

(Antonellini et al., 1994; Aydin, 1978; Du Bernard, 2002; Mollema <strong>and</strong> Antonellini,<br />

1996). Due to the internal loss <strong>of</strong> porosity <strong>and</strong> permeability they accommodate,<br />

deformation b<strong>and</strong>s can significantly affect bulk s<strong>and</strong>stone permeability when present as<br />

pervasive arrays, with important implications for the management <strong>of</strong> both groundwater<br />

<strong>and</strong> hydrocarbon resources (Sternl<strong>of</strong> et al., 2004).<br />

39


The term compaction b<strong>and</strong> (CB) was coined by Mollema <strong>and</strong> Antonellini (1996) to<br />

describe “…tabular zones <strong>of</strong> localized deformation…that accommodate pure<br />

compaction” in s<strong>and</strong>stone, <strong>and</strong> which constitute “…analogs for anticracks…such as<br />

pressure solution surfaces,” following the analysis <strong>of</strong> Fletcher <strong>and</strong> Pollard (1981).<br />

Mollema <strong>and</strong> Antonellini described two types <strong>of</strong> compactions b<strong>and</strong>s—straight, thick<br />

b<strong>and</strong>s <strong>and</strong> wavy, or crooked b<strong>and</strong>s—both exhibiting porosity loss in the absence <strong>of</strong> shear<br />

accommodated by granular rearrangement, grain cracking, indentation <strong>and</strong> limited<br />

comminution. They identified these structures as forming primarily within the<br />

compressional quadrants <strong>of</strong> deformation b<strong>and</strong>-style faults (Aydin, 1978; Aydin <strong>and</strong><br />

Johnson, 1978) in the Jurassic æolian Navajo s<strong>and</strong>stone <strong>of</strong> the Kaibab Monocline in<br />

Utah, finding them to be on the order <strong>of</strong> meters in trace length, segmented, <strong>and</strong> restricted<br />

to a limited outcrop area on the downthrown limb <strong>of</strong> the monocline in relatively coarse<br />

grained, high porosity layers.<br />

Extensive arrays <strong>of</strong> CBs pervade the upper half <strong>of</strong> the Jurassic Aztec s<strong>and</strong>stone<br />

exposed in <strong>and</strong> around the Valley <strong>of</strong> Fire State Park <strong>of</strong> southeastern Nevada (Figure 2.1)<br />

where, however, they bear no immediate genetic relationship to local faults <strong>and</strong> are tens<br />

to more than one hundred <strong>of</strong> meters in trace length (Sternl<strong>of</strong> et al., 2004). The Aztec is a<br />

1,400-m-thick æolian s<strong>and</strong>stone composed <strong>of</strong> greater than 90% detrital quartz <strong>and</strong><br />

typified by large-scale tabular <strong>and</strong> trough cross-bedding, average porosity <strong>of</strong> about 20%,<br />

<strong>and</strong> a mean grain diameter <strong>of</strong> about 0.25 mm within a range <strong>of</strong> about 0.1 mm to 0.5 mm<br />

(Flodin et al., 2005). Throughout the weakly lithified upper Aztec, two distinct phases <strong>of</strong><br />

deformation b<strong>and</strong>s are present. Steeply dipping, cm-thick CBs <strong>of</strong> the first phase comprise<br />

the oldest <strong>structural</strong> fabric. The CBs were crosscut <strong>and</strong> <strong>of</strong>fset by relatively low-angle<br />

shear b<strong>and</strong>s <strong>of</strong> the second phase. Overprinting by joints <strong>and</strong> the evolution <strong>of</strong> joint-based<br />

strike-slip faults associated with Basin <strong>and</strong> Range extension followed (Eichhubl et al.,<br />

2004; Flodin <strong>and</strong> Aydin, 2004; Hill, 1989; Myers <strong>and</strong> Aydin, 2004; Sternl<strong>of</strong> et al., 2004;<br />

Taylor et al., 1999). Noting their relative age <strong>and</strong> dominant orientation—north-northwest<br />

trend <strong>and</strong> steep dip—Hill (1989) first suggested that the CBs formed in response to east-<br />

northeast-directed regional compression associated with the Cretaceous Sevier orogeny.<br />

Building on this earlier work, we present a conceptual <strong>and</strong> mechanical model <strong>of</strong> CBs<br />

in porous s<strong>and</strong>stone as anticrack-inclusions [Sternl<strong>of</strong> <strong>and</strong> Pollard, 2001; Sternl<strong>of</strong> <strong>and</strong><br />

40


NV UT<br />

CA<br />

AZ<br />

Park Road<br />

Route 169<br />

0 km 5<br />

Map<br />

Detail<br />

Park Boundary<br />

*<br />

Silica Dome<br />

Park Office<br />

NEVADA<br />

ARIZONA<br />

10 km<br />

Valley<br />

<strong>of</strong> Fire<br />

State Park<br />

N<br />

LAKE<br />

MEAD<br />

115 o 30'<br />

Figure 2.1. Location <strong>of</strong> the Valley <strong>of</strong> Fire State Park, southeastern Nevada (inset), where<br />

more than 20 square kilometers <strong>of</strong> the 1,400-m-thick æolian Jurassic Aztec S<strong>and</strong>tone are<br />

extensively exposed. Photo shows view northward from the location marked * on the<br />

park detail. Throughout the upper half <strong>of</strong> the Aztec, compaction b<strong>and</strong>s crop out in<br />

positive relief as sub-parallel, centimeter-thick, north-northwest-trending, steeply eastdipping<br />

tabular fins spaced from centimeters to meters apart.<br />

41<br />

36 o 15'


Pollard, 2002). The particular utility <strong>of</strong> this hybrid interpretation lies in combining the<br />

analytical insights <strong>of</strong> Eshelby inclusion theory (Eshelby, 1957; Eshelby, 1959) with the<br />

power <strong>of</strong> boundary element method (BEM) simulations (Crouch <strong>and</strong> Starfield, 1983) to<br />

underst<strong>and</strong> the <strong>mechanics</strong> <strong>of</strong> CB evolution. Starting with targeted field <strong>and</strong> petrographic<br />

observations <strong>of</strong> CBs from the Aztec s<strong>and</strong>stone in the Valley <strong>of</strong> Fire, we build a<br />

mechanical analysis <strong>of</strong> these structures as Eshelby-type inclusions to establish the<br />

validity <strong>and</strong> utility <strong>of</strong> a two-dimensional anticrack treatment <strong>of</strong> CBs using a displacement<br />

discontinuity BEM.<br />

The basic hypothesis underlying this research is that a fundamental mechanism <strong>of</strong><br />

incremental, quasi-static <strong>and</strong> (anti)crack-like <strong>propagation</strong> was at work throughout the<br />

Aztec such that individual CBs tended to form orthogonal to the most compressive<br />

remote stress. The diversity <strong>of</strong> outcrop patterns observed therefore reflects the<br />

interactions <strong>of</strong> CBs with local material <strong>and</strong> stress heterogeneities through this<br />

fundamental <strong>propagation</strong> mechanism. Beyond contributing to the underst<strong>and</strong>ing <strong>of</strong> an<br />

interesting <strong>and</strong> generally overlooked mode <strong>of</strong> <strong>structural</strong> failure in porous, granular earth<br />

materials, this research is motivated by the practical goal <strong>of</strong> using mechanical<br />

underst<strong>and</strong>ing <strong>of</strong> the phenomenon to forecast its occurrence <strong>and</strong> associated fluid-flow<br />

effects in active reservoirs <strong>and</strong> aquifers. Although the research presented here is built on<br />

detailed CB data acquired from a specific outcrop <strong>of</strong> Aztec s<strong>and</strong>stone in the Valley <strong>of</strong><br />

Fire, we suggest that the intrinsic attributes identified are common to all CBs in the Aztec<br />

<strong>and</strong>, by extension, to other porous s<strong>and</strong>stones with similar depositional <strong>and</strong> tectonic<br />

histories.<br />

3. Methods<br />

We focused on a “type-locality” outcrop <strong>of</strong> CBs in Aztec s<strong>and</strong>stone near Silica Dome<br />

in the Valley <strong>of</strong> Fire State Park (Figure 2.1, inset) where elementary patterns <strong>of</strong><br />

approximately parallel b<strong>and</strong>s are well exposed. Using a steel tape <strong>and</strong> calipers, we<br />

measured tip-to-tip thickness pr<strong>of</strong>iles <strong>of</strong> the two-dimensional b<strong>and</strong> traces in outcrop<br />

(Figure 2.2), focusing on one example as representing the prototypical compaction b<strong>and</strong>.<br />

Eleven cores were collected along the trace <strong>of</strong> one CB from tip to middle, as well as<br />

from essentially undisturbed s<strong>and</strong>stone nearby. These were vacuum-impregnated with<br />

optical-grade blue epoxy <strong>and</strong> used to make 37 st<strong>and</strong>ard <strong>and</strong> eight polished thin sections.<br />

42


Figure 2.2. Left-h<strong>and</strong> photo shows an outcrop <strong>of</strong> widely spaced, relatively planar <strong>and</strong><br />

parallel compaction b<strong>and</strong>s along the northeast flank <strong>of</strong> Silica Dome. Arrows indicate<br />

opposite tips <strong>of</strong> a single b<strong>and</strong> 62 m long <strong>and</strong> up to 15 mm thick (illusory gaps in b<strong>and</strong><br />

continuity are due to outcrop topography <strong>and</strong> breaks in the telltale fin). A total <strong>of</strong> 16 tipto-tip<br />

thickness pr<strong>of</strong>iles were measured using a steel tape <strong>and</strong> calipers. Right-h<strong>and</strong> photo<br />

illustrates that, even when closely spaced, compaction b<strong>and</strong>s in this locale tend to remain<br />

planar (arrow indicates tip).<br />

43


Hue-based image analysis using MATLAB® was performed on more than 3,500 color<br />

pictures taken from the st<strong>and</strong>ard sections with a digital camera mounted on a<br />

petrographic microscope. Intensity-based image analysis was performed on over 200<br />

grayscale digital pictures collected with a scanning electron microscope (SEM). The<br />

particular SEM techniques used included backscatter electron imaging (BEI) <strong>and</strong><br />

cathodoluminescence.<br />

The data acquired from the field <strong>and</strong> petrographic methods were used to inform three<br />

mechanical assessments <strong>of</strong> the idealized anticrack-inclusion conceptual model.<br />

Comparisons <strong>of</strong> the near-tip states <strong>of</strong> stress calculated with each <strong>of</strong> these continuum<br />

approaches constitute the heart <strong>of</strong> this paper, in which we use the convention <strong>of</strong><br />

compression, closing-mode displacement discontinuity <strong>and</strong> volume-reduction strain as<br />

positive.<br />

The first approach considers the special case <strong>of</strong> an oblate ellipsoidal heterogeneity in<br />

the limit as its aspect ratio goes to zero, using the embedded layer analysis <strong>of</strong> Cocco <strong>and</strong><br />

Rice (2002) to construct algebraic expressions for the state <strong>of</strong> stress immediately ahead <strong>of</strong><br />

the tip. The second approach employs an exact, closed-form solution to the general three-<br />

dimensional Eshelby problem (Eshelby, 1957; Eshelby, 1959; Mura, 1987) for the oblate<br />

ellipsoidal heterogeneity to calculate how the state <strong>of</strong> stress varies with distance from a<br />

model CB matching the idealized physical parameters <strong>of</strong> the field-based conceptual<br />

model. Finally, we turn to a numerical, two-dimensional BEM code in MATLAB® to<br />

investigate the state <strong>of</strong> stress induced by an anticrack representation <strong>of</strong> the CB using only<br />

the trace length <strong>and</strong> effective closing-mode displacement data.<br />

4. Field <strong>and</strong> petrographic analysis<br />

4.1. Geological setting <strong>and</strong> paleostress state<br />

The Aztec s<strong>and</strong>stone, deposited during early to middle Jurassic time in a back-arc<br />

basin setting, comprises the western edge <strong>of</strong> a continuous æolian system that blanketed<br />

much <strong>of</strong> the intermountain southwest <strong>and</strong> includes the Navajo <strong>and</strong> Nugget s<strong>and</strong>stones<br />

(Blakey, 1989; Marzolf, 1983). In the Valley <strong>of</strong> Fire area, the 1,400-m-thick Aztec is<br />

directly <strong>and</strong> unconformably overlain by some 1,300 m <strong>of</strong> Cretaceous sediments, which<br />

comprise a generally upward-coarsening sequence <strong>of</strong> forel<strong>and</strong> basin deposits derived<br />

44


from Aztec highl<strong>and</strong>s riding atop the two closest Sevier thrust sheets to the west. The<br />

smaller <strong>and</strong> easternmost <strong>of</strong> these—the Willow Tank thrust—placed lower Aztec on upper<br />

Aztec <strong>and</strong> as much as the first 600 m <strong>of</strong> Cretaceous deposits. The Willow Tank thrust<br />

sheet subsequently was overridden by the regionally extensive, kilometers-thick Muddy<br />

Mountain thrust, which placed Palezoic carbonates over Aztec <strong>and</strong> provided the source<br />

for the upper Cretaceous units deposited atop the eastern reaches <strong>of</strong> the lower thrust sheet<br />

(Armstrong, 1968; Bohannon, 1983; Brock <strong>and</strong> Engleder, 1979; Carpenter <strong>and</strong> Carpenter,<br />

1994; Longwell, 1949).<br />

Silica Dome is situated approximately 800 m above the bottom <strong>of</strong> the Aztec, at the<br />

lower end <strong>of</strong> the CB-rich upper half <strong>of</strong> the formation <strong>and</strong> just above a regional, sub-<br />

horizontal alteration front separating uniformly red, hematite-stained s<strong>and</strong>stone below<br />

from bleached s<strong>and</strong>stone above (Taylor <strong>and</strong> Pollard, 2000). The hematite staining is<br />

interpreted as syndepositional, while CBs are observed to be older than the bleaching<br />

front, which represents the lower extent <strong>of</strong> an alteration event attributed to the upward<br />

<strong>and</strong> eastward expulsion <strong>of</strong> reducing basinal brines by the advancing Muddy Mountain<br />

thrust sheet (Eichhubl et al., 2004). This fluid-flow event, along with at least one<br />

subsequent episode, is responsible for the vividly colorful patterns <strong>of</strong> iron mineralization<br />

for which the Valley <strong>of</strong> Fire is justly named. Eichhubl et al. (2004) cite a preponderance<br />

<strong>of</strong> evidence to argue that thrust emplacement stopped just west (Muddy Mountain thrust)<br />

<strong>and</strong> northwest (Willow Tank thrust) <strong>of</strong> more than 20 km 2 <strong>of</strong> Aztec outcrops, which now<br />

comprise the heart <strong>of</strong> the Valley <strong>of</strong> Fire <strong>and</strong> include the study area. They also present<br />

diagenetic evidence from the base <strong>of</strong> the Aztec indicating that it has never been buried<br />

much deeper than 3 km, a depth which matches the current combined thickness <strong>of</strong> Aztec<br />

<strong>and</strong> Cretaceous to lower Miocene deposits preserved in the area.<br />

Vertical loading <strong>of</strong> the Aztec during CB formation in the area thus appears to have<br />

been due solely to synorogenic deposition. A strong constraint on the upper limit <strong>of</strong><br />

burial can be deduced from the relative timing <strong>of</strong> shear b<strong>and</strong> formation. Everywhere they<br />

occur in the Aztec s<strong>and</strong>stone, relatively low-angle shear b<strong>and</strong>s <strong>of</strong> the second phase <strong>of</strong><br />

deformation <strong>of</strong>fset <strong>and</strong> thus postdate CBs (Flodin <strong>and</strong> Aydin, 2004a; Hill, 1989; Taylor<br />

<strong>and</strong> Pollard, 2000). The dominant orientation <strong>and</strong> consistent top-to-the-east sense <strong>of</strong> shear<br />

displacement on these b<strong>and</strong>s mimics that <strong>of</strong> the overlying Willow Tank thrust, suggesting<br />

45


a direct genetic relationship (Hill, 1989). In fact, we observe that phase-two shear b<strong>and</strong>s<br />

extend into the Cretaceous deposits exposed directly beneath the thrust, while phase-one<br />

CBs are confined to the Aztec s<strong>and</strong>stone <strong>and</strong> all types <strong>of</strong> deformation b<strong>and</strong>s are absent<br />

higher in the Cretaceous section. From these observations we infer that CB formation<br />

predated deposition <strong>of</strong> at least the upper 700 m <strong>of</strong> Cretaceous sediments.<br />

We conclude that the depth <strong>of</strong> Silica Dome at the time <strong>of</strong> CB formation ranged from<br />

600 m (thickness <strong>of</strong> overlying Aztec s<strong>and</strong>stone) to 1,200 m (Aztec plus up to 600 m <strong>of</strong><br />

additional Cretaceous deposits). This corresponds to a range in vertical compressive<br />

stress from about 13 MPa to 27 MPa, given an average overburden density <strong>of</strong> 2.25 g/cm 3 .<br />

Assuming that well-drained, near-surface water-table conditions prevailed, as suggested<br />

by geologic, diagenetic <strong>and</strong> paleoclimatic evidence (Eichhubl et al., 2004; Marzolf,<br />

1983), pore pressure would have ranged from less than 6 MPa up to 12 MPa.<br />

Calculations based on Coulomb critical failure criteria (Zoback <strong>and</strong> Healy, 1984) <strong>and</strong><br />

born out by a catalog <strong>of</strong> insitu stress measurements (Townend <strong>and</strong> Zoback, 2000) indicate<br />

that the ratio <strong>of</strong> maximum to minimum regional principal stress does not exceed about<br />

two under hydrostatic pore pressure conditions. Therefore, taking the Andersonian view<br />

(Anderson, 1951) that the minimum principal stress is vertical in thrust-faulting<br />

environments, both the maximum remote horizontal (tectonic) stress <strong>and</strong> the minimum<br />

horizontal stress could have ranged anywhere from 13 MPa up to 54 MPa.<br />

For the purposes <strong>of</strong> the mechanical analyses, we posit a “best estimate” paleo stress<br />

state (σ1 > σ2 > σ3) for Silica Dome <strong>of</strong> 40 MPa, 30 MPa <strong>and</strong> 20 MPa. The value <strong>of</strong> σ3<br />

corresponds to an intermediate burial depth <strong>of</strong> 900 m (600 m <strong>of</strong> Aztec plus 300 m<br />

overlying Cretaceous deposits). The value <strong>of</strong> σ1 corresponds to the implied limit <strong>of</strong> 2σ3<br />

for active thrust-fault environments, <strong>and</strong> is presumed to have acted parallel to the east-<br />

vergent tectonic transport direction <strong>and</strong> perpendicular to the dominant northerly CB<br />

trend. The value <strong>of</strong> σ2 is simply taken as the average <strong>of</strong> σ1 <strong>and</strong> σ3. The mean (lithostatic)<br />

stress for this scenario is 30 MPa.<br />

4.2. Outcrop analysis<br />

Viewed individually at Silica Dome <strong>and</strong> elsewhere, CBs are seen to comprise sharply<br />

delineated tabular, bounded, <strong>and</strong> grossly penny-shaped structures that tend to weather out<br />

in positive relief as distinctive fins. They are generally between 1 cm <strong>and</strong> 2 cm thick in<br />

46


the middle <strong>and</strong> tens <strong>of</strong> meters in planar extent. Spacing between CBs ranges from<br />

centimeters to more than a meter (Figure 2.1). Truly isolated CBs are rare <strong>and</strong> close<br />

interactions between adjacent b<strong>and</strong>s are common. Frequently though, as at Silica Dome,<br />

adjacent CB traces remain markedly straight <strong>and</strong> parallel even in close proximity (Figure<br />

2.2). The appearance <strong>of</strong> CBs in outcrop, individually <strong>and</strong> in aggregate, is reminiscent <strong>of</strong><br />

distributed opening-mode fractures, <strong>and</strong> CBs might easily be mistaken for veins or s<strong>and</strong><br />

dikes by the casual observer. Closer inspection, however, reveals that they are composed<br />

<strong>of</strong> the same detrital material as the surrounding s<strong>and</strong>stone. In fact, depositional bedding is<br />

commonly preserved across the b<strong>and</strong>s with no detectable shear <strong>of</strong>fset (Figure 2.3).<br />

Sedimentary architecture in the Aztec is observed to affect CB distribution (Eichhubl<br />

et al., 2004; Hill, 1989; Sternl<strong>of</strong> et al., 2004). B<strong>and</strong>s <strong>and</strong> patterns <strong>of</strong> b<strong>and</strong>s frequently<br />

warp <strong>and</strong>/or terminate at or near cross-bed boundaries, <strong>and</strong> almost always terminate at<br />

major inter-dune contacts. Also, while many if not most dune packages contain abundant<br />

CBs, some do not.<br />

Thickness pr<strong>of</strong>iles <strong>of</strong> 16 tip-to-tip individual CB traces ranging from one to 62 meters<br />

long in outcrop reveal a characteristic, approximately elliptical shape (Figure 2.4a).<br />

While midpoint thickness for these pr<strong>of</strong>iles generally increases with increasing trace<br />

length, it does so along a decreasing trend that suggests a plateau <strong>of</strong> maximum attainable<br />

thickness independent <strong>of</strong> trace length (Figure 2.4b). In order to capture the typical<br />

characteristics <strong>of</strong> a well-developed CB, we focused on one particularly isolated <strong>and</strong><br />

symmetric outcrop trace <strong>of</strong> sufficient length <strong>and</strong> thickness that it could reasonably be<br />

assumed to represent a sub-horizontal pr<strong>of</strong>ile through the middle <strong>of</strong> the structure. This<br />

CB trace, designated CB-A (Figure 2.4c), is 24.75 m in length with a maximum midpoint<br />

thickness <strong>of</strong> 11.4 mm, yielding an aspect ratio (midpoint thickness/trace length) <strong>of</strong> 4.6 x<br />

10 -4 <strong>and</strong> an eccentricity <strong>of</strong> 0.99. Represented as an elliptical pr<strong>of</strong>ile fit by the least<br />

squares method (correlation index = 0.87), the midpoint thickness drops to 9.38 mm,<br />

thereby slightly decreasing the aspect ratio <strong>and</strong> increasing the eccentricity (Figure 2.4c).<br />

While this dimensional data was gathered on relatively low-angle outcrop faces, steep<br />

cliffs in the vicinity <strong>of</strong> Silica Dome <strong>and</strong> elsewhere in the Valley <strong>of</strong> Fire reveal that CBs<br />

can extend at least 10 m vertically. Coupled with the fact that they occur throughout a<br />

700-m-thick section <strong>of</strong> the Aztec s<strong>and</strong>stone, this observation supports what intuition<br />

47


Compaction b<strong>and</strong> fin<br />

Depositional<br />

bedding<br />

Cross-bed boundary<br />

Figure 2.3. Close-up <strong>of</strong> a typical, well-developed compaction b<strong>and</strong> fin in outcrop. Note<br />

that depositional bedding extends relatively undisturbed across the b<strong>and</strong>, <strong>and</strong> is clearly<br />

visible on the fin.<br />

48


Normalized thickness<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

-1.0 -0.5 0.0 0.5 1.0<br />

(a) Distance from center normalized by trace half-length<br />

Midpoint thickness (mm)<br />

(b)<br />

Thickness (mm)<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 5 10 15 20 25<br />

Trace half-length (m)<br />

30 35<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

(c)<br />

-1 -0.5 0 0.5<br />

Distance from center normalized by trace half-length<br />

1<br />

Figure 2.4. (a) Thickness pr<strong>of</strong>ile data for 16 tip-to-tip compaction b<strong>and</strong> traces (>1,700<br />

measurements). Position is normalized relative to center by the b<strong>and</strong> half length <strong>and</strong><br />

thickness is normalized by the midpoint value. The curved line represents an ellipse<br />

similarly normalized. The correlation index <strong>of</strong> the data to the ellipse is 0.79. (b) Plot <strong>of</strong><br />

midpoint thickness versus trace half-length for the 16 compaction b<strong>and</strong> pr<strong>of</strong>iles, star<br />

indicates CB shown in (c). Thickness generally increases with increasing length, but<br />

along a decreasing trend. (c) Thickness pr<strong>of</strong>ile for CB-A, the 24.75-m-long compaction<br />

b<strong>and</strong> trace featured in this paper. The curved line represents the least-squares best fit <strong>of</strong><br />

an ellipse to the data. The correlation index for this fit is 0.87.<br />

49


suggests—that to first approximation CBs are grossly penny shaped. Certainly, thickness<br />

is generally several orders <strong>of</strong> magnitude less than any in-plane, tip-to-tip dimension.<br />

4.3. Petrographic analysis<br />

Viewed in st<strong>and</strong>ard thin section, CB-A is every bit as sharply defined by its decreased<br />

porosity as it is in outcrop by its characteristic fin, <strong>and</strong> thickness measurements made<br />

under the microscope closely match those made with calipers in the field. Particularly<br />

striking is the absence <strong>of</strong> any well-developed <strong>structural</strong> fabric within the b<strong>and</strong>, <strong>and</strong> the<br />

degree to which depositional bedding is preserved within <strong>and</strong> across it (Figure 2.5). In<br />

fact, nowhere did we find appreciable evidence <strong>of</strong> bedding <strong>of</strong>fset across the b<strong>and</strong>, despite<br />

considerable natural variability in the relative orientation <strong>of</strong> bedding to b<strong>and</strong> between<br />

sample locations. Neither did we find any conclusive evidence <strong>of</strong> systematic bedding<br />

thickness variations within the b<strong>and</strong>. Together these observations indicate that the<br />

inelastic strain accommodated within CB-A as porosity-loss compaction was<br />

predominantly uniaxial <strong>and</strong> directed parallel to its shortest dimension (i.e. perpendicular<br />

to its trace).<br />

Despite the coarseness inherent to porosity measurements, particularly using two-<br />

dimensional image analysis <strong>of</strong> granular materials, pr<strong>of</strong>iles across thicker parts <strong>of</strong> CB-A<br />

clearly reveal the abrupt drop in porosity that defines the b<strong>and</strong> (Figure 2.6a). Mean<br />

porosity in the s<strong>and</strong>stone outside the b<strong>and</strong> was measured to be 24.5%, with a st<strong>and</strong>ard<br />

deviation <strong>of</strong> 2.9%. Mean porosity inside the b<strong>and</strong> was measured to be 11.9%, with a<br />

st<strong>and</strong>ard deviation <strong>of</strong> 3.0%. Toward the tip, the pr<strong>of</strong>iles indicate the absence <strong>of</strong> any<br />

obvious near-tip process zone recognizable as spatial variations in porosity (Figure 2.6b).<br />

Backscatter electron images (BEI) provide much greater definition than the st<strong>and</strong>ard<br />

color digital images. In particular, micro-porosity due to grain damage can be<br />

distinguished <strong>and</strong> clay, which absorbs epoxy <strong>and</strong> so looks blue in st<strong>and</strong>ard sections, can<br />

be differentiated (Figure 2.7). The clay occurs primarily as undeformed grain-bridging<br />

<strong>and</strong> pore-filling cements, both inside <strong>and</strong> outside the CB. Clay also frequently in-fills<br />

cracked grains. These observations indicate that clay should be ignored when determining<br />

the distribution <strong>of</strong> porosity related to mechanical compaction. Image analysis reveals that<br />

the Aztec averages about 4% clay by volume, while thicker parts <strong>of</strong> the b<strong>and</strong> <strong>and</strong> the<br />

s<strong>and</strong>stone immediately adjacent to it average about 7%. This preferential accumulation <strong>of</strong><br />

50


5 mm<br />

Figure 2.5. Composite photomicrograph <strong>of</strong> compaction b<strong>and</strong> sampled 6.0 meters from<br />

the tip, where it is about 9 mm thick (dotted black lines). Blue indicates epoxy-filled pore<br />

space (~ 25% outside the b<strong>and</strong>, ~ 10% inside). The plane <strong>of</strong> the section is vertical <strong>and</strong><br />

orthogonal to the trace <strong>of</strong> the b<strong>and</strong> in outcrop. This is a mosaic <strong>of</strong> images taken with<br />

plane-polarized transmitted light at 25X magnification using a digital camera mounted to<br />

a binocular microscope. Mineral constituents identifiable in this image are: detrital quartz<br />

(white to gray), detrital orthoclase (stained dark brown), <strong>and</strong> sparse hematite (black). As<br />

is typical in foreset dune deposits, bedding generally consists <strong>of</strong> alternating layers <strong>of</strong><br />

coarser <strong>and</strong> finer grains, here accentuated by the relative concentration <strong>of</strong> stained<br />

orthoclase in the finer-grained beds. Note that bedding is preserved, perhaps even<br />

visually enhanced within the b<strong>and</strong>. The three black arrows indicate a coarse-fine-coarse<br />

bedding package as it passes through the b<strong>and</strong> at high angle with no apparent change in<br />

thickness. The preservation <strong>of</strong> bed thickness is observed consistently in all thin sections,<br />

including those taken across the b<strong>and</strong> <strong>and</strong> orthogonal to the orientation shown above at<br />

four locations. This indicates that compaction strain within the b<strong>and</strong> is predominantly<br />

uniaxial <strong>and</strong> oriented parallel to the b<strong>and</strong> normal direction. Natural, mm-scale variations<br />

in bed thickness render more exact interpretations <strong>of</strong> bed distortion within the b<strong>and</strong><br />

inconclusive.<br />

51


Porosity<br />

Porosity<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

0 5 10 15 20 25 30<br />

(a)<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

Distance along pr<strong>of</strong>ile (mm)<br />

0<br />

0 5 10 15 20 25 30<br />

(b)<br />

Distance along pr<strong>of</strong>ile (mm)<br />

Figure 2.6. Porosity pr<strong>of</strong>iles taken orthogonally across the compaction b<strong>and</strong> 8.7 m from<br />

the tip (a) <strong>and</strong> at the tip (b). Each box plot represents a composite <strong>of</strong> 10 porosity<br />

measurements from immediately adjacent transects. The horizontal lines comprising the<br />

boxes indicate lower quartile, median <strong>and</strong> upper quartile values <strong>of</strong> the group. The<br />

whiskers extend up to 1.5 times the interquartile range to indicate the spread <strong>of</strong> additional<br />

data points. Outliers are indicated by “+”. A box plot representing the range <strong>of</strong><br />

background matrix porosity (minus outliers) appears to the right <strong>of</strong> each pr<strong>of</strong>ile. At its<br />

middle (a), the b<strong>and</strong> is clearly defined as a sharp, statistically significant drop in porosity.<br />

At its tip (b), however, while still visually apparent at a fraction <strong>of</strong> a millimeter thick, the<br />

b<strong>and</strong> can no longer be detected as a measurable reduction in porosity.<br />

52


Middle<br />

500 µm<br />

B<strong>and</strong> trend Tip<br />

Figure 2.7. Compositional electron backscatter images <strong>of</strong> a thick part <strong>of</strong> the compaction<br />

b<strong>and</strong> (top), adjacent s<strong>and</strong>stone less than 2 mm outside the b<strong>and</strong> (middle), <strong>and</strong> s<strong>and</strong>stone<br />

more than 1 m away from the nearest b<strong>and</strong> (bottom). Porosity is black, quartz is medium<br />

gray, feldspar is white, <strong>and</strong> clay (kaolinite) shows as clusters <strong>of</strong> dark gray speckles<br />

around the detrital grains. Note the dominant texture <strong>of</strong> micro-fracture accommodated<br />

plasticity inside the compaction b<strong>and</strong> versus the relative paucity <strong>of</strong> grain damage<br />

everywhere outside, even immediately adjacent to the b<strong>and</strong>.<br />

53


pore-clogging clay—due presumably to groundwater filtering by the compacted b<strong>and</strong><br />

coupled with mechanically enhanced feldspar degradation—at least partially explains the<br />

greater resistance to weathering exhibited by CBs, as well as their strong impact on fluid<br />

flow. That clay remains the primary cement in the weakly indurated Aztec today further<br />

suggests that cementation was light to nonexistent during CB formation.<br />

Six sets <strong>of</strong> backscatter images were collected at locations ranging from the middle to<br />

the tip <strong>of</strong> CB-A. Each set comprises images taken from inside <strong>and</strong> immediately outside<br />

the b<strong>and</strong> within the same depositional layer(s). Porosity-reduction volume loss related to<br />

CB formation at these locations ranged from 9.4% to 11.7% (Figure 2.8). These results<br />

suggest that mechanical compaction throughout the b<strong>and</strong> is uniform at about 10%,<br />

regardless <strong>of</strong> the local variations in s<strong>and</strong>stone porosity through which it propagated.<br />

Cathodoluminescence imaging indicates that redistribution <strong>of</strong> quartz via pressure solution<br />

is volumetrically insignificant <strong>and</strong> has not appreciably contributed to the spatial<br />

distribution <strong>of</strong> porosity now observed. This is not surprising given the shallow burial<br />

history <strong>of</strong> the study area (< 2 km), which, for any reasonable estimate <strong>of</strong> geothermal<br />

gradient, would have made quartz pressure solution an extremely slow to nonexistent<br />

process.<br />

Cracking <strong>of</strong> quartz grains at point contacts, due both to overburden <strong>and</strong> tectonic<br />

loading, is pervasive throughout the thin sections. Wholesale micro-fracture-<br />

accommodated plastic deformation <strong>of</strong> quartz grains, however, comprises the dominant<br />

micro-<strong>structural</strong> characteristic <strong>of</strong> the b<strong>and</strong>, <strong>and</strong> is the obvious mechanism by which<br />

granular rearrangement <strong>and</strong> porosity loss compaction was accommodated (Figure 2.9).<br />

Roughly half <strong>of</strong> the quartz grains within the b<strong>and</strong> exhibit plasticity, creating the hint <strong>of</strong> a<br />

mild shape fabric <strong>of</strong> deformed grains elongated parallel to the b<strong>and</strong> trace (i.e. orthogonal<br />

to the inferred direction <strong>of</strong> maximum compression). Far more striking is the relative<br />

absence <strong>of</strong> granular disaggregation, despite <strong>of</strong>ten intense micro-fracturing <strong>and</strong> distortion.<br />

Taken together, these observations corroborate the field interpretation <strong>of</strong> the b<strong>and</strong> as<br />

representing nearly pure uniaxial compaction with little or no shear, <strong>and</strong> suggest a stable,<br />

quasi-static process <strong>of</strong> plastic collapse.<br />

Isolated examples <strong>of</strong> grain plasticity <strong>and</strong> collapse can also be found outside the b<strong>and</strong><br />

(Figure 2.10). We interpret these as incipient CBs, representing an early stage <strong>of</strong><br />

54


0.30<br />

0.25<br />

0.20<br />

0.15<br />

0.10<br />

Porosity 0.35<br />

0.05<br />

0.00<br />

0.1 0.3 0.6 1.8 4.2 8.7<br />

Distance from tip (m)<br />

Figure 2.8. Clustered bar chart showing the distribution <strong>of</strong> porosity inside (middle bar)<br />

<strong>and</strong> immediately outside (left bar) the compaction b<strong>and</strong> with distance from the tip (clay<br />

included as porosity. The difference between (right bar) gives the relative volume strain<br />

attributable to mechanical compaction at each location. This compaction strain is<br />

remarkably uniform throughout the b<strong>and</strong> at about 10%.<br />

Figure 2.10. Electron backscatter image <strong>of</strong> localized quartz-grain plasticity <strong>and</strong> collapse<br />

(interlocked grains at center <strong>of</strong> image) that is not part <strong>of</strong> a well-developed compaction<br />

b<strong>and</strong>. Similar grain-scale features appear scattered throughout the s<strong>and</strong>stone <strong>and</strong> can be<br />

interpreted as incipient compaction b<strong>and</strong>s. Black is porosity, medium gray is quartz,<br />

white is feldspar.<br />

55


500 µm<br />

Figure 2.9. Electron backscatter images <strong>of</strong> micr<strong>of</strong>racture-accommodated quartz grain<br />

plasticity <strong>and</strong> interpenetration, which comprises the dominant mode <strong>of</strong> grain-scale<br />

deformation inside the compaction b<strong>and</strong> (white box in upper image indicates area <strong>of</strong><br />

detail shown in lower image). Despite intense grain damage <strong>and</strong> the absence <strong>of</strong><br />

appreciable quartz recrystallization, very little disaggregation <strong>of</strong> detrital grains is<br />

apparent.<br />

56


compaction around an original Griffith-type flaw—weak grain, irregular pore, etc.—<br />

which for any variety <strong>of</strong> reasons failed to develop further. Similar textural evidence <strong>of</strong><br />

incipient grain damage <strong>and</strong> collapse immediately around the b<strong>and</strong> <strong>and</strong> at the tip, however,<br />

is notably absent. Opening-mode fractures <strong>and</strong> faults typically exhibit a process zone <strong>of</strong><br />

transitional deformation that grades into the surrounding undamaged host material,<br />

particularly around the tip where induced stress concentrations are highest (Hoagl<strong>and</strong> et<br />

al., 1973; Peck et al., 1985). Despite considerable effort, we have as yet been unable<br />

positively to identify any analogous process zone around the b<strong>and</strong>, expressed either as<br />

grain damage or porosity variations.<br />

5. Anticrack-inclusion conceptual model<br />

Based on the observations, data <strong>and</strong> constraints presented above, we propose an<br />

idealized conceptual model <strong>of</strong> isolated compaction b<strong>and</strong>s in porous s<strong>and</strong>stone as highly<br />

eccentric, roughly axisymmetric ellipsoidal features <strong>of</strong> sharply defined inelastic porosity-<br />

loss compaction, which form in response to <strong>and</strong> generally perpendicular to the maximum<br />

remote compressive tectonic stress (Figure 2.11a). We further suggest that CBs initiate<br />

with the collapse <strong>of</strong> Griffith-type grain-scale flaws—weak grains, irregular pores, etc.—<br />

<strong>and</strong> then propagate outward to form tabular inclusions <strong>of</strong> nearly uniform uniaxial plastic<br />

strain <strong>and</strong> differing elastic moduli embedded within a surrounding elastic medium. This<br />

conceptual model is well represented mechanically as a contractile Eshelby inclusion<br />

(Figure 2.11b), which would generate near-tip compressive stress concentrations in the<br />

surrounding material consistent with self-sustaining in-plane <strong>propagation</strong>. Inside any<br />

ellipsoidal inclusion embedded within an infinite medium, the state <strong>of</strong> stress <strong>and</strong> strain is<br />

uniform (Eshelby, 1957).<br />

The combination <strong>of</strong> extreme aspect ratio (~10 -4 ) with significant uniaxial plastic<br />

strain (0.1) further suggests a close mechanical approximation <strong>of</strong> the idealized CB as an<br />

anticrack—that is, a sharp boundary across which a continuous distribution <strong>of</strong> closing-<br />

mode displacement discontinuity has occurred (Figure 2.11c). This virtual anticrack<br />

interpretation for compaction b<strong>and</strong>s is analogous to that <strong>of</strong> pressure solution surfaces<br />

(Fletcher <strong>and</strong> Pollard, 1981; Mollema <strong>and</strong> Antonellini, 1996), except that porosity loss<br />

takes the place <strong>of</strong> material dissolution <strong>and</strong> transport. In neither case does physical<br />

57


(a)<br />

(b)<br />

σ 3<br />

σ 1<br />

x 3<br />

x 1<br />

x 1<br />

(c) u 1<br />

Figure 2.11. Schematic representations <strong>of</strong> the idealized compaction b<strong>and</strong> model. (a)<br />

Axisymmetric geometry <strong>of</strong> the eccentric ellipsoidal b<strong>and</strong> aligned with the principal<br />

remote stresses. (b) Cross-sectional area <strong>of</strong> the b<strong>and</strong> (solid ellipse) relative to the precompacted<br />

area originally occupied by the same detrital grains (dashed ellipse). As<br />

inelastic compaction progresses, the boundary around the grains involved contracts as<br />

indicated by the displacement arrows (u1). This inward displacement <strong>of</strong> the elliptical<br />

boundary corresponds to the uniform uniaxial plastic strain <strong>of</strong> an Eshelby inclusion, <strong>and</strong><br />

the area between the dashed <strong>and</strong> solid ellipses corresponds to the volume loss associated<br />

with the compaction. (c) Two-dimensional anticrack representation <strong>of</strong> the model b<strong>and</strong> as<br />

an elliptical distribution <strong>of</strong> closing-mode displacement discontinuity equivalent to the<br />

uniform Eshelby compaction strain. In this virtual treatment, two material lines (solid<br />

lines) interpenetrate by an amount equivalent to the volume loss associated with the<br />

compaction (dashed ellipse) as shown by the displacement arrows (u1). Actual<br />

interpenetration does not occur. Because <strong>of</strong> its extreme eccentricity, however, solutions<br />

for the state <strong>of</strong> stress induced around the model b<strong>and</strong> using the Eshelby <strong>and</strong> anticrack<br />

approaches are substantially similar.<br />

58<br />

u 1<br />

x 2<br />

x 2<br />

u 1<br />

σ 2


interpenetration occur, rather the boundary demarcating volume loss migrates outward<br />

<strong>and</strong> into the surrounding material.<br />

Central to the anticrack-inclusion concept, <strong>and</strong> the mechanical analyses to follow, is<br />

the prescription <strong>of</strong> the model CB as an isolated feature <strong>of</strong> nonlinear inelastic strain that<br />

evolves quasi-statically in an infinite, homogeneous, isotropic, linear elastic continuum<br />

subject to uniform remote loading. All aspects <strong>of</strong> this model prescription warrant<br />

comment. Firstly, given the regional tectonic nature <strong>of</strong> the compression <strong>and</strong> a study area<br />

situated in the midst <strong>of</strong> a 1,400-m-thick s<strong>and</strong>stone deposit, the approximation <strong>of</strong> uniform<br />

remote stresses applied to an infinite material is reasonable. Given a ratio <strong>of</strong> trace length<br />

to spacing generally less than 0.1, it would be difficult to argue that any CB in the Aztec<br />

s<strong>and</strong>stone is truly isolated in a mechanical sense. Nonetheless, the specific field data on<br />

which the conceptual model is based comes from planar b<strong>and</strong>s that betray little reaction<br />

to their nearest neighbors.<br />

As with any granular material, the applicability <strong>of</strong> homogeneous continuity is scale<br />

dependent. In the Aztec s<strong>and</strong>stone, with an average grain diameter <strong>of</strong> 0.25 mm, this<br />

becomes reasonable at the cm-scale (Amadei <strong>and</strong> Stephansson, 1997), which represents a<br />

lower limit <strong>of</strong> resolution for interpreting the mechanical modeling results. To focus on<br />

grain-scale processes inside <strong>and</strong> immediately outside a CB, the distinct element method<br />

approach would be appropriate (Antonellini <strong>and</strong> Pollard, 1995; Morgan, 1999; Morgan<br />

<strong>and</strong> Boettcher, 1999). Also problematic at a larger scale is the application <strong>of</strong><br />

homogeneous isotropy to the complex æolian sedimentary architecture <strong>of</strong> the Aztec.<br />

Major dune boundaries influence CBs, so this conceptual model is based on data from an<br />

outcrop <strong>of</strong> b<strong>and</strong>s located within a single dune package. Given that CBs commonly cut<br />

across depositional bedding without apparent effect, we interpret the mechanical<br />

influence <strong>of</strong> such layering as negligible. Finally, the assumption <strong>of</strong> quasi-static CB<br />

<strong>propagation</strong> is based on the coherent nature <strong>of</strong> plastic quartz grain deformation within the<br />

b<strong>and</strong>s, which suggests stable <strong>propagation</strong> accommodated by slow, visco-elastic relaxation<br />

(Chester et al., 2004; Karner et al., 2003). By the same token, the paucity <strong>of</strong> plastic<br />

deformation outside the b<strong>and</strong>s suggests predominantly elastic behavior through time.<br />

59


6. Elastic properties<br />

Despite a long history <strong>of</strong> diagenesis <strong>and</strong> deformation, analysis <strong>of</strong> the Aztec s<strong>and</strong>stone<br />

reveals a material arguably similar in terms <strong>of</strong> depositional structure <strong>and</strong> cementation to<br />

what it was during CB formation—a weakly cemented, cross-bedded, porous quartz<br />

s<strong>and</strong>stone. That autochthonous clay derived from the degradation <strong>of</strong> sparse feldspar has<br />

replaced thin hematite grain coatings as the primary cement is unlikely to have<br />

significantly altered the bulk material properties <strong>of</strong> the Aztec’s close-packed quartz-grain<br />

skeleton. We therefore infer that the elastic properties <strong>of</strong> the paleo Aztec resemble those<br />

<strong>of</strong> the present day.<br />

Published values for apparent Young’s modulus in s<strong>and</strong>stones based on laboratory<br />

testing range from 10 GPa to 46 GPa (mean <strong>of</strong> 22 GPa), while apparent Poisson’s ratio<br />

ranges from 0.1 to 0.4 (mean <strong>of</strong> 0.24) (Bieniawski, 1984). We have calculated values <strong>of</strong><br />

apparent Young’s modulus for samples collected near Silica Dome from the results <strong>of</strong><br />

triaxial compression tests conducted at S<strong>and</strong>ia National Laboratories. The results were<br />

16.5 GPa <strong>and</strong> 21.0 GPa at confining pressures <strong>of</strong> 10 MPa <strong>and</strong> 50 MPa, respectively.<br />

Given the assumed lithostatic paleostress <strong>of</strong> 30 MPa, we therefore posit a Young’s<br />

modulus <strong>of</strong> 20 GPa <strong>and</strong> a Poisson’s ratio <strong>of</strong> 0.2 for the Aztec s<strong>and</strong>stone at Silica Dome<br />

during the period <strong>of</strong> CB formation.<br />

The CBs themselves, however, are a different matter. Preferential clay cementation <strong>of</strong><br />

the b<strong>and</strong>s subsequent to their formation has rendered them significantly more resistant<br />

<strong>and</strong>, presumably, stiffer than the friable host rock. Indeed, induration <strong>of</strong> the b<strong>and</strong>s can<br />

occasionally resemble that <strong>of</strong> quartzite, suggesting that local pressure-solution healing <strong>of</strong><br />

damaged grains may have occurred. Even if the current elastic properties <strong>of</strong> CBs could<br />

reliably be measured in the lab—a difficult challenge given their thin, tabular<br />

dimensions—we suggest that the results would not represent the b<strong>and</strong>s as they formed.<br />

Additionally, the texture <strong>of</strong> the b<strong>and</strong>s presents a paradox in that reduced porosity argues<br />

for increased shear stiffness, while intense micro-fracturing <strong>and</strong> effective grain-size<br />

reduction argues for increased shear compliance. It is reasonable to assume, however,<br />

that bulk modulus within CBs, which measures resistance to volume change, would<br />

increase as the result <strong>of</strong> both porosity loss <strong>and</strong> effective grain-size reduction. In the<br />

analyses that follow, we therefore consider a range <strong>of</strong> elastic properties for the model CB<br />

60


y fixing the ratio <strong>of</strong> b<strong>and</strong> to s<strong>and</strong>stone bulk moduli at 1.5. The interrelation <strong>of</strong> the<br />

essential elastic parameters—Young’s modulus (E), shear modulus (G), bulk modulus<br />

(K) <strong>and</strong> Poisson’s ratio (ν)—is given by E = 2G(1+ν) = 3K(1-2ν).<br />

7. Mechanical analysis<br />

A logical minimum requirement for in-plane <strong>propagation</strong> <strong>of</strong> the model CB is that<br />

compressive stress just ahead <strong>of</strong> the tip—whether the component normal to the b<strong>and</strong><br />

trace, the mean value, or some other measure—exceeds the remote value. If the<br />

compacted material inside the b<strong>and</strong> becomes elastically stiffer than the surroundings, the<br />

compressive stress at the tip will decrease. If the inside is s<strong>of</strong>ter, the compressive stress<br />

will increase. The other critical independent factor is the magnitude <strong>of</strong> the inelastic strain<br />

represented by the compaction, which would act to increase the magnitude <strong>of</strong> the near-tip<br />

compressive stress.<br />

In this section, we first use an embedded-layer model to develop these ideas in a<br />

quantitatively intuitive way, examining the relative influence <strong>of</strong> elastic moduli versus<br />

inelastic strain on the state <strong>of</strong> stress exactly at the tip <strong>of</strong> an infinitely thin b<strong>and</strong>. We then<br />

turn to a full analytical solution <strong>of</strong> the Eshelby problem to examine the state <strong>of</strong> stress<br />

induced around a b<strong>and</strong> matching the parameters <strong>of</strong> the field-based conceptual model.<br />

Finally, we compare the results obtained with the Eshelby solution to those derived from<br />

a two-dimensional (plane strain) BEM approximation <strong>of</strong> the model b<strong>and</strong> as an anticrack,<br />

using a distribution <strong>of</strong> closing-mode displacement discontinuity elements to represent the<br />

10% homogeneous uniaxial strain <strong>of</strong> the inclusion.<br />

7.1. Embedded layer model<br />

Expressions for the stress <strong>and</strong> strain fields in <strong>and</strong> around an Eshelby inclusion are<br />

given by Rudnicki (1999) in terms <strong>of</strong> the remote values, the elastic constants <strong>of</strong> the<br />

inclusion <strong>and</strong> surrounding material, <strong>and</strong> an array <strong>of</strong> factors depending on the geometry <strong>of</strong><br />

the inclusion <strong>and</strong> the Poisson’s ratio <strong>of</strong> the surroundings. He specifically treats the<br />

limiting case <strong>of</strong> interest here, where the aspect ratio <strong>of</strong> the inclusion goes to zero <strong>and</strong> its<br />

geometry approaches that <strong>of</strong> an embedded layer. The same result can be obtained more<br />

directly from the conditions given by Cocco <strong>and</strong> Rice (2002) for a planar fault zone.<br />

61


Comparison <strong>of</strong> the two approaches establishes the formal equivalence <strong>of</strong> the very<br />

eccentric ellipsoidal inclusion with the embedded layer.<br />

With the x1-axis normal to the layer—i.e. corresponding to the short axis <strong>of</strong> the<br />

inclusion (Figure 2.11a)—Cocco <strong>and</strong> Rice (2002) note that εij b = εij ∞ if neither i nor j is 1<br />

<strong>and</strong> σij b = σij ∞ if either i or j is 1, where the superscripts “b” <strong>and</strong> “∞” denote values <strong>of</strong><br />

stress (σ) <strong>and</strong> strain (ε) inside the model b<strong>and</strong> (layer/inclusion) <strong>and</strong> in the far field,<br />

respectively. The relations <strong>of</strong> particular relevance here are<br />

σ b<br />

ε b<br />

∞<br />

11 = σ 11<br />

∞<br />

22 = ε 22<br />

(1a)<br />

∞<br />

, ε ε<br />

(1b)<br />

b<br />

33 = 33<br />

Hooke’s law for the linear elastic relation between strain <strong>and</strong> stress is<br />

1 ⎧ ν ⎫<br />

ε ij = ⎨σ<br />

ij − σ kkδ<br />

ij ⎬<br />

2G<br />

⎩ 1+<br />

ν ⎭<br />

where the repeated index denotes summation, <strong>and</strong> δij =1 if i = j, <strong>and</strong> 0 if i ≠ j. The remote<br />

stresses <strong>and</strong> strains are related by (2). In the b<strong>and</strong>, (2) relates the stresses to the elastic<br />

strains. The total strain inside the b<strong>and</strong> is given by<br />

b total b elastic p<br />

[ ε ] [ ε ] + ε<br />

ij<br />

= (3)<br />

ij<br />

ij<br />

where εij p denotes the plastic strain representing compaction within the b<strong>and</strong>, which<br />

equals zero unless i = j = 1 as stipulated in the field-based conceptual model. Equations<br />

(1), (2) <strong>and</strong> (3) can be combined to eliminate the elastic strains <strong>and</strong> yield expressions for<br />

the stresses inside the model b<strong>and</strong> in terms <strong>of</strong> the far-field stresses <strong>and</strong> elastic constants:<br />

σ b<br />

σ<br />

∞<br />

11 = σ 11<br />

b<br />

22<br />

1+<br />

ν b =<br />

2<br />

G<br />

G<br />

b<br />

s<br />

( 1−ν<br />

)<br />

b<br />

⎧<br />

⎨<br />

⎩<br />

∞ ∞ ( σ −σ )<br />

22<br />

2<br />

( 1−ν<br />

s )<br />

( 1+<br />

ν )<br />

33<br />

s<br />

G<br />

G<br />

b<br />

s<br />

∞ ∞ ( σ + σ )<br />

22<br />

33<br />

∞ ⎡<br />

+ σ11⎢<br />

⎣<br />

62<br />

2ν<br />

b<br />

G<br />

−<br />

( 1+<br />

ν ) G ( 1+<br />

ν )<br />

b<br />

b<br />

s<br />

2ν<br />

s<br />

s<br />

⎤⎫<br />

⎥⎬<br />

+<br />

⎦⎭<br />

(2)<br />

(4a)<br />

(4b)


σ<br />

b<br />

33<br />

1+<br />

ν b =<br />

2<br />

G<br />

G<br />

b<br />

s<br />

( 1−ν<br />

)<br />

b<br />

⎧<br />

⎨<br />

⎩<br />

∞ ∞ ( σ −σ )<br />

22<br />

2<br />

( 1−ν<br />

s )<br />

( 1+<br />

ν )<br />

33<br />

s<br />

G<br />

G<br />

b<br />

s<br />

∞ ∞ ( σ + σ )<br />

22<br />

33<br />

∞ ⎡<br />

+ σ11⎢<br />

⎣<br />

2ν<br />

b<br />

G<br />

−<br />

( 1+<br />

ν ) G ( 1+<br />

ν )<br />

b<br />

b<br />

s<br />

2ν<br />

s<br />

s<br />

⎤⎫<br />

⎥⎬<br />

−<br />

⎦⎭<br />

where the subscripts “b” <strong>and</strong> “s” denote the elastic parameters <strong>of</strong> the b<strong>and</strong> <strong>and</strong><br />

surrounding s<strong>and</strong>stone, respectively.<br />

The state <strong>of</strong> stress at the point adjacent to <strong>and</strong> immediately outside the tip <strong>of</strong> the<br />

model b<strong>and</strong> along the x2-axis can now be obtained from the conditions <strong>of</strong> displacement<br />

<strong>and</strong> traction continuity at the interface:<br />

σ = σ<br />

(5a)<br />

t<br />

22<br />

t<br />

11<br />

b<br />

11<br />

b<br />

22<br />

t b<br />

ε = ε , ε = ε<br />

(5b)<br />

33<br />

33<br />

where the superscript “t” refers to the values just outside the tip. In the second equation<br />

<strong>of</strong> (5b), ε33 t can be replaced by ε33 ∞ from the second equation <strong>of</strong> (1b). Again, the elasticity<br />

relations can be used to eliminate the elastic strains <strong>and</strong> yield expressions for the stresses<br />

at the tip <strong>of</strong> the b<strong>and</strong> in terms <strong>of</strong> the uniform stresses inside, the elastic constants <strong>and</strong> the<br />

nonzero plastic strain component ε11 p :<br />

σ<br />

G<br />

ν<br />

⎛ G<br />

⎜ −<br />

⎝<br />

⎞<br />

⎟<br />

⎠<br />

G<br />

ν −ν<br />

t s b s b s s b s b<br />

s p<br />

11 = σ 11 + σ 22 1 σ kk<br />

ε11<br />

Gb<br />

( 1 ν s ) ⎜ G ⎟<br />

+<br />

+<br />

(6a)<br />

−<br />

b Gb<br />

( 1−ν<br />

s )( 1+<br />

ν b ) ( 1−ν<br />

s )<br />

σ = σ<br />

(6b)<br />

σ<br />

t<br />

22<br />

G<br />

b<br />

22<br />

ν<br />

⎛ G<br />

⎜ −<br />

⎝<br />

⎞<br />

⎟<br />

⎠<br />

G<br />

ν −ν<br />

t s b s b s s b s b<br />

s p<br />

33 = σ 33 + σ 22 1 σ kk<br />

ν sε11<br />

Gb<br />

( 1 ν s ) ⎜ G ⎟<br />

+<br />

+<br />

(6c)<br />

−<br />

b Gb<br />

( 1−ν<br />

s )( 1+<br />

ν b ) ( 1−ν<br />

s )<br />

By substituting (4) into (6), expressions for the stresses at the tip in terms <strong>of</strong> the<br />

remote values also can be derived. It can be seen from (6) that, for given values <strong>of</strong> the<br />

elastic parameters <strong>and</strong> remote stresses, σ11 t <strong>and</strong> σ33 t are linear functions <strong>of</strong> ε11 p , with<br />

slopes <strong>of</strong> 2Gs/(1-νs) <strong>and</strong> [2Gs/(1-νs)] νs, respectively, while σ22 t remains constant. In fact,<br />

for ε11 p > 10 -3 <strong>and</strong> any realistic range <strong>of</strong> elastic parameters <strong>and</strong> remote stresses, the final<br />

63<br />

2G<br />

2G<br />

(4c)


term in (6a) <strong>and</strong> (6c) begins to dominate the tip stress state, with σ11 t increasing linearly<br />

with ε11 p at a rate 1/ν ∞ times greater than σ33 t .<br />

Figure 2.12 illustrates this relationship using the putative remote stress state <strong>and</strong> elastic<br />

parameters defined above. Normal stress components at the tip are plotted as functions <strong>of</strong><br />

ε11 p from 0 to 0.1 for two cases: Gb = 2.182Gs <strong>and</strong> νb = 0.5νs (solid lines), <strong>and</strong> Gb =<br />

0.857Gs <strong>and</strong> νb = 2νs (dotted lines). Thus Eb = 2Es in the first case <strong>and</strong> 0.5Es in the<br />

second, while in both cases, Kb/Ks = 1.5. The two sets <strong>of</strong> lines lie virtually on top <strong>of</strong> each<br />

other over the entire range <strong>of</strong> ε11 p , indicating not only that the actual plastic strain <strong>of</strong> 0.1<br />

dominates the state <strong>of</strong> stress at the tip, but that the elastic properties inside the model<br />

b<strong>and</strong> have a negligible effect on the state <strong>of</strong> stress even as ε11 p approaches zero.<br />

The results <strong>of</strong> this analysis, however, pertain only to a point infinitesimally close to<br />

the tip <strong>of</strong> an infinitesimally thin ellipsoidal inclusion. The aspect ratio <strong>of</strong> the idealized<br />

ellipsoidal CB is small (10 -4 ) but not zero, <strong>and</strong> the approximation <strong>of</strong> the granular Aztec<br />

s<strong>and</strong>stone as a homogeneous continuum makes looking very close to the tip problematic.<br />

We therefore turn to a complete analytical solution <strong>of</strong> the Eshelby problem to match the<br />

exact physical parameters <strong>of</strong> the field-based conceptual model <strong>and</strong> extend the analysis<br />

away from the b<strong>and</strong>-s<strong>and</strong>stone interface.<br />

7.2. Eshelby inclusion model<br />

For this analysis, we use an exact, closed-form solution to the general three-<br />

dimensional Eshelby problem based on the equivalent inclusion method. The solution is<br />

coded in MATLAB® <strong>and</strong> was made available to us by Pradeep Sharma at the University<br />

<strong>of</strong> Houston. Within certain limitations, the code accepts any combination <strong>of</strong> ellipsoidal<br />

inclusion dimensions, internal <strong>and</strong> external elastic properties <strong>and</strong> remote loading to<br />

calculate the state <strong>of</strong> stress for any point outside the inclusion. The embedded layer stress<br />

results in Figure 2.12 can be reproduced to within a fraction <strong>of</strong> 1% using the Eshelby<br />

code, by inputting an axisymmetric inclusion with a 10 -7 aspect ratio <strong>and</strong> calculating the<br />

state <strong>of</strong> stress just outside (within machine precision) <strong>of</strong> the tip.<br />

To represent the field-based conceptual model, dimensions <strong>of</strong> 25 m by 25.1 m by 9.38<br />

mm are used, along with a fixed value <strong>of</strong> ε11 p = 0.1. The orientation <strong>of</strong> the model b<strong>and</strong><br />

with respect to the remote stress field, the range <strong>of</strong> elastic parameters inside <strong>and</strong> outside<br />

64


MPa<br />

2000<br />

1800<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

σ 11<br />

200<br />

σ22 0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

p<br />

Uniaxial Plastic Strain ( ε11) Figure 2.12. Normal stress components just outside the tip <strong>of</strong> the model compaction b<strong>and</strong><br />

as linear functions <strong>of</strong> the uniaxial plastic strain from the embedded layer solution. Two<br />

scenarios are plotted: one for a b<strong>and</strong> with Young’s modulus twice that <strong>of</strong> the surrounding<br />

material (solid lines) <strong>and</strong> one with Young’s modulus half that <strong>of</strong> the surroundings (dotted<br />

lines). In both cases, the b<strong>and</strong>/surroundings ratio <strong>of</strong> bulk moduli is 1.5 <strong>and</strong> the lines plot<br />

virtually on top <strong>of</strong> each other. The state <strong>of</strong> stress at the tip is dominated by the plastic<br />

strain <strong>and</strong> insensitive to differences in elastic properties.<br />

65<br />

σ 33


MPa<br />

MPa<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2<br />

(a)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

(b)<br />

Distance from tip (m)<br />

Distance from flank (m)<br />

σ 11<br />

σ 22<br />

σ 33<br />

σ 11<br />

σ 22<br />

σ 33<br />

0<br />

0 5 10 15 20 25<br />

Figure 2.13. Distributions <strong>of</strong> the normal stress components with distance from the model<br />

compaction b<strong>and</strong> tip (a) <strong>and</strong> flank (b) calculated using the Eshelby inclusion solution <strong>and</strong><br />

uniform uniaxial strain <strong>of</strong> 10%. Two scenarios are plotted: one for a b<strong>and</strong> with Young’s<br />

modulus twice that <strong>of</strong> the surrounding material (solid lines) <strong>and</strong> one with Young’s<br />

modulus half that <strong>of</strong> the surroundings (dotted lines). For both scenarios the<br />

b<strong>and</strong>/surroundings ratio <strong>of</strong> bulk moduli is 1.5 <strong>and</strong>, at both tip <strong>and</strong> flank, the lines plot<br />

virtually on top <strong>of</strong> each other. The state <strong>of</strong> stress is insensitive to differences in elastic<br />

properties <strong>and</strong> only significantly perturbed within a few cm <strong>of</strong> the tip.<br />

66


the b<strong>and</strong>, <strong>and</strong> the magnitudes <strong>of</strong> the remote stresses remain the same as in the embedded<br />

layer analysis. Figure 2.13 presents the resulting stress distributions outside the model<br />

b<strong>and</strong> away from the tip along the x2-axis (Figure 2.13a) <strong>and</strong> away from the midpoint <strong>of</strong><br />

the flank along the x1-axis (Figure 2.13b). Once again, there is a negligible difference<br />

between the results obtained over a realistic range <strong>of</strong> shear stiffness contrast between the<br />

b<strong>and</strong> <strong>and</strong> surrounding material for Kb/Ks fixed at 1.5. At about one mm from the tip, the<br />

compressive normal stresses are about double their remote values. This stress<br />

concentration dies away rapidly with distance, dropping to about 1.25 times the remote<br />

values at one cm from the tip, <strong>and</strong> decaying to a residual increase <strong>of</strong> about 5% within 20<br />

cm. Immediately adjacent to the flank <strong>of</strong> the b<strong>and</strong>, the compressive normal stresses drop<br />

by less than 3% relative to the remote values, <strong>and</strong> then smoothly rebound to effective<br />

background levels within about 10 m (σ22 <strong>and</strong> σ33) <strong>and</strong> 20 m (σ11).<br />

This Eshelby formulation provides the most accurate possible representation <strong>of</strong> the<br />

field-based conceptual model, <strong>and</strong> the results reinforce the conclusion from the embedded<br />

layer analysis that the perturbed state <strong>of</strong> stress induced is overwhelmingly a function <strong>of</strong><br />

the plastic strain accommodated. Even allowing the relative bulk modulus within the<br />

model b<strong>and</strong> to vary outside realistic bounds—from volumetrically very compliant (K =<br />

4.2 GPa) to very stiff (K = 66.7 GPa)—exerts a negligible effect on the resulting near-tip<br />

stress field when ε11 p = 0.1. Departures from the prescribed axisymmetric geometry also<br />

yield minimal effect. Holding the x2 dimension <strong>of</strong> the model b<strong>and</strong> constant at 25 m <strong>and</strong><br />

varying the x3 dimension from 6.25 m (¼x2)to 100 m (4x2) produces a variation <strong>of</strong> less<br />

than 5% in the magnitude <strong>of</strong> σ33 realized one cm in front <strong>of</strong> the tip.<br />

7.3. Anticrack model<br />

The preceding analyses indicate the primary importance <strong>of</strong> the uniaxial plastic strain<br />

accommodated within an idealized CB in determining the state <strong>of</strong> stress induced around<br />

it. This suggests that an anticrack treatment <strong>of</strong> the problem, with a specified distribution<br />

<strong>of</strong> closing-mode displacement discontinuity taking the place <strong>of</strong> internal plastic strain,<br />

would also capture the essential mechanical nature <strong>of</strong> the conceptual model. To test the<br />

validity <strong>of</strong> this proposition, we turn to a simple, two-dimensional displacement<br />

discontinuity representation <strong>of</strong> the idealized CB as an anticrack (Sternl<strong>of</strong> <strong>and</strong> Pollard,<br />

67


2002) <strong>and</strong> use a BEM approach (Crouch, 1976; Crouch <strong>and</strong> Starfield, 1983) coded in<br />

MATLAB® to solve it.<br />

In this analysis, we consider the principal x1-x2 (horizontal) plane through the middle<br />

<strong>of</strong> the CB, which roughly corresponds to the outcrop face from which the field data were<br />

collected. The 25-m-long CB trace is represented by 2,500 cm-long constant<br />

displacement discontinuity boundary elements laid end-to-end along the x2-axis from -<br />

12.5 m to +12.5 m. The closing-mode displacement discontinuity <strong>of</strong> each element is<br />

calculated from the elliptical relation<br />

D<br />

2<br />

⎛ Tmax<br />

⎞ ⎛ xi<br />

⎞<br />

i = ⎜ − Tmax<br />

⎟ 1−<br />

⎜ ⎟<br />

(7)<br />

⎝ 0.<br />

9<br />

⎠<br />

⎝ a ⎠<br />

where Di is the closing-mode displacement discontinuity <strong>of</strong> the ith element, Tmax is the<br />

maximum (midpoint) thickness <strong>of</strong> the CB trace (9.38 mm), xi is the x2-coordinate <strong>of</strong> the<br />

midpoint <strong>of</strong> the ith element, <strong>and</strong> a is the half length <strong>of</strong> the trace (12.5 m). This yields a<br />

step-wise distribution <strong>of</strong> closing-mode displacement discontinuity equivalent to the<br />

homogeneous uniaxial plastic strain <strong>of</strong> 10% used above. Because displacement is<br />

specified however, the elastic properties <strong>of</strong> the CB do not enter the formulation. Also, the<br />

assumption <strong>of</strong> plane strain dictates that, at every point in the x1-x2 plane, σ33 = νb(σ11 +<br />

σ22) <strong>and</strong> σ13 = σ23 = 0.<br />

Figure 2.14 shows the resulting anticrack distribution <strong>of</strong> normal stresses away from<br />

the tip along the x2-axis, <strong>and</strong> away from the flank along the x1-axis as compared to those<br />

computed with the Eshelby model. Within one mm <strong>of</strong> the tip, where the singularity <strong>of</strong> the<br />

anticrack solution produces stresses tending toward infinity, correspondence is poor<br />

(Figure 2.14a). At one cm, the two solutions correspond to within 5% for all normal<br />

stress components. Beyond 2cm the mismatch drops to less than 1%, with the anitcrack<br />

values always slightly above the Eshelby values. At the flank, correspondence between<br />

the two solutions is excellent, with less than a 1% mismatch for all normal stress<br />

components at any distance (Figure 2.14b).<br />

Figure 2.15 highlights the correspondence between the near-tip distributions <strong>of</strong> σ11<br />

for the anticrack <strong>and</strong> Eshelby models on a log-log plot. Within a distance (r) from the tip<br />

<strong>of</strong> less than 0.2 mm, the anticrack distribution <strong>of</strong> σ11 is dominated by the constant<br />

68


MPa<br />

MPa<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2<br />

(a)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 5 10 15 20 25<br />

(b)<br />

Distance from tip (m)<br />

Distance from flank (m)<br />

Figure 2.14. Comparison <strong>of</strong> the normal stress component distributions with distance<br />

from the model compaction b<strong>and</strong> tip (a) <strong>and</strong> flank (b) calculated using the BEM anticrack<br />

solution (solid lines) <strong>and</strong> Eshelby inclusion solution (dotted lines). Beyond 2 cm from the<br />

tip, the two solutions agree to within 1%. They increasingly diverge as the anticrack<br />

solution goes toward infinity at the tip, while the Eshelby solution remains finite.<br />

Adjacent to the flank, differences between the two solutions are less than 1% everywhere.<br />

69<br />

σ11 σ22 σ 33<br />

σ 11<br />

σ 22<br />

σ 33


MPa<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 −5 10 −4 10 −3 10 −2 10 −1<br />

10 0<br />

Distance from tip, r (m)<br />

Figure 2.15. Log-log plot <strong>of</strong> the distribution <strong>of</strong> σ11 with distance from the model<br />

compaction b<strong>and</strong> tip: BEM anticrack solution (solid line), Eshelby inclusion solution<br />

(dashed line), crack-like contours <strong>of</strong> 1/sqrt(r) (dotted lines), dislocation-like contours <strong>of</strong><br />

1/r (dash-dot lines). At the distance r = 1 cm from the tip, the two solutions effectively<br />

coincide. As r decreases, the Eshelby solution approaches the 1/sqrt(r) distribution,<br />

although it remains finite. The BEM anticrack solution, which is composed <strong>of</strong> a step-wise<br />

distribution <strong>of</strong> constant closing-mode displacement discontinuity elements, goes to<br />

infinity as 1/r.<br />

Distance (m)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

−0.1<br />

−0.2<br />

−0.3<br />

−0.4<br />

12.2 12.3 12.4 12.5 12.6 12.7 12.8 12.9 13<br />

Distance (m)<br />

Figure 2.16. Contour plot <strong>of</strong> mean normal stress at the tip <strong>of</strong> the model anticrack<br />

compaction b<strong>and</strong> (b<strong>and</strong> represented by white slot). The remote mean normal stress is 30<br />

MPa. Tick marks indicate the local orientation <strong>of</strong> the maximum principal stress.<br />

70<br />

MPa<br />

38<br />

37<br />

36<br />

35<br />

34<br />

33<br />

32<br />

31<br />

30


displacement (dislocation) tip element, varying roughly as 1/r (Weertman <strong>and</strong> Weertman,<br />

1964). For distances greater than 0.2 mm <strong>and</strong> less than about 2 mm, the anticrack<br />

distribution varies roughly as 1/sqrt(r), as expected for a true crack solution (Lawn <strong>and</strong><br />

Wilshaw, 1975). For distances greater than 2 mm, the magnitude <strong>of</strong> σ11 decreases more<br />

gradually toward the remote value <strong>of</strong> 40 MPa. The Eshelby distribution <strong>of</strong> σ11, on the<br />

other h<strong>and</strong>, varies much more gradually with distance from the tip, only beginning to<br />

approach 1/sqrt(r) inside about 0.01 mm. Beyond 1 cm from the tip, the anticrack <strong>and</strong><br />

Eshelby distributions <strong>of</strong> σ11 effectively coincide.<br />

The size, shape <strong>and</strong> magnitude <strong>of</strong> the stress perturbation generated at the tip <strong>of</strong> the<br />

model anticrack CB is entirely consistent with the field-based interpretation <strong>of</strong> self-<br />

sustaining in-plane <strong>propagation</strong> oriented generally orthogonal to the maximum remote<br />

principal stress. For example, elevated mean normal (lithostatic) stress, which would tend<br />

to favor compaction, is tightly concentrated immediately in front <strong>of</strong> the tip, while the<br />

maximum local principal stress acts everywhere nearly perpendicular to the trace <strong>of</strong> the<br />

existing b<strong>and</strong> (Figure 2.16).<br />

8. Discussion<br />

The results presented above demonstrate the efficacy <strong>of</strong> the two-dimensional,<br />

numerical BEM anticrack approach in representing the full three-dimensional Eshelby<br />

state <strong>of</strong> stress generated around an idealized CB. The only significant mismatch occurs<br />

within one cm <strong>of</strong> the tip, which in any case is the length scale at which the model<br />

prescription <strong>of</strong> homogeneous material continuity begins to depart from the granular<br />

reality <strong>of</strong> the Aztec s<strong>and</strong>stone. We argue therefore that the anticrack formulation is both<br />

sufficient <strong>and</strong> even preferable for the purposes <strong>of</strong> modeling CBs from two-dimensional<br />

outcrop data. Other advantages include increased computational efficiency, the ability to<br />

h<strong>and</strong>le asymmetric configurations <strong>of</strong> multiple CBs with non-uniform distributions <strong>of</strong><br />

closing-mode displacement (i.e. plastic strain), <strong>and</strong> the potential for modeling active<br />

<strong>propagation</strong> <strong>and</strong> pattern development at the outcrop scale (Sternl<strong>of</strong> et al., 2003).<br />

This study also highlights several interrelated <strong>and</strong> enigmatic issues that warrant<br />

further investigation: the apparent absence <strong>of</strong> a near-tip endzone expressed either as<br />

variations in porosity or concentrated grain damage; the observation that thickness<br />

pr<strong>of</strong>iles <strong>of</strong> isolated CBs are elliptical for meters to tens <strong>of</strong> meters from the tip; the marked<br />

71


concentration <strong>of</strong> quartz plasticity within the b<strong>and</strong>s <strong>and</strong> the spatial uniformity <strong>of</strong> the<br />

plastic strain thereby accommodated; <strong>and</strong> the apparent distribution <strong>of</strong> perturbed<br />

compressive stress that suggests grain damage could only occur within a few cm <strong>of</strong> the<br />

tip.<br />

Ultimately, we anticipate that more refined microscopic examinations will reveal the<br />

presence <strong>of</strong> a near tip CB process zone defined by quartz grain damage <strong>and</strong> incipient<br />

plasticity, <strong>and</strong> that the diameter <strong>of</strong> this s<strong>of</strong>tened damage zone may correlate to a<br />

maximum attainable b<strong>and</strong> thickness independent <strong>of</strong> trace length. Certainly, most <strong>of</strong> the<br />

uniform inelastic compaction accommodated within a CB occurs well behind the tip-line<br />

in an environment <strong>of</strong> reduced compressive stress that would seem to preclude thickening<br />

<strong>of</strong> the b<strong>and</strong> by lateral <strong>propagation</strong> into undamaged s<strong>and</strong>stone. We suggest that the<br />

characteristically elliptical pr<strong>of</strong>ile <strong>of</strong> an isolated b<strong>and</strong> could be due primarily to relatively<br />

rapid mechanical <strong>propagation</strong> <strong>of</strong> the tip-line, compared to relatively slow processes <strong>of</strong><br />

plastic relaxation <strong>and</strong> collapse that lead to progressive thickening within the rind <strong>of</strong><br />

damaged grains left in its wake.<br />

In terms <strong>of</strong> the bulk material <strong>and</strong> stress-strain conditions conducive to compaction<br />

localization in s<strong>and</strong>stone, as well as the mechanical characteristics <strong>of</strong> the phenomenon,<br />

this analysis <strong>of</strong> natural CBs in the Aztec s<strong>and</strong>stone presents a scenario seemingly at odds<br />

with results <strong>and</strong> interpretations reported in the recent experimental <strong>and</strong> theoretical rock<br />

<strong>mechanics</strong> literature. Firstly, regarding material <strong>and</strong> load conditions, we underst<strong>and</strong> the<br />

phenomenon to have occurred in well consolidated, saturated, but essentially uncemented<br />

s<strong>and</strong>stone at moderate mean compressive stresses consistent with burial <strong>of</strong> less than 2.5<br />

km in a thrust-faulting tectonic regime. Laboratory efforts to induce compaction<br />

localization in triaxial experiments have tended to focus on moderately to well-cemented<br />

s<strong>and</strong>stones (e.g. Berea, Bentheim <strong>and</strong> Castlegate) generally subjected to much greater<br />

confining pressures reaching as high as 300 MPa (e.g. Wong et al., 2001), which is<br />

equivalent to more than 12 km <strong>of</strong> overburden <strong>and</strong> enough to induce brittle-ductile<br />

transition behavior. The grain-scale textures <strong>of</strong> compaction resulting from such<br />

experiments also tend to involve intense grain crushing <strong>and</strong> comminution far in excess <strong>of</strong><br />

that observed in natural CBs.<br />

72


Secondly, the mechanical style <strong>of</strong> compaction induced in the laboratory differs<br />

fundamentally from that observed in the field, where discrete CBs appear to have<br />

initiated throughout the upper Aztec at widely distributed flaws, <strong>and</strong> then propagated<br />

outward along their tip lines to form a pervasive fabric grossly symmetric to the<br />

maximum remote compressive stress. The typical experiment, on the other h<strong>and</strong>, has<br />

produced a compaction front <strong>of</strong> damage that propagates across the specimen from one<br />

end cap toward the other, traveling in the direction <strong>of</strong> maximum compression <strong>and</strong> leaving<br />

fully compacted material in its wake (e.g Baud et al., 2004; Olsson <strong>and</strong> Holcomb, 2000).<br />

Such results have been convincingly interpreted in theoretical terms as a stress/strain<br />

bifurcation from homogenous to localized compaction (Issen <strong>and</strong> Rudnicki, 2000; Issen<br />

<strong>and</strong> Rudnicki, 2001; Olsson, 1999).<br />

We suggest, however, that this snowplow-like progression <strong>of</strong> compaction is an<br />

experimental phenomenon largely unrelated to natural compaction localization as<br />

observed in the Aztec s<strong>and</strong>stone at the Valley <strong>of</strong> Fire. Similarly, the practical<br />

applicability <strong>of</strong> bifurcation analysis is somewhat limited both by the stipulation <strong>of</strong> initial<br />

stress/inelastic strain homogeneity, <strong>and</strong> by the fact that it applies only to the moment <strong>of</strong><br />

localization without reference to any geometric constraint other than imposed planarity.<br />

By contrast, the observations <strong>and</strong> analysis reported here indicate that localization is the<br />

direct result <strong>of</strong> initial material heterogeneity—the concentration <strong>of</strong> stress at <strong>and</strong> failure <strong>of</strong><br />

pre-existing grain-scale Griffith flaws—<strong>and</strong> that the subsequent evolution <strong>of</strong> realistic CB<br />

arrays can be understood as the simultaneous <strong>propagation</strong> <strong>and</strong> interaction <strong>of</strong> many<br />

discrete b<strong>and</strong>s modeled as anticracks.<br />

Despite the temporal <strong>and</strong> physical limitations inherent to experimental modeling, we<br />

therefore suggest that a key to progress in compaction localization research lies in<br />

introducing stress-concentrating initial flaws to jumpstart the deformation process in<br />

more representative materials subjected to less extreme loading conditions. The rate <strong>and</strong><br />

state effects <strong>of</strong> pore water (Baud et al., 2000) <strong>and</strong> heat (Chester et al., 2004) might also<br />

catalyze improved results at laboratory scales <strong>and</strong> strain rates. A h<strong>and</strong>ful <strong>of</strong> recent<br />

experiments have incorporated the initial flaw idea (Sternl<strong>of</strong> <strong>and</strong> Pollard, 2002) as a<br />

circumferential notch designed to concentrate the applied load <strong>and</strong> localize compaction<br />

away from the end caps (Vajdova et al., 2003; Tembe et al., 2006). Although still<br />

73


conducted on well-cemented s<strong>and</strong>stones subjected to unrealistically high confining<br />

pressures, these experiments have produced some promising results. Another set <strong>of</strong><br />

experiments, although originally designed to investigate borehole breakouts in s<strong>and</strong>stone,<br />

have produced perhaps the most representative CBs yet achieved in the laboratory<br />

(Haimson, 2003; Haimson <strong>and</strong> Lee, 2004).<br />

9. Acknowledgements<br />

Our thanks to John Childs (Figure 2.1) <strong>and</strong> Frantz Maerten for their able assistance in<br />

the field, <strong>and</strong> to the Valley <strong>of</strong> Fire State Park staff <strong>and</strong> Superintendent Jim Hammons for<br />

access, support <strong>and</strong> permission to collect samples. We also are deeply indebted to Gaurav<br />

Chopra, Pradeep Sharma <strong>and</strong> Ole Kaven for their computational <strong>and</strong> coding work using<br />

MATLAB®. Chopra developed the BEM code used in this paper, while Sharma<br />

graciously provided the Eshelby code, which was modified for our use by Kaven. Tapan<br />

Mukerji provided invaluable assistance on issues <strong>of</strong> image analysis <strong>and</strong> rock physics,<br />

while Bob Jones enabled the SEM effort. Dave Lockner (U.S. Geological Survey), <strong>and</strong><br />

Bill Olsson <strong>and</strong> David Holcomb (S<strong>and</strong>ia National Laboratories) graciously allowed us to<br />

conduct preliminary testing <strong>of</strong> our ideas in the lab, while providing many useful insights.<br />

This work was funded under grants from the U.S. Department <strong>of</strong> Energy, Office <strong>of</strong> Basic<br />

Energy Science, Geosciences Research Program to John Rudnicki (DE-FG02-<br />

93ER14344), <strong>and</strong> to David Pollard <strong>and</strong> Atilla Aydin (DE-FG03-94ER14462), with<br />

additional support from the <strong>Stanford</strong> Rock Fracture Project.<br />

74


1. Abstract<br />

Chapter 3<br />

Energy-release model <strong>of</strong> compaction b<strong>and</strong> <strong>propagation</strong><br />

The elastic strain energy released per unit advance <strong>of</strong> a compaction b<strong>and</strong> in an infinite<br />

layer <strong>of</strong> thickness h is used to identify <strong>and</strong> assess quantities relevant to <strong>propagation</strong> <strong>of</strong><br />

isolated compaction b<strong>and</strong>s observed in outcrop. If the elastic moduli <strong>of</strong> the b<strong>and</strong> <strong>and</strong> the<br />

surrounding host material are similar <strong>and</strong> the b<strong>and</strong> is much thinner than the layer, the<br />

p<br />

energy released is simply σ ξhε , where σ + is the compressive stress far ahead <strong>of</strong> the<br />

b<strong>and</strong> edge, ξ h is the thickness <strong>of</strong> the b<strong>and</strong> <strong>and</strong><br />

+<br />

p<br />

ε is the uniaxial inelastic compactive<br />

strain in the b<strong>and</strong>. Using representative values inferred from field data yields an energy<br />

2<br />

release rate <strong>of</strong> 40 kJ/m , which is roughly comparable with compaction energies<br />

inferred from axisymmetric compression tests on notched s<strong>and</strong>stone samples. This<br />

suggests that a critical value <strong>of</strong> the energy release rate may govern <strong>propagation</strong>, although<br />

the particular value is likely to depend on the rock type <strong>and</strong> details <strong>of</strong> the compaction<br />

process.<br />

2. Introduction<br />

Mollema <strong>and</strong> Antonellini (1996) reported observations <strong>of</strong> tabular zones <strong>of</strong> localized<br />

compactive deformation without evident shear in the Navajo s<strong>and</strong>stone formation <strong>of</strong><br />

southern Utah. They called these structures compaction b<strong>and</strong>s (CBs). Similar structures<br />

have been observed in other field locations <strong>and</strong> Sternl<strong>of</strong> et al. (2005) have reported<br />

detailed measurements from CBs cropping out as extensive arrays in the Aztec s<strong>and</strong>stone<br />

<strong>of</strong> southeastern Nevada. Localized compaction has also been observed in laboratory<br />

axisymmetric compression tests on s<strong>and</strong>stones (Olsson, 1999; Olsson <strong>and</strong> Holcomb,<br />

2000; Baud et al., 2004; Klein et al., 2001) <strong>and</strong> in simulations <strong>of</strong> bore hole breakouts<br />

(Hamison, 2001; Haimson <strong>and</strong> Lee, 2004; Klaetsch <strong>and</strong> Haimson, 2002). There are a<br />

number <strong>of</strong> differences among the laboratory observations <strong>and</strong> between the laboratory <strong>and</strong><br />

field observations. Nevertheless, all the observed structures are characterized by localized<br />

porosity reduction in a roughly tabular zone that forms perpendicular to the maximum<br />

compressive stress. In addition to their inherent interest as a mode <strong>of</strong> localized<br />

deformation in porous rock that has been recognized only recently, CBs have potential<br />

75


~ 62 m<br />

compaction b<strong>and</strong> trend<br />

500µm<br />

B<strong>and</strong> tip<br />

B<strong>and</strong> tip<br />

Figure 3.1. Anatomy <strong>of</strong> a compaction b<strong>and</strong> in the Aztec s<strong>and</strong>stone. (top) Archetypical fin<br />

exposed in outcrop from tip-to-tip (15 mm max thickness). (bottom) Backscatter electron<br />

images <strong>of</strong> the uncompacted s<strong>and</strong>stone (left) <strong>and</strong> CB (right). Quartz is medium gray,<br />

feldspar is dirty white, clay is dark gray, hematite is bright white <strong>and</strong> porosity is black.<br />

Note intense micr<strong>of</strong>racture-accommodated plasticity <strong>of</strong> grains common in the b<strong>and</strong>.<br />

These images were collected from the same thin section collected across the b<strong>and</strong>.<br />

76


importance to applications because <strong>of</strong> their effect on permeability. Both field (Sternl<strong>of</strong> et<br />

al., 2004; Taylor <strong>and</strong> Pollard, 2000) <strong>and</strong> laboratory measurements (Holcomb <strong>and</strong> Olsson,<br />

2003; Vajdova et al., 2004) have shown that the permeability for flow across CBs is as<br />

much as several orders <strong>of</strong> magnitude smaller than that <strong>of</strong> the surrounding (host) material.<br />

Consequently, the formation <strong>of</strong> CBs in subsurface porous formations can strongly affect<br />

applications involving the injection or withdrawal <strong>of</strong> fluids, such as petroleum production,<br />

CO 2 sequestration, contamination clean-up <strong>and</strong> aquifer management.<br />

Viewed as two-dimensional pr<strong>of</strong>iles in outcrops <strong>of</strong> Aztec s<strong>and</strong>stone, individual CBs<br />

are generally long (10s <strong>of</strong> meters), thin (a few centimeters) zones <strong>of</strong> grains compacted<br />

through inelastic processes <strong>of</strong> porosity loss, primarily grain fracture <strong>and</strong> rearrangement<br />

(Figure 3.1). These CBs appear to have propagated outward along planes roughly<br />

orthogonal to the direction <strong>of</strong> maximum remote compressive stress (Sternl<strong>of</strong> et al., 2005).<br />

Experimental efforts to induce compaction localization in s<strong>and</strong>stone cylinders under<br />

axisymmetric compression (Olsson, 1999; Olsson <strong>and</strong> Holcomb, 2000; Baud et al., 2004;<br />

Klein et al., 2001) have tended to produce zones <strong>of</strong> compaction that appear to occur<br />

nearly instantaneously across the width <strong>of</strong> the specimen indicating that <strong>propagation</strong>, if it<br />

occurs, does so rapidly. On the other h<strong>and</strong>, in triaxial simulations <strong>of</strong> bore hole breakouts<br />

in s<strong>and</strong>stone (Haimson, 2001; Hamison <strong>and</strong> Lee, 2004; Klaetsch <strong>and</strong> Haimson, 2002), the<br />

tips <strong>of</strong> features resembling CBs have been identified extending from elongated, slot-<br />

shaped breakouts. In order to examine b<strong>and</strong> <strong>propagation</strong> in the laboratory, Vajdova <strong>and</strong><br />

Wong (2003) <strong>and</strong> Tembe et al., (2006) <strong>and</strong> conducted axisymmetric compression<br />

experiments on cylindrical specimens with an external, circumferential notch. In some<br />

cases, they observed compaction initiating at the notch <strong>and</strong> extending across the<br />

specimen in increments corresponding to small load drops.<br />

A critical element in reconciling different laboratory observations, differences<br />

between laboratory <strong>and</strong> field observations, <strong>and</strong> in making predictions for practical<br />

applications is a better underst<strong>and</strong>ing <strong>of</strong> conditions for <strong>propagation</strong>. Although theoretical<br />

studies (Olsson, 1999; Issen <strong>and</strong> Rudnicki, 2000, 2001; Detournay et al., 2003; Bésuelle<br />

<strong>and</strong> Rudnicki, 2004; Rudnicki, 2002, 2003, 2004) have had some success in identifying<br />

stress conditions <strong>and</strong> material behavior that are favorable for CB formation, there is less<br />

underst<strong>and</strong>ing <strong>of</strong> the conditions for <strong>propagation</strong> or arrest. Sternl<strong>of</strong> <strong>and</strong> Pollard (2002) <strong>and</strong><br />

77


Sternl<strong>of</strong> et al. (2005) have suggested a linear elastic anticrack model for CB <strong>propagation</strong>.<br />

An anticrack is the compressive, closing-mode counterpart <strong>of</strong> the more familiar tensile,<br />

opening mode crack but with the sign <strong>of</strong> the stresses <strong>and</strong> crack-surface displacements<br />

reversed. Although the model formally predicts interpenetration <strong>of</strong> the crack surfaces, for<br />

CBs this interpenetration is interpreted physically as the inelastic compaction<br />

accompanying porosity loss. In this model, a CB propagates because <strong>of</strong> the compressive<br />

stress concentration at the edge <strong>of</strong> the b<strong>and</strong>. Currently, little is known about the<br />

magnitude <strong>of</strong> the stress concentration required to induce <strong>propagation</strong> in either the field or<br />

laboratory. Vajdova <strong>and</strong> Wong (2003) <strong>and</strong> Tembe et al. (2006) also used the anticrack<br />

model to interpret their experiments on notched specimens <strong>of</strong> Berea <strong>and</strong> Bentheim<br />

s<strong>and</strong>stone. From the nominal stress vs. displacement curves, Vajdova <strong>and</strong> Wong (2003)<br />

2<br />

estimated a lower bound on the “compaction energy” for Bentheim <strong>of</strong> 16 kJ/m . Tembe<br />

2<br />

et al. (2006) estimated 43 kJ/m for Berea, but because the b<strong>and</strong> propagated at an angle<br />

<strong>of</strong> 55° from the plane ahead <strong>of</strong> the notch, it is likely to be a shear b<strong>and</strong>.<br />

The stress at the edge <strong>of</strong> an anticrack is singular, as it is for a tensile crack.<br />

Consequently, the quantity relevant to the <strong>propagation</strong> <strong>of</strong> the b<strong>and</strong> is the coefficient <strong>of</strong> the<br />

stress singularity, the stress intensity factor K , or, equivalently, the energy release rate E,<br />

related to K by<br />

E =<br />

(1 −ν<br />

)<br />

K<br />

2G<br />

2<br />

, (1)<br />

where G is the shear modulus <strong>and</strong> ν is Poisson’s ratio. At present the only estimates <strong>of</strong><br />

critical values <strong>of</strong> K or E appear to be those by Vajdova <strong>and</strong> Wong (2003) <strong>and</strong> Tembe et al.<br />

(2006) <strong>of</strong> compaction energies. In addition, it is unclear how these critical values relate to<br />

the parameters <strong>of</strong> the b<strong>and</strong>, in particular, the inelastic compactive strain accommodated.<br />

This article presents a very simple energy release model <strong>of</strong> CB <strong>propagation</strong> as a<br />

framework for identifying the relevant controlling parameters.<br />

3. Formulation<br />

The model is based on an example calculation <strong>of</strong> Rice (1968) illustrating the use <strong>of</strong><br />

the J-integral <strong>and</strong> is shown in Figure 3.2. Physically, it represents an idealized geometry<br />

based on outcrop observations from the Aztec s<strong>and</strong>stone (Figure 3.3 <strong>and</strong> Sternl<strong>of</strong> et al.,<br />

78


h<br />

strain energy = W -<br />

h<br />

strain energy = W +<br />

Figure 3.2. Model <strong>of</strong> a semi-infinite CB <strong>of</strong> maximum thickness ξh embedded in an<br />

infinite layer <strong>of</strong> thickness h. Boundaries <strong>of</strong> the layer are displaced toward each other by a<br />

total amount ∆. Propagation <strong>of</strong> the b<strong>and</strong> a unit distance reduces the energy <strong>of</strong> a vertical<br />

slice far ahead <strong>of</strong> the tip (W + ) to that <strong>of</strong> a slice far behind the b<strong>and</strong> tip (W - ).<br />

CB Tip<br />

Figure 3.3. Typical configuration <strong>of</strong> three parallel CBs in the Aztec s<strong>and</strong>stone, forming<br />

the basis for the model in Figure 3.2: left photo corresponds to vertical slice with elastic<br />

strain energy W - , right photo corresponds to vertical slice with energy W + , middle photo<br />

corresponds to transitional zone associated with the elliptical taper <strong>of</strong> the central b<strong>and</strong>’s<br />

tip. In the photos, the hypothetical boundaries <strong>of</strong> the model layer surrounding the central<br />

b<strong>and</strong> (Figure 3.2) fall between it <strong>and</strong> its two bounding neighbors, which are spaced 1. 2m<br />

apart. Thus, h in this case equals 06 . m.<br />

79


2005). It consists <strong>of</strong> a long (infinite in the x-direction) layer <strong>of</strong> material <strong>of</strong> thickness h (in<br />

the y-direction). The layer contains a long (semi-infinite) CB which attains a fixed<br />

thickness ξ h far behind the b<strong>and</strong> edge. Both the layer <strong>and</strong> the b<strong>and</strong> are uniform in z-<br />

direction (orthogonal to the xy-plane <strong>of</strong> Figure 3.2), at least over distances much larger<br />

than h. The layer boundaries are then displaced toward each other a total distance ∆. The<br />

geometry constrains all displacements <strong>and</strong> strains to be uniaxial in the vertical (y)<br />

direction. Because the configuration is self-similar (translationally invariant in the x-<br />

direction), <strong>propagation</strong> <strong>of</strong> the CB a unit distance in the x-direction must reduce the strain<br />

energy in a vertical slice (perpendicular to the xy-plane) far ahead <strong>of</strong> the b<strong>and</strong> edge to that<br />

<strong>of</strong> a vertical slice far behind the b<strong>and</strong> edge. Because displacement is fixed on the<br />

boundaries <strong>of</strong> the sample, this difference in strain energy exactly equals the energy<br />

released per unit area swept out by <strong>propagation</strong> <strong>of</strong> the CB a unit distance.<br />

Far ahead <strong>of</strong> the b<strong>and</strong>, the vertical strain is uniform <strong>and</strong> equal to ∆/ h . If the material<br />

is assumed to be elastic, then the vertical stress is M(∆/h) (horizontal stresses exist to<br />

enforce zero strain in these directions), where M is the modulus governing one-<br />

dimensional strain. For an isotropic material with shear modulus G <strong>and</strong> Poisson’s ratio<br />

ν , M = 2 G(1 − ν ) / (1− 2 ν ) . The strain energy per unit area <strong>of</strong> a vertical slice far ahead <strong>of</strong><br />

the CB is<br />

2<br />

+ 1 ⎛∆⎞ W = Mh⎜ ⎟ . (2)<br />

2 ⎝ h ⎠<br />

Far behind the b<strong>and</strong> edge, where the CB thickness is ξ h , the strain is uniform both<br />

inside ( ε b<strong>and</strong> ) <strong>and</strong> outside ( ε out ) the b<strong>and</strong>. The values are different but must be compatible<br />

with the total displacement <strong>of</strong> the boundary, given by<br />

ξhε + (1 − ξ) hε<br />

=∆. (3)<br />

b<strong>and</strong> out<br />

The stress outside the CB is given by σ out Mε<br />

out<br />

= . The material inside the CB is also<br />

assumed to be elastic, but with a different modulus M b , <strong>and</strong> to have undergone an<br />

p<br />

inelastic, vertical compactive strain ε . Hence, the stress is given by<br />

σ = M ε −ε<br />

⎛<br />

b<strong>and</strong> b ⎜<br />

⎝ b<strong>and</strong><br />

p ⎞<br />

⎟<br />

⎠<br />

. (4)<br />

80


Equilibrium requires that σ b<strong>and</strong> = σ out . This equation can be combined with (3) to<br />

determine the strains in terms <strong>of</strong> ∆/ h , the moduli, the geometry <strong>and</strong><br />

strain energy per unit area <strong>of</strong> a vertical slice far behind the edge <strong>of</strong> the CB is<br />

p<br />

ε . Therefore, the<br />

1<br />

p<br />

W h b<strong>and</strong> b<strong>and</strong> (1 ) h out out<br />

2 2<br />

1<br />

− ⎛ ⎞<br />

= ξ σ ⎜ε − ε ⎟+<br />

− ξ σ ε . (5)<br />

⎝ ⎠<br />

+ −<br />

Taking the difference W − W yields the energy release per unit area created <strong>of</strong> CB<br />

with thickness ξ h :<br />

⎧ 2<br />

⎫<br />

⎪ ⎛ ⎞<br />

2 ⎪<br />

⎨ ⎜ ⎟ ⎬<br />

⎪ ⎝ ⎠ ⎪<br />

⎩⎪ ⎭⎪<br />

1 Mh ⎛∆⎞ ⎛ M ⎞ p⎛∆⎞<br />

p<br />

Eb<strong>and</strong> =<br />

⎜ ⎟ξ⎜ − 1⎟+ 2ξε<br />

⎜ ⎟−<br />

ξε<br />

2( MM / b) ξ + (1 −ξ) ⎝ h⎠ ⎝Mb ⎠ ⎝ h⎠<br />

4. Special cases<br />

If the b<strong>and</strong> modulus vanishes (i.e. M b = 0 ), then (6) reduces to<br />

. (6)<br />

Eb<strong>and</strong> = 1<br />

2 M ⎡ ⎛∆⎞⎤ ⎢ ⎜ ⎟ ∆<br />

⎣ ⎝ h ⎠⎦<br />

⎥ , (7)<br />

where σ + = M ( ∆/ h)<br />

is the uniform stress ahead <strong>of</strong> the b<strong>and</strong>. This is equivalent to the<br />

result for a crack with zero tractions on the surfaces <strong>and</strong>, notably, does not depend on the<br />

inelastic compactive strain. Equating (7) <strong>and</strong> (1) yields the following expression for the<br />

stress intensity factor<br />

K<br />

b<strong>and</strong><br />

=∆<br />

MG<br />

, (8)<br />

(1 −ν<br />

) h<br />

which is independent <strong>of</strong> the b<strong>and</strong> length.<br />

If the elastic modulus <strong>of</strong> the CB remains the same as the material outside (i.e.<br />

M b = M ) <strong>and</strong> we neglect the term (ξε p ) 2 as inconsequentially small, then (6) reduces to<br />

p<br />

Eb<strong>and</strong> = σ + ξε h . (9)<br />

Thus, (9) has the simple interpretation <strong>of</strong> the stress multiplied by the compactive<br />

displacement accommodated within the CB.<br />

An indication <strong>of</strong> the effect <strong>of</strong> the difference in moduli can be obtained from Eb<strong>and</strong> by<br />

simplifying the expression (6) for ξ


⎡ ∆ ⎤ ⎧ p 1 ∆ ⎛ M ⎞⎫<br />

Eb<strong>and</strong> = M ξh ε<br />

1<br />

h 2 h Mb<br />

⎪ ⎛ ⎞ ⎛ ⎞ ⎪<br />

⎢ ⎜ ⎟⎥ ⎨ + ⎜ ⎟⎜<br />

− ⎟⎬.<br />

(10)<br />

⎣ ⎝ ⎠⎦ ⎪⎩ ⎝ ⎠⎝<br />

⎠⎪⎭<br />

Equation (10) suggests that the ratio <strong>of</strong> moduli is probably not significant unless it<br />

p<br />

exceeds two <strong>and</strong> ∆/ h is <strong>of</strong> the same order <strong>of</strong> magnitude as ε . This result is consistent<br />

with the finding <strong>of</strong> Sternl<strong>of</strong> <strong>of</strong> al. (2005) that the internal stiffness does not appreciably<br />

p<br />

affect the state <strong>of</strong> stress around a highly eccentric ellipsoidal CB when ε is on the order<br />

<strong>of</strong> 10 %. If the CB is modeled as a rigid inclusion, then Mb→∞<br />

<strong>and</strong> the last parenthesis in<br />

(10) becomes −1.<br />

If the inelastic compactive strain is much greater than the nominal<br />

p<br />

strain (i.e. ε >> ∆/ h ), then this reduces to the same expression as (9). Consequently,<br />

unless the stiffness <strong>of</strong> the CB material is much less than that <strong>of</strong> the surrounding material,<br />

(9) gives a good approximation to the energy release rate.<br />

5. Discussion<br />

The simple model presented here suggests that CB <strong>propagation</strong> can reasonably be<br />

considered to occur when the energy released per unit advance <strong>of</strong> the b<strong>and</strong> Eb<strong>and</strong> is equal<br />

to some critical value Ecrit that reflects the resistance <strong>of</strong> the material to compaction.<br />

Expressions (6), (7), (9), <strong>and</strong> (10) for Eb<strong>and</strong> give the energy released in terms <strong>of</strong> the<br />

compactive strain <strong>of</strong> the b<strong>and</strong>, the imposed strain, the elastic properties <strong>of</strong> the b<strong>and</strong> <strong>and</strong><br />

the surrounding material, <strong>and</strong> the thickness <strong>of</strong> the b<strong>and</strong> <strong>and</strong> the surrounding material<br />

(spacing between b<strong>and</strong>s). Assuming that b<strong>and</strong> <strong>propagation</strong> does occur when Eb<strong>and</strong> = Ecrit ,<br />

it is possible to estimate the minimum value <strong>of</strong> the material parameter Ecrit from<br />

representative values <strong>of</strong> the parameters derived from the Aztec s<strong>and</strong>stone (Sternl<strong>of</strong> et al.,<br />

2005). Taking σ + = 40 MPa, ξ h = 001 . m, corresponding to 1 cm thick CBs spaced 1<br />

p<br />

meter apart, <strong>and</strong> ε = 01 . , corresponding to a porosity reduction <strong>of</strong> 10 % , yields Eb<strong>and</strong><br />

=<br />

2<br />

40 kJ/m . Ecrit<br />

must be at least this large when <strong>propagation</strong> <strong>of</strong> the b<strong>and</strong> ceased, otherwise<br />

it would have continued advancing. Using the smallest <strong>and</strong> largest values for σ +<br />

estimated by Sternl<strong>of</strong> et al. (2005) <strong>and</strong> keeping the same values <strong>of</strong> ξ h <strong>and</strong><br />

2<br />

range <strong>of</strong> Ecrit from 10 to 60 kJ/m . For<br />

ν = 02 . , the value <strong>of</strong> σ + implies<br />

h<br />

p<br />

ε , yields a<br />

M = 22 GPa, corresponding to G = 8.3 GPa <strong>and</strong><br />

18 10<br />

−3<br />

∆/ = . × . However, ∆ is difficult to estimate <strong>and</strong><br />

82


h , the CB spacing, is the most variable physical parameter in outcrop, ranging from<br />

millimeters to meters.<br />

Despite many differences between the rock types, circumstances <strong>and</strong> morphologies <strong>of</strong><br />

2<br />

b<strong>and</strong> formation in the field <strong>and</strong> the laboratory, the value 40 kJ/m is similar to<br />

compaction energies inferred in the experiments on notched samples. Interestingly, the<br />

value inferred by Tembe et al. (2006) for Berea s<strong>and</strong>stone is about the same, although, as<br />

noted earlier, the localized deformation in this case likely involved significant shear. The<br />

2<br />

value <strong>of</strong> 16 kJ/m inferred by Vajdova <strong>and</strong> Wong (2003) for Bentheim s<strong>and</strong>stone is a<br />

factor <strong>of</strong> 25 . smaller. They did, however, report it as a lower bound. If Eb<strong>and</strong><br />

= 16 kJ/m<br />

<strong>and</strong> the stress is taken as 40 0 MPa, roughly the axial stresses at <strong>propagation</strong> for the<br />

Bentheim samples, then (9) yields 00 . 4 mm for the compactive displacement. For a<br />

plastic strain <strong>of</strong> 7% , the porosity change estimated by Klein et al. (2001), the b<strong>and</strong> width<br />

is 057 . mm, about the same as observed by Baud et al. (2004) (see their Figure 9). Thus,<br />

the expression (9) gives results that are consistent with the limited data. To determine<br />

whether there is systematic variation <strong>of</strong> Ecrit<br />

with different porous s<strong>and</strong>stones <strong>and</strong>,<br />

perhaps, differences in the microscale processes <strong>of</strong> compaction (e.g. fracturing vs. grain<br />

crushing) would require additional observations. Our results, however, indicate that the<br />

energy release rate provides a reasonable framework for comparison.<br />

p<br />

Since the term ξε h in this expression represents the net effective compactive<br />

displacement across the CB far behind the edge, an alternative criterion might be that<br />

<strong>propagation</strong> occurs at a critical value <strong>of</strong> this quantity. In its simplicity, the model does not<br />

specify whether this inelastic strain (or compactive displacement) is accumulated in a<br />

small zone near the edge or more gradually over a large zone behind the edge. The model<br />

only assumes that it eventually reaches a constant, asymptotic value far behind the edge.<br />

Far ahead <strong>of</strong> the edge the stress is σ + = M ( ∆/ h)<br />

. Far behind the edge the stress is<br />

p<br />

∆/ h −ξε<br />

σ − = M b<br />

. (11)<br />

ξ + (1 − ξ)<br />

M / M<br />

b<br />

Presumably, the stress increases from σ + far ahead <strong>of</strong> the CB edge to a higher value<br />

(theoretically infinite for the anticrack model) near the edge then decreases to σ − , at<br />

83<br />

2


which point the inelastic strain has attained a critical value (<br />

p<br />

ε )crit . This process is<br />

illustrated schematically in Figure 3.4, where the reduction <strong>of</strong> stress occurs over an<br />

endzone distance R . Further analysis <strong>of</strong> the spatial <strong>and</strong> temporal distribution <strong>of</strong> energy<br />

dissipation suggested by the more than 5 meters <strong>of</strong> elliptical taper near the ends <strong>of</strong> the<br />

b<strong>and</strong>s reported by Sternl<strong>of</strong> et al. (2005) is warranted <strong>and</strong> should contribute to a better<br />

underst<strong>and</strong>ing <strong>of</strong> the conditions needed for <strong>propagation</strong>.<br />

σ −<br />

σ<br />

σp<br />

σ −<br />

σ + = M∆/h<br />

R x<br />

σp<br />

σ −<br />

σ<br />

(ε ) crit<br />

p<br />

Figure 3.4. Schematic illustration <strong>of</strong> the hypothetical spatial distribution <strong>of</strong> stress near<br />

the edge (tip) <strong>of</strong> a CB. (a) As the tip is approached from the right (σ+ side), the stress<br />

increases to a maximum value σp , then drops below the remote background value <strong>of</strong> σ+<br />

to approach the relaxed value <strong>of</strong> σ— as the total inelastic compactive strain (ε p ) reaches<br />

its critical value (ε p )crit at a distance R behind the tip.<br />

6. Acknowledgements<br />

The authors thank Jim Rice for a suggestion concerning the model <strong>and</strong> Teng-fong<br />

Wong for discussion <strong>of</strong> his laboratory results. Principal financial support for this work<br />

was provided by the U. S. Department <strong>of</strong> Energy, Office <strong>of</strong> Basic Energy Science,<br />

Geosciences Research Program through grants to Northwestern University (John<br />

Rudnicki) <strong>and</strong> <strong>Stanford</strong> University (David Pollard <strong>and</strong> Atilla Aydin). Partial support<br />

during preparation <strong>of</strong> the initial manuscript while Rudnicki was a visitor at the Kavli<br />

Institute for Theoretical Physics was provided by the National Science Foundation under<br />

Grant No. PHY99-07949 (Preprint No. NSF-KITP-05-26). Additional support was<br />

provided by the <strong>Stanford</strong> Rock Fracture Project.<br />

84<br />

ε p


1. Abstract<br />

Chapter 4<br />

Propagation <strong>of</strong> compaction b<strong>and</strong>s in s<strong>and</strong>stone as anticracks:<br />

Field evidence, mechanical theory <strong>and</strong> numerical simulation<br />

Outcrop <strong>and</strong> petrographic observations <strong>of</strong> compaction b<strong>and</strong>s exposed in the Aztec<br />

s<strong>and</strong>stone <strong>of</strong> southeastern Nevada suggest that the <strong>mechanics</strong> <strong>of</strong> their <strong>propagation</strong> <strong>and</strong><br />

interaction can be approximated using anticrack theory. Our numerical simulations <strong>of</strong><br />

anticrack b<strong>and</strong> <strong>propagation</strong> using the boundary element method confirm this, while<br />

suggesting refinements to better account for the compacted-inclusion character <strong>of</strong> the<br />

b<strong>and</strong>s. Using realistic, field-based ranges for all physical parameters, we find that, as with<br />

opening-mode cracks, the degree to which adjacent anticrack b<strong>and</strong>s interact is inversely<br />

proportional to the magnitude <strong>of</strong> the remote differential stress acting upon them. This<br />

model observation suggests that the tendency toward anastomosis along the trend <strong>of</strong> a<br />

subparallel compaction b<strong>and</strong> array—<strong>and</strong> thus the extent to which it would impede fluid<br />

flow in that direction—can be predicted from knowledge <strong>of</strong> the remote stress state in<br />

which it formed. Conversely, the orientations <strong>and</strong> relative magnitudes <strong>of</strong> all three<br />

principal paleo stresses can be estimated from directional variations in the degree <strong>of</strong><br />

anastomosis revealed by a well-exposed (or imaged) compaction b<strong>and</strong> array.<br />

2. Introduction<br />

Compaction b<strong>and</strong>s (CBs) are a phenomenon <strong>of</strong> localized compressive failure<br />

commonly observed in outcrops <strong>of</strong> porous s<strong>and</strong>stone. They represent one kinematic end-<br />

member <strong>of</strong> a family <strong>of</strong> structures known collectively as deformation b<strong>and</strong>s (DBs), which<br />

also includes shear <strong>and</strong> dilation b<strong>and</strong>s, as well as mixed-mode combinations (Antonellini<br />

et al., 1994; Aydin, 1978; DuBernard et al., 2002; Mollema <strong>and</strong> Antonellini, 1996;<br />

Rudnicki <strong>and</strong> Sternl<strong>of</strong>, 2005; Sternl<strong>of</strong> et al., 2005). As thin, tabular features <strong>of</strong> porosity-<br />

loss compaction <strong>and</strong> order-<strong>of</strong>-magnitude permeability reduction that are millimeters to<br />

centimeters thick <strong>and</strong> tens <strong>of</strong> meters or more in planar extent, compactive DBs act as<br />

baffles to subsurface fluid flow under saturated conditions (Pittman, 1981; Freeman,<br />

1990; Antonellini <strong>and</strong> Aydin, 1994; Crawford, 1998; Gibson, 1998; Taylor <strong>and</strong> Pollard,<br />

2000; Sigda <strong>and</strong> Wilson, 2003; Sternl<strong>of</strong> et al., 2004). Where present as pervasive arrays,<br />

85


they can exert significant effects on fluid flow at scales relevant to the management <strong>of</strong><br />

both groundwater <strong>and</strong> hydrocarbon resources (Matthai et al., 1998; Sternl<strong>of</strong> et al., 2006).<br />

However, detecting the presence <strong>of</strong> CBs in the subsurface, let alone gathering the<br />

requisite geometrical data with which to assess their aggregate hydraulic impacts,<br />

presents a challenge. As small-scale structures, CBs are all but invisible to current<br />

seismic <strong>and</strong> borehole geophysical imaging techniques, while the coherent nature <strong>of</strong> the<br />

deformation they accommodate makes them difficult to see in both core <strong>and</strong> borehole<br />

televiewer logs. Even given limited information on CB occurrence, orientation <strong>and</strong><br />

spacing derived from sparse borehole data, knowledge <strong>of</strong> the greater, interconnected<br />

geometry <strong>of</strong> any subsurface array is essential to accurate modeling <strong>of</strong> its effects on fluid<br />

flow at production/injection scales (Sternl<strong>of</strong> et al., 2004; 2006). Geostatistical methods<br />

are commonly used for such spatial extrapolations <strong>and</strong> interpolations <strong>of</strong> data, although<br />

the results are only as good as the training models used to calibrate them. For robust<br />

predictions <strong>of</strong> subsurface CB geometry—or the geometry <strong>of</strong> any structure—we advocate<br />

projections based on a fundamental mechanical underst<strong>and</strong>ing <strong>of</strong> how individual b<strong>and</strong>s<br />

propagate <strong>and</strong> interact to form a through-going array.<br />

This paper exp<strong>and</strong>s on the anticrack-inclusion model for compaction b<strong>and</strong>s as<br />

proposed by Sternl<strong>of</strong> et al. (2005) <strong>and</strong> based on detailed observations <strong>of</strong> abundant CBs<br />

cropping out in the Aztec s<strong>and</strong>stone <strong>of</strong> southeastern Nevada (Figure 4.1). Specifically, we<br />

examine the linear elastic anticrack approximation for CB <strong>mechanics</strong> using the<br />

displacement discontinuity boundary element method (BEM) to simulate 2-D<br />

<strong>propagation</strong> <strong>and</strong> interaction. We then compare the simulation results to diagnostic<br />

patterns observed in Aztec outcrop. This effort is iterative, ins<strong>of</strong>ar as observations <strong>of</strong> CBs<br />

in the Aztec inspired the anticrack-inclusion model <strong>and</strong> now comprise the st<strong>and</strong>ard<br />

against which it can be tested <strong>and</strong> improved, while insights gleaned from the modeling<br />

effort suggest new perspectives for refined observations.<br />

Ultimately, we conclude that the anticrack model provides a valid first approximation<br />

for conceptualizing CBs, predicting their dominant orientation symmetric to the<br />

maximum remote compressive stress, <strong>and</strong> simulating their patterns <strong>of</strong> <strong>propagation</strong> <strong>and</strong><br />

interaction. We also find that, as with opening-mode cracks (e.g. Cruikshank et al., 1991;<br />

Olson <strong>and</strong> Pollard, 1989; Thomas <strong>and</strong> Pollard, 1993), the degree to which adjacent CBs<br />

86


NV UT<br />

CA<br />

AZ<br />

Park Road<br />

Map<br />

Detail<br />

Park Boundary<br />

Park Office<br />

Route 169<br />

0 km 5<br />

NEVADA<br />

ARIZONA<br />

10 km<br />

Valley<br />

<strong>of</strong> Fire<br />

State Park<br />

N<br />

LAKE<br />

MEAD<br />

115 o 30'<br />

Figure 4.1. Location <strong>of</strong> the Valley <strong>of</strong> Fire State Park approximately 60 km NE <strong>of</strong> Las<br />

Vegas, Nevada, where the 1,400-m-thick æolian Aztec S<strong>and</strong>tone is extensively exposed.<br />

Figure 4.2. Typical outcrop exposure <strong>of</strong> a sub-parallel compaction b<strong>and</strong> array in the<br />

Aztec s<strong>and</strong>stone. Individual b<strong>and</strong>s weather out in positive relief to form distinctive fins<br />

that are generally at high angle to depositional bedding. The b<strong>and</strong>s in this photo are <strong>of</strong> the<br />

dominant, through-going set, which trends NNW <strong>and</strong> dips steeply E (photo looking N).<br />

87<br />

36 o 15'


interact is inversely proportional to the magnitude <strong>of</strong> the remote differential stress acting<br />

upon them. This model observation suggests that the tendency toward anastomosis along<br />

the trend <strong>of</strong> a subparallel CB array—<strong>and</strong> thus the extent to which fluid flow in that<br />

direction may be impeded—can be predicted from knowledge <strong>of</strong> the remote stress state in<br />

which it formed. Conversely, the orientations <strong>and</strong> relative magnitudes <strong>of</strong> all three<br />

principal paleo stresses can be estimated from directional variations in the degree <strong>of</strong><br />

anastomosis revealed by a well-exposed (or imaged) CB array.<br />

3. Field evidence <strong>and</strong> interpretation<br />

Extensive arrays <strong>of</strong> generally sub-parallel, NNW-trending, steeply E-dipping CBs<br />

pervade the upper 600 meters <strong>of</strong> the 1,400-m-thick Aztec s<strong>and</strong>stone as exposed in <strong>and</strong><br />

around the Valley <strong>of</strong> Fire State Park <strong>of</strong> southeastern Nevada (Sternl<strong>of</strong> et al., 2005)<br />

(Figures 4.1 <strong>and</strong> 4.2). The Aztec is an æolian, subarkosic, chronostratigraphic equivalent<br />

<strong>of</strong> the early Jurassic Navajo <strong>and</strong> Nugget s<strong>and</strong>stones (Marzolf, 1983). It remains weakly<br />

lithified <strong>and</strong> is typified by large-scale tabular <strong>and</strong> trough cross-bedding, average porosity<br />

<strong>of</strong> 20-25%, <strong>and</strong> a mean grain diameter <strong>of</strong> 0.25 mm within a range <strong>of</strong> 0.1 mm to 0.5 mm<br />

(Flodin et al., 2005). The CBs, which comprise the oldest <strong>structural</strong> fabric present, are<br />

relatively thick at a centimeter or more, typically extend for tens <strong>of</strong> meters in outcrop<br />

trace length, <strong>and</strong> represent highly localized tabular inclusions <strong>of</strong> compacted detrital grains<br />

<strong>and</strong> secondary clay accumulation. With no obvious association to pre-existing structures,<br />

the b<strong>and</strong>s appear to represent a pervasive compressional tectonic fabric—in essence the<br />

kinematic <strong>and</strong> <strong>structural</strong> opposite <strong>of</strong> a regional joint set (Sternl<strong>of</strong> et al., 2005). Noting<br />

their relative age <strong>and</strong> dominant orientation, Hill (1989) first suggested that the CBs<br />

formed as a result <strong>of</strong> <strong>and</strong> perpendicular (symmetric) to ENE-directed regional<br />

compression associated with tectonic shortening during the Cretaceous Sevier orogeny.<br />

The CB fabric was subsequently crosscut <strong>and</strong> <strong>of</strong>fset by relatively low-angle shear b<strong>and</strong>s<br />

related to final, proximal emplacement <strong>of</strong> the overlying Sevier thrust sheets, while<br />

overprinting by joints <strong>and</strong> joint-based strike-slip faults associated with Basin <strong>and</strong> Range<br />

extension followed in mid Tertiary time (Eichhubl et al., 2004; Flodin <strong>and</strong> Aydin, 2004;<br />

Hill, 1989; Myers <strong>and</strong> Aydin, 2004; Sternl<strong>of</strong> et al., 2005; Taylor et al., 1999).<br />

Sternl<strong>of</strong> et al. (2005) present a coherent conceptual <strong>and</strong> mechanical model for CBs as<br />

anticrack inclusions that is grounded in detailed outcrop <strong>and</strong> petrographic observations<br />

88


made <strong>of</strong> the Aztec s<strong>and</strong>stone in the Valley <strong>of</strong> Fire. A synopsis <strong>and</strong> expansion <strong>of</strong> those<br />

observations <strong>and</strong> interpretations is presented below. Throughout the rest <strong>of</strong> this paper,<br />

only CBS in the Aztec are considered. All other younger <strong>structural</strong> fabrics—shear b<strong>and</strong>s,<br />

joints <strong>and</strong> joint-based faults—are ignored.<br />

3.1. Outcrop observations<br />

Compaction b<strong>and</strong>s in the otherwise weakly lithified Aztec s<strong>and</strong>stone tend to weather<br />

out in positive relief as sharply delineated tabular fins, rendering them readily visible in<br />

outcrop (Figure 4.2). In areas <strong>of</strong> outcrop relief, individual CBs are seen to be grossly<br />

penny-shaped (i.e. very thin, oblate discs). Expansive exposures over more than 10 km 2<br />

in the Valley <strong>of</strong> Fire represent an approximately horizontal observation plane through the<br />

3-D CB array, revealing a detailed 2-D pattern. Viewed in total, the impression is one <strong>of</strong><br />

a very consistent, NNW-trending, sub-parallel (anastomosing) pattern <strong>of</strong> b<strong>and</strong>s exhibiting<br />

meter to centimeter spacing <strong>and</strong> interspersed with areas up to tens <strong>of</strong> thous<strong>and</strong>s <strong>of</strong> m 2 that<br />

are relatively devoid <strong>of</strong> b<strong>and</strong>s. While the map-view observation plane does not actually<br />

cut exactly perpendicular to the dominant b<strong>and</strong> orientation, which currently dips eastward<br />

at 70-75°, for the purposes <strong>of</strong> basic pattern analysis, interpretation <strong>and</strong> the 2-D<br />

mechanical modeling to come, we make the simplifying assumptions that the current map<br />

view represents paleo horizontal at the time <strong>of</strong> CB formation <strong>and</strong> that the b<strong>and</strong>s<br />

themselves were vertical.<br />

In outcrop, b<strong>and</strong> patterns range from nearly straight <strong>and</strong> parallel, to strongly<br />

anastomosing, to essentially dendritic, to markedly uniform checkerboards comprised <strong>of</strong><br />

two distinct b<strong>and</strong> sets crosscutting each other at high angle (Figure 4.3). Truly parallel,<br />

evenly spaced b<strong>and</strong> patterns are not observed at any substantial scale, while the dendritic<br />

<strong>and</strong> checkerboard patterns occur as relatively rare, localized pockets within the dominant,<br />

sub-parallel pattern. In fact, the dominant, through-going b<strong>and</strong> trend—sometimes steeply<br />

W-dipping, but usually steeply E-dipping—is present in every b<strong>and</strong> pattern observed.<br />

The complex æolian sedimentary architecture <strong>of</strong> the Aztec also clearly influenced the<br />

extent <strong>and</strong> distribution <strong>of</strong> individual b<strong>and</strong>s <strong>and</strong> patterns <strong>of</strong> b<strong>and</strong>s, which are <strong>of</strong>ten<br />

observed to warp, terminate <strong>and</strong>/or change configuration at or near cross-bed boundaries,<br />

particularly major interdune contacts (Figure 4.4).<br />

89


(a) (b)<br />

(c) (d)<br />

Figure 4.3. Typical compaction b<strong>and</strong> patterns exposed in outcrops <strong>of</strong> the Aztec<br />

s<strong>and</strong>stone: (a) essentially parallel; (b) sub-parallel anastomosing (dominant pattern); (c)<br />

me<strong>and</strong>ering “dendritic”; (d) checkerboard.<br />

90


dune boundary<br />

dominant CB set<br />

cross-bed boundary<br />

secondary b<strong>and</strong> set<br />

dominant CB set<br />

Figure 4.4. Example <strong>of</strong> mechanical interaction between compaction b<strong>and</strong>s <strong>and</strong> æolian<br />

sedimentary architecture in the Aztec s<strong>and</strong>stone (view is to the NW). Relatively parallel<br />

compaction b<strong>and</strong>s <strong>of</strong> the dominant, steeply E-dipping set are plainly visible at both the<br />

top <strong>and</strong> bottom <strong>of</strong> the outcrop. At the fine-grained major dune boundary running<br />

diagonally across the bottom half <strong>of</strong> the photo, b<strong>and</strong>s <strong>of</strong> the dominant set either terminate<br />

or warp into a drastically different WNW-dipping orientation with a more wavy outcrop<br />

trace. At the upper cross-bed boundary, where the attitude <strong>of</strong> depostional bedding attitude<br />

reverts to that at the bottom, the dominant b<strong>and</strong> orientation, spacing <strong>and</strong> relatively<br />

smooth outcrop trace resumes.<br />

Thickness (mm)<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 6.25 12.5 18.75 25<br />

Distance in meters from left tip<br />

Figure 4.5. Tip-to-tip thickness pr<strong>of</strong>ile for a 25-m-long, relatively straight (planar)<br />

compaction b<strong>and</strong> (black squares are measurements). The dark gray curve represents the<br />

least-squares best fit <strong>of</strong> an ellipse to the data. The correlation index for this fit is 0.87, the<br />

midpoint thickness is 9.5 mm (from Sternl<strong>of</strong> et al., 2005).<br />

91


Viewed individually, CB traces tend to be straight (planar) at the scale <strong>of</strong> centimeters<br />

to meters <strong>and</strong>, where visible from tip to tip, reveal extremely eccentric ellipsoidal<br />

thickness pr<strong>of</strong>iles with aspect ratios <strong>of</strong> approximately 10 -4 (Figure 4.5). However, b<strong>and</strong>s<br />

are also commonly observed to curve, kink, zig-zag, me<strong>and</strong>er <strong>and</strong> veer toward (or away)<br />

from each other, while two or more b<strong>and</strong>s will frequently overlap <strong>and</strong> run parallel to each<br />

other in close (mm) proximity for meters to tens <strong>of</strong> meters (Figure 4.6). Separate CB<br />

traces do sometimes merge at oblique intersections into what appears in outcrop to be a<br />

single b<strong>and</strong>, but we rarely observe two CBs crossing each other except at high angle<br />

(>80°), usually as part <strong>of</strong> a checkerboard pattern. Noteworthy, because <strong>of</strong> analogous<br />

patterns documented for joint sets, is the ubiquitous presence <strong>of</strong> hooking-type interactions,<br />

where adjacent b<strong>and</strong>s spaced up to tens <strong>of</strong> centimeters curve sharply toward each other.<br />

These can involve two curving tips or one curving <strong>and</strong> one straight tip, as well as<br />

completely closed eye-shaped structures within a through-going b<strong>and</strong>. Straight tips are<br />

also observed in echelon configurations (Figure 4.7).<br />

3.2 Petrographic observations<br />

Viewed in thin section, a representative planar segment <strong>of</strong> the fins that define CBs in<br />

Aztec outcrops is clearly delineated by an abrupt decrease in porosity, which is directly<br />

attributable to inelastic mechanical compaction associated with micro-fracture<br />

accommodated plasticity <strong>of</strong> the quartz grains. The total volume change realized, as<br />

measured by the relative volume fraction occupied by detrital grains inside versus outside<br />

the b<strong>and</strong>, is about 10% <strong>and</strong> appears to be consistent along the b<strong>and</strong>s from tip to middle.<br />

Judging from the overall continuity <strong>of</strong> depositional bedding cutting across the b<strong>and</strong>s, <strong>and</strong><br />

the almost complete absence <strong>of</strong> granular disaggregation within them (despite intense<br />

grain fracturing), the volume change appears to have resulted from uniaxial compaction<br />

directed normal to the plane <strong>of</strong> the b<strong>and</strong> in the absence <strong>of</strong> appreciable shear<br />

accommodated across it (Figure 4.8).<br />

Notably lacking, or at least not yet detected by the petrographic analyses performed to<br />

date, is any halo <strong>of</strong> incipient grain damage <strong>and</strong>/or porosity loss around the elliptically<br />

tapered b<strong>and</strong> tip, as might be expected in analogy with veins, dikes, pressure solution<br />

surfaces <strong>and</strong> faults, which typically exhibit a process zone <strong>of</strong> transitional deformation<br />

grading into the undamaged host material (e.g. Delaney et al., 1986; Hoagl<strong>and</strong> et al.,<br />

92


(a)<br />

(b)<br />

(c) (d)<br />

(e) (f)<br />

(g) (h)<br />

Figure 4.6. Variety <strong>of</strong> compaction b<strong>and</strong> <strong>propagation</strong> behaviors observed in Aztec<br />

s<strong>and</strong>stone outcrop: (a) curving <strong>and</strong> intersecting; (b) complex “me<strong>and</strong>ering” interactions;<br />

(c) long, straight, sometimes closely spaced overlaps <strong>and</strong> straight, acute approaches; (d)<br />

veering <strong>and</strong> stepping between parallel b<strong>and</strong>s; (e) parallel <strong>propagation</strong> <strong>of</strong> multiple b<strong>and</strong>s<br />

spaced mm apart; (f) veering <strong>and</strong> zig-zag <strong>propagation</strong>; (g) touch-<strong>and</strong>-go intersections; (h)<br />

zig-zag <strong>propagation</strong> <strong>of</strong> closely spaced b<strong>and</strong>s.<br />

93


(a)<br />

(b)<br />

(c) (d)<br />

(e) (f)<br />

Figure 4.7. Examples <strong>of</strong> tip-to-tip interactions observed in Aztec s<strong>and</strong>stone outcrop: (a)<br />

closed eye structure between closely spaced b<strong>and</strong>s; (b) mutual hooking interaction<br />

without closure; (c) one hooking tip, one straight tip; (d) elongated eye; (e) one b<strong>and</strong><br />

hooking into another for behind its tip; (f) two straight b<strong>and</strong>s with a echelon overlap<br />

(very slight to no apprarent interaction).<br />

94


(a)<br />

(c)<br />

compaction b<strong>and</strong><br />

Porosity<br />

0.35<br />

0.30<br />

0.25<br />

0.20<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

(b)<br />

(d)<br />

0.1 0.3 0.6 1.8 4.2 8.7<br />

Distance from tip (m)<br />

Figure 4.8. Thin section image <strong>of</strong> a compaction b<strong>and</strong> within the Aztec s<strong>and</strong>stone taken<br />

with plane polarized light (quartz grains are white, feldspar grains are brownish, hematite<br />

is black, epoxy-impregnated porosity is blue). The b<strong>and</strong> as shown is approximately 9 mm<br />

thick <strong>and</strong> a coarse-fine-coarse bedding sequence running across it is indicated by the<br />

double-headed arrows. (a) Core from which this thin section was made. (b) Backscatter<br />

electron image <strong>of</strong> the (relatively) pristine s<strong>and</strong>stone matrix. (c) Backscatter electron<br />

image <strong>of</strong> the compaction b<strong>and</strong>. In the BEI images, quartz is gray, feldspar is white,<br />

speckles <strong>of</strong> dark gray are kaolinite, <strong>and</strong> black is porosity. Note the micro-fracture<br />

accommodated plasticity <strong>of</strong> the detrital quartz. (d) Clustered bar chart showing the<br />

distribution <strong>of</strong> porosity inside (middle bar) <strong>and</strong> immediately outside (left bar) the<br />

compaction b<strong>and</strong> with distance from the tip (clay included as porosity. The porosity<br />

difference (right bar) gives the relative volume strain attributable to mechanical<br />

compaction at each location. This compaction strain is remarkably uniform throughout<br />

the b<strong>and</strong> at about 10%.<br />

95


1973; Mardon, 1988; Peck et al., 1985). Given that our observations come from the tips<br />

<strong>of</strong> CBs that did naturally cease to propagate, however, raises the possibility that damage<br />

halos were originally present, but were consumed by subsequent compaction. On the<br />

other h<strong>and</strong>, small pods <strong>of</strong> grains involved in what appears to be incipient collapse can be<br />

distinguished outside <strong>of</strong> established CBs <strong>and</strong> interpreted as Griffith-type flaws or proto-<br />

b<strong>and</strong>s (Sternl<strong>of</strong> et al., 2005).<br />

3.3 Anticrack-inclusion interpretation<br />

The elliptical tip-to-tip pr<strong>of</strong>iles <strong>and</strong> uniform uniaxial internal compaction observed<br />

for straight (planar) CBs corresponds to a pure closing-mode sense <strong>of</strong> relative boundary<br />

displacement, albeit one distributed across a thin, but finite b<strong>and</strong> thickness rather than the<br />

knife-edge surface <strong>of</strong> displacement discontinuity that defines solution surfaces (Fletcher<br />

<strong>and</strong> Pollard, 1981; Mardon, 1988). This elliptical distribution <strong>of</strong> closing-mode (anti-mode<br />

I) displacement (Figure 4.9) defines the b<strong>and</strong>s kinematically <strong>and</strong> mechanically as<br />

anticracks (Mollema <strong>and</strong> Antonellini, 1996; Sternl<strong>of</strong> et al., 2005). In fact, Sternl<strong>of</strong> et al.<br />

(2005) show that the local state <strong>of</strong> stress induced in a linear elastic material by the plastic<br />

compaction <strong>of</strong> a perfectly planar CB oriented symmetric to the maximum remote<br />

compressive stress is, for all intensive purposes, independent <strong>of</strong> the evolving material<br />

properties inside the b<strong>and</strong>.<br />

Corroborating the anticrack interpretation is the fact that many <strong>of</strong> the characteristic<br />

CB patterns observed in Aztec outcrop—particularly the hook <strong>and</strong> eye configurations—<br />

are distinctly reminiscent <strong>of</strong> those described for opening-mode cracks (joints), many <strong>of</strong><br />

which can be explained as resulting from mechanical interactions between propagating<br />

tips within the framework <strong>of</strong> linear elastic fracture <strong>mechanics</strong> (e.g. Cotterell <strong>and</strong> Rice,<br />

1980; Cruikshank et al., 1991; Fleck, 1991; Olson <strong>and</strong> Pollard, 1989; Pollard <strong>and</strong> Aydin,<br />

1988; Segall <strong>and</strong> Pollard, 1983; Sempere <strong>and</strong> Macdonald, 1986; Sumi et al., 1985; Swain<br />

<strong>and</strong> Hagan, 1978; Thomas <strong>and</strong> Pollard, 1993). This suggests by analogy that CB patterns<br />

may also largely express the mechanical interactions <strong>of</strong> brittle cracks or, in this case,<br />

anticracks.<br />

Complicating this virtual anticrack approximation <strong>of</strong> CBs, however, is their physical<br />

reality as highly eccentric, oblate ellipsoidal inclusions <strong>of</strong> compacted grains, whose<br />

internal material properties must exert nonzero tractions along their boundaries with the<br />

96


(a)<br />

(b)<br />

(c)<br />

σ 3<br />

σ 1<br />

x 3<br />

x 1<br />

x 1<br />

Figure 4.9. Schematic representations <strong>of</strong> the idealized compaction b<strong>and</strong> model (from<br />

Sternl<strong>of</strong> et al., 2005). (a) Axisymmetric geometry <strong>of</strong> the eccentric ellipsoidal b<strong>and</strong><br />

aligned with the principal remote stresses. (b) Cross-sectional area <strong>of</strong> the b<strong>and</strong> (solid<br />

ellipse) relative to the pre-compacted area originally occupied by the same detrital grains<br />

(dashed ellipse). As inelastic compaction progresses, the boundary around the grains<br />

involved contracts as indicated by the displacement arrows (u1). This inward<br />

displacement <strong>of</strong> the elliptical boundary corresponds to the uniform uniaxial plastic strain<br />

<strong>of</strong> an Eshelby inclusion, <strong>and</strong> the area between the dashed <strong>and</strong> solid ellipses corresponds<br />

to the volume loss associated with the compaction. (c) Two-dimensional anticrack<br />

representation <strong>of</strong> the model b<strong>and</strong> as an elliptical distribution <strong>of</strong> closing-mode<br />

displacement discontinuity equivalent to the uniform Eshelby compaction strain. In this<br />

virtual treatment, two material lines (solid lines) interpenetrate by an amount equivalent<br />

to the volume loss associated with the compaction (dashed ellipse) as shown by the<br />

displacement arrows (u1). Actual interpenetration does not occur. Because <strong>of</strong> its extreme<br />

eccentricity, however, solutions for the state <strong>of</strong> stress induced around the model b<strong>and</strong><br />

using the Eshelby <strong>and</strong> anticrack approaches are substantially similar.<br />

97<br />

u 1<br />

u 1<br />

x 2<br />

x 2<br />

u 1<br />

σ 2


undamaged host rock. That CBs cannot exactly be represented as traction-free surfaces <strong>of</strong><br />

displacement discontinuity suggests that not all their interactions should be crack-like, as<br />

indeed some are not, notably their tendency to approach at low angle to within<br />

millimeters <strong>and</strong> then continue along parallel trends for meters. One can infer, for example,<br />

that the middle <strong>of</strong> a cm-thick CB, within which up to 50% <strong>of</strong> the available porosity has<br />

already been lost, represents a substantial physical barrier to the cross <strong>propagation</strong> <strong>of</strong> an<br />

adjacent, porosity-consuming b<strong>and</strong>. Certainly, the complex material evolution occurring<br />

inside a propagating <strong>and</strong> curving b<strong>and</strong> will affect the magnitude <strong>and</strong> distribution <strong>of</strong><br />

displacement discontinuity (shear <strong>and</strong> compaction) realized across it. Nonetheless, the<br />

linear elastic anticrack approach to modeling CB mechanical interactions at the outcrop<br />

scale is substantially justified in that the physical inclusions which define them only<br />

accumulate as a direct consequence <strong>of</strong> the predominant closing-mode sense <strong>of</strong><br />

displacement they accommodate.<br />

4. Mechanical theory<br />

Sternl<strong>of</strong> et al. (2005) provide a thorough development <strong>and</strong> justification for the 2-D<br />

(plane strain) mechanical model <strong>of</strong> an isolated, static CB as an anticrack within an<br />

infinite, homogeneous, linear elastic, isotropic material subjected to uniform principal<br />

remote compressive stress loading (σ11 r ≥ σ22 r , σ12 r = 0). So long as the scale <strong>of</strong><br />

observation exceeds a few millimeters, this constitutes a reasonable approximation <strong>of</strong><br />

CBs in the porous, granular Aztec s<strong>and</strong>stone as visible in the sub-horizontal plane <strong>of</strong><br />

outcrop exposure (Sternl<strong>of</strong> et al., 2005). In this section, we exp<strong>and</strong> the anticrack model to<br />

consider the theoretical implications for <strong>propagation</strong> <strong>and</strong> interaction between multiple<br />

b<strong>and</strong>s, with reference by analogy to classical elastic fracture <strong>mechanics</strong> (Irwin, 1960;<br />

Lawn <strong>and</strong> Wilshaw, 1975) as successfully applied to natural opening-mode fractures in<br />

rock <strong>and</strong> summarized by Olson <strong>and</strong> Pollard (1989) <strong>and</strong> Thomas <strong>and</strong> Pollard (1993).<br />

Throughout the rest <strong>of</strong> the paper, compression, compaction, <strong>and</strong> left-lateral senses <strong>of</strong><br />

shear stress <strong>and</strong> displacement are taken as positive, as illustrated in Figure 4.10.<br />

For an isolated, perfectly linear (planar) anticrack oriented symmetric to σ11 r , the<br />

local state <strong>of</strong> stress induced in the near-tip region <strong>and</strong> expressed in the global (remote<br />

principal) coordinate system resembles the example distributions shown in Figure 4.11.<br />

High concentrations <strong>of</strong> compressive normal stress (σ11 <strong>and</strong> σ22) are readily apparent<br />

98


σ22 r<br />

σ21 x 1<br />

x 3<br />

x 2<br />

r<br />

σ12 D s<br />

r<br />

σ11 D n<br />

θ σ θθ<br />

Figure 4.10. Schematic diagram <strong>of</strong> the global 3-D cartesian (x1, x2, x3), <strong>and</strong> the 2-D local<br />

anticrack tip cartesian (x1, x2) <strong>and</strong> polar (r,θ) coordinate systems. All components <strong>of</strong><br />

stress (σij) <strong>and</strong> displacement discontinuity (Dn, Ds) are shown as positive.<br />

99<br />

D s<br />

D n<br />

x 1 c<br />

r<br />

σ rθ<br />

x 2 c<br />

σ rr


0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

−0.1<br />

−0.2<br />

−0.3<br />

−0.4<br />

−0.5<br />

−0.5 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4 0.5<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

−0.1<br />

−0.2<br />

−0.3<br />

−0.4<br />

0<br />

−0.1<br />

−0.2<br />

−0.3<br />

−0.4<br />

(a)<br />

−0.5<br />

−0.5 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4 0.5<br />

(b)<br />

(c)<br />

−0.5<br />

−0.5 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4 0.5<br />

Figure 4.11. Contour plots <strong>of</strong> the near tip stress fields generated around an anticrack<br />

(white slot) oriented symmetric to the remote principal stresses (in this case an isotropic<br />

state <strong>of</strong> remote stress). Dimensions are in meters, stress magnitude are normalized by the<br />

remote value: (a) σ11. (b) σ22 (c) σ12.<br />

100<br />

0<br />

2.8<br />

2.6<br />

2.4<br />

2.2<br />

2<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

2.8<br />

2.6<br />

2.4<br />

2.2<br />

2<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

−0.1<br />

−0.2<br />

−0.3<br />

−0.4


immediately ahead <strong>of</strong> the tip, while shear stress (σ12) is distributed in an anti-symmetric<br />

quadrant pattern around it. Because each new increment <strong>of</strong> growth extends a small<br />

distance (r) from the existing tip, it is natural to think in terms <strong>of</strong> a local polar coordinate<br />

system centered at the tip (Figure 4.10), with <strong>propagation</strong> occurring in the direction for<br />

which the circumferential normal stress (σθθ) is a maximum, provided that this exceeds<br />

the compressive yield strength <strong>of</strong> the material. Expressed in terms <strong>of</strong> opening-mode<br />

cracks (tension positive), this is the maximum circumferential stress criterion <strong>of</strong> Erdogan<br />

<strong>and</strong> Sih (1963).<br />

For the symmetric anticrack configuration described above, representing pure mode I<br />

loading <strong>and</strong> displacements, the distribution <strong>of</strong> the polar-coordinate stresses as a function<br />

<strong>of</strong> θ resembles those illustrated in Figure 4.12a, where σθθ max occurs at θ = 0. This<br />

indicates that <strong>propagation</strong> would continue along a straight path. If even a small amount <strong>of</strong><br />

positive mode II displacement (left-lateral shear) is resolved across the tip <strong>of</strong> the<br />

anticrack, however, the next increment <strong>of</strong> <strong>propagation</strong> will diverge at a substantial angle<br />

to the negative (clockwise) side <strong>of</strong> the θ = 0 axis (Figure 4.12b). In effect, the anticrack<br />

tip attempts to reorient itself to remain symmetric to the perturbed local maximum<br />

principal compressive stress (i.e. along a path for which the local shear stress is zero).<br />

Successive small increments <strong>of</strong> shearing along the <strong>propagation</strong> path will result in a<br />

smoothly curving CB.<br />

Excluding variations in effective remote loading (e.g. due to tectonic changes,<br />

regional faulting or slip along major dune boundaries), there are two primary ways for<br />

shear to become resolved on the model anticrack tip: either it deviates from its symmetric<br />

path due to some grain-scale heterogeneity, or it propagates into the perturbed stress field<br />

generated by an adjacent anticrack. The first case results in the resolution <strong>of</strong> shear on the<br />

deviant tip from the remote loading. If σ11 r > σ22 r (i.e. a significant remote differential<br />

stress exists), then the sense <strong>of</strong> shear resolved rotates σθθ max back toward the original<br />

symmetric <strong>propagation</strong> path inhibiting further deviation (Figure 4.13a). As σ11 r<br />

approaches σ22 r in compressive magnitude, however (minimal remote differential stress),<br />

little or no path-correcting shear is resolved on the tip (Figure 4.13b). This analysis<br />

corresponds to that <strong>of</strong> Cotterell <strong>and</strong> Rice (1980) on the effect <strong>of</strong> remote crack-parallel<br />

stress on the <strong>propagation</strong> stability <strong>of</strong> opening-mode cracks except that, because anticracks<br />

101


Normalized stress magnitude<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

(a)<br />

−1.5<br />

−200 −150 −100 −50 0 50 100 150 200<br />

θ in degrees (local tip coordinate system<br />

Normalized stress magnitude<br />

4<br />

3<br />

2<br />

1<br />

0<br />

−1<br />

−2<br />

(b)<br />

−3<br />

−200 −150 −100 −50 0 50 100 150 200<br />

θ in degrees (local tip coordinate system<br />

Figure 4.12. Plots <strong>of</strong> the near-tip stress magnitudes expressed in polar coordinates as a<br />

function <strong>of</strong> θ for a fixed r (see Figure 4.10). (a) Distribution <strong>of</strong> stresses for a pure closingmode<br />

sense <strong>of</strong> displacement discontinuity. (b) Distribution <strong>of</strong> stresses for mixed-mode<br />

displacement discontinuity (shear component positive). Blue lines are σθθ, red lines are<br />

σrr <strong>and</strong> green lines are σrθ.<br />

(a)<br />

r<br />

σ22 (b)<br />

r<br />

σ22 r<br />

σ11 r<br />

σ11 Figure 4.13. Schematic representation <strong>of</strong> the effect <strong>of</strong> remote differential stress on<br />

<strong>propagation</strong> path stability. (a) When the least compressive remote stress acting parallel to<br />

the anticrack b<strong>and</strong> is significantly less than the maximum compressive remote stress<br />

(high differential stress) any deviation from symmetric <strong>propagation</strong> results in shear<br />

stresses being resolved on the deviant tip such that it turns back toward the original path.<br />

(b) When the remote stress state is nearly isotropic, the tendency for self-correction is<br />

diminished.<br />

102<br />

α<br />

α<br />

β<br />

β


form perpendicular to σ11 r , increasing σ22 r acts to reduce the effective differential stress<br />

<strong>and</strong> inhibits, rather than enhances path stability.<br />

In the second case, when the stress perturbations generated by adjacent anticracks<br />

impinge, the effect can be understood by holding one stationary <strong>and</strong> considering how its<br />

near-tip shear stress field (σ12) resolves onto the other tip propagating along a parallel<br />

path. For example, if tip B approaches tip A from the upper right (Figure 4.14a), it first<br />

enters an area where the sense <strong>of</strong> shear imposed on it by tip A is mildly positive (left<br />

lateral). This rotates the direction <strong>of</strong> σθθ max in front <strong>of</strong> B slightly clockwise, causing it to<br />

curve gently away from A. As B continues on its now sub-parallel path to the left, it<br />

enters the area where the sense <strong>of</strong> shear induced by A becomes distinctly negative (right<br />

lateral), rotating the direction <strong>of</strong> σθθ max abruptly counterclockwise <strong>and</strong> causing B to curve<br />

strongly toward A. Conversely, if tip B overtakes tip A from the upper left (Figure 4.14b),<br />

the sense <strong>and</strong> magnitude <strong>of</strong> shear imposed at first cause it to curve strongly away from<br />

the plane <strong>of</strong> A (counterclockwise), <strong>and</strong> then gently toward it (clockwise). The net result is<br />

that, if B approaches A, it steps closer to it along a curving path, while if B overtakes A,<br />

it steps away. In both cases, the senses <strong>of</strong> curving <strong>and</strong> stepping are the same (mirror<br />

images) if B propagates along a parallel path on the other side <strong>of</strong> A. If both A <strong>and</strong> B<br />

propagate, then both paths also curve as described.<br />

The path altering effects <strong>of</strong> the local <strong>and</strong> remote stress fields superimpose, generally<br />

in opposition: local effects acting to curve <strong>propagation</strong> paths out <strong>of</strong> symmetry with the<br />

remote principal stress state; <strong>and</strong> the remote differential stress acting to maintain<br />

symmetry. Three additional contributors to path stability for a given CB anticrack would<br />

be the geometry <strong>of</strong> the path it has already traveled, the distribution <strong>of</strong> inelastic<br />

(unrecoverable) relative boundary displacements realized along that path, <strong>and</strong> the<br />

physical properties (elastic or otherwise) <strong>of</strong> the compacted material inside it. All three <strong>of</strong><br />

these would tend to enhance <strong>propagation</strong> path stability, ins<strong>of</strong>ar as shear tractions imposed<br />

on a curving tip would be transmitted into the existing b<strong>and</strong> <strong>and</strong> resisted by its fixed<br />

physical attributes. The nature <strong>of</strong> these effects can be thought <strong>of</strong> as the inertia <strong>of</strong> a CB to<br />

changes in its path. Furthermore, the nonzero shear tractions imposed by the compacted<br />

material inside a CB on its boundaries with the surrounding pristine s<strong>and</strong>stone preclude<br />

the local principal state <strong>of</strong> stress from being symmetric to those boundaries. This fact<br />

103


helps to explain why the oblique approach <strong>of</strong> one b<strong>and</strong> to the flank <strong>of</strong> another rarely leads<br />

to intersection, as the local direction <strong>of</strong> σθθ max can never be parallel to the b<strong>and</strong> flank <strong>and</strong><br />

thus cannot lead the propagating tip on a direct collision course with the adjacent b<strong>and</strong><br />

flank.<br />

Finally we note that, unlike opening-mode cracks, which generally are modeled as<br />

shear traction-free boundary surfaces pressed open by an internal fluid pressure that<br />

constitutes an independently variable driving force for <strong>propagation</strong> (e.g. Cotterell <strong>and</strong><br />

Rice, 1980), anticrack CBs are compaction structures whose <strong>propagation</strong> can only be<br />

driven by near-tip compressive stress concentrations generated as a consequence <strong>of</strong><br />

previous compaction, which must ultimately result from the failure <strong>of</strong> grain-scale flaws<br />

within the ambient remote stress field (Sternl<strong>of</strong> et al., 2005). This amounts to a stringent<br />

constraint on initiation <strong>and</strong> <strong>propagation</strong> that renders the dense variety <strong>of</strong> b<strong>and</strong> patterns<br />

pervading the Aztec s<strong>and</strong>stone all the more remarkable.<br />

5. Numerical model<br />

In order to examine the stress <strong>and</strong> strain fields generated by CBs <strong>and</strong> model their<br />

<strong>propagation</strong> <strong>and</strong> interaction as anticracks, we developed a MATLAB® code based on the<br />

displacement discontinuity boundary element method <strong>of</strong> Crouch <strong>and</strong> Starfield (1983). In<br />

our formulation <strong>of</strong> this 2-D, linear elastic plane strain model, which corresponds well<br />

with the character <strong>of</strong> the field data <strong>and</strong> interpretations described above, we consider an<br />

infinite whole space subjected to uniform principal stresses at infinity <strong>and</strong> represent the<br />

b<strong>and</strong>s as segmented traces <strong>of</strong> linear boundary elements (Figure 4.15). Across each <strong>of</strong><br />

these elements, two independent <strong>and</strong> uniformly distributed components <strong>of</strong> displacement<br />

discontinuity, or dislocation, can be accommodated: Dn (normal) <strong>and</strong> Ds (shear). The<br />

displacements realized on each element affect those on all others <strong>and</strong> depend on the<br />

remote loading, as well as the traction boundary conditions (normal <strong>and</strong> shear) which<br />

must be specified for every element. In order to determine the two unknown components<br />

<strong>of</strong> displacement for all N elements, while matching the specified traction boundary<br />

conditions, requires the formulation <strong>and</strong> solution <strong>of</strong> a traction boundary value problem<br />

consisting <strong>of</strong> a linear system <strong>of</strong> 2N equations that can be summarized in matrix notation<br />

as:<br />

tn = E·dn + F·ds (1)<br />

104


0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

−0.1<br />

−0.2<br />

−0.3<br />

−0.4<br />

(a)<br />

−0.5<br />

−0.5 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4 0.5<br />

0.5 0.4<br />

0.4<br />

0.3<br />

0.3<br />

0.2<br />

0.2<br />

0.1<br />

0.1<br />

0<br />

0<br />

−0.1 −0.1<br />

−0.2<br />

−0.2<br />

−0.3<br />

−0.3<br />

−0.4<br />

−0.4<br />

−0.5<br />

−0.5 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4 0.5<br />

Figure 4.14. Schematic illustration <strong>of</strong> how one anticrack b<strong>and</strong> tip influences the<br />

<strong>propagation</strong> path <strong>of</strong> an approaching tip through the action <strong>of</strong> the near-tip shear stress field<br />

it generates. (a) A b<strong>and</strong> entering from the upper right first experiences a slight positive<br />

shear stress which causes its path to diverge. When it enters the zone stronger, negative<br />

shear stress, the approaching b<strong>and</strong> turns to follow a strongly convergent path. (b) A b<strong>and</strong><br />

entering from the upper left at first strongly diverges <strong>and</strong> then weakly converges. In both<br />

cases, if the b<strong>and</strong> approached from the bottom side, the resulting <strong>propagation</strong> paths would<br />

mirror those shown.<br />

r<br />

σ22 x 1<br />

x 2<br />

Figure 4.15. Schematic illustration <strong>of</strong> the essential aspects <strong>of</strong> the displacement<br />

discontinuity boundary element method model.<br />

(b)<br />

105<br />

r<br />

σ11 D n<br />

D s<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

−0.1<br />

−0.2<br />

−0.3<br />

−0.4


<strong>and</strong><br />

ts = G·ds + H·dn (2)<br />

where tn <strong>and</strong> ts are column vectors containing the normal <strong>and</strong> shear traction boundary<br />

conditions prescribed for the midpoint <strong>of</strong> each successive element 1 through N; dn <strong>and</strong> ds<br />

are column vectors representing the unknown components <strong>of</strong> displacement discontinuity,<br />

Dn <strong>and</strong> Ds., for each successive element 1 through N; <strong>and</strong> E, F,G <strong>and</strong> H are N x N<br />

matrices containing the influence coefficients that relate displacement on one element to<br />

traction on another. For example, the component Eij <strong>of</strong> E gives the normal traction on the<br />

ith element due to a unit normal displacement occurring on the jth element.<br />

Analytical functions relate the displacement discontinuity on each element to the<br />

stress field it generates in the surrounding elastic medium (Crouch <strong>and</strong> Starfield, 1983).<br />

The complete state <strong>of</strong> stress at any point within the medium can be determined by<br />

superimposing the contributions from each anticrack boundary element <strong>and</strong> the remote<br />

values. Thus, once Dn <strong>and</strong> Ds have been determined for all N elements representing a<br />

given b<strong>and</strong> geometry, the code uses them to determine the magnitude <strong>and</strong> location <strong>of</strong><br />

σθθ max along a circle <strong>of</strong> specified radius centered on the model b<strong>and</strong>. If this value exceeds<br />

the prescribed threshold limit for <strong>propagation</strong>, an additional displacement discontinuity<br />

element is appended in the appropriate direction, creating a new system <strong>of</strong> 2(N+1)<br />

equations to be solved. By iterating this process, <strong>propagation</strong> is simulated.<br />

As currently written, the code allows for any combination <strong>of</strong> traction <strong>and</strong><br />

displacement discontinuity boundary conditions to be specified for each element. If only<br />

boundary displacements are specified, then the problem is fully constrained <strong>and</strong> the<br />

solution is already in h<strong>and</strong>. If one or both displacement discontinuities are not specified,<br />

compensating traction boundary conditions are determined during the solution step based<br />

on elastic modulii that must be prescribed for each element. Other schemes for applying<br />

boundary conditions on the elements without assuming linear elastic behavior (e.g.<br />

nonlinear elasticity, plasticity <strong>and</strong> frictional sliding) would allow for the prescription <strong>of</strong> a<br />

greater variety <strong>of</strong> potentially more realistic internal CB behaviors. These improvements<br />

are currently being developed <strong>and</strong> implemented.<br />

Finally we re-emphasize that, while the code treats positive Dn as a material<br />

interpenetration across the elements, actual CBs represent porosity loss across a b<strong>and</strong> <strong>of</strong><br />

106


thin, but finite thickness. It is the extreme elliptical eccentricity <strong>of</strong> the b<strong>and</strong> trace that<br />

justifies the anticrack approximation in a BEM treatment (Sternl<strong>of</strong> et al., 2005).<br />

Nonetheless, nothing in the theory or implementation <strong>of</strong> linear elastic fracture <strong>mechanics</strong><br />

precludes the virtual interpenetration <strong>of</strong> anticrack walls, as represented in our treatment<br />

by positive Dn. By the same token, the code works just as well for opening-mode cracks,<br />

with the signs <strong>of</strong> the stresses <strong>and</strong> displacements reversed.<br />

6. Propagation simulation<br />

As a first step toward realistic simulation <strong>of</strong> CB <strong>propagation</strong>, interaction <strong>and</strong> pattern<br />

development, we used the model code to investigate the <strong>propagation</strong> behavior <strong>of</strong> a few<br />

simple configurations comprised <strong>of</strong> one or two anticrack b<strong>and</strong>s. In order to generate<br />

results comparable to the field observations, only values for the model parameters<br />

corresponding to estimates for the Aztec s<strong>and</strong>stone during CB formation are used<br />

(Sternl<strong>of</strong> et al., 2005). For the remote stress conditions, these are σ11 r = 40 Mpa, 20 Mpa<br />

≤ σ22 r ≤ 40 Mpa, <strong>and</strong> σ12 r = 0. For the elastic properties <strong>of</strong> the pristine (CB-free) infinite<br />

medium, these are Young’s modulus (Es) equals 20 GPa <strong>and</strong> Poisson’s ratio (νs) equals<br />

0.2, such that the shear modulus (Gs) equals 8.3 GPa. Finally, we normalized by 40 MPa<br />

to produce a problem dimensionless in stress, such that σ11 r = 1, 0.5 ≤ σ22 r ≤ 1, σ12 r = 0<br />

<strong>and</strong> Es = 500.<br />

6.1. Model calibration <strong>and</strong> stability testing<br />

As described in Section 2, the anticrack model for CBs is based on detailed<br />

observations, <strong>and</strong> thickness <strong>and</strong> porosity measurements <strong>of</strong> relatively isolated, straight<br />

(planar) b<strong>and</strong>s. These data established the physical reality <strong>of</strong> the anticrack-like elliptical<br />

distribution <strong>of</strong> pure closing-mode displacement for such CBs (Sternl<strong>of</strong> et al., 2005).<br />

However, in order to produce unconstrained <strong>and</strong> potentially predictive simulations <strong>of</strong><br />

<strong>propagation</strong> <strong>and</strong> interaction that are not perfectly symmetric to the remote stress field, we<br />

must leave all boundary element displacements unspecified <strong>and</strong> let the model determine<br />

their equilibrium values. Currently, this means specifying linear elastic modulii (Eb <strong>and</strong><br />

νb) for every element. Given the inelastic nature <strong>of</strong> compaction apparent inside CBs, this<br />

approach to applying boundary conditions has limited appeal others are being developed,<br />

as mentioned above. Nonetheless, if the rate <strong>of</strong> <strong>propagation</strong> was fast relative to the rate <strong>of</strong><br />

107


plastic compaction, as suggested by the elliptical b<strong>and</strong> pr<strong>of</strong>iles, then the assumption <strong>of</strong> at<br />

least quasi-elastic behavior inside the b<strong>and</strong>s on the time scale <strong>of</strong> <strong>propagation</strong> might be<br />

justified (Sternl<strong>of</strong> et al., 2005).<br />

Despite these uncertainties, the distribution <strong>of</strong> Dn measured for the 25-m-long CB<br />

pr<strong>of</strong>ile featured by Sternl<strong>of</strong> et al. (2005) can be used to estimate an approximate effective<br />

elastic stiffness (Eb) for the anticrack elements (Figure 4.16). For Eb = 0 (i.e. a perfect,<br />

traction-free anticrack), the model calculates a distribution <strong>of</strong> Dn two orders <strong>of</strong> magnitude<br />

too high (~10 cm versus ~1 mm in the middle <strong>and</strong> ~1 cm versus ~.1 mm at the tips). The<br />

addition <strong>of</strong> even a small amount <strong>of</strong> element stiffness, however, drastically reduces Dn,<br />

<strong>and</strong> we found that setting Eb = 15 (3% <strong>of</strong> Es) produced a slight over estimate. Of course,<br />

the measured distribution <strong>of</strong> Dn could be matched exactly by independently varying Eb<br />

for every element, but the benefits <strong>of</strong> such specificity do not justify the effort, particularly<br />

as the assumption <strong>of</strong> elastic behavior is dubious at best. For our immediate purposes, it is<br />

enough that the near-tip magnitudes <strong>of</strong> Dn calculated by the model generally coincide<br />

with those measured for natural b<strong>and</strong>s, since these displacement discontinuities, in<br />

concert with any Ds accommodated, dictate the stress perturbations produced outside the<br />

b<strong>and</strong>.<br />

The next aspect <strong>of</strong> essential model calibration involves choosing the characteristic<br />

distance r at which σθθ max will be calculated. As demonstrated by Sternl<strong>of</strong> et al. (2005),<br />

because the model represents CBs as trains <strong>of</strong> constant displacement dislocations, stress<br />

magnitudes calculated in the very near-tip field (r < 0.1 mm) vary as 1/r (Figure 4.17).<br />

For r ranging from about 0.5 mm to 5 mm, the calculated stresses vary along a more a<br />

crack-like 1/√r trend. Beyond 5mm, the BEM model-generated stresses basically<br />

coincide with those calculated analytically using the Eshelby inclusion method (Eshelby,<br />

1959; Mura, 1987), which we consider the most accurate representation <strong>of</strong> the ideal,<br />

isolated CB based on the field data, <strong>and</strong> which predicts a normalized σθθ max <strong>of</strong> about 8.75<br />

as r → 0. The stress singularity generated by the model at r = 0 has no significance in<br />

physical reality, both because the nature <strong>of</strong> deformation at the tip is actually plastic, <strong>and</strong><br />

because the assumption <strong>of</strong> homogeneous, isotropic elasticity cannot be presumed to hold<br />

as r drops below the grain size <strong>of</strong> the s<strong>and</strong>stone. In fact, with an average grain size <strong>of</strong><br />

~0.2 mm, any r less than about 5 mm becomes suspect (Amadei <strong>and</strong> Stephansson, 1997).<br />

108


Distribution <strong>of</strong> Dn in millimeters<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

Eb=10<br />

Eb=12.5<br />

Eb=15<br />

Eb=17.5<br />

Eb=20<br />

Field data<br />

0<br />

−15 −10 −5 0 5 10 15<br />

Distance in meters from center <strong>of</strong> b<strong>and</strong> pr<strong>of</strong>ile<br />

Figure 4.16. Distributions <strong>of</strong> closing-mode displacement discontinuity (Dn) predicted by<br />

the BEM model as a function <strong>of</strong> varying internal element stiffness (Eb) for a 25-m-long<br />

b<strong>and</strong> trace. Best-fit ellipse to Dn data as calculated from field measurements shown by<br />

thick black curve.<br />

MPa<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 −5 10 −4 10 −3 10 −2 10 −1<br />

10 0<br />

Distance from tip, r (m)<br />

Figure 4.17. Log-log plot <strong>of</strong> the distribution <strong>of</strong> σ11 with distance from the model<br />

compaction b<strong>and</strong> tip: BEM anticrack solution (solid line), Eshelby inclusion solution<br />

(dashed line), crack-like contours <strong>of</strong> 1/sqrt(r) (dotted lines), dislocation-like contours <strong>of</strong><br />

1/r (dash-dot lines). At the distance r = 1 cm from the tip, the two solutions effectively<br />

coincide. As r decreases, the Eshelby solution approaches the 1/sqrt(r) distribution,<br />

although it remains finite. The BEM anticrack solution, which is composed <strong>of</strong> a step-wise<br />

distribution <strong>of</strong> constant closing-mode displacement discontinuity elements, goes to<br />

infinity as 1/r (from Sternl<strong>of</strong> et al., 2005).<br />

109


To determine the sensitivity <strong>of</strong> <strong>propagation</strong> path to the choice <strong>of</strong> r, we performed a<br />

series <strong>of</strong> stability tests using the same model b<strong>and</strong> that produced Figures 4.16 <strong>and</strong> 17. We<br />

added a 2-cm element to one tip at a variable kink angle <strong>of</strong> α, <strong>and</strong> plotted the relationship<br />

between α <strong>and</strong> the angle at which the next increment <strong>of</strong> <strong>propagation</strong> was predicted to<br />

occur (β) for r = 1, 5 <strong>and</strong> 10 mm, <strong>and</strong> the bracketing remote differential stress magnitudes<br />

(σ d = σ11 r -σ22 r ) <strong>of</strong> zero <strong>and</strong> 0.5 (Figure 4.18). The response in every case is to correct the<br />

deviation such that β is negative for positive α along a nearly linear trend. When the<br />

remote stress state is isotropic, |β| ≤ |α| <strong>and</strong> symmetry with σ11 r is nearly restored with the<br />

next increment <strong>of</strong> <strong>propagation</strong>. When σ d is the maximum <strong>of</strong> 0.5, |β| ≥ |α| <strong>and</strong> the next<br />

increment <strong>of</strong> <strong>propagation</strong> over-corrects for the original deviation. The effect <strong>of</strong> varying r<br />

at the two stress states is also opposite: for σ d = 0, |β| varies directly with r; for σ d = 0.5,<br />

|β| varies inversely with r. Discarding r = 1 mm as unrealistically small for the reasons<br />

stated above, we observe that β never varies by more than 4° for a given α between r = 5<br />

mm <strong>and</strong> r = 10 mm. Due to this limited effect, <strong>and</strong> to preserve the plausibility <strong>of</strong> the<br />

linear elastic continuum assumption, we chose r = 5 mm for all simulations.<br />

Finally, as the model uses linear segments to approximate CB paths that generally are<br />

smoothly curving, we examined the stability <strong>of</strong> the <strong>propagation</strong> direction as a function <strong>of</strong><br />

the dislocation element length used. Naturally, the shorter the elements, the better the<br />

model fit to the ideal (i.e. smooth) path. As a practical matter <strong>of</strong> computational efficiency,<br />

however, longer elements are desirable. We performed a second suite <strong>of</strong> kink-angle<br />

stability tests for tip-element lengths <strong>of</strong> 2, 6 <strong>and</strong> 10 cm at r = 5 mm (Figure 4.19). Again,<br />

in every case β acts to oppose α along a nearly linear trend, with the response being<br />

strongest for σ d = 0.5. The sense <strong>of</strong> the effect at the two stress states is also opposite,<br />

with |β| varying directly with element length for σ d = 0.5, <strong>and</strong> indirectly for σ d = 0. The<br />

spread in the directional variability is about three times greater at σ d = 0.5 <strong>and</strong> is equal to<br />

about 85% <strong>of</strong> α. These results are entirely consistent with those reported in previous<br />

studies <strong>of</strong> <strong>propagation</strong> path stability for segmented opening-mode cracks using the<br />

kinked-tip approach (Broberg, 1987; Cotterell <strong>and</strong> Rice, 1980; Melin, 1983, 1987;<br />

Thomas <strong>and</strong> Pollard, 1993), which indicate that the misorientation <strong>of</strong> successive<br />

110


Correction angle (β) in degrees<br />

-100<br />

-90<br />

-80<br />

-70<br />

-60<br />

-50<br />

-40<br />

-30<br />

-20<br />

-10<br />

r=1mm, Sx=.5Sy<br />

r=5mm, Sx=.5Sy<br />

r=10mm, Sx=.5Sy<br />

r=1mm, Sx=Sy<br />

r=5mm, Sx=Sy<br />

r=10mm, Sx=Sy<br />

D n<br />

0<br />

0 5 10 15 20 25 30 35 40 45<br />

Kink angle (α) in degrees<br />

Figure 4.18. Relationship between the angle <strong>of</strong> incremental deviation from symmetric<br />

<strong>propagation</strong> (kink angle α) to the subsequent angle <strong>of</strong> incremental path correction (β) as a<br />

function <strong>of</strong> both r <strong>and</strong> the remote state <strong>of</strong> principal stress (Sy = σ1, Sx = σ2). Note that the<br />

sign <strong>of</strong> β is negative for positive α in all cases.<br />

Correction angle (β) in degrees<br />

-100<br />

-90<br />

-80<br />

-70<br />

-60<br />

-50<br />

-40<br />

-30<br />

-20<br />

-10<br />

2cm, Sx=.5Sy<br />

6cm, Sx=.5Sy<br />

10cm, Sx=.5Sy<br />

2cm, Sx=Sy<br />

6cm, Sx=Sy<br />

10cm, Sx=Sy<br />

0<br />

0 5 10 15 20 25 30 35 40 45<br />

Kink angle (α) in degrees<br />

Figure 4.19. Relationship between the angle <strong>of</strong> incremental deviation from symmetric<br />

<strong>propagation</strong> (kink angle α) to the subsequent angle <strong>of</strong> incremental path correction (β) as a<br />

function <strong>of</strong> both element length <strong>and</strong> the remote state <strong>of</strong> principal stress (Sy = σ1, Sx = σ2).<br />

Note that the sign <strong>of</strong> β is negative for positive α in all cases, <strong>and</strong> that geometric<br />

formulation <strong>of</strong> problem is as shown in Figure 4.18.<br />

111<br />

D n<br />

σ 1<br />

α<br />

Ds<br />

β<br />

+<br />

−<br />

D n<br />

σ 2


segments will self-correct to provide a reasonable, zig-zag approximation to the ideal<br />

path, certainly for |β| less than ~10° so long as σ d ≠ 0.<br />

In all the simulations presented below, we therefore adopted fixed values <strong>of</strong> Eb = 15<br />

<strong>and</strong> r = 5 mm, <strong>and</strong> use 10 cm <strong>propagation</strong> elements so long as β < 10°. Ins<strong>of</strong>ar as<br />

necessary to maintain numerical stability, <strong>and</strong> capture details <strong>of</strong> strong curvature <strong>and</strong> the<br />

close approach <strong>of</strong> adjacent b<strong>and</strong>s, we reduced the added element length to 2 cm. We also<br />

verified that at no point in any simulation did Dn on any element become positive (i.e.<br />

opening-mode). Finally, given the current lack <strong>of</strong> data on a critical σθθ threshold for<br />

<strong>propagation</strong> to occur, we used a model threshold <strong>of</strong> σθθ max > σ11 r = 1. In practice, the<br />

normalized magnitude <strong>of</strong> σθθ max never dropped below about 1.25.<br />

6.2. Symmetric <strong>propagation</strong><br />

In order to examine the tendency for CBs to form symmetric (orthogonal) to σ11 r , we<br />

simulated the <strong>propagation</strong> <strong>of</strong> a short, incipient b<strong>and</strong> (three elements, 6 cm) over a range<br />

<strong>of</strong> misalignment angles α (relative to the global coordinate system) for each <strong>of</strong> the remote<br />

stress states σ d = 0.05, 0.25 <strong>and</strong> 0.5 (Figure 4.20). Not surprisingly, the results indicate<br />

that the tendency <strong>of</strong> a misaligned incipient CB to propagate into symmetry with the<br />

remote stress field decreases as σ d → 0. For a perfectly isotropic state <strong>of</strong> remote stress,<br />

all b<strong>and</strong> trends are preferred equally. For the maximum probable remote differential<br />

stress (σ d = 0.5), the b<strong>and</strong> trend orthogonal to σ11 r is strongly preferred. Even at σ d =<br />

0.25 (σ22 r = 0.75σ11 r ), the symmetric orientation is clearly preferred, particularly if one<br />

allows for the fact that 6 cm greatly exaggerates the length a misaligned incipient CB<br />

would attain without starting to curve.<br />

Another factor to consider in assessing the tendency toward symmetric <strong>propagation</strong> is<br />

how σθθ max at the tips <strong>of</strong> the incipient model CB varies with α (Figure 4.20d). Even for<br />

σ d = 0.05, a slight but appreciable increase <strong>of</strong> 5.4% in σθθ max is realized as α goes from<br />

90° to 0°. For σ d = 0.25 <strong>and</strong> 0.5, the increase is 33.3% <strong>and</strong> 100%, respectively. These<br />

results demonstrate that, even as σ d → 0, the <strong>propagation</strong> <strong>of</strong> incipient b<strong>and</strong>s oriented<br />

more nearly orthogonal to σ11 r is preferred, likely str<strong>and</strong>ing those oriented at higher α.<br />

112


Distance in meters<br />

Distance in meters<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

−0.05<br />

−0.1<br />

0.1<br />

0.05<br />

0<br />

−0.05<br />

−0.1<br />

σ 1<br />

α<br />

σ 2<br />

(a)<br />

1 deg.<br />

5 deg.<br />

15 deg<br />

30 deg.<br />

45 deg.<br />

60 deg.<br />

75 deg.<br />

85 deg.<br />

89 deg.<br />

−0.15<br />

−0.15 −0.1 −0.05 0 0.05<br />

Distance in meters<br />

0.1 0.15<br />

0.15<br />

(c)<br />

1 deg.<br />

5 deg.<br />

15 deg<br />

30 deg.<br />

45 deg.<br />

60 deg.<br />

75 deg.<br />

85 deg.<br />

89 deg.<br />

−0.15<br />

−0.15 −0.1 −0.05 0 0.05 0.1 0.15<br />

Distance in meters<br />

Distance in meters<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

−0.05<br />

−0.1<br />

(b)<br />

1 deg.<br />

5 deg.<br />

15 deg<br />

30 deg.<br />

45 deg.<br />

60 deg.<br />

75 deg.<br />

85 deg.<br />

89 deg.<br />

−0.15<br />

−0.15 −0.1 −0.05 0 0.05<br />

Distance in meters<br />

0.1 0.15<br />

3<br />

2.8<br />

(d)<br />

Maximum tangential stress<br />

2.6<br />

2.4<br />

2.2<br />

2<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

Diff. stress=0.5<br />

Diff. stress=0.25<br />

Diff. stress=0.05<br />

1<br />

0 10 20 30 40 50 60 70 80 90<br />

Misalignment angle (α) in degrees<br />

Figure 4.20. Propagation <strong>of</strong> an incipient anticrack b<strong>and</strong> as a function <strong>of</strong> misalignment<br />

angle α from the symmetric orientation for normalized remote differential stress<br />

magnitudes <strong>of</strong> 0.5 (a), 0.25 (b) <strong>and</strong> 0.05 (c). A schematic <strong>of</strong> the model geometry is shown<br />

in the upper left <strong>of</strong> plot (a). Each <strong>propagation</strong> simulation started from an initially straight<br />

b<strong>and</strong> 6 cm long <strong>and</strong> centered at the origin. Plot (d) shows the normalized value <strong>of</strong> the<br />

maximum tangential stress at the tip <strong>of</strong> the initial b<strong>and</strong> as a function <strong>of</strong> α <strong>and</strong> the remote<br />

differential stress magnitude.<br />

113


6.3. Approaching tip interactions<br />

Aside from the effects <strong>of</strong> heterogeneous stress fields <strong>and</strong>/or mechanical anisoptropies<br />

related to sedimentary architecture (e.g. shear along fine-grained deflation-plane<br />

boundaries separating major, teardrop-shaped dune sequences), which this paper does not<br />

attempt to address, the primary source <strong>of</strong> mechanical interaction for a propagating CB is<br />

the close approach <strong>of</strong> an adjacent tip. Within the framework <strong>of</strong> our simple model, there<br />

are two primary variables that can affect the mechanical interaction <strong>of</strong> straight b<strong>and</strong>s<br />

propagating toward each other along parallel trends symmetric to σ11 r : the spacing<br />

between them <strong>and</strong> the magnitude <strong>of</strong> the remote differential stress. In order to examine the<br />

interplay <strong>of</strong> these variables in determining the degree <strong>of</strong> interaction realized between<br />

b<strong>and</strong>s, we performed a suite <strong>of</strong> simulations for σ d = 0, 0.25 <strong>and</strong> 0.5 at spacing intervals<br />

(s) <strong>of</strong> 2.5 cm, 25 cm <strong>and</strong> 2.5 m (Figure 4.21), starting with two 25-m-long b<strong>and</strong>s identical<br />

to that used in the calibration section above.<br />

In all cases the model results match our qualitative expectations based on theoretical<br />

considerations—the tips at first curve slightly away from each other <strong>and</strong> then curve more<br />

strongly toward each other, while the interaction is distinctly more pronounced for σ d = 0<br />

<strong>and</strong> s = 2.5 cm. The absolute magnitudes <strong>of</strong> path divergence from the established linear<br />

trend, however, can not be intuited, <strong>and</strong> the results reveal several interesting relationships.<br />

When the remote stress state is isotropic, interaction is quite pronounced, even at s = 2.5<br />

m for which the final overlapped spacing drops by half to 1.25 m before parallel<br />

<strong>propagation</strong> is restored. At s = 25 cm <strong>and</strong> 2.5 cm, the overlapped b<strong>and</strong> tips end up on<br />

intersecting trends, although the approach is distinctly more asymptotic for s = 25 cm.<br />

When σ d = 0.5, interaction is negligible except at s = 2.5 cm, when the final overlapped<br />

spacing drops by about one third to 1.6 cm. Even then, parallel <strong>propagation</strong> resumes<br />

before intersection occurs. Perhaps most interesting is that the results for σ d = 0.25 are<br />

all but indistinguishable from those <strong>of</strong> σ d = 0.5. Again, even at s = 2.5 cm, parallel<br />

<strong>propagation</strong> resumes prior to intersection. In fact, because the initial repulsive effect is<br />

more pronounced than for σ d = 0.5, the final overlapped spacing ends up the same at 1.6<br />

cm. In sum, these results suggest that strongly anastomosing b<strong>and</strong> patterns, particularly<br />

those involving very close encounters <strong>and</strong> the effective merging <strong>of</strong> b<strong>and</strong>s, can only occur<br />

114


when the ambient stress state is nearly isotropic <strong>and</strong>/or b<strong>and</strong>s are spaced less than a meter<br />

apart.<br />

One other factor mentioned in the section on mechanical theory as a potential<br />

influence on the <strong>propagation</strong> paths <strong>of</strong> interacting CBs is the effective drag or “inertia”<br />

imposed by their pre-existing length. As a test <strong>of</strong> the general validity <strong>of</strong> this notion, we<br />

undertook a series <strong>of</strong> three simulations identical in every respect except for the initial<br />

lengths <strong>of</strong> the model b<strong>and</strong>s: two 25-m b<strong>and</strong>s (as above <strong>and</strong> in Figure 4.21), one 25 m<br />

b<strong>and</strong> <strong>and</strong> one 5 m b<strong>and</strong>, <strong>and</strong> two 5 m b<strong>and</strong>s. In order to ensure adequate space <strong>and</strong><br />

<strong>propagation</strong> reactivity to render variations in path distinguishable on a single plot, a<br />

spacing <strong>of</strong> 2 m <strong>and</strong> an isotropic remote stress state were chosen (Figure 4.22). When both<br />

b<strong>and</strong>s are 25 m long, their mutual influence is substantially similar to that when spacing<br />

is 2.5 m. When one b<strong>and</strong> is 5 m long, its impact on the continued <strong>propagation</strong> <strong>of</strong> the<br />

longer b<strong>and</strong> is perceptibly diminished, both in terms <strong>of</strong> the initial repulsion <strong>and</strong> the<br />

subsequent attraction. The effect <strong>of</strong> the longer b<strong>and</strong> on the shorter one, however, is even<br />

more noticeably enhanced. When both b<strong>and</strong>s are 5 m long, their responses again mirror<br />

each other, but are completely different from either <strong>of</strong> the first two cases. In particular,<br />

they initially repel <strong>and</strong> attract each other somewhat less robustly than the pair <strong>of</strong> 25-m-<br />

long b<strong>and</strong>s, but, at the point when parallel <strong>propagation</strong> resumes in both earlier examples,<br />

attraction between the shorter b<strong>and</strong>s re-intensifies <strong>and</strong> they propagate toward intersection<br />

at a relatively high angle <strong>of</strong> approach (> 30°). The final path has a slight hitch in it that<br />

produces a mild, s-shaped curvature. This simple example demonstrates both that b<strong>and</strong><br />

length can influence the strength <strong>of</strong> tip interactions, <strong>and</strong> that the most general case<br />

involving different length b<strong>and</strong>s interacting is likely to produce noticeably asymmetric<br />

patterns <strong>of</strong> hooking.<br />

Finally, we note that, although the simulations presented in Figures 4.21 <strong>and</strong> 4.22<br />

generally predict the resumption <strong>of</strong> parallel <strong>propagation</strong> paths following initial overlap,<br />

the normalized magnitude <strong>of</strong> σθθ max rapidly drops toward one. This suggests that, for any<br />

reasonable critical threshold value, <strong>propagation</strong> would cease. Outcrop observations,<br />

however, demonstrate that long, sub-parallel overlaps between b<strong>and</strong>s spaced even<br />

millimeters apart are common.<br />

115


0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

−0.05<br />

−0.1<br />

−0.15<br />

−0.2<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

6<br />

4<br />

2<br />

0<br />

−2<br />

−4<br />

−0.2 −0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8<br />

−3 −2 −1 0 1 2 3 4 5 6<br />

−10 −5 0 5 10<br />

Figure 4.21a. Propagation simulations for two 25-m-long anticrack b<strong>and</strong>s approaching<br />

each other along parallel paths under isotropic remote stress conditions <strong>and</strong> initial spacing<br />

intervals <strong>of</strong> 2.5 cm (top), 25 cm (middle) <strong>and</strong> 2.5 m (bottom). Black lines indicate starting<br />

configurations, colored lines indicate simulated paths, all dimensions in meters.<br />

116<br />

15


0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

−0.05<br />

−0.1<br />

−0.15<br />

−0.2<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

6<br />

4<br />

2<br />

0<br />

−2<br />

−4<br />

−4<br />

−0.2 −0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8<br />

−3 −2 −1 0 1 2 3 4 5 6<br />

−10 −5 0 5 10<br />

Figure 4.21b. Propagation simulations for two 25-m-long anticrack b<strong>and</strong>s approaching<br />

each other along parallel paths (orthogonal to the maximum remote compressive stress) at<br />

a normalized differential stress <strong>of</strong> 0.25 <strong>and</strong> initial spacing intervals <strong>of</strong> 2.5 cm (top), 25 cm<br />

(middle) <strong>and</strong> 2.5 m (bottom). Black lines indicate starting configurations, colored lines<br />

indicate simulated paths, all dimensions in meters.<br />

117<br />

15


0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

−0.05<br />

−0.1<br />

−0.15<br />

−0.2<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

6<br />

4<br />

2<br />

0<br />

−2<br />

−4<br />

−0.2 −0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8<br />

−3 −2 −1 0 1 2 3 4 5 6<br />

−10 −5 0 5 10<br />

Figure 4.21c. Propagation simulations for two 25-m-long anticrack b<strong>and</strong>s approaching<br />

each other along parallel paths (orthogonal to the maximum remote compressive stress) at<br />

a normalized differential stress <strong>of</strong> 0.5 <strong>and</strong> initial spacing intervals <strong>of</strong> 2.5 cm (top), 25 cm<br />

(middle) <strong>and</strong> 2.5 m (bottom). Black lines indicate starting configurations, colored lines<br />

indicate simulated paths, all dimensions in meters.<br />

118<br />

15


4<br />

3<br />

2<br />

1<br />

0<br />

−1<br />

−2<br />

25 & 5m<br />

5 & 5m<br />

25 & 25m<br />

−4 −2 0 2 4 6 8<br />

Figure 4.22. Propagation simulations for two anticrack b<strong>and</strong>s approaching each other<br />

along parallel paths under isotropic remote stress conditions <strong>and</strong> initial spacing interval <strong>of</strong><br />

2 meters. Black lines indicate starting configurations, colored lines indicate simulated<br />

paths, <strong>and</strong> all dimensions are in meters. Three simulations are shown: two 25-m-long<br />

b<strong>and</strong>s (blue paths); two 5-m-long b<strong>and</strong>s (red paths); <strong>and</strong> a 25-m-long b<strong>and</strong> entering from<br />

the upper left <strong>and</strong> a 5-m-long b<strong>and</strong> entering from the lower right.<br />

7. Discussion <strong>and</strong> conclusions<br />

The fundamental postulate <strong>of</strong> this research is that all CBs in the Aztec—<strong>and</strong> by<br />

analogy s<strong>and</strong>stone in general—stem from the same fundamental mechanism <strong>of</strong> Griffith-<br />

flaw induced grain-scale collapse <strong>and</strong> subsequent anticrack-like <strong>propagation</strong> within a<br />

consistent remote stress field. This amounts to an assumption <strong>of</strong> mechanical simplicity.<br />

That being said, CBs clearly are not simple anticracks by any strict definition, rather they<br />

are ellipsoidal inclusions <strong>of</strong> compacted grains <strong>and</strong> porosity loss realized across a finite<br />

thickness. As such, they lack the distinct surfaces <strong>of</strong> displacement discontinuity <strong>and</strong><br />

discrete apertures characteristic <strong>of</strong> opening-mode cracks. However, ins<strong>of</strong>ar as CBs do<br />

also represent elliptical distributions <strong>of</strong> uniaxial volume change, <strong>and</strong> have very small<br />

thickness-to-length ratios, the stress <strong>and</strong> strain fields they induce within the surrounding,<br />

undamaged <strong>and</strong> nominally elastic s<strong>and</strong>stone can reasonably be modeled as resulting from<br />

virtual anticracks.<br />

The preliminary modeling results <strong>of</strong> <strong>propagation</strong> <strong>and</strong> interaction presented in this<br />

paper largely corroborate our theoretical expectations <strong>and</strong> are distinctly reminiscent <strong>of</strong><br />

119


the hooking patterns commonly observed in outcrop. We conclude that the anticrack<br />

model provides an adequate approximation to the outcrop-scale <strong>mechanics</strong> <strong>of</strong> compaction<br />

b<strong>and</strong>s <strong>and</strong> suggest specifically that:<br />

1. CBs propagate as a consequence <strong>of</strong> self-generated compressive stress<br />

concentrations along paths that attempt to maintain incremental symmetry<br />

with (orthogonality to) the maximum circumferential compressive stress<br />

immediately ahead <strong>of</strong> their tips;<br />

2. In the absence <strong>of</strong> external influences, CBs form symmetric with (orthogonal<br />

to) the maximum remote principal stress <strong>and</strong> will continue to propagate in<br />

plane;<br />

3. Small-scale directional perturbations <strong>of</strong> the tip <strong>of</strong> a well-developed b<strong>and</strong><br />

caused by material heterogeneities are self-correcting, regardless <strong>of</strong> the<br />

magnitude <strong>of</strong> the remote differential stress;<br />

4. Large-scale heterogeneity <strong>and</strong> anisotropy in the surrounding material aside,<br />

the primary cause <strong>of</strong> curving (out-<strong>of</strong>-plane) CB <strong>propagation</strong> is mechanical<br />

interaction between adjacent b<strong>and</strong>s, particularly between tips;<br />

5. The degree <strong>of</strong> these interactions is acutely sensitive to the magnitude <strong>of</strong> the<br />

remote differential stress—nearly isotropic stress conditions allow significant<br />

interaction, while realistically high values <strong>of</strong> differential stress virtually<br />

eliminate interaction;<br />

6. The degree <strong>of</strong> interaction is also sensitive to the material properties inside the<br />

propagating CBs, as well as their length (absolute in-plane dimensions)—<br />

lower internal resistance to shear enhances interaction as does shorter b<strong>and</strong><br />

lengths.<br />

Despite the general success <strong>of</strong> the current anticrack numerical model in producing<br />

realistic <strong>propagation</strong> behavior, considerable room for improvement exists. Among the<br />

physical phenomena that cannot as yet adequately be addressed are long overlaps<br />

between b<strong>and</strong>s spaced sometimes mere millimeters apart, the checkerboard patterns<br />

comprised <strong>of</strong> two distinct CB sets crossing each other at high angle, <strong>and</strong> the regular zig-<br />

zag behavior exhibited by both individual b<strong>and</strong>s <strong>and</strong> clusters <strong>of</strong> b<strong>and</strong>s. As the scope <strong>of</strong><br />

observation shrinks to focus on these issues at the grain scale, thereby challenging the<br />

120


assumption <strong>of</strong> an elastic continuum, distinct element method modeling techniques will<br />

become increasingly important (e.g. Antonellini <strong>and</strong> Pollard, 1995; Morgan, 1999;<br />

Morgan <strong>and</strong> Boettcher, 1999).<br />

Nonetheless, we believe that progress can be made toward underst<strong>and</strong>ing the outcrop-<br />

scale implications <strong>of</strong> grain-scale effects within the framework <strong>of</strong> the current model by<br />

implementing refined petrographic observations as more sophisticated schemes for<br />

applying boundary conditions on the b<strong>and</strong> elements <strong>and</strong> as more nuanced <strong>propagation</strong><br />

criteria. The key to these refined observations lies in detecting <strong>and</strong> describing the shape<br />

<strong>and</strong> nature <strong>of</strong> the damage zone around the uniformly compacted inclusion <strong>of</strong> grains<br />

currently recognized as a CB (Figure 4.23) or, alternatively, conclusively demonstrating<br />

the absence <strong>of</strong> any such damage envelope. In either case, the mechanical implications at<br />

both the grain <strong>and</strong> outcrop scales would be pr<strong>of</strong>ound.<br />

These various shortcomings aside, the current model simulation results do suggest<br />

<strong>and</strong> support several interpretations <strong>and</strong> inferences.<br />

1. The maximum normalized near-tip compressive stress concentration (σθθ at r<br />

= 5 mm) that can be realized for a realistic distribution <strong>of</strong> Dn is no more than<br />

about 2.5 (equivalent to 100 MPa). Even taking this value as a lower threshold<br />

because the data is derived from field measurements <strong>of</strong> CBs that stopped<br />

propagating, it appears that large stresses are not required to promote the<br />

quasi-static progression <strong>of</strong> compaction as accommodated by relatively<br />

coherent quartz grain plasticity in the Aztec s<strong>and</strong>stone.<br />

2. Judged in comparison to our simulation results, the variety <strong>of</strong> CB interactions<br />

observed in Aztec outcrop—ranging from gently anastomosing sub-parallel<br />

arrays <strong>of</strong> b<strong>and</strong>s spaced up to a meter apart to the near ubiquity <strong>of</strong> hooking tip<br />

<strong>and</strong> eye structures to strongly me<strong>and</strong>ering patterns <strong>of</strong> anastomosis—suggest<br />

that the prevailing remote paleo stress state was close to isotropic (σ22 r ≈<br />

0.9σ11 r , σ d ≈ 0.1).<br />

3. Although the remote state <strong>of</strong> sub-horizontal paleo stress in the Aztec was<br />

apparently nearly isotropic (σ d ≈ 0.1 in the x1-x2 plane), the state <strong>of</strong> stress in<br />

the x1-x3 plane, where σ33 r was essentially due to overburden, was decidedly<br />

anisotropic (σ d = 0.5 by definition in our model set up). Our simulation<br />

121


(a)<br />

(b)<br />

(c)<br />

σ1<br />

compaction b<strong>and</strong><br />

maximum thickness<br />

compaction b<strong>and</strong><br />

damage envelope<br />

<strong>propagation</strong> <strong>of</strong> damage zone<br />

<strong>propagation</strong> <strong>of</strong> compacted tip<br />

Figure 4.23. Schematic end-member conceptual models for the distribution <strong>of</strong><br />

compaction b<strong>and</strong> strain <strong>and</strong> damage. (a) B<strong>and</strong> pr<strong>of</strong>ile as currently recognized--elliptical<br />

inclusion <strong>of</strong> uniform uniaxial compactive strain = 0.1. (b) Hypothetical b<strong>and</strong> pr<strong>of</strong>ile as<br />

suggested by mechanical logic—same compacted b<strong>and</strong> as in (a), but surrounded by an<br />

envelope <strong>of</strong> grain damage <strong>and</strong> related s<strong>of</strong>tening generated at the tips. Uniform<br />

compaction within the damage zone accumulates from the inside out to form the b<strong>and</strong> as<br />

recognized, resulting in a maximum thickness once all the damaged grains have been<br />

accreted, <strong>and</strong> creating a layered inclusion <strong>of</strong> material properties that vary widely from<br />

each other <strong>and</strong> from the surrounding pristine s<strong>and</strong>stone. (c) Simultaneous <strong>propagation</strong> <strong>of</strong><br />

the compacted b<strong>and</strong> <strong>and</strong> the surrounding damage halo from the tip.<br />

122


esults therefore suggest that the generally anastomosing pattern <strong>of</strong> b<strong>and</strong>s<br />

captured in the x1-x2 plane would appear as a more consistently parallel, less<br />

interconnected pattern in the x1-x3 plane (Figure 4.24). This interpretation has<br />

yet to be confirmed by observing outcrop exposures with the requisite<br />

orientation, which are rare at best.<br />

4. Strongly parallel b<strong>and</strong> <strong>propagation</strong> in the x1-x3 plane would tend also to retard<br />

CB interactions in the x1-x2 plane. Given the degree <strong>of</strong> anastomosis<br />

nonetheless present in Aztec outcrops, this suggest that the paleo state <strong>of</strong><br />

stress in the x1-x2 plane was likely more nearly isotropic than our 2-D<br />

simulations suggest.<br />

5. These interrelationships <strong>of</strong> magnitude <strong>and</strong> direction between the components<br />

<strong>of</strong> principal paleo stress <strong>and</strong> the resulting patterns <strong>of</strong> b<strong>and</strong> <strong>propagation</strong> have<br />

significant implications for predicting 3-D permeability in aquifers <strong>and</strong><br />

reservoirs containing CBs. Specifically, because permeability along the trend<br />

<strong>of</strong> a b<strong>and</strong> pattern decreases with increasing connectivity, <strong>and</strong> the minimum<br />

principal permeability (k3 ≤ k2 ≤ k1) is directed orthogonal to the dominant<br />

b<strong>and</strong> orientation (Sternl<strong>of</strong> et al., 2004), k1 due to a subsurface b<strong>and</strong> array will<br />

parallel the presumed direction <strong>of</strong> paleo σ33 r , k2 will parallel σ22 r <strong>and</strong> k3 will<br />

parallel σ11 r (Figure 4.24). The relative magnitudes <strong>of</strong> the principal<br />

permeabilities can likewise be inferred from the presumed relative magnitudes<br />

<strong>of</strong> the paleo principal stresses.<br />

6. By the same token, the orientations <strong>and</strong> relative magnitudes <strong>of</strong> all three<br />

principal paleo stresses can be inferred from directional variations observed in<br />

the degree <strong>of</strong> anastomosis revealed by a well-exposed (or imaged) CB array.<br />

8. Acknowledgements<br />

We are deeply indebted to John Childs for his ever cheerful <strong>and</strong> able assistance in the<br />

field, <strong>and</strong> thank Ovunc Mutlu for help with the numerical code. This work was funded by<br />

the U.S. Department <strong>of</strong> Energy, Office <strong>of</strong> Basic Energy Science, Geosciences Research<br />

Program under grant DE-FG03-94ER14462, awarded to David Pollard <strong>and</strong> Atilla Aydin<br />

at <strong>Stanford</strong> University. Additional support was provided by the <strong>Stanford</strong> Rock Fracture<br />

Project.<br />

123


σ 3<br />

σ 2<br />

σ 1<br />

Figure 4.24. Schematic 3-D relationship between the dominant trend <strong>of</strong> a compaction<br />

b<strong>and</strong> array, <strong>and</strong> its connectivity along that trend, to the orientations <strong>and</strong> relative<br />

magnitudes <strong>of</strong> both the principal remote stresses (σ1 > σ2 > σ3) in which it formed <strong>and</strong> <strong>of</strong><br />

the corresponding principal permeabilities that result (k1 > k2 >k3).<br />

124<br />

k 3<br />

k 1<br />

k 2


1. Abstract<br />

Chapter 5<br />

Computational estimation <strong>of</strong> compaction b<strong>and</strong> permeability:<br />

From thin-section estimations to reservoir implications<br />

Permeability measurements can be difficult to obtain when sample availability is<br />

restricted, dimensions are limited, or materials poorly consolidated. With subsurface<br />

cores <strong>of</strong> s<strong>and</strong>stone containing thin, tabular compaction b<strong>and</strong>s, all three challenges could<br />

arise. Methods for estimating permeability from thin-section provide an alternative. We<br />

evaluate a new physics-based computational technique, in which Lattice-Boltzmann flow<br />

simulations are conducted on stochastic realizations <strong>of</strong> 3-D pore structure generated from<br />

digital thin-section images. Applied to a representative thin section from the Aztec<br />

s<strong>and</strong>stone <strong>of</strong> southeastern Nevada, an exhumed analog for b<strong>and</strong>-rich s<strong>and</strong>stone aquifers<br />

<strong>and</strong> reservoirs, the method yields estimates that agree well with available data—a few<br />

millidarcys (b<strong>and</strong>) to a few Darcys (s<strong>and</strong>stone)—capturing the range <strong>of</strong> both matrix <strong>and</strong><br />

compaction-b<strong>and</strong> permeability from a single thin section. Extracted from a subsurface<br />

equivalent <strong>of</strong> the Aztec, such data could prove invaluable, as pervasive arrays <strong>of</strong><br />

compaction b<strong>and</strong>s in s<strong>and</strong>stone have been shown capable <strong>of</strong> exerting substantial fluid-<br />

flow effects at scales relevant to aquifer <strong>and</strong> reservoir management.<br />

2. Introduction<br />

In porous, granular rocks such as the Aztec s<strong>and</strong>stone in the Valley <strong>of</strong> Fire State Park<br />

<strong>of</strong> southeastern Nevada (Figure 5.1), compaction b<strong>and</strong>s (CBs) crop out as thin, tabular<br />

features <strong>of</strong> porosity-loss compaction accommodated by grain damage, rearrangement <strong>and</strong><br />

preferential clay accumulation (Sternl<strong>of</strong> et al. 2005). Generally up to a few centimeters in<br />

thickness <strong>and</strong> tens <strong>of</strong> meters in planar extent, they represent the kinematic subset <strong>of</strong><br />

deformation b<strong>and</strong>s dominated by closing-mode displacement <strong>and</strong> oriented perpendicular<br />

to the local direction <strong>of</strong> maximum compression (Mollema <strong>and</strong> Antonellini 1996; Du<br />

Bernard et al. 2002; Borja <strong>and</strong> Aydin 2004; Sternl<strong>of</strong> et al. 2005). Reduced porosity, pore<br />

connectivity <strong>and</strong> average pore-throat diameter conspire to decrease deformation-b<strong>and</strong><br />

permeability by one to four orders <strong>of</strong> magnitude relative to the host rock matrix (Pittman<br />

1981; Freeman 1990; Antonellini <strong>and</strong> Aydin 1994; Crawford 1998; Gibson 1998; Taylor<br />

125


NV UT<br />

CA<br />

AZ<br />

Park Road<br />

Map<br />

Detail<br />

Park Boundary<br />

*<br />

Route 169<br />

0 5<br />

km<br />

NEVADA<br />

ARIZONA<br />

10 km<br />

Valley<br />

<strong>of</strong> Fire<br />

State Park<br />

N<br />

LAKE<br />

MEAD<br />

115 o 30'<br />

Figure 5.1. Schematic location map for the Valley <strong>of</strong> Fire State Park in southeastern<br />

Nevada. Colorful alteration patterns <strong>of</strong> iron-oxide staining in the Aztec s<strong>and</strong>stone lend<br />

the park its name (* marks approximate location <strong>of</strong> photographs shown in Figures 5.2 <strong>and</strong><br />

5.3).<br />

126<br />

36 o 15'


<strong>and</strong> Pollard 2000; Lothe et al. 2002), rendering them impediments to fluid flow under<br />

saturated conditions (Figure 5.2).<br />

Sternl<strong>of</strong> et al. (2004) have shown that anastomosing arrays <strong>of</strong> subparallel, subvertical<br />

CBs abundantly exposed in the Aztec s<strong>and</strong>stone (Figure 5.3) would significantly reduce<br />

bulk permeability <strong>and</strong> induce permeability anisotropy at outcrop scales. Part <strong>of</strong> a<br />

widespread Jurassic æolian system that includes the Navajo <strong>and</strong> Nugget s<strong>and</strong>stones<br />

(Marzolf 1983; Blakey 1989), the 1,400-m-thick Aztec is a medium-grained sub-arkose<br />

that is typified by large-scale cross stratification (Bohannon 1983) <strong>and</strong> comprises an<br />

exhumed analog for aquifers <strong>and</strong> reservoirs in similar lithologies. The pervasive CB<br />

arrays, which cut across depositional bedding as the oldest <strong>structural</strong> fabric present,<br />

formed in response to compression associated with the Cretaceous Sevier Orogeny (Hill<br />

1989; Eichhubl et al. 2004; Sternl<strong>of</strong> et al. 2005). If present in an active aquifer or<br />

reservoir, these b<strong>and</strong> arrays would cause substantial fluid-flow effects at scales relevant<br />

to production <strong>and</strong> management (Sternl<strong>of</strong> et al. in review).<br />

The specific hydraulic impacts exerted by CBs depend strongly on their porosity <strong>and</strong><br />

permeability relative to the surrounding s<strong>and</strong>stone matrix, as well as on geometrical<br />

factors such as average b<strong>and</strong> thickness, spacing <strong>and</strong> connectivity (Sternl<strong>of</strong> et al. 2004).<br />

Their limited thickness, however, renders direct measurement <strong>of</strong> CB permeability<br />

difficult <strong>and</strong> potentially inaccurate under even optimal conditions (Antonellini <strong>and</strong> Aydin<br />

1994). Given the generally limited availability <strong>of</strong> potentially unconsolidated core samples<br />

from active aquifers <strong>and</strong> reservoirs, problems proliferate. Methods for estimating<br />

permeability from epoxy-impregnated thin-sections using computational techniques<br />

provide an alternative (e.g. Berryman <strong>and</strong> Blair 1987; Adler et al. 1990; Bakke <strong>and</strong> Øren<br />

1997; Blair et al. 1996).<br />

In this paper, we evaluate a new physics-based algorithm for estimating permeability<br />

from thin section images (Keehm et al. 2004). In order to simulate the realistic situation<br />

<strong>of</strong> limited subsurface sample supply, we apply the method to a single, representative thin<br />

section from the Aztec (Figure 5.4) <strong>and</strong> compare the resulting estimates <strong>of</strong> CB <strong>and</strong> matrix<br />

permeability to available measurements reported for the Navajo <strong>and</strong> Aztec s<strong>and</strong>stones<br />

(Antonellini <strong>and</strong> Aydin 1994; Flodin et al. 2005).<br />

127


compaction b<strong>and</strong><br />

Figure 5.2. Typical compaction b<strong>and</strong> exposed as a tabular, cm-thick fin at high angle to<br />

depositional bedding in the Aztec s<strong>and</strong>stone. Sharply decreased internal porosity <strong>and</strong><br />

permeability due to grain crushing, rearrangement <strong>and</strong> preferential clay accumulation<br />

render the b<strong>and</strong> an impediment to fluid flow, as illustrated by its obvious impact on the<br />

distribution <strong>of</strong> iron-oxide staining.<br />

128


Figure 5.3. Extensive arrays <strong>of</strong> dominantly north-northwest trending, steeply eastdipping<br />

compaction b<strong>and</strong>s pervade the upper 600 m <strong>of</strong> the Aztec s<strong>and</strong>stone, creating a<br />

sub-parallel, anastomosing system <strong>of</strong> low-permeability baffles capable <strong>of</strong> affecting fluid<br />

flow over hundreds <strong>of</strong> meters. View is to the northwest, b<strong>and</strong>s crop out as resistant fins.<br />

129


3. Computational method<br />

The methodology workflow involves creating a 2-D binary map <strong>of</strong> porosity from a<br />

digital thin-section image, transforming this into multiple, equally probable stochastic<br />

realizations <strong>of</strong> the full 3-D pore structure, <strong>and</strong> finally applying a Lattice-Boltzman<br />

method flow simulation to calculate an effective permeability for each realization (Figure<br />

5.5).<br />

3.1. Image processing<br />

The first step in estimating permeability from thin-section is to convert representative<br />

digital images <strong>of</strong> the material into binary (pure black <strong>and</strong> white) images (Figures 5.5a, b).<br />

These binary conversions are used for porosity estimation, variogram modeling <strong>and</strong> as<br />

conditional data for the stochastic 3-D pore-structure realizations. Two types <strong>of</strong> initial<br />

digital images are common: true-color RGB (red-green-blue) images captured on<br />

st<strong>and</strong>ard petrographic microscopes using transmitted light; <strong>and</strong> grayscale (intensity)<br />

images captured with scanning electron microscopy (SEM). In both cases, the goal is to<br />

produce accurate binary maps representing the spatial distribution <strong>of</strong> pores <strong>and</strong> grains.<br />

Impregnation with colored epoxy (generally blue or red) prior to sectioning renders<br />

porosity distinct in RGB images. The original method <strong>of</strong> Keehm et al. (2004) creates a<br />

suite <strong>of</strong> index colors from which an operator selects those corresponding to porosity.<br />

With this approach, a degree <strong>of</strong> uncertainty in setting color thresholds to differentiate<br />

grains from pores is unavoidable. For grayscale images, there is only the intensity plane,<br />

with low intensity (black) corresponding to porosity. This reduces the binary conversion<br />

to a simple choice <strong>of</strong> threshold value, generally on a scale <strong>of</strong> 1 (pure black) to 256 (pure<br />

white). SEM images, particularly those collected in backscattered electron composition<br />

(BEC) mode, therefore provide generally sharper resolution than RGB images, but do<br />

require specialized equipment to obtain.<br />

3.2. Pore-structure realization<br />

Various methods for constructing 3-D pore-structure realizations from 2-D binary<br />

images have been developed, including those <strong>of</strong> Adler et al. (1990), Yeong <strong>and</strong> Torquato<br />

(1998) <strong>and</strong> Øren <strong>and</strong> Bakke (2002). A brief comparison <strong>of</strong> these is presented by Keehm<br />

et al. (2004), who <strong>of</strong>fer a geostatistical approach based on the sequential indicator<br />

simulation (SIS) method <strong>of</strong> Deutsch <strong>and</strong> Journel (1998). The SIS method produces<br />

130


stochastic 3-D realizations based on the two moments <strong>of</strong> a binary image—its porosity <strong>and</strong><br />

a two-point correlation function (variogram). A r<strong>and</strong>om path is followed through the<br />

nodes <strong>of</strong> a model 3-D grid. At each node, a local conditional cumulative distribution<br />

function (ccdf) is estimated by indicator kriging. Each successive ccdf is conditioned to<br />

the binary image <strong>and</strong> to all previous node selections, <strong>and</strong> a r<strong>and</strong>om selection from the<br />

function gives the value—pore or grain—for the current node. A 3-D (cubic) binary field<br />

with the same pore-grain spatial characteristics as the original 2-D image results (Figure<br />

5.5c). Comparing multiple realizations generated from a single image provides a means<br />

<strong>of</strong> assessing pore-structure variability.<br />

3.3. Permeability estimation<br />

Permeability is estimated by conducting numerical flow simulations on the 3-D pore-<br />

structure realizations, which tend to be geometrically complex <strong>and</strong> may contain statistical<br />

noise due to the stochastic nature <strong>of</strong> their construction (Keehm et al. 2004). The Lattice-<br />

Boltzmann (LB) method is a robust technique that simulates flow according to simple<br />

rules governing local interactions between individual particles (Doolen 1990) <strong>and</strong><br />

recovers the Navier-Stokes equations at the macroscopic scale (Ladd 1994). Prime<br />

advantages <strong>of</strong> the LB method are its ability to h<strong>and</strong>le any discrete geometry without<br />

simplification (Cancelliere et al. 1998), <strong>and</strong> its accuracy in describing flow through<br />

porous media (Ladd 1994; Bosl et al. 1998; Keehm et al. 2001). Artifacts can occur in the<br />

local velocity fields (Manwart et al. 2002), but applying time-averaged velocities<br />

mitigates these effects (Ladd 1994). The LB simulations are accomplished by assigning a<br />

pressure gradient (∇P) across opposite faces <strong>of</strong> the cubic pore-structure realization. From<br />

the local flux (Figure 5.5d), a volume-averaged flux 〈q〉 can be calculated. The<br />

macroscopic or bulk permeability (κ) is then calculated according to Darcy’s Law:<br />

κ<br />

〈 q〉 = − ∇P<br />

η<br />

where η is the dynamic viscosity <strong>of</strong> the fluid. As a statistical measure <strong>of</strong> κ variability,<br />

LB simulations can be conducted on multiple 3-D pore-structure realizations generated<br />

from the same image. The final permeability estimate is then the average <strong>of</strong> the values<br />

from all the simulations.<br />

131<br />

(1)


compaction b<strong>and</strong><br />

A<br />

5 mm<br />

A‘<br />

compaction b<strong>and</strong><br />

compaction b<strong>and</strong><br />

depositional bedding<br />

Figure 5.4. Representative thin section <strong>of</strong> a tabular compaction b<strong>and</strong> in the Aztec<br />

s<strong>and</strong>stone. White grains are quartz, brown grains are stained feldspar, black grains are<br />

iron oxide, blue is epoxy-filled porosity. Photo inset (upper left) shows the sample core<br />

from which the thin section was made. Transects A-A’ <strong>and</strong> B-B’ (normal to bedding)<br />

pertain to the calculations shown in Figure 5.8.<br />

132<br />

B<br />

B‘


0.5 mm<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Figure 5.5. Permeability estimation methodology workflow: A) raw backscattered<br />

electron composition image; B) binary conversion map <strong>of</strong> porosity (black) versus mineral<br />

constituents (white); C) stochastic 3-D pore-structure realization; D) parallel slices<br />

through the pore-structure realization showing local mass flux predicted by Lattice-<br />

Boltzmann flow simulation (values normalized by maximum).<br />

133


4. Application to the Aztec s<strong>and</strong>stone<br />

In order to test this computational estimation method under real-world conditions <strong>of</strong><br />

limited sample supply, we applied the algorithm to a single, representative thin section <strong>of</strong><br />

Aztec s<strong>and</strong>stone showing a cm-thick CB surrounded by relatively undeformed matrix<br />

(Figure 5.4). As is typical with foreset deposition in æolian s<strong>and</strong>stones, bedding consists<br />

<strong>of</strong> alternating, well-sorted coarse-grain (~0.4 mm) <strong>and</strong> fine-grain (~0.1 mm) layers up to<br />

several millimeters thick. Poorly sorted mixed-grain layers are also common. Within each<br />

<strong>of</strong> the three distinct bed types, the distribution <strong>of</strong> grains <strong>and</strong> pores is relatively<br />

homogeneous, both in the matrix <strong>and</strong> in the CB.<br />

Digital images were collected in BEC mode on a st<strong>and</strong>ard SEM at 15 kV <strong>and</strong> a<br />

magnification <strong>of</strong> 80x. Two images were collected from each bed type, with mosaic<br />

composites used as necessary to ensure representative coverage in coarse-grain beds. A<br />

consistent intensity threshold <strong>of</strong> 72 out <strong>of</strong> 256 was used for the binary conversions, <strong>and</strong><br />

ten 3-D pore-structure realizations were computed for each <strong>of</strong> the 12 binary images. Flow<br />

simulations were then conducted on each pore-structure realization to yield 120 total<br />

estimates <strong>of</strong> permeability relative to porosity. Figure 5.6 shows an example BEC image<br />

with its binary conversion for each bed type from both the matrix <strong>and</strong> the CB.<br />

4.1. Simulation results<br />

The resulting estimates <strong>of</strong> porosity <strong>and</strong> permeability are plotted in Figure 5.7; the<br />

averaged permeability results are summarized in Table 5.1.<br />

Porosity <strong>and</strong> permeability estimates for the matrix range from about 20-27% <strong>and</strong> 200-<br />

3,500 millidarcys (mD). Not surprisingly, average porosity for the pore-structure<br />

realizations derived from the mixed-grain bed images is lowest (22%), while average<br />

permeability estimated for the coarse-grain pore-structure realizations is highest (1,392<br />

mD). The fine-grain beds, which exhibit the tightest grain-size distribution, produced the<br />

highest average porosity estimate (25.25%) <strong>and</strong> the lowest average permeability estimate<br />

(406 mD). The combined average estimates for the matrix are 23.6% <strong>and</strong> 776 mD. This<br />

simple average bulk permeability value, however, does not account for the inherently<br />

anisotropic, layered nature <strong>of</strong> the s<strong>and</strong>stone. Figure 5.8 presents a more realistic approach<br />

to estimating directional permeability in the matrix, yielding best-estimate values <strong>of</strong><br />

1,053 mD parallel to bedding <strong>and</strong> 804 mD normal to bedding.<br />

134


S<strong>and</strong>stone matrix<br />

Compaction b<strong>and</strong><br />

Coarse-grain Fine-grain Mixed-grain<br />

Figure 5.6. Backscattered electron composition (BEC) images <strong>of</strong> the three depositional<br />

bedding types in the Aztec s<strong>and</strong>stone (coarse-grain, fine-grain <strong>and</strong> mixed-grain) shown<br />

above their corresponding binary conversions for both the s<strong>and</strong>stone matrix (top set) <strong>and</strong><br />

compaction b<strong>and</strong> (bottom set). In the BEC images black is porosity, dark gray is<br />

kaolinite, medium gray is detrital quartz <strong>and</strong> light gray is detrital feldspar.<br />

135<br />

BEC Binary BEC Binary


Permeability (mD)<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

Coarse-grain<br />

Fine-grain<br />

Mixed-grain<br />

Mean coarse<br />

Mean fine<br />

Mean mixed<br />

Mean CB<br />

(48)<br />

(33)<br />

Porosity (fraction)<br />

Matrix<br />

Compaction b<strong>and</strong><br />

10<br />

0 0.05 0.1 0.15 0.2 0.25 0.3<br />

0<br />

Figure 5.7. Semi-log scatter plot <strong>of</strong> permeability versus porosity for the computational<br />

estimates <strong>of</strong> matrix <strong>and</strong> compaction b<strong>and</strong> derived from the thin section shown in Figure<br />

5.4. The adjacent box plots represent the corresponding ranges <strong>and</strong> distributions <strong>of</strong><br />

available permeability measurements for the Aztec <strong>and</strong> Navajo s<strong>and</strong>stones, with the total<br />

number <strong>of</strong> data points in parentheses.<br />

Table 5.1. Summary <strong>of</strong> permeability estimates<br />

S<strong>and</strong>stone matrix Compaction b<strong>and</strong><br />

Coarse Fine Mixed Coarse Fine Mixed<br />

Mean 1,392 mD 406 mD 530 mD 5.22 mD 4.29 mD 5.77 mD<br />

Min 879 mD 253 mD 258 mD 2.04 mD 1.36 mD 1.20 mD<br />

Max 3,459 mD 655 mD 1,574 mD 10.67 mD 9.12 mD 19.73 mD<br />

136


Porosity <strong>and</strong> permeability estimates for the CB range from about 9-12% <strong>and</strong> 1-20<br />

mD, with the results from the three bed types substantially overlapping (Figure 5.7). This<br />

suggests that permeability within the b<strong>and</strong> is approximately homogeneous <strong>and</strong> isotropic.<br />

The combined average estimates for CB porosity <strong>and</strong> permeability are 10.6% <strong>and</strong> 5.1<br />

mD. It turned out that exactly half <strong>of</strong> the 3-D pore-structure realizations for the CB—12<br />

for the coarse-grain beds <strong>and</strong> nine each for the fine-grain <strong>and</strong> mixed-grain beds—yielded<br />

no connected flow paths <strong>and</strong> so returned zero permeability. Given that the minimum non-<br />

zero permeability result was 1.2 mD, we deemed all zero values to be unrepresentative<br />

<strong>and</strong> discarded them (see Conclusions). Our average estimate for CB permeability <strong>of</strong> 5.1<br />

mD is therefore a maximum, <strong>and</strong> we suggest that the true value is probably around 3 mD,<br />

given that the average including zeros is 2.5 mD.<br />

4.2. Comparison to measured values<br />

Our computational approach yields permeability estimates for the Aztec s<strong>and</strong>stone <strong>of</strong><br />

about 1,000 mD along bedding, 800 mD across bedding <strong>and</strong> 5 mD within CBs, which<br />

generally cut across bedding at high angle. Permeability reduction due to compaction<br />

within the b<strong>and</strong>s is therefore estimated to be at least two orders <strong>of</strong> magnitude—the<br />

middle <strong>of</strong> the range reported for deformation b<strong>and</strong>s in general (see Introduction).<br />

Antonellini <strong>and</strong> Aydin (1994) report minipermeameter measurements for both matrix <strong>and</strong><br />

compactive deformation b<strong>and</strong>s from the Navajo s<strong>and</strong>stone <strong>of</strong> southeastern Utah,<br />

recognized as a depositional <strong>and</strong> chronological equivalent <strong>of</strong> the Aztec (Marzolf 1983;<br />

Blakey 1989). Their results do not distinguish between bedding parallel <strong>and</strong> bedding<br />

normal permeability. Flodin et al. (2005) provide a h<strong>and</strong>ful <strong>of</strong> additional lab<br />

measurements for the Aztec matrix, also without reference to bedding orientation.<br />

Together, these measurements—ranging from 123 mD to 5,991 mD in the matrix (mean<br />

<strong>of</strong> 1,768 mD, <strong>and</strong> from 1.35 mD to 38.33 mD in the b<strong>and</strong>s (mean <strong>of</strong> 14.14 mD)—<br />

correspond well with our computational estimates, while being somewhat higher (Figure<br />

5.7). We ascribe this discrepancy to spatial permeability variation within the s<strong>and</strong>stone,<br />

as much as to any systematic inaccuracy in the computational method, noting again that<br />

the estimates derive from a single sample location.<br />

137


B‘<br />

A‘<br />

A<br />

c<br />

f<br />

m<br />

c<br />

m<br />

c<br />

f<br />

c<br />

f<br />

c<br />

f<br />

c<br />

f<br />

4.8 mm<br />

0.9 mm<br />

1.8 mm<br />

3.3 mm<br />

B<br />

4.5 mm<br />

0.9 mm<br />

0.6 mm<br />

1.5 mm<br />

0.9 mm<br />

1.2 mm<br />

1.5 mm<br />

7.2 mm<br />

0.9 mm<br />

K n<br />

L = 3 cm<br />

=<br />

K n<br />

=<br />

K p<br />

k f k m<br />

K p<br />

kc lc + kf lf + km L<br />

l c<br />

k c k f k m L<br />

+ kc km lf +<br />

l m<br />

k c k f<br />

l m<br />

Bulk Permeability Estimates<br />

Kp Kn Mean<br />

Minimum<br />

Maximum<br />

1,053 mD 804 mD<br />

655 mD 482 mD<br />

2,614 mD 1,785 mD<br />

Figure 5.8. Computation <strong>of</strong> bulk permeability parallel (Kp) <strong>and</strong> normal (Kn) to<br />

depositional bedding for the composite transect A-A’—B-B’ shown in Figure 5.4. Kp <strong>and</strong><br />

Kn are calculated as weighted arithmetic <strong>and</strong> harmonic averages, respectively (Deutsch,<br />

1989). Coarse-grain permeability <strong>and</strong> the total thickness <strong>of</strong> coarse-grain beds along the<br />

transect are connoted by kc <strong>and</strong> lc, respectively. Similarly, kf, lf, km, <strong>and</strong> lm represent these<br />

values for fine <strong>and</strong> mixed-grain beds. The mean, minimum <strong>and</strong> maximum bulk<br />

permeability estimates are computed using the corresponding mean, minimum <strong>and</strong><br />

maximum values <strong>of</strong> kc, kf <strong>and</strong> km (Figure 5.7).<br />

138


5. Conclusions<br />

The method applied in this paper comprises a convenient <strong>and</strong> robust computational<br />

technique for estimating permeability from thin-section images. It is particularly well<br />

suited to small structures <strong>and</strong> samples that do not lend themselves to direct measurement,<br />

<strong>and</strong> to situations where sample supplies are limited or unconsolidated materials requiring<br />

impregnation cannot otherwise be analyzed. We have demonstrated that both matrix <strong>and</strong><br />

compaction-b<strong>and</strong> permeability estimates derived from a single thin section <strong>of</strong> Aztec<br />

s<strong>and</strong>stone reliably reflect the range <strong>of</strong> measured values reported for comparable samples.<br />

Furthermore, we were able to isolate variations in permeability due to grain-size <strong>and</strong><br />

sorting differences between the characteristic depositional layers—well-sorted coarse,<br />

well-sorted fine <strong>and</strong> poorly sorted mixed—allowing us to assess the inherent permeability<br />

anisotropy due to bedding. Application <strong>of</strong> the method to low-porosity CBs, however,<br />

does highlight a resolution limitation warranting further comment.<br />

The stochastic 3-D realization algorithm produces pore structures lacking through-<br />

going connectivity with increasing frequency as porosity drops below about 12%. Half <strong>of</strong><br />

our CB pore-structure realizations, with a mean porosity <strong>of</strong> 10.6%, yielded zero<br />

permeability. Sensitivity testing reveals that below about 7% porosity, the probability <strong>of</strong><br />

achieving a nonzero permeability estimate becomes negligible. The reality for clastic<br />

materials is that permeability does not begin to disappear until a minimum porosity<br />

threshold <strong>of</strong> about 2-4% (Mavko <strong>and</strong> Nur 1997). This limitation <strong>of</strong> the method could be<br />

mitigated in a variety <strong>of</strong> ways, such as conducting pore-structure realizations until the<br />

desired number <strong>of</strong> non-zero effective permeability results is achieved; increasing<br />

resolution <strong>of</strong> the 3-D grid (i.e. decreasing node spacing), which however would also<br />

greatly increase computation time; or using a different 3-D pore-structure construction<br />

technique, such as multi-point realization.<br />

Nonetheless, we conclude that representative permeability estimations for porous,<br />

granular materials—even from a single bedding layer or compaction b<strong>and</strong> in s<strong>and</strong>stone—<br />

can be derived from thin-section images using the method demonstrated in this paper. In<br />

fact, the permeability data gleaned in our test case from a single thin section would be <strong>of</strong><br />

great value in anticipating the potential impact <strong>of</strong> CBs on fluid flow in a subsurface<br />

139


equivalent <strong>of</strong> the Aztec s<strong>and</strong>stone, with important implications for aquifer <strong>and</strong> reservoir<br />

management.<br />

6. Acknowledgements<br />

The authors wish to thank David Pollard <strong>and</strong> Gary Mavko for their comments,<br />

insights <strong>and</strong> support; John Childs for able assistance in the field; <strong>and</strong> Robert Jones for<br />

guidance on the SEM. Primary funding for this work was provided by the U.S.<br />

Department <strong>of</strong> Energy, Office <strong>of</strong> Basic Energy Sciences, Geosciences Research Program<br />

under grants DE-FG03-94ER14462 (David Pollard <strong>and</strong> Atilla Aydin) <strong>and</strong> DE-FG02-<br />

03ER15423 (Amos Nur <strong>and</strong> Gary Mavko). Additional support came from the <strong>Stanford</strong><br />

Rock Fracture Project (RFP) <strong>and</strong> the <strong>Stanford</strong> Rock Physics <strong>and</strong> Borehole Geophysics<br />

Project (SRB).<br />

140


1. Abstract<br />

Chapter 6<br />

Permeability effects <strong>of</strong> deformation b<strong>and</strong> arrays in s<strong>and</strong>stone<br />

We apply established numerical modeling techniques to the calculation <strong>of</strong> effective<br />

permeability for porous s<strong>and</strong>stone containing systematic arrays <strong>of</strong> low permeability<br />

deformation b<strong>and</strong>s (DBs). The numerical method, derived from homogenization theory,<br />

can produce effective permeability results for characteristic DB geometries, <strong>and</strong> provides<br />

a quantitative framework with which to exp<strong>and</strong> these effects to the reservoir-simulation<br />

scale from spatially limited data. The method is demonstrated in two dimensions for each<br />

<strong>of</strong> three characteristic DB patterns—parallel, cross-hatch <strong>and</strong> anastomosing—exposed in<br />

the Aztec s<strong>and</strong>stone at the Valley <strong>of</strong> Fire, Nevada, which provides an excellent exhumed<br />

analog for active s<strong>and</strong>stone reservoirs. Our analysis indicates that these extensive,<br />

systematic DB patterns can reduce effective permeability by as much as two orders <strong>of</strong><br />

magnitude at scales relevant to reservoir production, while inducing similar magnitudes<br />

<strong>of</strong> permeability anisotropy. The potential permeability impacts <strong>of</strong> DB arrays on reservoir<br />

production thus rival those routinely attributed to depositional heterogeneity—e.g.<br />

bedding <strong>and</strong> shale streaks. We suggest therefore that properly accounting for the<br />

aggregate effects <strong>of</strong> DBs where they occur could be as important to optimal reservoir<br />

simulation <strong>and</strong> production management in s<strong>and</strong>stone as accounting for sedimentary<br />

architecture.<br />

2. Introduction<br />

Sedimentary <strong>and</strong> <strong>structural</strong> heterogeneities both large (e.g. shale lenses, seismically<br />

detectable faults) <strong>and</strong> small (e.g. bedding, fractures) have been shown to pr<strong>of</strong>oundly<br />

influence effective permeability in s<strong>and</strong>stone. Small features, however, cannot explicitly<br />

be modeled in st<strong>and</strong>ard flow simulations using coarse (10 to 100 m) grid blocks. For<br />

porous, granular materials like s<strong>and</strong>stone, the simplifying assumption <strong>of</strong> isotropic<br />

background permeability is commonly assumed, with perhaps an anisotropy related to<br />

bedding. To more realistically account for the aggregate influence <strong>of</strong> small-scale features,<br />

an effective permeability must be calculated <strong>and</strong> assigned to each simulation block. This<br />

141


effective permeability represents an average bulk or “upscaled” permeability for the host<br />

rock <strong>and</strong> small-scale heterogeneities taken together.<br />

A variety <strong>of</strong> numerical techniques have been developed to compute effective<br />

permeability for use in coarse scale modeling. Extensive discussion <strong>of</strong> these can be found<br />

in the reviews <strong>of</strong> Wen <strong>and</strong> Gomez-Hern<strong>and</strong>ez (1996) <strong>and</strong> Renard <strong>and</strong> de Marsily (1997).<br />

Many <strong>of</strong> the techniques also have been used to quantify the effects <strong>of</strong> fine scale features<br />

on large scale permeability. For example, several researchers have computed effective<br />

permeability values for finely interbedded s<strong>and</strong>-shale sequences, finding that shale streak<br />

volume fraction (Vsh ) is the key controlling parameter. Desbarats (1987) modeled<br />

s<strong>and</strong>/shale sequences composed <strong>of</strong> homogeneous, isotropic s<strong>and</strong>stone (permeability=kss)<br />

<strong>and</strong> homogeneous, isotropic shale lenses (permeability=ksh). For ksh/kss=10 -4 , he found the<br />

effective vertical permeability to be an order <strong>of</strong> magnitude less than kss for Vsh ≈30%, <strong>and</strong><br />

2 orders <strong>of</strong> magnitude less for Vsh ≈65%. He found the effective horizontal permeability<br />

to be reduced by an order <strong>of</strong> magnitude for Vsh ≈70%. Deutsch (1989) determined that for<br />

a relative shale/s<strong>and</strong>stone permeability <strong>of</strong> ksh/kss=10 -5 , effective isotropic permeability<br />

drops by an order <strong>of</strong> magnitude for Vsh ≈50% <strong>and</strong> 2 orders <strong>of</strong> magnitude for Vsh ≈65%.<br />

These results suggest that for Vsh greater than 65%, effective horizontal <strong>and</strong> vertical<br />

permeability drop by about 1 <strong>and</strong> 2 orders <strong>of</strong> magnitude, respectively, thus inducing an<br />

order-<strong>of</strong>-magnitude permeability anisotropy.<br />

Previous studies also have demonstrated that other types <strong>of</strong> small scale sedimentary<br />

<strong>and</strong> <strong>structural</strong> heterogeneities can exert substantial permeability effects in pure s<strong>and</strong>stone.<br />

Durl<strong>of</strong>sky (1992) showed that cross bedding in æolian s<strong>and</strong>stones can reduce effective<br />

permeability by an order <strong>of</strong> magnitude <strong>and</strong> induce an anisotropy <strong>of</strong> maximum to<br />

minimum permeability (kmax/kmin) greater than 5. Similarly, it has been shown that the<br />

presence <strong>of</strong> open joints can increase effective permeability by 2 or more orders <strong>of</strong><br />

magnitude, <strong>and</strong> that the presence <strong>of</strong> low porosity deformation b<strong>and</strong>s (DBs) can reduce<br />

effective permeability by 2 or more orders <strong>of</strong> magnitude (Antonellini <strong>and</strong> Aydin, 1994;<br />

Lothe et al., 2002; Taylor <strong>and</strong> Pollard, 2000; Taylor et al., 1999). As a natural<br />

consequence <strong>of</strong> their dominant tabular/planar geometry, both joints <strong>and</strong> DBs also can be<br />

expected to induce permeability anisotropy (Antonellini <strong>and</strong> Aydin, 1995).<br />

142


In this paper, we use numerical methods <strong>and</strong> computer codes developed by Durl<strong>of</strong>sky<br />

(1991; 1992) to quantify the influence <strong>of</strong> realistic DB patterns on bulk s<strong>and</strong>stone<br />

permeability in two dimensions (2-D). We restrict ourselves to 2-D analyses so as not to<br />

obscure the essence <strong>of</strong> the approach—or the potentially important impact <strong>of</strong> DBs on bulk<br />

s<strong>and</strong>stone permeability—with the additional conceptual <strong>and</strong> computational complexity<br />

inherent to 3-D treatments. Also, ins<strong>of</strong>ar as DBs are themselves grossly tabular/planar<br />

features, with most arrays consisting <strong>of</strong> just one or two dominant orientations, the DB<br />

pattern visible on any 2-D observation plane (e.g. outcrop face) that is oriented at high<br />

angle to the b<strong>and</strong>s will be substantially similar for all adjacent parallel planes. That is,<br />

pattern <strong>and</strong> permeability continuity in the third dimension can reasonably be assumed.<br />

Nonetheless, the numerical methods <strong>and</strong> analyses presented here could naturally be<br />

extended to 3-D.<br />

We apply the 2-D method to three distinct <strong>and</strong> characteristic DB patterns—parallel,<br />

cross-hatch <strong>and</strong> anastomosing—as recognized <strong>and</strong> mapped within the æolian Jurassic<br />

Aztec s<strong>and</strong>stone <strong>of</strong> the Valley <strong>of</strong> Fire State Park, Nevada (Figure 6.1). The Aztec<br />

s<strong>and</strong>stone constitutes an excellent exhumed analog for active æolian s<strong>and</strong>stone reservoirs<br />

potentially affected by DBs, such as the Anschutz Ranch East Field in the Nugget<br />

s<strong>and</strong>stone <strong>of</strong> Wyoming, Idaho <strong>and</strong> Utah (Lindquist, 1988). The Nugget is considered a<br />

chronostratigraphic <strong>and</strong> depositional equivalent <strong>of</strong> the Aztec (Marzolf, 1986), <strong>and</strong><br />

Anschutz Ranch comprises a large anticlinal trap within the Nugget. Production evidence<br />

for Anschutz Ranch indicates that lowered porosities <strong>and</strong> permeabilities related to<br />

abundant DB fabrics (observed in core <strong>and</strong> outcrop) have degraded overall reservoir<br />

quality, introduced flow barriers <strong>and</strong> produced anisotropic behavior (Lewis, 1993).<br />

Production problems attributed to deformation b<strong>and</strong>s have also been reported for the<br />

Nubian S<strong>and</strong>stone in Egypt <strong>and</strong> for s<strong>and</strong>stone reservoirs located <strong>of</strong>fshore from Nigeria<br />

(Olsson et al., 2003).<br />

3. Deformation b<strong>and</strong>s<br />

Deformation b<strong>and</strong>s are thin, tabular, bounded features that accommodate pore-loss<br />

compaction, <strong>of</strong>ten in association with shear displacement, in s<strong>and</strong>stones via granular<br />

rearrangement, cataclasis <strong>and</strong> chemical diagenesis (Aydin, 1978; Aydin <strong>and</strong> Johnson,<br />

1983; Du Bernard, 2002; Engelder, 1974; Jamison <strong>and</strong> Stearns, 1982) (Figure 6.2).<br />

143


NV UT<br />

CA<br />

AZ<br />

Park Road<br />

Map<br />

Detail<br />

Park Boundary<br />

Route 169<br />

0 5<br />

km<br />

NEVADA<br />

ARIZONA<br />

10 km<br />

Valley<br />

<strong>of</strong> Fire<br />

State Park<br />

N<br />

LAKE<br />

MEAD<br />

115 o 30'<br />

Figure 6.1. Schematic location map for the Valley <strong>of</strong> Fire State Park in Nevada, about 60<br />

km northeast <strong>of</strong> Las Vegas. The colorful Aztec s<strong>and</strong>stone comprises the dominant<br />

lithology <strong>and</strong> primary tourist attraction in the park.<br />

144<br />

36 o 15'


DB tip<br />

DB<br />

DBs<br />

Grains<br />

Pores<br />

Figure 6.2. Compactive deformation b<strong>and</strong>s (DBs) as seen in the Aztec s<strong>and</strong>stone, Valley<br />

Clay<br />

<strong>of</strong> Fire, NV. These DBs are tabular, bounded features <strong>of</strong> pore-loss compaction that<br />

typically crop out as systematic <strong>and</strong> <strong>of</strong>ten extensive arrays <strong>of</strong> resistant, positive-relief<br />

ridges within a relatively undeformed s<strong>and</strong>stone matrix (top). Individual b<strong>and</strong>s are<br />

approximately planar, displaying distinct tips even when closely spaced—width <strong>of</strong><br />

measuring tape is one inch (bottom left). Viewed in thin section, grain cracking,<br />

rearrangement <strong>and</strong> pore loss are readily apparent, illustrating why DBs act as barriers to<br />

fluid flow. Unlike many DBs, however, those in the Aztec tend to exhibit limited<br />

cataclasis <strong>and</strong> little or no net shear <strong>of</strong>fset (bottom right).<br />

145


Commonly from ~1 mm to ~1.5 cm in thickness <strong>and</strong> ~1 to ~100+ m in extent, DBs<br />

generally exhibit displacements, both closing-mode (compaction) <strong>and</strong> shear, on the order<br />

<strong>of</strong> millimeters (Antonellini et al., 1994; Aydin <strong>and</strong> Johnson, 1978; Mollema <strong>and</strong><br />

Antonellini, 1996). All types <strong>of</strong> DBs are characterized by the absence <strong>of</strong> any distinct<br />

plane <strong>of</strong> displacement discontinuity. The petrophysical changes accommodated within<br />

DBs—compaction, grain size reduction, chemical diagenesis <strong>and</strong> clay infiltration—<br />

typically produce internal porosity reductions <strong>of</strong> about an order <strong>of</strong> magnitude—<strong>of</strong>ten<br />

down to a residual <strong>of</strong> only 1 or 2% (Ahlgren, 2001; Antonellini <strong>and</strong> Aydin, 1994). The<br />

porosity loss, in combination with related reductions in mean pore throat diameter <strong>and</strong><br />

connectivity, can produce drastic drops in permeability within DBs, <strong>and</strong> significant<br />

efforts have been made to quantify this phenomenon.<br />

Freeman (1990) used gas permeameter readings on plugs <strong>of</strong> Aztec s<strong>and</strong>stone to<br />

determine that the presence <strong>of</strong> DBs can cause 2-order-<strong>of</strong>-magnitude reductions in<br />

effective permeability relative to DB-free rock. Antonellini <strong>and</strong> Aydin (1994) used a gas<br />

injection mini-permeameter to measure individual DB permeability in a variety <strong>of</strong><br />

s<strong>and</strong>stones, finding DBs to be 2 to 4 orders <strong>of</strong> magnitude less permeable than the host<br />

rock, with an average permeability drop <strong>of</strong> 3 orders <strong>of</strong> magnitude. They also found that<br />

the relative permeability drop is greater for DBs in higher porosity s<strong>and</strong>stones (Entrada<br />

<strong>and</strong> Navajo at >20%) than in lower porosity s<strong>and</strong>stones (Morrison <strong>and</strong> Chinle at


did not report permeability values for individual b<strong>and</strong>s, Taylor <strong>and</strong> Pollard (2000) used<br />

his data to estimate that average DB permeability is about 2.3 orders <strong>of</strong> magnitude less<br />

than in the host rock. Gibson (1998) found that the permeability <strong>of</strong> DBs depends on the<br />

clay content <strong>of</strong> the parent s<strong>and</strong>stone <strong>and</strong> can vary from 1 to 4 orders <strong>of</strong> magnitude, with<br />

an average value <strong>of</strong> ~2.2 orders <strong>of</strong> magnitude.<br />

4. Characteristic DB patterns in the Aztec s<strong>and</strong>stone<br />

The Valley <strong>of</strong> Fire State Park lies about 60 km northeast <strong>of</strong> Las Vegas <strong>and</strong> about 30<br />

km west <strong>of</strong> the Utah border (Figure 6.1). The predominant outcrop unit within the park is<br />

the Jurassic Aztec s<strong>and</strong>stone, commonly considered correlative with the Navajo<br />

s<strong>and</strong>stone. The Aztec is an æolian, subarkosic s<strong>and</strong>stone up to 2 km thick typified by<br />

large-scale cross-bed sets up to 10+ meters thick (Marzolf, 1983; Myers, 1999). The<br />

Aztec has a mean grain size <strong>of</strong> about 0.25 mm, an average porosity <strong>of</strong> around 20%<br />

(Taylor <strong>and</strong> Pollard, 2000) <strong>and</strong> an average matrix permeability ranging from 0.1 to 1.0<br />

darcys (Antonellini <strong>and</strong> Aydin, 1994; Myers, 1999). The Aztec exhibits a variety <strong>of</strong><br />

brittle <strong>structural</strong> fabrics that include DBs, joints <strong>and</strong> faults, with DBs consistently<br />

appearing to be the oldest features (Hill, 1989; Myers, 1999; Taylor <strong>and</strong> Pollard, 2000).<br />

Hill (1989) hypothesized that DB formation within the Aztec was related to widespread<br />

compression during emplacement <strong>of</strong> the Summit/Willow Tank <strong>and</strong> Muddy Mountain<br />

thrusts during the Early Cretaceous Sevier orogeny.<br />

Fieldwork in the Valley <strong>of</strong> Fire has revealed at least three wide spread, systematic <strong>and</strong><br />

volumetrically significant patterns <strong>of</strong> dominantly compactive DBs in the Aztec s<strong>and</strong>stone<br />

over an area <strong>of</strong> more than 10 km 2 . All <strong>of</strong> these characteristic patterns—parallel, cross-<br />

hatch <strong>and</strong> anastomosing—have previously been described at various locations, though<br />

generally using different nomenclature (Antonellini <strong>and</strong> Aydin, 1994, 1995; Aydin <strong>and</strong><br />

Reches, 1982; Burhannudinnur <strong>and</strong> Morley, 1997; Davis, 1998; Hill, 1989; Jamison <strong>and</strong><br />

Stearns, 1982; Mollema <strong>and</strong> Antonellini, 1996; Taylor <strong>and</strong> Pollard, 2000; Underhill <strong>and</strong><br />

Woodcock, 1987; Woodcock <strong>and</strong> Underhill, 1987). Within the Aztec, the 2-D density <strong>of</strong><br />

DBs exposed in any given outcrop (area DBs/outcrop area) ranges from ~1% up to ~20%.<br />

Assuming reasonable continuity in the third dimension, as is clearly suggested by field<br />

observations in outcrop areas <strong>of</strong> high relief, the volume percent density <strong>of</strong> DBs also<br />

ranges between 1% <strong>and</strong> 20%.<br />

147


Although systematic arrays <strong>of</strong> DBs pervade the middle <strong>and</strong> upper Aztec, DB-poor<br />

exposures tens to hundreds <strong>of</strong> meters or more on a side are also common. The reasons for<br />

this patchwork quality to the DB fabric at a variety <strong>of</strong> scales is as yet poorly understood,<br />

but may be related to the æolian sedimentary architecture <strong>of</strong> the Aztec <strong>and</strong> differences in<br />

how distinct dune packages responded to tectonic loading during DB formation.<br />

Nonetheless, any given characteristic DB pattern recognized in outcrop can be seen to<br />

persist at length scales ranging from meters (the thickness <strong>of</strong> small cross-bed packages)<br />

to tens <strong>of</strong> meters (the thickness <strong>of</strong> large cross-bed packages) to hundreds <strong>of</strong> meters (for<br />

DBs that pass through multiple dune boundaries).<br />

4.1. Parallel<br />

Patterns <strong>of</strong> approximately parallel DBs crop out extensively throughout the Aztec<br />

over areas ranging from tens to hundreds <strong>of</strong> meters on a side. Individual b<strong>and</strong>s within the<br />

parallel sets commonly range up to ~1.5 cm thick <strong>and</strong> are spaced anywhere from<br />

centimeters to meters apart (Figure 6.3). It is not uncommon, however, to see groups <strong>of</strong><br />

individually distinct DBs running parallel to each other only millimeters apart, sometimes<br />

merging into what appears in outcrop to be a single b<strong>and</strong> 10 cm or more thick. Also,<br />

groups <strong>of</strong> closely spaced DBs <strong>of</strong>ten form distinct sets spaced a meter or more from the<br />

next adjacent, closely spaced set. DBs cropping out in parallel patterns throughout the<br />

Aztec cluster around three mean strike/dip orientations—350°/70°, 210°/60° <strong>and</strong><br />

240°/30° (using the convention <strong>of</strong> dip direction oriented 90° clockwise from the strike<br />

azimuth). The first two <strong>of</strong> these sets cut across depositional bedding at a high angle,<br />

while the third generally runs sub-parallel to bedding. The north-trending, steeply east-<br />

dipping DB orientation (350°/70°) strongly dominates in terms <strong>of</strong> abundance, persistence<br />

<strong>and</strong> distribution, <strong>and</strong> is the one shown in both Figures 6.2 <strong>and</strong> 6.3.<br />

4.2. Cross-hatch<br />

Cross-hatch patterns <strong>of</strong> DBs, which generally crop out over areas from meters to 10s<br />

<strong>of</strong> meters on a side, consist <strong>of</strong> two distinct sets <strong>of</strong> parallel b<strong>and</strong>s that commonly intersect<br />

each other at a high angle (Figure 6.4). In the Aztec, individual b<strong>and</strong>s within the cross-<br />

hatch patterns exhibit the same range <strong>of</strong> widths as b<strong>and</strong>s in the parallel patterns <strong>and</strong> the<br />

same tendency to occur as closely spaced pairs. Spacing between b<strong>and</strong>s ranges from 2 cm<br />

148


0 meters<br />

2<br />

Figure 6.3. A typical outcrop pattern <strong>of</strong> approximately parallel DBs in the Aztec<br />

s<strong>and</strong>stone. View is to the north, bedding is approximately horizontal.<br />

cm<br />

0 8<br />

1 meter<br />

Figure 6.4. A typical cross-hatch pattern <strong>of</strong> DBs in the Aztec s<strong>and</strong>stone. B<strong>and</strong>s in the two<br />

sets are at high angle to each other <strong>and</strong> to bedding (approximately coincident with the<br />

outcrop face (main photo). Looking down the axis <strong>of</strong> intersection <strong>of</strong> a cross-hatch set<br />

illustrates how planar DBs can be, <strong>and</strong> how sharp their intersections (inset photo).<br />

149


to 2 m, with both sets in a cross-hatch pattern typically exhibiting similar spacing. The<br />

cross-hatch pattern is both less abundant <strong>and</strong> less persistent than the parallel <strong>and</strong><br />

anastomosing DB patterns, <strong>and</strong> <strong>of</strong>ten at least one DB set within a cross-hatch pattern<br />

terminates at the nearest dune boundary. The two sets <strong>of</strong> DBs within the cross-hatch<br />

patterns exhibit mean orientations generally similar to those <strong>of</strong> any two <strong>of</strong> the three<br />

dominant orientations described above for the parallel pattern case, <strong>and</strong> commonly<br />

intersect with an acute angle <strong>of</strong> ~80°.<br />

4.3. Anastomosing<br />

Anastomosing patterns <strong>of</strong> DBs consist <strong>of</strong> sub-parallel, intersecting b<strong>and</strong>s generally<br />

exhibiting the same dominant, high-angle-to-bedding orientation described above for the<br />

parallel pattern, but with trends <strong>and</strong> dips varying by 25 degrees or more over the scale <strong>of</strong><br />

centimeters to meters (Figure 6.5). Deformation b<strong>and</strong>s within the anastomosing patterns<br />

exhibit ranges <strong>of</strong> thickness (up to ~1.5 cm) <strong>and</strong> spacing (centimeters to meters) similar to<br />

those reported above for the parallel case. Anastomosing b<strong>and</strong> sets have previously been<br />

described in the literature, but always as existing in narrow zones <strong>of</strong> very closely spaced<br />

b<strong>and</strong>s—typically associated with faults having distinct slip planes (Aydin, 1978; Aydin<br />

<strong>and</strong> Johnson, 1983). In the Aztec, however, anastomosing DB arrays commonly cover<br />

areas up to hundreds <strong>of</strong> meters on a side, while passing through multiple dune<br />

boundaries. Indeed, the anastomosing pattern <strong>of</strong> steeply dipping, generally north-trending<br />

DBs dominates the Aztec, <strong>and</strong> the arrays <strong>of</strong> approximately parallel DBs described above<br />

can reasonably be considered as simply more orderly sub zones within the greater,<br />

anastomosing configuration.<br />

5. Numerical modeling methods<br />

Numerical techniques can be used to calculate effective permeability for any<br />

characteristic DB pattern that can be represented as a finely meshed grid <strong>of</strong> cells, each<br />

cell specifying either DB or s<strong>and</strong>stone matrix permeability (Figure 6.6). The method<br />

involves solving a single fluid phase, steady state pressure equation subject to spatially<br />

periodic boundary conditions over the entire grid to yield this upscaled or effective<br />

permeability. The fluid flux <strong>and</strong> pressure boundary conditions are said to be spatially<br />

periodic ins<strong>of</strong>ar as they implicitly assume that the domain block in question is surrounded<br />

150


0 meter<br />

Figure 6.5. A typical anastomosing pattern <strong>of</strong> DBs located along trend, but about 200 m<br />

south <strong>of</strong> the outcrop shown in Figure 6.3. View is to the north, bedding is very nearly<br />

coincident with the outcrop face in the foreground.<br />

1<br />

151


y identical blocks all subject to the same regional pressure gradient that drives flow.<br />

This modeling approach derives from an application <strong>of</strong> homogenization theory as<br />

presented in detail by Durl<strong>of</strong>sky (1991; 1992).<br />

5.1. Governing equations<br />

Single-phase, steady state, incompressible flow through a heterogeneous porous<br />

media in the absence <strong>of</strong> sources or sinks can be described by Darcy’s law <strong>and</strong> the<br />

requirements <strong>of</strong> continuity as:<br />

1<br />

u = − k ⋅∇p<br />

µ<br />

∇⋅u = 0 (2)<br />

where u is the local fluid velocity vector, ∇ p is the local fluid pressure gradient<br />

vector, µ is the fluid viscosity, <strong>and</strong> k is the tensor representing directional permeability.<br />

The units <strong>of</strong> k are length squared, as can be demonstrated by dimensional analysis <strong>of</strong><br />

Equation (1). In a 2-D x-y coordinate system, k can be represented as:<br />

⎡kxx<br />

kxy⎤<br />

k = ⎢ ⎥ . (3)<br />

⎣kyx<br />

kyy⎦<br />

where kxy is the permeability component relating fluid velocity in the x-direction to<br />

pressure gradient in the y-direction. Permeability is commonly measured in darcys (1<br />

darcy = 9.87 x 10 -9 cm 2 ), <strong>and</strong> the native permeability <strong>of</strong> uniform s<strong>and</strong>stone ranges from<br />

about 10 -2 to 10 2 darcys in any given direction (Freeze <strong>and</strong> Cherry, 1979).<br />

The homogenization approach considers k to vary over two spatial scales—a fine<br />

scale, f, <strong>and</strong> a coarse scale, c, with f


1 meter 1 meter<br />

Figure 6.6. An idealized DB pattern displaying characteristics <strong>of</strong> both the cross-hatch<br />

<strong>and</strong> anastomosing types at the scale <strong>of</strong> one square meter (left); <strong>and</strong> a representation <strong>of</strong><br />

this pattern on a 40 x 40 cell finite difference/finite element grid where black corresponds<br />

to DB permeability <strong>and</strong> white indicates native s<strong>and</strong>stone permeability (right).<br />

f-scale<br />

f 2<br />

c 1 ~10 to >100m c-scale<br />

f1<br />

100m c-scale<br />

f1<br />


contiguous areas. This type <strong>of</strong> characteristic pattern replication lends itself naturally to<br />

numerical solution using spatially periodic boundary conditions, as detailed below.<br />

It is important to note that spatial periodicity does not imply fractal scaling. Based on<br />

our observations <strong>of</strong> DBs in the Aztec, we can state that their characteristic patterns<br />

definitely are not fractal. Also, it is not necessary for actual pattern periodicity in nature<br />

to be nearly as distinct <strong>and</strong> regular as idealized in Figure 6.7 for the numerical methods<br />

described here to work effectively.<br />

Using the two-scale representation <strong>of</strong> k <strong>and</strong> combining Darcy’s law with continuity<br />

yields the pressure equation for steady state, incompressible flow:<br />

[ ( c,f<br />

) ⋅ ∇ = 0<br />

∇ ⋅ k p]<br />

(4)<br />

where k varies on both the fine <strong>and</strong> coarse scales. Our intent is to calculate an<br />

effective permeability tensor, k * , that varies only on the c-scale but that implicitly<br />

accounts for permeability variations on the f-scale. This allows us to model the system<br />

using an equation <strong>of</strong> the form <strong>of</strong> (4) but with k * (c) replacing k (c,f). The solution to this<br />

“upscaled” equation is much less computationally dem<strong>and</strong>ing than the solution <strong>of</strong><br />

equation (4), because variations on the fine scale f need not be resolved. The essential<br />

step is to transform k (c,f) to k * (c) by calculating the effective permeability tensor <strong>of</strong> a c-<br />

scale block containing f-scale heterogeneities (DBs). It has been shown (e.g. Bourgeat,<br />

1984)) that homogenized upscaling is exact when two distinct spatial scales <strong>of</strong><br />

permeability variation exist. Subsequent work (e.g. Durl<strong>of</strong>sky, 1991) has demonstrated<br />

that upscaling can also be applied with reasonable accuracy to more realistic situations<br />

such as represented by DBs in the Aztec, where distinct spatial scales <strong>of</strong> permeability<br />

variation cannot be defined.<br />

5.2. Boundary conditions <strong>and</strong> solution<br />

In order to drive large-scale flow, pressure variation on the coarse scale c is imposed.<br />

This can be expressed generally as:<br />

p = p 0 + G ⋅ (c − c 0 ) (5)<br />

where c0 is the center <strong>of</strong> the region under study, p0 is the fluid pressure at c0, <strong>and</strong> G is<br />

the local fluid pressure gradient ( ∇p ) over the c-scale at c0.<br />

The average flow through a<br />

154


periodic, f-scale block, denoted , is related to the c-scale pressure gradient by the<br />

effective permeability tensor k * :<br />

k G<br />

* 1<br />

u = - ⋅ . (6)<br />

µ<br />

To calculate k * , we solve the pressure equation (4) over the f-scale block <strong>and</strong> then<br />

compute the average flux through it. The boundary conditions for this solution<br />

assume periodicity <strong>of</strong> the system up to the c-scale, as well as the imposition <strong>of</strong> the<br />

pressure gradient G. The boundary specifications require that flow across the domain<br />

boundaries be periodic over the f-scale block, <strong>and</strong> that the pressure be periodic but with a<br />

jump due to the large-scale gradient. Boundary conditions can be specified explicitly for<br />

an f-scale block (Figure 6.8) by resolving G into its two components: G = G1i1 + G2i2,<br />

where i1 <strong>and</strong> i2 are unit vectors in the coordinate directions f1 <strong>and</strong> f2, respectively.<br />

Comparing pressures <strong>and</strong> velocities on opposing sides with unit normals n1 <strong>and</strong> n2 yields:<br />

yields:<br />

p ( f , f 0)<br />

= p(<br />

f , f = b)<br />

− G × b<br />

(7a)<br />

1<br />

2 = 1 2<br />

2<br />

( 1 , f 2 = 0)<br />

⋅n<br />

1 = −u(<br />

f1,<br />

f 2 = b)<br />

n2<br />

u f ⋅ . (7b)<br />

Comparing pressures <strong>and</strong> velocities on opposing sides with unit normals n3 <strong>and</strong> n4<br />

p ( f 0,<br />

f ) = p(<br />

f = a,<br />

f ) − G × a<br />

(7c)<br />

1 = 2<br />

1 2 1<br />

( 1 = 0, f 2 ) ⋅n<br />

3 = −u(<br />

f1<br />

= a,<br />

f 2 ) n4<br />

u f ⋅ . (7d)<br />

To compute the complete effective permeability tensor k * , two problems must be<br />

solved, one for which G1=0, G2≠0 <strong>and</strong> one for which G1≠0, G2=0. By specifying a<br />

reference value for pressure at a single point in the domain, equation (4) can be solved<br />

subject to equations (7). The flux through the f-scale block can then be determined by<br />

integrating the velocities, <strong>and</strong> finally k * computed from equation (6). Durl<strong>of</strong>sky (1991)<br />

presents this computational approach in full detail.<br />

155


(0,b)<br />

n 3<br />

(0,0)<br />

f 2<br />

n 2<br />

n 1<br />

(a,0)<br />

(a,b)<br />

Figure 6.8. Gridded block representing permeability variation on the fine scale to which<br />

periodic fluid flux <strong>and</strong> pressure boundary conditions are applied in solving for effective<br />

permeability. The lower left corner is the origin (0,0) <strong>and</strong> the block has dimensions a in<br />

the f1 coordinate direction <strong>and</strong> b in the f2 direction, while n1 through n4 are unit normals<br />

to its sides (adapted from Durl<strong>of</strong>sky, 1991).<br />

156<br />

n 4<br />

f 1


5.3. Finite difference/finite element method<br />

For the purposes <strong>of</strong> computational efficiency, a coupled, two-step finite<br />

difference/finite element numerical method was used to compute k * . In order to<br />

adequately resolve centimeter-thick DBs within meter-scale patterns, finely discretized<br />

grids a few hundred cells on a side were generally necessary, with each cell being<br />

assigned one <strong>of</strong> two isotropic permeability values: kmatrix or kb<strong>and</strong> (e.g. Figure 6.6). Again<br />

for the sake <strong>of</strong> computational efficiency, two coarsening steps were used to compute the<br />

final, upscaled permeability tensor. For example, to compute k * for a pattern meshed at<br />

400 × 400 cells, the first coarsening step might be to a 20 x 20 cell system (yielding 400<br />

intermediate permeability tensors) <strong>and</strong> then to the final k * in the second step. An efficient<br />

finite difference method was used to make the first coarsening step. The second step to k *<br />

required a finite element procedure that is less efficient, but capable <strong>of</strong> h<strong>and</strong>ling the<br />

matrix <strong>of</strong> full permeability tensors produced during the first step. Repeat testing revealed<br />

that the final k * computed is relatively insensitive to the size <strong>of</strong> the coarsening steps used,<br />

varying by less than 10% for the patterns studied.<br />

6. Effective permeability results<br />

The governing equations, boundary conditions <strong>and</strong> numerical methods detailed above<br />

were used to calculate 2-D effective principal permeabilities for each <strong>of</strong> the three<br />

characteristic DB patterns described for the Aztec. For the simpler, more regular parallel<br />

<strong>and</strong> cross-hatch examples, we used idealized patterns to assess how effective<br />

permeability varies as a function <strong>of</strong> the relevant physical variables—b<strong>and</strong> thickness,<br />

spacing <strong>and</strong> intersection angle—<strong>and</strong> present analytical solutions for comparison where<br />

possible. For the more generally irregular anastomosing pattern, which exhibits<br />

substantial spatial variability that defies analytical treatment, we assess effective<br />

permeability for a representative outcrop pattern. In order to emphasize reductions in<br />

permeability attributable to DBs, all permeability values reported below <strong>and</strong> in the<br />

figures are normalized by that <strong>of</strong> the undeformed s<strong>and</strong>stone matrix (~ 0.1 to 1.0 darcys)<br />

to yield dimensionless ratios <strong>of</strong> relative permeability. Thus in the analyses that follow,<br />

DBs are assigned isotropic internal permeability values <strong>of</strong> 10 -2 <strong>and</strong> 10 -3 , representing the<br />

mid range <strong>of</strong> values for DBs as reported by other workers <strong>and</strong> summarized above.<br />

157


6.1. Parallel<br />

Effective permeability for any idealized parallel DB pattern can be calculated<br />

analytically, as previously reported (Antonellini <strong>and</strong> Aydin, 1994; Taylor et al., 1999).<br />

Our numerical modeling results matched these exact analytical solutions for parallel<br />

patterns, so only the latter are described here for the purpose <strong>of</strong> building intuition. The<br />

analytical methods are based on the general analysis <strong>of</strong> flow in layered media by Freeze<br />

<strong>and</strong> Cherry, pages 30-34 (1979).<br />

B<strong>and</strong>-parallel effective permeability, k , can be calculated as a weighted arithmetic<br />

average <strong>of</strong> the b<strong>and</strong> <strong>and</strong> matrix permeability values:<br />

w k + * b b ( W − wb)km k = p<br />

W<br />

*<br />

p<br />

where kb is the DB permeability, km is the matrix permeability, wb is cumulative DB<br />

thickness, <strong>and</strong> W is the width <strong>of</strong> the area studied. Note that using wb/W as a measure <strong>of</strong><br />

volume fraction assumes DB continuity in the third dimension, as suggested by field<br />

observations in the Aztec. B<strong>and</strong>-normal effective permeability can be calculated as the<br />

weighted harmonic average:<br />

k<br />

*<br />

n<br />

=<br />

( ( w / k ) + ( W − w ) / k )<br />

b<br />

b<br />

W<br />

b<br />

m<br />

The b<strong>and</strong>-parallel <strong>and</strong> b<strong>and</strong>-normal effective permeabilities represent the maximum<br />

<strong>and</strong> minimum (i.e. principal) values, respectively. An effective permeability tensor, k * ,<br />

for the parallel system thus reads:<br />

*<br />

* ⎡k<br />

⎤ p 0<br />

k = ⎢ * ⎥<br />

(10)<br />

⎣ 0 kn<br />

⎦<br />

B<strong>and</strong>-normal <strong>and</strong> b<strong>and</strong>-parallel effective permeability values were calculated for kb/km<br />

ratios <strong>of</strong> both 10 -2 <strong>and</strong> 10 -3 <strong>and</strong> DB volume fractions (wb/W) from 0% to 100% (Figure<br />

6.9). B<strong>and</strong>-normal effective permeability drops much more quickly at low wb/W than<br />

does b<strong>and</strong>-parallel effective permeability. For kb/km=10 -3 , b<strong>and</strong>-normal effective<br />

permeability is already reduced by an order <strong>of</strong> magnitude at wb/W=1% (1 mm b<strong>and</strong>s<br />

spaced at 10 cm) <strong>and</strong> 2 orders <strong>of</strong> magnitude at wb/W=10% (1cm b<strong>and</strong>s spaced at 10 cm).<br />

158<br />

(8)<br />

(9)


Relative Effective Permeability<br />

1.000<br />

0.100<br />

0.010<br />

kn (kb/km= 10-2)<br />

kp (kb/km= 10-2)<br />

kn (kb/km= 10 ) -3<br />

kp (kb/km= 10 ) -3<br />

0.001<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Volume Fraction Deformation B<strong>and</strong>s (%)<br />

Figure 6.9. B<strong>and</strong>-normal <strong>and</strong> b<strong>and</strong>-parallel effective permeability values (equivalent to<br />

the principal permeability values) as a function <strong>of</strong> DB volume fraction for perfectly<br />

parallel patterns <strong>and</strong> DB/s<strong>and</strong>stone matrix permeability ratios <strong>of</strong> 10 -2 <strong>and</strong> 10 -3 . The range<br />

<strong>of</strong> DB volume fraction observed in the Aztec s<strong>and</strong>stone is highlighted in gray.<br />

159


By contrast, b<strong>and</strong>-parallel effective permeability is not reduced by an order <strong>of</strong> magnitude<br />

until wb/W=90%, a DB volume density far exceeding the approximate upper limit <strong>of</strong> 20%<br />

observed in the Aztec. Thus for parallel DB patterns within the realistic wb/W range, a<br />

permeability anisotropy <strong>of</strong> up to 2 orders <strong>of</strong> magnitude can be expected, with the<br />

maximum permeability oriented in the b<strong>and</strong>-parallel direction (Figure 6.9). Note also that<br />

b<strong>and</strong>-parallel effective permeability for both kb/km=10 -2 <strong>and</strong> 10 -3 are virtually<br />

indistinguishable below wb/W=80%.<br />

6.2. Cross-hatch<br />

Cross-hatch DB sets in the Valley <strong>of</strong> Fire exhibit a consistent acute intersection angle<br />

<strong>of</strong> ~80°. In other locales, intersection angles ranging from ~20° to ~90° have been<br />

reported (Aydin <strong>and</strong> Reches, 1982; Hill, 1989; Jamison <strong>and</strong> Stearns, 1982; Underhill <strong>and</strong><br />

Woodcock, 1987). We begin our analysis <strong>of</strong> the effective permeability <strong>of</strong> cross-hatch<br />

patterns by studying those <strong>of</strong> evenly spaced, orthogonal DB sets as a function <strong>of</strong> kb/km<br />

<strong>and</strong> wb/W. Figure 6.10 compares our numerical modeling results for kb/km=10 -2 to an<br />

analytical approximation presented by Taylor <strong>and</strong> Pollard (2000) <strong>and</strong> b<strong>and</strong>-normal<br />

effective permeability as calculated by harmonic averaging. The numerical <strong>and</strong> harmonic<br />

averaging results match well over the entire wb/W range, but correspond to the analytical<br />

approximation only for wb/W below about 5%. For kb/km=10 -3 the wb/W range <strong>of</strong><br />

correspondence drops to


Relative Effective Permeability<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

Taylor <strong>and</strong> Pollard (2000)<br />

Numerical Model<br />

Harmonic Average<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Volume Fraction Deformation B<strong>and</strong>s (%)<br />

Figure 6.10. Isotropic effective permeability for cross-hatch patterns <strong>of</strong> evenly spaced,<br />

perpendicular DB sets as a function <strong>of</strong> DB volume fraction for the DB/s<strong>and</strong>stone matrix<br />

permeability ratio <strong>of</strong> 10 -2 . Results from the finite difference-finite element method used<br />

in this paper (FD-FE) modeling are plotted against both the analytical approximation <strong>of</strong><br />

Taylor <strong>and</strong> Pollard (2000) <strong>and</strong> the permeability computed normal to just one <strong>of</strong> the DB<br />

sets using the harmonic averaging method. The FD-FE <strong>and</strong> harmonic averaging methods<br />

agree well over the entire volume fraction range, while the approximation holds only for<br />

volume fractions up to about 5%. The range <strong>of</strong> DB volume fraction observed in the Aztec<br />

s<strong>and</strong>stone is highlighted in gray.<br />

161


Relative Effective Permeability<br />

1.00<br />

0.10<br />

Volume Fraction Deformation B<strong>and</strong>s (%)<br />

o<br />

kmax (15 )<br />

kmax (30 o)<br />

kmax (60 o)<br />

k-isotropic (90 o)<br />

kmin (60 o)<br />

kmin (30 o)<br />

o kmin (15 )<br />

0.01<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Figure 6.11. Principal effective permeability values (maximum <strong>and</strong> minimum) for evenly<br />

spaced cross-hatch DB patterns with a DB/s<strong>and</strong>stone matrix permeability ratio <strong>of</strong> 10 -2 .<br />

Relative permeability is plotted as a function <strong>of</strong> DB volume fraction for acute intersection<br />

angles <strong>of</strong> 15 o , 30 o , 60 o <strong>and</strong> 90 o . The range <strong>of</strong> DB volume fraction observed in the Aztec<br />

s<strong>and</strong>stone is highlighted in gray. Inset figure (lower left) illustrates the 2-D permeability<br />

ellipse for an evenly spaced cross-hatch set <strong>of</strong> DBs intersecting at an acute angle θ.<br />

162


where kmax <strong>and</strong> kmin are the maximum <strong>and</strong> minimum principal effective permeability<br />

values, kp is the permeability parallel to the b<strong>and</strong>s when θ=0°, kn is the permeability<br />

normal to the b<strong>and</strong>s when θ =0°, <strong>and</strong> k+ is the isotropic permeability when half <strong>of</strong> the<br />

b<strong>and</strong>s are set orthogonal to the other half (θ =90°). Since the analytical solution for k+<br />

only holds for low percentages <strong>of</strong> DBs, we use k+ as calculated by our numerical method<br />

in the equations above to find the maximum <strong>and</strong> minimum effective permeability values<br />

as functions <strong>of</strong> wb/W, θ <strong>and</strong> kb/km.<br />

Figure 6.11 shows the results for kb/km=10 -2 over the entire range <strong>of</strong> wb/W for<br />

intersection angles <strong>of</strong> θ =15°, 30°, 60°<strong>and</strong> 90°. Both principal effective permeability<br />

values decrease as wb/W increases, with the reduction in kmin being most pronounced. The<br />

degree <strong>of</strong> this permeability anisotropy also decreases as θ increases, with an isotropic<br />

permeability drop attained at θ =90°. For kb/km=10 -3 , the anisotropy introduced is even<br />

more pronounced, approaching 2 orders <strong>of</strong> magnitude as θ approaches 0° <strong>and</strong> the pattern<br />

becomes essentially parallel. For the generally high intersection angles (~80) <strong>and</strong> low<br />

volume percents (~10%) exhibited for cross-hatch DB patterns in the Aztec, relatively<br />

isotropic permeability reductions ranging from about 0.7 to1.5 orders <strong>of</strong> magnitude<br />

would be expected.<br />

6.3. Anastomosing<br />

No analytical solutions exist for the systematic, but irregular anastomosing pattern, so<br />

the purely numerical method was used to calculate effective permeability tensors for a<br />

finely discretized version <strong>of</strong> a real, 12 x 15 m DB outcrop pattern mapped in the Aztec on<br />

a low altitude balloon photograph (Figure 6.12). Not surprisingly, the maximum effective<br />

permeability is oriented roughly parallel to the mean trend <strong>of</strong> the b<strong>and</strong>s, while minimum<br />

permeability is directed roughly perpendicular. We also compared the effective<br />

permeability <strong>of</strong> this real anastomosing pattern to an idealized parallel pattern using the<br />

same average spacing <strong>and</strong> wb/W ratio <strong>of</strong> 15%. As expected, effective b<strong>and</strong>-parallel<br />

*<br />

p<br />

permeability (k ) was somewhat lower for the anastomosing pattern—0.79 versus 0.85 at<br />

kb/km=10<br />

-2 —due presumably to b<strong>and</strong> connectivity <strong>and</strong> associated compartmentalization<br />

*<br />

along trend. On the other h<strong>and</strong>, effective b<strong>and</strong>-normal permeability ( k ) was significantly<br />

higher for the anastomosing pattern—0.26 versus 0.06 at kb/km=10<br />

-2 —due presumably to<br />

163<br />

n


(a) (b)<br />

(c)<br />

0 meters 3<br />

= 10-2 kb /km k b /k m = 10 -3<br />

0.26<br />

0.04<br />

0.79<br />

0.77<br />

kmax log = 0.5 ( kmin)<br />

kmax log = 1.3 ( kmin)<br />

Figure 6.12. Modeling results for a characteristic anastomosing DB pattern from the<br />

Aztec s<strong>and</strong>stone: (a) low altitude balloon photograph <strong>of</strong> outcrop used as base map; (b)<br />

digitized version <strong>of</strong> DB pattern used for numerical modeling as a grid <strong>of</strong> 500 by 375<br />

cells; (c) permeability ellipses <strong>and</strong> order-<strong>of</strong>-magnitude permeability anisotropy values for<br />

DB/s<strong>and</strong>stone matrix permeability ratios <strong>of</strong> 10 -2 <strong>and</strong> 10 -3 .<br />

164


the discontinuous nature <strong>of</strong> the pattern <strong>and</strong> the availability <strong>of</strong> preferred flow paths<br />

crossing a minimum number <strong>of</strong> DBs. Nonetheless, the anastomosing parallel pattern<br />

studied still could induce a significant effective permeability anisotropy between the<br />

maximum (b<strong>and</strong>-parallel) <strong>and</strong> minimum (b<strong>and</strong>-normal) directions. For kb/km=10 -2 about a<br />

0.5 order-<strong>of</strong>-magnitude anisotropy results, for kb/km=10 -3 the anisotropy increases to<br />

about 1.3 orders <strong>of</strong> magnitude.<br />

7. Summary<br />

We have presented a numerical method for calculating the effective permeability<br />

induced by any given pattern <strong>of</strong> deformation b<strong>and</strong>s (DBs), <strong>and</strong> then used it to assess the<br />

2-D effective permeability for volumes <strong>of</strong> porous s<strong>and</strong>stone containing each <strong>of</strong> three<br />

characteristic DB patterns present in the Aztec s<strong>and</strong>stone <strong>of</strong> the Valley <strong>of</strong> Fire State Park,<br />

Nevada—parallel, cross-hatch <strong>and</strong> the dominant anastomosing.<br />

For substantially parallel patterns, we found that b<strong>and</strong>-parallel effective permeability<br />

remains relatively unaffected for realistic DB volume fractions up to 20%. Likewise,<br />

b<strong>and</strong>-parallel effective permeability is only slightly affected by changes in the<br />

permeability contrast between the DBs <strong>and</strong> the host rock. By contrast, b<strong>and</strong>-normal<br />

effective permeability drops even for low volume fractions <strong>of</strong> DBs <strong>and</strong> is extremely<br />

sensitive to varying internal DB permeability. Parallel b<strong>and</strong> patterns such as those<br />

observed in the Aztec can induce permeability anisotropy <strong>of</strong> up to 2 orders <strong>of</strong> magnitude,<br />

with the maximum principal permeability aligned in the b<strong>and</strong>-parallel direction. For<br />

idealized parallel patterns, our modeling results exactly match those computed<br />

analytically.<br />

Cross-hatch b<strong>and</strong> patterns can be analyzed numerically, or in a somewhat simplified<br />

way using a combination <strong>of</strong> numerical <strong>and</strong> analytical methods. They yield highly<br />

anisotropic effective permeability reductions for acute intersection angles close to 0°, <strong>and</strong><br />

isotropic effective permeability for angles close to 90°. For relatively regular, evenly<br />

spaced patterns, the direction <strong>of</strong> maximum permeability bisects the acute angle <strong>of</strong><br />

intersection. For DB volume fractions <strong>of</strong> ~10% <strong>and</strong> acute intersection angles <strong>of</strong> ~80 o ,<br />

such as found in the Aztec, the cross-hatch pattern can produce up to a 1.5-order-<strong>of</strong>-<br />

magnitude reduction in bulk permeability <strong>and</strong> more than an order-<strong>of</strong>-magnitude<br />

permeability anisotropy.<br />

165


Though grossly systematic, anastomosing DB patterns are strongly irregular in detail<br />

<strong>and</strong> so their permeability impact can only be assessed using the numerical method. For<br />

this study, a single representative outcrop array from the Aztec was studied. The resulting<br />

permeability effects were substantially similar to those <strong>of</strong> a comparable, idealized parallel<br />

pattern, though somewhat muted. For the anastomosing pattern, DB connectivity along<br />

trend acts to reduce effective b<strong>and</strong>-parallel permeability, while lack <strong>of</strong> continuity across<br />

the pattern enhances b<strong>and</strong>-normal permeability. Nonetheless, for the representative<br />

pattern studied, permeability anisotropies <strong>of</strong> up to 1.3 orders <strong>of</strong> magnitude are possible,<br />

with bulk permeability reductions <strong>of</strong> more than an order <strong>of</strong> magnitude.<br />

8. Discussion <strong>and</strong> conclusions<br />

In numerically computing the effective permeability <strong>of</strong> deformation b<strong>and</strong> (DB) arrays<br />

in s<strong>and</strong>stone, we have demonstrated, their potential to induce order-<strong>of</strong>-magnitude bulk<br />

permeability reductions <strong>and</strong> anisotropies at scales relevant to reservoir production. But do<br />

such permeability reductions <strong>and</strong> anisotropies influence fluid flow in actual field<br />

practice? While it is beyond the scope <strong>of</strong> our current research effort to address this issue<br />

directly—by correlating DB occurrence with pump test or production data, for example—<br />

such an effort clearly is indicated. We encourage petroleum geologists to be vigilant for<br />

evidence <strong>of</strong> DBs in core, well bore logs <strong>and</strong> outcrop, <strong>and</strong> to investigate the connection<br />

between DB occurrence <strong>and</strong> otherwise unexplained departures from expected production<br />

performance.<br />

This being said, evidence does exist for the substantial impact on permeability <strong>and</strong><br />

fluid flow <strong>of</strong> DBs in s<strong>and</strong>stone. For example, in the Aztec patterns <strong>of</strong> diagenetic<br />

alteration fronts indicative <strong>of</strong> paleo fluid flow clearly demonstrate that DBs <strong>and</strong> DB<br />

arrays acted as baffles to flow at a variety <strong>of</strong> scales (Eichhubl et al., 2003; Taylor <strong>and</strong><br />

Pollard, 2000). In addition, direct production evidence for the influence <strong>of</strong> geological<br />

heterogeneities—including DBs—has been noted in the Aztec-equivalent Nugget<br />

s<strong>and</strong>stone <strong>of</strong> the Anschutz Ranch East Field in Wyoming, Idaho <strong>and</strong> Utah (Lewis, 1993).<br />

Production problems attributed to deformation b<strong>and</strong>s have also been reported for the<br />

Nubian S<strong>and</strong>stone in Egypt <strong>and</strong> for s<strong>and</strong>stone reservoirs located <strong>of</strong>fshore from Nigeria<br />

(Olsson et al., 2003).<br />

166


Nonetheless, DBs are commonly ignored in the practice <strong>of</strong> s<strong>and</strong>stone reservoir<br />

management, even though their potential order-<strong>of</strong>-magnitude permeability effects rival<br />

those routinely attributed to depositional heterogeneities—e.g. bedding <strong>and</strong> shale<br />

fraction. We suggest therefore that recognizing <strong>and</strong> characterizing systematic DB patterns<br />

in the subsurface, <strong>and</strong> then accounting accurately for their aggregate influence on bulk<br />

permeability with an upscaling process such as presented here, should become a regular<br />

component <strong>of</strong> reservoir simulation <strong>and</strong> production management in s<strong>and</strong>stone. Finally, we<br />

observe that, even with only limited DB data gleaned from scattered core studies,<br />

valuable adjustments to the underlying permeability structure <strong>of</strong> s<strong>and</strong>stone reservoirs can<br />

be approximated based on the results presented here for the characteristic DB patterns <strong>of</strong><br />

the Aztec s<strong>and</strong>stone.<br />

9. Acknowledgements<br />

The authors would like to thank Eric Flodin <strong>and</strong> Nick Davatzes for their assistance in<br />

the field, <strong>and</strong> Herve Jourde for his help in the field <strong>and</strong> with the numerical modeling. This<br />

work was supported by the U.S. Department <strong>of</strong> Energy, Office <strong>of</strong> Basic Energy Science,<br />

Geosciences Research Program under grant DE-FG03-94ER14462, awarded to David<br />

Pollard <strong>and</strong> Atilla Aydin.<br />

167


168


1. Abstract<br />

Chapter 7<br />

Flow <strong>and</strong> transport effects <strong>of</strong> compaction b<strong>and</strong>s in s<strong>and</strong>stone<br />

at scales relevant to aquifer <strong>and</strong> reservoir management<br />

Thin, tabular, low-porosity, low-permeability compaction b<strong>and</strong>s form pervasive, sub-<br />

parallel, anastomosing arrays that extend over square kilometers <strong>of</strong> exposure in the Aztec<br />

s<strong>and</strong>stone <strong>of</strong> southeastern Nevada—an exhumed analog for active aquifers <strong>and</strong> reservoirs.<br />

In order to examine the potential flow <strong>and</strong> transport effects <strong>of</strong> these b<strong>and</strong> arrays at scales<br />

relevant to production <strong>and</strong> management, we performed a suite <strong>of</strong> simulations using an<br />

innovative discrete-feature modeling technique to capture the exact pattern <strong>of</strong> compaction<br />

b<strong>and</strong>s mapped over some 150,000 m 2 <strong>of</strong> contiguous outcrop. Significant impacts related<br />

to the presence <strong>of</strong> the b<strong>and</strong>s <strong>and</strong> their dominant trend are apparent: the average pressure<br />

drop required to drive flow between wells exceeds that for b<strong>and</strong>-free s<strong>and</strong>stone by a<br />

factor <strong>of</strong> three, <strong>and</strong> is 10% to 50% higher across the b<strong>and</strong>s versus along them; reservoir<br />

production efficiency varies up to 56% for a typical five-spot well array, depending on its<br />

orientation relative to the dominant b<strong>and</strong> trend; <strong>and</strong> contaminant transport away from a<br />

point source within an aquifer tends to channel along the b<strong>and</strong>s, regardless <strong>of</strong> the regional<br />

gradient direction. We conclude that accounting for the flow effects <strong>of</strong> similar<br />

compaction-b<strong>and</strong> arrays would prove essential for optimal management <strong>of</strong> those<br />

s<strong>and</strong>stone aquifers <strong>and</strong> reservoirs in which they occur.<br />

2. Introduction<br />

The potential for deformation b<strong>and</strong> arrays at the outcrop scale to reduce bulk effective<br />

2-D permeability by more than an order <strong>of</strong> magnitude, <strong>and</strong> to induce significant (10x)<br />

permeability anisotropy, has been established (Sternl<strong>of</strong> et al., 2004). Their potential to<br />

cause asymmetric patterns <strong>of</strong> well drawdown has also been demonstrated (Matthai et al.,<br />

1998). However, despite growing recognition that deformation b<strong>and</strong> fabrics are common<br />

in both exhumed (e.g. Underhill <strong>and</strong> Woodcock, 1987; Edwards et al., 1993; Davis, 1998;<br />

Swierczewska <strong>and</strong> Tokarski, 1998; Cashman <strong>and</strong> Cashman, 2000; Shipton <strong>and</strong> Cowie,<br />

2001; Davatzes et al., 2005) <strong>and</strong> subsurface (e.g. Fossen <strong>and</strong> Hesthammer, 1998;<br />

Hesthammer et al., 2000; Hesthammer <strong>and</strong> Henden, 2000) s<strong>and</strong>stones, their ability to<br />

169


100 meters<br />

5 meters<br />

NV UT<br />

CA<br />

AZ<br />

Park Roa d<br />

Route 16 9<br />

0 km 5<br />

Map<br />

Detail<br />

Air Photo<br />

NEVADA<br />

ARIZON A<br />

10 km<br />

Valley<br />

<strong>of</strong> Fire<br />

State Park<br />

N<br />

LAKE<br />

MEAD<br />

115 o 30'<br />

Figure 7.1. Aerial photograph showing a largely bare <strong>and</strong> level outcrop area <strong>of</strong> Aztec<br />

s<strong>and</strong>stone in the Valley <strong>of</strong> Fire State Park <strong>of</strong> southeastern Nevada. The geographical<br />

location <strong>of</strong> the study area is indicated in the map (inset, upper right). A high-resolution<br />

digital scan <strong>of</strong> the negative made it possible to identify (photo inset lower left) <strong>and</strong> map<br />

(photo inset lower right) the patterns <strong>of</strong> individual, cm-thick compaction b<strong>and</strong>s cropping<br />

out as distinct fins over an area <strong>of</strong> more than 150,000 m 2 .<br />

170<br />

36 o 15'


exert appreciable fluid flow <strong>and</strong> transport effects at scales <strong>of</strong> practical relevance to<br />

resource management in aquifers <strong>and</strong> reservoirs has remained an open question.<br />

To begin addressing this issue, we present a series <strong>of</strong> flow simulation scenarios that<br />

honor the exact details <strong>of</strong> individual compaction b<strong>and</strong> arrays mapped over an outcrop area<br />

<strong>of</strong> some 150,000 m 2 (37 acres) in the Aztec s<strong>and</strong>stone <strong>of</strong> southeastern Nevada (Figure<br />

7.1). The simulations were conducted using a discrete-feature model in which the b<strong>and</strong>s<br />

are represented as one-dimensional elements assigned realistic petrophysical properties.<br />

This effort represents the application <strong>of</strong> state-<strong>of</strong>-the-art flow modeling to a data set <strong>of</strong><br />

unprecedented detail <strong>and</strong> scale, which was derived from a real geological system that<br />

serves as an exhumed analog for active s<strong>and</strong>stone aquifers <strong>and</strong> reservoirs. Analog<br />

outcrop-based permeability <strong>and</strong> flow-modeling studies, whether focused on gross<br />

sedimentological heterogeneities (e.g. White <strong>and</strong> Barton, 1999; Stephen <strong>and</strong> Dalrymple,<br />

2002) or on fine-scale <strong>structural</strong> features (e.g. Taylor et al., 1999; Sternl<strong>of</strong> et al., 2004),<br />

can provide valuable insights into the natural variability <strong>of</strong> inter-well flow <strong>and</strong> the<br />

balance <strong>of</strong> parameters governing optimal (or at least improved) well placement.<br />

The Aztec s<strong>and</strong>stone as exposed in the Valley <strong>of</strong> Fire State Park lies about 60 km<br />

northeast <strong>of</strong> Las Vegas (Figure 7.1, map inset). The Aztec is a fine to medium-grained,<br />

sub-arkosic æolian s<strong>and</strong>stone some 1,400 m thick, which was deposited as large<br />

amplitude cross beds in a back arc basin setting from Late Triassic to Middle Jurassic<br />

time (Bohannon, 1983; Marzolf, 1983). It exhibits a mean grain size <strong>of</strong> about 0.25 mm,<br />

porosity ranging from 15% to 25% <strong>and</strong> permeability ranging up to several Darcys<br />

(Antonellini <strong>and</strong> Aydin, 1994; Flodin et al., 2005). Extensive arrays <strong>of</strong> sub-parallel,<br />

anastomosing compaction b<strong>and</strong>s (CBs) pervade the upper 600 m <strong>of</strong> the Aztec exposed<br />

over an area <strong>of</strong> more than 10 km 2 . Apparently associated with tectonism <strong>of</strong> the<br />

Cretaceous Sevier orogeny, these b<strong>and</strong>s at high angle to bedding comprise the oldest<br />

<strong>structural</strong> fabric present, which subsequently was overprinted by shear b<strong>and</strong>s, joints,<br />

sheared joints <strong>and</strong> strike-slip faults (Hill, 1989; Taylor <strong>and</strong> Pollard, 2000; Eichhubl et al.,<br />

2004; Flodin <strong>and</strong> Aydin, 2004; Myers <strong>and</strong> Aydin, 2004; Sternl<strong>of</strong> et al., 2005).<br />

As observed in the Aztec <strong>and</strong> similar s<strong>and</strong>stones, CBs are discrete, tabular, bounded<br />

features <strong>of</strong> porosity-loss compaction accommodated by granular rearrangement, grain<br />

crushing <strong>and</strong> chemical diagenesis that exhibit little or no net shear <strong>and</strong> tend to weather<br />

171


compaction b<strong>and</strong> trend<br />

500µm<br />

Figure 7.2. Typical outcrop pattern <strong>of</strong> sub-parallel CB fins in the Aztec s<strong>and</strong>stone (top).<br />

Backscatter electron images from just outside (bottom left) <strong>and</strong> inside (bottom right) a<br />

CB—white is hematite, light gray is feldspar, medium gray is quartz, dark gray is clay,<br />

black is porosity. Porosity drops from >20% in the s<strong>and</strong>stone to 1,000 mD (s<strong>and</strong>stone) to


out in positive relief as distinct fins (Sternl<strong>of</strong> et al., 2005; Mollema <strong>and</strong> Antonellini, 1996)<br />

(Figure 7.2). Millimeters to centimeters thick <strong>and</strong> tens <strong>of</strong> meters in planar extent, CBs<br />

represent a kinematic end member <strong>of</strong> the suite <strong>of</strong> structures known collectively as<br />

deformation b<strong>and</strong>s, which also includes shear <strong>and</strong> dilation b<strong>and</strong>s (Aydin, 1978;<br />

Antonellini et al., 1994; Mollema <strong>and</strong> Antonellini, 1996; Du Bernard et al., 2002). The<br />

petrophysical changes accommodating mechanical compaction inside CBs, in conjunction<br />

with subsequent preferential cementation, can reduce porosity within the b<strong>and</strong>s to a few<br />

percent. Decreased porosity, <strong>and</strong> corresponding reductions in pore throat diameter <strong>and</strong><br />

connectivity, can in turn reduce saturated permeability within the b<strong>and</strong>s by 1 to 4 orders<br />

<strong>of</strong> magnitude relative to that <strong>of</strong> the surrounding host rock (Pittman, 1981; Freeman, 1990;<br />

Antonellini <strong>and</strong> Aydin, 1994; Crawford, 1998; Gibson, 1998; Taylor <strong>and</strong> Pollard, 2000;<br />

Sigda <strong>and</strong> Wilson, 2003).<br />

Permeability anisotropy due to depositional bedding in the Aztec <strong>and</strong> Navajo<br />

s<strong>and</strong>stones is generally less than a factor <strong>of</strong> two (Freeman, 1990; Antonellini <strong>and</strong> Aydin,<br />

1994; Flodin et al., 2005), minimal when compared to the reduced permeability<br />

represented by the CBs. There is no data to suggest that significant anisotropy related to<br />

bedding persists inside the b<strong>and</strong>s, <strong>and</strong> no reason to expect that the consequences <strong>of</strong> any<br />

such remnant anisotropy would be significant. In the analyses to follow, we therefore<br />

make the simplifying assumption <strong>of</strong> homogeneous, isotropic permeability for both the<br />

s<strong>and</strong>stone <strong>and</strong> the CBs.<br />

3. Mapping<br />

The fins <strong>of</strong> CBs cropping out in the Aztec s<strong>and</strong>stone are plainly visible in low-altitude<br />

aerial photographs. A high-resolution (3,000 dpi) digital scan made from one such<br />

st<strong>and</strong>ard (23 cm x 23 cm) aerial photo color negative reveals an extensive exposure <strong>of</strong><br />

nearly continuous Aztec outcrops at a scale <strong>of</strong> 1:2,400 (Figure 7.1). This scan captured a<br />

level <strong>of</strong> detail down to the grain <strong>of</strong> the negative, producing a digital image with a pixel<br />

dimension relative to outcrop <strong>of</strong> about 2 cm. Loading this image into Adobe Illustrator®<br />

<strong>and</strong> enlarging it, we were able to recognize <strong>and</strong> trace individual cm-thick CBs (or in many<br />

cases, closely space clusters <strong>of</strong> CBs) for tens to hundreds <strong>of</strong> meters (Figure 7.1, inset<br />

photos). This photo mapping effort was guided by our familiarity with the appearance <strong>of</strong><br />

CBs in outcrop <strong>and</strong> was checked extensively in the field.<br />

173


Cover<br />

Outcrop<br />

B<strong>and</strong><br />

N<br />

100 meters<br />

20 meters<br />

Figure 7.3. Compaction b<strong>and</strong> trace map representing b<strong>and</strong>s unequivocally identifiable as<br />

fins on the air photo (Figure 7.1). Gaps not due to sedimentary cover, fault damage or<br />

shadows represent outcrop areas where b<strong>and</strong>s are scarce, absent or lack identifiable fins.<br />

Much <strong>of</strong> the b<strong>and</strong>-trend variability apparent on the map is due to topography. The density<br />

<strong>of</strong> b<strong>and</strong>s captured in the map is approximately one third that visible in outcrop.<br />

174


To produce a consistent <strong>and</strong> representative CB map for flow modeling, we applied the<br />

simple criteria that only b<strong>and</strong> traces that could unequivocally be identified on the high-<br />

resolution scan were included on the map. In particular, no attempt was made to<br />

interpolate b<strong>and</strong> patterns through areas <strong>of</strong> poor exposure due to sediment cover or the<br />

damage caused by younger, mid-Tertiary strike-slip faulting. Nor was any attempt made<br />

to extrapolate b<strong>and</strong> patterns in order to increase the area <strong>of</strong> coverage. In this way, a<br />

consistent <strong>and</strong> conservative st<strong>and</strong>ard <strong>of</strong> objectivity was maintained while complications<br />

<strong>and</strong> uncertainties associated with pattern matching <strong>and</strong> infilling were avoided.<br />

The resulting 150,000 m 2 CB map (Figure 7.3) represents more than 2000 individual<br />

b<strong>and</strong> traces ranging from a few meters to more than 100 m in length, with an average<br />

spacing <strong>of</strong> about 1.5 to 2 m. Successive intersections between b<strong>and</strong>s occasionally form<br />

through-going trends hundreds <strong>of</strong> meters long. Average b<strong>and</strong> thickness measured in the<br />

field is about 1 cm while spacing averages about 0.5 to 0.75 m—roughly one third that<br />

captured on the map. Clustering <strong>of</strong> CBs (mm spacing) is also common, resulting in<br />

composite b<strong>and</strong>s up to 10 cm <strong>and</strong> more in total thickness. Field checking indicated that<br />

most gaps within the mapped array not attributable to sedimentary cover or fault damage<br />

do represent real interruptions in the continuity <strong>of</strong> the b<strong>and</strong> fabric, although the absence<br />

<strong>of</strong> distinct fins <strong>and</strong>/or poor image resolution was occasionally to blame. All areas <strong>of</strong><br />

b<strong>and</strong>s identified <strong>and</strong> mapped on the scan were found to exist in outcrop.<br />

4. Flow simulation model<br />

In the scenarios considered below, both single-phase <strong>and</strong> two-phase flow simulations<br />

are performed. The basic governing equation for incompressible single-phase flow in<br />

porous media is obtained by combining Darcy's law with conservation <strong>of</strong> mass:<br />

⎛ k ⎞<br />

∇ ⋅⎜<br />

∇p⎟<br />

+ Q = 0 , (1)<br />

⎝ µ ⎠<br />

where p is pressure <strong>and</strong> Q represents sources <strong>and</strong>/or sinks (e.g., production/injection<br />

wells). The fluid is described by its viscosity µ , while k represents the absolute<br />

(intrinsic) permeability <strong>of</strong> the rock. Consistent with core measurements for the s<strong>and</strong>stone<br />

systems considered here, we take k to be locally isotropic for both the compaction b<strong>and</strong>s<br />

175


(CBs) <strong>and</strong> the matrix rock in all simulations (although these permeabilities are locally<br />

isotropic, large-scale anisotropy will result from the CB orientation).<br />

Equation (1) can be extended to accommodate two-phase flow by considering Darcy's<br />

law <strong>and</strong> mass conservation for each phase separately:<br />

( φS<br />

)<br />

∂<br />

∂t<br />

( φS<br />

)<br />

∂<br />

∂t<br />

⎛ kk<br />

= ∇⋅<br />

⎜<br />

⎝<br />

∇<br />

⎞<br />

⎟<br />

⎠<br />

n rn<br />

⎜ pn<br />

⎟ + Qn<br />

µ n<br />

⎛ kk<br />

= ∇⋅<br />

⎜<br />

⎝<br />

∇<br />

⎞<br />

⎟<br />

⎠<br />

w rw<br />

⎜ pw<br />

⎟ + Q w<br />

µ w<br />

, (2)<br />

, (3)<br />

where subscripts n <strong>and</strong> w refer to nonwetting <strong>and</strong> wetting phases. Saturation (phase<br />

volume fraction) S is defined for each phase. The porosity <strong>of</strong> the rock is designated byφ<br />

,<br />

<strong>and</strong> in addition to the absolute permeability , a relative permeability, k , is defined for<br />

k r<br />

each phase. A capillary pressure (pc)<br />

relationship expresses the difference in pressure<br />

between the two phases as a function <strong>of</strong> S; i.e., pc( Sw)<br />

= pn<br />

− pw<br />

. An additional<br />

constraint requires Sn<br />

+ Sw<br />

= 1 .<br />

For all the flow scenarios, the governing equations are solved using a specialized<br />

finite-volume discretization technique, namely a discrete-feature model, developed by<br />

Karimi-Fard et al. (2004). Although originally designed to model flow in fractured porous<br />

media, where the discrete features provide high-permeability channels, the generality <strong>of</strong><br />

the method allows low-permeability features such as CBs to be treated equally well. The<br />

technique is designed for use with unstructured grids (in both two <strong>and</strong> three dimensions),<br />

which can accurately represent complex natural geometries such as the CB map (Figure<br />

7.3). A particular strength <strong>of</strong> the method is that it allows small-scale linear features such<br />

as CBs to be represented <strong>and</strong> modeled using control volumes that are <strong>of</strong> the same<br />

thickness as the feature, meaning that CBs need not be resolved in the transverse (thin)<br />

direction by the grid. This reduces the overall number <strong>of</strong> cells required <strong>and</strong> greatly<br />

simplifies the gridding procedure. In addition, inefficiencies caused by the proliferation <strong>of</strong><br />

small control volumes at CB intersections are mitigated using a special connectivity<br />

transformation, which improves numerical stability <strong>and</strong> allows for larger time steps in<br />

transport problems in some cases (Karimi-Fard et al., 2004).<br />

176


(c)<br />

(a)<br />

(b)<br />

l<br />

d lm<br />

A j<br />

m<br />

d m<br />

A j<br />

A j<br />

m<br />

b<br />

ll bb<br />

d lm<br />

l m<br />

Figure 7.4. Schematic diagrams depicting quantities relevant to the transmissibility<br />

calculations for (a) matrix to matrix flow, (b) matrix to compaction b<strong>and</strong> flow <strong>and</strong> (c)<br />

compaction b<strong>and</strong> to compaction b<strong>and</strong> flow. The b<strong>and</strong> thickness is exaggerated in (b) <strong>and</strong><br />

(c).<br />

177


The finite volume discretization for single-phase flow is derived by integrating<br />

Equation (1) over each control volume (designated Ω) with k assumed constant within the<br />

control volume. The flux q across the control volume boundary ∂Ω is given by:<br />

⎛ k ⎞<br />

q = − ⎜ ∇p⎟<br />

⋅n<br />

dS<br />

= u⋅<br />

n dS<br />

, (4)<br />

∫∂ Ω⎝<br />

µ ⎠ ∫∂Ω<br />

where n is the outward-pointing unit normal on the surface <strong>of</strong> ∂Ω <strong>and</strong> u is the Darcy<br />

velocity. For triangular control volumes, as are considered here, the sum <strong>of</strong> the fluxes<br />

through the three faces (qj, j =1,2,3) is balanced by the source term; i.e., q + q + q = −Qˆ<br />

,<br />

∫ Ω<br />

where Qˆ = Q dV<br />

. In a two-point flux approximation (TPFA), qj<br />

is expressed in terms <strong>of</strong><br />

the geometries, permeabilities, <strong>and</strong> cell-centered pressures <strong>of</strong> the two triangles that share<br />

edge j.<br />

With reference to Figure 7.4a, for flow between two matrix control volumes l <strong>and</strong> m,<br />

this gives<br />

Tj<br />

q j = ( pl<br />

− pm)<br />

, (5)<br />

µ<br />

where Tj is the transmissibility <strong>and</strong> pl <strong>and</strong> pm designate the pressures at the barycenters <strong>of</strong><br />

the two control volumes. Transmissibility is given by:<br />

k<br />

A<br />

j j<br />

T j = , (6)<br />

dlm<br />

where kj is in general the weighted harmonic average <strong>of</strong> kl <strong>and</strong> km (though here matrix<br />

permeability is taken as a constant), Aj is the area (length times thickness) <strong>of</strong> the shared<br />

interface, <strong>and</strong> dlm = |dlm| is the length <strong>of</strong> the line connecting the two cell centers. When dlm<br />

is not exactly orthogonal to the shared edge, we take appropriate projections, as described<br />

in Karimi-Fard et al. (2004).<br />

Analogous expressions for Tj can be developed for flow between matrix <strong>and</strong> CB<br />

control volumes <strong>and</strong> for flow between two CB control volumes. For the first case, with<br />

reference to Figure 7.4b, Tj is again given by Equation (6), but now kj is computed via the<br />

weighted harmonic average <strong>of</strong> the matrix <strong>and</strong> CB permeabilities (km <strong>and</strong> kb):<br />

178<br />

1<br />

2<br />

3


k<br />

2d<br />

m b<br />

j = , (7)<br />

lb<br />

2d<br />

m<br />

k<br />

b<br />

+<br />

+ l<br />

k<br />

m<br />

where lb is the CB thickness. In addition, dlm in Equation (6) is replaced by m b , the<br />

center to center distance (see Figure 7.4b). The presence <strong>of</strong> k<br />

d + l / 2<br />

b -1 in the denominator <strong>of</strong><br />

Equation (7) results in kj < km. It is in this way that the permeability-barrier effect <strong>of</strong> the<br />

CBs enters the matrix-CB transmissibility <strong>and</strong> thus the flow equations. For flow between<br />

two CB control volumes, Equation (6) is again applied, with quantities as defined in<br />

Figure 7.4c. We note that permeability anisotropy in the CBs (i.e., different permeability<br />

along <strong>and</strong> across the b<strong>and</strong>s) could be readily modeled within the current formulation,<br />

though this is not necessary here because CB permeability is assumed to be isotropic.<br />

Writing the discrete mass balance ( q + q + q = −Qˆ<br />

) for each cell, <strong>and</strong> expressing the qj<br />

1<br />

2<br />

via Equation (5), gives the finite volume representation for Equation (1). Solution <strong>of</strong> the<br />

resulting linear system provides the cell-center pressure unknowns.<br />

In the case <strong>of</strong> two-phase flow, finite volume discretization <strong>of</strong> Equations (2) <strong>and</strong> (3)<br />

entails mass conservation statements for each component (phase). The net flux out <strong>of</strong> the<br />

control volume in this case is balanced by both the source <strong>and</strong> accumulation terms. The<br />

flux <strong>of</strong> a particular phase (e.g., water) across edge j is now given by:<br />

w<br />

w w<br />

q = λ T ( p − p )<br />

(8)<br />

j<br />

w<br />

j<br />

l<br />

m<br />

w<br />

w<br />

where λw is the upstream-weighted phase mobility (λw = krw/µw) <strong>and</strong> p <strong>and</strong> p are the<br />

water pressures in blocks l <strong>and</strong> m. The transmissibility Tj<br />

is the same for both phases <strong>and</strong><br />

is again given by Equation (6). Expressing the discrete mass balances for each component<br />

using Equation (8) <strong>and</strong> introducing time discretizations for the accumulation terms<br />

provides the finite volume representation. Pressure <strong>and</strong> saturation for each cell at each<br />

time step can then be computed.<br />

3<br />

The two-point flux approximation described above is applied for all connections. This<br />

is strictly valid only when the grid is “orthogonal,” which in this context means that dlm<br />

(the line connecting the two cell centers) is orthogonal to the shared interface.<br />

Nonorthogonality (or full-tensor effects, not an issue here since permeability is taken as<br />

179<br />

l<br />

m


isotropic) can lead to numerical discretization errors when TPFA is applied. Such errors<br />

can be eliminated through use <strong>of</strong> multipoint flux approximations (see Karimi-Fard et al.,<br />

2004, for discussion <strong>of</strong> this issue). Errors resulting from TPFA are, however, reduced<br />

when the grid is formed via a Delaunay triangulation. Given a set <strong>of</strong> points (vertices), a<br />

Delaunay triangulation (Shewchuk, 1996) satisfies a Max-Min property, which means<br />

that, for all possible triangulations <strong>of</strong> the set <strong>of</strong> points, the Delaunay procedure maximizes<br />

the minimum angle <strong>of</strong> the triangulation. This tends to avoid very small or very large<br />

angles, which lead to large discretization errors when TPFA is applied, <strong>and</strong> gives<br />

triangles that are more nearly equilateral. For a grid comprised <strong>of</strong> equilateral triangles<br />

(<strong>and</strong> isotropic permeability), a two-point flux approximation is strictly valid.<br />

The overall solution procedure is compatible with general-purpose flow simulators<br />

that apply connection lists for the definition <strong>of</strong> cell to cell connectivity. The<br />

transmissibilities defined above along with the associated cell volumes are input to the<br />

<strong>Stanford</strong> General Purpose Research Simulator (GPRS) for flow simulation. See Cao<br />

(2002) for a detailed description <strong>of</strong> this simulator.<br />

The CB map (Figure 7.3) was gridded using a st<strong>and</strong>ard Delaunay triangulation<br />

(Shewchuk, 1996) technique. The resulting 2-D grid (Figure 7.5) contains 146,682<br />

triangular <strong>and</strong> segment control volumes representing 1,948 discrete b<strong>and</strong> traces <strong>and</strong> the<br />

surrounding s<strong>and</strong>stone matrix. To avoid spurious boundary effects, an oval, no-flow<br />

exterior boundary was established well outside the mapped pattern, while all flow effects<br />

are evaluated inside the pattern. We note that these simulations are relatively dem<strong>and</strong>ing<br />

computationally, <strong>and</strong> that it is not currently practical to perform such runs on full-field<br />

models. However, by considering sectors, as is done here, it is possible to quantify the<br />

impact <strong>of</strong> CBs on large-scale flow <strong>and</strong> transport.<br />

In all <strong>of</strong> the simulations presented below, we specify a uniform b<strong>and</strong> thickness (lb) <strong>of</strong><br />

3 cm, b<strong>and</strong> porosity <strong>of</strong> 10%, homogeneous <strong>and</strong> isotropic internal b<strong>and</strong> permeability <strong>of</strong> 1.5<br />

mD, s<strong>and</strong>stone matrix porosity <strong>of</strong> 25%, <strong>and</strong> homogeneous, isotropic matrix permeability<br />

<strong>of</strong> 1.5 Darcys. The stipulation <strong>of</strong> 3-cm b<strong>and</strong> thickness approximately accounts for the<br />

under-representation <strong>of</strong> b<strong>and</strong> density in the final map as discussed above. The<br />

b<strong>and</strong>/matrix permeability ratio <strong>of</strong> 10 -3 falls in the middle <strong>of</strong> the range reported in the<br />

literature (Sternl<strong>of</strong> et al., 2004).<br />

180


150 meters<br />

40 meters<br />

Figure 7.5. Unstructured Delaunay triangular discretization <strong>of</strong> the compaction b<strong>and</strong> map<br />

shown in Figure 7.3, with an oval no-flow boundary added well outside the b<strong>and</strong> pattern<br />

to delimit the model reservoir/aquifer (left). In this discrete-feature representation, the<br />

exact trace <strong>of</strong> each b<strong>and</strong> is captured as a series <strong>of</strong> linear segments defined by the shared<br />

sides <strong>of</strong> adjacent matrix control volumes (bold, segmented lines in the grid enlargement to<br />

the right).<br />

181


5. Simulations<br />

Three sets <strong>of</strong> 2-D simulations were performed. The first set prescribes single-phase,<br />

incompressible well-to-well flow for a series <strong>of</strong> well pairs to examine how the b<strong>and</strong><br />

pattern affects the pressure drop required to maintain a given flow rate. The second set <strong>of</strong><br />

simulations prescribes two-phase, incompressible flow for a st<strong>and</strong>ard five-spot petroleum<br />

reservoir production scenario with central injector to examine how the orientation <strong>of</strong> the<br />

well array relative to the b<strong>and</strong> pattern influences production efficiency. For this<br />

production scenario, the effects <strong>of</strong> relative permeability are considered with water as the<br />

wetting phase <strong>and</strong> oil as the nonwetting phase. The third set <strong>of</strong> simulations prescribes<br />

incompressible contaminant transport (mobility ratio = 1) for a point-source release to<br />

examine the relationship <strong>of</strong> regional gradient direction to b<strong>and</strong> pattern in determining the<br />

shape <strong>and</strong> distribution <strong>of</strong> the resulting plume.<br />

5.1. Well-to-well flow<br />

In order to characterize the bulk effects <strong>of</strong> the CB array as mapped, we first performed<br />

a series <strong>of</strong> 210 single-phase, incompressible well-to-well flow simulations for a fixed<br />

flow rate using injector-producer well pairs spaced 25, 50, 75, 100, 125, 150 <strong>and</strong> 200 m<br />

apart. At 15 midpoint locations scattered within the b<strong>and</strong> pattern, two simulations were<br />

performed for each well-spacing—one oriented generally parallel to (along) the dominant<br />

b<strong>and</strong> trend <strong>and</strong> one oriented generally normal to (across) it (Figure 7.6). For every<br />

simulation, the pressure drop measured between the wells in the presence <strong>of</strong> the b<strong>and</strong>s<br />

was normalized by the corresponding CB-free (homogeneous, isotropic s<strong>and</strong>stone matrix)<br />

value. The resulting normalized pressure drop data (∆P) are presented graphically in<br />

Figure 7.7.<br />

5.1.1. Results<br />

The most basic observation is that the pressure drop required to drive flow in the<br />

presence <strong>of</strong> the b<strong>and</strong>s always exceeds that for the CB-free case (Figure 7.7a). With a<br />

range from 1.7 to 4.2 <strong>and</strong> a mean <strong>of</strong> about 2.9, there is a slight trend toward increasing ∆P<br />

with decreasing well spacing. Also, the pressure drop required to drive flow across the<br />

dominant b<strong>and</strong> trend (∆Pn) generally exceeds that required to drive flow along it (∆Pp) by<br />

an average <strong>of</strong> about 25% (Figure 7.7b). In other words, a modest degree <strong>of</strong><br />

182


150 meters<br />

I<br />

I<br />

P<br />

P<br />

Well-pair midpoint<br />

I Injection well<br />

P Production well<br />

Figure 7.6. Fifteen locations within the compaction b<strong>and</strong> pattern around which well-towell<br />

flow simulations were performed—one set roughly normal to the dominant b<strong>and</strong><br />

trend <strong>and</strong> one roughly parallel to it—for spacings <strong>of</strong> 25, 50, 75, 100, 125, 150 <strong>and</strong> 200 m.<br />

183


Normalized ∆P<br />

∆P ratio (n/p)<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

(a) Well spacing (meters)<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

(b)<br />

n<br />

p<br />

n<br />

p<br />

n<br />

p<br />

n<br />

p<br />

25 50 75 100 125 150 175 200<br />

25 50 75 100 125 150 175 200<br />

Well spacing (meters)<br />

Figure 7.7. Pressure drop (∆P) results for the 210 well-to-well flow simulations shown as<br />

boxplots, each representing the 15 well-pair midpoint locations shown in Figure 7.6. The<br />

top, middle <strong>and</strong> bottom lines <strong>of</strong> each box give the upper quartile, median <strong>and</strong> lower<br />

quartile values, respectively, for the corresponding set <strong>of</strong> data; the whiskers extend to the<br />

maximum <strong>and</strong> minimum values. (a) Boxplot pairs representing ∆P roughly normal (n)<br />

<strong>and</strong> parallel (p) to the dominant b<strong>and</strong> trend as a function <strong>of</strong> well spacing (all values<br />

normalized by the corresponding homogeneous, isotropic, b<strong>and</strong>-free result). (b) Boxplots<br />

representing the ∆P ratio (n/p), or anisotropy, at each well-pair midpoint location as a<br />

function <strong>of</strong> spacing.<br />

184<br />

n<br />

p<br />

n<br />

p<br />

n p


large-scale anisotropy is associated with the presence <strong>of</strong> the b<strong>and</strong>s <strong>and</strong>, except for a drop<br />

to about 10% at 200 m, this ∆P ratio (∆Pn/∆Pp) appears to be relatively insensitive to well<br />

spacing.<br />

5.1.2. Discussion<br />

It is <strong>of</strong> interest to compare the results illustrated in Figure 7.7 with those <strong>of</strong> Sternl<strong>of</strong> et<br />

al. (2004), who found that the anastomosing pattern <strong>of</strong> CBs characteristic <strong>of</strong> the Aztec<br />

s<strong>and</strong>stone at the outcrop scale can reduce bulk effective 2-D permeability by about an<br />

order <strong>of</strong> magnitude, while inducing a similar degree <strong>of</strong> anisotropy with minimum<br />

permeability directed normal to the dominant b<strong>and</strong> trend. The greater magnitude <strong>of</strong> the<br />

effects detected by Sternl<strong>of</strong> et al. (2004) can largely be attributed to the relatively small<br />

scale <strong>of</strong> their modeling effort (~10 m), <strong>and</strong> their application <strong>of</strong> local boundary conditions<br />

to force flow through the b<strong>and</strong> pattern in prescribed directions. Our simulations represent<br />

a generalization <strong>of</strong> this earlier effort, ins<strong>of</strong>ar as larger scales defined by realistic well<br />

spacings are considered, while remotely applied (global) boundary conditions do not<br />

constrain local flow pathways.<br />

It is no surprise that ∆P decreases with increasing spacing between wells. As spacing<br />

increases, so does the number <strong>of</strong> minimally obstructed flow pathways through the<br />

intervening b<strong>and</strong> pattern, even as the average length <strong>of</strong> those pathways relative to the<br />

straight-line distance between the wells decreases. The relatively modest ∆P ratio, or<br />

anisotropy, can be understood in light <strong>of</strong> the sinuous, anastomosing nature <strong>of</strong> the CB<br />

pattern, which ensures that any straight-line path between two wells will intersect<br />

multiple b<strong>and</strong>s regardless <strong>of</strong> its orientation relative to the dominant (mean) b<strong>and</strong> trend.<br />

Of practical importance to the hydrogeologist or petroleum engineer is the possibility<br />

that unexpectedly high pump-test pressures in what otherwise appears to be high-<br />

permeability s<strong>and</strong>stone might indicate the presence <strong>of</strong> CBs. In the absence <strong>of</strong><br />

corroborating core or analog outcrop data, <strong>and</strong> given the current lack <strong>of</strong> down-hole<br />

geophysical techniques for remote CB detection, the relatively modest ∆P ratio<br />

(anisotropy) induced by a b<strong>and</strong> pattern could be used to both confirm its presence <strong>and</strong><br />

determine its dominant geometry.<br />

185


P<br />

P<br />

P<br />

P<br />

I<br />

I<br />

P<br />

P<br />

Case 1<br />

P<br />

P<br />

Case 2<br />

I Injection well<br />

P Production well<br />

150 meters<br />

Figure 7.8. Well configurations used for the five-spot production scenarios—square<br />

pattern 150 m on a side with central injector <strong>and</strong> producers at the corners. In Case 1, one<br />

diagonal <strong>of</strong> the five-spot is aligned with the dominant b<strong>and</strong> trend (left). In Case 2, both<br />

diagonals are oriented obliquely (~45°) to the b<strong>and</strong>s.<br />

186


5.2. Reservoir production<br />

A st<strong>and</strong>ard five-spot production-well configuration—square, with central injector <strong>and</strong><br />

producers at the corners—was used to simulate two-phase (oil-water) incompressible<br />

flow within the b<strong>and</strong> pattern. Two case scenarios were modeled using the same 5.5-acre<br />

pattern (150 m on a side)—the only difference being a 45° rotation <strong>of</strong> the production-<br />

wells about a fixed injector location (Figure 7.8). In Case 1, the straight-line path from<br />

injector to two <strong>of</strong> the producers runs generally along the dominant b<strong>and</strong> trend, while the<br />

path to the two other producers runs generally across it. In Case 2, the straight-line paths<br />

from injector to all four producers run oblique to the dominant b<strong>and</strong> trend. For both cases,<br />

we specified complete initial saturation <strong>of</strong> the pore volume with oil, exact balance at all<br />

times between the total volume <strong>of</strong> water injected <strong>and</strong> the total volume <strong>of</strong> fluid produced,<br />

<strong>and</strong> equal pressure at all four production wells. A total injection/pumping rate <strong>of</strong> 15.9<br />

m 3 /day was used <strong>and</strong> both simulations were run for 250 days, yielding a total pore volume<br />

injected (PVI) <strong>of</strong> 0.15 (i.e. 15% <strong>of</strong> the total model pore volume). Relative permeability<br />

was accounted for using representative data for light crude oil (viscosity 2 cp) as the<br />

nonwetting phase <strong>and</strong> formation water (viscosity 1 cp) as the wetting phase in granular<br />

media similar to the Aztec s<strong>and</strong>stone (Table 1). Capillary pressure is neglected in these<br />

simulations, as attempting to include it produced computational problems. The efficient<br />

h<strong>and</strong>ling <strong>of</strong> strong capillary effects may require the modification <strong>of</strong> some <strong>of</strong> our<br />

numerical treatments. The simulation results are presented in Figures 7.9 <strong>and</strong> 7.10.<br />

5.2.1. Results<br />

Figure 7.9 shows equivalent saturation-map snapshots for the two simulations at<br />

approximately 0.03, 0.09 <strong>and</strong> 0.15 PVI. In both cases, the elliptical patterns <strong>of</strong> water<br />

infiltration that develop suggest a strong tendency toward preferred transport along the<br />

dominant b<strong>and</strong> trend. Judging from the aspect ratios <strong>of</strong> the infiltration patterns, the<br />

transport rate along the b<strong>and</strong> trend exceeds that across it by at least a factor <strong>of</strong> two. This<br />

effect results in early breakthrough <strong>of</strong> water to the production wells oriented along the<br />

b<strong>and</strong> trend from the injector in Case 1 at < 0.03 PVI (Figure 7.9a). For Case 2, significant<br />

breakthrough does not occur until 0.09 PVI (Figure 7.9b). These diverging results are<br />

clearly reflected at 0.15 PVI (Figure 7.9c), when the substantially greater area <strong>of</strong> water<br />

infiltration for Case 2 illustrates that 56% more oil has been produced than in Case 1.<br />

187


Case 1<br />

Case 1<br />

Case 1<br />

(a)<br />

(b)<br />

(c)<br />

Injection well<br />

Production well<br />

200 meters<br />

188<br />

Case 2<br />

Case 2<br />

Case 2


Figure 7.9. (facing page) Saturation-map snapshots for the five-spot scenarios shown in<br />

Figure 7.8: (a) at 0.03 <strong>of</strong> the total pore volume injected (PVI); (b) at 0.09 PVI; <strong>and</strong> (c) at<br />

0.15 PVI. Injected water shows as dark gray, with shades <strong>of</strong> gray denoting a two-phase<br />

mixture with oil. The pattern <strong>of</strong> water infiltration is initially similar for the two cases (a),<br />

but diverges as early breakthrough <strong>of</strong> water occurs at the production wells located along<br />

the dominant b<strong>and</strong> trend from the injector for Case 1 (b), culminating in a larger <strong>and</strong> more<br />

efficient water sweep for Case 2 (c).<br />

Pore volume produced (oil)<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

0 0.05 0.1<br />

Case 1<br />

Case 2<br />

Ideal<br />

0.15<br />

Pore volume injected (water)<br />

Figure 7.10. Oil production for the two scenarios as a function <strong>of</strong> the pore volume<br />

injected (PVI) <strong>of</strong> water. The solid line denotes the ideal case where the volume <strong>of</strong> oil<br />

produced equals the volume <strong>of</strong> water injected. Production efficiency drops more quickly<br />

for Case 1 than for Case 2. At PVI = 0.15, the volume <strong>of</strong> oil produced for Case 2 exceeds<br />

that for Case 1 by 56%.<br />

Table 7.1. Relative permeability data for reservoir production simulations<br />

Fraction<br />

S<strong>and</strong>stone Matrix Compaction B<strong>and</strong>s<br />

Water Saturation Rel. Perm. Water Rel. Perm. Oil Rel. Perm. Water Rel. Perm. Oil<br />

0.0 0.000 0.850 0.000 0.850<br />

0.2 0.020 0.544 0.010 0.544<br />

0.4 0.080 0.306 0.040 0.306<br />

0.6 0.180 0.136 0.090 0.136<br />

0.8 0.320 0.034 0.160 0.034<br />

1.0 0.500 0.000 0.250 0.000<br />

189


5.2.2. Discussion<br />

The elliptical patterns <strong>of</strong> water infiltration shown in Figure 7.9 demonstrate the<br />

dominant impact <strong>of</strong> the CB pattern on the gross direction <strong>of</strong> fluid transport. In fact, the<br />

saturation maps for the two cases are at first strikingly similar (Figure 7.9a), indicating<br />

that the production wells exert a minor influence on transport when compared to<br />

channeling <strong>of</strong> flow along the dominant b<strong>and</strong> trend. These directional transport effects are<br />

more pronounced than the relatively modest ∆P ratios presented in Figure 7.7 would<br />

suggest.<br />

The 56% increase in production efficiency gained in Case 2 for a 45° rotation <strong>of</strong> the<br />

five-spot well configuration is a significant (but not optimized) result that could vary<br />

substantially within the CB pattern <strong>and</strong> for different production configurations. In<br />

addition, the scale <strong>of</strong> our five-spot, as constrained by the available mapped area, is<br />

smaller than would typically be installed in a real reservoir production situation.<br />

Nonetheless, the petroleum engineer can glean two valuable insights from our results—<br />

that the presence <strong>of</strong> CB arrays can strongly affect production efficiency in reservoirs, but<br />

also that relatively simple adjustments in well placement can be made to mitigate <strong>and</strong>/or<br />

harness those effects.<br />

5.3. Contaminant transport<br />

In order to examine how the CB array would affect contaminant transport <strong>and</strong><br />

influence the evolution <strong>of</strong> a point-source plume, we established a fixed leak location for<br />

each <strong>of</strong> three regional pressure gradient directions: parallel to the dominant b<strong>and</strong> trend,<br />

oblique (45°) to the trend, <strong>and</strong> normal to the trend (Figure 7.11). We considered<br />

incompressible contaminant transport (mobility ratio = 1), <strong>and</strong> neglected physical<br />

dispersion. A consistent regional gradient <strong>of</strong> 0.23 Pa/m (enough to yield an average flow<br />

rate <strong>of</strong> approximately 7.7 m/year in the absence <strong>of</strong> CBs) was established in each <strong>of</strong> the<br />

three directions by specifying pressures for an array <strong>of</strong> eight wells distributed around the<br />

perimeter <strong>of</strong> the model domain (just inside the oval no-flow boundary). Each simulation<br />

represents 30 years at a constant leak rate <strong>of</strong> 0.08 m 3 /day. For each scenario, a<br />

corresponding CB-free control simulation was conducted. The results <strong>of</strong> all the<br />

simulations are presented in Figures 7.12, 7.13 <strong>and</strong> 7.14.<br />

190


N<br />

Leak<br />

150 meters<br />

P<br />

O<br />

Figure 7.11. Configuration <strong>of</strong> contaminant leak location with three directions <strong>of</strong> equal<br />

regional pressure gradient used in examining the impact <strong>of</strong> the compaction b<strong>and</strong> pattern<br />

on plume migration (P=parallel to the dominant b<strong>and</strong> trend, O=oblique to the trend,<br />

N=normal to the trend).<br />

191<br />

O<br />

P<br />

N


5.3.1. Results<br />

With the regional gradient directed roughly parallel to the dominant b<strong>and</strong> trend, the<br />

resulting contaminant plume is narrower <strong>and</strong> longer than would be the case in the absence<br />

<strong>of</strong> the CB array (Figure 7.12a) <strong>and</strong> extends some 60 m up-gradient from the leak location.<br />

The long axis <strong>of</strong> the two plumes (with <strong>and</strong> without CBs) remains coincident, however,<br />

<strong>and</strong> parallel to the b<strong>and</strong> trend. Looking in detail, the plume has clearly channeled along<br />

the b<strong>and</strong> array <strong>and</strong> been constrained from spreading laterally (Figure 7.12b).<br />

With the regional gradient directed obliquely to the dominant b<strong>and</strong> trend, the shape<br />

<strong>and</strong> orientation <strong>of</strong> the resulting plume departs substantially from that <strong>of</strong> the CB-free<br />

control case, remaining constrained by <strong>and</strong> aligned with the b<strong>and</strong> array (Figure 7.13a).<br />

The plume is, however, wider <strong>and</strong> shorter than in the parallel scenario, extends 80 m<br />

obliquely up-gradient from the leak (along the b<strong>and</strong> trend), <strong>and</strong>, where gaps in the CB<br />

pattern allow, turns into the regional down-gradient direction (Figure 7.13b).<br />

With the regional gradient directed normal to the dominant b<strong>and</strong> trend, the resulting<br />

plume bears scant resemblance to the CB-free control case, remaining distinctly elongated<br />

along the b<strong>and</strong> array (Figure 7.14a). The plume is shorter <strong>and</strong> broader again than in the<br />

oblique scenario, <strong>and</strong> turns strongly into the regional down-gradient direction at a major<br />

gap in the b<strong>and</strong> pattern (Figure 7.14b, upper left). Finally, the trailing edge <strong>of</strong> the plume is<br />

essentially straight, having pushed uniformly almost 20 m in the regional up-gradient<br />

direction.<br />

5.3.2. Discussion<br />

As with the reservoir-production simulations above, the plume evolution results<br />

reveal a dominating influence on fluid transport by the CB pattern, regardless <strong>of</strong> the<br />

regional gradient direction. Limited sensitivity testing also indicated that realistic<br />

variations in the magnitudes <strong>of</strong> the regional gradient <strong>and</strong>/or contaminant leak rate do not<br />

substantially affect the results. For the hydrogeologist designing a pump-<strong>and</strong>-treat<br />

remediation system therefore, recognizing <strong>and</strong> accounting for the effects <strong>of</strong> the CBs<br />

would prove crucial, particularly with 100% contaminant capture as the goal.<br />

For example, the st<strong>and</strong>ard cleanup approach for a leaking underground storage tank is<br />

to install the minimum number <strong>of</strong> wells that can be spaced immediately down gradient<br />

from the plume front to provide complete capture, assuming radial drawdown at an<br />

192


(a)<br />

(b)<br />

Leak<br />

200 meters<br />

Regional<br />

gradient<br />

No CBs CBs<br />

100 meters<br />

Figure 7.12. Contaminant plumes (dark gray) after 30 years with the regional gradient<br />

directed parallel to the dominant compaction b<strong>and</strong> trend (scenario P from Figure 7.11).<br />

(a) Plume distribution in the absence <strong>of</strong> b<strong>and</strong>s (left) <strong>and</strong> with b<strong>and</strong>s present (right). (b)<br />

Detail <strong>of</strong> the b<strong>and</strong>-impacted plume (shades <strong>of</strong> gray denote two-phase mixing).<br />

193


(a)<br />

(b)<br />

Leak<br />

200 meters<br />

Regional<br />

gradient<br />

No CBs CBs<br />

100 meters<br />

Figure 7.13. Contaminant plume (dark gray) after 30 years with the regional gradient<br />

directed obliquely to the dominant compaction b<strong>and</strong> trend (scenario O from Figure 7.11).<br />

(a) Plume distribution in the absence <strong>of</strong> b<strong>and</strong>s (left) <strong>and</strong> with b<strong>and</strong>s present (right). (b)<br />

Detail <strong>of</strong> the b<strong>and</strong>-impacted plume (shades <strong>of</strong> gray denote two-phase mixing).<br />

194


(a)<br />

(b)<br />

Leak<br />

200 meters<br />

No CBs<br />

Regional<br />

gradient<br />

CBs<br />

100 meters<br />

Figure 7.14. Contaminant plume (dark gray) after 30 years with the regional gradient<br />

directed normal to the dominant compaction b<strong>and</strong> trend (scenario N from Figure 7.11).<br />

(a) Plume distribution in the absence <strong>of</strong> b<strong>and</strong>s (left) <strong>and</strong> with b<strong>and</strong>s present (right). (b)<br />

Detail <strong>of</strong> the b<strong>and</strong>-impacted plume (shades <strong>of</strong> gray denote two-phase mixing).<br />

195


anticipated sustainable pumping rate based on the results <strong>of</strong> simple slug tests <strong>and</strong> the<br />

assumption <strong>of</strong> homogeneous, isotropic permeability. Given any <strong>of</strong> the CB scenarios<br />

presented in Figures 7.12, 7.13 <strong>and</strong> 7.14, this approach would lead to misplaced wells,<br />

both in terms <strong>of</strong> their location relative to the effective contaminant transport direction <strong>and</strong><br />

the spacing needed for complete capture given elliptical drawdown patterns (Matthai et<br />

al., 1998). Failure to recognize the transport-distorting effects <strong>of</strong> CBs would also<br />

complicate efforts to locate the unknown source <strong>of</strong> a contaminant plume, <strong>and</strong>/or identify<br />

its extent based on limited well data.<br />

6. General discussion<br />

While the model simulation scenarios presented above in no way constitute a<br />

complete treatment <strong>of</strong> the flow <strong>and</strong> transport effects <strong>of</strong> CBs in the Aztec s<strong>and</strong>stone, the<br />

results do illustrate a clear potential for substantial impacts. Given the broad practical<br />

implications <strong>of</strong> these findings, a few key interpretations <strong>and</strong> assumptions underlying this<br />

work warrant further comment.<br />

Firstly, as acknowledged in the section on mapping, limitations <strong>of</strong> outcrop exposure,<br />

image resolution <strong>and</strong> map scale prevent the final CB map (Figure 7.3) from capturing the<br />

full abundance <strong>and</strong> detail <strong>of</strong> the b<strong>and</strong> array present in the field. The map does, however,<br />

accurately represent the relative abundance, distribution, orientation <strong>and</strong> connectivity <strong>of</strong><br />

the exposed pattern <strong>of</strong> oldest CBs that dominates the area <strong>and</strong>, as such, is adequate to the<br />

purpose <strong>of</strong> modeling flow <strong>and</strong> transport effects at relatively coarse scales <strong>of</strong> practical<br />

interest. In fact, ins<strong>of</strong>ar as the true abundance, distribution <strong>and</strong> connectivity <strong>of</strong> the b<strong>and</strong>s<br />

are under-represented in the map, the substantial flow <strong>and</strong> transport effects realized here<br />

can reasonably be considered as conservative, low-end estimates.<br />

Also, our 2-D approach to an undeniably 3-D problem is dictated by the nature <strong>of</strong> the<br />

outcrop exposure, which presents a roughly horizontal slice through the CB array to<br />

reveal the detailed pattern as mapped. Given accurate 3-D data <strong>and</strong> sufficient computing<br />

power, the full spatial effects <strong>of</strong> the entire CB array could be modeled. Nonetheless,<br />

abundant field observations throughout the upper 600 m <strong>of</strong> the Aztec s<strong>and</strong>stone,<br />

particularly in areas <strong>of</strong> topographic relief, attest to the vertical persistence <strong>of</strong> the overall<br />

CB array. We suggest therefore that, to a reasonable first approximation, the<br />

anastomosing pattern <strong>of</strong> b<strong>and</strong>s in Figure 7.3 can be assumed to extend hundreds <strong>of</strong> meters<br />

196


vertically such that both the magnitude <strong>and</strong> directionality <strong>of</strong> the flow <strong>and</strong> transport effects<br />

realized in 2-D would substantially persist into 3-D.<br />

While it might be interesting <strong>and</strong> useful to assess whether current dispersion theory<br />

could capture the gross plume behavior exhibited in Figures 7.12-7.14, it is unlikely that<br />

this approach would provide quantitatively accurate results. Even for idealized (Gaussian,<br />

log-normal, etc.) permeability descriptions, it has been demonstrated that, as log<br />

permeability variance increases, so does the gap between simulation results <strong>and</strong><br />

theoretical predictions (Naff et al., 1998; Dentz et al., 2002). It may nonetheless be<br />

possible to describe the observed plume behavior qualitatively, or even semi-<br />

quantitatively using effective dispersivity methods.<br />

Accounting for capillary pressure effects in our two-phase simulations would, <strong>of</strong><br />

course, also be desirable. Although the modeling methods used in this study can in theory<br />

h<strong>and</strong>le these effects, as stated earlier, attempting to include them in the five-spot<br />

production simulations led to significant degradation in computational efficiency. We<br />

believe that this occurs as a result <strong>of</strong> the simulator attempting to resolve the very short<br />

capillary time scales associated with the b<strong>and</strong>s. Further investigation, testing <strong>and</strong> possible<br />

modification <strong>of</strong> the numerical methods will be required to address this issue adequately. It<br />

is important to note, however, that in many practical cases viscous (convective) effects<br />

dominate capillary effects, so the neglect <strong>of</strong> capillary pressure is physically reasonable in<br />

such systems.<br />

Although not addressed in this paper, coarse-scale descriptions (i.e. upscaled models)<br />

for detailed CB systems such as considered here would comprise the actual input<br />

parameters to real-world aquifer- <strong>and</strong> reservoir-scale simulations. We note, however, that<br />

coarse-scale flow attributes such as permeability <strong>and</strong> transmissibility might not be<br />

independent <strong>of</strong> the flow process under consideration, but rather could depend on specified<br />

parameters such as boundary conditions, well locations <strong>and</strong> flow rates. This sort <strong>of</strong><br />

process-dependent behavior has previously been observed in highly heterogeneous<br />

systems with permeability features that are <strong>of</strong> a length scale comparable to the system<br />

size or well spacing, as is the case here (Chen et al., 2003). For such systems, accurate<br />

upscaling may require the incorporation <strong>of</strong> global flow information, which can<br />

significantly complicate the computations.<br />

197


Finally, there is the issue <strong>of</strong> forecasting where volumetrically extensive compaction<br />

<strong>and</strong> deformation b<strong>and</strong> arrays are most likely to occur. This is an active area <strong>of</strong> research<br />

that to date has emphasized discrete deformation b<strong>and</strong> style faults in s<strong>and</strong>stone (e.g.<br />

Aydin <strong>and</strong> Johnson, 1983, Shipton <strong>and</strong> Cowie, 2001; Davatzes et al., 2005) over the<br />

distributed style <strong>of</strong> tectonic fabric formed by CBs in the Aztec (Sternl<strong>of</strong> et al., 2005).<br />

Much <strong>of</strong> the research into both types <strong>of</strong> b<strong>and</strong>s has focused on their occurrence in<br />

Mesozoic æolian s<strong>and</strong>stones <strong>of</strong> the southwestern U.S., although deformation b<strong>and</strong>s are<br />

increasingly being identified in a wide variety <strong>of</strong> porous, granular geologic materials<br />

ranging from nonwelded ignimbrites (Wilson et al., 2003) to unlithified Holocence beach<br />

s<strong>and</strong>s (Cashman <strong>and</strong> Cashman, 2000). Compaction b<strong>and</strong>s, however, have yet to be<br />

positively identified outside <strong>of</strong> the Navajo (Mollema <strong>and</strong> Antonellini, 1996) <strong>and</strong> Aztec<br />

(Sternl<strong>of</strong> et al., 2005) s<strong>and</strong>stones. As <strong>of</strong> this writing therefore, the most that can be said is<br />

that CB arrays such as exposed in the Aztec might reasonably be suspected in any high-<br />

porosity s<strong>and</strong>stone that has been subjected to a differential tectonic stress prior to<br />

substantial lithification.<br />

7. Conclusions<br />

In this study, we performed flow modeling using a state-<strong>of</strong>-the-art discrete-feature,<br />

finite-volume simulator applied to a CB data set <strong>of</strong> unprecedented detail <strong>and</strong> scale, which<br />

was derived from a real geological system that serves as an exhumed analog for active<br />

s<strong>and</strong>stone aquifers <strong>and</strong> reservoirs. Our results demonstrate the potential for CB arrays,<br />

such as abundantly exposed in the Aztec s<strong>and</strong>stone at the Valley <strong>of</strong> Fire, Nevada, to exert<br />

pr<strong>of</strong>ound effects on subsurface flow at scales relevant to aquifer <strong>and</strong> reservoir modeling<br />

<strong>and</strong> production. Specifically, we found that:<br />

• The CBs as mapped cause an average three-fold increase in the pressure drop<br />

required to maintain a given flow rate between two wells, as compared to CB-free<br />

s<strong>and</strong>stone;<br />

• There is a mild directional aspect to this pressure effect—an anisotropy—with ∆P<br />

across the dominant b<strong>and</strong> trend exceeding that along the trend by an average <strong>of</strong><br />

25%; <strong>and</strong><br />

198


• The tendency for transported fluids to channel preferentially along the dominant<br />

b<strong>and</strong> trend can overpower the effects <strong>of</strong> regional pressure gradients, both natural<br />

(contaminant plume scenarios) <strong>and</strong> induced (reservoir production scenarios).<br />

Compaction b<strong>and</strong>s, <strong>and</strong> other types <strong>of</strong> deformation b<strong>and</strong>s with similar flow<br />

characteristics, almost certainly exist unrecognized in many s<strong>and</strong>stone aquifers <strong>and</strong><br />

reservoirs. Reliable borehole geophysical techniques for detecting these types <strong>of</strong><br />

structures in the subsurface <strong>and</strong> assessing their gross geometry, density, connectivity <strong>and</strong><br />

petrophysical characteristics are needed. Nonetheless, our modeling suggests that simple<br />

steps based on even limited data can mitigate the impact <strong>of</strong> CB arrays. While not an<br />

exhaustive study, the character <strong>and</strong> magnitude <strong>of</strong> the flow <strong>and</strong> transport effects modeled<br />

here demonstrate that accounting for such b<strong>and</strong> fabrics could prove essential to the<br />

optimal management <strong>of</strong> s<strong>and</strong>stone aquifers <strong>and</strong> reservoirs in which they occur.<br />

8. Acknowledgements<br />

Our thanks go to John Childs for his acute <strong>and</strong> cheerful assistance in the field.<br />

Primary funding for this work was provided by the U.S. Department <strong>of</strong> Energy, Office <strong>of</strong><br />

Basic Energy Sciences under grant DE-FG03-94ER14462 awarded to David Pollard <strong>and</strong><br />

Atilla Aydin at <strong>Stanford</strong> University. Additional support was provided by the <strong>Stanford</strong><br />

Reservoir Simulation Consortium (SUPRI-B) <strong>and</strong> the <strong>Stanford</strong> Rock Fracture Project<br />

(RFP).<br />

199


200


References<br />

Adler, P. M., Jacquin, C. G., <strong>and</strong> Quiblier, J. A., 1990, Flow in simulated porous media:<br />

International Journal <strong>of</strong> Multiphase Flow, v. 16, p. 691-712.<br />

Ahlgren, S. G., 2001, Exploring the formation, microtexture <strong>and</strong> petrophysical properties<br />

<strong>of</strong> deformation b<strong>and</strong>s in porous s<strong>and</strong>stone: AAPG Bulletin, v. 85, no. 11, p. 2046.<br />

Amadei, B., <strong>and</strong> Stephansson, O., 1997, Rock Stress <strong>and</strong> its Measurement: London,<br />

Chapman & Hall, 490 p.<br />

Anderson, E. M., 1951, The Dynamics <strong>of</strong> Faulting <strong>and</strong> Dyke Formation with Application<br />

to Britain: Edinburgh, Oliver <strong>and</strong> Boyd Ltd., 206 p.<br />

Antonellini, M., <strong>and</strong> Aydin, A., 1995, Effect <strong>of</strong> faulting on fluid flow in porous<br />

s<strong>and</strong>stones: geometry <strong>and</strong> spatial distribution: AAPG Bulletin, v. 79, no. 5, p.<br />

642-671.<br />

Antonellini, M. A., <strong>and</strong> Aydin, A., 1994, Effect <strong>of</strong> faulting on fluid flow in porous<br />

s<strong>and</strong>stones: petrophysical properties: AAPG Bulletin, v. 78, no. 3, p. 355-377.<br />

Antonellini, M. A., Aydin, A., <strong>and</strong> Pollard, D. D., 1994a, Microstructure <strong>of</strong> deformation<br />

b<strong>and</strong>s in porous s<strong>and</strong>stones at Arches National Park, Gr<strong>and</strong> County, Utah: Journal<br />

<strong>of</strong> Structural Geology, v. 16, p. 941-959.<br />

Antonellini, M. A., Aydin, A., Pollard, D. D., <strong>and</strong> D'Onfro, P., 1994b, Petrophysical<br />

study <strong>of</strong> faults in s<strong>and</strong>stone using petrographic image analysis <strong>and</strong> x-ray<br />

computerized tomography: Pure <strong>and</strong> Applied Geophysics, v. 143, p. 181-201.<br />

Antonellini, M. A., <strong>and</strong> Pollard, D. D., 1995, Distinct element modeling <strong>of</strong> deformation<br />

b<strong>and</strong>s in s<strong>and</strong>stone: Journal <strong>of</strong> Structural Geology, v. 17, no. 8, p. 1165-1182.<br />

Armstrong, R. L., 1968, Sevier orogenic belt in Nevada <strong>and</strong> Utah: Geol. Soc. Amer. Bull.,<br />

v. 79, p. 429-458.<br />

Aydin, A., 1978, Small faults formed as deformation b<strong>and</strong>s in s<strong>and</strong>stone: Pure <strong>and</strong><br />

Applied Geophysics, v. 116, p. 913-930.<br />

Aydin, A., <strong>and</strong> Johnson, A. M., 1978, Development <strong>of</strong> faults as zones <strong>of</strong> deformation<br />

b<strong>and</strong>s <strong>and</strong> as slip surfaces in s<strong>and</strong>stone: Pure & Applied Geophysics, v. 116, p.<br />

931-942.<br />

-, 1983, Analysis <strong>of</strong> faulting in porous s<strong>and</strong>stones: Journal <strong>of</strong> Structural Geology, v. 5, p.<br />

19-31.<br />

Aydin, A., <strong>and</strong> Reches, Z., 1982, Number <strong>and</strong> orientation <strong>of</strong> fault sets in the field <strong>and</strong> in<br />

experiments: Geology, v. 10, p. 107-112.<br />

201


Bakke, S., <strong>and</strong> Øren, P. E., 1997, 3-D pore-scale modeling <strong>of</strong> s<strong>and</strong>stones <strong>and</strong> flow<br />

simulations in the pore networks: Society <strong>of</strong> Petroleum Engineers Journal, v. 2, p.<br />

136-149.<br />

Baud, P., Klein, E., <strong>and</strong> Wong, T.-f., 2004, Compaction localization in porous<br />

s<strong>and</strong>stones: Spatial evolution <strong>of</strong> damage <strong>and</strong> acoustic emission activity: Journal <strong>of</strong><br />

Structural Geology, v. 26, p. 603-624.<br />

Baud, P., Zhu, W., <strong>and</strong> Wong, T.-f., 2000, Failure mode <strong>and</strong> weakening effect <strong>of</strong> water<br />

on s<strong>and</strong>stone: Journal <strong>of</strong> Geophysical Research, v. 105, p. 16371-16390.<br />

Berryman, J. G., <strong>and</strong> Blair, S. C., 1987, Kozeny-Carman relations <strong>and</strong> image processing<br />

methods for estimating Darcy's constant: Journal <strong>of</strong> Applied Physics, v. 60, p.<br />

1930-1938.<br />

Bésuelle, P., <strong>and</strong> Rudnicki, J. W., 2004, Localization: Shear b<strong>and</strong>s <strong>and</strong> compactions<br />

b<strong>and</strong>s, in Guéguen, Y., <strong>and</strong> Boutéca, M., eds., Mechanics <strong>of</strong> Fluid Saturated<br />

Rocks: International Geophysics Series: New York, Elsevier, p. 219-321.<br />

Bieniawski, Z. T., 1984, Rock Mechanics Design in Mining <strong>and</strong> Tunneling: Rotterdam, A.<br />

A. Balkema, 272 p.<br />

Bilodeau, W. L., <strong>and</strong> Keith, S. B., 1986, Lower Jurassic Navajo-Aztec equivalent<br />

s<strong>and</strong>stones in southern Arizona <strong>and</strong> their paleogeographic significance: American<br />

Association <strong>of</strong> Petroleum Geologists Bulletin, v. 70, no. 6, p. 690-701.<br />

Blair, S. C., Berge, P. A., <strong>and</strong> Berryman, J. G., 1996, Using two-point correlation<br />

functions to characterize microgeometry <strong>and</strong> estimate permeabilities <strong>of</strong><br />

s<strong>and</strong>stones <strong>and</strong> porous glass: Journal <strong>of</strong> Geophysical Research, v. 101, no. B9, p.<br />

20359-20376.<br />

Blakey, R. C., 1989, Triassic <strong>and</strong> Jurassic <strong>geology</strong> <strong>of</strong> the southern Colorado Plateau, in<br />

Jenny, J. P., <strong>and</strong> Reynolds, S. J., eds., Geologic evolution <strong>of</strong> Arizona: Tuscon,<br />

Arizona Geological Society, p. 369-396.<br />

Bohannon, R. G., 1977, Geologic map <strong>and</strong> sections <strong>of</strong> the Valley <strong>of</strong> Fire region, North<br />

Muddy Mountains, Clark County, Nevada: U. S. Geological Survey<br />

Miscellaneous Field Studies, Map MF-849, 1:25,000 scale, 2 sheets.<br />

-, 1983, Mesozoic <strong>and</strong> Cenozoic tectonic development <strong>of</strong> the Muddy, North Muddy, <strong>and</strong><br />

northern Black Mountains, Clark County, Nevada: GSA Memoir, v. 157, p. 125-<br />

148.<br />

-, 1983b, Geologic map, tectonic map <strong>and</strong> structure sections <strong>of</strong> the Muddy <strong>and</strong> northern<br />

Muddy Mountains, Clark County, Nevada: U. S. Geological Survey<br />

Miscellaneous Investigations Series, Map I-1406, 1:62,500 (2 sheets).<br />

202


-, 1984, Non-marine sedimentary rocks <strong>of</strong> Tertiary age in the Lake Mead region,<br />

southeastern Nevada <strong>and</strong> northwestern Arizona: U.S. Geological Survey<br />

Pr<strong>of</strong>essional Paper 122, 72 p.<br />

Bohannon, R. G., Grow, J. A., Miller, J. J., <strong>and</strong> Blank, R. H., 1993, Seismic stratigraphy<br />

<strong>and</strong> tectonic development <strong>of</strong> the Virgin River depression <strong>and</strong> associated basins,<br />

southeastern Nevada <strong>and</strong> northwestern Arizona: Geological Society <strong>of</strong> America<br />

Bulletin, v. 105, p. 501-520.<br />

Borja, R. I., <strong>and</strong> Aydin, A., 2004, Computational modeling <strong>of</strong> deformation b<strong>and</strong>s in<br />

granular media, I: Geological <strong>and</strong> mathematical framework: Computer Methods<br />

in Applied Mechanics <strong>and</strong> Engineering, v. 193, no. 26-29, p. 2667-2698.<br />

Bosl, W. J., Dvorkin, J., <strong>and</strong> Nur, A., 1998, A numerical study <strong>of</strong> pore structure <strong>and</strong><br />

permeability using a Lattice-Boltzmann simulation: Geophysical Research Letters,<br />

v. 25, p. 1475-1478.<br />

Bourgeat, A., 1984, Homogenized behavior <strong>of</strong> two-phase flows in naturally fractured<br />

reservoirs with uniform fracture distribution: Computational Methods in Applied<br />

Mechanical Engineering, v. 47, p. 273-286.<br />

Broberg, K. B., 1987, On crack paths: Engineering Fracture Mechanics, v. 28, no. 5/6, p.<br />

663-679.<br />

Brock, W. G., <strong>and</strong> Engleder, T., 1979, Deformation associated with the movement <strong>of</strong> the<br />

Muddy Mountain overthrust in the Buffington window, southeastern Nevada:<br />

Geological Society <strong>of</strong> America Bulletin, v. 88, p. 1667-1677.<br />

Burchfiel, B. C., Cowan, D. S., <strong>and</strong> Davis, G. A., 1992, Tectonic overview <strong>of</strong> the<br />

Cordilleran orogen in the western Unite States, in Burchfiel, B. C., Lipman, P. W.,<br />

<strong>and</strong> Zoback, M. D., eds., The Cordilleran Orogen: Conterminous U.S.: Boulder,<br />

CO, Geological Society <strong>of</strong> America, p. 407-479.<br />

Burhannudinnur, M., <strong>and</strong> Morley, C. K., 1997, Anatomy <strong>of</strong> growth fault zones in poorly<br />

lithified s<strong>and</strong>stones <strong>and</strong> shales: implications for reservoir studies <strong>and</strong> seismic<br />

interpretation: part 1, outcrop study: Petroleum Geoscience, v. 3, p. 211-224.<br />

Campagna, D. J., <strong>and</strong> Aydin, A., 1994, Basin genesis associated with strike-slip faulting<br />

in the Basin <strong>and</strong> Range, southeastern Nevada: Tectonics, v. 13, p. 327-341.<br />

Cancelliere, A., Chang, C., Foti, E., Rothman, D. H., <strong>and</strong> Suci, S., 1998, The<br />

permeability <strong>of</strong> a r<strong>and</strong>om medium: Comparison <strong>of</strong> simulation with theory:<br />

Physics <strong>of</strong> Fluids A, v. 2, p. 2085-2088.<br />

Cao, H., 2002, Development <strong>of</strong> techniques for general-purpose simulators [Ph.D. thesis<br />

thesis]: <strong>Stanford</strong> University, 184 p.<br />

203


Carpenter, D. G., <strong>and</strong> Carpenter, J. A., 1994, Fold-thrust structure, synorogenic rocks,<br />

<strong>and</strong> <strong>structural</strong> analysis <strong>of</strong> the North Muddy <strong>and</strong> Muddy Mountains, Clark County,<br />

Nevada, in Dobbs, S. W., <strong>and</strong> Taylor, W. J., eds., Structural <strong>and</strong> stratigraphhic<br />

investigations <strong>and</strong> petroleum potential <strong>of</strong> Nevada, with special emphasis south <strong>of</strong><br />

the Railroad Valley producing trend: Nevada Petroleum Society Conference II, p.<br />

65-94.<br />

Cashman, S., <strong>and</strong> Cashman, K., 2000, Cataclasis <strong>and</strong> deformation-b<strong>and</strong> formation in<br />

unconsolidated marine terrace s<strong>and</strong>, Humboldt County, California: Geology, v. 28,<br />

no. 2, p. 111-114.<br />

Chen, Y., Durl<strong>of</strong>sky, L. J., Gerritsen , M., <strong>and</strong> Wen, X. H., 2003, A coupled local-global<br />

upscaling approach for simulating flow in highly heterogeneous formations:<br />

Advances in Water Resources, v. 26, p. 1041-1060.<br />

Chester, J. S., Lenz, S. C., Chester, F. M., <strong>and</strong> Lang, R. A., 2004, Mechanisms <strong>of</strong><br />

compaction <strong>of</strong> quartz s<strong>and</strong> at diagenetic conditions: Earth <strong>and</strong> Planetary Science<br />

Letters, v. 220, no. 3-4, p. 435-451.<br />

Cocco, M., <strong>and</strong> Rice, J. R., 2002, Pore pressure <strong>and</strong> poroelasticity effects in Coulomb<br />

stress analysis <strong>of</strong> earthquake interactions: Journal <strong>of</strong> Geophysical Research, v.<br />

107, B22030, doi: 10.1029/2000JB000138.<br />

Cotterell, B., <strong>and</strong> Rice, J. R., 1980, Slightly curved or kinked cracks: International<br />

Journal <strong>of</strong> Fracture, v. 16, p. 155-169.<br />

Crawford, B. R., 1998, Experimental fault sealing: shear b<strong>and</strong> permeability dependency<br />

on cataclastic fault gouge characteristics, in Coward, M. P., Daltaban, M. P., <strong>and</strong><br />

Johnson, T. S., eds., Structural <strong>geology</strong> in reservoir characterization, Geological<br />

Society <strong>of</strong> London Special Publications, v. 127, p. 27-47.<br />

Crouch, S. L., 1976, Solution <strong>of</strong> plane elastictiy problems by the displacement<br />

discontinuity method: International Journal <strong>of</strong> Numerical Methods in Engineering,<br />

v. 10, p. 301-343.<br />

Crouch, S. L., <strong>and</strong> Starfield, A. M., 1983, Boundary Element Methods in Solid<br />

Mechanics: with applications in rock <strong>mechanics</strong> <strong>and</strong> geological engineering:<br />

London, George Allen & Unwin, 322 p.<br />

Cruikshank, K. M., Zhao, G., <strong>and</strong> Johnson, A. M., 1991, Analysis <strong>of</strong> minor fractures<br />

associated with joints <strong>and</strong> faulted joints: Journal <strong>of</strong> Structural Geology, v. 13, p.<br />

865-886.<br />

Davatzes, N. C., Eichhubl, P., <strong>and</strong> Aydin, A., 2005, The <strong>structural</strong> evolution <strong>of</strong> fault<br />

zones in s<strong>and</strong>stone: The Moab fault, SE Utah: Geological Society <strong>of</strong> America<br />

Bulletin, v. 117, no. 1/2, p. 135-148.<br />

Davis, G. H., 1998, Fault-fin l<strong>and</strong>scape: Geological Magazine, v. 135, p. 283-286.<br />

204


Delaney, P. T., Pollard, D. D., Ziony, J. I., <strong>and</strong> McKee, E. H., 1986, Field relations<br />

between dikes <strong>and</strong> joints: emplacement processes <strong>and</strong> paleostress analysis:<br />

Journal <strong>of</strong> Geophysical Research, v. 91, no. B5, p. 4,920-4,938.<br />

Dentz, M., Kinzelbach, H., Attinger, S., <strong>and</strong> Kinzelbach, W., 2002, Temporal behavior <strong>of</strong><br />

a solute cloud in a heterogeneous porous medium, 3: Numerical simulations:<br />

Water Resources Research, v. 38, no. 7, p. 1118, doi: 10.1029/2001WR000436.<br />

Desbarats, A. J., 1987, Numerical estimation <strong>of</strong> effective permeability in s<strong>and</strong>-shale<br />

formations: Water Resources Research, v. 23, no. 2, p. 273-286.<br />

DeTournay, C., Cunall, P., <strong>and</strong> Parra, J., 2003, A study <strong>of</strong> compaction b<strong>and</strong> formation<br />

with the double yield model, in Brummer, R., ed., FLAC <strong>and</strong> Numerical<br />

Modeling in Geo<strong>mechanics</strong>: Proceeding <strong>of</strong> the Third International FLAC<br />

Symposium: Lisse, Netherl<strong>and</strong>s, Swets & Zeitlinger, p. 27-33.<br />

Deutsch, C., 1989, Calculating effective absolute permeability in s<strong>and</strong>stone/shale<br />

sequences: Society <strong>of</strong> Petroleum Engineers Formation Evaluations, v. 4, p. 343-<br />

348.<br />

Deutsch, C. V., <strong>and</strong> Journel, A. G., 1998, GSLIB: Geostatistical S<strong>of</strong>tware Library <strong>and</strong><br />

User's Guide: New York, Oxford University Press, 369 p.<br />

Doolen, G. D., 1990, Lattice Gas Methods for Partial Differential Equations: Redwood<br />

City, California, Addison-Wesley, 554 p.<br />

Du Bernard, X., Eichhubl, P., <strong>and</strong> Aydin, A., 2002, Dilation B<strong>and</strong>s: A New Form <strong>of</strong><br />

Localized Failure in Granular Media: Geophysical Research Letters, v. 29, no. 24,<br />

doi: 1029/2002GLO15966.<br />

Duebendorfer, E. M., Beard, L. S., <strong>and</strong> Smith, E. I., 1998, Restoration <strong>of</strong> Tertiary<br />

deformation in the Lake Mead region, southern Nevada: the role <strong>of</strong> strike-slip<br />

transfer faults, in Faulds, J. E., <strong>and</strong> Stewart, J. H., eds., Accommodation zones<br />

<strong>and</strong> tranfer zones: Regional segmentation <strong>of</strong> the Basin <strong>and</strong> Range province,<br />

Geological Society <strong>of</strong> America Special Paper 323, p. 127-148.<br />

Durl<strong>of</strong>sky, L. J., 1991, Numerical calculation <strong>of</strong> equivalent grid block permeability<br />

tensors for heterogeneous porous media: Water Resources Research, v. 27, p.<br />

699-708.<br />

-, 1992, Representations <strong>of</strong> grid-block permeability in coarse-scale models <strong>of</strong> r<strong>and</strong>omly<br />

heterogeneous porous media: Water Resources Research, v. 28, p. 1791-1800.<br />

Edwards, E. H., Becker, A. D., <strong>and</strong> Howell, J. A., 1993, Compartimentalization <strong>of</strong> an<br />

eolian s<strong>and</strong>stone by <strong>structural</strong> heterogeneities: Permo-Triassic Hopeman<br />

S<strong>and</strong>stone, Moray Firth, Scotl<strong>and</strong>, in North, C. P., <strong>and</strong> Prosser, D. J., eds.,<br />

Characterization <strong>of</strong> fluvial <strong>and</strong> eolian reservoirs: London, Geological Society <strong>of</strong><br />

London, p. 339-365.<br />

205


Eichhubl, P., Taylor, W. L., Pollard, D. D., <strong>and</strong> Aydin, A., 2004, Paleo-fluid flow <strong>and</strong><br />

deformation in the Aztec S<strong>and</strong>stone at the Valley <strong>of</strong> Fire, Nevada--Evidence for<br />

the coupling <strong>of</strong> hydrogeological, diagenetic <strong>and</strong> tectonic processes: Geological<br />

Society <strong>of</strong> America Bulletin, v. 116, no. 9-10, p. 1120-1136.<br />

Engelder, T., 1974, Cataclasis <strong>and</strong> the generation <strong>of</strong> fault gouge: Geol. Soc. America<br />

Bull., v. 85, p. 1515-1522.<br />

Erdogan, F., <strong>and</strong> Sih, G. C., 1963, On crack extension in plates under plane loading <strong>and</strong><br />

transverse shear: Journal <strong>of</strong> Basic Engineering—Transactions <strong>of</strong> American<br />

Society <strong>of</strong> Mechanical Engineers, v. 85, p. 519-527.<br />

Eshelby, J. D., 1957, The determination <strong>of</strong> the elastic field <strong>of</strong> an ellipsoidal inclusion <strong>and</strong><br />

related problems: Royal Society <strong>of</strong> London, Proceedings, v. A241, p. 376-396.<br />

-, 1959, The elastic field outside an ellipsoidal inclusion: Proceedings <strong>of</strong> the Royal<br />

Society <strong>of</strong> London, v. A252, p. 561-569.<br />

Fleck, N. A., 1991, Brittle fracture due to an array <strong>of</strong> microcracks: Proceedings <strong>of</strong> the<br />

Royal Society <strong>of</strong> London, v. A432, p. 55-76.<br />

Fleck, R. J., 1970, Tectonic style, magnitude <strong>and</strong> age <strong>of</strong> deformation in the Sevier<br />

orogenic belt in southern Nevada <strong>and</strong> eastern California: Geological Society <strong>of</strong><br />

America Bulletin, v. 81, p. 1705-1720.<br />

Fletcher, R. C., <strong>and</strong> Pollard, D. D., 1981, Anticrack model for pressure solution surfaces:<br />

Geology, v. 9, p. 419-424.<br />

Flodin, E. A., 2003, Structural evolution, petrophysics <strong>and</strong> large-scale permeability <strong>of</strong><br />

faults in s<strong>and</strong>stone, Valley <strong>of</strong> Fire, Nevada [Ph.D. thesis]: <strong>Stanford</strong> University,<br />

180 p.<br />

Flodin, E. A., <strong>and</strong> Aydin, A., 2004, Evolution <strong>of</strong> a strike-slip fault network, Valley <strong>of</strong><br />

Fire, southern Nevada: Geological Society <strong>of</strong> America Bulletin, v. 116, p. 42-59.<br />

Flodin, E. A., Gerdes, M., Aydin, A., <strong>and</strong> Wiggins, W. D., 2005, Petrophysical properties<br />

<strong>and</strong> sealing capacity <strong>of</strong> fault rock from sheared-joint based faults, Aztec<br />

S<strong>and</strong>stone, Nevada, in Tsuji, Y., <strong>and</strong> Sorkhabi, R., eds., Fault seals <strong>and</strong> petroleum<br />

traps: American Association <strong>of</strong> Petroleum Geologists Memoir, v. 85, p. 197-217.<br />

Flodin, E. A., Prasad, M., <strong>and</strong> Aydin, A., 2003, Petrophysical constraints on deformation<br />

styles in Aztec S<strong>and</strong>stone: Pure <strong>and</strong> Applied Geophysics, v. 160, p. 1589-1610.<br />

Fossen, H., <strong>and</strong> Hesthammer, J., 1998, Deformation b<strong>and</strong>s <strong>and</strong> their significance in<br />

porous s<strong>and</strong>stone reservoirs: First Break, v. 16, no. 1, p. 21-25.<br />

Freeman, D. H., 1990, Permeability effects <strong>of</strong> deformation b<strong>and</strong>s in porous s<strong>and</strong>stones<br />

[Master's thesis]: University <strong>of</strong> Oklahoma, 90 p.<br />

206


Freeze, R. A., <strong>and</strong> Cherry, J. A., 1979, Groundwater: Englewood Cliffs, New Jersey,<br />

Prentice-Hall, Inc., 604 p.<br />

Gibson, R. G., 1998, Physical character <strong>and</strong> fluid-flow properties <strong>of</strong> s<strong>and</strong>stone-derived<br />

fault zones, in Coward, M. P., Daltaban, T. S., <strong>and</strong> Johnson, H., eds., Structural<br />

Geology in Reservoir Characterization: Geological Society <strong>of</strong> London Special<br />

Publications, v. 127, p. 83-97.<br />

Haimson, B. C., 2001, Fracture-like borehole breakouts in high-porosity s<strong>and</strong>stone: Are<br />

they caused by compaction b<strong>and</strong>s?: Physics <strong>and</strong> Chemistry <strong>of</strong> the Earth, Part A, v.<br />

26, p. 15-20.<br />

-, 2003, Borehole breakouts in Berea s<strong>and</strong>stone reveal a new fracture mechanism: Pure<br />

<strong>and</strong> Applied Geophysics, v. 160, p. 813-831.<br />

Haimson, B. C., <strong>and</strong> Lee, H., 2004, Borehole breakouts <strong>and</strong> compaction b<strong>and</strong>s in two<br />

high-porosity s<strong>and</strong>stones: International Journal <strong>of</strong> Rock Mechanics <strong>and</strong> Mining<br />

Science, v. 41, p. 287-301.<br />

Hesthammer, J., <strong>and</strong> Henden, J. O., 2000, Information on fault orientation from<br />

unoriented cores: AAPG Bulletin, v. 84, no. 4, p. 472-488.<br />

Hesthammer, J., Johansen, T. E. S., <strong>and</strong> Watts, L., 2000, Spatial relationships within fault<br />

damage zones in s<strong>and</strong>stone: Marine <strong>and</strong> Petroleum Geology, v. 17, p. 873-893.<br />

Hill, R. E., 1989, Analysis <strong>of</strong> deformation b<strong>and</strong>s in the Aztec S<strong>and</strong>stone, Valley <strong>of</strong> Fire<br />

State Park, Nevada [Master’s thesis]: University <strong>of</strong> Nevada, 68 p.<br />

Hoagl<strong>and</strong>, R. G., Hahn, G. T., <strong>and</strong> Rosenfield, A. R., 1973, Influence <strong>of</strong> microstructure<br />

on fracture <strong>propagation</strong> in rock: Rock Mechanics, v. 5, p. 77-106.<br />

Holcomb, D. J., <strong>and</strong> Olsson, W. A., 2003, Compaction localization <strong>and</strong> fluid flow:<br />

Journal <strong>of</strong> Geophysical Research, v. 108, no. B6, p. 2290, doi:<br />

10.1029/2001JB000813.<br />

Irwin, G. R., 1960, Fracture Mechanics, in Structural Mechanics—1st Symposium on<br />

Naval Structural Mechanics, Proceedings, p. 557-592.<br />

Issen, K. A., <strong>and</strong> Rudnicki, J. W., 2000, Conditions for compaction b<strong>and</strong>s in porous rock:<br />

Journal <strong>of</strong> Geophysical Research, v. 105, p. 21,529-21,536.<br />

-, 2001, Theory <strong>of</strong> compaction b<strong>and</strong>s in porous rock: Physics <strong>and</strong> Chemistry <strong>of</strong> the Earth,<br />

Part A, v. 26, no. 1-2, p. 95-100.<br />

Jamison, W. R., <strong>and</strong> Stearns, D. W., 1982, Tectonic deformation <strong>of</strong> Wingate S<strong>and</strong>stone,<br />

Colorado National Monument: American Association <strong>of</strong> Petroleum Geologists<br />

Bulletin, v. 66, no. 12, p. 2584-2608.<br />

207


Karimi-Fard, M., Durl<strong>of</strong>sky, L. J., <strong>and</strong> Aziz, K., 2004, An efficient discrete-fracture<br />

model applicable for general-purpose reservoir simulators: SPE Journal, v. 9, no.<br />

2, p. 227-236.<br />

Karner, S. L., Chester, F. M., Kronenberg, A. K., <strong>and</strong> Chester, J. S., 2003, Subcritical<br />

compaction <strong>and</strong> yielding <strong>of</strong> granular quartz s<strong>and</strong>: Tectonophysics, v. 377, no. 3-4,<br />

p. 357-381.<br />

Katsman, R., Aharonov, E., <strong>and</strong> Scher, H., 2005, Numerical simulation <strong>of</strong> compaction<br />

b<strong>and</strong>s in high-porosity sedimentary rock: Mechanics <strong>of</strong> Materials, v. 37, no. 1, p.<br />

143-162.<br />

Katsman, R., <strong>and</strong> Aharonov, E., 2006, A study <strong>of</strong> compaction b<strong>and</strong>s originating from<br />

cracks, notches <strong>and</strong> compacted defects: Journal <strong>of</strong> Structural Geology, v. 28, p.<br />

508-518.<br />

Keehm, Y., Mukerji, T., <strong>and</strong> Nur, A., 2001, Computational rock physics at the pore scale:<br />

Transport properties <strong>and</strong> diagenesis in realistic pore geometries: The Leading<br />

Edge, v. 20, p. 180-183.<br />

-, 2004, Permeability prediction from thin sections: 3D reconstruction <strong>and</strong> Lattice-<br />

Boltzmann flow simulation: Geophysical Research Letters, v. 31, p. L04606, doi:<br />

10.1029/2003GL018761.<br />

Klaetsch, A. R., <strong>and</strong> Haimson, B. C., 2002, Porosity-dependent fracture-like breakouts in<br />

St. Peter s<strong>and</strong>stone, in Hammah, R., ed., Mining <strong>and</strong> Tunneling Innovation <strong>and</strong><br />

Opportunity: Toronto, University <strong>of</strong> Toronto Press, p. 1356-1371.<br />

Klein, E., Baud, P., Reuschlé, T., <strong>and</strong> Wong, T.-f., 2001, Mechanical behavior <strong>and</strong> failure<br />

mode <strong>of</strong> Bentheim s<strong>and</strong>stone under triaxial compression: Physics <strong>and</strong> Chemistry<br />

<strong>of</strong> the Earth, Part A, v. 26, p. 21-25.<br />

Ladd, A. J. C., 1994, Numerical simulations <strong>of</strong> particulate suspensions via a discretized<br />

Boltzmann equation, Part 2: Numerical Results: Journal <strong>of</strong> Fluid Mechanics, v.<br />

271, p. 311-339.<br />

Lawn, B. R., <strong>and</strong> Wilshaw, T. R., 1975, Fracture <strong>of</strong> Brittle Solids: Cambridge,<br />

Cambridge University Press, 204 p.<br />

Lewis, H., <strong>and</strong> Couples, G. D., 1993, Production evidence for geological heterogeneities<br />

in the Anschutz Ranch East Field, western USA: Geological Society Special<br />

Publications, v. 73, p. 321-338.<br />

Lindquist, S. J., 1988, Practical characterization <strong>of</strong> eolian reservoirs for development:<br />

Nugget S<strong>and</strong>stone, Utah-Wyoming thrust belt: Sedimentary Geology, v. 56, no. 1-<br />

4, p. 315-339.<br />

208


Longwell, C. R., 1921, Geology <strong>of</strong> the Muddy Mountains, Nevada: American Journal <strong>of</strong><br />

Science, v. 1, no. 5, p. 1921.<br />

-, 1949, Structure <strong>of</strong> the northern Muddy Mountain area, Nevada: Geological Society <strong>of</strong><br />

America Bulletin, v. 60, p. 923-967.<br />

-, 1960, Possible explanation <strong>of</strong> diverse <strong>structural</strong> patterns in southern Nevada: Amer.<br />

Jour. Sci., v. 258-A, no. 192-203.<br />

Lothe, A. E., Gabrielsen, R. H., Bjornevoll-Hagen, N., <strong>and</strong> Larsen, B. T., 2002, An<br />

experimental study <strong>of</strong> the texture <strong>of</strong> deformation b<strong>and</strong>s; effects on the porosity<br />

<strong>and</strong> permeability <strong>of</strong> s<strong>and</strong>stones: Petroleum Geoscience, v. 8, no. 3, p. 195-207.<br />

Manwart, C., Aaltosalmi, U., Koponen, A., Hilfer, R., <strong>and</strong> Timonen, J., 2002, Lattice-<br />

Boltzmann <strong>and</strong> finite-difference simulations <strong>of</strong> permeability for three-dimensional<br />

porous media: Physical Review E, v. 66, p. 16702-16802.<br />

Mardon, D., 1988, Localization <strong>of</strong> pressure solution <strong>and</strong> the formation <strong>of</strong> discrete<br />

solution seams [Ph.D. thesis]: Texas A&M University.<br />

Marzolf, J. E., 1983, Changing wind <strong>and</strong> hydrologic regimes during deposition <strong>of</strong> the<br />

Navajo <strong>and</strong> Aztec s<strong>and</strong>stones, southwestern United States, in Brookfield, M. E.,<br />

<strong>and</strong> Ahlbr<strong>and</strong>t, T. S., eds., Eolian Sediments <strong>and</strong> Processes: Amsterdam, Elsevier,<br />

p. 635-660.<br />

-, 1986, Nugget-Navajo-Aztec s<strong>and</strong>stone: interaction <strong>of</strong> eolian s<strong>and</strong> sea with Andean-type<br />

volcanic arc: AAPG Bulletin, v. 70, no. 5, p. 616.<br />

-, 1990, Reconstruction <strong>of</strong> extensionally dismembered early Mesozoic sedimentary<br />

basins: Southwestern Colorado Plateau to the eastern Mojave Desert, in Wernicke,<br />

B. P., ed., Basin <strong>and</strong> Range extensional tectonics near the latitude <strong>of</strong> Las Vegas,<br />

Nevada: Geological Society <strong>of</strong> America Memoir, v. 176, p. 477-500.<br />

Matthai, S. K., Aydin, A., Pollard, D. D., <strong>and</strong> Roberts, S. G., 1998, Numerical simulation<br />

<strong>of</strong> departures from radial drawdown in a faulted s<strong>and</strong>stone reservoir with joints<br />

<strong>and</strong> deformation b<strong>and</strong>s, in Jones, G., Fisher, Q. J., <strong>and</strong> Knipe, R. J., eds., Faulting,<br />

Fault Sealing <strong>and</strong> Fluid Flow in Hydrocarbon Reservoirs: Geological Society <strong>of</strong><br />

London Special Publications, v. 147, p. 157-191.<br />

Mavko, G., Mukerji, T., <strong>and</strong> Dvorkin, J., 1998, The Rock Physics H<strong>and</strong>book: Cambridge,<br />

U.K., Cambridge University Press, 329 p.<br />

Mavko, G., <strong>and</strong> Nur, A., 1997, The effect <strong>of</strong> percolation threshold in the Kozeny-Carmen<br />

relation: Geophysics, v. 62, p. 1480-1482.<br />

Melin, S., 1983, Why do cracks avoid each other?: International Journal <strong>of</strong> Fracture, v. 23,<br />

p. 37-45.<br />

209


-, 1987, Fracture from a straight crack subjected to mixed mode loading: International<br />

Journal <strong>of</strong> Fracture, v. 32, p. 257-263.<br />

Mollema, P. N., <strong>and</strong> Antonellini, M. A., 1996, Compaction b<strong>and</strong>s: a <strong>structural</strong> analog for<br />

anti-mod I cracks in eolian s<strong>and</strong>stone: Tectonophysics, v. 267, p. 209-228.<br />

Morgan, J. K., 1999, Numerical simulations <strong>of</strong> granular shear zones using the distinct<br />

element method, 2. Effect <strong>of</strong> particle size distribution <strong>and</strong> inter-particle friction on<br />

mechanical behavior: Journal <strong>of</strong> Geophysical Research, v. 104, p. 2721-2732.<br />

Morgan, J. K., <strong>and</strong> Boettcher, M. S., 1999, Numerical simulations <strong>of</strong> granular shear zones<br />

using the distinct element method. 1. Shear zone kinematics <strong>and</strong> micro<strong>mechanics</strong><br />

<strong>of</strong> localization: Journal <strong>of</strong> Geophysical Research, v. 104, p. 2703-2719.<br />

Mura, T., 1987, Micro<strong>mechanics</strong> <strong>of</strong> defects in solids: Hague, Netherl<strong>and</strong>s, Martinus<br />

Nijh<strong>of</strong>f.<br />

Myers, R., 1999, Structure <strong>and</strong> hydraulics <strong>of</strong> brittle faults in s<strong>and</strong>stone [Ph.D. thesis]:<br />

<strong>Stanford</strong> University, 176 p.<br />

Myers, R., <strong>and</strong> Aydin, A., 2004, The evolution <strong>of</strong> faults formed by shearing across joint<br />

zones in s<strong>and</strong>stone: Journal <strong>of</strong> Structural Geology, v. 26, no. 5, p. 947-966.<br />

Naff, R. L., Haley, D. F., <strong>and</strong> Sudicky, E. A., 1998, High-resolution Monte Carlo<br />

simulation <strong>of</strong> flow <strong>and</strong> conservative transport in heterogeneous porous media, 2:<br />

Transport results: Water Resources Research, v. 34, no. 4, p. 679-698, doi:<br />

10.1029/1997WR02711.<br />

Olson, J., <strong>and</strong> Pollard, D. D., 1989, Inferring paleostresses from natural fracture patterns:<br />

A new method: Geology, v. 17, p. 345-348.<br />

Olsson, W. A., 1999, Theoretical <strong>and</strong> experimental investigation <strong>of</strong> compaction b<strong>and</strong>s in<br />

porous rock: Journal <strong>of</strong> Geophysical Research, v. 104, p. 7219-7228.<br />

Olsson, W. A., <strong>and</strong> Holcomb, D. J., 2000, Compaction localization in porous rock:<br />

Geophysical Research Letters, v. 27, p. 3537-3540.<br />

Olsson, W. A., Lorenz, J. C., <strong>and</strong> Cooper, S. P., 2004, A mechanical model for multiplyoriented<br />

conjugate deformation b<strong>and</strong>s: Journal <strong>of</strong> Structural Geology, v. 26, no. 2,<br />

p. 325-338.<br />

Øren, P. E., <strong>and</strong> Bakke, S., 2002, Process-based reconstruction <strong>of</strong> s<strong>and</strong>stones <strong>and</strong><br />

prediction <strong>of</strong> transport properties: Transport in Porous Media, v. 46, p. 311-343.<br />

Peck, L., Barton, C. C., <strong>and</strong> Gordon, R. B., 1985a, Microstructure <strong>and</strong> the resistance <strong>of</strong><br />

rock to tensile fracture: Journal <strong>of</strong> Geophysical Research, v. 90, p. 11,533-11,546.<br />

210


Peck, L., Nolen-Hoeksema, R. C., Barton, C. C., <strong>and</strong> Gordon, R. B., 1985b, Measurement<br />

<strong>of</strong> the resistance <strong>of</strong> imperfectly elastic rock to the <strong>propagation</strong> <strong>of</strong> tensile cracks:<br />

Journal <strong>of</strong> Geophysical Research, v. 90, p. 7,827-7,836.<br />

Pittman, E. D., 1981, Effect <strong>of</strong> fault-related granulation on porosity <strong>and</strong> permeability <strong>of</strong><br />

quartz s<strong>and</strong>stone, Simpson Group (Ordovician), Oklahoma: American<br />

Association <strong>of</strong> Petroleum Geologists Bulletin, v. 65, p. 2381-2387.<br />

Pollard, D. D., <strong>and</strong> Aydin, A. A., 1988, Progress in underst<strong>and</strong>ing jointing over the past<br />

century: Geological Society <strong>of</strong> America Bulletin, v. 100, p. 1,181-1,204.<br />

Renard, F., <strong>and</strong> de Marsily, G., 1997, Calculating equivalent permeability: A review:<br />

Advances in Water Resources, v. 20, p. 253-278.<br />

Rice, J. R., 1968, A path independent integral <strong>and</strong> the approximate analysis <strong>of</strong> strain<br />

concentration by notches <strong>and</strong> cracks: Journal <strong>of</strong> Applied Mechanics, v. 35, p. 379-<br />

386.<br />

Rudnicki, J. W., 1999, Alteration <strong>of</strong> regional stress by reservoirs <strong>and</strong> other<br />

inhomogeneities: Stabilizing or destabilizing?, in Proceedings <strong>of</strong> the Ninth<br />

International Congress on Rock Mechanics, Paris, p. 1629-1637.<br />

-, 2002, Conditions for compaction <strong>and</strong> shear b<strong>and</strong>s in a transversely isotropic material:<br />

International Journal <strong>of</strong> Solids <strong>and</strong> Structures, v. 39, p. 3741-3756.<br />

-, 2003, Compaction b<strong>and</strong>s in porous rock, in Labuz, J. F., <strong>and</strong> Drescher, A., eds.,<br />

Bifurcations <strong>and</strong> Instabilities in Geo<strong>mechanics</strong>: Lisse, Netherl<strong>and</strong>s, Swets &<br />

Zeitlinger.<br />

-, 2004, Shear <strong>and</strong> compaction b<strong>and</strong> formation on an elliptic yield cap: Journal <strong>of</strong><br />

Geophysical Research, v. 109, p. B03402, doi: 1029/2003JB002633.<br />

Rudnicki, J. W., <strong>and</strong> Sternl<strong>of</strong>, K. R., 2005, Energy release model for compaction b<strong>and</strong><br />

<strong>propagation</strong>: Geophysical Research Letters, v. 32, no. L16303, doi:<br />

10.1029/2005GL023602.<br />

Segall, P., <strong>and</strong> Pollard, D. D., 1983, Joint formation in granitic rock <strong>of</strong> the Sierra Nevada:<br />

Geological Society <strong>of</strong> America Bulletin, v. 94, p. 563-575.<br />

Sempere, J.-C., <strong>and</strong> Macdonald, K. C., 1986, Overlapping spreading centers: Implications<br />

from crack growth simulation by the displacement discontinuity method:<br />

Tectonics, v. 5, p. 151-163.<br />

Shewchuk, J. R., 1996, Triangle: Engineering a 2D quality mesh generator <strong>and</strong> Delaunay<br />

triangulator, in First Workshop on Applied Computational Geometry,<br />

Philadelphia, PA, U.S.A., p. 124-133.<br />

211


Shipton, Z. K., <strong>and</strong> Cowie, P. A., 2001, Damage zone <strong>and</strong> slip-surface evolution over<br />

micrometre to kilometre scales in high-porosity Navajo s<strong>and</strong>stone, Utah: Journal<br />

<strong>of</strong> Structural Geology, v. 23, p. 1825-1844.<br />

Sigda, J., <strong>and</strong> Wilson, J., 2003a, Are faults preferential flow paths through semiarid <strong>and</strong><br />

arid vadose zones?: Water Resources Research, v. 39, no. 8, 1225, doi:<br />

10.1029/2002WR001406.<br />

Stephen, K. D., <strong>and</strong> Dalrymple, M., 2002, Reservoir simulations developed from an<br />

outcrop <strong>of</strong> incised valley fill strata: AAPG Bulletin, v. 86, no. 5, p. 797-822.<br />

Sternl<strong>of</strong>, K., <strong>and</strong> Pollard, D., 2001, Deformation b<strong>and</strong>s as linear elastic fractures:<br />

Progress in theory <strong>and</strong> observation: EOS Trans. Amer. Geophys. Union, v. 82, no.<br />

47, Fall Meeting Supplemental Abstract T42E-04.<br />

-, 2002, Numerical modeling <strong>of</strong> compactive deformation b<strong>and</strong>s as granular anti-cracks:<br />

EOS Trans. Amer. Geophys. Union, v. 83, no. 47, Fall Meeting Supplemental<br />

Abstract T11F-10.<br />

Sternl<strong>of</strong>, K. R., Chapin, J. R., Pollard, D. D., <strong>and</strong> Durl<strong>of</strong>sky, L. J., 2004, Permeability<br />

effects <strong>of</strong> deformation b<strong>and</strong> arrays in s<strong>and</strong>stone: American Association <strong>of</strong><br />

Petroleum Geologists Bulletin, v. 88, no. 9, p. 1315-1329.<br />

Sternl<strong>of</strong>, K. R., Karimi-Fard, M., Pollard, D. D., <strong>and</strong> Durl<strong>of</strong>sky, L. J., 2006, Flow effects<br />

<strong>of</strong> compaction b<strong>and</strong>s in s<strong>and</strong>stone at scales relevant to aquifer <strong>and</strong> reservoir<br />

management: Water Resources Research, in press, doi: 10.1029/2005WR004664.<br />

Sternl<strong>of</strong>, K. R., Rudnicki, J. W., <strong>and</strong> Pollard, D. D., 2005, Anticrack-inclusion model for<br />

compaction b<strong>and</strong>s in s<strong>and</strong>stone: Journal <strong>of</strong> Geophysical Research, v. 110, B11403,<br />

doi: 10.1029/2005JB003764.<br />

Stewart, J. H., <strong>and</strong> Carlson, J. E., 1978, Geologic Map <strong>of</strong> Nevada: U.S. Geological<br />

Survey Map MF-930, 1:500,000 (2 sheets).<br />

Sumi, Y., Nemat-Nasser, S., <strong>and</strong> Keer, L. M., 1985, On crack path stability in a finite<br />

body: Engineering Fracture Mechanics, v. 22, p. 759-771.<br />

Swain, M. V., <strong>and</strong> Hagan, J. T., 1978, Some observations <strong>of</strong> overlapping interacting<br />

cracks: Engineering Fracture Mechanics, v. 10, p. 299-304.<br />

Swierczewska, A., <strong>and</strong> Tokarski, A. K., 1998, Deformation b<strong>and</strong>s <strong>and</strong> the history <strong>of</strong><br />

folding in the Magura Nappe, western outer Carpathians, Pol<strong>and</strong>: Tectonophysics,<br />

v. 297, no. 1-4, p. 73-90.<br />

Taylor, W. L., 1999, Fluid flow <strong>and</strong> chemical alteration in fractured s<strong>and</strong>stones [Ph.D.<br />

thesis]: <strong>Stanford</strong> University, 411 p.<br />

212


Taylor, W. L., <strong>and</strong> Pollard, D. D., 2000, Estimation <strong>of</strong> in-situ permeability <strong>of</strong> deformation<br />

b<strong>and</strong>s in porous s<strong>and</strong>stone, Valley <strong>of</strong> Fire, Nevada: Water Resources Research, v.<br />

36, p. 2595-2606.<br />

Taylor, W. L., Pollard, D. D., <strong>and</strong> Aydin, A., 1999, Fluid flow in discrete joint sets-field<br />

observations <strong>and</strong> numerical simulations: Journal <strong>of</strong> Geophysical Research, v. 104,<br />

p. 28,983-29,006.<br />

Tembe, S., Vajdova, V., Wong, T.-f., <strong>and</strong> Zhu, W., 2006, Initiation <strong>and</strong> <strong>propagation</strong> <strong>of</strong><br />

strain localization in circumferentially notched samples <strong>of</strong> two porous s<strong>and</strong>stones:<br />

Journal <strong>of</strong> Geophysical Research, v. 111, p. B02409, doi: 10.1029/2005JB003611.<br />

Thomas, A. L., <strong>and</strong> Pollard, D. D., 1993, The geometry <strong>of</strong> echelon fractures in rock:<br />

implications from laboratory <strong>and</strong> numerical experiments: Journal <strong>of</strong> Structural<br />

Geology, v. 15, p. 323-334.<br />

Townend, J., <strong>and</strong> Zoback, M. D., 2000, How faulting keeps the crust strong: Geology, v.<br />

28, no. 5, p. 399-402.<br />

Underhill, J. R., <strong>and</strong> Woodcock, N. H., 1987, Faulting mechanisms in high-porosity<br />

s<strong>and</strong>stones; New Red S<strong>and</strong>stone, Arra, Scotl<strong>and</strong>, in Jones, M. E., <strong>and</strong> Preston, R.<br />

M. F., eds., Deformation <strong>of</strong> Sediments <strong>and</strong> Sedimentary Rocks: Geological<br />

Society <strong>of</strong> London Special Publications, v. 29, p. 91-105.<br />

Vajdova, V., <strong>and</strong> Wong, T.-f., 2003, Incremental <strong>propagation</strong> <strong>of</strong> discrete compaction<br />

b<strong>and</strong>s: Acoustic emission <strong>and</strong> micro<strong>structural</strong> observations on circumferentially<br />

notched samples <strong>of</strong> Bentheim s<strong>and</strong>stone: Geophysical Research Letters, v. 30, no.<br />

14, 1775, doi: 10.1029/2003GL017750.<br />

-, 2004, Permeability evolution during localized deformation in Bentheim s<strong>and</strong>stone:<br />

Journal <strong>of</strong> Geophysical Research, v. 109, p. B10406, doi: 10.1029/2003JB002942.<br />

Weertman, J., <strong>and</strong> Weertman, J. R., 1964, Elementary Dislocation Theory: New York,<br />

Macmillan Company, 213 p.<br />

Wen, X.-H., <strong>and</strong> Gomez-Hern<strong>and</strong>ez, J. J., 1996, Upscaling hydraulic conductivities in<br />

heterogeneous media: An overview: Journal <strong>of</strong> Hydrology, v. 183, p. 9-32.<br />

White, C. D., <strong>and</strong> Barton, M. D., 1999, Translating outcrop data to flow models, with<br />

applications to the Ferron s<strong>and</strong>stone: SPE Reservoir Evaluation & Engineering, v.<br />

2, no. 4, p. 341-350.<br />

Wilson, J. E., Goodwin, L. B., <strong>and</strong> Lewis, C. J., 2003, Deformation b<strong>and</strong>s in nonwelded<br />

ignimbrites: Petrophysical controls on fault-zone deformation <strong>and</strong> evidence <strong>of</strong><br />

preferential fluid flow: Geology, v. 31, no. 10, p. 837-840.<br />

Wong, T., Baud, P., <strong>and</strong> Klein, E., 2001, Localized failure modes in compactant porous<br />

rock: Geophysical Research Letters, v. 28, p. 2521-2524.<br />

213


Wong, T. F., David, C., <strong>and</strong> Zhu, W., 1997, The transition from brittle faulting to<br />

cataclastic flow in porous s<strong>and</strong>stones: mechanical deformation: Journal <strong>of</strong><br />

Geophysical Research, v. 102, p. 3009-3025.<br />

Woodcock, N. H., <strong>and</strong> Underhill, J. R., 1987, Emplacement-related fault patterns around<br />

the Northern Granite, Arran, Scotl<strong>and</strong>: Geological Society <strong>of</strong> America Bulletin, v.<br />

98, p. 515-527.<br />

Yeong, C. L. Y., <strong>and</strong> Torquato, S., 1998, Reconstructing r<strong>and</strong>om media, II: Threedimensional<br />

media from two-dimensional cuts: Physical Review E, v. 58, p. 224-<br />

238.<br />

Zoback, M. D., <strong>and</strong> Healy, J. H., 1984, Friction, faulting <strong>and</strong> "in situ" stress: Annales<br />

Geophysicae, v. 2, no. 6, p. 689-698.<br />

214

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

Saved successfully!

Ooh no, something went wrong!