mechanics of faulting - 國立臺灣大學 · 2017. 4. 26. · critical-taper wedge mechanics...

66
Mechanics of faulting Jyr-Ching Hu, Dept. Geosciences National Taiwan University http://www.sanandre asfault.org

Upload: others

Post on 27-Jan-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

  • Mechanics of faulting

    Jyr-Ching Hu, Dept. Geosciences National Taiwan University

    http://www.sanandre

    asfault.org

  • Strengths of active thrust-belt wedges & their basal

    detachments: directly determined from the covariation of

    surface slope a with detachment dip b , without strong

    assumptions about the specific strength-controlling

    Test: Niger delta thrust belt, the active Taiwan mountain

    belt, and the thrust that slipped in the M = 7.6 Chi-Chi

    earthquake

    Suppe, Geology 2007

  • Absolute fault & crustal strength from wedge tapers

    Basal detachments: exceedingly weak, with effective

    coefficients of friction (0.04–0.1) that are an order of

    magnitude less than most laboratory friction coefficients

    (0.6–0.85)

    Weak faults & strong crust: wedges are moderately

    strong internally, within the range of pressure-dependent

    strengths in deep boreholes

  • Frictional resistance = b x Weight

    The classic thrust-fault problem…

    The breakup Maximum length ~20 km

    Courtesy of John Suppe

  • Critical-taper wedge mechanics

  • Courtesy of John Suppe

  • Critical-taper wedge

    taper = ab

    Courtesy of John Suppe

  • Critical-taper wedge mechanics

    Courtesy of John Suppe

    Actively deforming fold-and-thrust belts & accretionary wedges:

    simultaneously at regional failure internally & along their base

    Mechanical equilibrium: between the critical taper α + β of a

    wedge & the strength of the wedge & its base, where α is the

    surface slope and β is the dip of the detachment

    taper = ab

    Davis, 1983

  • Wedge theory

    1 (1 )

    sin1 2(1 )

    1 sin

    f b b b

    f

    S gH

    C gH

    b a b

    Wedge theory: infer the magnitudes of strength parameters that

    are consistent with observed tapers, e.g., internal & basal friction

    coefficients (μ = tanφ, μb) & depth-normalized pore-fluid pressures

    (λ = Pf /ρgH).

    Mechanically homogeneous Wedge (Dahlen, 1990, equation 99):

    Sb & C: non-pressure-dependent parts of the fault and wedge strength H: thickness

  • Wedge theory

    Equations contain a number of average regional-scale fault

    & crustal strength parameters, but unfortunately have little

    direct constraint in actively deforming regions.

    Equations simpified:

    (1 )b b bF S gH

    Fault-strength terms: 2(1 ) sin 1 sinW

    C gH

    Wedge-strength terms

    1

    1

    f

    f

    F

    W

    b a b

    ( , )f F Wa b

    [1 − (ρf /ρ)]: ratio of the density of the overlying fluid (seawater or air)

    to the mean density of rock & is 1 for subaerial wedges & ~0.6 for

    submarine wedges

  • Wedge theory

    1

    1

    f

    f

    F

    W

    b a b

    (see Dahlen, 1990, equations 88, 90, 91, 97)

    F gH

    F: regional normalized basal shear traction

    1 3W gH

    W: normalized differential stress

  • Wedge theory

    If F & W are homogeneous then a & b are linearly related

    1 1f fF W

    W Wa b

    0 sba a b

  • Get strengths from co-variation of a & b

    1

    sW

    s

    0F Wab

    11

    f

    sW

    s

  • Application to active wedges

    Dry-sand wedges on a Mylar base,

    Two active geologic wedges, Taiwan and

    the Niger delta

    Approximate the assumption of large-scale

    homogeneity

    1. Approximate linear covariation of α and β

    2. Rather thick (H = 5 – 12 km)

  • Application to active wedges

    Not mechanically homogeneous: thin toes (H < ~1

    km) of active accretionary wedges such as the

    Nankai trough & Barbados show surface slopes α

    that decrease away from the toe, with no

    associated change in detachment dip β

    Have horizontal gradients in wedge strength, given

    the strong lateral variation in porosity, lithification, &

    hence cohesion, & probably fluid pressure.

  • Application to active wedges

    Basal coefficient of friction of F = μb = 0.27 & a wedge

    strength W = 1.9, which corresponds to a cohesionless

    internal friction of μ = tanφ = 0.57.

    predicts µ = 0.57

    measured µ = 0.58

    predicts µb = 0.27

    measured µb = 0.3

    Davis et al. (1983)

  • h=v

    Taiwan Main Detachment

    Carena et al.,

    geology

    2002

    Stepping down to deeper detachments to East…

    Courtesy of John Suppe

  • Linear regressions of taper measurements

    Carena et al. (2002)

    Bilotti & Shaw (2005)

    Deep-water

    compressive toe of

    the Niger delta

    Central Taiwan

  • Summary

    W = (σ1 − σ3)/ρgH based on the regression slopes &

    obtain similar results for both wedges.

    Taiwan gives W = 0.6 & the Niger delta gives W = 0.7

    Normalized basal shear traction F = σ /ρgH:

    F = 0.08 for Taiwan and F = 0.04 for the Niger delta.

    Observed ratio of fault strength to wedge strength F/W

    = σ /(σ1 − σ3):

    0.13 for Taiwan and 0.06 for the Niger delta.

    These results show that the basal detachments are

    exceedingly weak absolutely and relative to the wedge

    strengths.

  • Comparison with deep borehole data

    Borehole stress measurements

    SAFOD pilot hole: strong

    decrease with depth; suggesting

    that the measurements, which

    are at a depth of 1–2 km in

    granite, are still within the near-

    surface boundary layer in which

    cohesion dominates

    Cohesive strength C = ~46 MPa:

    A factor of four less than the

    borehole-scale cohesion estimated

    for the SAFOD pilot hole at 197–

    212 MPa (Hickman & Zoback,

    2004).

  • Comparison with deep borehole data Borehole stress measurements

    W* is relatively

    constant

    as a function of depth,

    indicating that the KTB

    region is dominated by

    pressure-dependent

    strength, with W* = 1.0

    ± 0.2 to a depth of 8 km.

    KTB borehole σ2 is vertical, whereas in compressive wedges

    σ3 is vertical

  • Constraint of a single taper

    1 ( )fF Wa a b 0 sba a b

  • Constraint of a single taper

    Courtesy of John Suppe

  • Wedge-strength constraints

    Courtesy of John Suppe

  • Courtesy of John Suppe

  • Thermal anomaly in post Chi-Chi boreholes

    Courtesy of John Suppe

  • Constraint of Chi-Chi thermal anomaly

    Tanaka et al. GRL 2006

  • Post Chi-Chi borehole stress measurements

    Hung et al. Tectonophysics

    2009

    W*=0.75-0.95

  • Constraint of borehole thermal anomaly

    Courtesy of John Suppe

  • Summary Upper bound on upper-crustal strength: Byerlee’s law (μ

    = 0.85) with hydrostatic pore-fluid pressures (λ = 0.4),

    then W ≤ 2.2 & F ≤ 0.21, which is a weak detachment

    Chinshui Shale detachment: exceedingly weak, & best

    estimate is in the range F = σ /ρgH = 0.07–0.11

    Chelungpu thrust ramp is even weaker based on shear

    tractions σ estimated from post Chi-Chi borehole

    thermal anomalies and W* observed by Tanaka et al.

    (2006), Hung et al. (2009), and Kano et al. (2006) (F* =

    σ /σn = 0.03–0.05).

  • Summary

    These extreme fault weaknesses are especially striking

    in light of the observation that the regional pore-fluid

    pressures surrounding the Chinshui Shale detachment

    and thrust ramp are hydrostatic (λ = 0.4) (Yue, 2007).

    Therefore, the static ambient Hubbert & Rubey (1959)

    fluid-pressure hypothesis is not the cause of the

    weakness of the Chinshui Shale detachment or thrust

    ramp.

    Furthermore, the wedge is strong in spite of the very

    weak thrust ramp within it, presumably because of the

    internal strength of the thrust sheets in bending

  • Why are the faults so weak & the crust so strong???

    W 1 3

    gH 0.6 .0.95

    F gH 0.040.09

    F

    W

    1 3

    0.04 0.15

  • Are pore-fluid pressures the solution to the weak fault problem???

    The classic Hubbert-Rubey hypothesis…

    F (1 )b

    where

    Pf gz

    need

    (1 ) 0.1 or

    0.9

  • The Chinshui shale detachment is above fluid-retention depth ZFRD

    …therefore not classic Hubbert & Rubey fluid-pressure mechanism

    Courtesy of John Suppe

  • Relationship between ZFRD and Hubbert & Rubey effect….

    (1 ) 0.6ZFRD

    Z

    need Z > 5ZFRD or ~10-15 km for Taiwan

    Courtesy of John Suppe

  • Chi-Chi earthquake…

    Yue, Suppe & Hung, 2005

    Earthquake slip is confined to geometric segments…

  • Coseismic folding in Chi-Chi earthquake…

    Fault bends must be the locus of crustal strength…

    Courtesy of John Suppe

  • Coseismic folding in Chi-Chi earthquake…

    Courtesy of John Suppe

  • Continual deformation of new rock along axial surfaces….

    Weak faults and strong crust…

    Courtesy of John Suppe

  • Coseismic folding in Chi-Chi earthquake…

    Fault bends must be the locus of crustal strength…

    Courtesy of John Suppe

  • Contrasting crustal strengths…

    Courtesy of John Suppe

  • Areas of very thick deforming sediments…

    Gulf of Mexico Niger delta Borneo Sumatra Nankai trough Cascadia Bangladesh/Myanmar Makran Gulf of Alaska New Zealand Taiwan

    Courtesy of John Suppe

  • Low strength of deep San Andreas fault gouge from

    SAFOD core

    David A. Lockner, Carolyn Morrow, Diane

    Moore & Stephen Hickman

    Nature, 472, 82–85, 2011

  • Weakness of the San Andreas Fault Zone

    Absence of a heat flow anomaly (Brune et al.,

    1969; Lachenbruch & Sass, 1980; Williams et

    al., 2004)

    Stress orientation across the fault (Zoback et

    al., 1987; Mount and Suppe, 1987),

  • Hypotheses

    Fault zone consists of clay gouge (Wu et al., 1975; Wu,

    1978; Wang et al., 1978), especially a montmorillonite rich

    clay gouge that has frictional coefficients as low as ~0.1

    (e.g., Wang and Mao, 1979; Chu et al., 1981; Carpenter et

    al., 2011; Lockner et al., 2011)

    Fault zone has a normal frictional coefficient but is

    dynamically weakened during earthquakes by shear heating

    & other physicochemical processes (e.g., Lachenbruch,

    1980; Di Toro et al., 2011)

    Frictional coefficient of the fault is ‘normal’, but high

    porepressure in the fault zone lowers the effective normal

    stress on the fault and thus its frictional resistance to sliding

    (Rice, 1992; Byerlee, 1990)

  • SAFOD: San Andreas

    Fault Observatory at

    Depth

    Study the physical &

    chemical processes

    controlling faulting &

    earthquake generation

    along an active, plate-

    bounding fault at depth

    SAFOD

  • San Andreas Fault

  • Fault Contact at 10,063 ft

    Highly Deformed Siltsone

    Granite Cobble

    Conglomerate

    Clay Gouge

    2.5 cm

  • Direct measurement

    of the processes

    that control

    earthquakes

    San Andreas Fault Observatory at Depth (SAFOD)

    North

    American

    Plate

    A 15-year effort of

    Mark Zoback,

    Steve Hickman,

    and Bill Ellsworth

    Steve Steve

  • Location of SAFOD site SAFOD site: located at the NW end of the rupture zone of the 1966

    and 2004 M 6 Parkfield earthquakes, in the transition between the

    creeping and locked sections of the SAF

    At surface, the fault is creeping

    at a rate of 1.8 cm/yr.

    Numerous earthquakes occur

    directly on the SAF at 3-12 km

  • Geophysical logs & generalized lithology from phase 2 of the SAFOD project

    SAF is a broad

    zone of

    anomalously low

    P- and S-wave

    velocity and

    resistivity

  • Why Parkfield? Transition between the locked portion of the fault to the SE & the segment of the fault to the NE where slip dominantly occurs by aseismic creep

  • Geologic cross-section parallel to the trajectory of the SAFOD borehole

    San Andreas Fault damage zone: SDZ, CDZ, & NBF

  • 56

    2007 => creeping strands:

    southwest deforming zone (SDZ)

    central deforming zone(CDZ)

    Depth:2.7 km,

    damage zone:200 m wide

  • Methods

    •1.6 m and 2.6 m fault gouge of 31 m of core

    •Powder(

  • X-Ray Diffraction

    SDZ & CDZ:

    porphyroclasts of serpentinite

    and sedimentary rock

    dispersed in a matrix of Mg-

    rich clays

  • X-Ray Diffraction

    CDZ:Sap (>60%)

    SDZ:Sap, Cor, Q , F

    The two gouge zones: product of

    shearing-enhanced metasomatic

    reactions between serpentinite &

    adjoining sedimentary rocks.

    Q = quartz

    Cc = calcite

    K = K feldspar

    Pl = plagioclase

    (albite?)

    Chl = chlorite

    Km = K - micas

    Srp = serpentine

    Sap = saponite

    Cor = corrensite

    SW side

    NE side

  • X-Ray Diffraction

    CDZ:Sap (>60%)

    SDZ:Sap, Cor, Q , F

    The two gouge zones: product of

    shearing-enhanced metasomatic

    reactions between serpentinite &

    adjoining sedimentary rocks.

    Q = quartz

    Cc = calcite

    K = K feldspar

    Pl = plagioclase

    (albite?)

    Chl = chlorite

    Km = K - micas

    Srp = serpentine

    Sap = saponite

    Cor = corrensite

    SW side

    NE side

  • Sample strength: in situ

    μ:coefficient of friction τ :shear stress σn :effective normal stress p:pore pressure

    σn = 122 MPa

    V = 1.15 μm/s

    Compositional

    change

    SW => NE

    SDZ & CDZ:

    0.13≦ μ ≦0.21

    Saponite (μ

    ~0.05)

  • Strength & sliding rate

    Serpentinite porphyroclast from

    the SDZ: a − b = +0.0004 ± 0.0014

    All other core samples have

    positive rate sensitivity:

    1. Outside the foliated gouge zone:

    +0.001  o => stable creep

  • Low frictional strength (µ  ≈ 0.15) of foliated gouge:

    (1) lack of an observed heat flow anomaly

    (2) maximum compressive stress oriented at a high angle to the fault trace

    (3) no evidence of pore pressure elevated in fault zone

    The positive dependence of strength on slip rate of the fault gouge material is consistent with deformation by creep rather than by earthquakes.

    Stable creep and low strength

    Boness & Zoback, 2006

    Argument

  • Stress state

    Tembe et al., 2009 77 °

    Depth : 2.7km

    τ = 17 MPa , σn = 122 MPa p = 27

  • Summary

    The laboratory strength measurements of the SAFOD

    fault core materials at in situ conditions, demonstrating

    that at this locality & this depth the San Andreas fault is

    profoundly weak (μ = 0.15) owing to the presence of

    the smectite clay mineral saponite, which is one of the

    weakest phyllosilicates known.

    This Mg-rich clay is the low-temperature product of

    metasomatic reactions between the

    quartzofeldspathic wall rocks and serpentinite blocks in

    the fault.

    Deformation of the mechanically unusual creeping

    portions of the San Andreas fault system is controlled

    by the presence of weak minerals rather than by high

    fluid pressure or other proposed mechanisms

  • References 11. Zoback, M. D., et al. (1987), New evidence on the state of stress of

    the San Andreas Fault, Science, 238, 1105–1111.

    2. Scholz, C. H. (2000), Evidence for a strong San Andreas fault,

    Geology, 28, 163– 166.

    3. Zoback, M. D. (2000) Strength of the San Andreas. Nature 405, 31–32

    4. Carpenter, B. M., C. Marone, and D. M. Saffer (2011) Weakness of the

    San Andreas Fault revealed by samples from the active fault zone,

    Nature Geoscience, doi: 10.1038/ngeo1089.

    5. Wang, C-Y (2011) High pore pressure, or its absence, in the San

    Andreas Fault, Geology, 39, 1047-1050, doi: 10.1130/G32294.1.

    6. Zoback, M., Hickman, S. Ellsworth, W. and the SAFOD Science Team

    (2011) Scientific Drilling Into the San Andreas Fault Zone — An Overview

    of SAFOD’s First Five Years, doi:10.2204/iodp.sd.11.02.2011

    7. Collettini, C., Niemeijer, A., Viti C. & Marone, C. (2009) Fault zone

    fabric and fault weakness, Nature 462, 36,