9. overburden rock, temperature, and heat flow_d. deming

Upload: andre-jhosep

Post on 06-Jul-2018

222 views

Category:

Documents


1 download

TRANSCRIPT

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    1/22

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    2/22

      66 eming

    2

    3

    ~ 4

    w

    ~

    w

    0 2

    ~

    >

    E 3

    CIJ->::

    4

    5

    6

    1

    m/m.y.

    AGE Ma)

    2

    1

    AGE Ma)

    200

    00

    0 ~ _ . ~ ~ r ~ ~ ~ ~ ~

    00

    0

    ~

    1

    2

    3

    4

    5

    SSIVE

    M RGINS

    ~ ~ ~ ~ ~ ~ = = ~ = = ~ = = ~

    600

    5

    5

    AGE

    Ma)

    400

    .;;;;;;;::

    Williston Basin

    INTR CR TONIC B SINS

    400

    3

    2

    3

    ~

    4

    3

    5

    6 ~ ~ = ~ = : ~ = = ~

    5

    6

    ~

    2

    3

    4

    5

    6

    CE

    FOREL ND

    B SINS

    9 ~ ~ ~ 1 0 0

    500

    4

    2

    3

    ~

    5

    6

    AGE Ma)

    Figure

    9.1.

    Representative tectonic subsidence histories for basins from different tectonic settings. The top graph shows the

    slopes of a

    range

    of sedimentation rates after compaction and is provided for reference. After Angevine et al.,

    1990.)

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    3/22

      100-50 m/m.y.) to a

    passive

    margin basin 20-10

    m/m.y.). The lowest sedimentation rates -10 m/m.y.)

    are found in intracratonic basins such as the Michigan,

    Illinois, and Williston basins in North America Sleep,

    1971; Schwab, 1976; Sleep et al., 1980). Strike-slip and

    forearc basins are characterized by much higher rates

    1000-100

    m/m.y.). Foreland basins experience the most

    varied sedimentation rates, but generally occupy the

    middle ground. The highest sedimentation rates are

    found in areas of rapidly prograding river deltas e.g.,

    U.S. Gulf Coast basin), where sediment deposition can be

    as much as 1000-5000

    m/m.y.

    Sharp and Domenico,

    1976; Bethke, 1986; Bredehoeft et al.,

    1988).

    Geothermal gradients in sedimentary basins also vary

    widely, from

    as low

    as 10 -15°C/km to as high as

    50°-60°C/km. Part of this variation can be attributed to

    differences in the background thermal state of the crust

    on which the basin rests. However, the thermal proper

    ties of sediments e.g.,

    thermal

    conductivity) and

    physical processes acting within basins

    e.g.,

    sedimenta

    tion and groundwater flow) are also important determi

    nants.

    Why do

    some

    basins

    accumulate

    sediment

    much

    faster than others? What controls temperature in sedi

    mentary basins

    and its variation between and within

    basins? How does sedimentation itself affect the thermal

    regime? The purpose of this chapter

    is

    to address these

    questions by describing why and how sedimentary

    basins form and the physical properties and processes

    that control temperature within them.

    FORMATION

    OF

    SEDIMENTARY BASINS

    Sedimentation and Subsidence

    A

    sedimentary

    basin is any downwarped area of the

    continental or oceanic crust where sediments accumulate

    and compact

    with

    burial into sedimentary rock. The

    accumulation and removal of these rocks defines the life

    cycle of a basin, from the initial event that creates the

    basin through senescence, culminating in eventual uplift

    and destruction.

    A sedimentary basin forms when a topographic low is

    created in the basement rock through either tectonic

    subsidence or sedimentation subsidence, or both. Sedi-

    mentation

    subsidence

    can be defined as the

    downward

    movement of

    the

    basement rock-sedimentary rock

    contact in response to sediment loading e.g., a major

    river delta), while tectonic subsidence

    is

    the subsidence

    of

    basement rock that occurs, or would occur, in the

    absence of sedimentation e.g., the deep ocean basins).

    In general, both tectonic subsidence and sedimenta

    tion are necessary for the creation of a sedimentary basin.

    Sediments accumulate only in topographic lows, thus a

    basin must generally exist before the fill. Conversely,

    sedimentation reinforces the tectonic subsidence that was

    initiated

    by

    a basin-forming event. The load due to accu

    mulated sediments is capable of increasing total basin

    depth

    by

    a factor of

    two

    or three Turcotte,

    1980).

    The

    9. Overburden

    Rock

    Temperature and

    eat Flow

    167

    sediment tion

    Pm

    >

    Pc

    >Ps

    sedimentary

    rock

    crust

    mantle

    Figure

    9 2

    Schematic illustration

    of

    isostatic subsidence

    following crustal thinning and sedimentation Terms are as

    follows:

    c

    = crustal density; Ps = sediment density; Pm =

    mantle; h depth of compensation

    relative importance of tectonic subsidence and sedimen

    tation as driving forces for the creation of sedimentary

    basins varies according to the circumstances involved.

    Major river deltas e.g., Mississippi, Amazon, and Niger)

    are primary examples in which sedimentation itself plays

    a major role

    in

    forcing subsidence and increasing the

    depth of a basin.

    On

    the opposite extreme, the abyssal

    plains of

    the

    oceanic basins are relatively

    sediment

    starved. They

    owe

    their existence to the cooling and

    subsidence of oceanic lithosphere as it moves away from

    the site of its creation at a mid-oceanic spreading ridge;

    sedimentation is limited and plays an insignificant role in

    determining total subsidence.

    Isostasy and Flexure

    What controls the subsidence of a sedimentary basin?

    l f we assume that the lithosphere has no lateral strength,

    the principle of isostasy applies. Isostasy is the funda

    mental principle governing the

    development and

    evolution of topography

    on the

    earth s surface. A

    succinct

    mathematical

    statement

    of

    isostasy is

    that

    density

    p)

    integrated

    over

    an

    imaginary column

    extending from the surface to the depth of compensation

    remains constant:

    hp z = constant

    1)

    where z is

    depth and

    h

    the

    depth of compensation,

    commonly taken as near the base of the lithosphere, or

    about

    100 km

    Figure

    9.2).

    More simply stated, equation

    1 is a mass balance equation. The total mass of material

    in

    a column

    between the surface and

    the depth

    of

    compensation must be constant. If the mass or weight)

    of the column increases, the column must sink, or isostat

    ically subside. As

    the column

    sinks, relatively h igh

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    4/22

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    5/22

    McKenzie's model is often applied (and misapplied)

    to estimate the timing of hydrocarbon generation. The

    usual procedure is to ''backstrip sedimentary basin fill

    for the purpose of separating tectonic subsidence from

    the total subsidence. This is done by

    applying

    the

    principle of isostasy and compensating

    for

    factors such

    as

    sediment compaction

    and

    changes in

    sea level

    (Steckler and Watts,

    1978;

    Sclater

    and

    Christie, 1980;

    Sclater et al., 1980). The estimated tectonic subsidence

    curve is then compared to McKenzie's theoretical predic

    tions and a ''best value for the stretching factor found.

    Once

    is

    known, heat flow can

    e

    estimated, tempera

    ture calculated, and source rock maturity

    predicted

    (provided that the location of the source rock in the basin

    fill

    is known). This is a straightforward approach, but

    there are many ancillary determinants that must also be

    taken into consideration

    if

    meaningful estimates of the

    thermal history are to be made. These

    include the

    depression of heat flow by

    sedimentation (De

    Bremaecker, 1983), the thermal conductivity of rocks

    within the basin (Blackwell and Steele, 1989), the surface

    temperature, and the possible influence of groundwater

    flow. The relative importance of these intrabasin factors

    grows with passing time as the influence of the initial

    basin-forming event wanes.

    Intracratonic asins

    Intracratonic, or platform, basins form on continental

    interiors e.g., the Michigan, illinois, and Williston basins

    of North America; Figure 9.1). They are typically a few

    hundred kilometers wide and contain a few kilometers

    of flat-lying sedimentary rocks recording continuous

    subsidence and sediment deposition over periods of time

    greater than

    100

    m.y. (Sleep et al., 1980). Sleep 1971) was

    the first to note that the subsidence of these basins was,

    like oceanic basins, proportional to the square root of

    time, with a time constant of about 50 m.y. This led to

    speculation that the formation of these basins, like rift

    basins, was

    controlled by some

    type of heating

    or

    thermal event followed by thermal contraction (Sleep,

    1971; Sleep and Snell, 1976; Ahern and Mrkvicka, 1984;

    Nunn

    et al., 1984; Klein and Hsui, 1987).

    For an intracratonic basin to be formed by thermal

    contraction, isostasy requires that a considerable amount

    of crustal erosion occur during the initial heating, uplift,

    and thermal expansion phase. For example, i the basin

    fill is 3 km deep, it would be necessary to first remove

    about 1 km of the continental crust

    through

    erosion.

    However, in many instances, there is little evidence that

    this type of dramatic erosion ever occurred (Sleep et al.,

    1980). Recognition of this problem has led to the

    proposal of several alternative hypotheses. These include

    (1) an increase in density of the crust due to one or more

    phase transitions,

    2)

    rifting,

    3)

    mechanical subsidence

    caused by an isostatically uncompensated excess mass of

    igneous intrusions,

    4)

    tectonic reactivation along older

    structures, or

    5)

    some combination of these or other

    theories

    see

    review by Klein,

    1991).

    For example, Klein

    (1991)

    suggests

    that intracratonic

    basins in

    North

    9. Overburden Rock Temperature and Heat Flow 69

    America initially underwent fault-controlled mechanical

    subsidence in response to rifting. The initial phase of

    basin formation was followed by thermal subsidence

    and subsidence due to the isostatically uncompensated

    mass of a cooled igneous intrusion.

    Although the subsidence history of intracratonic

    basins is apparently consistent with some

    type

    of thermal

    mechanism, the exact nature of the initial thermal event,

    its subsequent evolution, and the role of other factors in

    basin genesis and development are apparently not well

    understood at the present time.

    oreland asins

    Foreland basins (Beaumont, 1981) are asymmetric,

    wedge-shaped accumulations of sedimentary rock that

    form adjacent to

    fold thrust

    belts. Migration of

    the

    fold-thrust sheet loads the lithosphere, causing isostatic

    subsidence

    underneath the

    core of

    the

    orogen

    and

    flexural

    downwarping in

    the adjacent foreland. The

    foredeep that forms next to the orogenic belt rapidly

    fills

    with sediment eroded from the adjacent mountains.

    Sedimentation

    amplifies flexural subsidence, and a

    foreland basin is formed (Figure 9.1).

    The foreland basin process continues until the forces

    driving uplift and orogeny

    cease. Erosion

    then

    dominates, reducing the weight of the mountain chain,

    leading to uplift and further erosion. The life cycle of a

    foreland basin is thus typically one of fairly rapid burial

    and

    subsidence followed

    by

    a much longer period of

    uplift and erosion. Most source rocks buried by the

    foreland basin

    fill

    probably

    go

    through a relatively short

    heating

    and maturation

    phase, followed

    by

    a longer

    cooling phase.

    Thermal events play a minor role in the formation of

    foreland basins. However, the thermal state of the lithos

    phere influences its flexural strength, thereby exerting an

    indirect control

    on

    the structural evolution of foreland

    basins (Watts et al.,

    1982).

    Other Types of asins

    Many other

    types

    of basins can

    e

    defined; these

    types

    are potentially as numerous as the heterogeneous crust

    of the earth. Some of these include strike-slip, forearc,

    and backarc (Figure

    9.1).

    Strike-slip or pull-apart basins

    are formed by lateral movement along transform faults,

    literally pulling the crust apart and creating a void that

    fills with sediment e.g., the Los Angeles basin) (Turcotte

    and Ahern, 1977; Turcotte and McAdoo,

    1979).

    Backarc

    and

    forearc basins

    form

    in back

    of and

    in

    front

    of

    volcanic arcs, respectively, near subduction zones.

    Backarc basins may form from active seafloor spreading

    and rifting, in which case they exhibit high heat flow. In

    other

    cases,

    backarc basins

    are

    apparently

    passive

    features that may merely represent trapped segments of

    old oceanic crust.

    Forearc basins

    are the result of

    sediments filling the topographic low created by

    subduction.

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    6/22

    17 Deming

    STRUCTUR L ND THERM L

    EVOLUTION OF SEDIMENT RY

    B SINS

    Are the structural and thermal evolution of sedimen

    tary basins linked? In some cases the answer is yes, but

    on the scale of a petroleum system, this

    fact

    may have

    little utility in attempts to understand the temperature

    history of the basin

    fill.

    For example, the formation of

    rift

    basins is well understood through relatively simple theo

    retical models that invoke an initial extensional event. In

    these cases, it is possible to demonstrate with a fair

    amount of confidence a link between the thermal

    and

    structural evolution of these basins. However, this may

    be

    of limited relevance in estimating temperature of the

    basin fill at a time when

    hydrocarbons

    are being

    thermally generated. The magnitude

    of

    the initial

    thermal event associated with the creation of a rift basin

    decays with passing time (Figure 9.3). Thus, by the time

    sufficient overburden accumulates for hydrocarbon

    maturation to begin, the influence of the initial basin

    forming thermal event may be

    relatively

    small

    in

    comparison to other factors.

    An

    example of a

    rift

    basin in which the initial thermal

    event had little influence on the maturation of hydrocar

    bons

    is

    the Gulf Coast basin

    of

    the southeastern United

    States. This basin formed by rifting in Late Triassic-Early

    Jurassic time ( 180 Ma), but it was relatively sediment

    starved up to about

    40

    Ma. Rapid accumulation of

    sediments since that time has increased the burial depth

    and temperature of the source rocks. Cretaceous and

    early Tertiary age source rocks are estimated to presently

    be in the oil generation window Nunn and Sassen,

    1986). Because the thermal anomaly associated with the

    rifting of the Gulf Coast basin has been decaying for the

    last

    180

    m.y., the degree to which the lithosphere was

    extended or rifted has a negligible influence on the

    present-day thermal state (Figure 9.3). Factors such as

    lateral

    variations in

    overburden thickness and

    the

    depression of heat flow by sedimentation have a greater

    influence on source rock temperature. For example,

    Nunn and Sassen 1986) estimate that present-day heat

    flow in the Gulf Coast basin is depressed

    -30

    below its

    equilibrium value by high rates of sedimentation.

    On the scale of the petroleum system, the influence

    of

    initial basin-forming thermal events

    is

    thus of indirect or

    limited importance in determining temperature of the

    basin

    fill

    at

    the

    time

    hydrocarbons

    are

    generated.

    Temperature of the sedimentary basin

    fill is

    more likely

    to be sensitive to intrabasin factors such as thermal

    conductivity,

    groundwater

    flow, sedimentation,

    and

    surface temperature (Table

    9.1).

    The following sections

    discuss the importance of these four factors in more

    detail.

    M THEM TIC L DESCRIPTION OF

    HE T TR NSPORT

    Sedimentary basins are never in complete thermal

    equilibrium, and groundwater flow may drastically

    change the distribution of thermal energy within a basin.

    Table 9 1 Factors Determining Temperature in

    Sedimentary Basin Fill

    Importance

    Factor

    Order)

    Qualifications

    Overburden thickness 1st

    Always important

    Heat flow 1st Always important

    Thermal conductivity 1st

    Always important

    Surface temperature

    2nd

    Always important

    Sedimentation

    1st

    >100 m/m.y.

    2nd

    100 m/m.y.

    3rd

    60 Ma)

    Nevertheless, steady-state conductive heat transport is a

    useful first order approximation that provides a starting

    point from which one may later consider departures.

    Fourier's law of heat conduction is

    q=kg

    2)

    where

    q is

    heat flow, k is thermal conductivity, and

    g is

    the thermal gradient. Applying this to the analysis of

    temperature within sedimentary basins, we obtain

    T

    =

    T

     

    +

    q/k)

    dz:

    3)

    where T

    is subsurface

    temperature, T

    is the mean

    annual surface temperature, and ru

    is

    thickness of the

    overburden. Thus, heat flow, thermal conductivity, and

    overburden thickness are of equal importance in deter

    mining subsurface temperature. However, heat flow is

    generally a more useful measure of the thermal state

    of

    sedimentary basins than temperature gradient alone

    because the geothermal gradient,

    g = q/k),

    varies

    according to thermal conductivity, which can change by

    as much as factor of three or four among common rock

    types.

    A more generalized description of heat transport can

    be obtained by considering departures from steady-state

    conditions and including advection of heat by moving

    fluids. The change of temperature with respect to time

    )T j )t)

    is then described by

    pC )T

    /iJt)

    =d/dz[kz )T /iJz)]- VzPwCw )T /dz) +A

    4)

    where z is depth, p and C are the bulk density and heat

    capacity, respectively, of a porous rock,

    Pw

    is fluid

    density, Cw

    is

    fluid heat capacity, Vz

    is

    the Darcy velocity

    of a fluid moving

    through

    a

    porous

    medium, k

    2

    is

    thermal conductivity, and

    A

    is

    radioactive heat genera-

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    7/22

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    8/22

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    9/22

    9. Overburden Rock Temperature and Heat Flow

    173

    100

    a)

    80

    ATI

    ~ ~ H I N K P J

    - T

    1----

      ~

    ----  r _l_

    NIN

    I ETK DRP

    60

    SBE KOL I

    ----r---- IKP

    40

    20

    LBN

    0

    A b)

    A

    SOH ETK JWD

    SBE KOL

    IN

    NIN FCK NKPATI DRP

    0

    E

    2

    :::.

    --

    -

    i

    3

    so

    75

    >

    -

     

    tO

    Q

    r

    :1:

    ---

     

    Q

    5

    .0

    -

      160

    c

    Q.6Q

    50 km

    l

    7

    2 0 0

    B

    •. ,

    Figure

    9.6.

    Estimated

    (a) shallow heat flow (in mW/m2)and (b) subsurface temperature (in 'C) from the North

    Slope

    basin,

    Alaska . Dotted lines

    are stratigraphic

    units

    of cross

    section;

    solid lines are

    isotherms.

    Numbers on trace of well bore are

    corrected bottom-hole temperatures. Abbreviations for wells:

    A

    Tl, Atigaru Point

    1;

    DRP,

    Drew

    Point 1; EKU, East

    Kurupa

    1; ETK,

    ast

    Topagoruk 1;

    FCK,

    Fish Creek 1; IKP, lkpikpuk 1; INI, lnigok

    1;

    JWD, J. W. Dalton

    1;

    KOL,

    Koluktak 1;

    LBN,

    Lisburne

    1; NIN,

    North

    lnigok

    1;

    NKP, North

    Kalikpik 1;

    SBE, Seabee

    1;

    and SOH, South

    Harrison

    ay 1. After

    Deming et al., 1992.)

    For the North Slope basin, Deming et al. 1992) used

    the

    method

    of variable bias a conceptually simple

    algorithm designed to extract the maximum amount of

    information

    from

    the data

    while

    simultaneously

    averaging the noise. The method involves the sequential

    estimation of temperature at different spatial locations

    through a series of weighted least-squares regressions.

    Based on an estimate of the magnitude of error in the

    data, a decision

    is

    made that n

    BHTs

    are to be averaged

    through an interpretive model. For each point at which

    temperature

    is

    to be estimated, the algorithm searches

    through

    three-dimensional space until

    it

    locates the

    closest

    n

    BHTs.

    These are given substantial weighting in

    the regression analysis; distant data are given much

    lower weightings. Instead of all of the data from a basin

    being

    averaged simultaneously

    only data in the

    immediate vicinity of the estimation point are averaged.

    If data density is locally high, local features of the

    temperature field are thus resolved.

    If

    data density is

    low, it is impossible to resolve detailed features of the

    temperature field, and the data are averaged over a

    wider area. Thus, the balance

    between the need

    to

    resolve the temperature field and the need to reduce

    noise

    by

    averaging is largely determined by the data

    themselves. A complete description of this method is

    given by Deming et al.

    1990a).

    Figure 9.6 shows

    temperature

    in the North Slope

    basin estimated from the method of variable bias along a

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    10/22

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    11/22

    0

    TERTIARY

    ..

    .

    .

    I

    I

    ..

    .

    GANNET

    I

    .

    •I

    ..

    .,

    .

    PREUSS

    .:

    2

    - . .

    I , •.. .••

    I . - .. •

    TWIN CREEK

    E

    NUGGET

    --·.....1 ----1

    .

    .¥.

    I

    ..

    r-:-..;-•-- . J

    .

    .

    NKAREH

    ·

    . .

    L--- •1

    ••

    ••

    J:

    3

    I

    ...

    ·

    I

    THAYNES

    I •

    .

    .

    a..

    L •

    LLJ

    ..-7·

    0

    WOOD.

    DJNW

    I ¥

    PHOSPHORIA

    L __ . . ·

    .

    .

    4

    • •

    : 1·

    ,

    I

    I

    I

    .

    WE ER

    I

    .

    I

    .

    I

    .

    I

    5

    I•

    I

    r

    t

    1

    ..

    MADISON

    .

    .

    ·· ··

    I

    I

    I

    I 2 3 4

    6

    7

    THERMAL CONDUCTIVITY (W

    /m

    K)

    Figure

    9.7. Thermal conductivity data,

    Anschutz

    Ranch

    well

    34-02, Utah-Wyoming

    thrust belt. Dots

    are matrix

    conduc

    tivities

    measured

    at 20°C in the laboratory; dashed lines

    are estimated n

    s tu

    thermal conductivity. After Deming

    and

    Chapman,

    1988.)

    data measured on samples from the Anschutz Ranch 34-

    02 well in the Utah-Wyoming thrust belt Deming and

    Chapman,

    1988).

    The discrete points represent matrix

    conductivities measured in the laboratory on drill chips;

    the dashed lines show estimated in situ thermal conduc

    tivities. The in situ estimates are lower than the labora

    tory measurements due to the effects of porosity

    and

    temperature, and range from about 2 to 4 WIm K. The

    wide scatter of

    measurements

    for

    any

    formation is

    partially due to errors in measurement, but most of the

    scatter can be attributed to changes in lithology

    and

    mineralogy.

    The number of measurements needed to determine

    the average thermal conductivity of a geologic unit to an

    acceptable level of precision depends on its lithologic

    heterogeneity. For some marine Paleozoic units that are

    lithologically uniform over hundreds of kilometers, it

    may be possible to make only 10-20 measurements for

    an entire basin. However, large spatial variations in

    thermal conductivity are more typical because most sedi

    mentary rocks tend to have facies changes that occur

    both vertically and laterally. It is therefore difficult in

    most cases to collect enough data to estimate how the

    thermal

    conductivity of a geologic unit changes

    throughout a basin.

    To overcome this difficulty, concerted efforts have

    been made to

    estimate

    thermal conductivity from

    9.

    Overburden Rock Temperature and Heat Flow 75

    geophysical well logs. In many instances, strong correla

    tions have been found between thermal conductivity and

    one or more log parameters such as resistivity, seismic

    velocity, and density Houbolt and Wells, 1980; Reiter et

    al., 1980; Vacquier et al., 1988; Blackwell and Steele,

    1989). In other cases, mineralogy has been estimated

    from well logs, and the thermal conductivity of the bulk

    rock

    estimated

    from

    laboratory-derived

    values for

    different mine ralogies Brig a

    ud

    and Vasseur, 1989;

    Brigaud et al., 1990; Demongodin et al.,

    1991).

    The limita

    tion of all of these methods is the lack of an accurate

    mineralogy log. Matrix thermal conductivity

    is

    deter

    mined by mineralogic composition; correlations and

    inferences found to be valid in specific instances cannot

    be generalized. Thus, at the present time, there is no

    simple algorithm for estimating thermal conductivity

    from well logs that

    is

    demonstrably accurate. However,

    well logs may prove useful in interpolating between

    measurement sites when log parameters can be cali

    brated by laboratory measurements.

    CONTROLS ON TEMPERATURE IN

    SEDIMENTARY BASINS

    Heat Flow and Thermal Conductivity

    Because the primary mode of heat transport in the

    crust is conduction, both heat flow determined from

    equation 2) and thermal conductivity measured

    directly) are of first-order and equal importance in deter

    mining temperature in sedimentary basin fill.

    Heat

    flow is inversely correlated to tectonic age

    Vitorello

    and

    Pollack, 1980; Morgan, 1984)

    and

    is

    depressed by sedimentation see later discussion). Heat

    flow in young

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    12/22

      76 Deming

    G

    0

    + ~

    ..

    .

    +

    1000

    2000

    Elevation m)

    3000

    Figure

    9.8.

    Mean annual air temperatures +) from rnetero

    logic data collected by author), and mean annual ground

    temperatures •) estimated from extrapolation of borehole

    temperature logs from the north-central Colorado

    Plateau.

    Made

    by

    Bodell and Chapman, 1982.)

    Surface Temperature

    Although it

    is

    often ignored, surface temperattli'e T

    0

    )

    is

    an important boundary

    condition on geothermal

    conditions. The

    temperature

    at the

    earth's

    surface is

    determined by climate and has diurnal and annual cycles

    that rapidly attenuate in the subsurface.

    By

    applying

    equation 5, it can be seen that the annual variation of

    temperature propagates no deeper than about

    10m

    into

    the subsurface. Thus, the quantity of interest in geot

    hermal studies (T ) is a long term

    mean,

    a fictional

    quantity that

    is

    usually estimated by the linear extrapola

    tion of a borehole temperature log to the surface. Obser

    vations have shown that extrapolated borehole tempera

    tures are closely related to mean annual air temperatures,

    but that

    ground

    temperatures are

    always

    higher

    by

    about 2 -3°C (Figure 9.8). This discrepancy is commonly

    attributed to the insulating effect of

    snow cover

    in

    winter. However,

    the

    offset

    between

    mean annual

    ground and air temperatures is also found at low latitude

    sites (e.g., Howard

    and

    Sass, 1964). Mean

    annual

    air

    temperature on

    the

    earth's surface is -16°C, varying

    from -25°C at the equator, to --22°C at the poles (Gross,

    1993). Air temperatures also decrease with elevation;

    lapse rates typically range from -4 to -10°C/km.

    Air and ground temperatures vary not only spatially

    but also have short

    and

    long term temporal trends. From

    about 1400 to 1900, air temperatures were about o.soc

    colder than present day (the little ice age ). Before that,

    from

    about

    1000 to 1400 A.D., there was a Medieval

    warm period when air temperatures were about O SOC

    higher than present day . Over the past 1 m.y., tempera

    tures have fluctuated about 5-6°C as the glaciers

    retreated and advanced in a series of ice ages. The last

    such ice age ended about

    10,000

    yr ago as temperatures

    rose

    5--6°C,

    coincident with the emergence of civilization

    (Folland et al., 1990). Since Late Cretaceous time 70 Ma),

    k, >k,

    3:

    k,< k,

    5

    Time

    0

    sediment

    u::

    c

    ii

    15

    /

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    13/22

    9.

    Overburden Rock Temperature and Heat Flow 77

    -

      .

    E

    -

     

    10

    2

    -

    o

    J

    0

    .> .>

    lo..

    .

    t

    q

    .)

    ::: :l

    -

    t

    c:

    v

    Q:

    c

    0

    g:

    -

    10

    3

    c.

    Q

    ..,;;;

    '10-14 m2) aquifers and conspicuous signs of under

    ground flow e.g., artesian wells) are not a prerequisite.

    In areas of high relief and rugged topography, the

    presence of groundwater flow is nearly ubiquitous,

    making it difficult to obtain accurate estimates of back

    ground thermal conditions in these locations.

    Groundwater

    moves

    1) in response to potential

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    14/22

      78 eming

    gradients Hubbert,

    1940)

    or

    2)

    as a result of free convec

    tion. Strictly speaking, it is impossible to

    define

    a

    hydraulic potential for a fluid whose density is variable.

    However, the use of a hydraulic potential

    or

    pseudopo

    tential is a useful concept

    that

    can

    be used

    to obtain

    insight into geologic problems for which an exact answer

    is unobtainable.

    Common geologic mechanisms for creating potential

    gradients

    are

    sediment

    compaction and elevation

    gradients. Groundwater

    flow

    driven

    by sediment

    compaction

    and

    pore collapse, however, is a relatively

    inefficient mechanism for heat and mass transport unless

    pore fluids are focused spatially or temporally Cathles

    and Smith, 1983; Bethke, 1985; Deming et al., 1990b).

    Flow velocities tend to be very low,

    and

    the total amount

    of water is limited to the water contained in the original

    sediments. Few direct observations of thermal anomalies

    have been ascribed to compaction-driven flow. Bodner

    and Sharp 1988) speculated tha t relatively high geot

    hermal gradients associated with fault zones in the Gulf

    Coast basin in south Texas may be due to the upward

    flow of pore water along the faults. However, they were

    unable to rule out alternative explanations for the high

    thermal gradients, such as changes in thermal conduc

    tivity.

    In

    contrast

    to

    compaction-driven

    flow,

    regional

    groundwater flow over distances of 100-1000

    km

    due to

    potential gradients arising from elevation differences has

    been documented for several sedimentary basins. There

    is virtually no doubt

    that such flow systems exist e.g.,

    Bredehoeft et al., 1982). Correspondingly, the thermal

    regime must be disturbed, with the nature of the thermal

    disturbance depending upon the

    depth

    and velocity of

    groundwater

    flow. In foreland basins,

    the

    following

    pattern is typical. In the foothills of the mountain range

    where water infiltrates at high elevations, the geothermal

    gradient and surface heat flow are depressed as heat is

    advected downward by moving groundwater. Near the

    midpoint

    of

    the basin

    axis of

    the

    basin fill), flow is

    largely horizontal and the effect

    on

    basin temperature is

    minimal. At the distal edge of the basin, flow is forced

    upward by the basin geometry, leading to a high geo

    thermal gradient

    and

    high surface heat flow. This pattern

    has been

    observed in

    the

    Western

    Canadian basin

    Majorowicz and Jessop, 1981; Hitchon, 1984;

    Majorowicz,

    1989);

    the Kennedy, Denver, and Williston

    basins of the Great Plains province of the central United

    States Gosnold,

    1985, 1990);

    the Uinta basin in western

    United States Chapman et al.,

    1984);

    the Great Artesian

    basin in Australia Cull and Conley,

    1983);

    and the North

    Slope basin in Alaska Deming et al., 1992). The thermal

    anomalies

    associated with these

    flow systems

    can

    dramatically influence the

    temperature-dependent

    generation of oil and gas, and the flow systems them

    selves may play a role in oil

    and

    gas migration Toth,

    1988;

    Carven,

    1989;

    Bethke et al.,

    1991;

    Meissner, 1991).

    Free convection in sedimentary basins

    may

    conceiv

    ably

    arise from density gradients

    due to

    thermal

    expansion or the presence of solutes. For free convection

    to occur, the permeability of the porous medium must be

    sufficiently high and a density inversion must exist, with

    higher density fluid overlying less dense. In most sedi

    mentary basins, however, free convection is probably

    inhibited because the increase of salinity and density)

    with depth overshadows the decrease in density due to

    increased

    temperature

    and concomitant

    thermal

    expansion.

    Relatively little is

    known about the

    occurrence

    or

    significance of free convection in sedimentary basins;

    there are no direct observations of the process. However,

    the presence of extensive quartz cementation in sand

    stones found in the geopressured zone of the Gulf Coast

    basin of the southeastern United States requires

    much

    higher fluid volumes than could possibly be supplied

    by

    pore

    water

    alone Land, 1991). One solution to this

    discrepancy would

    be

    to invoke free convection within

    compartmentalized cells in

    the geopressured

    zone.

    Within each cell, the pressure gradient would be hydro

    static and fluid would be free to circulate Cathles,

    1990).

    n

    interesting new hypothesis

    by Nunn 1992)

    attributes

    episodic subsidence in the Michigan basin to the cata

    strophic release

    of

    heat by periodic

    episodes

    of free

    convection in the upper 10

    km

    of the continental crust, as

    originally envisaged by Deming 1992). At the present

    time, however, the possible occurrence of free convection

    in sedimentary basins and the continental crust remains

    speculative. For example, little

    is

    known about the extent

    to which fracture permeability exists in the crust and

    what

    stress states and tectonic processes could create

    sufficient permeability to allow for free convection.

    THERM L

    HISTORY

    OF A WELL FROM

    NORTH

    SLOPE

    BASIN ALASKA

    The importance of some of the factors that determine

    temperature in sedimentary basins and thus source rock

    maturation in the petroleum system) can be illustrated

    by considering

    an

    example from the North Slope basin,

    Alaska.

    During the years

    1977-1984,

    the

    U.S. Geological

    Survey drilled

    28

    deep 1-6

    km)

    petroleum exploration

    wells in the North Slope basin. See

    Gryc,

    1988;

    Tailleur

    and Weimer,

    1987;

    and Bird, Chapter

    21,

    this volume, for

    comprehensive discussions of the exploration history

    and regional geology.) A large amount of geologic and

    geophysical data were collected as part of this drilling

    program. Thermal

    data

    include thermal conductivity

    measurements on core

    and

    drill chip samples Deming et

    al., 1992), equilibrium ±0.1°C) temperature logs in the

    upper sections

    0--600

    m, average) of 21 of

    28

    boreholes

    Lachenbruch et al., 1987, 1988), and

    23

    reliable ±3°C)

    estimates of formation temperatures at depths of about

    1-4.5

    km

    obtained from extrapolation of series of BHTs

    measured during geophysical logging

    runs

    Blanchard

    and Tailleur,

    1982;

    Deming et

    al., 1992).

    Vitrinite

    reflectance measurements were also made on core and

    drill chip samples Magoon

    and

    Bird,

    1988).

    One of the wells from which temperature, thermal

    conductivity,

    and

    vitrinite reflectance data are available

    is the

    Ikpikpuk

    well latitude 70.46°N,

    longitude

    154.33°W) Figure 9.11). The history of burial and

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    15/22

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    16/22

    180  emin

    g

    Table 9.2. De pos i

    tional History, lkp

    ikpuk Well, North

     Slope o f Alaska

     

    Thickness

    Age of Dep

    osition or Event

    Sedimen

    tation Rate

    Ge

    ologic Unit(s) or Ev

    ent

    (m) (Ma) 

    (m/m

    .y.)

    Basement at

     surface

    350 and old

    er

    En

    dicott Group

    235

    350-3

    35

    16

    lisburne Group 94 6 

    335-258

    12

    Sa

    dlerochit Group

    320 258-23

    5

    14

    Shublik Fo

    rmation/

    Sa g River Sand

    stone

    168 235-2

    08

    6

    Kingak Sha

    le

    721 208-133

    10

    Nondeposit ion/er

    osion (?)

    0

    133

    -122

    Pebble

    shale  unit

    76  

    122-116

    13

    Tor ok Formatio

    n

    1304

    116--106

    130

    Nanushu

    k Group

    86

    9

    106--95

     

    79

    Colvil le Grou p/

    Sag

    avanirktok Forma

    tion

    -1000 

    9

    5-55 ?) 

    2

    5

    Erosion

    -100

    0

    55-1 ?)

    Gub

    ik Formation

    10  

    1-Q

    10

     

    Vitri

    nite Ref le

    ctance(

    Ro)

    interval

    heat perm

    eability

    0.2  

    1

    - 2

    -

     

    Q)

    3

    4

    5

    Nanushuk

    Torok 

    ebble

     

    s h l e ~

    \

    Sadleroch

    it

    -

    - ,

    \

    \

    \

    \

    f low mW/m

    2

     

    10-

    1

    4

    m2

    4 0

    31

    10

    44

    .79

    153 

    4.5

    392

    2.9

    -

    293

    Lisburn e

     

    40

    60

    80 100

    \

    \

    \

    \

    \ BHT

    \

    \

    \

    61

    .16

    ndicott

    bas

    ement

    T

    emperatu

    re Equiva

    lent (°C)

    F

    igure 9.12. Avera

    ge vitrinite reflect

    ance determined from

    a piecewise

     linear regression

     

    on

    measuremen

    ts (solid line), and

     

    predicted for s

    teady-state condu

    ctive hea t flows o

    f

    40

    ,

    60,

    80, and 100 mW

    /m2 (dashe

    d lines), lkpikpuk

     well, North Slo pe

     

    basin, Alaska. Also shown

    are

    interval heat flow s ca lculated from vitrinite reflectance

    and

    thermal conductM ty data alon

    g

    with average perm

    eabilites me asur

    ed parallel to bed

    ding.

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    17/22

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    18/22

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    19/22

    sedimentary

    basins that could be responsible for fluid

    circulation at the

    Ikpikpuk

    well is free convection. Free

    convection is normally inhibited in

    sedimentary

    basins

    because increasing salinity with depth

    more than

    offsets

    the decrease of fluid density due to thermal expansion.

    However,

    original brines

    in

    the

    North

    Slope

    basin

    have

    apparently

    been flushed

    out by

    the regional

    ground

    water flow system (see analyses

    of

    formation waters by

    Kharaka

    and

    Carothers, 1988;

    Woodward,

    1987).

    t

    may

    therefore

    be

    conceivable

    that

    the

    Ikpikpuk well

    is located

    in the descending section of one

    or more

    convection cells.

    f

    so, other wells

    in

    the

    North

    Slope

    basin may be

    located

    in

    upwelling sections

    of

    the same

    or other

    convection

    cells. Vitrinite reflectance

    data

    from

    these wells should

    show the opposite pattern to that found in the Ikpikpuk

    well-high paleothermal

    gradients near

    the

    top of the

    well

    and

    low paleothermal gradients near the bottom.

    Although it is possible to demonstrate that a partic

    ular hypothesis groundwater flow) is consistent

    with

    data from the Ikpikpuk well, the proposed hypothesis

    3)

    cannot

    be shown

    to be

    the

    only

    one that

    satisfies the

    observations. At our current level of understanding, the

    estimation of any thermal history is a complex problem

    that usually cannot

    be

    brought to a unique conclusion.

    Data errors

    and the

    difficulties inherent in inferring pale

    otemperatures from geothermometers such

    as

    vitrinite

    reflectance

    make it

    difficult to reconstruct

    the

    tempera

    ture

    history

    of

    potential

    source

    rocks accurately.

    Similarly,

    the potential of different mechanisms

    (e.g.,

    sedimentation, groundwater flow,

    and

    conductive heat

    refraction) to

    lead

    to identical

    thermal

    anomalies

    makes

    it

    difficult

    to uniquely designate

    specific

    physical

    processes as important factors

    in

    specific instances.

    Acknowledgments I would

    like to thank

    my

    colleagues and

    friends at the

    Branch of

    Petroleum

    Geology

    in the U.S. Geolog-

    ical

    Suroey

    at

    Menlo

    Park California. Ken Bird provided the

    estimated

    burial history for the Ikpikpuk well

    and

    Les Magoon

    made

    substantial contributions that improved

    the

    manuscript.

    Jud

    Ahern George

    Klein Dan McKenzie Jeffrey Nunn

    Gerard Demaison Peter van

    de

    Kamp and John

    T.

    Smith

    reviewed the manuscript

    and

    made suggestions for its improve-

    ment.

    References ited

    Ahem, J. L., and S. R. Mrkvicka,

    1984,

    A mechanical and

    thermal model for the evolution of the Williston basin:

    Tectonics, v. 3, p. 79-102.

    Angevine, C L., P. C Heller, and

    C

    Paola, 1990, Quantitative

    sedimentary basin modeling: AAPG Continuing Education

    Course Notes Series 32, 133

    p.

    Barker, C. E., and M. J. Pawlewicz, 1986, The correlation of

    vitrinite reflectance with maximum temperature in humic

    organic matter, in G. Buntebarth and

    L.

    Stegna, eds., Paleo

    geothermics: evaluation of geothermal conditions in the

    geological past, Lecture Notes in Earth Sciences: Berlin,

    Springer-Verlag, p. 79-93.

    9.

    Overburden

    Rock Temperature and Heat Flow

    183

    Beaumont, C., 1981, Foreland basins: Geophysical Journal of

    the Royal Astronomical Society,

    v. 65, p. 291-329.

    Bethke, C. M., 1985, A numerical model of compaction-driven

    groundwater flow and heat transfer and its application to

    the paleohydrology of intracratonic basins: Journal of

    Geophysical Research, v. 90, p. 6817-6828.

    Bethke, C. M., 1986, Inverse hydrologic analysis of the distrib

    ution and origin of Gulf Coast-type geopressured zones:

    Journal of Geophysical Research, v. 91, p.

    6535--6545.

    Bethke, C. M., J.D. Reed, and D. F. Oltz,

    1991,

    Long-range

    petroleum migration in the Illinois basin: AAPG Bulletin,

    v. 75, p. 925-945.

    Birch,

    F., 1948,

    The effects of Pleistocene climatic variations

    upon geothermal gradients: American Journal

    of

    Science,

    v. 246, p 729-760.

    Birch,

    F.,

    1950, Flow of heat in the Front Range, Colorado:

    GSA

    Bulletin,

    v. 61,

    p.

    567-620.

    Birch,

    F.,

    and H. Clark, 1940, The thermal conductivity of

    rocks and its dependence upon temperature and composi

    tion: American Journal of Science,

    v.

    238,

    p.

    529-558,

    613-635.

    Bird,

    K. J.,

    1988,

    Structure-contour and isopach maps of the

    National Petroleum Reserve in Alaska, in G. Gryc, ed.,

    Geology and exploration of the National Petroleum

    Reserve in Alaska, 1974

    to 1982:

    U.S.G.S.

    Professional

    Paper

    1399,

    p.

    355-377.

    Blackwell,

    D. D.,

    and J.

    L

    Steele,

    1989,

    Thermal conductivity

    of rocks: measurement and significance, in

    N.

    D. Naeser

    and T.

    H.

    McCulloh, eds., Thermal history of sedimentary

    basins: methods and case histories: New York, Springer

    Verlag, p.

    13-36.

    Blackwell, D. D.,

    J.

    L Steele, and C. A Brott,

    1980,

    The terrain

    effect on terrestrial heat flow: Journal of Geophysical

    Research,

    v. 85, p.

    4757-4772.

    Blanchard, D C., and

    I

    L. Tailleur, 1982, Temperatures and

    interval geothermal gradients determinations from wells in

    National Petroleum Reserve in Alaska:

    U.S.G.S.

    Open-File

    Report

    82-391,

    79 p.

    Bodell,

    J M.,

    and D. S. Chapman,

    1982,

    Heat flow in the

    north-central Colorado Plateau: Journal of Geophysical

    Research, v.

    87,

    p. 2869-2884.

    Bodner, D.P., and J. M Sharp, Jr., 1988, Temperature varia

    tions in South Texas subsurface: AAPG Bulletin, v. 72, p.

    21-32.

    Bredehoeft, J.D., W.

    Back,

    and

    B B

    Hanshaw,

    1982,

    Regional

    ground-water flow concepts in the United States: historical

    perspective, in T.

    N.

    Narasimhan, ed., Recent trends in

    hydrogeology:

    GSA

    Special Paper

    189, p. 297-316.

    Bredehoeft, J.D., and

    I

    S. Papadopulos,

    1965,

    Rates of vertical

    groundwater movement estimated from the earth s

    thermal profile: Water Resources Research,

    v.

    1,

    p. 325-328.

    Bredehoeft, J.D., R. D. Djevanshir, and K. R. Belitz, 1988,

    Lateral fluid flow in a compacting sand-shale sequence:

    South Caspian basin: AAPG Bulletin, v. 72, p. 416-424.

    Brigaud, F., and G. Vasseur,

    1989,

    Mineralogy, porosity and

    fluid control on thermal conductivity of sedimentary rocks:

    Geophysical Journal,

    v. 98, p. 525-542.

    Brigaud, F., D. S. Chapman, and S. Le Douaran, 1990, Esti

    mating thermal conductivity in sedimentary basins using

    lithologic data and geophysical well logs: AAPG Bulletin,

    v. 74, p. 1459-1477.

    Bullard, E.

    C.,

    1947, The time necessary for a borehole to attain

    temperature equilibrium: Monthly Notes of the Royal

    Astronomical Society,

    v.

    5,

    p. 127-130.

    Cathles,

    L.

    M.

    III, 1990, Scales and effects of fluid flow in the

    upper crust: Science,

    v.

    248,

    p.

    323-329.

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    20/22

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    21/22

  • 8/17/2019 9. Overburden Rock, Temperature, And Heat Flow_D. Deming

    22/22