Lob Kov Sky 1991

Embed Size (px)

Citation preview

  • 8/17/2019 Lob Kov Sky 1991

    1/32

    Tectonophysics, 199 1991) 343-314

    Elsevier Science publishers B.V., Amsterdam

    343

    A two level concept of plate tectonics: application to geod~a~cs

    L.I. Lobkovsky a and V.I. Kerchman b

    y

    nstitute of Oceanology The U.S.S.R. Academy

    of Scien ces

    Moscow ii 7238, USSR

    h The Interuniversity Computer Center, Kishinev 279003, USSR

    (Received August 15.1989; revised version accepted July 15.1990)

    ABSTRACT

    Lobkovsky, L.I. and Kerchman, V.I., 1991. A two-level concept of plate tectonics: application to geodynamics. In: L.P.

    Zonenshain (Editor), The Achievements of Plate Tectonics in the U.S.S.R. Tectonophysics, 199: 343-374.

    A twolevel plate tectonics concept is developed on the basis of data on lithosphere rbeological stratification. This

    approach differentiates between crust and subcrust plate ensembles separated by a lower-crust viscoplastic asthenolayer.

    Similarly to classical plate tectonics, three types of boundaries are distinguished in the lower layer which do not always

    coincide with crust-plate boundaries (especially for continents). Applications of this concept to geodynamics are considered,

    and a corresponding quantitative analysis for several important

    processes is carried out.

    A

    quantitative mode1 of mountain

    formation and collision-plateau origins is proposed.

    Also, a geodynamic model of the evolution of passive margins, taking into

    account a lower-crust viscous flow, is considered and its geological consequences are discussed. A mechanism

    of

    rifting, taking

    into consideration rheological lithospheric layering and its vertical movements caused by extension, is developed. Both a

    qualitative scheme and quantitative analysis of the slow evolution of intracraton structures of “shield-basin” type, taking into

    account erosion and sedimentation processes, are worked through. Also, historical aspects of plate tectonics are discussed from

    the point of view of the proposed concept.

    Intmduetion

    The orthodox theory of plate tectonics seems to

    have certain restrictions on its application. Thus,

    tectonic processes of a regional scale, that is of

    several hundreds of kilometres, cannot be de-

    scribed sufficiently well by standard plate-tectonic

    models. This is of special concern for the conti-

    nents (Molar, 1988; Lobkovsky, 1988a) because,

    when one analyses regional processes, those inho-

    mogeneities (both horizontal and vertical) and dis-

    tributions of intraplate strains which have not

    been considered on a global scale (since they have

    virtually been averaged) become the main objects

    of the research. Since many researchers have been

    guided by the idea of scale restrictions imposed on

    the plate-tectonic processes, they have put this

    above the quantitative study of plate non-rigidity

    and patterns of intraplate strains and stresses

    (Molnar and Tapponier, 1978; England and Mc-

    Kenzie, 1982; Vilotte et al., 1982; Cloetingh et al.,

    1984; England et al., 1985; Khain, 1986; Bruhn,

    1987; Kirby and Kronenberg, 1987).

    As for geomechanics, the analysis of the prob-

    lem mentioned is reduced to the description of the

    rheological behaviour of lithospheric rocks under

    the real P-T conditions and for the various regi-

    mes of tectonic strain.

    In geodynamics, an “effective strength)* is usu-

    ally used as a generalized rheological characteristic

    of the lithosphere (Ranalli and Murphy, 1987;

    Kirby and Kronenberg, 1987); this may be inter-

    preted in different ways, depending upon the

    P-T

    conditions and mechanisms of rock strain. For the

    domain of quasi-elastic strains and brittle fractur-

    ing, this characteristic coincides with the defini-

    tion of a material strength as used in mechanics

    (Byerlie, 1968; Sibson, 1974; Brace and Kohlstedt,

    1980). For the ductile (non-linear viscous) flow of

    the medium, the notion of “creep strength” has

    been introduced (Ranalli and Murphy, 1987;

    Kirby, 1983).

    ~-1951/91/ 03.50 6 1991 - Elsevier Scienw Publishers

    B.V. All rights reserved

  • 8/17/2019 Lob Kov Sky 1991

    2/32

    344

    L..l. LOBKOVSKY AND ‘4.1. KtRtHMAN

    a

    Temperature

    rengh

    b Tempera ure

    Fig. 1. Model strength profiles (solid lines): (a) for normal

    continental lithosphere with a 15 km thick granitic upper crust

    and a 25 km thick mafic lower crust; (b) for normal oceanic

    lithosphere with a 5 km thick basaltic upper crust and a 2 km

    thick serpentinitic lower crust. Rashed lines denote the geo-

    therms for the two cases. The dotted line denotes the strength

    envelope. The strain fate is lO_” s-‘.

    Ahhough each particular profile of an “effec-

    tive strength” compiled from the data of labora-

    tory tests, depends upon the assumed composition

    of the Earth’s crust and upper mantle, and upon

    temperature distributions, deformation rate, water

    saturation of the media, and so on, all these pro-

    files plainly indicate the main feature of the rheo-

    logical s~at~~ca~on of the exosphere; th8t is, the

    occurrence of an extremely low strength (viscosity)

    of the medium in the lower crust in contrast to the

    stronger and more brittle layers of the upper crust

    and the subcrustal part of the lithosphere.

    As an example, in Fig. la we illustrate a typical

    profile of the generalized strength of the litho-

    sphere, which is composed of an upper “granitic”

    crust (Z) - 15 km thick, a lower “mafic” crust

    (II) - 25 km thick and an underlying olivine

    mantle (III). This plot has been drawn from a

    typical continental geotherm that corresponds to

    the average heat flow on the surface; that is, about

    50 mW/m* {Morgan and Sass, 1984), in accor-

    dance with experimental rheological data (Byerlie,

    1968; Brace and Kohlstedt, 1980; Kirby, 1983;

    Ranalli and Murphy, 1987), the deformation rate

    being constant at C? 1O-‘5 s-j. The envelope line

    shows the resistance of the medium to brittle

    fracturing (Byerlie, 1968), and corresponds to the

    generalized strength of almost the entire upper

    crustal layer (I), a small part of a “cold” lower

    crust (II’) and a quasi-rigid core of the subcrustal

    mantle (III ‘). The main part of the lower crust

    (II) - that is, the crustal asthenolayer - shows

    ductile properties which may be described by a

    creep law (Kriby, 1983). The lower quasi-viscous

    lithosphere (III) is a tradition to the mantle

    asthenosphere.

    The profile of the generalized strength for a

    normal oceanic lithosphere has a similar character

    (Fig. lb). Here, the lower serpentinite layer of the

    oceanic crust is analogous to the crustal

    asthenolayer (Raleigh and Paterson, 1965;

    Lobkovsky et al., 1986). Note that each particular

    profile of a generalized strength (limiting shear

    stress) varies considerably from region to region as

    the heat regime and tectonic strain of the media

    vary (for instance, as the rates of various processes

    change in different layers).

    The present-day concepts described concerning

    the rheological stratification of geological media,

    distinguishing quasi-rigid and ductile layers in the

    crust, are analogous to the traditional geophysical

    concept of the existence of the lithosphere and

    asthenosphere in the upper mantle, the concept on

    which the orthodox theory of plate tectonics rests.

    The understanding of the geodynamic essence of

    this analogue has resulted in the fo~ula~on by

    Lobkovsky (1987a,b; 1988a,b) of the pfincipaily

    new concept of two-level plate tectonics. The main

    idea of this concept lies in the recognition of two

    main levels (crustal and subcrustal) and in the

    introduction of different scales in the conventional

    theory of plate tectonics. When global horizontal

    motions of several thousands of kilometres (a typi-

  • 8/17/2019 Lob Kov Sky 1991

    3/32

    A TWO-LEVEL CONCEPT OF PLATE TECTONICS

    345

    cal scale) are described, it is the lower subcnustal

    (upper-mantle) level of the system that works,

    whereas the processes in the upper-crustal level

    are subordinate. However, to describe adequately

    the majority of the regiond te&m.ic processes, the

    typical scales of which are only some hundreds of

    kilometres and even less, it is necessary to turn to

    the upper-crustal level of the geodynamic model

    considered. Within this level, the upper brittle

    crustal sublayer is dissected into individual micro-

    plates and geoblocks, their dimensions being from

    several hundreds to several tens of kilometres.

    Such crustal blocks are capable of horizontal dis-

    placements over the underlying crustal astheno-

    layer relative to the mantle part of the lithosphere.

    If one ignores these displacements - in the case

    of relatively strong coupling of crustal blocks with

    the mantle and at lesser differential stresses - it

    is possible to hold to the standard plate-tectonic

    constructions. If differential stresses are suffi-

    ciently great, as in the case of collision and con-

    ~nent~-~ft~g zones? the upper-crnstal layer may

    move and be deformed independently, to a con-

    siderable extent, on the subcrustal layer of the

    system due to the developed ductile flow of the

    lower crust. One such situation is illustrated in

    Fig. 2. A general two-level scheme for the continental c&&ion

    of India and Eurasia (after Lobkovsky, 19 8a; see text

    for

    explanation). I - Upper brittle crust; 2 = lower ductile crust;

    3 = subcrustal Lithosphere; 4 = mantle asthenosphere; 5 =

    thrust-type (convergent) crusti boundaries; 6 = strikaslip-type

    (transform) txustal bounties; 7 = she-ax- elo@ity distributio+x

    ill a cross-section of lowex crust; 8 = direetioll of sllbcrustal

    lithosphere movement; 9 = subduction zone.

    Fig. 2: the continental collision of India and

    Eurasia (Lobkovsky, 1988a,b). This shows the

    mosaic of the upper-crustal microplates, the rela-

    tive

    displacements of which are determined both

    by the indentation of the Hindustan to Eurasia

    boundary (at crustal level) and by the dragging of

    crustal blocks by the mantle part of the litho-

    sphere7 which is slipping beneath the crust. It is

    also determined by the flow of the lower crust

    caused by the above-mentioned process. As a re-

    sult of the viscous flow of the lower crust, its

    material may be forced (injected) into the vicinity

    of the suture zone, thus causing thickening of the

    lower crust and formation of the roots of the

    mount~s and isostatic uplifts of the territory.

    The present paper not only gives a quantitative

    analysis of the collision process (according to the

    scheme mentioned above) (Lobkovsky, 1988a,b;

    1990; Khain and Lobkovsky, 3990), but also con-

    siders how a two-level concept of plate tectonics

    can be applied to some geodynarnic problems. It

    describes the evolution of passive continental

    margins and the formation of rift zones at their

    rear (Lobkovsky, 1989; Lobkovsky and Khain,

    1989); it gives the general scheme of continental

    rifting at two levels and analyses the geodynamic

    behaviour of the crust within cratons with regard

    to the ‘~erosion-sedimen~tion-rnet~o~~srn-

    flow in the lower crust” rnate~~-circula~on cycle

    (Lobkovsky, 1989); finally, it considers some his-

    torical aspects of two-level evolution of the litho-

    sphere.

    Geological-geophysical grounds and general ideas

    of a two-level, plate-tectonic concept

    First, we briefly consider the factual data that

    provide evidence for a two-layered (in first ap-

    proximation) rheological structure of the Earths

    crust. The results of deep seismic studies, by pro-

    jects such as COCORP (Allmendinger et al., 1987),

    ECDRS (Choukroune and Garridq 1989) and

    EUGENO-S (h4eissner et al., 1987), have allowed

    their authors to determine a fine-layered structure

    of the lower continental crust. We believe that

    such a stratification probabIy originates from the

    horizontal flow of ductile mater&I in the Iower-

    crustal asthenolayer.

  • 8/17/2019 Lob Kov Sky 1991

    4/32

    346

    L 1. LOBKOVSKY AND V.I. KEKCHMAN

    Considerable progress has been achieved in in-

    strumental seismology in the past decade. As a

    result, the accuracy with earthquake hypocentres

    can be determined has increased by one order of

    magnitude (Chen and Molnar, 1983). From the

    studies that have been carried out, it turns out that

    the hypocentres of continental earthquakes are

    predominantly concentrated in the uppermost 20

    km of the crust. The lower layer of the continental

    crust, which is 15-20 km thick, is virtually

    aseismic; earthquake foci appear again only in the

    subcrustal horizons of the upper mantle in active

    tectonic regions of the Earth. Thus, the lower

    Continental crust can be considered to be an

    aseismic layer (Chen and Molnar, 1983; Meissner,

    1985; Jackson, 1987) and this verifies its astheno-

    spheric nature.

    A comparison of the observed distribution pat-

    terns of earthquake hypocentres in the lithosphere

    and the temperature field in the crust and upper

    mantle, made for various regions of the Earth, has

    shown that seismicity within the continental crust

    is limited by about the 3SO” C isotherm, whereas

    seismicity in the upper mantle is controlled by the

    700 o C isotherm (Chen and Molnar, 1983; Wiens

    and Stein, 1983). The temperatures mentioned

    correspond to the transition from a quasi-brittle to

    ductile deformation regime for the geomaterial of

    the crust (at

    - 350 o C) and mantle (at - 700 0 C)

    obtained in laboratory tests (DeRito et al., 1986;

    Jackson, 1987).

    The observed pattern of the distribution of

    seismicity serves as independent evidence for the

    rheological stratification of the Earth’s crust into

    the lithosphere and asthenosphere. Note that the

    correlation between the heat regime of the litho-

    sphere and the character of seismicity in one re-

    gion or another allows us to draw conclusions on

    the lateral variability of the properties of the

    crustal asthenosphere; and on the changes in its

    thickness, effective viscosity, etc. in particular.

    It follows from a proposed pattern of plate

    tectonics (Lobkovsky, 1988a) that the process of

    isostatic equilibrium should occur at two levels at

    least; namely, at the crustal and hthosphe~c levels.

    As for the lateral scales of isostatic compensation

    of several tens (or perhaps hundreds) of kilo-

    metres, the main role is played by the crustal

    asthenosphere, this being verified by analysis of

    local isostasy on the continents (McAdoo, 1985).

    When the scale is increased to several hundreds of

    kilometres, the mantle asthenosphere acquires an

    important meaning. The given supposition of a

    two-level isostatic compensation is verified by

    analysing the isostasy of each particular region.

    On the basis of such an analysis, it is possible to

    show which part of the isostatic compensation is

    taken up by the crust (it is usually 70%) and how

    much remains in the mantle lithosphere and

    asthenosphere (Artemjev and Kaban, 1987).

    The rheologieal statification of the crust and

    lithosphere is verified by numerous geological data

    concerning their tectonic layering in each particu-

    lar region &nipper and Ruzhentsev, 1977; Peive

    et al., 1983; Ranalli and Murphy, 1987). Extensive

    granite-gneiss allochthons with displacement am-

    plitudes of hundreds of kilometres, in the Alpine

    belt, Appalachian Mountains and other fold belts,

    are the direct geological evidence for the occur-

    rence of the lower-crustal asthenolayer (Cook et

    al., 1979; Hsii, 1979; Peive et al., 1983).

    We shall now formulate some general state-

    ments initially proposed by Lobkovsky (1987a,

    1988a) which complement the plate-tectonic ap-

    proach to a two-level pattern. At each level we

    distinguish a geodynamic system with its own

    ensemble of plates (~croplates) which, generally

    speaking, do not coincide. This is the principal

    difference in the proposed model from the ortho-

    dox plate-tectonic concept. The system of plates

    and microplates of the upper-crustal “deck” coin-

    cides with the orthodox plate-tectonic concept (Le

    Pichon et al., 1973; Zonenshain and Savostin,

    1979). The tectonic system of the lower (mantle>

    floor of the lithosphere can be explained by the

    occurrence of a quasi-rigid subcrustal layer III’

    (Fig. la) dissected into large plates, the size of

    which is on the scale of thousands of kilometres.

    They are dissected by boundaries of three types, in

    a similar manner to the orthodox concept, namely

    divergent, convergent and transform.

    If these plates and the boundaries between

    them coincide (the character of motion included),

    the plates may be considered as a monolith and

    their boundaries as a common one. In this case,

    we can use constructions of conventional plate

  • 8/17/2019 Lob Kov Sky 1991

    5/32

    A TWO-LEVEL CONCEPT OF PLATE TECTONICS

    341

    tectonics: this is true for the most part of the

    oceanic lithosphere, since in the oceans, the

    governing behaviour is that of the lower subcrustal

    lithosphere with a quasi-rigid “core” III’ (- 20-

    40 km thick), which is considerably thicker and

    stronger compared to the brittle overlying basalt

    crust (3-5 km).

    If the plates (and their boundaries) of the upper

    and lower floors do not ocincide, a more complex

    two-level model of tectonics originates, where

    various motions of crustal plates (blocks) are pos-

    sible relative to each other and relative to the

    mantle part of the lithosphere (due to the flow of

    a ductile lower-crustal layer). The collision scheme

    for such motions is shown in Fig. 2.

    A two-level tectonic structure is most distinct

    in the continental regions where, in contrast to

    oceans, the total strength and thickness of the

    upper-crustal brittle layer and the quasi-rigid sub-

    crustal core of the lithosphere are comparable. In

    this connection, a problem arises as to how to

    single out individual plates of the lower level

    under the continents, the boundaries of which do

    not coincide with the traditional ones.

    As for the divergent plate boundaries of the

    lower level, it seems most logical to attribute them

    to the extensive linear zones of basaltic (alkaline

    and tholeiitic) magmatism (Milanovsky, 1975),

    which usually occurs long before the rifting in the

    system at the upper level and which produces a

    complete splitting up of the plates in the tradi-

    tional sense. So, for instance, it is known

    (Razvalyaev, 1984) that during the past 800 Ma

    intrusions of alkaline basalts happened many times

    (at around 770-450, 290, 185, 120 and 80 Ma)

    along the East African rift system. The periods of

    basaltic magmatism alternated with the periods of

    secondary alkaline-granitic magmatism (570-450,

    185 and 50 Ma) (Razvalyaev, 1984). In addition to

    the East African rift system, a spatial coincidence

    of the linear zones of alkaline magmatism with rift

    structures has been revealed in the Gardar prov-

    ince (Greenland), the St. Lawrence rift, the Baikal

    rift zone and a number of others. In many cases,

    such magmatism occurred before not only the

    Cenozoic but also the Mesozoic rifting.

    In the framework of a two-level plate-tectonic

    concept, it is natural to consider such prolonged

    linear zones of alkaline magmatism as divergent

    plate boundaries in the lower level. It is difficult

    to trace these boundaries through the entire area

    of the continents, as the magmatism attributed to

    them is not always exposed on the Earth’s surface;

    for example, due to the compression regime in the

    upper level of the system. As an example of a

    divergent plate boundary in the lower level, we

    can draw a linear belt extending from the North

    Sea to the Lower and Upper Rhine grabens, to the

    grabens of the Seine and Rh&e, then continuing

    through the Mediterranean (Tunisia strait, grabens

    in Pantelleria and Malta), further on into Africa

    to the Gulf of Guinea, and along the Cameroon

    line into the South Atlantic. The East African rift

    system and the East Asian system from Anadyr

    Bay to Tungting Hu Lake in Southern China may

    be attributed to the same category of divergent

    plate boundaries (Lobkovsky and Khain, 1989).

    As in orthodox plate tectonics, the divergent seg-

    ments of plate boundaries in the lower level might

    be connected by lower-layer transform segments,

    although there is little direct evidence of this.

    The convergent plate boundaries of the lower

    level usually occur in collision belts where in-

    tracontinental subduction processes develop. Ex-

    amples are the Alpine-Himalayas zone of sub-

    crustal subduction (Fig. 2) and possible in-

    tracontinental subduction beneath the Rocky

    Mountains during the Laramide orogeny. As a

    rule, lower-level subduction is not marked by mid-

    dle- and deep-foci earthquakes, although rare ex-

    ceptions occur in individual segments with a nar-

    row frontal part (Calabria, Hindu Kush, etc.). The

    aseismic behaviour of such subduction zones can

    be explained first by the fact that cold hydrated

    crust is not subducted, as in the case of the

    oceanic plates and, second, by the considerable

    frictional heating of the subducted lithosphere (this

    mechanism will be considered in detail below).

    The aseismic regions of convergent plate

    boundaries in the lower level can be revealed using

    seismic tomography data. The inclined high-veloc-

    ity layers and the Q-factor correspond to such

    zones (Spakman, 1986). Post-collisional granitic

    magmatism generated by frictional heating may

    serve as a geological indicator of collisional sub-

    duction (Debon et al., 1986; Lobkovsky, 1988a;

  • 8/17/2019 Lob Kov Sky 1991

    6/32

    34x

    1, I. LOBKOVSKY AND V.I. KEKCHMAN

    Khain and Lobkovsky, 1990; Kerchman and

    Lobkovsky, 1990b).

    The two-level concept of plate tectonics poses a

    general problem, how to recognize the plate

    boundaries of the lower layer using the mass of

    complex geological and geophysical data availa-

    ble, including gravity field, heat flow and geo-

    chemical information (in particular, the zones of

    mantle helium).

    The possible structure of divergent plate

    boundaries in the lower layer may be predicted by

    analogy with the central parts of slowly expanding

    mid-oceanic ridges (Lobkovsky, 1989). Thus, they

    should show a bilaterally symmetrical inhomo-

    geneity of the subcrustal Lithosphere thickness. An

    example of a divergent plate boundary in the

    lower level may be the asthenospheric bulge under

    the Sayan-Baikal arch uplift and its northeast and

    southwest extensions (Rogozhina and Kozhevni-

    kov, 1979; Logachev and Zorin, 1988). Another

    example of a divergent boundary in the subcrustal

    layer is an extension of an anomalous mantle

    bulge from the Rio-Grande rift into the inner

    parts of the North American continent. In con-

    trast to the oceanic divergent boundaries, which

    are characterized by extension only, crustal struc-

    tures of both extension (Baikal rift) and compres-

    sion (the Mongolian Altai, Han-hai, the Gobi AI-

    tai and Eastern Sayan) can be located over the

    divergent boundaries of the lower level. Within the

    African continent, the latter may be represented

    by a considerably thinner subcrustal lithosphere

    (Fairhead and Reeves, 1977; Kazmin, 1987).

    As has been noted above, the divergent

    boundaries of the lithospheric lower layer are part

    of the global system of plate boundaries. In par-

    ticular, they continue from the continental to the

    oceanic regions and vice versa. The continuation

    of the East Pacific Rise under the crustal layer of

    North America in the Basin and Range Province

    and the ‘“Cameroon line” that continues the is-

    lands of the South Atlantic into the lower-level

    extension boundary under West Africa are typical

    of this feature (Lobkovsky, 1989).

    Plate bound~es of the subcrustal layer seem to

    be as stable in time as the principal boundaries of

    large oceanic plates, i.e. mid-oceanic ridges and

    subduction zones. They can change due to changes

    in the mantle convection regime, and also due to

    large-scale geodynamie processes which occur at

    both levels of the lithosphere (collision, for in-

    stance).

    A two-level plate-tectonic model allows us to

    incorporate the principal mechanisms of rift prop-

    agation (Co~tillot, 1982; Martin, 1984). In par-

    ticular, the propagation of rifts from oceanic basins

    to the upper-crustal layer of continents may be

    due to a pre-existing divergent boundary in their

    lower layer.

    This new view of the plate tectonics of the

    lithosphe~c lower layer allows us to consider hot

    spots within the continents as zones where mag-

    matic activity is localized at triple junctions of

    divergent and transform lower-level boundaries,

    where the quasi-rigid layer of subcrustal litho-

    sphere is completely absent. Examples of such

    isolated points are Yellowstone Park in the U.S.A.,

    and the volcanic areas of Tibesti, Darfur and

    Achaggar in Central and Western Africa (Lobkov-

    sky, 1989). The tectonic evolution of the Earth in

    the earliest historical stages (Archean and Pro-

    terozoic) is of great interest. At present, the majot-

    ity of researchers believe that plate-tectonic

    processes began on the Earth as late as in the

    Proterozoic (Khain and Mikhailov, 1985; Khain

    and Bozhko, 1988). According to the concept de-

    scribed above, the plate-tectonic regime may begin

    when a general cooling of the Earth occurs, the

    intensity of mantle convection decreases and, as a

    result, a continuous subcrustal layer is formed in

    the lithosphere. The quasi-rigid core of the lower

    lithospheric layer prevents chaotic motion and

    decouples the previously formed massifs of sialic

    protocrust from the unsteady convective flow of

    the mantle. Such “capturing” of individual proto-

    crustal blocks by the lithospheric lower layer may

    explain the formation of stable cratons and con-

    tinental cores. The mosaic pattern of Archean

    permobile tectonics may be explained by the lack

    of a quasi-rigid frame in the lower level. Similar

    conclusions can be reached about the oceanic

    proto-crust. This concept of the evolution of the

    Earth’s tectonic regime from a hot chaotic state to

    an organized plate-tectonic one by the formation

  • 8/17/2019 Lob Kov Sky 1991

    7/32

    A TWO-LEVEL CONCEPr OF PLATE TECTONKS

    of a continuous lower-level “skeleton” (frame)

    may be applied to the evolution of all terrestrial

    planets.

    We note that in a two-level plate-tectonic model

    the main driving forces are applied to the plates of

    the lower level, either by coupling of subcrustal

    plates with convective mantle flows or by a “pull-

    ing” effect that works on the subducted mantle

    parts of these plates or, finally, by spreading of

    anomalous mantle and “ridge push” (Forsyth and

    Uyeda, 1975; Cloetingh and Wortel, 1985; Seki-

    guchi, 1985). The forces driving the upper-crustal

    layer are determined by their coupling with the

    plates of the lower level; they also may be con-

    nected with inhomogeneities of crustal thickness

    and density. One other special aspect of the two-

    level model is that plate interactions at the crustal

    level play a greater role in the geodynamics of

    crustal blocks and microplates than the corre-

    sponding plate interactions at the lower level.

    Mechanical aspects of the two-level plate-tectonic

    model

    To analyse the mechanical problems which arise

    in two-level tectonic processes, we first consider

    briefly some aspects of the rheological behaviour

    of the continental lithosphere under various con-

    ditions and strain regimes. The ductile flow law

    derived experimentally for crustal and upper-man-

    tle rocks at high temperature is:

    t=Ar”exp[-Q/R(T+273)]

    (1)

    where P is the strain rate, 7 = u1 - a3 is twice the

    shear stress, T is the temperature in *C; and A, n

    and Q are material constants depending upon the

    composition, structure and water content of the

    rocks (Kirby, 1983).

    In Figs. 3a and b are presented (a) the modified

    curves of generalized strength (limiting stress) for

    different temperature regimes and (b) the non-ho-

    mogeneous strain-rate dist~bution in sublayers of

    viscous flow that correspond to real tectonic regi-

    mes; for instance, to the slip of a lithospheric

    subcrusml layer during intr~ntinental subduc-

    tion. This difference from the idealized curves

    (solid lines, Figs. 1 and 3) should be taken into

    account when analysing each particular geody-

    namic situation qu~titatively.

    Strmgth

    I 60

    %

    4

    80

    .

    b

    st rain rote

    Fig. 3. Model strength profiles of continental lithosphere with

    a 15 km thick granitic upper crust and a 25 km thick mafic

    lower crust: (a) for different temperature regimes correspond-

    ing to g-therms with heat flows of 45 mW/m’ (solid line) and

    60 mW/m* (dashed line); (b) for ho~~~us (solid line) and

    non-homogeneous (dashed line) strain-rate distribution in the

    lithosphere (dotted line).

    We now consider some mathemati~l models in

    which the main role is played by ductile flow of

    the material in the lower crust due to changes in

    its thickness.

    Consider a two-dimensional mathematical

    model in which the x axis is oriented along the

    dominantly horizontal motion and the z axis is

    oriented vertically upwards (Fig, 4). We assume

    that there is a viscous deformable asthenolayer

    (II) of the lower crust, with varying thickness,

    which is overlain by an elastic-brittle upper crust

    (I) and underlain by a horizontally moving

    quasi-rigid plate (III’) of the lower level. For

    simplicity, we ignore the bending rigidity of the

    upper layer (I) and assume it to be non-deforma-

  • 8/17/2019 Lob Kov Sky 1991

    8/32

    350

    L.I. LOBKOVSKY

    AND V.I. KERCHMAN

    Fig. 4. Main layers interacting with each other within the

    framework of the two-level plate-tectonic concept (see text for

    explanation). I = upper brittle crust; 2 = lower ductile crust;

    3 = subcrustal quasi-rigid lithosphere.

    ble in a horizontal direction. If we assume a

    Newtonian viscosity of layer (II) and local iso-

    static compensation of the crust (due to a rather

    rapid response of the subcrustal lithosphere), the

    equation for the evolution of the lower-crustal

    thickness h is (Lobkovsky, 1988a):

    - ; (Uh)

    where K = (pM -

    &~cg/lh.m; P,,., ad pc are

    the densities of the mantle and crust, respectively,

    g is the acceleration due to gravity, q, is an

    effective mean viscosity of the crustal lower layer,

    and U is the horizontal velocity of the subcrustal

    part of the lithosphere relative to the upper-crustal

    layer. We consider this in greater detail in the

    Appendix.

    When the behaviour of the lower crust follows

    a more realistic rheological law (1) with index

    n = 3, the equation of the evolution of the layer in

    an isothermic approximation for a fixed sub-

    crustal basement is (see Appendix):

    Here:

    /3=bB

    Pc(Ph4 - Pc)g

    3

    PM

    1

    (3)

    (4)

    where

    B = I

    exp[ - Q/R(T + 273)] and

    b

    is a

    normalizing factor that depends on the behaviour

    of the upper crust: when this layer is horizontally

    rigid (flow “under a cap”), b = l/80; when the

    upper layer is fractured and blocks of the upper

    crust can move freely in a hotintal direction,

    b = l/S (Kerchman, 1990).

    In the following sections we shall consider

    mathematical models of particular geodynamic

    processes, using eqns. (2) and (3).

    A two-layer collision model at an intrawutiuental

    subduction zone

    According to orthodox plate-tectonic theory,

    the mountain fold belts of the Earth are formed

    by the collision of large continental plates and by

    the accretion of smaller crustal blocks (terranes) in

    zones of lithospheric convergence (Dewey and

    Bird, 1970; Ben-Avraham et al., 1981; Zonenshain,

    1986). Various mechanisms for crustal thickening

    and uplift of mountain areas have been proposed:

    namely, frontal compression, shortening and

    warping of the crust and lithosphere (Dewey et al.,

    1988); partial subhorizontal subduction of one

    continental plate under another to cause a

    mechanical doubling of the crust (Powell and

    Conaghan, 1973); penetration of rigid continental

    “indentors” (Adria, Arabia and Hindustan) into

    the elastoplastic (Molnar and Tapponnier, 1978)

    or viscoplastic (England and McKenzie, 1982;

    England and Houseman, 1986) body of an ad-

    jacent plate (Eurasian lithcsphere); and various

    types of piling up and delamination of the crust

    (Oxburgh, 1972; Bird, 1978; Hsti, 1979).

    Although the above-mentioned models each

    have their own merits and may accurately describe

    certain aspects of collision belts, they do not ex-

    plain some important special features of their

    structure and evolution. For example, they do not

    explain the data on thickening of the continental

    crust due to accretion of its lower layer (Giese,

    1980; Choukroune and Garrido, 1989), the high

    geothermal gradient of collision belts (Artyush-

    kov, 1979; Morgan and Sass, 1984), post-collision

    granitic magmatism (Debon et al., 1986), or the

    lower convergence rate of continental blocks in

    the collision zone in comparison with that of the

    lithosphere plates which carry them (collision of

    Hindustan with Eurasia; Trifonov, 1987). As yet,

    there is no explanation for phenomena such as the

    absence of mid- and deep-focal earthquakes in the

    greater part of the Alpine-Himalayas collision

    belt, although the entire geological history of the

    Tethys closure (Sborschikov, 1988; Kazmin et al.,

  • 8/17/2019 Lob Kov Sky 1991

    9/32

    A TWO-LEVEL CONCEFT OF PLATE TECTONICS

    351

    1986 and the recent data from seismic tomogra-

    phy (Spakrnan, 1986) provide evidence for sub-

    duction of the lithosphere under Eurasia.

    A new model of continental collision was

    worked out by Lobkovsky (1988a,b). It arises from

    the assumption of two-level plate tectonics, by

    which the above aspects of the structure and

    evolution of collision belts are explained (Fig 2).

    The traditional analysis of the plate-tectonic

    evolution of the Alpine-Himalayas belt is based

    on the assumption of a mosaic pattern of litho-

    spheric microplates (- 100-200 km thick) and on

    the standard kinematic models (McKenzie, 1972,

    Zonenshain and Savostin, 1979). In contrast to

    this approach, the two-level plate-tectonic model

    (Fig. 2) supposes that the microplates that we

    observe at the surface are of crustal origin and

    may move relative to the mantle part of the litho-

    sphere, undergoing large rotations and non-elastic

    deformations. Palaeomagnetic data actually give

    evidence of remarkable rotations of microplates

    and blocks in the process of collision (Klootwijk

    et al., 1986).

    In accordance with this new approach (Bird,

    1978; Lobkovsky, 1988a,b), at an early stage in

    the collision between continental plates (preceded

    by subduction of the oceanic part of the plate,

    which carries a “climbing” continent; Hindustan,

    for instance), a sharp reduction in the convergence

    velocity of the upper brittle layer of the crust

    occurs. The mantle part of the lithosphere con-

    tinues to move and to subduct under the forces of

    convective “dragging” and “pulling” of the sink-

    ing edge of the plate (Figs. 2 and 5).

    At the same time, an intensive shear flow

    evolves in the lower plastic layer of the crust. This

    Fig. 5. Two-lewl model of mountain formation in the wmsse of

    continental collision (after Lobkovsky, 1988a; see text for

    explanation).

    flow embraces one area after another (squeezed

    between the upper crust and mantle part of the

    lithosphere) as the front of deceleration of the

    upper crust advances toward the inner parts of the

    overriding subcontinent. This propagation of the

    deformation front in the crust can be described by

    Elsasser’s equation (Lobkovsky, 1988a). Non-elas-

    tic compression of the upper crust in the vicinity

    of a suture zone is manifested as a system of

    thrusts that form the front of the developing

    orogen. The principal agent of crustal thickening

    and regional uplift is the pumping of the plastic

    material of the lower crustal layer into the vicinity

    of the subduction zone by the movement of the

    underlying mantle part of the lithosphere (Figs. 2

    and 5). A similar model of the detachment of

    subducted lithosphere from the crust in a colli-

    sional orogen, but with a different crustal thicken-

    ing mechanism, is considered in papers by Mat-

    tauer (1986) and Dewey et al. (1988).

    Consider a quantitative analysis of this ap-

    proach (Lobkovsky,

    1988a; Kerchman and

    Lobkovsky, 1990b). Changes in thickness h of the

    lower-crustal layer in the approximation of New-

    ton quasi-isomers rheology are described by

    eqn. (2). The co-ordinate system moves with the

    conventional non-deformed upper crust of the col-

    liding continent (India, for instance) (Figs. 2 and

    5). The slip rate U =

    U, of the subcrustal litho-

    sphere relative to the upper crust increases from

    zero, starting from the moment of detachment

    after the first collision phase (which is manifested

    by frontal compression thrusts and a slight thick-

    ening of the crust), as shown in Fig. 6b.

    The following parameter values were assumed

    in order to solve eqn. (2) numerically: pc = 2.8-2.9

    g/cm3 (for the lower crust), PM = 3.3-3.35 g/cm3

    and nc = (0.3-1.5)

    X

    lo*’ Pa s; therefore K =

    (0.5-3) X lo-’ km-’ yr-‘. Time variations in the

    rate of slip have been adopted to correspond with

    the conditions of the intracontinental subduction

    of the lower layer of the Indian plate (Fig. 6b).

    In Fig. 6a are shown calculated curves for the

    subsequent thickening of the lower-crustal layer

    from the moment when the shear flow started.

    Thus, the total crustal thickness (its brittle upper

    layer is 18-25 km thick) increases to about 70 km

    after a time of 40 Ma in the collision zone, which

  • 8/17/2019 Lob Kov Sky 1991

    10/32

    352

    L I LOBKOVSKY AND V I KERCHMAN

    IO

    20

    30 40

    t , Ma

    time

    Fig. 6. (a) The curves of succe.ssivehickeningof the lower ductile crust in the wurse of continentaJcolIisionobtained by numerical

    simulation; values near the curves denote post-collision time in Ma. (b) Time dependence of subduction velocity of subcrustal

    litbosphere elative o the upper crust since the beginningof the collision. See text for explanation.)

    agrees well with data on the crustal thickness in

    the Himalayas. The typical lateral dimension of

    such an area with thickened crust (formation of

    the Himalayas and Tibet “roots”) is 600-700 km,

    corresponding well with the actual area of isostatic

    uplift of the territory.

    The calculated asymmetry of relief about the

    suture zone, which implies a steeper topography of

    the frontal erogenic area, a considerably more

    gradual slope at its rear, and the existence of a

    collisional plateau, is more pronounced in the

    calculations when the non-linear viscous rheologi-

    cal model of eqn. (1) and the evolutionary equa-

    tion (3) are used. In Fig. 7 is shown the evolution

    of a collisional plateau due to the northward prop-

    agation of the material “pumped” under the orog-

    eny.

    We should mention that the model described

    above is based on the principle of local isostasy

    bending rigidity of the lithospheric lower layer

    (III’) should be included in estimates of the

    isostatic subsidence of the subcrustal part of the

    lithosphere under the load of a thick crust (Karner

    h.

    km

    L

    L

    I

    1W 200 300 400 ml

    600

    X.km

    dlstonce from orogen ax,s

    Fig. 7. Succe&on of the cakzuiated urves showiiq the evolu-

    tion of a collision rdateaa values near the curves denotina

    (see eqn. (Al) in the Appendix). In reality, the

    post-colliSionime n Ma.

  • 8/17/2019 Lob Kov Sky 1991

    11/32

    A TWO-LEVEL CONCEPT OF PL4TE TECTONICS

    Fig. 8. Thermal scheme of a three-layered lithosphere witbin an

    intracontinental subduction zone, based on the model of con-

    tinental collision (see text for explanation). 1 = Granite upper

    crust; 2 = “basalt” lower crust, 3 = olivine subcrusti litho-

    sphere; 4 = direction of subduction; 5 = horizontal velocity

    distribution in the layered lithosphere; 6 = frontal thrusts.

    and Watts, 1983; Lyon-Caen and Mohmr, 1985).

    Therefore, the above dynamic model of a colli-

    sional orogeny should be generalized to include

    quasi-elastic bending of the hthospheric lower

    layer (III’) under the influence of a thickening

    crust. With such a model it would be possible to

    study the dynamics of the formation of foredeep

    basins, and to have a more realistic view of the

    theoretical history of uplift.

    We shah now analyse the thermal regime of the

    crust and subcrustal lithospheric layer in the zone

    of a continental collision, using the above geody-

    namic model and considering the additional factor

    of dissipative heating of the media. Assuming that

    all displacements are horizontal, we consider a

    tw~~~ion~ thermal model of a three-layered

    lithosphere that includes a “granite” upper layer

    of the crust (I), 0 G z B h,(x), a “basalt” lower

    crustal layer (II), kl (x)gz G k2(x), and an

    olivine subcrustal part of the lithosphere (Fii. 8;

    Taylor and McLennan, 1985). Taking into account

    the motion of the lower layers (the upper layer of

    the crust is assumed to be locked and therefore

    stable), dissipation in the ductile layer of the lower

    crust and radioactive heat generation in the crust

    lead to the following non-stations

    heat conduction:

    aT

    2

    a+fi

    37 =‘I ax2

    (

    ar2

    1

    +Q,=P -z/M

    wP)1

    O

  • 8/17/2019 Lob Kov Sky 1991

    12/32

    354

    lL.l. LOBKOVSKY AND V.I. KERCHMAN

    km further beneath the Himalayas and Tibet; Fig.

    8). The thickness of the “granite” layer was as-

    sumed to increase from 12.5 km in the south to 20

    km in the north.

    The rheological parameters of the rocks com-

    posing the deformable “basalt” crustal layer were

    taken from experimental data on diabase (Kirby

    and Kronenberg, 1987): n = 3.4; A = 3 x lo6

    GPa-” s-’ and Q = 260 kJ/mol in the tempera-

    ture range from 400 to 800°C and at strain rates

    from 10-‘4-10-‘3 s-i.

    The temperature distribution in a “normal”

    continental lithosphere (Morgan and Sass, 1984;

    Taylor and McLennan, 1985) was taken as the

    initial state. These temperatures were held con-

    stant for the entire period of the modelling process

    at the boundary of the study area. Numerical

    estimates were made using an explicit finite-dif-

    ference scheme with steps of Ax = 6 km; AZ = 2.5

    km, At = 0.05 Ma. The numerical analysis de-

    scribed here used the following parameter values:

    X, = 2.5 W/mK,

    (pc,,), = 3 x lo6 J/m3K,

    X, = 2 W/mK,

    (PC,), = 3 x lo6 J/m3K,

    X, = 3.5 W/mK, (PC,), = 4 x lo6 J/m3K,

    Q,=2X10e6 W/m3;

    qb = 0.5

    x

    1O-6 W/m3.

    In Fig. 9 is shown the evolution of the calcu-

    lated geotherms (for 30 Ma) of the plastic flow

    developed in the lower crust, with the relative

    velocity of the subcrustal lithosphere given by Fig.

    6b. It can be seen that the temperature in the

    lower-crustal layer increases to 680-700 o C as a

    result of dissipative heating. The mantle heat flow

    is screened by the anomalously hot lower crust,

    leading to temperature increases in the mantle: in

    the adjacent 10 km thick layer of the subcrustal

    lithosphere, for t = 15-20 Ma, the increase is 80-

    150’ C (up to 700-750” C) and, below that, it is

    40-80” C (up to 750-800” C). Thus, the litho-

    sphere subducted under the Himalayas is already

    heated to 700-800” C in its upper part, and so

    brittle rupture does not occur in the zone of high

    shear strain (Molnar and Chen, 1983; Jackson,

    1987). This may explain the aseismic behaviour of

    this entire intracontinental subduction zone. Mid-

    c

    km

    Fig. 9. Calculated lithosphere gmtherrns for different times

    since the beginning of the collision. I = Initial geotherm; 2 =

    geotherm for 10 Ma; 3 = geotherm for 20 Ma; 4 = geotherm

    for 35 Ma.

    depth and deep-focal seismicity may occur in some

    regions of a subducted lithosphere due to some

    local underheating (for instance, because of the

    sensitivity of the dissipation process to rheological

    properties of the media, or because of the ex-

    istence of more brittle lithosphere as a result of

    water release during deserpentinization of the sub-

    siding parts of the oceanic crust (Lobkovsky et al.,

    1986).

    The estimated dissipative heating of the lower

    crust is sufficient to generate collisional and post-

    collisional magmatism, since the melt temperature

    for damp (hydrous) granites at depths of 15-25

    km is 650-700°C (S&mid and Wood, 1976;

    Dobretsov, 1980; Reverdatto and Kalinin, 1980;

    Taylor and McLerqmn, 1985).

    We note that for a period typical of a continen-

    tal collision, i.e. 30-50 Ma, the thermal dis-

    turbance in the lower crust has only a slight effect

    on the surface heat flow. Nevertheless, the high

    mean heat flow of Hinchrstan (especiaRy in the

    north; &I&, 1985) may he a result of dieraipative

    heating of the media; local variations caused by

    various types of radiogenic heat release in the

    surface crystal layer and by other factors are not

    considered in this paper.

  • 8/17/2019 Lob Kov Sky 1991

    13/32

    A TWO-LEVEL CONCEPT OF PLATE TECTONICS

    Geodynaniics

    ,of

    passive tlwghw taking

    into

    account te49onic flow of the lower crust and sub-

    lithosphere

    3nombm mantle

    The general features of the genesis of passive

    margins have been explained within the frame-

    work of the classical theory of plate tectonics (Le

    Pichon et al., 1973; Burke and Drake, 1974). How-

    ever, certain details of their structure and evolu-

    tion cannot, as yet, be explained sufficiently well.

    In particular, these phenomena are (1) the forma-

    tion of elongated uplifts on the continents, parallel

    to passive margins (Ollier, 1985); (2) the existence

    and evolution of a system of rift zones at the rear

    of passive margins (Milanovsky, 1976; Ziegler,

    1982); (3) the process of tearing away of crustal

    blocks (such as microcontinents and terranes) from

    large continental massifs - a process which, in

    fact, is the opposite of accretion tectonics (Sengor,

    1984; Vink et al., 1984; Kazmin, 1989); and (4)

    the occurrence of a belt of crystalline basement

    (100-200 km wide) with an anomalous P-wave

    velocity of about 7 km/s between the ~ntinent~

    slope and normal oceanic crust (Emery and

    Uchupi, 1984).

    The two-level geodynamic model of the evolu-

    tion of passive continental margins proposed by

    Lobkovsky (1989) and Lobkovsky and Khain

    (1989) explains the features listed above qualita-

    tively. We shall consider this model briefly and

    then describe the evolution of passive margins

    within its framework quantitatively. To a first

    appro~mation, the spreading of the ~thosphere

    subsequent to the process of continental rifting

    leads to the formation of passive margins. This

    spreading is determined by the interaction of the

    four main layers of the crust and upper mantle,

    namely (I + II’), (II), (111’) and (M), which

    were described in the Introduction (Fig. la;

    Lobkovsky, 1989).

    An accumulation of anomalous mantle first

    occurs under the pre-rift uplift during the exten-

    sion of the ~thosph~e. This accusation results

    in the breaking through of a portion of partially

    melted mantle into the crustal level of the geody-

    namic system. Thus, a typical continental rift

    structure develops that includes upper subcrustal

    (IV) and lower sub~~osphe~c (V) lenses of

    mf @2 @j-j@ m4 loo;;lS /36 B7

    Fig. 10. Successive stages of rifting and spreading processes:

    (a) continental rifting; (b) initial stage of ocean spreading; (c)

    mature stage of ocean spreading (after Lobkovsky, 1989).

    I = Upper brittle crust; 2 = lower ductile crust; 3 = subcrusti

    lithosphere; 4 = anomalous mantle; 5 = normal mantle; 6 =

    owau crust; 7 = sediments (see text for explanation).

    anomalous mantle (Fig. lOa). The size of the

    former lens is several tens or hundreds of kilo-

    metres, whereas that of the latter is, at first, several

    hundreds of kilometres; later, in the stage of ac-

    tive spreading, becoming thousands of kilometres.

    The appearances of a subcrustal diapir of anoma-

    lous mantle and related additional local extension

    and heating of the crust lead to thinning of the

    lower plastic layer of the crust due to the outflow

    of its material off the rift axis, resulting in an

    &static near-axial subsidence. The development

    of extensional fissures and normal faults in the

    brittle upper layer of the crust that forms the

  • 8/17/2019 Lob Kov Sky 1991

    14/32

    356

    L.I. LOBKOVSKY AND V.I. KERCHMAN

    structure of the rift graben can be observed (the

    lower crust and anomalous mantle propagating

    under the peripheral zones of the rift lift their

    “shoulders” simultaneously (Fig. 10a). Thus, the

    dynamics of the crustal layer governs this system’s

    interaction with the subcrustal mantle lens: the

    sublithospheric anomalous mantle causes an uplift

    of the terrane and acts as a feeding chamber for

    the upper subcrustal lens.

    Continental rifting is completed by extreme

    thinning of the lower crust and rupturing of the

    upper brittle layer, which is also considerably

    thinned by that time. A final stage of developed

    spreading follows this one. In the final stage, the

    separated parts of a continental plate move away

    from each other and undergo further evolution

    (Figs. lob and c).

    The previously cited theoretical papers devoted

    to the evolution of the continental passive margins

    are mainly concerned with the problem of vertical

    subsidence due, first, to the cooling and densifica-

    tion of anomalous mantle preserved under the

    crust (Artyushkov, 1979; Meissner and Kopnick,

    1988) and, second, to accretion of a “heavy”

    oceanic lithosphere “brazed” onto the continental

    lithosphere along the line of the prime rupture

    (Sorokhtin, 1979). Note that the subsidence of the

    basement of the passive margin is enhanced by the

    additional loading of a rapidly accumulating sedi-

    mentary cover (Cloetingh et al., 1984), whereas rift

    expansion and subsidence of a continental edge is

    accompanied by the development of listric faults

    (Le Pichon and Sibuet, 1981, Kazmin, 1987).

    In addition to the phenomena described above,

    within the passive transitional zones from conti-

    nents to oceans, new processes arise that are in-

    duced by horizontal motions in various directions

    of the media in the upper and lower levels of the

    heterogenous system illustrated in Figs. lob and c.

    The main process at the lower level is the propa-

    gation of a large lens of sublithospheric mantle

    away from the axis of a mid-oceanic ridge, which

    is usually accompanied by “dragging” through

    convection. The formation of wide swells along

    passive margins seems to be connected with this

    process. It occurs about 30-100 Ma after the

    onset of relative spreading (Lobkovsky and

    Khain,

    1989).

    We now consider a quantitative model of this

    phenomenon (Figs. lob and c). The equation de-

    scribing the evolution of a rheologically homoge-

    neous layer of anomalous mantle with variable

    thickness h and viscosity na (which is consider-

    ably less than that of the over- and underlying

    media) reduces to eqn. (2) in view of the observed

    isostatic condition (Artyushkov, 1979; Lobkovsky,

    1988a).

    The coefficient of eqn. (2), K = pM -

    Pa)Pag/12Plvrnla, where pM and p, are the densi-

    ties of the normal and anomalous sublithospheric

    mantle, respectively. Typical values of the parame-

    ters are as follows: plLl= 3.4 g/cm3, pM -

    Pa =

    0.1-0.5 g/cm3; and the viscosity qa of anomalous

    mantle is - lo’* Pa s. For the coefficient

    K we

    thus have an estimate K = (l-5) X lo-’ km-’

    yr-‘.

    First consider the simplest model problem,

    which ignores the convective motion of the sub-

    lithospheric mantle (this motion seems to be im-

    portant only in a narrow belt, several hundreds of

    kilometres wide, along the margin). In this case, if

    the axis x is oriented horizontally towards the

    motion of the anomalous mantle and the reference

    point x = 0 is located under the crest of the coastal

    slope of the margin, eqn. (2) is valid only if x >, 0.

    Using the equation:

    -=KK-

    (6)

    we solve the problem of the propagation of the

    anomalous mantle lens beneath the continent, sub

    ject to the boundary condition that its thickness is

    constant at x = 0.

    hl,_,,=H, hl,=,=O

    (7)

    We choose an initial condition that represents the

    early stage of the under-flow of a localized massif

    of anomalous mantle, its width being I, = 100-200

    km:

    (usually H =

    50-80 km,

    the dimension of a study

    area S being 2500-5000 km; see Fig. 1Oc).

    The solution of eqn. (6) with such initial condi-

    tions is characterized by a finite rate of propa-

  • 8/17/2019 Lob Kov Sky 1991

    15/32

    ATWO-LEVELCONCEI'TOFPLATETECTONICS

    I 1

    I

    /

    I

    I

    Q

    I‘? A4

    d6 A8 mc

    Fig. 11. Graph of the function f(t) for a self-similar asymp-

    totic law of anomalous mantle lens propagation.

    gation of the frontal boundary of the disturbed

    area X(t) (Barenblatt, 1980). The boundary con-

    ditions (7) permit self-similar “intermediate”

    asymptotics, such as:

    h=f fW, E= -

    Substituting (8) into eqn. (6) we obtain an

    ordinary differential equation for the function

    f 5):

    the conditions being:

    fl

    +,=l

    and

    f

    It_co=O

    (10)

    Here to = lim, _ f0

    X(t)/dKH3tl

    so that f I ,to

    = 0; the function d f 4/dc should be continuous at

    the point t = &, to satisfy the flw-continuity con-

    dition of anomalous mantle.

    A plot of the numerical solution

    of this

    boundary-value problem is presented in

    Fig. 11.

    To estimate the time of propagation

    for the

    351

    anomalous mantle front at a distance L, from the

    edge of the continental margin, we use the

    asymptotic law of propagation of the front:

    t=o %Li/~H~

    (11)

    When

    H =

    70 km (which agrees with an isostatic

    uplift of S = 2 km) and uplift width L, = 1200

    km, we have (for K = 0.15 km-’ Ma-‘) I* = 30

    Ma.

    The general solution for the full eqn. (2) has

    been obtained by an explicit finite-difference

    scheme. This method permits us to consider a

    non-stationary boundary condition at the left-hand

    edge of the study area, reflecting, for instance, a

    local thinning of an anomalous mantle lens due to

    the subsidence of an oceanic plate margin under

    the increasing load of accumulating sediments.

    The initial form of the protruding massif of

    anomalous mantle under the continental margin is

    assumed to be:

    ho(x)=

    i

    Hew[-0.2(x/H)2], O~XGI,,

    0, l,

  • 8/17/2019 Lob Kov Sky 1991

    16/32

    358

    L.I LOBKOVSKY AND V.I. KERCHMAN

    The evolution of the anomalous asthenospheric

    layer beneath the continent is presented in Fig. 12.

    The self-similar law (11) is a good description of

    the propagation of the lens for a time t > 10 Ma.

    Thus, the flow of anomalous mantle in the

    lower level of the system uplifts the edges of

    continental passive margins, and lags considerably

    behind the primary rifting (Lobkovsky and Khain,

    1989; Kerchman and Lobkovsky, 1990a). As flow

    beneath the continental passive margin reaches

    maturity (Fig. lOc), the margin undergoes various

    vertical motions: (1) gradual uplift as a wide

    marginal swell is formed as a result of the flow of

    asthenospheric matter under the continent

    (another competing uplift mechanism may be con-

    nected with thermal erosion of the lithosphere

    basement, its thinning and isostatic uplift); and

    (2) very sharp, large-amplitude, subsidence which

    occurs in a much narrower belt, caused mainly by

    the cooling and compaction of the upper lens of

    anomalous mantle, probably accompanied by

    phase transitions of the basalt-eclogite type

    (Artyushkov, 1979).

    The two main mechanisms presented cannot

    completely explain the evolution of continental

    passive margins within which, in accordance with

    a two-level model of plate tectonics (Lobkovsky,

    1989; Lobkovsky and Khain, 1989), one more

    geodynamic process affecting the development of

    its structure should occur. This is the flow in the

    lower ductile layer of the crust within a continen-

    tal margin at the spreading stage (Bott, 1972)

    (Figs. lob and c). Such a flow, from the continent

    towards the ocean, starts when the continental

    crust splits. By that time, as was mentioned earlier,

    both layers of the continental crust in the rift zone

    are already considerably thinned. When the conti-

    nents begin to move apart, the ductile lower crust

    at the edge of the continental margin tends to be

    squeezed out towards the ocean, as it is subjected

    to an uncompensated horizontal loading. Taking

    into account isostasy on the mantle surface, the

    outflow of the non-linear viscous material of the

    lower crust towards the ocean can be described by

    the evolution equation (3) of the “Mechanical

    aspects” section, derived in the Appendix.

    Let us consider the problem of a self-squeezed,

    semi-infinite layer of the lower crust when it is

    I

    I

    I

    I

    I

    I I

    -\

    ,

    -42 -60

    -48 -46 -44

    -42 a 42

    44 E

    Fig. 13. Oceanward and continent-ward propagation of a lower

    crust thickness inhomogeneity on a continental passive margin.

    (a) Numerical solution (values near the curves are time in Ma).

    (b) Asymptotic self-similar solution: I, for unfractured upper

    crust “cap”; 2, for dissected upper-crust cover of the “tongue”.

    (See text for explanation.)

    able to propagate horizontally after the continents

    start to move apart; that is, under conditions such

    as:

    H, xx> -1, (hl,i,

    h(,=,=O, hl,,_,=H

    A numerical solution of the problem of the

    flow for H = 18 km, /I = 10e9 kmP3 yr-’ and a

    small initial wedge or “ tongue” (I, = 5 km, I, = 10

    km) is shown in Fig. 13a. For longer times (t 2 20

    Ma), the propagation of the “tongue” towards the

    ocean and the depression front towards the conti-

    nent are well described by the asymptotic self-sim-

    ilar solution (Kerchman and Lobkovsky, 1990a):

    h=Hf(E),

    E= “-“‘:,a

    (BH7f)

    where function f satisfies the equation:

    4;r;i

    fS

    3 3++-J

    [ i

    )I

    (15)

    in the domain &, f 5) = 0 and for t < t2,

    f(E) = 1 (corresponding to an undisturbed state of

    the media into which the depression wave propa-

    gates at a finite rate). Moving towards the con-

  • 8/17/2019 Lob Kov Sky 1991

    17/32

    A TWO-LEVEL CONCEPT OF PLATE TECTONICS

    59

    tinental interior, such a front captures new por-

    tions of the lower crust in accordance with the law

    ,=W t)

    ‘/4, from which it is seen that the

    expansion of the flow of the lower crust decel-

    erates sharply with time. The function f(t) is

    given in Fig. 13b. Its construction is described in

    Kerchman (1990).

    The above analysis of ductile flow in the lower

    layer of the crust predicts the appearance of ten-

    sional stresses in the more brittle layers of the

    overlying “granite” crust and underlying rigid

    lithosphere. These tensional-stress maxima de-

    velop in the vicinity of the disturbed wavefront as

    it propagates towards the continental interior. The

    maximum tensional stress in the upper brittle layer

    of the crust is estimated as:

    1

    I

    x2

    e=-

    d

    rdx

    x1

    I Pch - P&( If2 -q

    %.&

    ,

    h1= Wx,) 07)

    where d is the thickness of the upper, mechani-

    cally strong layer of the brittle crust; X, is the

    section separating the non-fractured part of the

    brittle crust from the frontal, oceanward “frac-

    tured” area, including the zone of listric faults

    (Figs. 14a and b); and x2 is the continent-ward

    km

    b

    Fig. 14. Structure and presumable evolution of continental passive margin: (a) North Atlantic passive margin, Flemish Cap bank

    (after Emery and Uchupi, 1984); (b) model pattern of passive-margin dynamics (after Lobkovsky, 1989) (see text for explanation).

    I = Synrift sediments; 2 = post-rift sediments; 3 - upper “granitic” crust; 4 = lower “mafic” crust; 5 = subcrustal lithosphere;

    6 - anomalous mantle fens; 7 = normal mantle (asthenosphere.); 8 = fractured zone in subcrustal lithosphere; 9 = lower-crustal flow;

    10 = filtration in partially melted astbenosphere. (Values on the upper figure denote P-wave velocities in different layers.)

  • 8/17/2019 Lob Kov Sky 1991

    18/32

    MO

    L 1 LOBKOVSKY AND V 1. KERC’HMAN

    front of the flow propagation in the lower crust.

    For

    d =

    8-10 km, H = 15-20 km and h, = 5-10

    km, the mean tensional stress in the upper crust

    reaches u = 20-40 MPa. A long-term tension of

    this size seems sufficient to cause brittle rupture of

    the upper weaker part of the crust (Sawyer, 1985)

    rifting, and the formation of faults and effective

    “removal” of the blocks of the upper crust ocean-

    ward (Fig. 14).

    The proposed rifting mechanism operates

    through the lateral inhomogeneity and rheological

    stratification of the crust within the transition

    zone. It seems natural to attribute to it the ob-

    served evolution of rift structures that are sub-

    parallel to continental passive margins - near the

    continental slope, in the early stages of ocean

    opening, and at the rear of the transition zone in

    the later evolutionary stages (Fig. 14b). The same

    mechanism acts throughout the geological history

    of a passive margin, causing the splitting and

    subsequent break-off of continental blocks from

    the margins of large continents (Vink et al., 1984;

    Lobkovsky and Khain, 1989).

    An obvious result of the creep-propagation

    model of the lower-crustal “tongue” is the ap-

    pearance of a layer of anomalous basement rocks

    with a P-wave velocity of about 7.0 km/s (typical

    of the lowermost continental crust) in the transi-

    tion zone between the continental slope and nor-

    mal oceanic crust (Lobkovsky and Khain, 1989).

    It is clear from seismic cross-sections of the con-

    tinental passive margins that such an anomalous

    rock layer does, in fact, underly sediments in the

    transition zone, occupying a belt about 100-200

    km wide (Emery and Uchupi, 1984). Geochemical

    data also exist which provide evidence for the

    continental origin of the lower crust in the transi-

    tion zone of passive margins, even in cases where

    the upper-crustal layers are composed of basalts

    (Morton and Taylor, 1987). The model presented

    also explains the quiet magnetic field in this tran-

    sition zone (Boillot, 1983).

    Once local fracturing and the formation of a

    fault structure have occurred in the upper crust,

    the boundary conditions on the top of the ductile

    lower layer change. In particular, it shifts to the

    regime of a quasi-free horizontal displacement of

    the top boundary, leading to changes in the effec-

    tive coefficient in equations of the type (2) or (3).

    The coefficient ratio becomes /?,//? - 10 for the

    fractured, oceanward-moving tongue (see the Ap-

    pendix). This, in turn, causes an approximately

    double acceleration of the flow under the newly

    fractured part of the brittle crust, and the effective

    transport of the corresponding block toward the

    ocean, as well as substantial thinning of the dis-

    sected crust at the rear of the block and the

    consequent formation of a sedimentary basin there

    (Figs. 14a and b). The scheme described thus

    causes a discrete sequence of events: when the

    previous oceanward block of the upper continen-

    tal crust is faulted off, a new cycle of preparation

    for the subsequent rifting stage starts within its

    continent-ward part. According to the above anal-

    ysis, the first stages of marginal rifting (at an early

    stage in the ocean opening) often break off “small”

    blocks such as the Blake plateau, the Flemish Cup

    bank, the Rockall plateau, the Vijring plateau, the

    San Paulo plateau, the Exmouth plateau and

    others. When the “squeezing-out” of the lower-

    crustal layer decelerates (due to the cooling and

    blocking of the front of the tongue), the flow wave

    propagating toward the continent behind decel-

    erates. As the pressure gradients on the roof of the

    lower crust decrease, a strain sufficient for rifting

    accumulates at considerably greater distances from

    the continental edge and over the considerably

    greater time of hundreds of millions of years. This

    time can, for instance, even exceed the duration of

    one Wilson cycle in the Atlantic. The graben

    systems of China and the Rhine, the Labrador rift

    zone and others may be due to this process of

    rifting at large distances from the ocean. Very

    often, such distant rift zones are located in ancient

    weak sections of the subcrustal lithosphere (or of

    the crust), especially along suture zones (Vink et

    al., 1984; Dunbar and Sawyer, 1988). The ad-

    ditional tension of a quasi-rigid subcrustal litho-

    spheric core, caused by the flow of the lower crust,

    might reactivate the lower-level plate boundaries

    (in particular, transform and suture boundaries

    might change to divergent).

    This might cause complete splitting up of the

    lithosphere through formation of a divergent litho-

    spheric boundary in the lower level. This may

    occur long before the rupturing of the crust. The

  • 8/17/2019 Lob Kov Sky 1991

    19/32

    A TWO LEVEL CONCEPi OF PLATE TECTONICS

    361

    possibility of the complete disintegration of the

    continental lithosphere and a jump of an oceanic

    spreading axis to a new position at the continental

    margin arises when the above rifting rn~h~srn is

    superimposed on the general (“background”) ten-

    sion of the lithosphere. Such a background tension

    may be created by horizontal flow in the astheno-

    sphere and/or by a pull due to a subduction zone

    on the opposite edge of the oceanic part of the

    plate (Bott, 1982). This situation may have oc-

    curred during the structural evolution (destruc-

    tion) of Gondwanaland, when a powerful horizon-

    tal mantle flow (from the south to the north) and

    the “pulling” of a subducted oceanic part of the

    plate under Laurasia, together with the continu-

    ously acting mechanism of the marginal continen-

    tal rifting, caused the onset of new rift zones at

    the northern boundary of the Gondw~a super-

    continent. This mechanism may even have caused

    the subsequent break-off of blocks of microconti-

    nents, lithosphere terranes and corresponding

    jumps of the spreading axes to a new southward

    position (Sengiir, 1984; Kazmin, 1989; Lobkovsky

    and Khain, 1989).

    We note that the problem of the structure and

    evolution of passive continental margins has been

    studied extensively in the past decade. Various

    models for passive margins have been developed,

    particularly: models of different types of crustal

    and lithospheric extension (McKenzie, 1978; Le

    Pichon and Sibuet, 1981; Beaumont et al., 1982;

    Wernicke, 1985; Lister et al., 1986); models incor-

    porating a different rheology of the upper and

    lower parts of the continental crust during the

    extension (Bott, 1971, 1982; Meissner, 1985);

    “ volcanic” models of the continental margins

    (Royden et al., 1980; Mutter et al., 1988; Meissner

    and Kiipnick, 1988); models of isostasy of passive

    margins (Karner and Watts, 1982); thermomecha-

    nical models of the evolution of passive margins,

    taking sedimentation into account (Cloetingh et

    al., 1984) and others. All of these models consider

    various aspects of the structure and evolution of

    passive margins, many aspects of which may also

    be interpreted within the framework of the pro-

    posed two-level scheme.

    Modei fur ~~ q~-sy~ extension

    of

    the continental lithosphereand some featuresof

    rifting

    We now consider several aspects of continental

    rifting from the viewpoint of the two-level plate-

    tectonic model. A great deal of attention has re-

    cently been devoted to the problem of the out-

    crops of ultramafic mantle rocks on the Earth’s

    surface (oceanic bottom) as a result of the com-

    plete tectonic denudation of the continental crust

    during rifting. In fact, the data show the existence

    of anomalous ophiolitic sequences in which the

    basalt layer and dyke complex are completely

    missing, where oceanic sediments directly overlie

    serpentinized peridotites which are sometimes as-

    sociated with gabbro. Such incomplete ophiolitic

    sequences are typical of the Ligurian segment of

    the Mesozoic Tethys and occur, for instance, in

    the Alps, Apennines and Corsica (Lemoine et al.,

    1987). A similar anomalous structure of the oc-

    eanic crust has been seen in a few places on recent

    continental passive margins; particularly, on the

    Galicia Bank and in the North Atlantin (Boillot et

    al., 1987) on Zabargad island in the Red Sea

    (Bonatti et al., 1986), and on the Sardinia passive

    margin in the Tyrrhenian Sea (Lemoine et al.,

    1987), where serpentinites or peridotites have been

    found beneath sediments by deep-sea drilling.

    These data suggested to some geologists that, at

    least in the early stages of ocean opening, a special

    geodynamic regime exposed the ultramafic base-

    ment between diverging blocks of the continental

    lithosphere (Boillot et al., 1987; Lemoine et al.,

    1987). A model of the splitting and asymmetrical

    divergence of continental lithosphere along a gen-

    tle, through-going fault has been proposed to ex-

    plain the formation of rather extended segments

    of oceanic bottom (now composed of se~ent~iz~

    peridotites) at divergent plate boundaries. This

    model was first introduced by Wernicke (1981) in

    an analysis of the tectonic situation in the Basin

    and Range Province in the western U.S.A., and it

    has subsequently been widely used for various

    schemes of rifting (Wemicke, 1985; Lister et al.,

    1986).

  • 8/17/2019 Lob Kov Sky 1991

    20/32

    362

    From the standpoint of mechanics, one cannot

    ignore the fact that in the Wemicke scheme the

    postulated complete fracture of the lithosphere

    along the single, gently dipping fault contradicts

    the entire sum of our knowledge of the rheological

    stratification of the crust and lithosphere. It seems

    more realistic to suppose that the fracture and

    deformation of the various layers of the crust and

    lithosphere in the course of the rifting occur

    according to the particular flow laws acting in

    each layer.

    Another purely mechanical objection to the

    Wernicke scheme arises from the fact that the

    lithosphere cannot bend sharply without sec-

    ondary faulting during the normal slip of the

    adjacent walls of the fixed inclined fault. As early

    as 1958, Heiskanen and Vening-Meinesz, in their

    L..l LOBKDVSKY AND V.I. KkKCHMAN

    fundamental work, showed that during crustal ex-

    tension, for instance, a second fault appears due

    to the elastic bending of the crustal layer and

    results in the formation of a graben structure. It

    follows from this argument that even if we assume

    an initial fracture of the lithosphere in accordance

    with the Wemicke model - a single, gently dip-

    ping fault - the further development of extension

    would follow another course (see below) and would

    be determined mainly by the appearance of an

    additional fault in the lithosphere and by the

    motion of a lithospheric block cut off by it (Us-

    sami et al., 1986).

    We now describe a new model of the develop-

    ment of continental rifting, proposed by Lobkov-

    sky (1989). This scheme is guided by the estab-

    lished rheological and tectonic stratification of the

    b

    d

    Fig. 15. Two-level model of continental rifting: successive stages (after Lobkovsky, 1989).

    1 =

    Upper brittle crust; 2 = lower ductile

    crust; 3 = subcrustd lithosphere; 4 = asthenosphere; 5 = volcanics. (See text for explanation.)

  • 8/17/2019 Lob Kov Sky 1991

    21/32

    A ‘IWO-LEVEL CONCEPT OF PLATE TECTONICS

    lithosphere. Derived from general mechanical con-

    siderations and from the results of physical mod-

    elling of the extension of an elastic-plastic litho-

    sphere floating on a liquid basement (Shemenda,

    1984) this model postulates that the loss of stabil-

    ity and strain localization in the quasi-rigid mantle

    core of the lithosphere would cause the formation

    of two conjugate inclined shear planes. These shear

    planes bound a central subcrustal lithospheric

    block of a wedge-like or of a trapeziform geome-

    try, which would rise under the influence of the

    applied system of forces and squeeze the viscous

    material of the lower-crustal layer away the axis

    (Lobkovsky, 1989) (Figs. 15a and b). This, in turn,

    would cause thinning of the crust (neck-forma-

    tion), as well as additional extension and isostatic

    subsidence of the upper brittle layer. It is im-

    portant to stress that in this scheme the material

    flow that leads to the thinning of the lower ductile

    layer of the crust is induced not so much by the

    external tensional force applied to the crust but by

    the squeezing effect of the rising mantle block

    (Fig. 15b).

    Unlike the majority of the proposed schemes

    (McKenzie, 1978; Le Pichon and Sibuet, 1981), in

    this theory the crust is thus not subject to uniform

    363

    extension during continental rifting. This permits

    us to explain the observed discrepancy between

    the degree of extension of the crust, determined

    from the system of faults, and the estimates of the

    extension of the crustal layer derived from its

    thinning and the isostatic subsidence of the surface

    (Artyushkov, 1988). In fact, the analysis of the

    structure of recent and ancient large depressions

    within continents shows that in the majority of

    cases crustal rifting involves extensions of only

    several percent, whereas the thinning of a con-

    solidated crust determined from seismic data is

    50-lOO’%, i.e. approximately one order of magni-

    tude greater.

    Let us consider the above-described model in

    detail, highlighting the principal stages of rifting

    (Lobkovsky, 1989). These stages, which are char-

    acterized by the occurrence of two inclined fault

    planes, dipping from the axis of the future rift,

    and formed in the mantle part of the lithosphere

    during its extension, are illustrated in Figs. 15a-c.

    As extension increases, a trapeziform block of

    subcrustal lithosphere, bounded by fault planes,

    begins to move upward under the influence of the

    non-u~fo~y applied stresses (Fig. 16). This

    movement was verified by physical modelling ex-

    b

    II..*....

    . . . . . .

    * -   . . . . , . s .

    . . . . . . . .

    Fig. 16. Rift scheme for cakulation of (a} balanceof forces and (b) uplift amplitudeof the axial sub~~phe~ block {see text for

    explanation). I = Weight of the axial block 1rl; 2 = upper brittle crust; 3 = lower ductile crust; 4 = subcrustal lithosphere;

    S = asthenosphere.

  • 8/17/2019 Lob Kov Sky 1991

    22/32

    364

    I..1 LOBKOVSKY AND V.I. KFKCHMAIU

    perimems, taking similarity criteria into account

    (Shemenda, 1984).

    The qualitative explanation of this phenome-

    non is as follows. When the ~thosphere blocks Z

    and ZZ move away from the central block ZZZ see

    Fig. 16a), the average pressure on its inclined

    planes AB and CD drops and, as a result of the

    buoyancy force applied to the base of block ZZZ, t

    becomes unbalanced. A subsequent drop in the

    pressure, transmitted from the overhanging parts

    of blocks I and ZZ, as they move away, occurs due

    to their own elasticity.

    Let us consider the uplift of block ZZZquantita-

    tively and estimate its ultimate amplitude. We

    assume that block ZZZ protrudes into the plastic

    layer of the lower crust by an amount Ah. Then

    the total hydrostatic pressure on the base of the

    block equals:

    where ZZ, and H, are the thicknesses of the crust

    and of the mechanically strong subcrustal part of

    the lithosphere, respectively; p, and p, are their

    densities; 1 is the width of the upper surface of the

    uplifted block; and a is the inclination of the side

    planes of the block (Fig. 16a). The weight of block

    ZZZ per unit of length in the plane, perpendicular

    to the figure) is p&Z 4 H, cot cu)ZZ,; the pres-

    sure on its upper plane is p,g(ZZ, - Ah)Z. The

    vertical component of the pressure of the over-

    hanging lithospheric blocks Z and ZZ on the side

    planes of block ZZZcan be calculated as:

    [2p,H,H, - pc(Ahj2 -t P&Z,,, - Ah)‘] g cot a

    i-2

    J

    Hm-Ah(r - Ao cot a) dt

    0

    where the first term of this expression contains the

    pure hydrostatic part of the forces, whereas the

    second term reflects the deviation from the hydro-

    statics on the side planes of block ZZZ n the lower

    quasi-rigid lithospheric layer due to (a) the drop

    Au of pressure transmitted from the overhanging

    regions; and (b) the friction r along the inclined

    faults. From the balance of forces applied to block

    ZZZ and its weight, it is seen that at Ah = 0 (the

    initial state after the inclined faults are formed;

    see (Fig. 16a) the net force affecting block ZZZ is

    negative (oriented upward), if the following condi-

    tion is satisfied:

    Aa > 7 tan cy

    It should be noted that the magnitude ha is

    proportional to the geomaterial strength up to its

    fracture, whereas r has the meaning of a residual

    shear strength; then from eqn. (19) it follows that

    the uplift of the block may start when the angle a

    is not very large, i.e. approximately 0 2 60 ‘.

    On the other hand, angle LT annot be too small

    (as is assumed in the Wemicke model; Wemicke,

    1981); hence, if angles (Y are small, bending of

    relatively extended near-fault parts of blocks Z

    and ZZof a quasi-rigid lithosphere begins to play a

    main role, and this would cause the secondary

    faulting of the lithosphere. Such secondary faults

    are proposed in the paper by Ussami et al. (1986).

    Let us estimate the ultimate amplitude of the

    uplift of mantle block ZZZ. First we note that, as

    the block uplifts, friction 7 in the faults would

    decrease due to the frictional heating of the media

    and other factors. Ignoring friction 7 at the ma-

    ture uplift stage of the block and assuming the

    magnitude of Aa being approximately equal to the

    mean strength a, of a quasi-rigid subcrustal litho-

    sphere, we would write the condition of the force

    balance as:

    2u,( ZZ, - Ah) cot IY=

    Apg[iAh + (Ah)2 cot CY]

    (20)

    where Ap = pm - pc.

    For typical values of the parameters which are

    used in eqn. (20) (Ap 5=0.4 g/cm3, g = lo3 cm/s*,

    I = 20-30 km, a* = 100-300 MPa, ZZ,,,= 30-50

    km), we have the uplift amplitude of block Ah -

    lo-25 km. This means that, in the course of

    rifting, a mantle block may be uplifted to the base

    of the upper brittle layer of the crust. When a

    quasi-rigid mantle block is uplifted, it squeezes

    aside plastic matter of the lower crust, the flow of

    which - in turn - causes additional extension of

    the upper brittle crustal layer (Fig. 15b). Thus,

    non-uniform extension of the crust occurs and its

    lower ductile layer undergoes almost complete de-

    nudation, whereas the upper brittle layer is ex-

    tended and becomes thinner, although only by

    several percent (Lobkovsky, 1989).

  • 8/17/2019 Lob Kov Sky 1991

    23/32

    A TWO-LEVEL CONCEPT OF PLATE TECTONICS

    365

    If the lithospheric extension ceases at

    this stage,

    a structure typical of a continental depression of

    rift origin is formed, which has a sharply thinned

    consolidated crust and a slight extension on its

    surface. Some large aulacogens within continents

    seem to have been formed according to this model.

    Another situation is possible when much more

    extension of the upper brittle layer of the crust

    leads to its complete splitting. This may lead to an

    exposure of mantle rocks directly on the bottom

    surface of the opening oceanic basin (Fig. 1%).

    Thus the proposed model may explain the ex-

    istence of mantle rocks on some parts of the

    oceanic bottom which correspond to the initial

    stages of the ocean’s formation (in particular, on

    passive continental margins). From this stand-

    point, the Ligurian ophiolites seem to be attri-

    buted to a rather narrow oceanic basin of the type

    described.

    An important consequence of the described

    non-magmatic model of continental rifting is as

    follows. Since the uplift amplitude of the mantle

    block is proportional to the thickness of the

    quasi-rigid core of a subcrustal lithosphere H,,

    thinning of this strong core may hinder a rather

    considerable uplift of the central block. In such a

    case, a further divergence of lithospheric blocks I

    and II cannot be compensated by the uplift of

    block 111, and their adjacent sides would move

    apart. The asthenospheric matter would flow into

    the gaps formed as a consequence of this diver-

    gence (Fig. 15d). As a result, two large channels

    would be formed at the lateral sides of central

    block 117, which would provide for the uplift of

    the as~enosphe~c matter to the base of the crust

    and further onto the Earth’s surface as alkaline

    basalt outflows (Lobkovsky, 1989) (Fig. 15d).

    It should be noted that the existence of mag-

    matic channels along the margins of central block

    111 may explain a very curious feature of rift

    volcanism: the majority of large volcanoes are

    usually attributed not to the central, but to the

    peripheral parts of the rift zone. Powerful perip