13 Intracratonic Basins PDF

You might also like

Download as pdf
Download as pdf
You are on page 1of 20
Intracratonic Basins 13 George D. Klein Introduction. Intracratonic basins occur within continental interiors away from plate margins. They are oval in plan and saucer-shaped in cross section. Intracratonic basins are floored with continental crust, and in most instances, are also underlain by failed or fossil rifts (sce Chapters 2 and 3). Their evolution involves a combination and succession of _basin-forming processes, which include continental extension, ther- mal subsidence over a wide area, and later isostatic readjustment. Controversy has surrounded understanding of the origin and evolution of intracratonic basins. Eleven hypotheses (Klein, 1991) for the formation of intracratonic basins have been proposed and include: (1) increase in density of the crust by an eclogite phase transformation (Michigan basin, Haxby et al., 1976; Williston basin, Fowler and Nisbet, 1985), (2) nifting associated with impingement of a thermal plume at the base of the lithosphere (Illinois basin, Burke and Dewey, 1973; McGinnis et al., 1976), (3) thermal metamorphism of the lower crust to bound- ary conditions of the greenschist and amphibolite facies (Australian intracratonic basins, Middleton, 1980), (4) mechanical subsidence caused by isostati- cally uncompensated excess mass of igneous intru- sions (Williston basin, DeRito et al., 1983), (5) tec- tonic reactivation along older structures (Michigan basin, Howell, 1989; Parana, Maranhao, and Ama- zonas basins, Brazil, DeBrito Neves et al., 1984), (6) thermal subsidence (Illinois basin, Sleep, 1976: Sleep and Snell, 1976; Sleep et al., 1980; Heidlauf et al., 1986; Williston basin, Ahem and Mrkvicka, 1984; Ahern and Ditmars, 1985; Michigan basin, Sleep et al, 1980; Nunn et al, 1984; Nunn, 1986; Hudson Bay basin, Quinlan, 1987; Amadeus basin, Australia, Lindsay and Korsch, 1989), (7) partial melting in lower crust and drainage of resulting igneous melt to mid-ocean ridges by volcanism, resulting in basin subsidence above the zone from which igneous melt ‘was removed (Sloss and Speed, 1974), (8) changes in intraplate stress, assuming a visco-elastic plate (Aus- tralian intracratonic basins, Lambeck, 1983a,b; Shaw et al., 1991) or an elastic plate (all intracratonic basins, Karner, 1986), (9) thermal subsidence follow- ing intrusion of anorogenic granites in response to changing heat flow during supercontinent breakup (all intracratonic basins, Klein and Hsui, 1987), (10) thermal subsidence followed by subsidence caused by an isostatically uncompensated excess mass of cooling igneous intrusions (Illinois basin, Treworgy et al., 1989; Kolata and Nelson, 1990a,b; Hamdani et al., 1991), and (11) subsidence caused by tectonic events at adjacent plate margins (Howell and van der Pluijm, 1990; Leighton and Kolata, 1990). In North America (Fig. 13.1), subsidence analysis has shown that initiation of subsidence of the Illinois, Michigan, and Williston basins, and initiation of sub- sidence of latest Precambrian and earliest Paleozoic passive margins were coeval with the age of breakup of a late Precambrian supercontinent (Bond and Kominz, 1984, 1991; Bond et al., 1984; Klein and Hsui, 1987). In Australia, initial formation of late Pro- terozoic cratonic basins was coeval with the breakup of a supercontinent (Lindsay et al., 1987; Korsch and Lindsay, 1989; Lindsay and Korsch, 1989). Similar timing of subsidence followed breakup of Pangea in the Paris basin (Brunet and LePichon, 1982; Perrodon and Zabek, 1990), the North Sea (Ziegler, 1977, 1982, 1987, 1988), and Gondwana basins of India (Kailasham, 1976). Sloss (1972, 1979) and Sloss and ‘Scherer (1975) observed that systematic synchronous 459 460 Chapter 13 r wr ae a a wd ‘CANADA a \ oo \ Q cso 8 Witten (3S eosin wo Minos Bash oo extco™ Lg Ree aoe vor tage tor ee KLEIN - InvacratoneFigue 19.1 Fig. 13.1 Map of North Amer, showing location of major nacre tone basins changes in sediment volume were characteristic of intracratonic basins in North America, Brazil (Soares et al,, 1978; DeBrito Neves et al., 1984), the Russian platform, and Africa (Peters, 1979). Moreover, conti: nentwide interregional unconformities characteristic of Sloss’s (1963, 1988a) cratonic sequence boundaries established in North America are found in other intracratonic settings, including eastern Europe and the former Soviet Union (Sloss, 1972, 1976), Brazil (Soares et al., 1978; DeBrito Neves et al., 1984; Zalan et al., 1990), and Africa (Peters, 1979). These strati- graphic observations indicate a commonality in time and space of formation of intracratonic basins in response to a global process, which Sloss (1988a,b) emphasized is tectonic. This chapter focuses on some of these problems and suggests an hypothesis not only for the origin of intracratonic basins, but also for their global com- monality. This chapter summarizes recent observa- tions that clarify some of the earlier explanations for mechanisms of formation of intracratonic basins listed above, as well as new data that permit resolu- tion of much of the controversy regarding the forma- tion of intracratonic basins. This chapter also summa- Tizes what is known about North American Paleozoic intracratonic basins, including the Illinois basin (Sleep, 1971, 1976; Sleep et al., 1980; Heidlauf et al, 1986; Treworgy et al., 1989; Kolata and Nelson, 1990), the Michigan basin (Sleep and Sloss, 1978; Nunn et al., 1984; Nunn, 1986; Ahern and Dikeou, 1989; Howell, 1989), and the Williston basin (Ger- hard et al., 1982, 1990; DeRito et al, 1983; Ahern and Mrkvicka, 1984; Ahem and Ditmars, 1985; Kent, 1987; Gerhard and Anderson, 1988), as well as selected intracratonic basins, such as the Amadeus basin of Australia, and the Parana basin of Brazil Intracratonic Stratigraphic Sequences Within intracratonic basins and adjacent platforms, stratigraphic subdivision differs from international stratigraphic subdivisions. Intracratonic stratigraphic subdivision follows the now well established concept of “Cratonic Sequences” of Sloss (1963, 1988a; also see Leighton and Kolata, 1990). Sloss (1963) recog: nized that the stratigraphy of the North American cra- ton is subdivided into six cratonic sequences, which are bounded by major interregional (continent-wide) unconformities. These cratonic sequences are (from oldest to youngest): Sauk (590 to 488 Ma), Tippeca- noe (488 to 401 Ma), Kaskaskia (401 to 330 Ma), Absaroka (330 to 186 Ma), Zuni (186 to 60 Ma), and Tejas (60 Ma to present). Each cratonic sequence records a nearly complete trangressive-regressive sedi- mentary cycle, Internally, these cratonic sequences are subdivided into subsequences separated by lesser regional unconformities; these sequences include basal onlap and are capped by minor offlapping sequences below the unconformities. It is the writer’s view that Sloss’s North American cratonic sequences are comparable to the classic European geological sys- tems and are unique to intracratonic settings in other regions of the world, including the Russian platform (Sloss, 1972, 1976, 1979), Brazil (Soares et al., 1978; DeBrito Neves et al., 1984), and Africa (Peters, 1979). Intracratonic areas outside North America show coeval, continentwide interregional unconformities, and paralle! temporal trends to areal distribution and sediment volume (Sloss, 1972, 1976, 1979). Although the cratonic sequences were defined from a single tec- tonic setting (intracratonic) in North America, they are observed globally (Soares et al., 1978; Peters, 1979; Sloss, 1979). By contrast, the classic European, type sections were recognized and established in di fering tectonic settings (e.g., active margins, rift basins, platforms, intractatonic basins, and foreland basins), a fact which has contributed to never-ending arguments over global correlation of system bound- aries in different tectonic domains. Summary of North American Intracratonic Basins Four major intracratonic basins are known from North America (Fig. 13.1). Of these, the Illinois, Michigan, and Williston basins are summarized below. Mlinois Basin ‘The Illinois basin is oval in plan (Figs. 13.1, 13.2). Six- thousand meters of sedimentary strata accumulated within this basin during Paleozoic time (Figs. 13.3, 13.4; Kolata and Nelson, 1990a,b; Nelson, 1990a,b). Initially, sediments accumulated on a south- facing embayed ramp (Reelfoot aulacogen; McGinnis, 1970; McGinnis et al., 1976; Kolata and Nelson, 1990a,b; Nelson, 1990a,b). Uplift of the Pascola arch on its south side closed this embayment between Pennsylvanian and Cretaceous time (Sterns, 1957; Marcher, 1961; Bethke, 1985). Controversy exists concerning basement structure. Magnetic and gravity anomalies and the location of Holocene earthquake epicenters (e.g. McGinnis, 1970; McGinnis et al, 1976; Braille et al., 1982, 1986; Hildenbrand, 1985) suggest that the Reelfoot aulacogen (Fig. 13.2) was part of a three-arm rift sys- tem, Moreover, seismic data suggest that the crust is thin below the depocenter and the Reelfoot aulaco- gen (Braille et al., 1986). More recent seismic surveys (eg., Pratt et al., 1989; Bretagne and Leising, 1990; Heigold, 1990; Heigold and Oltz, 1990) and structural mapping (e.g., Nelson, 1990a,b) show that the base- ment consists of bedded volcanics; only one of the proposed three arms, the Rough Creek graben, could be identified as extending east of the depocenter from the Reelfoot aulacogen (Fig. 13.2). To what extent Intracratonic Basins 461 or 925 50 a Edje of Chesterian Rough Creek Graben KLEIN - Intracratonie - Figure 13.2 Fig. 13.2. Map showing location of theee wells used for tectonic subsidence analysis (Fg. 13.5) in lino basin (basin boundary marked by edge of Cheserian Series) Location of Resfoot rit and Rough Creek staben, both stippled, afte Nelson (19900). 1 = lis IN = Indiana Y= Kentucky: MO = Misou the boundaries of the other two arms (mapped by earlier geophysical studies) represent localized faults, remains problematic. Even the origin of the Rough Creek “graben” now appears to be in dispute because recent work by Goetz et al. (1992) shows it to have been formed in a strike-slip zone involving transpres- sion throughout its history. I recomputed the tectonic-subsidence history of the Ilinois basin using stratigraphic data from Heidlauf ct al, (1986; their tables 1-3) and the Harland et al. (1990) time scale. Methods used for this computation were those summarized in Sclater and Christie (1980) and Bond and Kominz (1984). Tectonic-subsidence analysis of three deep wells in the basin depocenter (Fig. 13.5) shows that the basin formed by initial, rapid fault-controlled subsidence associated with rift- ing and crustal thinning, followed by a period of rapid thermal subsidence that lasted about 60 my (shown by an upward concave profile beyond the initial steep, SERIES _ SEO. arn Grew” Tenrany [ova PLEISTOCENE PLIQCENE | TEJAS ZUNI PERM] WOLFCAMPIAN ‘CARBONIFEROUS luPpenl VIRGILIAN MISSOURIAN sub-Zuni uncontormity, MIDDLE LOWER DESMOINESIAN ATOKAN MORROWAN CHESTERIAN VALMEYERAN ‘ABSAROKA, ‘CHAUTAUQUAN: —\SENECAN ERIAN cele ULSTERIAN SILURIAN [DEVONIAN] MISSISSIPPIAN | _ PENNSYLVANIAN CAYUGAN NIAGARAN ‘ALEXANDRIAN KINDERHOOKIAN| KASKASKIA ‘ORDOVICIAN LOWER CINCINNATIAN TIPPECANOE ™.] UPPER [LOWER] U. CHAMPLAINIAN CANADIAN ‘CAMBRIAN UPPER ST. CROMKAN ALBERTAN owen | mDOLE | WAUCOBAN SAUK ‘sub-Tippecance i ‘unconformity New Magra it Complex ‘sub-Sauk unconformity Porsiaui Fig. 13.3 Schematiesratigraph section showing relationship of strat sraphic megagroups,lholaces, series and cratonc sequences of Sloss (1963) in the Ilinois basin, (Redraw alter Buschbach and Kola, 1990.) line). Thermal subsidence was verified by comparing the elevation of basement depth with t* (Fig. 13.5); the linear trend that was established showed strong Positive correlation coefficients ranging from 0.95 to 0.98 (Heidlauf et al., 1986), as outlined by McKenzie (1978a). This phase of thermal subsidence began around 510 Ma and continued until 450 Ma. Heidlauf et al. (1986) observed that, during this 60 my thermal-subsidence phase, subsidence rates sud- denly increased at various times. These changes in subsidence rates were attributed originally to changes in sea level (e.g., Sleep and Snell, 1976; Sleep et al., 1980; Heidlauf et al., 1986) or inaccuracies in earlier time scales (e.g., Heidlauf et al., 1986). However, sea level was rising continuously during this time interval (Watso and Klein, 1989), so sea-level changes appear not to have caused these departures from subsidence rate. Davis (1987, 1990), Clendenin (1989), and Watso and Klein (1989) suggested, instead, that dur- ing the initial phase of rapid thermal subsidence, regional transfer faulting (Clendenin, 1989) and asso- ciated normal faulting contributed to additional subsi- dence. During the earliest phase of thermal subsi- dence, an interval of overlapping, but diminished, mechanical subsidence occurred: this faulting may be attributed 10 progressively increasing lithospheric rigidity with cooling (Karner et al., 1983). Quinlan (1987) reported a similar combination of mechanical and thermal subsidence. tes ost 00 MoM a PRECMORIN Fig, 13 Westcast stratigraphic cos section of linols basin, approximately 100 km north of depocenter. (Redrawn ater Collinson tal, 1988) Intracratonic Basins 463, Rapid thermal subsidence was followed by slower subsidence, which Heidlauf et al. (1986) also inter- preted as thermal subsidence (cf., McKenzie, 1978a; Korsch and Lindsay, 1989). More recently, Treworgy et al. (1989) and Kolata and Nelson (1990a,b) sug- gested that this slower phase of subsidence, which lasted from approximately 450 to 360 Ma, was caused by an isostatically uncompensated excess mass (cf. DeRito et al., 1983). This model of DeRito et al. (1983) requires large-scale tectonic stresses. Tectonic events known in the Illinois basin started during latest Mississippian and Pennsylvanian time (Nelson and Krause, 1981; Nelson and Lumm, 1984; Kolata and Nelson, 1990a,b). Alternatively, in my view, after the relatively rapid phase of thermal subsidence lasting 60 my, progressive cooling of crustal magma would cause a phase change below the basin and subsidence would occur in response to a different style of isostatic compensation (cf., Ham- dani et al., 1991), particularly as lithospheric rigidity increased (Karner et al., 1983); it would be accompa- nied also by a concurrent slower rate of thermal subsi- dence and decay (cf., McKenzie, 1978a). Thus, I con- clude that this period of subsidence (450-360 Ma) was caused by a combination of late stages of decaying thermal subsidence and an isostatically uncompen- sated excess mass caused by a phase change associated with rift-related magmatic cooling (e.g., Hamdani et al., 1991), independent of large-scale compression. During Carboniferous time, the rate and magni- tude of basin subsidence increased a second time (Heidlauf et al., 1986). Collision tectonics of the Alleghenian-Hercynian-Ouachita orogeny occurred along the eastern and southern margins of North America and caused widespread regional compressive stresses (Craddock and van der Pluijm, 1989), which yielded a consistent regional fracture trend, associ- ated strike-slip faulting in the Ilinois basin (Nelson and Krause, 1981; Nelson and Lumm, 1984; Nelson, 1990a,b), and altered sedimentation patterns (Greb, 1989). Subsidence was reactivated by a far-field tec- tonic effect involving tectonic forces at the margin, which influenced a widespread area in the interior of, the continent (e.g., Kluth and Coney, 1981; Kluth, 1986; Ziegler, 1987, 1988; Leighton and Kolata, 1990); in this case, thrust loading on eastern and concurrent 464° Chapter 13 ig. 135 Let decompaced and o o backstripped tectonic subsidence for Texas Pacific Co No #1 Fa z ley Weil upper), Texas Pace = Co. No. AT Stich Wel i ga 1 dle) and Exxon Comp. No. # 8 Choice Duncan Well ower, nos basin. Data from Held #1 Farley Farley laut etl (1986), ecomputed 2 according to time scale of Har- 00 do aa aaa + + TE hand etal. (1990). Right: Com: —_ pason of tectonicsubsidence Age (Ma) ‘Sqrt. Time (Vmy ) ‘with respect to t? for Farley (upper, Streich (idae), and Duncan (ower) wel ins asin. See Fig. 142 for locations) o 9 £ £4 a #1 Streich “aft Steicn “595 ao at ata + $ 3 ‘Age (Ma) Saget. Time (VAY) o = g4 8 _|_Stoice Duncan #1 -a|__Shoice Duncan #1 a + + 7% 76 ‘Age (Ma) Sqr. Time (V7AV) southern Laurussia caused major tectonic deforma- tion and subsidence in the continental interior of North America (Quinlan and Beaumont, 1984; Beau- mont et al., 1987). Thrust loading caused widespread flexural stress (ef. Craddock and van der Pluijm, 1989), which was transmitted through the litho- sphere; the associated flexural wave reactivated and focused basin subsidence along older trends in the Ilinois and Michigan basins, as well as in the central Appalachian foreland basin (cf., also Karner et al., 1983; Lambeck, 1983a,b; Quinlan and Beaumont, 1984; Karner, 1986). This thrust loading may have caused uplift and inversion of a Precambrian rift basin on the Cincinnati arch (Shrake et al., 1991). Wide- spread intracratonic basin subsidence in North Amer- ica occurred during three periods of subsidence asso- ciated with Late Ordovician, Late Silurian-Early Devonian, and Carboniferous orogenic events (Willard and Klein, 1990). The thermal age of the crust below the Appalachian foeland basin is approxi- mately 600 Ma (Bond et al., 1984; Willard and Klein, 1990); lithospheric rigidity increased gradually with time and during each successive orogenic event, as confirmed by decreasing rates of subsidence through- out Paleozoic time (Willard and Klein, 1990). Conse- quently, foreland-basin subsidence was shallower and broader during the Alleghenian-Hercynian orogeny (see Chapters 9 and 11). Distribution of sedimentary facies in the central Appalachian basin fits such a geo- dynamic model, inasmuch as the basin was filled with submarine turbidites during the Late Ordovician, tur- idites overlain by deltas during the Late Devonian, and fluvio-deltaic sediments during the Pennsylva- nian (Tankard, 1986; Willard and Klein, 1990). Such shallower and broader subsidence favored lateral transmission of compressive stresses, which reactivated subsidence in the Illinois basin during Middle and Late Carboniferous time, and reactivated Precambrian faults, causing uplift along the eastern boundary of the basin, the Cincinnati arch (Shrake et, al,, 1991). This subsidence was associated also with deposition of Pennsylvanian cyclothems (Tankard, 1986), which were influenced by concurrent tectonic and climatically induced cyctic processes (Klein and Willard, 1989; Klein, 1992a; Klein and Kupperman, 1992). This second phase of subsidence in the Illino basin was associated with intrusion of Late Pennsyl- vanian and Permian alnoites, derived by partial melt- ing of the lower crust (Zartman et al., 1967). Their presence suggests renewed extension in the Reelfoot aulacogen (Klein and Hsui, 1987). Sedimentary systems in the Illinois basin range from fluvial to coastal to shelf, with local turbidite units below deltaic strata (see Heidlauf et al., 1986, their Fig. 6; Kolata and Nelson, 1990a, their Fig. 18-1; Leighton et al., 1990, p. 75-164, for recent sum- maries and syntheses). Cambrian fluvial and coastal deposits occur in the Reelfoot aulacogen, and onlap- ping strata occur immediately above each of the cra- tonic sequences reported by Sloss (1963, 1988), Collinson et al. (1988), and Leighton et al. (1990, p.75-164). Cambrian fluvial-coastal deposition was associated with rifting and thermal subsidence of the Ilinois basin (Heidlauf et al., 1986); as thermal subsi- dence diminished, deposition of carbonate platform tacies was dominant. The source of lower and middle Paleozoic clastic deposits appears to have been the Canadian shield (Potter and Pryor, 1960). Intracratonic Basins 465 During Silurian time, extensive pinnacle reefs developed; these are major reservoirs for petroleum and form a linear trend along the basin axis (Whi- taker, 1988). Carbonate-platform deposition persisted until latest Mississippian time. During Alleghenian- Hercynian orogenic movements in the Appalachian and Ouachita regions, the Illinois basin was linked with the continental margin by flexural tectonics (Quinlan and Beaumont, 1984; Beaumont et al., 1987) and the depositional motif was dominated by coal-bearing cyclothems, particularly during Pennsylvanian time (Klein and Willard, 1989). Deposition of these cyclothems was influenced concurrently by regional flexural tectonics, glacial eustasy, and long-term cli- mate change (Klein and Willard, 1989; Cecil, 1990; Klein, 1992a,b; Klein and Kupperman, 1992), Fluvial systems controlled by concurrent faulting (e.g., Greb, 1989) and extensive deltas characterized deposition of these cyclothems. The principal source of clastic com- ponents of these cyclothems appears to have been the Appalachians (Potter and Pryor, 1960; Greb, 1989). The Illinois basin has been a major oil-producing province since oil was discovered in 1886 (Leighton et al., 1990). Principal source beds are in the upper- most Devonian to lowest Mississippian New Albany Shale (Cluff and Byrnes, 1990) and the Ordovician Maquoketa Shale, both of anoxic relatively deep- water origin, Other sources include Pennsylvanian coals, and the Cambrian intertidal Eau Claire Shale Reservoirs occur primarily in the Silurian (pri- marily reefs), Devonian (shelf carbonates), Missis- sippian and Pennsylvanian systems (Howard, 1990; Seyler and Cluff, 1990). The Mississippian reser- voirs are diverse, ranging from oolitic shoals in the Salem Limestone, to tidal sand bodies of the Aux Vases Sandstone. Pennsylvanian channel and delta- distributary deposits comprise remaining reservoir types. ming and patterns of oil migration have been modelled relatively recently (Bethke, 1985; Bethke and Marchak, 1990; Bethke et al., 1991). Because the Ilinois basin was a cratonic embayment through most of Paleozoic time (Kolata and Nelson, 1990a,b), it did not assume its present shape until the Pascola arch on its southern boundary was uplifted as early as Permian time (Bethke, 1985; Bethke et al., 1991) 466 Chapter 13 With this uplift, topographically driven gravity flow of basin fluids was initiated to the north. These fluids were heated during migration into the basin depo- center, where temperatures were higher, and this thermal effect caused maturation of oil from the prin- cipal source beds. Lateral migration into reservoirs ‘occurred over a wider region (Fig. 13.6). ‘Trapping mechanisms show considerable variabil- ity (Seyler and Cluff, 1990), with structural traps (ie., anticlines and faults) being dominant; facies changes and lateral pinchouts provide important stratigraphic traps, particularly in the Pennsylvanian and the Mississippian, and also in Silurian pinnacle reefs. Seals consist mostly of overlying shale, or shale bounding one of the walls of a fault-bounded trap. Williston Basin The Williston basin is also saucer-shaped, with an oval plan view (Fig. 13.1). It is filled with 3700 m of sedimentary strata, which are also subdivided accord- ing to Sloss’s (1963, 1988a) cratonic sequences (Fig. 13.7), Strata consist mostly of carbonate rocks and evaporites, and smaller volumes of sandstone and Petroleum Production Pig. 13.6 Areal distribution of petroleum production and thermally mature, oil:prone New Albany Shale (Upper Devonian - Lower Misi sippian,llinois basin. Edge of Iinls basin represented by heavy black line with white markings Redrawn from Bethke and Marshak, 1990) shale; representing all periods of the Phanerozoic (eg. Gethard et al., 1982, 1990; Peterson and Mac- Cary, 1987; Gerhard and Anderson, 1988). These strata include coastal to shallow-shelf deposits, and Devonian pinnacle reefs. Drilling to basement indi- cates that the basin is underlain by Archean base- ment (see Hoffman, 1988), Proterozoic suture zones, a major transform fault, and a province of anorogenic granite (Kent, 1987; Gerhard et al., 1990). The Williston basin underwent thermal subsidence (Fig. 13.8) starting from about 520 to 500 Ma (Gerhard et al, 1982, 1990; Ahern and Mrkvicka, 1984; Ahern and Ditmars, 1985; Fowler and Nisbet, 1985; Gerhard and Anderson, 1988). This rapid phase of thermal sub- sidence was followed by both a lower rate of thermal subsidence coupled with subsidence due to an isostati- cally uncompensated excess mass during middle Paleo- zoic time (e.g., DeRito et al,, 1983). A Proterozoic failed rift underneath the Williston basin has been inferred both by Stewart (1972) and Kent (1987); this rift was, reactivated (DeRito et al, 1983; Quinlan, 1987). A transtensional basin precursor may also have been associated with reactivation of Precambrian orogenic zones (Gethard et al., 1990). The reactivation that ‘caused the thermal event at about 500 Ma appears to have been associated either with a phase change at the base of the crust (Fowler and Nisbet, 1985) or with a mantle intrusion into the crust (Ahern and Ditmars, 1985). Both these interpretations are based on analysis of tectonic subsidence. Foreland flexural subsidence, (e-g., see Chapters 9 and 11) similar to Carboniferous subsidence in the Ulinois and Michigan basins, is inferred to have developed from post-Jurassic time through the Late Eocene, presumably in response to convergent tectonics in the North American Cordillera The Williston basin is one of North America’s majot petroleum provinces (Gerhard et al, 1990). The prin pal source bed is the Upper Devonian to Lower Mi sissippian Bakken Formation, which formed during transgressive anoxic episode (Gerhard et al., 1990) Subsidiary source beds include the Ordovician Win: nipeg Group (also of anoxic, relatively deep-water conditions) and parts of the Pennsylvanian Tyler Fig. 19.7 Generalized sratigrophy, Willston basin, (Redraw after Get hard etal, 1990), 8 PERMAN Ta LMINNEKAHTA ‘SYSTEMS 2 ROCK UNITS < OPECHE QUATERNARY @ | PLEISTOCENE PENNSYLVANIAN 5 wSDEN ge WHITE RIVER, 2 meeeY@ GROUP S RATCLIFFE | e ussssrowy | 2 |_[_Rrewme $ &| FROBISHER |= TLSTON MEL CREEK 5 INTERVAL (i roxws — a NTERVAL i nnn IRDBEAR cretaccous | 2 53] |__anpeesn_o ® eae < DUPEROW @ NIOBRARA & DEVONIAN ~~ s ‘SOURIS cane S| men = [oavsowan © rouge FY 23 WINNIPEGOSIS @ ASHEN @ INVAN KARA ‘SILURIAN | f a é INTERLAKE @ wet § © [sorewat RIERDON. ORDOVICIAN ‘STONY MTN. £ p—__} | — : ransse | gS | sven g camnocoro. | § [reco cnour® £5 nm PERMIAN 28 | ‘unrestricted [__Precamarian DEADWOOD @ 468 Chapter 13 CAMBRIAN PROOWCIAN SIL | DEV | wS5 [PENN] PERN ‘SHUR | TIPPECANOE. [KASKASKIA] ~ABSAROKA_ ee aaa 3 550, 500,450.00 sso Wa 500. vee] mt 100 vf papain gs! 2 000] E1500 venga in 2 a0! 9 550___S00__450__490_ S50__ 290 Me s00} sf 10} 20% tis asin aso! Fig. 13.8 Backstipped and decompacced tectonic subsidence for the ‘Willston, Michigan, and tkinos basin, showing similar Hime of basin formation around 550 to 530 Ma, and initiation of thermal subsidence between $30 and 500 Ma. 5 tigraphlc data for these curves came from rks Well a Michigan basin (Hinze et al, 1978), and regional stratigraphic summaries for other two basins (see Bond and Kominz 1991, for detalls). (Redraw from Bond and Kominz, 1991, Fig. 28.) Meclure-5 Formation (mostly coastal deposits). Reservoirs are primarily Devonian and include the Winnipegosis Formation (major reef trend), and the Duperow and Dawson Bay formations (both stromatoporoid banks). The Pennsylvanian Tyler Formation contains productive reservoirs, which are mostly barrier-island and deltaic-distributary reservoirs. A recent play has focused on the Sherwood Formation (Mississippian), where oil has accumulated in stratigraphic traps at carbonate facies pinchouts into mudstones and evap- orites (Shirley, 1991). Petroleum migration was mostly vertical, along regional lineaments and vertical faults (Gerhard et al., 1990). Most of the traps are structural, and even the reefs and stromatoporoid reservoirs are fault-bounded. Michigan Basin ‘The Michigan basin (Fig. 13.1) also is saucer-shaped in cross section and oval in plan view. It is filled with 4500 m of sedimentary strata, which consist domi- nantly of carbonate rocks and evaporites, with subor- dinate shale and sandstone (Fisher et al., 1988; Cata- cosinos etal, 1990; Howell and van der Pluijm, 1990). These strata represent coastal to shallow-shelf depositional settings. Silurian evaporites and fringing pinnacle reefs represent one of the most unique set of facies in this basin and serve as major economic resources (see Fisher et al., 1988 and Catacosinos et al., 1900 for detailed reviews). Sedimentary strata renge in age from Cambrian through Carboniferous, and are capped with Jurassic red shale (Fig. 13.9). All six cratonic sequences of Sloss (1963, 1988a) occur in this basin. The Michigan basin is underlain by an arm of the Midcontinent rift system, whose age is 1.1 Ga (Van Schmus and Hinze, 1985). The relationship between ‘mechanical subsidence associated with this rifting and the onset of thermal subsidence from around 520 to 500 Ma (e.g., Sleep and Sloss, 1978; Nunn et al, 1984; Nunn, 1986; Ahern and Dikeou, 1989) remains a problem because of the 600 my interval between the two tectonic events. This problem was resolved recently by Howell (1989) and Howell and van der Pluijm (1990), who proposed that intracratonic-basin formation and subsidence, especially in the Michigan basin, were associated with reactivation of earlier rift systems caused by decoupling of the upper crust and upper mantle because of weakening of the lower crust. Howell's (1989) and Howell and van der Pluijm’s (1990) model is supported by discovery of approximately 670 Ma thermal metamorphism (McCallister et al., 1978) of Keweenawan (1.1 Ga) basalts recovered by drilling from the Michigan basin depocenter, and by secondary magnetization of this basement following deposition of overlying sedi- ments, which retained their primary magnetism (Van der Voo and Watts, 1978). It is suggested, therefore, that earliest Paleozoic tectonic reactivation of subsi dence in the Michigan basin was controlled in part by late Proterozoic structural trends of the Midcontinent rift system (cf., DeBrito Neves et al., 1984; Leighton and Kolata, 1990). Thermal subsidence followed this system MEDON|sEquenck [sere ER FORMATION SI a ale Tertiany | '® eias + —_ i coe JURASSIC. Zn TTT 2008 oD 286 Grand Fiver 3 | Absaroka | & ‘saghaw & 3204 5 e 2 ju — | é ae 360 SX gene 5 _—— Y z Kaskaskia [EERO hen Zz ‘374 ‘Traverse Ls. "Bell Sh. gs Me. 8 Legs L ade ind poe eee t 408 — oa ae 2 | cwmeme See z 1 et ; 2 {J [emraarane————| z as soe_fttcan = = Y ‘Tica éu = > ‘Black Fiver Q iM See ; 6 [7% For ———— 505 — Trempeaiea a = z ’ B pre ee id 2 4 a El sr mM PRECAMBRIAN | Pre-K. Semon ig 13.9 suratigraph subdivision of tichigan basin. (Reerawn from Ctacosinos eta. 1990,) reactivation rifting; it began coevally with thermal subsidence in the Illinois and Williston basins (Fig. 13.8). This observation is consistent with the sugges- tion of Sloss (1988a) and Howell and van der Pluijm (1990) that around 500 Ma, the Michigan basin was a northern extension of the Reelfoot aulacogen underlying the Wlinois basin; thus, thermal subs: dence of the Michigan basin was linked to thermal subsidence in the Illinois basin. The Michigan basin was characterized by renewed subsidence during the middle Devonian (Fig. 13.8), Intracratonic Basins 469 and also, during the late Mississippian and Pennsylvan- ian (Middle and Late Carboniferous) due to far-field flexural subsidence in response to the Alleghenian- Hercynian orogeny (Quinlan and Beaumont, 1984; Beaumont et al., 1987; Leighton and Kolata, 1990) The flexural stress associated with that orogeny caused not only shallow and broad subsidence of the Appalachian basin, but also transmission of these stresses into the intracratonic interior, reactivating subsidence of intracratonic basins (cf., Ziegler, 1987). Howell and van der Pluijm (1990) also provided evi- dence suggesting that subsidence rates increased dur- ing Late Ordovician and Late Silurian to Early Devo- nian time because basin subsidence was renewed by reactivation along inherited crustal trends due to similar far-field influences during early and middle Paleozoic orogenic movements in the Appalachians (Taconic and Acadian orogenies, respectively) Although petroleum is produced from the Michi- gan basin, source and reservoir units appear to be less favorable than in either the Illinois or the Williston basins. Principal source beds are the Pre- cambrian lacustrine Nonesuch Shale, several Ordovician through Mississippian shales of anoxic origin (Nunn et al., 1984; Catacosinos et al., 1990), and possibly Silurian horizons. Reservoirs are con- fined to marine carbonate-platform beds of the Tra- verse Formation (Devonian) and Niagaran (Silu- rian) pinnacle reefs. Migration of petroleum appears to have been lateral (Catacosinos et al., 1990), although modeling studies by Nunn et al. (1984) suggest a vertical component of oil migra- tion, Principal traps and seals appear to be carbonate mudstone enclosing pinnacle reefs, and unconfon ties and fault-bounded features associated with the Traverse Formation. Other Intracratonic Basins Amadeus Basin The Amadeus basin of Australia is a late Proterozoic to early Paleozoic intracratonic basin (Fig. 13.10) that shows an evolutionary history comparable to other intracratonic basins. The basin is an elongate ellipse in plan, with dimensions of 800 x 300 km; strata 470 Chapter 13 5o reach a maximum thickness of 14 km. The discussion that follows is based on recent work by Lindsay (1987), Lindsay et al. (1987), Korsch and Lindsay (1989), and Shaw et al. (1991). ‘The stratigraphy of the basin (Fig. 13.11) is domi- nated by clastic rocks, mostly of coastal-plain, deltaic and continental-shelf facies, and some of tidal origin, Locally, salt deposits are known, Basal Upper Pro- terozoic units are mostly sandstone with interbedded basalt; they appear to represent a rift sequence, Overlying widespread strata contain quartz arenite, carbonate rocks, and evaporites. This sequence is continuous in the basin depocenter and overiaps basement; the boundary represents an equivalent to a breakup unconformity of a passive margin (see Chapter 4), This sequence is overlain unconformably by fluvial, glacial, and shelf facies. The basal marine onlapping beds of this third unit are arkosic and phosphatic; locally, a deep-water facies occurs at its base. Latest Proterozoic and Cambrian sedimentation (Stage 2 in Fig. 13.11) occurred in major subbasins of the Amadeus basin. Sedimentation began with depo- sition of deltaic, coastal-plain and shelfal sandstone, followed by evaporites. The entire succession is capped by a widespread Ordovician to Silurian unconformity. Sequence 2 in the Amadeus basin is coeval with the Sauk Sequence (cf., Sloss, 1963, 1988a). It is overlain by marine Devonian strata Fig. 13.10 Location map ‘Amadeus basin, central ‘Australia, Peterman Renges om southwest ide of basin were principal source terrane for clastic sediments during basi filling. “Trough” refers to basin depocenter. (Redrawn ater Lindsay and Korsch, 1989) throughout the basin; these are capped by a fluvial conglomerate. Onlap curves were derived from field stratigraphic relations, and a relative sea-level curve was con- structed (Fig. 13.11). Several regular cycles are evi- dent. Drops in relative sea level were observed and found to correlate to subsidence (Fig. 13.12) and tec- tonic evolution of the basin, Lindsay and Korsch (1989) discriminated between larger-order cycles con- trolled by regional tectonics, and smaller-order cycles controlled by both climate and local tectonics. The climatic-eustatic sequences may have been influenced by global tectonic patterns (Lindsay and Korsch, 1989) during evolution of the Amadeus basin. Subsidence analysis was completed in three loca~ tions, one froma deep well in the platform zone adjoining the basin, and the other two from within the basin (Fig. 13.12). The subsidence curves show an early history of crustal extension and thermal subsi- dence (Stages 1 and 2; Fig. 13.11) followed by regional thrusting (Stage 3). The subsidence history is similiar to that reported from the Illinois basin (Heid- lauf et al., 1986; Treworgy et al., 1989; Kolata and Nelson, 1990b), the Michigan basin and the Williston, basin (Howell and van der Pluijm, 1990; Bond and Kominz, 1991). Lambeck (1983a,b) and Shaw et al. (1991) argued, however, from model calculations and unconformity mapping, that the principal driving mechanisms for formation of the Amadeus basin Intracratonic Basins 471 aE ig 13.11 Stratigraphic syn Low thesis, onlap curves. interpreted relative sea level and crustal evolution, Amadeus basin, ‘nustalia. (Redrawn ater sues | Siete = = “SAY CREEK [Lowa ooveen Forman) a Lindsay and Korsch, 1989.) see Sraching supe Srechng HEAVTREE QUART. SRD NERDS ‘Arona Comet (BaseiaeNis ~~ || [I were regional intraplate stresses caused by compres- sional tectonic activity at continental margins (ct. also Leighton and Kolata, 1990), and not thermal subsidence. The presence of early extensional faults (Korsch and Lindsay, 1989; Lindsay and Korsch, 1989) argues against this interpretation. Devonian subsidence, associated with regional thrust faulting, may have been more widespread (also see, Quinlan and Beaumont, 1984; Shaw et al., 1991). ‘Age (Ma) 700600500400 os 10 15 20 25 30 35 40 te | soe 45 Stage 1 Stage2 Stage 3 Fig. 1.12 Tectonic subsidence from (A) platform setting. (B) Missionary iains trough, and (C) Carmichael subbasin, Amadeus basin, Australia, (be Fig. 13.10), (Redravn after Lindsay and Korsch, 1989.) Depth (km) Parana Basin The intracratonic Parana basin incorporates an area of 1,400,000 sq. km in southern Brazil, Paraguay, northeast Argentina, and Uruguay (Figs. 13.13, 13.14), The discussion that follows is derived from DeBrito Neves et al. (1984), Stanley et al. (1985), and Zalan et al. (1985, 1990). The basin is filled with sedimentary and volcani strata nearly 6000 m thick; 1700 m of the basin sec- tion consist of Jurassic and Cretaceous volcanic rocks, which cover two-thirds of the basin’s outcrop area. ‘Most of the basin sedimentary rocks are siliciclas- tic, and are arranged into five stratigraphic sequences bounded by unconformities (Fig. 13.15). The lower two sequences are equivalent to Sloss’s (1963, 1988a) Tippecanoe sequence, whereas the remainder are equivalent to the Absaroka and Zuni sequences (also see, Soares et al., 1978). This dominance of clastics is attributed to the basin’s paleoposition in polar lati- tudes (Zalan et al, 1990). The Paleozoic sequences represent transgressive-regressive cycles, whereas the Mesozoic Sequences are all nonmarine. The Mesozoic also contains a thick (1700 m) volcanic succession, which appears to be the thickest volcanic sequence reported from an intracratonic basin (Zalan et al., 1990). These lavas range in age from 147 to 119 Ma (Zalan et al., 1990). Each of the sequences can be correlated to major regional tectonic events 472 Chapter 13 Fig. 13.13 Index map of Brazil, shoving location of Parana basin, (Redrawn alter Zalan et al, 1990.) (Fig. 13.15). Eruption of the lavas was associated with rifting of South America from Africa, ‘The basin is criss-crossed by regional normal faults and lineaments, some of which are known to extend into basement (DeBrito Neves et al., 1984; Zalan et al., 1985, 1990). These are oriented primarily northeast- southwest, and northwest-southeast. All tend to fol- low older (Proterozoic) tectonic trends within the basin and probably were formed by propagation from and reactivation of basement faults. Reactivation of these basement-controlled features may have con- trolled formation and evolution of the Parma basin (deBrito Neves et al., 1984). Reactivation by exten- sion facilitated recurring basin subsidence (Fig. 13.16). These subsidence events coincided with accu- mulation of sedimentary sequences during Silurian- Devonian, Permo-Carboniferous, and Late Jurassic- Early Cretaceous times. Fig. 13.14 Isopach map, incorporating both sedimentary and volcanic basin il, Parana basin, Braz (Redeawn after Zalan eta, 1990,) / Silurian-Devonian subsidence was thermal and was associated with a westward facing passive margin (Zalan et al., 1990). Permo-Carboniferous reactiva- tion of subsidence was associated with rifting (Zalan et al., 1990). Jurassic-Cretaceous subsidence appears to have been caused entirely by volcanic loading. This subsidence accounts for the low slope at this time in the tectonic-subsidence curve (Fig. 13.16) and was critical for hydrocarbon formation, inasmuch as Per- mian source beds underwent thermal maturation during this volcanic loading event (Zalan et al, 1990). The presence of these thick volcanics contrasts with other intracratonic basins. System PS°LiTHOSTRATIGRAPHY™™) Shes | eveuney CRETACEOUS Gumsaa [ele Pa ena = IE setesrmo oo harsFu xara Fig 13.15 Chronosiraigraphic diagram of Parana basin compared to regional tectonic events and to interpreted changes in se level (Redeawn alter Zalan etal, 1990.) No commercial petroleum exists in this basin, although drilling has found gas in the basin center and oil in the southern part of the basin. Tar sands and seeps are known in the eastern part of the basin. ‘Two shale beds, the Permian Irati Formation and the Devonian Ponta Grossa Formation constitute the best source beds. Reservoirs are known from multiple pay zones, including Permian fluvial sand, and Triassic fluvial and colian sands. A Silurian prospect exists in marine sandstones (Zalan et al., 1990). Vertical migration (both up and down) of hydrocarbons has occurred along faults and lineaments. Trapping geometries are poorly known, but include fault boundaries, and basaltic sills and dikes as boundaries (Zalan ct al., 1990). The role of late Mesozoic igneous activity on limiting the petroleum potential of this basin remains unknown Mechanisms of Formation of Intracratonic Basins The above summary of the geology and geodynamics of intracratonic basins demonstrates that the Illinois, Williston, Michigan, Amadeus, and Parana basins formed by a similar sequence of processes. These processes, in order of occurrence were: 1) lithospheric stretching, 2) mechanical, fault-controlled subsidence, 3) thermal subsidence and contraction, and 4) merging of slower thermal subsidence with reactivated subsi- Intracratonic Basins 473, Z 2 oneraceous , TecToMi sussioence a ? BASEMENT SUBSIDENCE Fig 13.16 Tectonic and basement subsidence, north-central Parana basin, Bani, (Redeawn attr Zalan etal, 1990.) dence due to isostatically uncompensated excess mass (eg. Sleep, 1976; Sleep and Snell, 1976; Sleep and Sloss, 1978; Sleep et al., 1980; DeRito et al., 1983; Ahem and Mrkvicka, 1984; Nunn et al., 1984; Ahem and Ditmars, 1985; Heidlauf et al., 1986; Klein and Hsui, 1987; Ahern and Dikeou, 1989). These events were followed by flexural foreland subsidence in some of the basins (e.g, Quinlan and Beaumont, 1984; Beaumont et al., 1987; Quinlan, 1987; Howell and van. der Pluijm, 1990; Leighton and Kolata, 1990). More- over, timing of initiation of thermal subsidence in the Illinois, Michigan, and Williston basins is remarkably close (Fig. 13.8; 530 to 500 Ma). Mesozoic and Ceno- zoic intracratonic basins of western Europe and India also share a common timing of initiation of thermal subsidence (see summary in Klein and Hsui, 1987) To explain these basin-forming processes, concur- rent initiation of thermal subsidence of North Ameri can Paleozoic intracratonic basins, and concurrent ini tiation of thermal subsidence of Mesozoic/Ceno: intracratonic basins of Eurasia, several common char- acteristics of the basins must be considered. First and foremost are the cause and localization of rifting and reactivation of the basins (Ahern and Dikeou, 1989; Howell, 1989). Secondly, major Paleozoic intracra- tonic basins of North America, Europe, Africa, and South America appear to have been initiated during breakup of major supercontinents, either during the Latest Proterozoic/Early Cambrian (Bond et al., 1984; 474 Chapter 13 Klein and Hsui, 1987) or during the Late Triassic/Early Jurassic (Ziegler, 1977, 1978, 1982, 1987, 1988). Thirdly, sedimentary sequences of intracratonic basins in North America, Africa, Asia, Europe, and South America show similar ages for interregional unconfor- mities that separate intracratonic sedimentary sequences, and similar trends in thickness and volume (Sloss, 1963, 1972, 1976, 1979: Sloss and Scherer, 1975; Soares et al., 1978; Peters, 1979; DeBrito Neves et al., 1984; Zalan et al., 1990). These observa- tions suggest a common global explanation, and the near synchroneity of the ages of formation of both North American (earliest Paleozoic) and Eurasian (early Mesozoic) intracratonic basins implies a large- scale process that caused both rifting and thermal subsidence Since 2 Ga, four major episodes of supercontinent accretion have been followed by supercontinent breakup (Condie, 1982b; Bond et al., 1984; Worsley et al., 1984, 1986). During periods when continents were dispersed, heat loss from the earth’s interior was dissipated through mid-ocean ridges (Fischer, 1984; Worsley et al., 1984), whereas during times when continents were merged into a supercontinent, ridge volcanism and associated heat loss diminished consid- erably, partially beacuse spreading rates may have been reduced as well (Fischer, 1984; Worsley et al., 1984). Consequently, supercontinents may have acted as heat lenses (Anderson, 1982; Worsley et al., 1984, 1986). As a consequence of developing super- continents, the rate of partial melting of continental lithosphere would have increased in response to focused heat flow, particularly in the lower crust and upper mantle. Breakup of a supercontinent into smaller plates would have changed the heat-flow regime to one more like the present. This cyclic repe- tition would also have been influenced by the chang- ing convective history of the mantle, as plates shifted from separate geoid highs to merge over geoid lows (Gurnis, 1988, 1990a.b). Research concerning the origin and nature of Pro- terozoic granite of the North American craton has documented two contrasting modes of granitic crustal formation (e.g., Anderson and Cullers, 1978; Emslie, 1978; Anderson, 1980, 1983; Anderson et al., 1980). During plate convergence and associated mountain building, crustal formation included emplacement of orogenic granite. Intrusion of anorogenic granite postdated these orogenic intrusive events by up to several hundred-million years (Anderson and Cullers, 1978; Emslie, 1978; Anderson 1980, 1983; Anderson et al., 1980). Extensive chemical petrologic analysis of anorogenic granite from several continents demon- strates that they formed by partial melting of older orogenic granitic crust (Anderson and Cullers, 1978; Emslie, 1978; Anderson et al., 1980; Nurmi and Haa- pala, 1986). Continent-wide intrusion of anorogenic granite (e.g,, 1.5 to 1.4 Ga) may have been associated with supercontinent development and_ regional extension (Silver et al, 1977; Anderson and Cullers, 1978; Emslie, 1978; Anderson, 1983; Bickford et al. 1986; Nurmi and Haapala, 1986; Haapala and Ramo, 1987). The simultaneous partial melting of such scat- tered coeval anorogenic granite requires a common mechanism (Fig. 13.17A), postulated to be the buildup of mantle-derived heat below a superconti- nent (Anderson, 1982; Worsley et al., 1984; Ander- son, 1987). Alternatively, coeval development of a mafic igneous body in the upper mantle could also produce partial melting in the lower crust (Emslie, 1978; Anderson, 1983; Nurmi and Haapala, 1986) One consequence of partial melting is that the remainder of the crust must have been thinned (cf., Howell, 1989; Howell and van der Pluijm, 1990). Intrusion of anorogenic granite on a supercontinent scale would have created major lateral physical continuities in properties of the continental crust. When stress fields changed from compressional to extensional, or vice versa, or in direction and magni- tude, major ruptures and associated surficial defor- mation would be expected on supercontinents, espe- cially at crustal inhomogeneities. Such changes in stress also would have favored the reactivation of basin subsidence (DeRito et al., 1983; Lambeck, 1983a,b; DeBrito Neves et al., 1984; Quinlan, 1987; Howell, 1989; Howell and van der Pluijm, 1990; Leighton and Kolata, 1990; Zalan et al., 1990; Shaw et al., 1991). Extension may have focused on zones within the lithosphere where anorogenic granite intruded (Anderson and Cullers, 1978; Emslie, 1978; Nurmi and Haapala, 1986; Haapala and Ramo, 1987), thus causing both thermal doming and weakening of vanse nem ae cana cast een ate er tathoteo) (icing cceani erst Fig 13.17 Evolution and simultaneous formation of intracratone basins 2s eras of supercontinent i partially melted and anorogenic granite intrudes crust during stage when supercontinent acted as heat lens (A), followed by rifting and assocated fault-conttlled, mechanical subsidence (B). These rift zones underwent thermal subsidence, forming saucer shaped inirctatone basins ae anorogenic bodies cooled (C) Consequently intracratonic basins occur above anorogenic-granite bodies Intracratonic basins and thelr global regional unconformite, parallel tecionle-subsidence histories, and parallel changing sediment volumes, Schematie representation superposed, in parton continental-heat-lens and breakup model of Worsley ta. (1984, Fig. 8). ocess accounts for nearly simultaneous init crust (Fig. 13.178). Crust immediately above partially melted zones would have been thinned, thus favor- ing localized and regional stretching. This configura- tion would have accounted for stretching within the crust above areas of anorogenic granite intrusion and partial melting during supercontinent breakup. Because of increased viscosity of these granite mag- mas, the partially melted bodies should have split, separated, and accelerated uplift on the flanks of the rifted basin (Neugebauer and Reuther, 1987) during continued extension. With continued stretching, the magma bodies themselves may have fractured; such fractures would have become conduits for subsequent intrusion of upwelled igneous materials derived from the upper ‘mantle (Fig. 13.17C). It would be expected that asso- Intracratonic Basins 475 ciated rift valleys that formed in response to stretching associated with supercontinent breakup would be floored by anorogenic granite intruded by tholeiitic basalt derived from the upper mantle. Such an associ- ation of anorogenic granite and tholeiitic basalt of approximately coeval age (577 to 500 Ma) has been. documented in the southern Oklahoma aulacogen (e.g, Gilbert, 1983; Lanbert et al., 1988). Several younger Phanerozoic examples of rift-related, paired, anorogenic granite and tholeiitic basalt are known, induding the Jurassic-Cretaceous White Mountain. magma series of New Hampshire, Jurassic granite of, Nigeria, and Miocene-Pliocene volcanics of the Aden Sea region, the latter being involved with relatively recent breakup and rifting of Africa and the Arabian Peninsula (Anderson and Cullers, 1978, p. 311; Emslie, 1978). In the Precambrian of south Africa, Cooper (1990) documented several cyclic recurrences of paired anorogenic granite and tholeiitic basalt asso- ciated with eat basin evolution. A similar succession of intracratonic basin development after intrusion of anorogenic gran- ite has been reported from the Precambrian of Finland by Gaal and Gorbatschev (1987). Three North American intracratonic basins, the Ili- nois, Williston, and Michigan basins, formed following rifting (Howell, 1989; Gerhard et al., 1990) during a period of Late Precambrian breakup of a superconti nent, which was followed by thermal basin subsidence (Sleep and Sloss, 1978; Braille et al., 1982; Keller et al. 1983; Ahern and Mrkvicka, 1984; Fowler and Nisbet, 1984; Nunn et al., 1984; Ahem and Ditmars, 1985; Heidlauf et al., 1986; Ahern and Dikeou, 1989). Thus, following Klein and Hsui (1987), the location of intracratonic basins was controlled by rifting above partial melting of lower crust and intrusion of anoro- genic granite during Late Precambrian breakup of a supercontinent (Fig. 13.17C; cf., also Cooper, 1990), Intracratonic basins display both a common time of formation, and synchroneity of interregional uncon- formities and sediment volume changes. Crustal zones which were domed upward above anorogenic granite are susceptible to rupture by rifting, which is why intracratonic basins occur in these zones. Cooling of anorogenic granite, basalt, and asthenosphere to form lithosphere caused subsequent thermal subsidence. st stages of Precambrian intracratonic 476 Chapter 13 Bally (1989) and Sloss (1988b) objected to this model on several grounds. Bally (1989) rejected the concept of supercontinent accretion and breakup cycles, and questioned the existence of anorogenic granite bodies below intracratonic basins (also see Sloss, 1988). Moreover, Bally (1989) argued that subsidence histories of intracratonic basins are inde- pendent, and cited Quinian’s (1987) discussion of subsidence history of the Hudson Bay basin, which clearly subsided thermally at least 100 my after the other three North American basins. Both Sloss (1988b) and Bally (1989) also called attention to a 600my delay between Keweenawan rifting below the Michigan basin and early Paleozoic thermal subsi- dence. They failed to consider arguments concerning reactivation of such rift systems (Quinlan, 1987), thermal metamorphism of the Keweenawan basalts in the depocenter (McCallister et al., 1978), and sec- ondary magnetization of the basalts in the depocenter (Van der Voo and Watts, 1978), and were unaware of subsequent work establishing reactivation in that basin by Howell (1989). Moreover, Bond and Kominz (1991) demonstrated that subsidence histories of the Michigan, Williston, and Illinois basins (Fig. 13.12) are nearly synchronous, with minor devi differential subsidence rates in each basin (also see Sloss, 1988a). It should be remembered that supercontinent breakup is not an instantaneous process. The ages of rifting associated with late Proterozoic and early Pale- ‘ozoic aulacogens and interior rifts differ as one moves away from continental margins where initial breakup occurred. Klein and Willard (1989, Fig. 2) compared subsidence histories of the Ilinois and central Appalachian basins, and showed, as did Bond et al., (1984), that the late Proterozoic-early Paleozoic pas- sive margin of eastern North America rifted around 620 Ma (Willard and Klein, 1990), whereas rifting below the Illinois basin started around 530 Ma. It is my interpretation that as a supercontinent breaks up, the areas of greatest heat flow migrate away from supercontinents both to mid-ocean ridges (also see Sloss and Speed, 1974) and toward interiors of remaining continents. It is, for instance, well known that the breakup of Pangea occurred in successive stages over a period in excess of 120 my (e.g., Watts ons due to et al., 1982; Ziegler, 1982, 1988, among many oth- es), and some of these areas of later breakup con- tained anorogenic granite and tholeiitic basalt (Anderson and Cullers, 1978). Therefore, later ther- mal subsidence of the most interior intracratonic basins would be expected. This change in timing may account also for evidence of localized differential sub- sidence reported by Sloss (1988a). ‘The proposed basin-forming mechanism for the origin of intracratonic basins poses two consequences, First, below older Precambrian intracratonic basins, where the root zones are exposed, anorogenic grat should occur. Gaal and Gorbatschev (1987) demon- strated that in the Jomian intracratonic basin of southwestern Finland, the basal Jotnian Sandstone overlies anorogenic granite (Laitakari, 1983; Nurmi and Haapala, 1986). In an unnamed intracratonic basin in northeastern central Finland along the Gulf of Bothnia, the basal arkosic Muhos Formation over- lies anorogenic granite (Veltheim, 1969; Paakola, 1983). Recent seismic surveys have demonstrated that the crust is thin below the anorogenic granite (Gaal and Tuomi, in press). The basement-sediment contact in the Williston basin (Kent, 1987) shows anorogenic granite below part of the basal Phanerozoic succes- sion, In southern Africa, Cooper (1990) reported paired anorogenic granite and tholeiitic basalt below successive Precambrian intracratonic basins, ‘A second consequence of this hypothesis is that formation of intracratonic basins follows major coll sional orogenic events. Both examples from Finland are postorogenic basins (Nurmi and Haapala, 1986). Intracratonic basins were observed to have developed India starting during latest Permian time following the breakup of Gondwana (Kailasham, 1976). The Paris basin and the Viking graben of the North Sea show subsidence patterns that are similar to those of North American Paleozoic intracratonic basins; ther- mal subsidence of the former started during latest Permian and Triassic time (Sclater and Christie, 1980; Brunet and LePichon, 1982; Ziegler, 1982, 1988; Per- rodon and Zabek, 1990). These findings support the interpretation advanced in this chapter. Although the stages of subsidence of intracratonic basins are sequential, overlapping of subsidence mechanisms has occurred (Fig. 13.18). In the Ilinois basin, Davis (1987), Clendenin (1989), and Watso and Klein (1989) recognized that Cambrian and Ordovi- cian normal faulting occurred when thermal subsi- dence was dominant. Following a 60my period of rapid thermal subsidence, a period of slower thermal subsidence overlapped a time of increasing subsidence due to isostatically uncompensated excess mass (cf., DeRito, et al., 1983; Treworgy et al., 1989; Kolata and Nelson, 1990). This slow thermal subsidence was associated with Devonian faulting (Davis, 1987). Late phases of subsidence are documented in the three US Paleozoic intracratonic basins (Illinois, Williston, and Michigan) in response to foreland tectonics and lateral, transmission of intraplate stress during times of oroge- nesis along continental margins of North America (Quinlan and Beaumont, 1984; Beaumont et al., 1987; also see Craddock and van der Pluijm, 198 Klein and Willard, 1989; Willard and Klein, 1990), During times of orogenesis, partial melting may occur, locally, giving rise to igneous intrusions, such as the late Paleozoic alnoites reported from the Illinois basin (Zartman et al., 1967). Such localized partial melting may have caused shifting of depocenters, as reported from the Illinois, Williston, and Michigan basins by Heidlauf et al. (1986) and Sloss (1987). Partial melting is likely to be associated with reactivation of older rifts me, \ wreacraTonc esi \ evoUUTION orm a rastcsercg™ ‘Subsidence _ "ara bene | = Swsence due \sBiateay Unconpenssieg Excess Na Fig. 13.18 Generalized backstripped and decompacted tectonic subsidence, showing sequential stages of subsidence processes for Snracratonc basins Intracratonic Basins 477 and tectonic belts beneath intracratonic basins, such as those reported by DeBrito Neves et al. (1984), Quinlan (1987), Howell (1989), Howell and van der Pluijm (1990), and Leighton and Kolata (1990). Finally, it must be emphasized that this model of basin formation of intracratonic basins requires fur- ther testing. Since the original proposal of the model (Klein and Hsui, 1987), additional data add plausabil- ity; recurring cycles of intrusions of anorogenie gran- ite preceding intracratonic-basin evolution in south- em Africa (Cooper, 1990), and the importance and verification of reactivation along older tectonic trends have been demonstrated (DeBrito Neves et al., 1984; Quinlan, 1987; Howell, 1989; Howell and van der Pluijm, 1990; Leighton and Kolata, 1990). The major difficulty with the hypothesis is that Phanerozoic intracratonic basins have not been drilled to base- ment and inadequate sampling leaves the problem unresolved. I predict that if ultradeep scientific drilling is undertaken in the Illinois basin, for instance, then anorogenic granite ranging in age from 600 to 510 Ma should be recovered, below the lay- ered crustal structure reported by Pratt et al. (1989). ‘The proposed model for intracratonic basins accounts for all the common properties and features of intracratonic basins. Acknowledgments ‘My research in cratonic basins has been supported by various sources, including the National Science Foun- dation (EAR-90-01448), the University of Illinois Research Board, and the Department of Geology at the University of Minois, Discussions with JJ. Eidel, LO. Gerhard, RD. Howell, A:T. Hsui, D.R. Kolata, ‘MW, Leighton, W.J. Nelson and BJ. van der Pluijm have proved helpful, In particular, 1 thank M.W. Leighton for making arrangements for me to obtain preprints of chapters in AAPG Memoir 51. Michael R. Leeder, Jeffrey A. Nunn, and the editors, Cathy Busby and Raymond V. Ingersoll, are thanked for their constructive comments on earlier manuscript versions of this chapter. 478 Chapter 13 Further Reading ‘ond GC, Kominz MA, 1991, Disentangling middle Paleozoic se level and tectonic events in cratonic margins and eratonic basins of North America: Journal of Geophysical Research, v.94, p. 6619-6638. Hamdani ¥, Mareschal JC, Arkani-Hamed J, 1991, Phase changes and ‘thermal subsidence in iniracontinental sedimentary basins: Geophysical Journal Ieratonal,w. 106, p. 637-665, Heidlauf DT, Hsu AT, Klein Gdev, 1986, Tectonic subsidence analysis of the llinois basin: Journal of Geog, v.94, p- 779-794, Klein, GdeV, 1991, Origin and evolution of North American cratonle basins South African Journal of Geology, 94 p. 3-18. Leighton MW, Kolata DR, Ole DE, Eide 3 (eds), 1990. Interior eatonie. basins: American Assocation of Petroleum Geologists Memoir 51, 819p. Sleep NH, Nunn JA, Chou L, 1980, Platform basins: Annual Review of arth and Planetary Sens, v. 8. P.17-34 Sloss LL, 19888, Tectonic evolution of the craton in Phanerozoic time, in Sloss L. (ed), Sedimentary cover - North American craton: U.S. (The Geo ogy of North America, v.D-2]: Geological Sctety of America. Boulder, P.25-51.

You might also like