In this chapter 25.1 Global topography and plate tectonics 973 Global geological maps: Introduction 973 Global geological cartography: Selected milestones 974 Stratigraphic nomenclature and the geological time scale 978 References 979 25.2 Global topography and bathymetry: The face of old earth reworked and modified by present processes 982 Introduction 982 Global relief models: Onshore and submarine morphology and plate tectonic regimes 983 References 987 25.3 Neotectonics; earthquakes and conventional (i.e., rigid) versus diffuse plate boundaries 988 Neotectonics: Introduction 988 Global earthquake distribution 989 Well-defined versus diffuse plate boundaries 990 Neotectonic plate motions: Their relation to a fixed Eurasia and to Cenozoic/ Mesozoic fold belts 992 References 997 25.4 Global stress maps and paleostress studies 998 References 1001 25.5 The continental lithosphere and continental crust 1003 Introduction 1003 The continental lithosphere 1004 The continental crust 1005 Crustal layers, rheological models, and conclusions 1007 References 1008 Plates for global topography, neotectonics, the continental lithosphere and crust: Segments 25.1–25.3 and 25.5 1010 25.6 Tectonic maps of the world 1027 Introduction to tectonic maps 1027 Recent advances in alpine tectonics: An example of the scope of larger scale tectonic maps 1028 Simplified tectonic maps of the world 1029 About Phanerozoic plate tectonic reconstructions 1031 References 1032 Polar tectonic maps: Introduction 1036 Arctic tectonic map 1036 Antarctic tectonic map 1037 References 1038 25.7 Cenozoic/Mesozoic and Paleozoic orogenic systems and their fold and thrust belts (FTBs) 1042 Orogeny versus epeirogeny 1042 Subduction, sutures, and orogens 1043 Active margin fold and thrust belts (AMFTBs) 1050 Foreland fold and thrust belts (FFTBs) 1051 Normal faulting in foreland fold and thrust belts (FFTBs) 1054 References 1055 25.8 Age of Continental basement 1059 Introduction to basements, that is, the “residual” peneplaned former fold belts 1059 Merging the global tectonic map with a Precambrian basement map 1061 References 1063 25.9 Hot spots, linear island chains, large igneous provinces (LIPs), and radiating dike swarms; active volcanoes 1066 Introduction 1066 Large igneous provinces (LIPs) 1068 Giant radiating dike swarms 1071 Is there a “canonical progression of tectonic themes” preceding and/or following the emergence of a plume? 1071 The distribution of active volcanoes 1074 References 1074 25.10 Tectonic settings of mafic/ultramafic oceanic and intra-oceanic arc system crust, LIPs, rifted and volcanic passivemargins, tectonic setting and discussion of equivalent allochthonous “ophiolitic” fragments in orogens 1079 Introduction 1079 Subducted oceanic plateaus 1080 Allochthonous accreted oceanic plateaus and intra-oceanic island arc terranes 1081 Allochthonous fragments, oceanic and intra-oceanic arc systems, and lower crust and uppermost mantle of hyper-extended passive margins 1081 Allochthonous, exhumed continental crust-mantle transitions and the Ivrea-Verbano zone 1085 Conclusion 1086 References 1088 Plates for tectonics, orogenic systems, hot spots, lips, volcanoes: Segments 25.6–25.10 1092 25.11 Sedimentary basins and rifts (including Rifts) 1108 Introduction 1109 References 1111 Rift systems on relatively stable/rigid lithosphere 1112 References 1114 Passive margins on relatively stable/rigid lithosphere 1114 References 1119 Cratonic basins on relatively stable/rigid lithosphere 1121 References 1125 Basins on the periphery of orogens: Deep sea trenches and foreland basins 1127 References 1131 Basins located within orogens (Episutural basins of Bally and Snelson, 1980, or epi-eugeosynclines and successor basins of earlier authors) 1133 References 1139 Oceanic basins formed by spreading ridges (incl. oceanic back-arc basins) 1141 General reference 1141 Sedimentary basins and rifts: Segment 25.11 1142 importance and high quality of the work of the Commission forthe Geological Map of the World (CGMW) (www.cgmw.net or
[email protected]). For an understanding of the regional context, their maps are an outstanding, regularly updated source of information. The Geological Atlas of the World (Choubert and Faure-Muret, 1976/1983), earlier editions of the Geological Map of the World (Bouysse, 2000) nowmodified and re- issued in a third edition (Bouysse, 2009, 2010), and the Exxon TectonicMap (Exxon Production Research, 1985) were the principal sources from which simplified ele- ments of this set of world maps were derived. Many additional sources were used to complete themap series and are acknowledged in the text and/or in the captions. For easy comparison, our global maps are all on the same, commonly used Mercator or else Polar projections. The greatly simplified maps are scaled to fit a standard double-page size and drawn in Adobe Acrobat> Auto CAD files. Begin- ning with simplified coastlines, the cartographic accuracy of boundaries, trends, 2255 Tectonic and Basin maps of the world A.W. Bally,* D.G. Roberts,{ D. Sawyer,* A. Sinkewich{ *Department of Earth Science, Rice University, Houston, Texas, USA {Department of Earth Sciences, Royal Holloway, University of London, Egham, Surrey, United Kingdom {Anton Sinkewich Studio, Houston, Texas, USA 2255..11 Global topography and plate tectonics Global geological maps: Introduction Geological maps and regional studies integrate a wide variety of geological, geophysical, and geochemical information. Today, this calls for astute use of computerized information systems. In the final analysis, most geological maps remain interpretations of observations that involve numerous assumptions. For this chapter, a series of global maps have been compiled to illustrate a limited choice of regionally relevant themes and show their inter-relationships. We note here the Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps DOI: 10.1016/B978-0-444-56357-6.00024-X Copyright © by Elsevier B.V. All rights of reproduction in any form reserved. 973 and locations of the maps is limited. Thus, these maps should be viewed as 974 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps a collection of “sketch maps.” Ideally, a modified and GIS-compatible version of our maps that still preserves our simple graphics would be desirable. For more background information, readers are encouraged “to dig deeper” andwith greater care by consultingmore detailed published maps. Today’s ever-increasing “information overload” requires thematic organization to preserve the extraordinary legacy of worldwide geological information compiled over two centuries in the formofmaps by individual earth scientists, as well as pub- lic and private organizations. It is hoped that our maps may also serve as a graphic background to aid the computerized retrieval of some of that legacy by themes. Modern data management systems do require precise definitions. However, tra- ditional earth sciences are burdened with a jargon that includes a bewildering plethora of terms. For an understanding of the pre-plate tectonics terminology, Dennis (1967) is particularly useful, while Neuendorf et al. (2008) provide a detailed glossary of modern terminology. Earth science terminology has evolved with time. Yet, today, with ever-increasing specialization, the meaning of many terms often differs substantially among specialized earth scientists. Simplification is therefore important, not only to communicate among colleagues with different backgrounds but also when explaining the earth sciences to others. The rapid suc- cess of modern GIS systems indicates that first-order earth science communica- tions of the future will best be done with the aid of maps, cross-sections, and 3D displays that, in essence, will speak for themselves. To sum up, the simplified thematic global maps in this chapter are intended to provide a first-order background to the tectonic setting of sedimentary basins and foreland/folded belts. Like other global geological/tectonic maps, this set of thematic maps, clearly, does not provide proof of any theoretical concept or model. However, in the context of any given region, the maps may help not only to inform but also to determine whether specific theoretical models may apply or else be seriously “off-target.” It is hoped that these global maps may furthermore attempt to thematically organize a very large legacy of geological maps. Eventu- ally, all our maps will be superseded by cartographically more accurate, yet still mutually compatible, GIS-compatible displays of selected geological themes. Global geological cartography: Selected milestones Over more than 200 years, geological and tectonic maps of the world have recorded the progress of our regional geological understanding. Geological mapping was first promoted by Guettard (1746) and Guettard and Monnet (1780) who published a geological map of France. Guettard (1752) also published the first geological map of North America and compared it with one of Switzerland. His map of North America extends from Hudson Bay to the Gulf of Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Mexico. The map was compiled on the basis of the reports and corre- spondence of earlier explorers. Guettard differentiated only three units, that is, “sandy, marly and schistose bands.” He showed these lithologic units to occur in both North America and Switzerland. J.E Guettard may well be considered to be the “father of comparative regional geology.” William Smith’s seminal map of England, Wales, and Southern Scotland (Smith, 1815; Winchester, 2001) set a high standard and the way forward for modern geolog- ical maps. Ami Boue´ (Johnston, 1856) compiled the earliest geological map of the world notably, on the now commonly used Mercator projection, which is also used for this map series. Marcou’s (1875) Geological Map of theWorld reviewedmerely one third of the world’s continental areas as known to geologists at that time. Some 20 years later, the Geological Atlas compiled by Berghaus (1892) covered about two-thirds of the continents and included geomorphologic, hypsometric, and other thematic figures that already foresaw many current research themes. For example, the Berghaus Atlas displayed the Circum–Pacific volcanic “Ring of Fire,” earthquake maps, reconstructions of past ice covers, and the advance of the tsunami generated by the eruption of Krakatoa in 1883. The atlas also included separate geological maps for each of the continents and some evenmore detailed maps. The time of publication of the Berghaus Atlas overlaps with the publication of Suess’ (1885–1909) “Das Antlitz der Erde” (i.e., The Face of the Earth— translated into English by Sollas, 1904, 1909; La Face de la Terre—translated into French by de Margerie, 1921). This updated French translation is the best illustrated version of this classic treatise. Moving fast forward, Choubert and Faure-Muret (1976/1983) compiled the mas- terful 1:10,00,00,000 Geological Atlas of the World. This was followed by the 1:25,000,000 Geological Map of the World (Bouysse, 2000, 2009, 2010). Over recent years, many geological maps of the continents and subcontinents have been published. Recent examples include Africa (Choubert and Faure-Muret, 1990), the Arctic (Harrison et al., 2008), Asia (China Geoscience Institute, 1975, 2004), the Middle East (Aghanabati, 1993), Europe (Asch, 2005), North America (Reed et al., 2005a,b), and South America (Schobenhaus, 2001/2005). While providing a good overview, these important small-scalemaps do not resolve the details that larger scale maps offer. However, many of the larger scale regional maps are often confined by the political boundaries of countries, and other political/administrative subdivisions, that disrupt the regional continuity of the maps. Most continent-wide geological maps tend to follow internationally codi- fied stratigraphic, igneous, and metamorphic nomenclature. On the other hand, the legends of the more interpretative tectonic maps substantially reflect compilers’ views. The second half of the past century also saw the publication of many tectonic maps, some covering single continents and others the world. Initially some of these emphasized structural trends and the age of deformation (e.g., Staub, 975 1928; Umbgrove, 1947). Often, early tectonic maps either also supported now obsolete tectonic concepts such as the geosynclinal theory or else reflected the then-wide acceptance of Stille’s (1924) notion of short-lived orogenies punctuat- ing otherwise times of tectonic quiescence, that is, a “Pulse of the Earth” (Umbgrove, 1947). Examples of tectonic maps that ushered the transition from the preplate tec- tonic era into the overall acceptance of plate tectonics are the maps of Eurasia compiled by Yanshin (1966) and Peive and Yanshin (1980). In the former USSR, a significant part of debate surrounding the acceptance of the paradigm involved various attempts to reconcile the “Wilson cycle” with Stille’s tectonic phases as concisely summarized by Puchkov (1994). King’s (1969) Tectonic map of North America is another example of an outstanding tectonic map that, however, to a significant degree, reflects a preplate tectonics vision of the North American continent. Even if these old concepts are now outdated, the substance of these maps still remains important as they also display a lot of valuable basic information. Table 25.1 provides an “informal conversion” of Table25.1 Approximate correlation of “old” geosynclines with current sedimentary basins Taphrogeosyncline/Aulacogen Rift Autogeosyncline Cratonic or Intracratonic basin Miogeosyncline Passive margin or Atlantic-type margin Paraliageosyncline Delta-dominated passive margin Leptogeosyncline Starved oceanic deepwater basin 976 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Exogeosyncline Foredeep or foreland basins The mountain-derived clastic sequence was often referred to as a “Clastic wedge.” Orogen-derived turbidites were included under the poorly defined term “Flysch.” Deltaic and prodeltaic clastics were called “Molasse” in which conglomerates were emphasized as they shed light on the unroofing history of a folded belt. Eu-geosynclines (mostly in Europe) Ultramafics, pillow lavas, radiolarites (i.e., the old “Steinmann Trinity”), and overlying sediments, that is, the oceanic crust Eu-geosynclines (in N. America) Island arc/active margin systems including volcanics and associated island arc sedimentary basins Epi-Eugeosynclines Forearc and backarc basins The suggested equivalencies of this table are only approximate and reflect common past usage, rather than the original definitions of various authors. For more elaborate discussion see Dennis (1967), Dietz and Holden (1974), Dott (1974), Dott and Shaver (1974), and Neuendorf et al. (2008). some of the outdated geosynclinal terms into more contemporary equivalents. Understanding the old geosynclinal terminology is a key to many unique and Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps well-documented stratigraphic studies of the preplate tectonic era and their paleontological documentation. It is important not to lose this part of our geological legacy. Cox (1973), Hallam (1973), McKenzie (1977), and Oreskes (2003) offer a sense of the early days of plate tectonics. Beginning in the late sixties, following the wide acceptance of plate tectonics, new geological and tectonic maps were published culminating with the Exxon Production Research’s (1985) 1:10 M TectonicMap of the World. A selection of examples of recently published maps that are firmly grounded in plate tectonic principles include: the Tectonic Map of North America (Muehlberger, 1992), the Tectonic Map of Europe (Khain and Leonov, 1996/ 1998), the Seismotectonic Map of the World (Haghipour, 2001, 2002, 2006), the Geological Map of North America (Reed et al., 2005a,b), and the Structural Map of Eastern Eurasia (Pubellier, 2008). The Geological Map of North America is at the same scale and projection as the tectonic map (Muehlberger, 1992), the gravity map (Committee for the Gravity Anomaly map of North America, 1987), the magnetic map (Committee for the Magnetic Anomaly map of North America, 1987), the geothermal map (Blackwell and Steele, 1992), the seismicity map (Engdahl, 1988), and the stress map (Zoback et al., 1987). The Stratigraphic Atlas of North America (Cook and Bally, 1977) is not at the same scale as the above listed maps but uses the same projection. All the maps of the Stratigraphic Atlas should be digi- tized and regularly updated in a GIS compatible format. The World Atlas of the Russian Academy of Geosciences (1998) is a unique and comprehensive information source that, among other maps, illustrates global and regional geological processes with numerous world maps as they pertain to the evolution of landscapes, the lithosphere, the atmosphere, and the hydro- sphere. The Geological Atlas of Africa (Schlu¨ter and Trauth, 2008) provides a simplified first order overview that is organized by themes and countries. Contrast this atlas with the Geological Atlas of China (Ma Lifang, 2003), a comprehensive compendium that is introduced by Geological Maps of the World, Asia, and China, followed by about 15 thematic maps of China and some 40 geological maps of regions and provinces. The above-listed three examples are all widely different in scope, yet they all are very creative and useful. Modern global geological and tectonic maps and many more detailed maps now cover the geology of all the oceans including the Arctic Ocean, and all continents, including Antarctica. These maps, as well as numerous thematic Geological and Geophysical Atlases that are not listed here, are also valuable resources that should be consulted when analyzing the geology of large regions. In particular, cartographic advances in the earth sciences have been spectacular 977 and they are constantly evolving with the development of modern computer- able to go back to the original, paleontologically controlled stage names, as a 978 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps guide to conversion to the updated time scale. based GIS systems. Stratigraphic nomenclature and the geological time scale Stratigraphic nomenclature has evolved over time. Its often confusing complexity baffles earth scientists who had little exposure to the history of stratigraphy. Typi- cally, only specialists are familiar with the esoteric names of over 80 Phanerozoic Stages, or else over 30 Phanerozoic Series/Epochs and 10 Proterozoic Systems/ Periods. In addition, traditional stratigraphy is also burdened by the widespread use of litho-stratigraphic formation names. Despite Krynine’s comment that “lithostratigraphic nomenclature is the triumph of terminology over common sense,” there always were and still remain many practical and plausible reasons that justify the introduction (and retention) of formation names to correlate lithos- tratigraphic intervals over a wide region. Charts are also available to correlate var- ious named formations from one region to another. Even so, the ubiquitous and bewildering number of formation names often impedes communication from one region to another and also between earth scientists with differing specializations. Murphy and Salvador (1994) published a sensible abridged version of the official International Stratigraphic Guide. There is a consensus regarding systems/periods, series/epochs, stage/ages, and theirnames.Thecolorcodeshownonthe International Stratigraphic Chart has been widely used for decades and corresponds to codes used by Commission of the Geological Map of theWorld (CGMW). The International CommissiononStratigraphy-ICS (2010)hasproducedaonepage International Strati- graphic Chart that is widely available on the web and that is regularly updated. That chart is based on the expanded and carefully documented Geologic Time Table of Gradstein et al. (2005). A shorter version by Ogg et al. (2008) is supplemented with outstanding illustrations and a selection of plate tectonic reconstructions. While there is a consensus on the names of almost all major stratigraphic sub- divisions, time boundaries in million years (Mas) have changed over the years because of progressively tighter age control. Worldwide correlation of formations based onmarine pelagic fossils, is often preferred as the first order control for long- distance correlations while the tie from and to continental and lacustrine deposits is best based on palynological constraints. It is important to be aware of paleogeo- graphic and paleoclimatic influences on fossil assemblages. When judging ages based on fossils, it is important to recognize the processes leading to the disartic- ulation of skeletons and shells, before and after burial, as well as the reworking and transportation of fossils that are all part of the “taphonomy” of fossil assemblages. When reading older publications that use an outdated time scale, it is often desir- Stratigraphywas radicallychangedwith the introductionofmodern sequencestratig- China Geoscience Institute, 2004. Geological Map of West China and Adjacent Areas. Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps 1:2 500 000. In 8 sheets and explanatory notes. Geological Press, Beijing. Choubert, G., Faure-Muret, A., 1976/1983. Geological Atlas of the World. 1:10M, 22 sheets with text. CGMW (Commission of the Geological Map of theWorld)/UNESCO, Paris (also available in ArcInfo format). Choubert, G., Faure-Muret, A., 1990. International Geological Map of Africa. 1: 5 000 000M. In 6 Sheets. CGMW (Commission of the Geological Map of the World), UNESCO, Paris. Committee for the Gravity Anomalymap of North America, 1987. Gravity AnomalyMap of North America. Geological Society of North America, Boulder, CO. Decade of North American Geology Continental Scale Maps 002. 1:5 000 000. 4 sheets. Committee for the Magnetic Anomaly Map of North America, 1987. Magnetic Anomaly Map of North America. Geological Society of North America, Boulder, CO.Decade of North American Geology Continental Scale Maps 003. 1:5 000 000. 4 sheets. raphy, based on the interpretation of reflection seismic profiles (e.g., Mitchum et al., 1977; Vail, 1975; Vail et al., 1977). For an outstanding recent overview of modern sequence stratigraphy, see Catuneanu (2006). More of this is also discussed in Chapters4Vol.1A (RobertsandBally), 13Vol. 1A (Bertram)and22Vol. 1A (Gallagher) of this volume. However, in the context of this introduction to a global map series, it is important to realize that in addition to structural information and the distribution of igneous andmetamorphic rocks, geologicalmaps also include a very large amount of stratigraphic information. An understanding of all the information shown on regional and more detailed local geological maps and their legends is an obvious prerequisite for an in-depth understanding of the evolution of larger regions. References to25.1.Global Topography and Plate Tectonics Aghanabati, A., 1993. GeologicalMapof theMiddle East. 1:5 000 000, 1 Sheet. Geological Survey of Iran. CGMW (Commission of the Geological Map of the World), Paris. Asch, K., 2005. Geological Map of Europe and Adjacent Areas (IGME 5000). 1:5 000000 (reduced version 1:10 000 000 also available), second ed. BGR#. CGMW (Commission of the Geologi- cal Map of the World), Paris. Berghaus, H., 1892. Atlas der Geologie der Welt. 15 coloredmaps and text. Justus Perthes, Gotha. Blackwell, D.D., Steele, J.L., 1992. Geothermal Map of North America. Geological Society of North America, Boulder, CO. Decade of North American Geology Continental Scale Map 006. 1:5 000 000. 4 sheets. Bouysse, P., 2000. Geological Map of the World. 1:25 000 000, third ed. (In Mercator and polar stereographic projections). In 3 sheets. CGMW (Commission of the Geological Map of the World)/UNESCO, Paris (also available in Arc Info format). Bouysse, P., 2009. Geological Map of the World. 1:50 000 000, third ed. (In Mercator and polar stereographic projections). Sheet 1: Physiography, volcanoes and astroblemes (centered on Pacific Ocean); Sheet 2 Geology and Structure (centered on Atlantic Ocean). CGMW (Com- mission of the Geological Map of the World)/UNESCO. Bouysse, P., 2010. Geological Map of the World. 1:25 000 000, third ed. (In Mercator and Polar stereographic projections). In 3 sheets (free choice of centering sheets) with explanatory notes. CGMW (Commission of the Geological Map of the World) UNESCO, Paris. China Geoscience Institute, 1975. Geological Map of Asia. 1:5 000 000. In 20 sheets. Geological Press, Beijing. 979 Cook, T.D., Bally, A.W., 1977. Stratigraphic Atlas of North and Central America. Princeton Univer- sity Press. 188 Maps, 272 pp. 980 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Cox, A., 1973. Plate Tectonics and Geomagnetic Reversals. Readings with instructions. W.H. Freeman and Co, San Francisco, 702 pp. de Margerie, E., 1921. La face de la Terre (authorized French translation of Suess 1885–1909), vols. 1–4. Librairie Armand Colin, Paris, 3079 pp. Dennis, J.C., 1967. International Tectonic Dictionary, English Terminology. AAPG Memoir 7, 196 pp. Dietz, R.S., Holden, J.C., 1974. Collapsing continental rises: actualistic concept of geosynclines—a review. In: Dott, R.H., Shaver, R.H. (Eds.), Modern and Ancient Geosynclinal Sedimentation, vol. 19. Societyof EconomicPaleontologists andMineralogists. Special Publication, pp.15–25. Dott, R.H., 1974. The geosynclinal concept. In: Dott, R.H., Shaver, R.H. (Eds.), Modern and Ancient Geosynclinal Sedimentation, vol. 19. Society of Economic Paleontologists and Mineralogists. Special Publication, pp. 1–14. Dott, R.H., Shaver, R.H. (Eds.), 1974. Modern and Ancient Geosynclinal Sedimentation, vol. 19. Society of Economic Paleontologists and Mineralogists. Special Publication, 380 pp. Engdahl, E.R., 1988. Seismicity Map of North America. Geological Society of North America, Boulder, CO. Decade of North American Geology Continental Scale Maps 004.5 Sheets Scale 1:5 000 000. Exxon Production Research, 1985. Tectonic Map of the World, World Mapping Project. Scale 1:5,000,000, 20 panels. American Association of Petroleum Geologists. Gradstein, F.M., Ogg, J.G., Smith, A.G., et al., 2005. A Geological Time Scale 2004. Cambridge University Press, Cambridge, 610 pp. See also: www.stratigraphy.org. Guettard, J.E., 1746. Me´moire et carte mine´ralogique sur la nature et la situation des terrain qui la France et l’Angleterre. Memoires Acade´mie Royale des Sciences, pp. 363–393. 2 plates. Guettard, J.E., 1752. Me´moire dans lequel on compare le Canada a la Suisse, par rapport a ses mine´raux. Histoire de l’Acade´mie Royale des Sciences, In 3 parts, pp. 189–220, 323–360 and 524–538. 4 plates and 2 maps. Guettard, J.E., Monnet, A.G., 1780. Me´moire et carte ge´ologique de la France. Entrepris par l’ordre du roi France, 214 pp. 31 maps. Hallam, A., 1973. A Revolution in the Earth Sciences: From Continental Drift to Plate Tectonics. Clarendon Press, Oxford, 127 pp. Haghipour, A., 2001. Seismotectonic Map of the World at 1:25 000 000 in 3 Sheets. CGMW (Commission of the Geological Map of the World)/UNESCO, Paris. Haghipour, A., 2002. Seismotectonic Map of the World at 1:50 000 000 in 1 sheet. CGMW (Commission of the Geological Map of the World)/UNESCO, Paris. Haghipour, A., 2006. Structural and Kinematic Map of the World at 1:50 000 000 in 1 sheet (digital version in high resolution format). CGMW (Commission of the Geological Map of the World)/UNESCO, Paris. Harrison, C. et al., 11 lead compilers, 2008. GeologicalMap of the Arctic. 1: 5 000 000 in 5 Sheets. Ottawa Geological Survey of Canada (GSC Open File 5816) and CGMW Commission of the Geological Map of the World//UNESCO, Paris. International Commission on Stratigraphy-ICS, 2010. International Stratigraphic Chart. ccgm .freefr/accueilgb.html. Johnston, A.K., 1856. The Geological Structure of the Globe According to Ami Boue´, with Addi- tions to 1855. Blackwood and Son’s, Edinburgh and London. I sheet. Khain, V.E., Leonov, Y., 1996/1998. International Tectonic Map of Europe at 1:5 M, third ed. CGMW (Commission of the Geological Map of the World)/UNESCO, Paris. (Only available in high resolution pdf. format). King, P.B., 1969. Tectonic Map of North America 1:5 000 000 M. U.S. Geological Survey. Ma Lifang, 2003. Geological Atlas of China. 59 map sheets, tables and photos, Geological Pub- lishing House, Beijing, 348 pp (in English). McKenzie, D.P., 1977. Plate tectonics and its relationship to the evolution of ideas in the geologic sciences. Daedalus 106, 97–124. Marcou, J., 1875. Explication d’une seconde edition de la Carte Ge´ologique de la Terre. One colored map. Wurster/Stanford/Savy, Zurich/London/Paris, 222 pp. Muehlberger, W.R., 1992. Tectonic Map of North America. American Association of Petroleum Geologists, Tulsa, OK, 4 sheets, l:5 000 000. Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Murphy, M.A., Salvador, A.C., 1994. IUGS and ICG: international stratigraphic guide—an abridged version. Episodes 22, 255–271. Neuendorf, K.E., Mehl, J.P., Jackson, J., 2008. Glossary of Geology, fifth ed., 800 pp. Ogg, J.G., Ogg, G., Gradstein, F.M., 2008. The Concise Geological Timescale, Cambridge Univer- sity Press, 177 pp. Oreskes, N. (Ed.), 2003. Plate Tectonics, An Insider’s History of the Modern Theory of the Earth. Westview Press, p. 448. Peive, A.V., Yanshin, A.L., 1980. Tectonic Map of Northern Eurasia. Akademia Nauk SSSR. Puchkov, V.N., 1994. Tectonic phase and cycles in the context of plate tectonics. Geotectonics (English translation) 8, 267–271. Pubellier, M., 2008. Structural Map of Eastern Eurasia. 1:12 500 000, 1 Sheet. CGMW (Commis- sion of the Geological Map of the World)/UNESCO, Paris. Reed, J.C., Wheeler, J.O., Tucholke, B.E., 2005a. Geologic Map of North America, 1:5 000 000 in 2 Sheets. Geological Society of America, Boulder, CO. Reed, J.C., Wheeler, J.O., Tucholke, B.E., 2005b. Decade of North American Geology. Geologic Map of America—Perspective and Explanation. Geological Society of America, Boulder, CO, 28 pp. Russian Academy of Geosciences, 1998. Resource and Environment—World Atlas. Compiled by the Institute of Geography, Ho¨lzel, Vienna, 2 vols, 190 plates with comments. Schlu¨ter, T., Trauth, M.H., 2008. Geological Atlas of Africa, second ed. Springer, Berlin, 308 pp. Schobenhaus, C., (Compiler). 2001. (reprint 2005). Geological Map of South America. 1:5 5000 000. Geol. Survey of Brazil—DNPM; Available in printed and digital format: CGMW (Commission of the Geological Map of the World)/UNESCO, Paris. Smith, W., 1815. Geological Map of England, Wales and Southern Scotland. John Carey, London. Staub, R., 1928. Der Bewegungsmechanismus der Erde. Gebru¨der Borntraeger, Berlin, 1 coloured map 44 Figs, 270 pp. Stille, H., 1924. Grundfragen der vergleichenden Tektonik. Borntra¨ger, Berlin, 443 pp. Suess, E., 1885–1909. Das Antlitz der Erde. Prag and Vienna, Tempsky, vol. 1 (1885) 778 pp; vol. 2 (1888) 703 pp; vol. 3. part 1(1901) 508 pp; and part 2 (1909) 789 pp. Translated by Sollas, H.B.C., Sollas, W.J. (1904–1924). The Face of the Earth (authorized English translation) Oxford, Clarendon. vol. 1 (1904) 104 pp; vol. 2 (1906) 556 pp; vol. 3 (1908) 400 pp; and vol. 4 (1924) 170 pp; Translated by E. de Margerie E. (1921). (Authorized French translation) vol. 1–4, 3079 pp. Paris, Librairie Armand Colin, 4 vol, 3079 pp. Umbgrove, J.H.F., 1947. The Pulse of the Earth. Martinus Nijhoff, The Hague, 358 pp. 8 color plates. Winchester, S., 2001. The Map that Changed the World. Harper Viking, 352 pp. Yanshin, A.L., 1966. Tectonic Map of Eurasia. 1:5 000 000 Scale in 9 Sheets. Akademia Nauk SSSR or Izd Nauka. Zoback, M. et al., compilers, 1987. Stress Map of North America. Geological Society of North America, Boulder, CO. Decade of North American Geology. Continental scale maps 005 1:5 000 000. 4 sheets. 981 982 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps 2255..22 Global topography and bathymetry: The face of old earth reworked and modified by present processes Introduction In the late 1700s, James Hutton (1788, 1795) in his “Theory of the Earth” and Playfair (1802), proposed what much later was to be called the “geostrophic cycle”; that is, sediments were deposited, then deformed into folds, and subjected to high temperatures/pressures, finally leading to intrusions, thus forming what today we call folded belts or orogens. These, in turn, were eroded and truncated by unconformities that formed a base on which sediments were deposited thus marking the inception of a new cycle. Lyell (1830–1833) in his “Principles of Geol- ogy” expanded this into the important concept of uniformitarianism, which later on evolved into the popular dictum “the present is the key to the past.” Today this is commonly understood tomean that present day geological processes should be the key to past geological processes. Due emphasis on present processes as a key to the past needs to be amplified by observing that the present face of the Earth has been modified/reworked by these processes over long time spans ranging from over billions of years for the Precambrian to hundreds of million years for the Phanerozoic and even much shorter intervals for Tertiary thermotectonic events and Quaternary Glaciations. The conceptual “Wilson-cycle” (Wilson, 1966) has morphed into a didactic man- tra now often described as an “ocean opening and ocean closing” cyclical process beginning with rifting and ending with mountain building. Some authors also link the Wilson cycle to rock cycles. Notions of cyclicity, episodic global events, and/or episodic “revolutions” have permeated the geological literature of the past two centuries. That these concepts need to be taken with more than a grain of salt is best illustrated by the present day global topography. The present morphology of the Earth’s surface shows various stages of rifting, opening, and spreading of oceans as well as backarc opening and spreading. Coe- val subduction-related orogenic processes may begin with juvenile subduction of oceanic crust(s) eventually leading to processes that form subduction-related col- lisional orogens, to be followed by overall “epeirogenic” uplift of the orogen and its “cratonic foreland” eventually morphing into various stages of peneplanation processes and ending with coastal piedmont plains and shields. An, admittedly trivial, observation is that all these processes are going on simulta- neously today in various parts of the world. However, over time, these perennially ongoing processes shift and modify tectonic plates, climates, and ocean circula- Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Plate 25.1 is presented on the sameMercator projection asmost of the other maps of this series. Plate 25.2A and B are based on polar projections that are the same, but cover a much larger area than the polar tectonic maps (Plate 25.16A and B) and polar basin maps (Plate 25.32A and B). ModernGlobal ReliefMaps (e.g., Plates 25.1 and 25.2A,B) show that today a great variety of tectonic provinces and processes concurrently occur in various locations on our Earth; for example, relatively stable cratonic areas presently co-exist with continental rifts (e.g., East Africa), young oceans (e.g., Red Sea), mid-ocean ridges that were active only since the Tertiary (e.g., Northern Atlantic), or else since the Jurassic (e.g., Central Atlantic and much of the Pacific) or the Cretaceous (e.g., South Atlantic). tion patterns. Thus, global geomorphology and the distribution of plates and con- tinents have been changing in space and over time. Tectonic processes implicate four dimensions that are ubiquitous in time as well as in space. However, rarely, if ever, do they repeat themselves in the same location. Thus, new tectonic trends are frequently and likely to be superimposed discordantly on previous tectonic trends and the re-activation of old tectonic trends is selective, fortuitous, and far from common. While the classical Wilson cycle remains as an elegant simplifi- cation, it must be kept inmind that sequences of “Wilsonian themes” rarely repeat themselves within the same regional boundaries. Also, as in the case of rift systems, it is obvious that many rifts record ephemeral extensional events that frequently aborted and never led to the opening of an ocean. Global relief models: Onshore and submarine morphology and plate tectonic regimes (Plates 25.1 and 25.2A,B) The general surface topography andmarine bathymetry of the earth is best shownby global relief models. The present-day first order global geomorphology offers an important perspective on a “present that is re-working the past.” Themaps emphasize the broad topography of the world such as the submarine relatively elevated mid- ocean ridges, abyssal plains, deep sea trenches, the shallow shelves, low-relief active and passive margins, elevated plateaus associated with active rifts, and the high plateaus/elevationsassociatedwithcollisional/orogenic systemsorwithpolar icecaps. ETOPO 1 combines a global sea floor bathymetry based on satellite-based radar altimetry and gravimetry with satellite-derived topographic maps on land (Amante and Eakins, 2009; Sandwell and Smith, 1997; Sandwell, 1990). The data for our maps were taken from http://www.ngdc.noaa.gov/mgg/global/ global.html. 983 Complex “ocean-closing stages” in action can all be observed on the present day 984 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps global relief map (Plate 25.1). It is sensible to view the morphologic expression of various orogenic stages as an evolutionary continuum in both time and space. This in turn suggests that classical present day hypsographic domains lump tectoni- cally widely diverse provinces that may share similar elevation ranges. Based on a global hypsometric curve, Watts (2007) derives amodal topography of 0.303 km on land and 4.303 km under water to emphasize the basic difference between the continents and oceans. However, closer scrutiny of selected slices of the hypsometric curve reveals areas of widely differing origin that often share common elevations and suboceanic depth ranges as follows: (1) Elevations over 2000 m occupy about 5% of the solid surface of the Earth. These elevated areas include the ice caps of Greenland and Antarctica, the “stable cratonic areas” of South Africa, the Nubian Arabian shield and the associated East African rifts, the Mesozoic-Cenozoic folded belts including the Alpine/Himalayan, the Andean, and North American Cordilleras, and the associated, mostly late- to post-orogenic, “epeirogenic” plateau uplifts. The word “mountains” (say elevations over 2000 m) includes a great variety of tectonic/morphologic objects. However, the term “orogeny” (i.e., mountain building) is traditionally reserved for “folded belts” including all their coeval igneous and metamorphic associations. Some authors resolve the resulting semantic dilemma by broadening themeaning of “mountain building” to include most elevated land areas (e.g., Ollier, 2000, 2005). “Orogenic” events and pro- cesses are often, and perhaps confusingly, contrasted with their “epeirogenic” counterparts and it should be realized that broad regional uplifts often overlap in time and space with subduction-related orogenic areas. Dennis (1967) offers a concise historical explanation of these terms, and Sengo¨r (2003) presents an in-depth historical perspective on large wavelength deformation of the litho- sphere. In the context of this map series, we use the terms “folded belt” and “orogeny” loosely and interchangeably for the product of all subduction-related, penetrative deformation processes. Today, the physiographic expression of ongoing orogenic processes ranges from the heights of the Himalayas to the depths of oceanic deep-sea trenches. Included in this, are subduction-related intrusive and extrusivemagmatism aswell as a great variety of metamorphic processes. To conclude, the term “orogeny” is, pedantically speaking, a misnomer. Its mean- ing has now morphed over time from describing “mountain building processes” into “subduction-related mountain building processes and their subsequent tectonic evolution,” which also includes their transition into supra-regional peneplanation processes. Ultimately, all elevated land (i.e., “hard-rock areas”), in their quest for the ultimate base level, will end up being eroded to become the basement of future sedimentary basins. (2) About 20% of the surface of earth surface lies between 2000 and 200 m Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps above sea level. This large area includes the relatively stable continental inte- riors, as well as the onshore extension of passive and some active continental margins. Evidently, during Late Cretaceous, a greater part of this area was covered by shallow seas that were in continuity with parts of the areas now occupied by mountain ranges. While Late Cretaceous stratigraphy may be well explained by sea-level changes, the “continental cratons” of the world as well as the adjacent mountains were eventually elevated because of a variety of tectonic causes. (3) Isolating areas between>200 m above and>200 m below sea level outlines coastal plains and the neritic shelves of passive margins and active margins that occupy about 17%of the Earth’s surface. The stratigraphy of these wider coastal domains is modulated by eustatic sea-level changes that are generally thought to be the root cause of the alternating deposition of higher-ordered transgressive and regressive stratigraphic sequences (e.g., Catuneanu, 2006; Mitchum et al., 1977 and Payton 1977). Coastal and shallowmarine erosion associated with low sea levels creates regional unconformities that define the boundaries of stratigraphic sequences. However, further back in time, coastal plains and neritic shelves covered much wider continental areas as shown by the craton-wide distribution of Lower and Middle Ordovician sediments and also of Middle and Upper Cretaceous sediments. Sea-level changes are the immediate cause of changes in the areal distribution of coastal and shallow water sediments. Nonetheless, it will always remain a challenge to isolate the differing principal contributing factors that change eustatic sea levels to the stratigraphic record. These range from orbital causes, to changing dynamic topography, to climatic (e.g., subglacial, glacial, and/or erosional in deserts), and to a great variety of regional and local tectonic processes. There is no likelihood of there ever being a clearly identifiable, single, and overall dominant component responsible for eustatic sea-level changes, for example, ice ages and their demise, the deposition of thick wedges of sediments or the tectonic response to these processes, and most importantly plate tectonic processes themselves. Eustatic sea-level changes register the combined impact of all these processes. Consequently, the present hypso- metric curve does not easily permit the construction of hypothetical past hypsometric curves. (4) Mid-ocean ridges, many seamounts, and oceanic volcanic plateaus, as well as continental slopes along both active and passive margins range between 200 and 4000 m depths below sea level. The bathyal and upper abyssal regions of the oceans of the Earth occupy about another 17% of the Earth surface. On many passive margins, continental slope areas are dominated by gravitational, salt, and/or shale slope tectonics. Continental slopes are also the locus of widespread erosion in submarine canyons and also 985 by geostrophic, that is, submarine contour currents that erode the conti- 986 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps nental slopes rise to carve profound “slope” unconformities that are plainly recognizable on reflection seismic profiles. For an introduction to deep-water erosion see Heezen et al. (1966), Stommel (1958), Broecker (1991), Zenk (2008) and Rebesco and Camerlenghi (2008). Ideally, geological outcrop maps on continental slopes should reveal the impact of erosion along submarine slopes. However, as pointed out by Tucholke (in Reed et al. 2005), submarine outcrops on the continental slopes and also on the deeper parts of the ocean are often overlain by a thin superficial veneer of younger sediments that prevent identification, sampling, and determination of the age of elusive, “nearly outcropping” rocks. (5) About 45% of the ocean is located between water depths of 4000–7000 m that are commonly referred to as middle and lower abyssal depths. These depths coincide mostly with the ocean floor and the lower slopes of sea- mounts, oceanic ridges, and submarine plateaus. (6) Finally, less than 2% of the abyssal ocean floor below 7000 m is occupied by deep sea trenches, associated with subduction zones that define the outer parts of active margins. Typically, the deep sea trenches of active subduction zones are filled with sediments that are separated by an unconformity and overlap on the underlying, downflexed arc-ward dipping oceanic slab and its sediment cover. The suggestion is that the sedimentary infill of trenches is not only derived from nearby island arcs but also may include sediments redistributed by arc-parallel contour currents that sweep along the sides of the trench. Plate 25.2A andB showon separatemaps the topographyof theArctic andAntarctic regions with their continent-anchored icecaps and ice, and also the present topog- raphy of the sea floor. The bedrock map shows the formation of a supraregional/ subglacial unconformity as “awork inprogress” affecting anarea larger thanEurope. The present bedrock surface of polar ice caps does not take into account the effect of isostatic rebound that would be associated with the melting of the ice caps nor the impact on changing coastlines following the disappearance of the polar ice caps. A useful historical perspective by Kru¨ger (2008) offers an overview over advances made from the time of the recognition of ice ages to the present debate about global warming. Petit Maire and Bouysse (1999, 2000) image the effects of climate change with their reconstructions of the global environment during the last glacial maximum at 18,000 � 2000 years B.P. and during the following glacial optimum at 8000 � 1000 years B.P. By analogy, their maps provide a broader perspective onpresentlyongoingglobalwarming.Accordingly,duringthemaximumglaciation, with average temperatures 4.5 �C lower than the present, Arctic ice sheets covered most of North America down to the 45th parallel, as well as Northern Europe (down to say the 53rd parallel), and the low-lying areas of Western Siberia (down to say the 65thparallel). TheAntarctic ice sheet coveredmost of the Antarctic shelf areas. During the glacial optimum, with average temperatures 2 �C lower than the pres- ent, the Arctic ice retreated to Greenland,much of Ellesmere Island and a few addi- References to25.2.Global Topography Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Amante, C., Eakins, B., 2009. ETOPO1 Arc-Minute Global Relief model: Procedure, Data Sources and Analysis. NOAA Technical Memorandum NESDIS NGDC-24, 19. Broecker, W., 1991. The great ocean conveyor belt. Oceanography 4, 79–89. Catuneanu, O., (reprinted 2008). 2006. Principles of Sequence Stratigraphy. Elsevier B.V., London/Amsterdam, 375 pp. Dennis, J.C., 1967. International Tectonic Dictionary, English Terminology. AAPG Memoir 7, 196 pp. Heezen, B.C., Hollister, C.D., Ruddiman, W.F., 1966. Shaping the continental rise by deep geo- strophic contour currents. Science 152, 506–508. Hutton, J., 1788. Theory of the earth; or an investigation of the laws observable in the composi- tion, dissolution, and restoration of Land upon the globe. Trans. R. Soc. Edinb. 1, Part 2, p. 209–304, plates I and II. Hutton, J., 1795. Theory of the Earth, with Proofs and Illustrations, vols. I & II. Cadell & Davies, London; William Creech, Edinburgh; vol 3. Geological Society, London. Kru¨ger, T., 2008. Die Entdeckung der Eiszeiten. Rezeption und Konsequenzen fu¨r Versta¨ndnis der Klimageschichte. Benno Schwabe, Basel, 619 pp. Lyell, C., 1830–1833. Principles of geology, being an attempt to explain the former changes of the Earth’s surface, by reference to causes now in operation. 3 vols. Re-issued by. Secord, J.E. (Ed.) (1977). Penguin Classics. With Introduction. John Murray, London, p. 48, Text 475. Mitchum, R.M., Jr., Vail, P.R., Sangree, J.B., 1977. Seismic stratigraphy and global changes of sea level, Part 6. In: Payton, C.E. (Ed.), Seismic Stratigraphy—Applications to Hydrocarbon Exploration. American Association of Petroleum Geologists Memoir 26, pp. 117–134. tional islands. In the Antarctic, the ice receded to the onshore areas of the Antarctic and remained on the Ross and Ronne ice shelves. Aside from the short-lived “Little Ice Ages” between 1300 and 1870, the present earth has been on an increasing warming trend. The rising of the land underlying the glacial ice caps is likely to be accompanied by receding coastlines. Today, the polar ice regions have already receded more than during the Holocene Optimum, which was estimated as an average of 2 �C warmer than today. Petit Maire and Bouysse (1999 and 2000) clearly emphasize that their maps explore the effects of “natural warming” and natural variability of environmental factors, and not the effect of “unnatural changes” associatedwith human activities. The latter need to be evaluated against a background of natural change. To evaluate future global warming due to both “natural” and “human/unnatural” causes, we will have to rely on increasingly more refined models that also fully account for the effects of isostatic rebound due to the melting of ice and other factors. Global models and maps that attempt to predict the future will, at any given time, display both rising and drowning coastlines. As shown by Petit Maire and Bouysse (1999, 2000), the overall environ- mental effects such as changing patterns of vegetation or the reduction of perma- frost need to be predicted together with the expected impact of counteraction by mitigation efforts within a given realistic time frame for the future. 987 Ollier, C.D., 2000. Mountain building and orogeny on an expanding earth. In: Wezel, F.C. (Ed.), the World and ANDRA (French national Agency for Nuclear Waste), Paris. 988 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Petit Maire, N., Bouysse, P.h., 2000. Geological record of the recent past, a key to the near future world environments. Episodes 23, 230–243. Playfair, J., 1802. Illustrations of the Huttonian Theory of the Earth. William Creech, Edinburgh; Facsimile Reprint with an introduction by G.W.White 1956 Urbana: University of Illinois Press. Introduction p.1–18; text, 528 pp. Payton, C.E. (Ed.), 1977. Seismic Stratigraphy—Applications to Hydrocarbon Exploration, Ameri- can Association of Petroleum Geologists Memoir 26, 516 pp. Rebesco, M., Camerlenghi, A. (Eds.), 2008. Contourites. Developments in Sedimentology 80. Elsevier, Amsterdam, 663 pp. Reed, J.C., Wheeler, J.O., Tucholke, B.E., 2005. Decade of North American Geology. Geologic Map of America—Perspective and Explanation. Geological Society of America, Boulder, CO, 28 pp. Sandwell, D.T., 1990. Geophysical applications of Satellite Altimetry. Rev. Geophys. Suppl. 132–137. Sandwell, D.T., Smith, W.H.F., 1997. http:www.ngdc.noaa.gov/mgg/bathymetry/predicted/ explore.HTML. Sengo¨r, A.M.C., 2003. The LargeWave-Wavelength Deformations of the Lithosphere: Material for a History of the Evolution of Thought from the Earliest Times to Plate Tectonics. Geological Society of America Memoir, Boulder, CO, 365pp. Smith,W.H.F., Sandwell, D.T., 1997. Global sea floor topography from satellite altimetry and ship, depth soundings. Science 277, 1956–1962. Stommel, H., 1958. The abyssal circulation. Deep Sea Res. 5, 80–82. Watts, A.B., 2007. An overview. In: Watts, A.B. (Ed.), Crust and Lithosphere Dynamics. Vol. 6. of G.R. Schubert (Ed.), Treatise of Geophysics. Elsevier, Amsterdam, pp. 1–48. Wilson, T.J., 1966. Did the Atlantic close and then re-open? Nature 211, 676–681. Zenk, W., 2008. Abyssal and contour currents. In: Rebesco, M., Camerlenghi, A. (Eds.), Contour- ites. Elsevier, Amsterdam, pp. 37–58. 2255..33 Neotectonics; earthquakes and conventional (i.e., rigid) versus diffuse plate boundaries Neotectonics: Introduction Dramis and Tondi (2005) reviewed the use of the term “Neotectonics” in the con- text ofWestern Europe. Formost authors, “Neotectonics” includes presently active tectonics as manifested by earthquakes, present stresses, relative displacement Earth Dynamics Beyond the Plate Paradigm, vol. 5. Bolletino della Societa Geologica Italiana. Spec., pp. 169–176. Ollier, C.D., 2005. Mountain uplift and the neotectonic period. Annali di Geofisica 48 (Suppl.), 971–984, Chapter 9. Petit Maire, N., Bouysse, P.h., 1999. Map of the World Environment During the Last Two Climatic Extremes. Map 1: The last Glacial Maximum (18 000� 2000 years B.P); Map 2: The Holocene Optimum (8000� 1000 years) 1:50 000 000. CGMW (Commission of the Geological Map of ofplatesbasedon satellite-basedgeodeticmeasurements, etc.,Goingback in time, (1968). Diverging, converging, and transforming plate boundaries of rigid plates Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps became commonly accepted and illustrated by large-scale maps that also showed the corresponding earthquake mechanisms. The data shown on Plates 25.3–25.6 were obtained from the earthquake catalog archived by IRIS (Incorporated Research Institutions for Seismology at www.iris.org). Earthquake location, depth, and magnitude information were obtained from both the Bulletin of the Interna- tional Seismological Centre and the Monthly Hypocenter Data Files. Stein and Wysession (2010) present an in-depth discussion of the relationship of seismology and plate tectonics. Plates 25.3–25.6 include all events reported at magnitude 4 or greater occurring in 1997 through 2003 and all earthquakes with reported depths between 0 and 70 km, and separately all earthquakes with reported depth greater than 70 km. Haghipour (1992, 2001, 2002, 2006) com- piled regional (i.e., Middle East) and global tectonic wall maps based on orogenic regimes that also show the distribution of earthquakes reported over the past 5000 years. Two Geodynamic Maps of the Mediterranean prepared by Barrier et al. (2004) are another example of maps that combine neotectonics based on a mostly geodetic model (GDSRM 1 of Kreemer et al., 2003) with earthquake distribution and focal mechanism maps. Mesozoic and Cenozoic tectonics are recorded in detail in the oceanic crust (e.g.,Mu¨ller et al., 2008), the adjacent passivemargins, andon the continents. This is in great contrast to Paleozoic and Precambrian tectonics that are only preserved on continents and no longer have a continuous, preserved oceanic record. Neotectonics is the bridge between Present and Cenozoic/Mesozoic tectonics and neotectonicobservationsmaybeextrapolatedbackwith someconfidence to theEarly Miocene (e.g., Becker, 1992). However, within the context of regional geology, neo- tectonics can only credibly link the present to the past when the continuity of its pres- ent record, back into the Neogene and earlier times, is reasonably well documented. The waxing and waning of ice sheets during the Pleistocene is commonly viewed as a sequence of climatic events. However, the accompanying loading and later unloading of ice sheets are often viewed as tectonic consequences of climate changes. Similarly, ongoing/neotectonic regional uplifts associated with the unroofing of rising mountain ranges are also viewed as tectonic processes. This emphasizes the importance of precisely specifying the types of processes included in the term “Neotectonics” before projecting them into a more remote past. Global earthquake distribution (Plates 25.3–25.6) Mapping the global distribution of earthquakes led to the original definition and delineation of the tectonic plates proposed by Sykes (1966) and Isacks et al. 989 A review of the practical/engineering aspects associated with earthquake 990 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps seismology falls beyond the purview of this summary. For an overview, see Lee et al. (2002). Well-defined versus diffuse plate boundaries (Plates 25.7–25.10) Beginning in the mid-1960s, plate tectonics became the widely accepted para- digm of modern geology. Among many others, Cox (1973), Hallam (1973), McKenzie (1977), Oreskes (2003), Gordon (1998), and Bird (2003) offer insights into the still evolving history of the plate tectonics paradigm. Early on, a limited number of torsionally rigid, tectonic plates were thought to be circumscribed by well-defined boundaries. These plates were assumed to move with respect to one another on a spherical planet and theirmotionswere best described as rota- tions around Euler poles (McKenzie and Parker, 1967). Eventually, the basic assumptions of plate rigidity and the overall notion of narrowly defined plate boundaries were questioned. This was particularly the case for many continental interiors (e.g., Middle East and Central Asia) that underwent intense deformation duringmuch of the Tertiary and, even today, are characterized by an overall diffuse distribution of earthquake epicenters. The concept of wider, diffuse plate boundaries was introduced by Gordon (1998, 2000); see also Gordon and Stein (1992) and Stein and Sella (2002). Plates 25.7–25.10 are simplified and modified versions of maps published earlier by these authors. On Plate 25.8, the length of the arrows is proportionate to relative spreading and subduction rates. Selected numerical spreading and convergence rates were rounded up or down from Haghipour (2002, 2006). The introduction of diffuse plate boundaries based on platemotion studies also led to the recognition of smaller subplates some of which are shown on Plates 25.8–25.10. Bird (2003) has updated plate boundaries and lists 14 large plates that roughly correspond to the traditional large plates but also defines 38 small plates. Many of these are within, or at least partially overlapping, the diffuse plate boundaries of Gordon (2000). About 20 of these small plates are located in theWestern Pacific and Eastern Asia and between the Australian and the Southwestern Pacific plate. In particular, the small plates are well defined based on recent geodetic and earthquake observations. Much of the work on “quasi-rigid” plates and their boundaries led to tracking “geologically current plate motions” (e.g., Gripp and Gordon, 2002) that are based on Neogene seafloor spreading rates and the rapidly increasing satellite- based geodeticmotionmeasurementsmade over short time intervals. Earlier work by DeMets et al. (1990, 1994) was recently updated by the MORVEL l system developed and comprehensively introduced by DeMets et al. (2010) with a detailed description of plate boundaries and “geologically current plate motions.” These authors adopt the boundaries of Bird’s (2003) 25 large plates and also of a Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps selection of 14 additional smaller plates. Boundaries that today separate smaller plates in Eastern China are well defined. However, industry reflection seismic pro- files from China across many of the microplates and their boundaries also show clear and extensive evidence of late Paleogene to Miocene and younger exten- sional, transtensional, and transpressional tectonics. Thus, there may be limits as to how far back in time smaller, rigid plates and their boundaries might be pro- jected. In the context of regional geology, smaller plates may turn out to be more ephemeral than large plates. A reduced longevity of smaller plates may be relevant when evaluating regional tectonics backward in time over time spans in excess of, say, 25 Ma. The relatively short time span of “neotectonic” observations contrasts with a geological record that extends over some 545 million years for the Phanerozoic and an additional 4 billion years for a total of over 4.5 billion years. Regional geological and/or tectonic maps tend to agglomerate observations on mountain ranges and sedimentary basins that have evolved over very long time spans. Kreemer et al. (2003) specifically address present plate motions across boundaries adjacent to orogenic systems. Not surprisingly, their model shows the highest present strain concentrations along a broad band that coincides with the distribu- tion of Tertiary folded belts and active continental margins. A narrower band follows the mid-ocean ridges as defined by shallow focus earthquakes (e.g., Plates 25.3–25.6). Plate 25.9 is an admittedly generic attempt to show the relation of neotectonic plates and some of their diffuse plate boundaries to the distribution of Cenozoic/Mesozoic folded/orogenic belts (i.e., the Mz/Cz Megasuture of Bally and Snelson, 1980) and the distribution of Tertiary rift systems. There appears to be a rough correlation of neotectonic continental diffuse boundaries with Cenozoic/Mesozoic orogens for, say, the past 50 Ma. Going farther back in time to the Early Jurassic, Cenozoic/Mesozoic orogens, as well as rift systems, are better viewed as long-lasting and wide, diffuse plate boundaries. Notably, a number of oceanic, neotectonic diffuse boundaries extendwell into the Indian and Atlantic Oceans. These and additional observations have led Gordon and his co-workers to differentiate the Africa/Nubia, Arabia, Africa/Somalia, and Lwandle sub-plates, as well as the Capricorn subplate, that is, the westernmost segment of the former Australia plate (see also DeMets et al., 2010). The term “diffuse plate boundary” provides a welcome flexibility for various authors to circumscribe often somewhat differing boundaries (e.g., Chamot- Rooke and Rabaute, 2006). To conclude, particularly on continents and seen over much longer observed time spans of observation, the area of diffuse plate bound- aries broadens when combining the Recent with the Cenozoic/Mesozoic past. So much so that in older orogenic belts, pervasive penetrative deformation only 991 permits inference from reconstructions of the location of former rigid plates that gene time, smaller plates appear to be embedded in a matrix of diffuse plate 992 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps boundaries. Bird’s (2003) map shows wide areas as “orogens.” These in large part overlap with the diffuse plate boundaries of earlier authors. In Eastern Asia, a trio of small plates (i.e., Yangtze, Amuria, and Sundae) have been differentiated on Bird’s (2003)map and accepted by DeMets et al. (2010).While these authors are clearly justified in identifying these neotectonic smaller plates, the geological evidence also shows that theYangtzeandAmuriaplatesareseparatedbyabroad“taphrogenic” zone that is affected by a complex transtensional Paleogene/Neogene fault system, that is, the onshore North China platform, as well as the offshore China and Korean marginal seas (seeChapter8Vol.1A (LiDesheng)).Thus, seenovera longer timespan, a long-term rigorously rigid ancestry of these smaller plates is less evident. To conclude, our maps illustrate in a general way that indeed the Recent is the key to Neotectonics. However, going farther back in time, the rapidly decreasing pre- cision of regional geological generalizations and plate tectonic reconstructions is mainly due to the very large number and great variety of observations that are integrated over much longer geological time spans. Understanding and applying neotectonic insights to describe the tectonic setting of well-defined tectono- stratigraphic megasequences in sedimentary basins (see Chapter 4, Vol. 1A (Roberts and Bally)) also show that over time regional tectonic regimes will change episodically. However, because of shifting of plates and the assembly and dispersal of supercontinents over tens of million years or more and with changing stress regimes, the sequencing of tectonic processes and themes is much less likely to re-occur within the same geographic limits. Neotectonic plate motions: Their relation to a fixed Eurasia and to Cenozoic/Mesozoic fold belts (Plate 25.11) The relationship of Cenozoic–Mesozoic compressional/orogenic and extensional/ rifting domains to plate motion studies is based on the ocean spreading record and convergence rates at subduction boundaries that may be tied to a fixed were either subducted or else on the continents were involved in extensive ther- mal/tectonic reworking of the crust. Plate 25.10 shows a now commonly accepted distribution of all major lithospheric plates. Note that only the Pacific, the Philippine, and Scotia plates are entirely oce- anic plates, whereas all other plates are part continental and part oceanic. Abbre- viations also mark selected smaller plates that are either in part or wholly surrounded by diffuse boundaries, for example, the Lwandle, Africa/Nubia, Africa/Somalia, and the Arabian plates, and the smaller oceanic Caroline and Cap- ricorn plates were retained. In Eastern Asia, when viewed over most of the Neo- reference frame (e.g., selected hotspots, selected fixed continents, or else to fixed Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps regions and/or cratons that are judged to be relatively more stable). Recent plate motions based mostly on satellite-based geodetic observations measure instantaneous/ongoing displacements. Plates 25.9 and 25.11 use this information to compare the present plate motion regimes with regional Mesozoic/Cenozoic orogenic and taphrogenic regimes. These, to a degree, mimic the present and also suggest limitations to the use of simple direct projections and generalizations farther back in time. A compelling display of absolute place motions as they relate to a hot spot frame of reference by Gripp and Gordon (2002) shows a strong northwestward dis- placement at rates of over 100 mm per year in a band that extends from the Central East Pacific Rise (say, Easter Island) toward the Mariana islands. While maintaining their overall direction, these rates slow down to>100 to 50 mm/year in the NW and the SE Pacific. In the Indian Ocean, northerly convergence direc- tions dominate and range from about 110 mm/year west of Fiji to less than 30 mm/year in the Arabian Sea, showing an overall counterclockwise rotation of the Indian plate. A similar rotation projected backward to a more distant past is also reflected by the eastward, diachronous younging of the inception of the collision of India with Eurasia from Eocene in Pakistan to Pliocene in Northern Assam. Relatively smaller convergence rates are associated with the relative west- ward drift of the South and North America plates, respectively. Note that the African plate and the Antarctic plate rotate only nominally in a counterclockwise direction. Plate 25.11 is a much simplified and redrafted version based on “Plate tectonics from Space” (Chamot-Rooke and Rabaute, 2006), yet another fascinating neotec- tonic map. Mostly based on accurate satellite-based geodetic measurements, the motions of tectonic plates are here shown with respect to a fixed Eurasia. The length of the arrows is proportional to present motion rates ranging from about 10 to about 100 mm/year (i.e., “about the speed at which fingernails grow” to quote Stein and Wysession, 2010). To fit the small scale of Plate 25.11, the original, much larger scale, wall map of Chamot-Rooke and Rabaute (2006) was greatly simplified by omitting the smal- lest rates of motion (say Plates 25.11 and 25.30 explore the general relationship of “instantaneous” 994 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps neotectonic images to the distribution of Cenozoic/Mesozoic orogenic belts (including Tertiary rifting on Plate 25.30). Even allowing for its great simplification, Plate 25.11 allows the differentiation of a number of present megatectonic/ orogenic settings that were inherited from earlier Mesozoic and Cenozoic plate convergences. Table 25.2 attempts to classify the different neotectonic set- tings of orogens shown on Plate 25.11 and on the more detailed and accurate map of Chamot-Rooke and Rabaute (2006). Present motions with respect to Eurasia strongly suggest a variety of orogenic tectonic settings that are primarily influenced by the direction and the nature of converging crusts ranging from ocean–ocean convergence island arc systems to ocean–continent convergence and to various stages of continental collision. Today’s active orogens have a long ancestry that goes back to the Jurassic. Chamot-Rooke and Rabaute (2006) dis- play “present plate tectonics in action”while Cawood et al. (2009) give a structured overview of orogenic systems that span from the Archean until today. Given these twowidely different but important objectives, Table 25.2 aims to show the extent of ongoing present-day tectonics that may be tracked back in time. The present orogenic settings listed in Table 25.2 differentiate well over 11 oro- genic regions subdivided into 28 subregions. The differences between the regions is mainly due to varying convergence directions and rates that result in different strain partitioning, involving the different crustal types and ages of their respective “forelands” and “hinterlands.” In addition, the maturity of these orogenic processes as well as their transition into late postorogenic, large wavelength uplifts also varies from one region to another. Finally, from the low topography of the Paleozoic Appalachian and Ural systems (see Plate 25.1), it is evident that the complete development of any givenmountain system should also include that part of the system that was, and still is, eroded over a long time and is destined to become the “peneplaned” basement of a future basin. Today, a great variety of orogenic settings listed on Table 25.2 exist side by side and are also coeval with extensional rifts and ongoing ocean spreading (see Plate 25.30). This simple observation limits the scope of idealized cycles such as the “Wilson Cycle.” On the other hand, the purpose and scope of regional stud- ies based on a variety of geological, geophysical, and geochemical observations is likely to be greatly expanded by a greater emphasis on understanding the differ- ences among orogens. Neotectonic maps convincingly show the impact of present tectonic processes with substantial precision. However, combining present convergence regimes with a map of Cenozoic/Mesozoic orogens strongly suggests that, overall, pres- ent plate motions going back in time mimic past orogenic conditions with decreasing accuracy. Conversely, moving forward in time and, for example, beginning in the early Mesozoic with an inherited history that began in the Table 25.2 Plate motions with respect to a fixed Eurasia and Cenozoic/Mesozoic orogenic settings (Plates 25.11 and 25.30): Neotectonic equivalents of the orogen classification of Cawood et al. (2009). Western Pacific and Alpine orogens 1. W. Pacific orogens: Present NW converging Pacific plate: Present clockwise, NW-directed motion of Pacific plate with rates of 70–90 mm/year: Extensive Paleogene and Neogene backarc systems and rift systems (see Plate 25.30)—Retreating Accretionary Orogens 1.1. W. Pacific Intra-oceanic arc systems: Mariana/Bonin/Izu arc system 1.2. Inner W. Pacific Island arc systems: Koriak-Kamchatka-Japan arc systems 1.3. Transpressional Philippines ocean–ocean converging orogen 2. SW pacific transitional/transpressional boundary-related orogens: Present clockwise, NW-directed motion rates between 90 and 60 mm/year in the southwestern Central Pacific and the Southern Pacific and about 60 mm/year for the northerly moving Australian Plate (Retreating Accretionary Orogens) 2.1. Australia/New Guinea transpressional ocean–continent collision 2.2. Complex Bismarck/Solomon Islands-Vanuatu-Fiji Islands transpressional/transtensional ocean–ocean collision 2.3. Transpressional New Zealand-Tonga ocean–ocean and continent collision 3. SE and Central Asia orogens: Present anticlockwise N-NNE directed motion of India Plate in the southwest toward Central Asia and in the southeast towards the Indonesia/Arakan Yoma island system with convergence rates ranging between 50 in the east and 25 mm/year in the west 3.1. Australia/W. New Guinea-Timor/Australia continent/island arc collision (Retreating Accretionary Orogen) 3.2. Indonesia/Arakan/Yoma ocean/island arc collision (Retreating Accretionary Orogen) 3.3. India/SE and Central Asia orogen; southeasterly escape of SW China (Advancing Accretionary to Collisional Orogen) 4. Middle East collisional orogens of Arabian plate with SW Asia: Present anticlockwise N-directedmotion ranging from 25 in the southeast to 10 mm/year in the northwest suggests a nearly mature orogen in the northwest 4.1. Zagros and Central Iran orogens (Collisional Orogen) 4.2. Taurus and Anatolian orogens, presently escaping at, say, 45 mm/year toward the SE—Mediterranean Collisional/Retreating Accretionary (e.g., Aegean Orogen) 5. W-central Mediterranean/Alpine collisional orogens of Africa with Europe: PresentoverallanticlockwiseNW-directedmotionofAfricatowardEuropeof>1 mm/year isnominal suggestingrelatively“mature”orogens 5.1. Balkan-Hellenide orogen—Collisional Orogen 5.2. Pyrenean-Alpine-Carpathian orogen with southeasterly escape of Pannonian Basin—Overall Collisional Orogen, but Retreating Accretionary Orogen in the Eastern Carpathians 5.3. Apennine-Sicilian orogen with Tyrrhenian/Gulf of Lyon backarc systems—Retreating, Accretionary Orogen 5.4. North African Atlas-Betic orogen with Alboran backarc systems—Retreating, Accretionary Orogen Andean-Cordilleran orogens (including Scotia system) 6. Scotia orogen: A hairpin-shaped orogen bounding the Scotia plate with an eastern subduction boundary and two transform boundaries to the north and to the south. The enclosed backarc area involves two spreading centers. Present clockwise west-directed motion of the South America plate is about 20 mm/year while the Scotia plate moves with a rate of about 15 mm/year towards the west. The western boundary of this orogen is a transform fault that separates the Scotia plate from the relatively stable Antarctic plate— Overall, Retreating, Accretionary Orogen. 7. Andean orogens: South America continent/Pacific Ocean converging orogens: Present westwardmotion rate of S. America is about 20–25 mm/year and opposed by about 40 mm/year eastwardmotion of Nazca plate—Advancing, Accretionary Orogens. 7.1. Southernmost Chile and Argentina Andean orogen involving westerly motion of the Atlantic plate and an almost static Antarctic plate moving in a north/northeasterly direction at rates of > 10 mm/year. 7.2. Main Andean orogen extending from about 45� to 5� S latitudes Continued 995 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Table 25.2 Plate motions with respect to a fixed Eurasia and Cenozoic/Mesozoic orogenic settings (Plates 25.11 and 25.30): Neotectonic equivalents of the orogen classification of Cawood et al. (2009).—cont’d N.American - Cordille 8. Central American cord Present nominal westw plate—Advancing, Accr 8.1. Panama-Cost Rica 8.2. Nicaragua-Hondu 9. Mexico orogens 9.1. Southern Mexico continental Yucat Accretionary Orog 9.2. Central and Nort of North America associated with M Accretionary Orog 10. N. American/Cordille 10.1. Southern Cord accelerated in w San Andreas fa Paleogene/Rec adjacent forela Retreating, Accr 10.2. Northern Cord about 25 mm/ about 9 mm/ye 10.3. Alaska/Norther Southern Arctic year in the eas Orogen in the W 11. Caribbean orogens The Caribbean plate presently appears to move westward at ra Accordingly, the oro 11.1. The compressi 11.2. The transpress 11.3. The complex t 11.4. The Cuban tra 996 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps ran orogens (incl. Caribbean system) illeran orogen: ard motion of the Caribbean plate converging with about 10–70 mm/year northeasterly motion of Cocos etionary orogens orogen; two-sided ocean–ocean convergence ras; overall transpressional orogen orogen (between 20� and 25� N latitudes) with present 25 mm/year westward motion of an/Reforma opposed by about 7 mm northeastward motion of the Cocos plate—Advancing, ens � � Archean, the geological orogenic record becomes increasingly more complex. Thus, as seen “from space,” presently active orogenic belts show “a work that has been in progress” at least since the early Mesozoic but also much earlier and a work that today is in various stages of an ill-defined future completion. All this further underlines the futility of correlating and naming so-called “global orogenic phases.” Instead, the purpose of the regional tectonic analysis of sedi- mentary basins and orogens and their associated foreland folded belts is to better define the character and duration of a complex succession of tectonic and strati- graphic regimes. hern Mexico Orogen (between 25 and 35 N latitudes) with about 40 mm/year westward motion and northwest motion of about 80 mm/year of the Pacific and an overall transtensional setting exico Plateau-uplift that followed an earlier Laramide compressional setting—Retreating, ens ran orogens illera between 25� and 48� N latitudes with WSW-directed motion of about 25 mm/year of North America estern Nevada/California to about 30 mm/year, contrasting with amotion of about 60 mm/year west of the ult, and north of the Mendocino fault a slow, about 2 mm/year, northeast-directed motion. The late ent transtensional Basin and Range province is one aspect of complex plateau-like uplifts of the orogen and its nd (Plate 25.A-1). The transtensional uplift regime followed extensive Laramide and earlier compression. etionary Orogen illera between 48� and about 65� N latitudes with west-south westerly motion of North America ranging from year in the south to about 20 mm/year in the north opposing a northwesterly motion of the Pacific plate of ar. Advancing, Accretionary Orogen n Yukon Orogen (Between Gulf of Alaska and Arctic Ocean) with 20 mm/year southwesterly motion of the Ocean and a northwesterly motion of the Pacific plate toward the Aleutian arc ranging from about 50 mm/ t to, say, 120 mm/year in the west. Advancing, Accretionary Orogen in the East and Extending, Accretionary est with its eastern subduction boundary and its complex, northern and southern transpressional boundaries be “bypassed” to the north and the south by the North and South American plates, respectively, which both te of about 25 mm/year contrasting with the westward motion of about 11 mm/year of the Caribbean plate. gen may be subdivided into: onal Lesser Antillean Island arc system—Retreating Accretionary Orogen ional Southern Caribbean (Venezuela/Colombia)—Advancing/Retreating, Accretionary Orogen ranspressional St. Croix to Jamaica system—Advancing/Retreating, Accretionary Orogen nspressional system—Advancing/Retreating, Accretionary Orogen References 25.3. Neotectonics: Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps earthquakes and rigid versus diffuse plate boundaries Bally, A.W., Snelson, S., 1980. Realms of subsidence. In: Miall, A.D. (Ed.), Facts and Principles of World Petroleum occurrence. Can. Petrol. Geol. Mem.6, pp. 6–94. Barrier, E., Chamot-Rooke, N., Giordano, G., 2004. Geodynamic Map of the Mediterranean 1:13 000 000 in 2 Sheets; J.P. Cadet and R. Funiciello (Coordinators). Sheet 1. Tectonics and Kinematics, Sheet 2. Seismicity and Tectonics. CGMW (Commission of the Geological Map of the World) UNESCO, Paris. Becker, A., 1992. An attempt to define a “neotectonic period” for central and northern Europe. Int. J. Earth Sci. (Geol. Rundsch.) 81, 67–83. Bird, P., 2003. An updated digital model of plate boundaries. Geochem. Geophys. Geosyst. 4, 3. An Electronic Journal of the Geosciences 102. Cawood, P.A., Kro¨ner, A., Collins,W.J., Kusky, T.M.,Mooney,W.D.,Windley, B.F., 2009. Accretionary Orogens Through Earth History. Geological Society London Special Publications, 318, pp. 1–36. Chamot-Rooke, N., Rabaute, A., 2006. Plate Tectonics from Space. 1:50 000 000. CGMW (Com- mission of the Geological Map of the World) UNESCO, Paris. Cox, A. (Ed.), 1973. Plate Tectonics and Geomagnetic Reversals. San Francisco Ca. W. H. Freeman and Co., 702 pp. DeMets, C., Gordon, R.G., Argus, D.F., Stein, S., 1990. Current plate motions. Geophys. J. Int. 101, 425–478. DeMets, C., Gordon, R.G., Argus, D.F., Stein, S., 1994. Effects of recent revisions of the geomag- netic reversal scale on estimates of current motions. Geophys. Res. Lett. 21, 2191–2194. DeMets, C., Gordon, R.G., Argus, D.F., 2010. Geologically current plate motions. Geophy. J. Int. 181, 1–80. doi:10.1111/j.1365–246X.2009.04491.x; see also: http://www.geol.wisc.edu/ chuck/MORVEL/motionframe_mrvl.html. Dramis, F., Tondi, E., 2005. Neotectonics. In: Koster, E.A. (Ed.), The Physical Geography ofWestern Europe. Oxford University Press, pp. 25–37. Gordon, R.G., 1998. The plate tectonics approximation: plate non rigidity, diffuse plate and global plate reconstructions boundaries. Ann. Rev. Earth Planet. Sci. 26, 615–642. Gordon, R.G., 2000. Diffuse Plate boundaries: strain rates, vertically averaged rheology and com- parisons with narrow plate boundaries and stable plate interiors. In: Richards, M.A., Gordon, R.G., van der Hilst, R.D. (Eds.), The History and Dynamics of Global Plate Motions. Geophys. Mon.121. American Geophysical Union, Washington DC, pp. 143–159. Gordon, R.S., Stein, S., 1992. Global tectonics and space geodesy. Science 256, 333–342. Gripp, A.E., Gordon, R.G., 2002. Young tracks of hotspots and current plate velocities. Geophys. J. Int. 150, 321–361. Haghipour, A., 1992. Seismotectonic Map of the Middle East. 1:5 000 000. 1 Sheet (in English) Geological Survey of Iran. CGMW (Commission of the Geological Map of the World) UNESCO, Paris. Haghipour, A., 2001. Seismotectonic Map of the World at 1:25 000 000 in 3 Sheets. CGMW (Commission of the Geological Map of the World) UNESCO, Paris. Haghipour, A., 2002. SeismotectonicMap of theWorld at 1:50 000 000 in 1 Sheet. CGMW (Com- mission of the Geological Map of the World) UNESCO, Paris. Haghipour, A., 2006. Structural and Kinematic Map of theWorld at 1:50 000 000 in 1 Sheet (digi- tal version in high resolution format). CGMW (Commission of the Geological Map of the World) UNESCO, Paris. Hallam, A., 1973. A Revolution in the Earth Sciences: From Continental Drift to Plate Tectonics. Clarendon Press, Oxford. 127 pp. IRIS (Incorporated Research Institutions for Seismology), www.iris.org. 997 Isacks, B.L., Oliver, J., Sykes, L.R., 1968. Seismicity and the new global tectonics. J. Geophys. Res. 73, 5855–5899. Also Tectonophysics, 7, 527–541. 998 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps 2255..44 Global stress maps and paleostress studies The scale of our diagrammatic maps is too small to permit an informative and readable graphic reduction of the relevant stress maps. Consequently, a global stress map has not been included in this map series. However, because of the importance of stress maps in the context of regional tectonic studies, selected recent publications are briefly reviewed below. Zoback and Zoback (1989, 1991) and Zoback (1992) pioneered modern stress maps. Heidbach et al. (2004, 2007a, 2007b, 2008) and Sperner et al. (2008) later published and commented on a global stress map at scale of 1:46,000,000. Their up-to-date global stress map displays maximum horizontal compressional stress directions differentiated and based on earthquake focal mechanisms, breakouts, drill-induced fracturing, overcoring, hydro-fracturing, and geological indicators. The data are also quality ranked. Borehole breakouts are a major contributor for the upper 6 km of the crust. Below 6 km, earthquake focal mechanisms take over Kreemer, C., Holt, W.E., Haines, A.J., 2003. An integrated global model of present-day motions and plate boundary deformation. Geophys. J. Int. 145, 8–34. Lee, W.K., Kanamori, H., Jennings, P.C., 2002. Earthquake and Engineering Seismology. Interna- tional Handbook of Earthquake and Engineering Seismology. Parts A and B. Academic Press, London, 1973 pp. McKenzie, D.P., 1977. Plate tectonics and its relationship to the evolution of ideas in the geologic sciences. Daedalus 106, 97–124. McKenzie, D.P., Parker, R.L., 1967. The North Pacific: an example of tectonics on a sphere. Nature 216, 1276–1280. Mu¨ller, R.D., Sdrolias, M., Gaina, C., Roest, W.R., 2008. Age, spreading rates and spreading sym- metry of the worlds ocean crust. Geoche. Geophys. Geosyst. 9, Qo4006. doi: 10.10290/ 2007GC001713. Oreskes, N., 2003. Plate Tectonics: An insider’s History of the Modern Theory of the Earth. West- view Press, Boulder, CO, 255 pp. Stein, S., Sella, G.F., 2002. Plate boundary zones: concept and approaches. In: Stein, S., Freymu¨ller, J.T. (Eds.), Plate Boundary Zones.Geodynamics Series, vol. 30. AmericanGeophys- ical Union, Washington D.C., pp. 1–26. Stein, S., Wysession, M., 2010. Seismology and plate tectonics. In: Stein, S., Wysession, M. (Eds.), An Introduction to Seismology, Earthquakes and Earth Structure. Blackwell Publishing, Oxford, pp. 286–368, Chapter 5. Sykes, L.R., 1966. Seismicity and the deep structure of island arcs. J. Geophys. Res. 72, 2981–3006. U.S. Geological Survey, 2008. http://earthquake.usgs.gov/research/structure/crust/index.ph. (about 76% of the data). Following Zoback and Zoback (1989, 1991) and Zoback Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps (1992), Heidbach et al. (2007b), in a review of their world stress map, differentiate plate-wide/regional stress patterns as follows: 1st-order stress patterns (scales larger than 500 km) attributed to plate boundary forces, that is, “ridge push, slab pull, trench suction, collisional forces and traction at the base of the lithosphere.” 2nd-order stress patterns (100-m to 500-km scales) due to “lateral density con- trasts, associated with continental rifting, isostatic compensation and topography, de-glaciation effects and lithospheric flexure.” 3rd-order stress patterns (scales < 100 km) due to “active faults, seismically induced stress changes associated with large earthquakes, local density contrasts (e.g., salt diapirs and detachment horizons).” In practice, the differentiation of the three scale-related orders is likely to involve “common sense ”judgments guided by regional, local studies as well as paleo- stress studies. In the context of the evaluation of conventional as well as uncon- ventional hydrocarbon reservoirs, it is very important to understand present stress distribution (e.g., Tingay et al., 2005). Zoback and Zoback (2007) in their concise review of lithosphere stress and defor- mation conclude as follows: (1) The “remarkably uniform” present day stress distribution over large regions of the lithosphere is not related to past residual stresses associatedwith earlier tectonics. (2) Forces acting on plate margins are largely responsible for the stress within plates. (3) Stress measurements in deep wells and scientific research boreholes (up to 8.1 km) indicate that stress magnitudes within the crust are controlled by frictional strength. (4) The crust is in a state of “incipient frictional failure”and regional rates of deformation are controlled by the overall strength of the lithosphere. Thus, young hot lithosphere on active plate boundaries deforms more rapidly than the colder and stronger lithosphere of mid-plate regions. (5) Most of the stress is transmitted through the strong upper brittle crustal layer. Paleostress studies provide a perspective on past stress distributions and their relation to present stress studies. Lacombe (2007) noted the widely different methodologies used by the “contemporary stress community” and the “paleo- stress community,” respectively. Perhaps mildly exaggerating, he suggests that people working on present-day stress “don’t seem to speak the same language or deal with similar mechanical concepts” as the paleostress community. Paleostress methodology initially involved microfault slip measurements and basic rock mechanics as pioneered by Angelier (1984, 1989), Bergerat et al. (1985), and 999 Bergerat (1987) and reviewed by Lisle et al. (2006). In addition “paleopiezo- 1000 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps metry” reviewed by Burkhard (1993), Ferrill et al. 2003 and Lacombe (2007) involves measuring “the dislocation density of calcite, the dynamic recrystallisa- tion of calcite and the mechanical twinning of calcite dolomite.” In particular, the calcite twinning process has been experimentally calibrated and therefore permits estimation of the orders of magnitude of shear stress twinning. Lisle et al. (2006) reviewed earlier paleostress studies and concluded that the assumption that one of the stress axes is vertical is generally confirmed. How- ever, the steepest axis deviates by more than 25% from the vertical. Lisle et al. (2006) cite a number of reasons for the deviation including the possibility that the vertical axis assumption may not be valid at greater depths below the surface, effects due to topography, and the influence of nearby larger scale faults, in other words, many of the factors that Heidbach et al. (2007b) include in their 3rd Class. According to Lacombe (2007), an important issue involves “the difficulty of estab- lishing paleostress/paleodepth relationships.” In drill holes, present-day stresses are measured at a known depth. On the other hand, most paleostress studies involve surface observations and measurements on rocks that have been first buried and then exhumed. An evaluation of the paleodepth can be made from the degree of metamorphism of rocks, whereas shallow metamorphism can be estimated from the maturity of hydrocarbon source beds thought to be mainly linked to temperature and time. Paleodepth estimates of buried sediment are at best a useful approximation. Applying paleostress analysis to the evolution of sedimentary basins is particularly useful. Kleinspehn et al. (1989), using the example of Spitsbergen, trace the impact of alternating stratigraphically dated transpressional and transtensional regimes. Surprisingly, more systematic studies that assign paleostress regimes to tectono- stratigraphic megasequences in continental interior basins have yet to be under- taken. In particular, regional/supraregional unconformities (e.g., second order unconformities with hiatuses ranging from say 10 to over 80 Ma) separating stratigraphic sequences (Mitchum et al. 1977; Sloss, 1963; 1988a,b) are likely to be directly related to corresponding changes in regional and continent-wide paleostress regimes. The episodic tectonic evolution of sedimentary basins is also a response to varying paleostress regimes represented by successive tectonostratigraphic megase- quences that are bounded by supraregional unconformities (see Chapter 4, Vol. 1A (Roberts and Bally)). These unconformities are best displayed on long, if possible, latest vintage, regional seismic profiles calibrated by cross-sections of correlated wells. Regional seismic reflection profiles may also be supplemented by more local 3D seismic surveys as well as outcrop studies. In addition, supra- regional subcrop maps of major unconformities also reveal important clues to the structural regime that preceded the termination of a tectono-stratigraphic Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps megasequence (Bally, 1989; Cook and Bally, 1975; Levorsen, 1961). Most regional and supraregional unconformities branch out in a downdip direc- tion into wedges of multiple unconformities. These eventually may merge into conformable correlative stratigraphic boundaries. Changing paleostress regime boundaries are best determined within the wedges but the large hiatus of the merged unconformities will in general prevent accurate tectonostratigraphic dat- ing of changing paleostress regimes. To conclude, present-day/contemporary stress measurements and maps are useful in the context of earthquake studies and also in the context of hydrocar- bon exploration for unconventional plays that involve fractured hydrocarbon- rich shales and also coal beds. The relationship of the present day stress distribution to paleostress distribution is complementary and can be considered as an important adjunct to regional studies. Independent analysis of shallow near surface data from industry reflection seismic profiles complemented with shallow high frequency surveys where justified should be used as indepen- dent adjunct to recent stress studies. In the end, if we are going to build a bridge from the recent into the past, it will be important to provide the best possible data to link present stress distributions to paleostress systems. Given the uniformity of present-day regional stress distributions, it would eventually be desirable to compare global paleostress maps for plausibly defined global plate tectonic regimes with plate tectonic reconstructions of continents for these intervals. References 25.4. StressMaps and Paleostress Studies Angelier, J., 1984. Tectonic analyses of fault slip data sets. Fault behaviour and the earthquake gen- eration process. J. Geophys. Res. B 89, American Geophysical Union, Washington D.C., pp. 4835–5848. Angelier, J., 1989. From orientation to magnitudes in paleostress determinations using fault slip- data. J. Struct. Geol. 11, 37–50. Bally, A.W., 1989. Phanerozoic basins of North America. In: Bally, A.W., Palmer, E.R. (Eds.), The Geology of North America—An Overview. Geological Society of America, Boulder, CO. The Geology of North America, V.A., pp. 397–446. Bergerat, F., 1987. Stress fields in the European Platform at the time of Africa-Eurasia collision. Tectonics 6, 99–132. Bergerat, F., Bergues, J., Geyssant, J., 1985. Estimationsdespale´ocontraintes lie´es a` la formationde de´crochements dans la plateforme d’Europe du Nord. Geol. Rundsch. 74, 311–320. Burkhard, M., 1993. Calcite twins, their geometry, appearance and significance as stress-strain markers and indicators of tectonic regime: a review. J. Struct. Geol. 15, 351–368. Cook, T., Bally, A.W., 1975. Stratigraphic Atlas of North and Central America Princeton. Princeton University Press, New Jersey, 272 pp. Ferrill, D., Morris, A.P., Evans, M.A., Burkhardt, M., Groshong Jr., R.G., Onasch, C.M., 2003. Cal- cite twin morphology: a low-temperature deformation geothermometer. Struct. Geol. 1521–1529. Elsevier. 1001 Heidbach, O., Barth, A., Connolly, P., Fuchs, K., Mu¨ller, B., Reinecker, J., et al., 2004. Stressmaps in a minute: the 2004 Stress Map Release. EOS. Trans 85 (49), 521–529. http://www.world -stress.org. Heidbach, O., Fuchs, K., Mu¨ller, B., Reinecker, J., Sperner, B., Tingay, M., et al., 2007a. World 1002 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Stress Map 1:46 000 000, incl. CD-ROM with Digital Maps & Notes. CGMW (Commission of the Geological Map of the World) and Heidelberg Academy of Sciences and Humanities/UNESCO, Paris. Heidbach, O., Reinecke, J., Tingay, M., Mu¨ller, B., Sperner, B., Fuchs, K., et al., 2007b. Plate boundary forces are not enough: second and third order stress patterns highlighted in the World Stress Map data base. Tectonics 26, doi:10.1029/2007TC002133. Heidbach, O., Tingay,M., Barth, A., Reinecker, J., Kurfess, D.,Mu¨ller, B., 2008. (Available on line at www.world-stress-map-.org). Kleinspehn, K.L., Pershing, J., Teyssier, C., 1989. Paleostress stratigraphy: a new technique for ana- lyzing tectonic control on sedimentary basin subsidence. Geology 17, 253–256. Mitchum Jr., R.M., Vail, P.R., Sangree, J.B., 1977. Seismic stratigraphy and global changes of sea level, part 6. In: Payton, C.E. (Ed.), Seismic Stratigraphy—Applications to Hydrocarbon Explo- ration, American Association of Petroleum Geologists Mem. 26, pp. 117–134. Lacombe, O., 2007. Comparison of paleostress magnitude from calcite twins with contemporary stress magnitudes and frictional sliding criteria in the continental crust: mechanical implica- tions. J. Struct. Geol. 29, 86–99. Lisle, R.J., Orife, T.O., Arlegui, I., Liesa, C., Srivastava, D.C., 2006. Favoured states of paleostress in the earth’s crust evidence from fault slip data. J. Struct. Geol. 28, 1051–1066. Levorsen, A.I., 1961. PaleogeologicMaps.W.H. Freeman andCo. San Francisco and London. 174. Sloss, L.L., 1963. Sequences in the cratonic interior of North America. Geol. Soc. Am. Bull. 74, 93–114. Sloss, L.L., 1988a. Introduction. In: Sloss, L.L. (Ed.), Sedimentary Cover—North American Craton, U.S. The Geology of North America, Geological Society of America, Boulder, CO, D.-2, pp. 1–3. Sloss, L.L., 1988b. Tectonic evolution of the craton in Phanerozoic time. In: Sloss, L.L. (Ed.), Sedi- mentary Cover-North American Craton, U.S. The Geology of North America, Geological Soci- ety of America, Boulder, CO, D.-2, pp. 25–51. Sperner, B., Mu¨ller, B., Hrach, O., Delvaux, D., Reinecker, L., Fuchs, K., 2008. Tectonic stress in the earth’s crust: advances in the world stress map project. In: Nieuwland, D.A. (Ed.), New Insights in Structural Interpretation and Modeling, Geological Society London Special Publi- cation, 212, pp. 101–116. Tingay, M., Mu¨ller, B., Reinecker, J., Heidbach, O., Wenzel, F., Fleckenstein, P., 2005. The world stress map project “Present-day Stress in Sedimentary basins” Initiative; building a valuable resource to understand tectonic stress in the oil patch. Lead. Edge 24(12), 1276–1282. Zoback, M.L., 1992. First and second order patterns of stress in the lithosphere: the World Stress Map Project. J. Geophys. Res. 97, 11703–11728. Zoback, M.L., Zoback, M.D., 1989. Tectonic stress field of the conterminous Unites States. In: Pakiser, L.C., Mooney, W.D. (Eds.), Geophysical Framework of the Continental United States. Memoir, 172, Geological Society of America, Boulder, CO, pp. 523–539. Zoback, M.D., Zoback, M.L., 1991. Tectonic stress field of North America and relative plate motions. In: Slemmons, B.D., Engdahl, E.R., Zoback, M.D., Blackwell, D.D. (Eds.), Neotec- tonics of North America. Geological Society of America, Boulder. CO, pp. 339–366. Zoback, M.L., Zoback, M.D., 2007. 6.06 Lithosphere stress and deformation. In: Watts, A.B. (Ed.), Crust and Lithosphere Dynamics Vol. 6 of G.R. Schubert (Ed.), Treatise of Geophysics. Elsevier, Amsterdam, pp. 1–48. Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps 2255..55 The continental lithosphere and continental crust (Plates 25.12 and 25.13) Introduction Plate tectonics and its relation to regional andmore detailed regional studies are the justification for including small-scale continental lithosphereandcrust thicknessmaps (Plates 25.12 and 25.13). Fowler (2005), Romanowicz and Dziewonsky (2007), Mooney (2007) and Watts (2007) have published recent overviews of geophysical studies of the lithosphere and the crust. Geochemical and petrological perspectives on the crust and the lithosphere are well summarized by Rudnick and Gao (2005). The followingverybrief review ismostly basedon these references. For varyingdefini- tions of lithospheres see Anderson’s (2007) segment on “The many lithospheres.” Much of the Cenozoic/Mesozoic oceanic crust has already been subducted and only minor fragments of that crust have returned to the surface in the sutures of orogenic zones (see Chapter 27, Vol. 1A (Nicholas)). Consequently, the presently preserved oceanic crust and lithosphere constitute the repository of the remaining Cenozoic/Mesozoic ocean spreading record. About 90% of the corresponding oceanic lithosphere is mantle. In contrast, about 60% of the continental litho- sphere is mantle although the age of the mantle does not necessarily correspond to the perceived age of the overlying continental crust. This contrast between continents and oceans becomes even more striking with the realization that, over < 4 Ga, the continental crust has become the residual repository of a very complex process of continental growth and its thermo-mechanical reworking. Over time, additional recycling has also involved erosion and deposition of sediments, the formation of igneous and metamorphic rocks that eventually might become subducted. Finally, igneous, metamorphic, and sedimentary rocks can always be re-worked in the context of various thermo-mechanical processes that affect both the crust and the mantle and lead to a re-setting of the age of the Moho. Different geophysical methodologies are applied to decipher the structure of the lithosphere and the crust and include the following: (1) Passive source (i.e., earthquake) seismology (for anoverview seeRomanowicz and Dziewonsky, 2007). (2) Active explosion source seismology (for an overview see Levander et al., 2007; Mooney, 2007; and see Chapter 11, Vol. 1A (White)). The wide- ranging scale of these studies, ranging from sublithospheric scales to, say, tens of meters of the near surface allows for mapping the subsurface in vary- ing detail. When coupledwith geochemical/petrologic and stratigraphic and structural geological studies, seismic reflection profiles, at varying scales of 1003 boundaries and with added control points in a tightly documented paper, Arte- 1004 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps mieva (2006)presenteda thermal lithosphere thicknessmodelof all thecontinents. Theprincipal geological base for the earliermap aswell as thenewglobalmodel is a detailed compilation of crustal ages based on Goodwin (1996) and many other sources. In addition, geotherms derived frommantle xenolith studies are also con- sidered. The result shows a strong correlation between the age of the crust and the thermal state of the lithosphere that allows construction of a global model of the thermal lithosphere. The global continental lithospheric thickness (Artemieva, 2006; Fig. 2, and see Chapter 5, Vol 1A (Artemieva)), in general and with few exceptions, mimics the distribution of surface and near-surface crustal ages (Plate 25.20). Continental lithosphere ranges between say about 100 and 350 km. Combining a seismic approach for the continents and a thermal/crustal age-related approach for the oceans, Conrad and Lithgow-Bertelloni (2006) present a world lithospheric thickness map that is less detailed than the maps of Artemieva (2006) but also includes an oceanic lithosphere with thicknesses ranging from close to resolution, are particularly important for an understanding of the regional geological settings in their plate tectonic context. (3) Lithosphere/plate rheology and mechanics (e.g., Burov, 2007) (4) Plate deformation, ranging from postglacial rebound to coseismic and post- seismic deformation, to intraplate deformation, and to plate deformation associatedwith rifted areas as well as subduction zones (e.g., Sabadini, 2008) (5) Heat flow and thermal structure of the lithosphere (for an overview see Jaupart and Mareschal, 2007). When modeling the origin and evolution of sedimentary basins as well as the maturation process of hydrocarbon source beds, it is particularly important to understand their thermal evolution from the inception of burial all the way through possible late uplift stages. Our main interest here lies in folded belts and sedimentary basins that are located on continents and their active and passive margins. Therefore, our maps and the following comments focus on the continental lithosphere and crust. The continental lithosphere (Plate 25.12) For various definitions of lithospheres, see Anderson (2007, p. 37). A seismically defined “elastic” lithosphere ranges in thickness from about 300 to 95 km under continents and 95 to 4 km under the oceans. The base of the lithosphere is com- monly defined by a thermal boundary layer. With about 90% of its total thickness being mantle, typical oceanic lithosphere contrasts with, say, 60% of continental mantle of the continental lithosphere. Plate 25.20 is a re-drafted version of the thermal continental lithosphere thickness map of Artemieva andMooney (2001, 2002). Themap shows only the continental thermal lithosphere thickness in areas that have extensive heat flow coverage, that is, about 40% of the total continental area. Going well beyond these data-limited Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps zero at mid-ocean ridges to about 110 km along continent–ocean boundaries. On Gondwana-derived passivemargins, there is a striking contrast between overall sub- parallel oceanic lithosphere isopachs and discordantly truncated continental lithosphere isopachs. This simply reflects that new oceans often open discordantly to the structures produced by the closing of earlier oceans. In detail, uncertainties about the relatively narrow width of the continent–ocean crustal/lithospheric transition are related to the transitional nature of rifted/hyperextended or else volcanic/underplated passive margins. The continental crust (Plate 25.13) An in-depth review of earlier crustal studies was prepared by Meissner (1986). A comprehensive updated review was edited by Rudnick (2005). The following reviews some salient points selected from Mooney’s (2007) more recent overview of global crustal structure. The Moho (Mohorovicic) discontinuity sepa- rates silicic from ultramafic rocks and separates a lower crustal layer with p-wave velocities ranging between 6.6 and 7.6 km/s from the upper lithospheric mantle with velocities around 8.0–8.2 km/s. By area, about 59% of the crust is oceanic and 41% is continental. Submarine passive and active continental margins across their continental slopes often show an abrupt transition from about 30-km-thick continental crust to oceanic crust about 10 km in thickness. On our global maps, this transition is shown by the symbols for passive margins and the oceanic sub- duction boundaries, respectively (see Plate 25.14 legend). Taking the slope break of continental margins as a proxy for the continent–ocean boundary, it is esti- mated that about 31% of the continental area is below shallow oceans. Rudnick and Gao (2005) estimate that by volume only about 5% of the upper 16 km of the crust consists of sedimentary rocks. Yet, these sedimentary rocks host the bulk of our vital fresh water, energy, and fertilizer resources. This may explain the dispro- portionate anthropocentric attention given to Phanerozoic sediments, their stra- tigraphy, and their involvement in structural deformation. The oldest preserved oceanic crust has a Jurassic age of about 190 Ma that contrasts with amuch longer but, going back in time, increasingly rudimentary and fragmen- tary Paleozoic to Precambrian record (from 220 to over 4000 Ma). On average, the oceaniccrust is�6–7 kmthick.Fast-spreadingmid-oceanridgesareabout9 kmthick. Thecrustal thicknessof volcanicplateaus intheoceans rangesbetween10and30 km. This is in contrast to continental crust that ranges from 16 to 25 km in the rifted Afar triangleof E. Africa to theover 75-km-thick crust of the Tibetanplateau.The continen- tal crust includes the crystalline basement, the overlying sediments, and even ice. On the basis of seismic velocities, the continental crust is subdivided into upper, middle, and lower crust. Sediments are included in the upper crust that has com- pressional wave (p-wave) velocities that range widely between say 2 and 4 km/s for clastics and up to about 6 km/s for carbonates. The crystalline upper continen- tal crust is characterized by P-wave velocities ranging between 5.6 and 6.3 km/s that are associated with granitic and metasedimentary rocks. With increased 1005 1006 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps depths between 10 and 15 km, there is a transition to the middle crust with vel- ocities ranging between 6.4 and 6.7 km/s that are commonly associated with intermediate composition plutonic rocks and amphibolite facies metamorphic rocks. The lowermost crust has P-wave velocities ranging between 6.8 and 7.3 km/s, considered to be associated with granulite facies metamorphic rocks (Mooney, 2007). The Moho separates the crust from an underlying lithospheric mantle with velocities ranging from 8.0 to 8.2 km/s associated with peridotites or eclogites. Hammer et al. (2011) assembled several crustal reflection seismic pro- files into a Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps ments, crystalline upper, middle, and lower crust, and the uppermost mantle). CRUST 2 (Laske et al., 2009) is a later, improved model that is also available on the web. This newer model provides better resolution of continental margins and also of oceanic plateaus. By combining a crustal thickness map (Plate 25.13) with a generalized basement map, it is obvious that the continental crust in general thins from say 30-km to about 10-km continental crust on active and passive continental margins. The ocean–continent boundaries straddle the transi- tion between a 30-km continental crust and a 10-km oceanic crust. Larger continental and regional crustal thickness maps provide greater resolution than our small-scale global maps and show the relationships between tectonic provinces and crustal thickness. Examples include the crustal thickness map of North America (Mooney and Braile, 1998, updated by Chulik and Mooney, 2002), and Moho depth maps of Australia (Collins et al., 2003) and the European plate (Grad et al., 2008). For Western and Central Europe, Ziegler and De´zes (2006) did not apply corrections for the thickness of the sedimentary cover or sur- face topography. Therefore, their map is not a true continental thickness map and should not be compared with Plate 25.20. Even so, their tectonic analysis shows that the Proterozoic European craton and the Paleozoic Caledonian and Variscan (Hercynian) domains were overprinted by the following features: (a) Devonian and Permian wrench tectonics and rifting, (b) Arctic/North Atlantic Tethys-related rifting, (c) widespread inversion of these rifts associated with far- field impact of the Alpine collision, (d) Late Oligocene and Neogene rifting of the Rhine-Rhone systems and back-arc extension in the Mediterranean, and (e) deep crustal roots beyond 50 km depth associated with cores of the Alps, Pyrenees, and the Northern Apennines. The example of Western and Central Europe shows how crustal thickness is mod- ified over geological time and how the crust mantle boundary is periodically reset during a succession of changing tectonic regimes as noted in the exhumed lower crust of the Southern Alps, as summarized in Section “Allochthonous Fragments, Oceanic and Intra-oceanic Arc Systems, and Lower Crust and Uppermost Mantle of Hyper-Extended Passive Margins.” Crustal layers, rheological models, and conclusions Crustal and upper mantle layers that are based on seismic refraction and tomo- graphic studies are commonly accepted. Crustal reflection profiles often show a For this map series, the crustal thickness map (USGS, 2010) was redrafted and combinedwith a schematicmap of both outcropping and buried orogens. Amore detailed “basement map” is shown on Plate 25.19. The continental basement iso- pachs are from the CRUST 5.1 model of Mooney et al. (1998). The CRUST 5.1 model is based on over 560 refraction seismic measurements and the model includes an eight-layer subdivision (i.e., ice, water, soft sediments, hard sedi- 1007 Burov, E.B., 2007. Plate rheology andmechanics. In:Watts, A.B. (Ed.), Crust and LithosphereDynam- 1008 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps ics. Vol.6. of G.R. Schubert (Ed.), Treatise of Geophysics. Elsevier, Amsterdam, pp. 99–152. Christensen, N.I., Mooney, W.D., 1995. Seismic velocity structure and composition of the conti- nental crust: a global review. J. Geophys. Res. 100, 9761–9788. Collins, C.D.N., Drummond, B.J., Nicoll, M.G., 2003. Crustal thickness patterns in the Australian continent. In: Hillis, R.H.,Mu¨ller, R.D. (Eds.), Evolution of the Australian Plate. Geological Soci- ety of America Special Paper 372, pp. 121–128. Conrad, C.P., Lithgow-Bertelloni, C., 2006. Influence of continental roots and asthenosphere on plate mantle coupling. Geophys. Res. Lett. 33, L05312. doi:10.1029/2005GLO25621. Chulik, G.S.,Mooney,W.D., 2002. Seismic structure of the crust and theuppermostmantle ofNorth America and adjacent Ocean basin: a synthesis. Bull. Seismol. Soc. Am. 92, 2478–24592. Dohr, G., Lukic, P., Bachmann, C.H., 1983. Deep crustal reflections in the Northwest-German basin. In: Bally, A.W. (Ed.), Seismic expression of Structural Styles. AAPG Studies in Geology # 15. V. 1. American Association of Petroleum Geologists, 1.5-1 to 1.5-5. Dohr, G., Meissner, R., 1975. Deep crustal reflections in Europe. Geophysics 40, 25–39. Emmermann, R., Lauterjung, J., 1997. The german continental deep drilling program KTB. J. Geo- phys. Res. 102, 18197–18201. Fowler, C.M.R., 2005. The Solid Earth: An Introduction to Global Geophysics, Second ed. Cambridge University Press, Cambridge, U.K. 685 pp. Goodwin,A.M., 1996. Principlesof PrecambrianGeology. Academic Press Ltd, LondonU.K., 328pp. layered/laminated lower crust as well as dipping reflectors that are often inter- preted as faults, ductile shears, sills, etc. Rheological models need to be reconciled with these images. Ranalli (1997), Burov (2007) and a review by Jackson et al. (2008) all discuss the distribution of ductile and brittle layers in the crust and the uppermantle based onmineralogy, temperature and pressure gradients, pres- ence or absence of fluids, and other factors that influence rheological models. This very limited review shows that lithospheric and crustal rheological studies have an important impact on the evolution of a given regional setting. To fully understand this setting, it is important to realize that problems, questions, and models are often constrained by “outside the box” considerations. Conversely, geophysical and geochemical models need to be on target when dealing “inside the box” with the specific thermo-tectonic evolution of polyphase sedimentary basins or folded belts of all ages. Much of this involves a dialogue among spe- cialists that transcends the specific language/jargon of each group. Hence, the need for acceptable simplifications arises. References: 25.5 The Continental Lithosphere and The Continental Crust Anderson, D.L., 2007. The many lithospheres. In: Anderson, D.L. (Ed.), New Theory of the Earth. Cambridge University Press, pp. 36–38. Chapter 4, The outer shells of Earth. Artemieva, I.M., 2006. Global 1� � 1� thermal model TC1 for the continental lithosphere: impli- cations for the lithosphere secular evolution. Tectonophysics 416, 245–277. Artemieva, I.M.,Mooney,W.D., 2001. Thermal evolution of the Precambrian lithosphere: a global study. J. Geophys. Res. 106, 16387–16414. Artemieva, I.M., Mooney, W.D., 2002. On the relations between cratonic lithosphere thickness, plate motions and basal drag. Tectonophysics 358, 211–231. Bassin, C., Laske, G., Masters, G., 2000. The current limits of resolution of surface wave tomogra- phy in North America. EOS Trans. AGU 81, F897. Grad, M., Tira, T., ESC Working Group, 2008. The Moho depth map of the European Plate. Geo- Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps phys. J. Int. 176, 279–292. Hammer, P.T.C., Clowes, R.M., Cook, F.A., Vadsudevan, K., van der Velden, A.J., 2011. The big picture: a lithospheric cross section of the North American continent. GSA Today 1, 4–10. Jackson, J., McKenzie, D.P., Priestley, K., Emmerson, B., 2008. New views on the structure of the lithosphere. J. Geol. Soc. 165(2), 453–466. Jaupart, C., Mareschal, J.C., 2005. Constraints on crustal heat production from heart flow data. In: Rudnick, R.L. (Ed.), The Crust. Elsevier, Amsterdam, pp. 65–121. Jaupart, C., Mareschal, J.C., 2008. Heat flow and thermal structure of the lithosphere. In: Watts, A.B. (Ed.), Crust and Lithosphere Dynamics. Vol.6. of G.R. Schubert (Ed.), Treatise of Geophysics. Elsevier, Amsterdam, pp. 217–252. Juhlin, C., 1988. Interpretation of the seismic reflectors in the Graveberg-1 well. In: Boden, A., Erickson, K.G. (Eds.), Deep Drilling in Crystalline Bedrock. Springer, Berlin, 364 pp. Kozlovsky, Y.A. (Ed.), 1987. The Superwell of the Kola Peninsula (Exploration of the Deep Conti- nental Crust). Springer Verlag, Berlin, 558 pp. Laske, G., Masters, G., Reif, C., 2009. Crust 2: A New Global Crustal Model at 2�2 Degrees. http://igppweb.ucsd.edu/-gabi/crust2.html. Levander, A., Zelt, C.A., Symes, W.W., 2007. Crust and Lithosphere studies. In: Romanowicz, B., Dziewonsky, A. (Eds.), Seismology and the Structure of the Earth. of G.R. Schubert (Ed.), Trea- tise of Geophysics, vol. 1. Elsevier, Amsterdam, pp. 247–288. Meissner, R., 1986. The Continental Crust: A Geophysical Approach. Academic Press, San Diego, CA, 426 pp. Mooney, W.D., 2007. Crust and lithospheric structure-Global crustal structure. In: Romanowicz, B., Dziewonsky, A. (Eds.), Seismology and structure of the Earth, Vol.1. Treatise of Geophysics, Elsevier, Amsterdam, NL, pp. 361–471. Mooney, W.D., Braile, L.W., 1998. The seismic structure of the continental crust and uppermantle of North America. In: Bally, A.W., Palmer, A.R. (Eds.), The Geology of North America. Geologi- cal Society of America, Boulder, CO, pp. 39–52. Mooney, W.D., Laske, G., Masters, T.G., 1998. Crust 5.1: a global crustal model at 5˚-5˚. J. Geo- phys. Res. 103, 727–747. Papasikas, N., Juhlin, C., 1997. Interpretation of reflections from the central part of the Siljan Ring impact structurebasedon the results of the Stenberg1borehole. Tectonophysics 269,237–245. Pollack, H.N., 1982. The heat flow from the continents. Ann. Rev. Earth Plan. Sci. 10, 459–481. Pollack, H.N., Hurter, S.J., Johnson, J.R., 1993. Heat loss from the Earth’s interior; analysis of a global data set. Rev. Geophys. 31, 267–280. Ranalli, G., 1997. Rheology of the Lithosphere in Space and Time. Geol. Soc. Lond. Special Publi- cation 121, pp. 19–37. Romanowicz, B., Dziewonsky, A., 2007. Seismology and the structure of the earth. In: Schubert, G. (Ed.), Treatise of Geophysics, vol. 1. Elsevier, Amsterdam, 858. Rudnick, R.L., 2005. The crust. In: Holland, H.D., Turekian, K.K. (Eds.), Treatise of Geochemistry, vol. 3. Elsevier, Amsterdam, p. 683. Rudnick, R.L., Gao, S., 2005. Composition of the Continental Crust. In: Holland, H.D., Turekian, K.K. (Eds.), Treatise of Geochemistry, vol. 3. Elsevier, Amsterdam, pp. 1–64. Rudnick, R.L., Fountain, D.M., 1995. Nature and composition of the continental crust. Rev. Geo- phys. 33, 267–309. Sabadini, R., 2007. Plate deformation. In: Watts, A.B. (Ed.), Crust and Lithosphere Dynamics. In: G.R. Schubert (Ed.), Treatise of Geophysics, vol. 6. Elsevier, Amsterdam, pp. 153–216. USGS, 2010. Earthquake Hazards Program: The Earth’s Crust: http://earthquake.usgs.gov./ research/structure/crust/index.php. Watts, A.B., 2007. An overview. In: Schubert, G.R. (Ed.), Treatise of Geophysics, vol. 6. Crust and Lithosphere Dynamics. Elsevier, Amsterdam, 1–49. Ziegler, P.A., De`zes, P., 2006. Crustal evolution of Western and Central Europe. In: Gee, D.G., Stephenson, R.A. (Eds.), Geological Society London Memoir 32, 43–56. 1009 PLATES FOR GLOBAL TOPOGRAPHY, NEOTECTONICS, THE CONTINENTAL LITHOSPHERE ANDCRUST: Segments 25.1–25.3 and 25.5 (For online version of the plates/figures cited in this chapter, the reader is referred to http://www. elsevierdirect.com/companion.jsp? ISBN=9780444563576). 1010 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps -180˚ -180˚ -150˚ -150˚ -120˚ -120˚ -90˚ -90˚ -60˚ -60˚ -30˚ -30˚ 0˚ 0˚ 30˚ 30˚ 60˚ 60˚ 90˚ 90˚ 120˚ 120˚ 150˚ 150˚ 180˚ 180˚ -60˚ -60˚ -30˚ -30˚ 0˚ 0˚ 30˚ 30˚ 60˚ 60˚ -180˚ -180˚ -150˚ -150˚ -120˚ -120˚ -90˚ -90˚ -60˚ -60˚ -30˚ -30˚ 0˚ 0˚ 30˚ 30˚ 60˚ 60˚ 90˚ 90˚ 120˚ 120˚ 150˚ 150˚ 180˚ 180˚ -60˚ -60˚ -30˚ -30˚ 0˚ 0˚ 30˚ 30˚ 60˚ 60˚ -7000 -6500 -6000 -5500 -5000 -4500 -4000 -3500 -3000 -2500 -2000 -1500 -500 0 200 400 600 1000 1500 3500 7000 Credits: ETOPO1, Amante C. and B.W. Eakins 2009. Data downloaded from: http://www.noaa.gov/mgg/global/global.html 4 may 2010 Plate 25.1 Topography and Bathymetry (excl. Polar Areas). 1 0 1 1 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s 180˚ 150 ˚W 12 0˚ W 9 0 ˚ W 60˚W 30˚W 0˚ 30˚ E 60 ˚E 9 0 ˚ E 120˚E 150˚E 180 150 ˚W 12 0˚ W 9 0 ˚ W 60˚W 30˚W 0˚ 30˚ E 60 ˚E 9 0 ˚ E 120˚E 150˚E Credits: ETOPO1, Amante C. and B.W. Eakin 009. Data downloaded from: http://www.noaa.gov/ g/global/global.html 4 may 2010 Credits: ETOPO1, Amante C. and B.W. Eakins 2009. Data downloaded from: http://www.noaa.gov/mgg/global/global.html 4 may 2010 A Plate 25.2 (A) Arctic Topography and Bathymetry map: (incl. Ice surface) Arctic Topography and Bathym try map: (incl. Bedrock surface below ice). 1 0 1 2 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s ˚ s 2 mg e 180˚ 150˚W 120˚W 9 0 ˚ W 60 ˚W 30˚ W 0˚ 30˚E 60˚E 9 0 ˚ E 12 0˚ E 150 ˚E Credits: ETOPO1, Amante C. and B.W. Eakins 2009. Data downloaded from: http://www.noaa.gov/mgg/global/global.html 4 may 2010 Credits: ETOPO1, Amante C. and B.W. Eakins 2009. Data downloaded from: http://www.noaa.gov/mgg/global/global.html 4 may 2010 180˚ 150˚W 120˚W 9 0 ˚ W 60 ˚W 30˚ W 0˚ 30˚E 60˚E 9 0 ˚ E 12 0˚ E 150 ˚E B Plate 25.2 Cont’d (B) Antarctic Topography and Bathymetry map (incl. Ice surface) Antarctic Topography and Bathymetry map (incl. Bedrock surface below ice). 1 0 1 3 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s -180˚ -180˚ -150˚ -150˚ -120˚ -120˚ -90˚ -90˚ -60˚ -60˚ -30˚ -30˚ 0˚ 0˚ 30˚ 30˚ 60˚ 60˚ 90˚ 90˚ 120˚ 120˚ 150˚ 150˚ 180˚ 180˚ -60˚ -60˚ -30˚ -30˚ 0˚ 0˚ 30˚ 30˚ 60˚ 60˚ -180˚ -180˚ -150˚ -150˚ -120˚ -120˚ -90˚ -90˚ -60˚ -60˚ -30˚ -30˚ 0˚ 0˚ 30˚ 30˚ 60˚ 60˚ 90˚ 90˚ 120˚ 120˚ 150˚ 150˚ 180˚ 180˚ -60˚ -60˚ -30˚ -30˚ 0˚ 0˚ 30˚ 30˚ 60˚ 60˚ Credits: www.iris.org Plate 25.3 Distribution of Shallow-Focus Earthquakes (0-70km). 1 0 1 4 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s -180˚ -180˚ -150˚ -150˚ -120˚ -120˚ -90˚ -90˚ -60˚ -60˚ -30˚ -30˚ 0˚ 0˚ 30˚ 30˚ 60˚ 60˚ 90˚ 90˚ 120˚ 120˚ 150˚ 150˚ 180˚ 180˚ -60˚ -60˚ -30˚ -30˚ 0˚ 0˚ 30˚ 30˚ 60˚ 60˚ -180˚ -180˚ -150˚ -150˚ -120˚ -120˚ -90˚ -90˚ -60˚ -60˚ -30˚ -30˚ 0˚ 0˚ 30˚ 30˚ 60˚ 60˚ 90˚ 90˚ 120˚ 120˚ 150˚ 150˚ 180˚ 180˚ -60˚ -60˚ -30˚ -30˚ 0˚ 0˚ 30˚ 30˚ 60˚ 60˚ Credits: www.iris.org Plate 25.4 Distribution of Deep-Focus Earthquakes (deeper than 70km). 1 0 1 5 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s 180˚ 150 ˚W 12 0˚ W 9 0 ˚ W 60˚W 30˚W 0˚ 30˚ E 60 ˚E 9 0 ˚ E 120˚E 150˚E Credits: www.iris.org Credits: www.iris.org 180˚ 150 ˚W 12 0˚ W 9 0 ˚ W 60˚W 30˚W 0˚ 30˚ E 60 ˚E 9 0 ˚ E 120˚E 150˚E Plate 25.5 Arctic Shallow-Focus Earthquakes (0-70 km), Arctic Deep-Focus Earthquakes (deeper than 70 km). 1 0 1 6 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s 180˚ 150˚W 120˚W 9 0 ˚ W 60 ˚W 30˚ W 0˚ 30˚E 60˚E 9 0 ˚ E 12 0˚ E 150 ˚E 180˚ 150˚W 120˚W 9 0 ˚ W 60 ˚W 30˚ W 0˚ 30˚E 60˚E 9 0 ˚ E 12 0˚ E 150 ˚E Credits: www.iris.org Credits: www.iris.org Plate 25.6 Antarctic Shallow-Focus Earthquakes (0-70 km), Antarctic Deep-Focus Earthquakes (deeper than 70 km). 1 0 1 7 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s Conventional Oceanic Plate Boundaries & Motions: Oceanic Plates Oceanic Crust (undifferentiated) Mostly Oceanic Plates Associated with Backarcs Oceanic Subduction Continental Boundaries & Structures: Plate Velocities: Half Spreading Rates in mm/ year (Rates simplified from Haghipour 2005) Diverging Plate Boundaries/ Mid Ocean Ridges: Converging Rates Toward Boundaries: PA - Pacific Plate PH - Philippine Plate CA - Caribbean Plate SC - Scotia Plate NZ - Nazca Plate Continental (A) Subduction Boundary Passive Ocean / Continent Boundary Cenozoic/ Mesozoic Folded Belts Tertiary Rifts Basement - Involved Uplifts and/ or Rift Inversions CO - Cocos Plate JF - Juan De Fuca Plate CAL - Caroline Plate Mid Ocean Ridge & Transform Faults Continent/Ocean Plates associated with breakup of Gondwana parts of Pangea Continent/Ocean Plates associated with breakup of Gondwana parts of Pangea SA - South America NA - North America Y - Yangtse SU- Sunda B - Borneo EU - Eurasia* AF/NU - Africa / Nubia Plate AF/SOM - Africa / Somalia Plate Diffuse Plate Boundaries: On Continents and close to Subduction Zones (mostly subaerial) Deformation inferred from Seismicity, Topography, Faulting, etc. Submarine Regions where non-closure of plate circuits indicates measurable deformation; deformation mostly inferred from Seismicity. AR - Arabia IN - India AU - Australia CAP - Capricorn a. Paleozoic b. Cenozoic/ Mesozoic *Including Ill-defined sub-plates within diffuse plate boundaries as follows: O - Okhotsk T - Tarim AM - Amuria a b 15 87 LW - Lwandle Plate 25.7 Legends for Plates 8-10 Diffuse plate boundaries. 1 0 1 8 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s 35 39 30 46 15.7 7 777 7 7 7 22 32 36 72 77 60 35.5 2532.2 18 14 44 48 100/110 83 10.6 12 16.8 17 17 15 12 11 10 10 11 8 8.6 6.9 7.2 49 50 70 72 75 70 84 78 87 Credits: Modified from Gordon and Stein 1992, Gordon 1998 and 2000 Convergence and divergence rates simplified after Haghipour 2002 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W Plate 25.8 Diffuse Plate boundaries. 1 0 1 9 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s NA CO JF PA NZ SA AF/NU EU AR CAP T IN AF/SOM CA AN AU SU PH Y AM OK SC PA LW CAR Credits: Modified from Gordon and Stein 1992, Gordon 1998 and 2000, Convergence and divergence arrows simplified after Haighpour 2002, Combined with Bally 2010 (Plate 18) 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S Plate 25.9 Diffuse plate boundaries, Cenozoic/Mesozoic Folded Belt and Cenozoic Rifts. 1 0 2 0 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s NA CO JF PA NZ SA AF/NU EU AR CAP T IN SU AF/SOM CA AN AU PH Y AM OK SC PA LW CAR 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S Credits: Modified from Gordon and Stein 1992, Gordon 1998 and 2000 Combined with Bally 2010 (Plate 18) Plate 25.10 Diffuse Plate boundaries and Tectonic Plates. 1 0 2 1 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s 80° N 10 20 40 60 80 100 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 30° E 60° E 90° E 120° E 150° E 180° 150° W 120° W 90° W 60° W 30° W 30° E0° 30° E 60° E 90° E 120° E 150° E 180° 150° W 120° W 90° W 60° W 30° W 30° E0° Credits: Simplified and modified from Chamot Rooke and Rabaute, 2006 Combined with Bally 2010 Plates 9 and 18 Graphic Scale: mm / yr Plate 25.11A Plate motions with respect to Fixed Eurasia, centered on the Pacific Ocean. P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s 1 0 2 2 Credits: Simplified and modified from Chamot Rooke and Rabaute, 2006 Combined with Bally 2010 Plates 9 and 18 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 150° W 150° W 120° W 90° W 60° W 30° W 0° 30° E 150° W 120° W 90° W 60° W 30° W 0° 30° E 60° E 90° E 120° E 150° E 180° 60° E 90° E 120° E 150° E 180° F I X E D E U R A S I A 100 80 60 40 20 10 Graphic Scale: mm / yr Plate 25.11B Plate motions with respect to a fixed Eurasia centered on Eurasia.1 0 2 3 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s 150 150125 125 125125 150 100 125 175 175 200 200 100 100 100 100 200 150 250 125 175 150 200 300 25 0 250 350 175 17 5 125 100 175 175 200 17 5 150 250 175 15 0 200 125 125 125150175175 200 100 150 100 175 100 175 125 150 125 17 5 175200 200 300 350 17 5 100 125 Credit: Based on maps published by Artemieva and Mooney 2002 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W Plate 25.12 Thickness of Thermal Lithosphere. 1 0 2 4 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s 30 40 40 30 10 50 40 40 45 45 45 50 30 10 30 40 3010 10 45 40 10 10 10 10 10 30 70 45 40 40 40 30 30 30 10 30 40 10 30 30 10 5010 30 30 10 40 45 10 30 10 10 30 40 40 10 10 30 30 45 Paleozoic Cenozoic - Mesozoic Age of Continental Basement: Precambrian Thickness of Continental Crust in KM Credits: Based on Mooney et al (1998). (Note: a new more detailed model is available from Laske et al 2009) and Bally (2010) Plates 32 a & b. 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 20 Plate 25.13 Thickness of Continental Crust and Age of Basement. 1 0 2 5 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps 2255..66 Tectonic maps of the world Introduction to tectonic maps Earlier global tectonic maps emphasized structural trends and some authors pre- ferred to focus on rectilinear lineaments associated with preferred directions (e.g., Beaumont, 1852; Hobbs, 1904). Interest in lineaments continued over the years, particularly with photogeologists and also withmetallogenists engaged inmineral exploration, for example, Hodgson et al. (1974), Podwysocki and Earle (1976), O’Leary and Earle (1978), Gabrielsen et al. (1981), and Sabin (1996). Many of these researchers differentiate fractures, shear zone, and linear fault trends. More recently, Dekalb (2007) published a global map of diagonal shear lines on a Mer- cator projection. The reality and significance of all these patterns remain elusive, and explanations involve concepts that are difficult to reconcile with plate tectonic principles.Moreover, past lineament studies often do not clearly differentiate pres- ent stress patterns (e.g., Heidbach et al. 2007a, 2007b) from paleostress patterns that are known to change over geologic time, in response to evolving plate tec- tonic regimes (e.g., Angelier, 1984, 1989; Bergerat, 1987; for more see also 25.4 of this chapter). Beginning with Suess (1885–1909), several authors, for example, Argand (1922), Kober (1928), Staub (1928), and Umbgrove (1947) tried to simplify “patterns of folded belts.” Terms such as linkages, deflections, syntaxis, and virgations were used to describe the bundling and unraveling of curved fold belt trends. It was all part of the descriptive approach of the earth sciences prior to the advent of plate tectonics (for overviews, see Bucher, 1957; Dennis, 1967). Global tectonic trend maps served as background to support drifting, escaping, and subsequent conver- gence of continents toward and away from the poles (Staub, 1928) or else a “Pulse of the Earth” (Umbgrove, 1947). Large-scale tectonic wall maps covering entire continents or some of the oceans emerged during the 1950s and 1960s. Many of these maps were based on the now outdated geosynclinal concepts (see Dennis, 1967; Dott and Shaver, 1974; Table 25.1). Various types of geosynclines were combined with Stille’s (1924) orogenic concepts that involved“ episodic changes in rock fabric,” which, as pointed out by Dennis (1967), marked a conceptual change from making an orogeny an event as well as a process. Typical large-scale tectonic maps of this period are for Eurasia (Yanshin, 1966) and Peive and Yanshin (1980), for S. Amer- ica, Martin-Bellizia (1978), and for N. America, King (1969a). These and other similar maps are important milestones of a tectonic legacy that was largely based on extensive field observations that are still relevant today. Beginning in the late 1960s, following the wide acceptance of plate tectonics, new global and 1027 continent-wide geological and tectonic maps were published culminating with ern plate tectonics. An important milestone that related Alpine tectonics to the Dinarides and the Apennines was published by Bigi et al. (1983) and Bigi et al. 1028 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps (1990). For more on the history of Alpine interpretations, see Dal Piaz (2001) and De Graciansky et al. (2011). New perspectives on Alpine tectonics have emerged in recent years. Schmid et al. (2004a) published and updated tectonic maps of the Alps and also a map of their continuation in the Carpathians and Dinarides (Schmid et al., 2008). These maps the outstanding Exxon (1985) 1:10 M Tectonic map of the World. More recent tectonic maps that are firmly grounded in plate tectonic principles include the Tectonic Map of North America (Muehlberger, 1992), the Seismotectonic Map of the World (Haghipour, 2002), the Structural and Kinematic Map of the World (Haghipour, 2006), the Geological Map of North America (Reed et al., 2005), and the Structural Map of Eastern Eurasia (Pubellier, 2008). Recent advances in alpine tectonics: An example of the scope of larger scale tectonic maps The small scale of Plates 25.14–25.19 permits only a simple first order overview of the global plate tectonic setting of orogens that developed over hundreds of million years ago. Byway of contrast, Plate 25.11 is a simplified viewof the relation- ship of present plate tectonics with their current impact on Cenozoic/Mesozoic orogens. On the basis of these neotectonic data, Table 25.2 lists 30 different orogenic settings assigned to 11 fundamentally different groups. Clearly, this leaves little room for a unique orogenicmodel or a single “typical” orogenicmodel. The Alps and their continuation in the Carpathians are perhaps the best known, yet in some people’s mind they are an “atypical” orogenic system. The following briefly outlines the scope of the tectonic/geophysical documentation that is presently available for this specific folded belt. The present integration of Alpine tectonic and geophysical data is impressive and may serve as a desirable model for the documentation of other orogenic systems. Based on the pioneering work of Argand (1911, 1916) and his contemporaries, a detailed tectonic map by Staub (1924) of the Alps showed the distribution all major tectonics units. Nappes were correlated on the base of their common stratigraphic and tectonic evolution. Axial (i.e., “downplunge”) projections allowed correlation of nappes from one axial culmination across an axial depression into the next axial culmination that may appear in a tectonicwindow. Themaps and the cross-sections constituted a compel- ling tectonic image supporting a mobilistic perspective that explicitly accepted Wegener’s (1912, 1966) continental drift. As pointed out by Tru¨mpy (2001), it was the imaginative work of geophysicists and oceanographers, tested and cali- bratedbythe JoidesDeepSeaDrillingProject, thatpermittedtheformulationofmod- were complemented by the Metamorphic Maps of the Alps (Oberha¨nsli and Wortel, 2004), and with long lithospheric transects across the Alps and the Medi- Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps terranean (Cavazza et al., 2004; Schmid et al., 2004b) and reconstructions by Stampfli et al. (2001), Stampfli and Borel (2004). Handy et al. (2010) reconciled these tectonic overviews with seismic and tomo- graphic observations in a new set of reconstructions of the Alpine Carpathian system. These authors correlated the amount of subducted lithosphere derived from their reconstructions with the slab material imaged by seismic tomography. There appears to be more (i.e., about 10–30%) reconstruction-derived litho- sphere than indicated by the tomographic images. The authors propose a number of explanations for this discrepancy. However, in the context of regional geology, it is more important to note the extraordinary progression of tectonic mapping and understanding of the Alpine system from the early 1900s until today. The Alpine example shows the importance of synthesizing traditional detailed geolog- ical maps into modern tectonic maps and integrating crustal reflection and refrac- tion profiles for crustal transects supported by crustal and lithospheric tomography. By all accounts, the Alpine–Carpathian system is now one of the best and most intensely studied and documented orogens of the world. When comparedwith the Alpine/Carpathianmaps, the scope and simplification of our small-scale global tectonic and basin maps are very limited indeed, yet the need for a simple global overview remains if only to compare and contrast the detailed evolution of orogens and sedimentary basins. Simple small-scale thematic maps also help to communicate and to preserve and organize an earth science legacy that is accumulating at an ever-increasing pace. Simplified tectonicmaps of theworld (Plates 25.14–25.16A,B) The small-scale tectonic maps of the world (Plates 25.15 and 25.16A,B) and the legend (Plate 25.14) are an updated version of similar maps published earlier by Bally and Snelson (1980) and Bally et al. (1985). Spreading oceans, oceanic subduction zones, and transform faults are the key elements that circumscribe major plates. Goffe´, 2004; Oberha¨nsli et al., 2004) and recent overviews on orogenic processes in the Alpine Collision Zone (Froitzheim and Schmid, 2008). Crustal reflection seismic profiles were acquired across the Alps such as the ECORS-CROP traverse across the Western Alps (e.g., Roure et al., 1996), the NPR 20 project and its trans- ects across the Swiss Alps (e.g., Pfiffner et al., 1997), the TRANSALP traverse across the Eastern and Southern Alps (e.g., Castellarin et al., 2006; Lu¨schen et al., 2004), and the CELEBRATION Project (e.g., Grad et al., 2006, 2007; Tasa´rova´ et al., 2009) across the Carpathians. These profiles weremergedwith tomographic crustal/lith- ospheric maps (e.g., Hafkenscheid et al., 2006; Kissling et al., 2006; Spakman and 1029 Because today both oceanic and continental subduction are widely accepted, it is 1030 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps suggested that the terms A (Ampferer) and B (Benioff) subduction introduced by the above cited are no longer necessary, as long as the well-defined oceanic sub- duction is separated from less well-circumscribed continental subduction (for more see Section “Subduction, Sutures, and Orogens”). Based on geophysical and geodetical observations, diffuse plate boundaries (Plates 25.8 and 25.9), as well as pervasive strain partitioning within folded belts (orogens), are widely recognized. Therefore, there is also no longer a need for the term “Megasuture” because it is now accepted that most orogens represent a compressed record of earlier diffuse plate boundaries. Present and past orogens are the result of both ductile and rigid deformation processes affecting sedimentary sequences and metamorphic/igneous crustal basement rocks and synorogenic intrusions and extrusions. Earlier, it was re-affirmed that, despite inherent semantic problems, the terms oro- gen and folded belts will be used here interchangeably, because both terms are already deeply entrenched in the literature. Graphically simplifying the world’s major orogenic systems in themanner shown on Plates 25.15 and 25.16A,B allows use of them as tectonic background for our basin maps and basin classifications (Plates 25.29–25.36). Global tectonic maps are yet another interpretation of exhumed/outcropping orogenic belts and their mostly older subsurface equivalents, that is, the basement that underlies all sedimentary basins. Rift systems are also important major tectonic elements. However, to avoid undue clutter, rifts are not shown on our tectonic maps. Instead, rifts are included in the context of sedimentary basins (Plates 25.29 and 25.30). The tectonic maps (Plates 25.14–25.16A,B) show that since the early Jurassic, ocean spreading created the relatively rigid oceanic crust that underlies about two thirds of the Earth surface. During the same interval, Cenozoic/Mesozoic (Cz/Mz) folded belts were formed. Major boundaries of these orogens are oceanic subduction zones often associated with deep-sea trenches, and transform/strike slip boundaries (e.g., the San Andreas fault system of California or the Alpine fault system of New Zealand). The continental subduction boundary system is not rig- orously circumscribed and coincides with the cratonward boundaries of foreland fold and thrust belts (FFTBs). The tectonic maps lump into a single system of all Cenozoic/Mesozoic orogens, that is, (1) the arcs of the Western Pacific, the Andean–Cordilleran system of the East Pacific that are associated with intra- oceanic subduction or else with oceans subducted under active continental mar- gins and (2) the Alpine–Himalayan system, that is, the collision of continents and smaller terranes. A systematic classification of these Cenozoic/Mesozoic folded belts is listed on Table 25.2. Many Cenozoic/Mesozoic and older oceanic subduction boundaries are no longer associated with active margins. Instead, they were incorporated and exhumed in the inner parts of the orogen where they form “sutures” that may be marked by volume 1C). Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps About Phanerozoic plate tectonic reconstructions Although desirable, a detailed review of the many Phanerozoic plate tectonic reconstructions that were published in recent years goes well beyond the scope of these comments. Instead, a selection of published plate tectonic reconstruc- tions includes the following: Blakey (2010), Ford and Golonka (2003), UTIG PLATES (2007/2008), Hall (2002), Golonka (2000), Golonka et al. (2006), Scotese (2001 and updates), Lawver et al. (2000, 2003, 2007), Stampfli et al. (2001), Stampfli and Borel (2004), and Yilmaz et al. (1996). Super-regional plate tectonic reconstructions combined with paleo-lithological maps particularly useful in the context of global or supraregional overviews on sedimentary basins include Dercourt et al. (1993, 2000), Barrier and Vrielynck (2008), and Daukev (2002). The paleogeographic plates included in Bohacs et al. (Chapter 23, vol 1A) eloquently complement the set of Global Tectonic and Basins maps presented in this chapter . They illuminate the long and complex Phanerozoic plate tectonic evolution of the earth that preceded the present distribution of continent and oceans, their tectonic provinces and sedimentary basins. discontinuous outcrops of ophiolites and remnant of accretionary wedges (for more on “ophiolites” see Section 25.10). The Paleozoic orogenic systems shown on the tectonic maps also include selected late Proterozoic fold belts that appear to be immediate precursors of larger Paleozoic orogenic systems. Thus, the Timanide, Baikalian, Adelaidian, Caledonian, Hercynian, and other Paleozoic systems were all merged into a single system that was completed with the assembly of the Pangea Supercontinent. Precambrian orogenic systems are all lumped on Plates 25.15 and 25.16A,B, but outcropping Precambrian shields are differentiated from Precambrian basement underlying sedimentary basins. Precambrian orogenic systems are less preserved than younger orogenic systems. However, when preserved they offer crucial insights into older crustal tectonic processes, because secular erosion has commonly exhumed deeper crustal levels. A separate Base- ment Map (Plate 25.20) shows further subdivisions of Precambrian basements and orogenic systems by age, extending into the Archean, that is, over 4.5 giga- years (see Section 25.8). A larger scale tectonic map with much added detail was prepared by Haghipour (2006). Another useful simplified tectonic map (USGS, 2010) is based on thermo- tectonic ages and combines orogenic as well as taphrogenic thermotectonic events. For this map series, it was preferred to separate (taphrogenic) rifts from orogens and to include rifts as part of our basin maps (Plates 29 and 30 of this 1031 References25.6. TectonicMapsof the World 1032 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Note: References 25.6. References for arctic maps and antarctic maps are listed separately. Angelier, J., 1989. From orientation to magnitudes in paleostress determinations using fault slip data. J. Struct. Geol. 11, 37–50. Angelier, J., 1984. Tectonic analysis of fault slip data sets; Fault behaviour and the earthquake gen- eration process. J. Geophys. Res. B 89, 4835–5848. Argand, E., 1911. Les nappes de recouvrement des Alpes Pennines et leur prolongement structur- aux. Materiaux Carte Ge´ologique Suisse 31, 1–26. Argand, E., 1916. Sur l’arc des Alps Occidentales. Eclogae Geologicae Helveticae 14, 145–192. Argand, E., 1922. La tectonique de l’Asie. Compte Rendu du 13e`me Congre`s Ge´ologique International, Belgique. Vaillant-Carmanne, Lie`ge. pp. 171–372. Translation by Carozzi A.V. (ed.andtranslator)EmileArgand:TectonicsofAsia.NewYork:Hafner.218pp.and1TectonicMap. Bally, A.W., Snelson, S., 1980. Realms of subsidence. In: Miall, A.D. (Ed.), Facts and Principles ofWorld PetroleumOccurrence, Canadian Society of PetroleumGeologistsMemoir 6, 9–94. Bally, A.W., Catalano, R.E., Oldow, J., 1985. Elementi di Tettonica Regionale Evoluzione dei Bacini Sedimentari e delle Catene Montuose. Pitagora Editrice, Bologna, 271 pp. Barrier, E., Vrielynck, C.K. (Eds.), 2008. Paleotectonic Maps of the Middle East. Tectono- sedimentary Palinspastic Maps 14 Maps at 1:18 500 000. CGMW Paris: (Commission of the Geological Map of the World) UNESCO, Paris. Bigi, G., Cosentino, D., Parotto,M., Sartori, R., Scandone, P., 1983. Structural Model of Italy: Geo- dynamic Project C.N.R. Selca, Firenze, scale 1:500 000. 9 sheets. Bigi, G., Castellarin, A., Coli, M., Dal Piaz, G.V., Sartori, R., Scandone, P., et al., 1990. Structural Model of Italy, Sheets 1–2. CNR. Progetto Finalizzato Geodinamica, SELCA Firenze,
[email protected]. Scale 1:500 000 in 6 sheets. Sheets 1 and 2 (Western and Eastern Alps). Blakey, R., 2010. Regional Paleogeographic Views of Earth History. http://jan.ucc.nau.edu/~rcb7/ globaltext.html. Beaumont, E.d.e., 1852. Notice sur les Syste`mes des Montagnes, 3 vols, 1543 pp. Bergerat, F., 1987. Stress fields in the European Platform at the time of Africa-Eurasia collision. Tec- tonics 6, 99–132. Bucher, W., 1957. The Deformation of the Earth Crust. Hafner, 518 pp. Castellarin, A., Nicolich, R., Fantoni, R., Cantelli, L., Sella, M., Selli, L., 2006. Structure of the lith- osphere beneath the Eastern Alps (southern sector of TRANSALP transect. Tectonophysics 414, 259–282. Cavazza, W., Roure, F., Ziegler, P.A., 2004. The Mediterranean area and the surrounding regions: active processes, remnant of former Tethyan oceans and related thrust belts. In: Cavazza, W., Roure, F., Spakman, W., Stampfli, G., Ziegler, P.A. (Eds.), The Transmed Atlas—The Mediter- ranean Region from Crust to Mantle. Springer, Berlin, pp. 1–29. Chapter 1. Dal Piaz, G.V., 2001. History of tectonic interpretations of the Alps. J. Geodynamics 32, 99–114. doi:10.1016/S0264-3707(01)00019-9. De Graciansky, P.C., Roberts, D.G., Tricart, P., 2011. TheWestern Alps, From Rift to Passive Margin to Orogenic Belt: An Integrated Geoscience Overview. In: Schroder Jr., J.F. Developments in Earth Surface Processes, vol. 14. Elsevier, p. 398. Dekalb, H.F., 2007. Globalmap of diagonal shear lines. NewConcepts in Global Tectonics (NCGT) Newsletter 45, 3–5. Dennis, J.C., 1967. International tectonic dictionary, English terminology. Am. Assoc. Petrol. Geo- log. Mem. 7, 196. Dercourt, J., Ricou, L.E., Vrielynck, B. (Eds.), 1993. Atlas Tethys—Paleoenvironmental Maps. 14 Maps. Commission for the Geological Map of the World. Explanatory Notes 307. Dercourt, J., Gaetani, M., Vrielynk, B., Barrier, E., Biju-Duval, M.F., Brunet, J.P., et al., 2000. Atlas Peri-Tethys, Paleogeographical Maps. 24 Maps at 1:10 000 000 and Explanatory Notes. Gau- thier Villars, Paris, 269 pp. Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Dott, R.H., Shaver, R.H. (Eds.), 1974. Modern and Ancient Geosynclinal Sedimentation, Vol. 19. Society of Economic Paleontologists and Mineralogists. Spec. Publ. 380 pp. Exxon Production Research, 1985. Tectonic Map of the World, World Mapping Project, Scale 1:5000 000, 20 Panels. American Association of Petroleum Geologists. Ford, D., Golonka, J., 2003. Phanerozoic paleogeography, paleoenvironment and lithofacies in the circum-Atlantic margins. In: Golonka, J. (Ed.), Thematic Set of Paloegeographic Recon- struction and Hydrocarbon Basins. Atlantic, Caribbean. South America. Middle East, Russian Far East, Arctic. Marine and Petroleum Geology, vol. 20. 249–285. Froitzheim, N., Schmid, S.M. (Eds.), 2008. Orogenic Processes in the Alpine Collision zone. Swiss J. Geosci.101 (Suppl) 310. Birkha¨user, Basel. Gabrielsen, R.V., Ramberg, I.B., Roberts, D., Steinlein, O.A., 1981. Proceedings of the 4th Int. Con- ference on Basement Tectonics. IBTA publications, 365 pp. Golonka, J., 2000. Cambrian to Neogene Plate Tectonic Maps. Wydawynictwa Universytetu Jagiellonskiego, Krakow, 125 pp. Golonka, J., Krobicki, M., Pajak, J., Giang, N.V., Zuxhiewicz, W., 2006. Global Plate tectonics and the Paleogeography of SE Asia. AGH University of Science and Technology, Arcadia, 128 pp. Grad,M., Guterc, A., Keller, G.R., POLONAISE 97 andCELEBRATION2000WorkingGroups, 2007. Variations of the lithospheric structure across the margin of Baltica in Central Europe and the role of the Variscan and Carpathians orogenesis. In: Hatcher Jr., R.D., Carlson, M.P., McBride, J.H., Martı´nez Catala´n, J.R. (Eds.), 4-D Framework of Continental Crust, Geological Society of America Memoir 207, pp 341–356. doi:10.1130/2007.1200(17). Grad, M., Guterch, A., Keller, G.R., Janik, T., Hegdu¨s, E., Voza´r, J., et al., 2006. Lithosphere struc- ture beneath trans-Carpathian transect from Precambrian Platform to Pannonian basin; CEL- EBRATION 2000 seismic profile Cel.O5. J. Geophys. Res. 111, B03301. doi:10.1029/ 2005JB003647. Hafkenscheid, E., Wortel, M.J.R., Spakman, W., 2006. Subduction history of the Tethyan Region derived from seismic tomography and tectonic reconstructions. J. Geophys. Res. 111b, B08401. doi:10.1029/2005//B003791. Haghipour, A., 2002. SeismotectonicMap of theWorld at 1:50 000 000 in 1 Sheet. CGMW (Com- mission of the Geological Map of the World)/UNESCO, Paris. Haghipour, A., 2006. Structural and Kinematic Map of the World at 1:50 000 000 in 1 Sheet. (digital version in high resolution format). CGMW/(Commission of the Geological Map of the World) UNESCO, Paris. Hall, R., 2002. Cenozoic geological and plate tectonic evolution of SE. Asia and the SW Pacific: computer-based reconstructions and animations. J. Asian Earth Sci. 20, 353–431. Hanson, R.E., Crowley, J.L., Bowring, S.R., Ramezani, J., Goe, W.A., Dalziel, I.W.D., et al., 2004. Coeval large scale magmatism in the Kalahari and Laurentian cratons during Rodinia assem- bly. Science 304, 1126–1129. Handy, M.R., Schmidt, S.M., Bousquet, R., Kissling, E., Bernoulli, D., 2010. Reconciling plate tec- tonic reconstructions of Alpine Tethys with the geological, geophysical record of spreading and subduction in the Alps. Earth Sci. Rev. doi:10.1016/j. earscirev.2010.06.002. Heidbach, O., Fuchs, K., Mu¨ller, B., Reinecker, J., Sperner, B., Tingay, M., et al., 2007a. World Stress Map 1:46 000 000. Incl. CD-ROM with digital maps & notes. CGMW (Commission of the Geological Map of the World) and Heidelberg Academy of Sciences and Humanities/ UNESCO, Paris. Heidbach, O., Reinecke, J., Tingay, M., Mu¨ller, B., Sperner, B., Fuchs, K., et al., 2007b. Plate boundary forces are not enough: second and third order stress patterns highlighted in the World Stress Map data base. Tectonics 26, doi:10.1029/2007TC002133. 1033 Hobbs,W.H., 1904. Lineaments of the Atlantic Border region. Geological Society America. Bulletin 15, 483–506. Hodgson, R.A., Gay Jr., S.P., Benjamins, J.V., 1974. First International Conference on New Base- 1034 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps ment Tectonics. Utah Geological Association Publication 5, 359 pp. Kober, L., 1928. Der Bau der Erde, 2nd ed. Gebru¨der Borntra¨ger, Berlin, 500 pp. Kissling, E., Schmid, S.M., Lippitsch, R., Ansorge, J., Fu¨genschuh, B., 2006. Europe Alpine to Pres- ent. Lithosphere structure and tectonic evolution of the Alpine arc, new evidence from high resolution teleseismic tomography. Geol. Soc. Lond. Memo. 32, 129–145. King, P.B., 1969a. TectonicMap of North America 1:5 000 000 U.S. Geological Survey in 2 Sheets. Lawver, L.A., Coffin, M.F., Gahagan, L.M., Campbell, D.A., Royer, J.Y., 2000. Opening of the Indian Ocean. (18 frames 180 ma-Ma), UTIG—University of Texas Institute for Geophysics gw Indian_00810_301.ppt. Lawver, L.A., Dalziel, I.W.D., Gahagan, L.M., Kygar, R.M., Martin, K.M., Campbell, D., 2003. The plates 2003 Atlas of Plate Reconstructions (750Ma to Present Day. Pacific View. In: Plates Prog- ress Report No 280-0703. UTIG Technical Report No190, 79. Lawver, L.A., Dalziel, I.W.D., Gahagan, L.M., Kygar, R.M., Martin, K.M., Campbell, D., 2007. 2006 Atlas of Plate Reconstructions (750 to Present day). Frames 2–60 in Mollweide projec- tion; Frames 61–93 Arctic in Polar projection; 94–127 Antarctic in Polar projection. Lu¨schen, E., Lammerer, B., Gebrande, H., Millahn, K., Nicolich, R., TRANSALP Working Group, 2004. Orogenic structure of the Eastern Alps, Europe, from TRANSALP deep seismic reflection profiling. Tectonophysics 388, 85–102. Martin-Bellizia, C.F., 1978.Mapa tecto´nico, Norte de America del Sur. 1: 2 250 000 in Two Sheets. Caracas: Ministero de Energia y Minas, Direction. Muehlberger, W.R., 1992. Tectonic Map of North America. American Association of Petroleum Geologists, Tulsa, Oklahoma. 4 sheets, l:5, 000 000. Mu¨ller, R.D., Royer, J.Y., Lawver, L.A., 1993. Revised plate motions relative to the hotspots from combined Atlantic and Indian ocean hotspot tracks. Geology 21, 275–278. Mu¨ller, R.D., Sdrolias, M., Gaina, C., Roest, W.R., 2008. Age, spreading rates and spreading sym- metry of the world’s ocean crust. Geochem. Geophys. Geosyst. 9, Qo4006. doi:10.10290/ 2007GC00171. Oberha¨nsli, R., Bousquet, R., Engi, R., Goffe´, B., Gosso, G., Handy, M.R., et al., 2004. Metamor- phic structure of the Alps. Scale 1:1 000 000, Commission for the Geological Map of the World. UNESCO, Paris. Oberha¨nsli, R., Goffe´, B., 2004. Explanatory notes to the map: metamorphic structure of the alps. Introduction. Mitteilungen O¨sterreichische Mineralogische. Gesellschaft 149, 115–1123. O‘Leary, D.W., Earle, J.L. (Eds.), 1978. Proceedings of the 3rd Int. Conference on Basement tec- tonics. IBTA Publications, 400 pp. Peive, A.V., Yanshin, A.L., 1980. Tectonic Map of Northern Eurasia. Akademia Nauk SSSR. Pfiffner, A., Lehner, P., Heitzmann, P., Mueller, S.t., Steck, A. (Eds.), 1997. Deep Structure of the Alps: Results from NPR-20. Birkha¨user, Basel, p. 380. Podwysocki, M.H., Earle, J.L., 1976. Second International Conference on New Basement Tecton- ics, International Basement Tectonics Association (IBTA) Publications, 595 pp. Pubellier, M., 2008. Structural Map of Eastern Eurasia. 1:12 500 000, 1 sheet. CGMW (Commis- sion of the Geological Map of the World), UNESCO, Paris. Reed, J.C., Wheeler, J.O., Tucholke, B.E., 2005. Geologic Map of North America. 1:5 000000 in 2 Sheets. Geological Society of America, Boulder, CO. Roure, F., Bergerat, F., Damotte, B., Mugnier, J.L., Polino, R. (Eds.), 1996. The ECORS CROPAlpine Seismic Traverse, Bulletin de la Socie´te´ Ge´ologique de France 170, 1–113. Ronov, A., Khain, V.E., Seslavinski, A., 1984a. Atlas of Lithological and Paleogeographical Maps of the World: Late Precambrian and Paleozoic of the Continents. Leningrad USSR Academy of Sciences, p. 70. Ronov, A., Khain, V.E., Balukhovski, A., 1984b. Atlas of Lithological and Paleogeographical Maps of the World: Mesozoic and Cenozoic of the Continents. Leningrad USSR Academy of Sciences, p. 70. Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Sabin Jr., F.F., 1996. Remote Sensing: Principles and Interpretations. 3rd edition. W.H Freeman, London, 432 pp. Scotese, C., 2001 and updates. PALEOMAP project: http://www.scotese.com. Schmid, S.M., Fu¨genschuh, B., Kissling, E., Schuster, A., 2004a. Tectonic Map and overall archi- tecture of the Alpine orogen. Eclogae Geol. Helv. 97 (1), 93–117. Schmid, S.M., Fu¨genschuh, B., Kissling, E., Schuster, A., 2004b. TRANSMED transects IV, V and VI: Three lithospheric transects across the Alps and their forelands. CD. In: Cavazza, W., Roure, F., Spakman, W., Stampfli, G., Ziegler, P.A. (Eds.), The Transmed Atlas: The Mediterranean from Crust to Mantle. Springer, Berlin. Schmid, S.M., Bernoulli, D., Fu¨genschuh, B., Matenco, L., Schefer, S., Schuster, R., et al., 2008. The Alpine-Carpathian-Dinaridic orogenic system; correlation and evolution of tectonic units. Swiss J. Geosci. 101, 139–183. Sengo¨r, A.M.C., Natalin, B.A., 1996. Paleotectonics of Asia. Fragments of a synthesis. In: Yin, A., Harrison, T.M. (Eds.), TheTectonic EvolutionofAsia.CambridgeUniversity Press, pp. 486–640. Spakman, W., Wortel, M.J.R., 2004. A tomographic view of the Western Mediterranean geody- namics. Chapter 2 In: Cavazza, W., Roure, F.M., Stampfli, G.M., Ziegler, P.A. (Eds.), The Transmed Atlas—TheMediterranean Region from Crust toMantle. Springer, Berlin, pp. 31–52. Stampfli, G.M., Borel, G., 2004. The TRANSMED transects in time and space. In: Cavazza, W., Roure, F., Spakman, W., Stampfli, G., Ziegler, P.A. (Eds.), The Transmed Atlas: The Mediterra- nean from Crust to Mantle. Springer, Berlin, pp. 53–80. Stampfli, G.M.,Mosar, J., Favre, P., Pillevuit, A., Vannay, J., 2001. Permo-Mesozoic evolution of the western Tethys realm; theNeotethys EasternMediterraneanbasin connection. In: Ziegler, P.A., Cavazza, W., Robertson, A.H.F., Soleau, C.S. (Eds.), Peri-tethyan Rift/wrench Basins and Pas- sive Margins: Memoires du Museum d’Histoire Naturelle, vol. 186. Paris, pp. 51–108. Staub,R.,1924.Beitra¨gederGeol,KarteSchweis,N.F.52,2Plates,1MapDerBauderAlpen.Bern272. Staub, R., 1928. Der Bewegungsmechanismus der Erde, Berlin Gebru¨der Borntraeger, 1 colored map 44 Figs. 270. Stille, H., 1924. Grundfragen der vergleichenden Tektonik. Borntra¨ger, Berlin, p. 443. Suess, E., 1885–1909. Das Antlitz der Erde. V. 1. 1885. 778 pp; vol. 2. 1888. 703 pp.; vol. 3. part 1. 1901. 508 pp; and part 2. 1909. 789 pp. Prag and Vienna, Tempsky. Translation by Sollas, H.B.C., Sollas W.J. 1904–1924. The Face of the Earth (authorized English translation) Oxford Clarendon. vol. 1.1904. 104 pp; vol. 2. 1906. 556 pp; vol. 3. 1908. 400 pp. and vol. 4. 1925. 170 pp; Translation by E .deMargerie E. 1921. La face de la Terre (authorized French trans- lation). Librairie Armand Colin. 4 volumes 3079 pp. Tasa´rova´, A., Afonso, J.C., Bielik, M., Go¨tze, H.J., Ho´k, J., 2009. The lithospheric structure of the Western Carpathian–Pannonian Basin region based on the Celebration 2000 seismic experi- ment and gravity modeling. Tectonophysics 475, 454–469. Tru¨mpy, R., 2001.Why plate tectonics was not invented in the Alps. Int. J. Earth. Sci. 90, 477–483. Umbgrove, J.H.F., 1947. The Pulse of the Earth.MartinusNijhoff, TheHague, p. 358. 8 color plates. USGS, 2010. Geologic Province and Thermo-Tectonic Age Map Earthquake Hazards Program. http://earthquake.usgs.gov/research/structure/crust/map. UTIG (The University of Texas Institute of Geophysics) PLATES project, 2007/2008. 11 reconstruc- tions in Mollweide projection and 9 reconstructions in Orthographic projection http//www .plates.ig.utexas.edu. Wegener, A., 1912. Die Entstehung der Kontinente. Petermann Mitteilungen 58, 185–195, 253– 256 and 305–309. Wegener, A., 1966. TheOrigin of Continents andOceans. Translated from the 4thGerman edition by John Biram (Translator). Dover Publications, New York, p. 246. 1035 Yanshin, A.M., 1966. TectonicMap of Eurasia, 1:5 000 000MGeol Inst. SSSR andMinistry of Geol- 1036 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps et al. (2008). Interestingly, Ady et al. (2010) and Dinkelman et al. (2010) were unable to find “a good fit for the Amerasian basin” and propose a two-phase opening that began ogy USSR. Yilmaz, P.O.,Norton, I.O., Leary,D.C.,Chuchla,R.J., 1996.TectonicevolutionandPaleogeography of Europe. In: Ziegler, P.A., Horvath, F. (Eds.), Peri-TethysMemoir 2, Structure and Prospects of Alpine Basins and forelands, Memoires du Museum d’Histoire Naturelle, vol. 170, pp. 47–60. Polar tectonic maps: Introduction (Plate 25.16) Polar projections of the Arctic and Antarctic demonstrate the overall continuity of major tectonic provinces in the areas unduly distorted by the Mercator projection used for most of our global maps. Originally, basic data for the Arctic map (Plate 25.16A) as well as the Antarctic tectonic map (Plate 25.16B) were mostly gleaned from Choubert Faure-Muret (1976/1983). Later, it was attempted to update the early version of both maps. The Arctic Ocean and its surrounding con- tinents and Antarctic continent have had contrasting andwidely different tectonic evolutions. The presently melting polar ice, while being centered on the Antarctic Continent, is distributed over both the Arctic Ocean and Greenland Ice caps (see Plate 25.2A and B). Polar ice distributions are shown on Plate 25.2A and B. Polar earthquake distributions are shown on Plates 25.5 and 25.6. Arctic tectonic map (Plate 25.16A) The tectonic map of the Arctic is greatly simplified when compared with the outstanding and most recent compilations of Arctic geology (Grantz et al., 2009; Harrison et al. 2005). For a concise overview of Arctic geology, see Harrison (2005). Important issues relate to the opening of the Amerasian basin of the Western Arctic. Ziegler (1988,1989), Lawver et al. (2000,2002), andGolonkaet al. (2003)published plate tectonic reconstructions of the Arctic. Colpron and Nelson (2009) connect the Cordilleran Paleozoic terranes of California and southern Alaska with the northern Caledonides, the Timanides, the Polar Urals, and Taymir in a Caribbean–Scotia arc style, looped subduction system that was completed during the Upper Paleozoic, that is, prior to the opening of much of the present Arctic Ocean. For a number of years the opening of the Arctic Ocean and particularly the Amer- asian basin has been the subject of controversy. A majority of Arctic workers now prefer an opening of the Amerasian basin that is associated with a counterclock- wise rotation of Northern Alaska away from the Canadian Arctic Islands. Different perspectives on this were presented by Lane (1997) and Grantz et al. (1998). An improved geophysical characterization of the Western Arctic Ocean and a discus- sion of a number of models to account for its opening were published by Alvey with transtension in the Lower Cretaceous and a later phase that still involves delta area. The suggestion is that the Canadian Arctic continental marginmaywell Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps be a complex continent–ocean transform margin analogous to the shear margins as concisely summarized by Bird (2001). Plate 25.16A also shows a greatly simplified picture of the combined Alpha– Mendeleyev ridges interpreted as an “oceanic plateau” (Maher, 2001) (note here oversimplified “oceanicplateau” (in black). Formore on thisHighArctic LIP, see Jokat (2003) andBuchan and Ernst (2006a, 2006b). Clearly,models of the openingAmer- asian basin are likely to be modified by future studies of this remote area. The Arctic basin map (Plate 25.32A) shows that the outcropping basalts of the Tunguska plateau of eastern Siberia extend under most of the West Siberia basin as documented on the basis of regional seismic data calibrated by wells (Reichow et al., 2004; Vyssotski et al., 2006, this volume). Kuzmichev and Pease (2007) and Kuzmichev (2009) observed that tholeitic subvolcanic intrusions on the west- ernmost Siberian Islands “are identical in age” to those of Siberian traps. Accept- ing their interpretations would likely lead to important modifications of Plate 25.16A in the Arctic regions of Northeast Siberia on and offshore. To sum up, Arctic tectonics are fascinating and our understanding is still evolving. Antarctic tectonic map (Plate 25.16B) The Geologic History of Antarctica is reviewed in Chapter 2 of Anderson (1999a, 1999b). St. John (1991) edited a volume on the petroleum potential of Antarctica. In a concise overview of Antarctica, Storey (2005) notes its key position in plate tectonic reconstructions of the supercontinent of Rodinia (1000 Ma) and Neopro- terozoic assembly of Gondwana (540 Ma). East Gondwana includes the assembly of Africa/Arabia, India, Australia/S. China, and possibly N. America. This is illu- strated by a map that shows the relationship of Pre-Grenville, Grenville, and Pana- frican basements (see Yoshida et al., 1992, 2003a, 2003b). counterclockwise rotation of northern Alaska. Clearly, the views on the origin of the Arctic Ocean and, particularly, its Amerasian basin, as well the Alpha- Mendeleyev ridges, are currently evolving. Verzhbitsky et al. (2007) review the relations of the Pacific plate to the Bering Sea and conclude that the Aleutian basin represents a fragment of the Upper Cretaceous Kula Plate. Time did not allow incorporation of many of the latest insights into the geology of the Arctic on Plate 25.16A. Our Arcticmap thereforemerely shows the overall con- tinuity of some of the major elements of the tectonic map (Plate 25.15). A striking feature of Arctic tectonics is the near linear trend that extends the trans- form faults offshore W. Spitsbergen (De Geer fault system) and North Greenland, along the northern shelf of the Arctic islands all the way toward the Mackenzie 1037 Based on the geochemistry of Antarctic Paleozoic-Quaternary mafic igneous rocks, Leat et al. (2005) identify a stable/cratonic East AntarcticArchean toMid-Proterozoic 1038 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps subcontinental/lithosphericmantledomain.Onthebasisof seismic tomography, the lithosphere is “up to 250 km” thick. They also suggest that much of the region was reworked during the Panafrican orogeny. On the other hand, West Antarctica is characterizedbyMiddleProterozoic–Paleozoiccrustal rocks,andseismic information suggests the absence of deep lithospheric roots. Leat et al. (2005) further indicate that amalgamation of West Antactica involved terrane accretion after the Cambrian Ross orogeny. The process ended with the docking of the Antarctic Peninsula in the Mid-Cretaceous. In sharp contrast with the East Antarctic craton, the lithosphere thickness of West Antarctica is estimated to be no more than 50 km, that is, less than an 80–100-km-thick attenuated continental crust or an oceanic crust, or an oceanic crust that is older than 100 Ma. Leat et al. (2005) conclude that the thinning of crust is due to rifting pro- cesses associated with the breakup of Gondwana from the Mid-Jurassic to the Mid-Cretaceous and the Cretaceous to Tertiary rifts schematically shown on Plate 25.32B. By analogy to other circum-Pacific areas, one might also be tempted to speculate that the now inactive West Pacific margin of Western Ant- arctica might have been more active during the Late Mesozoic and early Tertiary causing xtension due to slab-rollback. Clearly, there is need for a more precise res- olution of the ages of rifting in W. Antarctica than the crude dates implied by Plate 25.32B. Much like the Arctic region, but in fundamentally differentways from it, Antarctica is also a remote frontier region where new information is likely to modify our tectonic understanding as new observations become available. Like the Arctic, plate tectonic reconstructions for Antarctica are particularly important to define remaining pro- blems and to check solutions that are proposed for them. A very careful, detailed reconstruction of Antarctic Paleogeography was published by Torsvik et al. (2007). These authors present eight paleogeographic maps from the Neopro- terozoic (750 Ma) Rodinia—just after breakup—to post-Panafrican, Late Cambrian (500 Ma), to Gondwana, Late Permian (250 Ma) and Early Jurassic (180 Ma) with reconstruction of Karroo and Ferrar volcanic provinces (179–184 ma), to the inception of Circum Atlantic (160 Ma) and Indian Ocean spreading (130 ma), and finally to the present stage of all the oceans surrounding Antarctica. Note that throughout the Late Permian and the Mesozoic, an active continental margin extended from Western South America to Western Antarctica. This again was further modified by the spreading and opening of the Scotia Sea. References to Polar Tectonic Maps Note: Someof the referencesbelowarenot specifically cited inourbriefArcticandAntarcticoverviews.They are listedherebecausetheywereconsulted inthecompilationof thePolar tectonicandbasinmapsand in the text of the short introduction of both the tectonic and the basin maps. Arctic Tectonic (Plate 69.16A) and Basin (PLATE 69.32A) Maps Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Ady, B., Dinkelman, M., Helwig, J., Kumar, N., 2010. Deformable Plate Tectonic Reconstructions Incorporating Interpretations from Newly Acquired Geophysical Data Support a Multi-Phase Plate Tectonic Model for the Origin and Evolution of the Amerasian Basin. Abstract AAPG ICE Conference, Calgary. Alvey, A., Gaina, C., Kusznir, N.J., Torsvik, T.H., 2008. Integrated crustal thickness mapping and plate reconstructions for the high Arctic. Earth Planet. Sci. Lett. 274, 310–321. Buchan, K.L., Ernst, R., 2006a. Giant dyke swarms and the reconstruction of Canadian Arctic islands, Greenland, Svalbard and Franz-Josef Land. In: Handski, E.S., Mertanen, S.T., Ramo¨, T., Vuollo, J. (Eds.), Dyke Swarms Time Markers of Crustal Evolution. Taylor and Francis/Balkema, pp. 27–49. Buchan, K.L., Ernst, R.E., 2006b. The High Arctic Large Igneous Province (HALIP) Evidence for an associated Giant Radiating Dyke Swarm. LIP of the Month http://www.largeigneouspro- vinces.org/06apr. Bird, D.E., 2001. Shear margins: continent-ocean transform and fracture zone boundaries. Lead. Edge 150–159. Colpron, M., Nelson, J.L., 2009. A Paleozoic Northwest passage: incursion Caledonian, Baltican and Siberian terranes into Eastern Panthalassa and the early evolution of the North American Cordillera. In: Cawood, P.A., Kro¨ner, A. (Eds.), Accretionary Systems in Space and Time, Geo- logical Society London Special Publication 318, 273–307. Choubert, G., Faure-Muret, A., (Eds.), 1976/1983. Geological Atlas of the World. 1:10M, 22 Sheets with Text. CGMW, Paris (Commission of the Geological Map of the World)/UNESCO, Paris (also available in ArcInfo format). Dinkelman,M., Ady, B., Helwig, J., 2010. EvaluatingCurrent Plate TectonicModels in Light ofNewly Acquired Geophysical Data in the Amerasian Basin. GeoCanada 2010, Calgary. Abstract. Eldholm, O., Skogseid, J., Sundvor, E., Myhre, A.M., 1990. The Norwegian Greenland Sea. In: Grantz, A., Johnson, L., Sweeney, J.F. (Eds.), The Arctic Ocean. The Geology of North America. L. Geological Society of America, Boulder, CO, pp. 351–364. Eldholm, O., Thiede, J., Taylor, B., et al., 1987. Proc. ODP. Init. Reports, Ocean Drilling Program, 104, College Station, TX. doi:10.2973/odp.proc.ir.104.1987. Lawver, L.A., Gahagan, L.M., Campbell, D.A., et al., 2000. Tectonic Evolution of the Arctic Region since the Ordovician. 450-0 Ma.
[email protected]. Grantz, A., Johnson, L., Sweeney, J.F. (Eds.), 1990. The Arctic Ocean Region. The Geology North America. L. Geological Society of America, p. 664. Grantz, A., Clark, D.l., Phillips, R.L., Srivastava, S.P., Blome, C.D., Gray, L.B., et al., 1998. Phanero- zoic stratigraphyofNorthwindRidge,magnetic anomalies in theCanadabasin, and thegeom- etryandtimingof rifting in theAmerasiabasin,Arcticocean.Geol. Soc.Am.Bull. 110,801–820. Grantz, R.A., Scott, S., Drachev, S., Moore, T.E., 2009.Maps showing the sedimentary successions of the Arctic region (58�-64� to 90� N) that may be prospective for hydrocarbons (incl. 78 pp. Explanatory Notes); American Association of Petroleum Geologists GIS-UDRIL Open-File Spa- tial Library; http://gisudril.aapg.org/gisdemo/. Gautier, D.L., Bird, K.J., Charpentier, R.R., Grantz, A., Houseknecht, D.W., Klett, T.R., et al., 2009. Assessment of Undiscovered Oil and Gas in the Arctic. Science 329, 1175–1179. U.S. Geolog- ical Survey, Circum-Arctic Resource Appraisal (north of the Arctic Circle) Assessment Units GIS Data (2009); http://energy.usgs.gov/arctic/. Golonka, J., Bocharova,N.Y., Ford,D., Edrich,M.E., Bednarczyk, J.,Wildharber, J., 2003. Paleogeo- graphic reconstructionsandbasindevelopmentof theArctic. In:Golonka, J. (Ed.), ThematicSet of PaloegeographicReconstruction andHydrocarbonBasins. Atlantic, Caribbean,SouthAmer- ica,Middle East, Russian Far East, Arctic.Marine and PetroleumGeology, vol. 20, pp. 211–248. Harrison, J.C., 2005. Geological History of the Arctic. In: Nuttall, M. (Ed.), Encyclopedia of the Arctic. Routledge, New York, pp. 711–722. 1039 Harrison, J.C., et al., 11 lead compilers 2005. Geological Map of the Arctic, 1:5 000 000 in 5 Sheets. Ottawa Geological Survey of Canada (GSC Open File 5816) and CGMWCommission 1040 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps of the Geological Map of the World/UNESCO, Paris. Jackson, H.R., Forsyth, D.A., Johnson, G.L., 1986. Oceanic affinities of the Alpha Ridge. Mar. Geol. 73, 243–261. Jokat, W., 2003. Seismic investigations along the western sector of the Alpha Ridge, central Arctic Ocean. Geophys. J. Int. 152, 185–201. Kuzmichev, A.B., 2009. Where does the south Anyui suture go into the New Siberian Islands and the Laptev Sea: implications for the Amerasia basin origin. Tectonophysics 463, 86–108. Kuzmichev, A.B., Pease, V.L., 2007. Siberian trap magmatism on the New Siberian Islands: con- straints for Arctic Mesozoic plate reconstructions. J. Geol. Soc. 164, 959–968. Lane, L.S., 1997. Canada basin, Arctic ocean: evidence against a rotational origin. Tectonics 16, 363–387. Lawver, L.A., Gahagan, L.M., Campbell, D.A., 2000. Tectonic evolution of the Arctic Region since the Ordovician. Austin UTIG 15 Powerpoint frames. contact L. Gahagan: plates@ig. texas.edu. Lawver, L.A., Grantz, A., Gahagan, L.M., 2002. Plate Kinematic Evolution of the Present Arctic Region Since the Ordovician. GSA Special Paper 360, pp. 333–358. Maher Jr., H.D., 2001. Manifestation of the cretaceous high arctic large igneous province. J. Geol. 109, 91–104. Moore, T.E., Wallace, W.K., Bird, K., Mull, C.J., Dillon, J.T., 1949. Geology of Northern Alaska. In: Plafker, G., Berg, H.C. (Eds.), The Geology of Alaska, DNAG Vol G-1. The Geology of North America. Geological Society of America, pp. 49–140. Nikishin, A.M., Sobornov, K.O., Prokopiev, A.V., Frolov, S.V., 2010. Tectonic evolution of East Siberia during the Vendian and Phanerozoic. Moscow Univ. Geol. Bull. Springer Link 65, 1–16. Olaussen, S., Steel, R.E., 2011. Getting Started #2-Arctic Geology. AAPG/Datapages CD-Rom. Reichow, M.K., Saunders, A.D., Ivanov, A.V., Puchkov, V.N., 2004. The Siberian large igneous province. Event 36 of LIP database. LIP of the Month 7. Spencer A.M., Embry A.F., Gauthier D.L., Stoupakova A.V., Sørensen K. (Eds), 2012. Arctic Petro- leum Geology. Geological Society Memoir No.35. London p. 818. Stein, R., 2008. Arctic ocean sediments: processes, proxies and paleoenviroment. Devel. Marine Geol. 2, 608. Trettin, H.P., 1989. The Arctic Islands. In: Bally, A.W., Palmer, A.R. (Eds.), The geology of North America. DNAG Vol. A. Geological Society of America, pp. 349–370. Trettin, H.P. (Ed.), 1991. Geology of the innuitian orogen and arctic platform of Canada and Greenland. DNASG vol. E Geological Survey of Canada, Geology of Canada, No 3; Geological Society of America: Vol. E1. Text 569 pp; Vol. E2 Plates (8 Maps). Verzhbitsky, E.V., Kononov, M.V., Kotelkin, V.D., 2007. Plate tectonics of the northern part of the pacific ocean. Oceanology 47, 705–717. Vyssotski, A.V., Vyssotski, V.N., Nezhdanov, A.A., 2006. Evolution of the West Siberian Basin. Mar. Petrol. Geol. 23, 93–126. Ziegler, P.A., 1988. Evolution of the Arctic North Atlantic and the Western. Tethys Am. Assoc. Petrol. Geol. Memo. 43, 198. Ziegler, P.A., 1989. Evolution of Laurussia—A Study in Late Paleozoic Plate Tectonics. Kluwer Academic Publishers, The Netherlands, p. 102 and 14 Plates. Antarctic Tectonic (Plate 69.16B) and Basin (Plate 69.32B) Maps Anderson, J.B., 1999a. Antarctic Marine Geology. In: Cambridge University Press, Cambridge, p. 289. Anderson, J.B., 1999b. Geologic history of Antarctica. In: Anderson, J.B. (Ed.), Antarctic Marine Geology, pp. 28–57. Black, L.P., Sheraton, J.W., Kinny, P.D., 1992. Archean events in Antarctica. In: Yoshiha, Y., Katsutada, K., Kazuyuki, S. (Eds.), Recent Progress in Antarctic Earth Science. Tokyo Terra Scient. Publ., pp. 1–7. Cooper, A.K., Barrett, P.J., Stagg, H., Storey, B.D., Stump, E., Wise, W., the 10th ISAES editorial team, (Eds.), 2007. Antarctica in a Changing World. Proceedings of the 10th International Symposium on Antarctic Earth Sciences. The National Academy Press, Washington D.C, Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps p. 150. Dalziel, I.W.D., 1992. Antarctica: a tale of Two Supercontinents? Annu. Rev. Earth Planet. Sci. 20, 501–526. Fu¨tterer, D.K., Damaske, D., Kleinschmidt, G., Miller, H., Tessensohn, F. (Eds.), 2006. Antarctica. Contributions to Global Earth Science. Springer, p. 478. Jacobs, J., Klemd, R., Fanning, C.M., Bauer, W., Colombo, F., 2003. Extensional collapse of the late Neoproterozoic-early Paleozoic East African-Antarctic Orogen in Central DronningMaud Land, East Antarctica. In: Yoshida, M., Windley, B.F., Dasgupta, S. (Eds.), Proterozoic East Gondwana Assembly and Breakup. Geological Society London Special Publication 206, pp. 271–287. Leat, P.T., Dean, A.A., Millar, I.L., Kelley, S.P., Vaughan, A.P.M., Riley, T.R., 2005. Lithospheric mantle domains beneath Antarctica. In: Vaughan, A., Leat, P.T., Pankhurst, R.J. (Eds.), Terrane Processes at the margins of Gondwana. Geological Society London Special Publication 246, 359–380. Satish-Kumar, M., Motoyoshi, Y., Osanai, Y., Hiroi, Y., Shiraishi, K. (Eds.), 2008. Geodynamic Evolution of East Antarctica: A key to the EastWest Gondwana Connection, Geological Society London Special Publication 308, 465 pp. St. John, B. (Ed.), 1991. Antarctica as an Exploration Frontier—Hydrocarbon Potential, Geology and Hazards. American Association Petroleum Geologists, Studies in Geology 31, p. 154. Stonehouse, B. (Ed.), 2002. Encyclopedia of Antarctica and the SouthernOceans. J.CWiley, p. 404. Storey, B.C., 2005. Antarctic. In: Selley, R.C., Cocks, L.R.M., Plimer, I.R. (Eds.), Encyclopedia of Geology, vol. 1. 132–140. Tingey, R.J., 1991. The regional geology of Archean and Proterozoic rocks in Antarctica. In: Tingey, R.J. (Ed.), The Geology of Antarctica. Clarendon Press, Oxford, pp. 1–58. Tingey, R.J. (Ed.), 1992. The Geology of Antarctica. Monographs on Geology and Geophysics (no 17). Clarendon Press, Oxford, UK, p. 680. Torsvik, T.H., Gaina, C., Redfield, T.F., 2007. Antarctica and Global Paleogeography: From, Through Gondwanaland and Pangea, to the Birth of the Southern Ocean and the Opening Gateways. In: Cooper, A.K., Barrett, P.J., Stagg, H., Storey, B., Stump, E., Wise, W., the 10th ISAES editorial team, (Eds.), Antarctica Keystone in a Changing World. Proceedings of the 10th International Symposium on Antarctic Earth Sciences. The National Academy Press, Washington D.C, pp. 125–140. Yoshida, M., Windley, B.F., Dasgupta, S. (Eds.), 2003a. Preface in Proterozoic East Gondwana Assembly and Breakup, Geological Society London Special Publication 206, pp. vii–x. Yoshida, M., Jacobs, J., Santosh, M., Rajeh, H.M., 2003b. Role of Panafrican events in the Circum-East Antarctic Orogen of Gondwana. In: Yoshida, M., Windley, B.F., Dasgupta, S. (Eds.), ProterozoicEastGondwanaAssemblyandBreakup,GeologicalSocietyLondonSpecialPublication 206, 57–75. Yoshida, Y., Katsutada, K., Kazuyuki, S. (Eds.), 1992. Recent Progress in Antarctic Earth Science. Terra Scientific Publication, p. 796. —also: http://www.terrapub.co.jp/e-library/aes/index.html. 1041 2255..77 1042 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Cenozoic/Mesozoic and Paleozoic orogenic systems and their fold and thrust belts (FTBs) Orogeny versus epeirogeny From the mid-nineteenth century onwards, the physiographic and the tectonic implications of the term “orogeny” were never clearly separated and the ambiguity of the term still persists today. Orogens involve compressional and transpressional deformation of sediments and their underlying basement. Folds, thrusts, as well as strike-slip faults dominate but normal faults do also occur. The accretion of various terranes, regionalmetamorphism (associatedwith the rise of metamorphic core complexes), and a host of differing igneous processes associated with mountain building have become topics of considerable research interest. The terms “orogenic” or “folded” belts are here used loosely, and inter- changeably, for “subduction-relatedmountain systems.” Formore, see Chapter: 6 Vol. 1A (Roeder). Gilbert (1890) proposed the term “epeirogeny” for the processes responsible for broad uplifts and depressions that affect large regions of the continents. Orogenic structures are often contrasted with regional “epeirogenic” uplifts and subsi- dence. Plate 25.1 displays plateau-like uplifts of some subduction-related/oro- genic mountains systems. However, other “long-wavelength uplifts” (Sengo¨r, 2003) may also be related to Neogene and/or presently active continental rift sys- tems or else be associated with the rebound of formerly glaciated areas, The for- mation of the “African Erosion surface” (Burke and Gunnell, 2008) is another example of a complex “epeirogenic” uplift history. There is no single explanation for “epeirogenic/long-wavelength uplifts,” just as much as there is no single cause for the origin of various sedimentary basins and the broad arches that separate them. Epeirogenic and orogenic processes often overlap. Orogens contrast with the relatively more stable continental cratons that typically formed and survived over hundreds of million years. A random surface traverse across any orogenic (folded) belts will not reveal rigid plate characteristics to the observers. Today, over wide areas, Cenozoic/Mesozoic orogenic belts coincidewithNeotectonic/Neogene diffuse plate boundaries (Plate 25.9). The evolution of Phanerozoic orogens indicates that wide, ever changing diffuse plate boundaries may have evolved over hundreds of millions years. Observations of neotectonic processes and process-oriented models are all quite informative. However, to quote Dewey (2007) “Ultimately the current, blinkered obsession with process must be tempered by the historical perspective of the record preserved in rocks.” That historical perspective includes the long time spans paleogeographic and tectonic evolutions. Ophiolites, ophiolitic me´langes, and Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps remnants of accretionary wedges have all been associatedwith suturing (formore, see Section 25.10: Mafic/ultramafic crust, etc., with examples of equivalent allochthonous fragments in orogens, Plate 25.26 and Table 25.7). Regional/supraregional scale sutures are shown as lines on many maps (e.g., Pubellier, 2008; Sengo¨r and Natalin, 1996). The Atlas of Central Asia by Daukeev of preorogenic, “ocean-opening” stages, followed by “ocean-closing” synoro- genic and postorogenic stages. Finally, high-relief mountain ranges will be pene- planed reducing them to cratonic shields and their subsurface equivalents, that is, the basements of sedimentary basins. It is sensible to visualize the evolution of any orogenic region through a sequence of periodically changing tectonic regimes, often also overprinted by climatic changes and their consequences. Subduction, sutures, and orogens (Plates 25.17 and 25.18) Roeder (2009, see p. 6–8,) reviewed the origin of the term subduction, which was introduced in the early 1970s (see Chapter 6, Vol. 1A (Roeder)). In an attempt to define boundaries of major orogens, Bally and Snelson (1980) differ- entiated B–Benioff subduction (i.e., oceanic subduction) from A-Ampferer subduction (i.e., continental subduction) as the boundary of their “megasuture.” Because at the time continental subduction was not widely accepted, these authors did follow the general practice to name these poorly understood bound- aries after well-known scientists. However, the main purpose of naming these was to define the boundaries of orogens, that is, the “megasuture” of these authors. Today continental subduction is accepted (e.g., Ernst et al., 1997; Ernst 2001; Lallemand, 1999 and Lallemand and Funicello 2009) and the terms “A- and B-subduction” and “megasuture” are now redundant and may best be dropped. The legends of Plates 25.7, 25.14, and25.20 refer to oceanic versus continental sub- duction boundaries with appropriate symbols. The complex relations of absolute platemotions, obliquity of convergence, slab dip, types of crust, etc., were reviewed by Lallemand et al. (2005, 2008). FFTBs are here used to draw outer boundaries of continent-verging orogens. However, the relationship of FFTBs to variousmetamor- phic structures within the internal zones of orogens including the ones revealed by exhumed metamorphic core complexes still needs further clarification. Orogenic systems do record the complex dynamic processes at compressional plate boundaries that welded (sutured) large or smaller continental blocks, as well as island-arc segments. The term suture is used widely, but not consistently, for major narrow zones often within folded belts (e.g., Dennis and Atwater, 1974) that are interpreted as the remnants of former oceans or their margins that separated continental crustal terranes with significantly different pre-collisional 1043 et al. (2002) illustrates the relevance of using sutures to subdivide the area 1044 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps in regional tectonic/stratigraphic provinces in support of supra-regional plate tectonic reconstructions. Four boundary types of orogenic system boundaries are differentiated: (1) Oceanic (former B-type of Bally and Snelson, 1980) subduction boundaries associated with subduction of oceanic lithosphere and with deep-sea trenches (2) Transform and/or strike-slip dominated boundaries (3) Continental (former A-type of Bally and Snelson, 1980) subduction bound- aries that suggest limited subduction of continental lithosphere. These boundaries coincide with the boundary separating FFTBs from the relatively less deformed, foreland basin and its underlying basement. (4) In Central Asia, an ill-defined boundary characterized by a diffuse envelope aroundmostlyMesozoic intrusions and volcanics that is also characterized by complex basement-involved intra-plate deformation Cawood et al. (2009) differentiate accretionary orogens and collisional orogens and further subdivide accretionary orogens into two: (1) Retreating orogens characterized by slab/trench roll-back relative to the overriding plate and associated backarc basins (e.g., present W. Pacific and its Island arc systems) and (2) Advancing orogens associated with compression and transpression (e.g., present Eastern Pacific and Andean andNorth American Cordilleras) (see also Lallemand et al., 2005, 2008). Condie (2007) prepared a global list of accreted terranes and superterranes. He differentiates simple accretionary orogens involving accretion of juvenile island arc and/or oceanic plateaus from complex accretionary orogens in- volving collisions of formerly rifted margins, or else of arc systems and/or exotic cratons. Metamorphism and associated subduction processes, like subduction-related igneous processes and metamorphism, are key aspects for a more complete understanding of orogenesis. However, the global distribution of various meta- morphic complexes and igneous intrusions are beyond the scope and resolution of the global maps introduced herein. For overviews on the subject, metamorphic patterns and their tectonic setting, see Brown (2010). Holdsworth et al. (2001) differentiate “reactivation,” that is, the rejuvenation of discrete structures from “reworking,” that is, repeated metamorphism, defor- mation, and magmatism of “previously tectonized crustal or lithospheric volumes.” Maps of radiometrically dated, reworked “basement terranes” help to untangle often thermotectonically reworked units, that are key elements of complex orogenic collages outcropping in mountains and on cratonic shields and also buried as the basements of sedimentary basins. The anatomy of accretionary orogens documents the progressive assembly and the subsequent Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps breakup of former supercontinents such as Columbia, Rodinia, Gondwana, and Pangea. Continental orogens range from the relatively simple continent/continent colli- sion of the Pyrenees to the more complex Alpine collision and a great variety of accretionary orogens, for example, the Himalayas and the Tibetan plateau, the Andes of South America, and the Western Cordillera of North America. The great variety of neotectonic orogenic settings (e.g., Plate 25.11, Table 25.2) and asso- ciated tectonic styles suggests different genetic configurations for each orogen. Presumably, Paleozoic and Precambrian orogens formed in a similar variety of tec- tonic settings, whereby outcropping older orogens today typically expose deeper levels of crustal deformation. Following the fine review by Miller and Snoke (2009), Phanerozoic crustal cross-sections call for a synthesis of deep seismic-reflection profiles, lab-based seismic velocity determinations, studies of xenoliths, and of exposed crustal cross-sections, These authors conclude “Perhaps the most apparent issue is that there is no “typical” cross-section of a continental crust” and emphasize polyphase magmatic and crustal histories that span over 10–100s of Ma followed by one or more extensional/transtensional events often associated with exhumation. Among causes of exhumation, Miller and Snoke (2009) list delamination of lithospheric mantle and in some cases of the lowermost crust leading to heating and uplift due to the rising warmer asthenosphere. As emphasized earlier, the vicissitudes of a polyphase crustal evolution are likely to vary from one plate tectonic setting to another. An idealized cartoon across only one type of folded belt (i.e., Cordilleran-type) merely illustrates a few selected critical elements of a Cenozoic/Mesozoic folded belts (Plate 25.17). On this diagram, the otherwise obvious importance of accreted terranes, metamorphic processes, and subsequent rise of core com- plexes and igneous intrusions are barely alluded to. “Ocean verging Exter- nides” are here contrasted with “Continent verging Externides.” In current jargon, the oceanward-verging Externides are called “active margins.” The bot- tom of the diagram also shows a continental mantle wedge overlying the sub- ducting slab. Plate 25.17 illustrates the idea that, with the possible exception of some igneous intrusions, all units of an orogen are likely to be allochthonous with respect to their oceanic or continental forelands. In other words, the orogen “floats” on a lower crust-mantle wedge (Laubscher, 1983; Oldow et al., 1991) that presumably was reworked/overprinted several times during its polyphase evolution. Among issues that need to be addressed are the age and nature of the under- lying continental foreland basement wedge that has been imaged by seismic reflection traverses across a number of orogens. This wedge often extends over 1045 hundreds of kilometers across strike below the deep internal parts of the 1046 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps orogen. For example, the LITHOPROBE seismic Atlas (e.g., LITHOPROBE, 2010a, 2011a,b) offers exquisite documentation of the deep reflection profiles acquired and interpreted in Canada (see also Chapter 7, Vol. 1A (Cook and Van der Velden)). Awestward thinning Precambrian eastern foreland basement wedge underlies the allochthonous terranes of theW. Canada Cordillera and can be followedwestward and across strike for well over 300 km (Cook et al., 1987; 1988; Hammer et al., 2011; LITHOPROBE, 2010a). The overlying westward thickening wedge of accreted allochthonous terranes implies that themantle that was originally under- lying these allochthonous terranes was detached and subducted and/or delami- nated prior to, during, and/or after their emplacement onto the buried “foreland basement wedge.” In a similar situation (i.e., the major tectonic units of the eastern of the Appalachians), Cook and Vadsudevan (2006) conclude that “terranes are detached thin flakes,” that is, the terranes of today’s accretionary orogens may best be viewed as lithospheric fragments or “flakes.” This example and many others confirm that the surface age of an orogen often may differ substantially from the age of crust/mantle boundary and its underlying mantle including some of the following possibilities: (1) the age of the underlying older foreland basement (2) the age of a plume-related uplift that preceded rifting and/or later volcanism prior to formation of a preorogenic passive margin, and (3) the age of late- to postorogenic extension/transtension and associated underplating. The theme of shingled “sequentially stacked orogens” has recently been docu- menteddramaticallywith a coast to coast “W-E lithospheric cross-section” that splices adjacent Canadian LITHOPROBE crustal profiles along distinct paleogeographic domains, or else as “suspect terranes” in the Western Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Cordillera (Coney et al., 1980; Monger et al., 1992; Schermer et al. 1984) and also tectonic units that were differentiated to unravel the complex collage of Asia (e.g., Pubellier, 2008; Sengo¨r and Natalin, 1996; Yin and Harrison, 2000). We now talk about “accretionary” orogens. Hatcher et al. (2007b), Cawood et al. (2009), and Murphy et al. (2009) assembled a number of papers that serve well as an introduction to significant recent work as well as current issues related to orogenesis. While acutely aware that all these studies are essential for complete understanding of any region, the scale of our global maps does not permit detailed portrayal of the major tectonic elements of orogens. The principal rea- son for differentiating only major orogenic/basement domains on the global maps was to greatly simplify the inherently complex regional geology of orogens to facilitate communication and at same time serve as background for our basin maps (Plates 25.31–25.36). FTBs, that is, the Externides, when compared to “Internides,” have gained perhaps disproportionate attention in the past because, though still complex, FTBs may have been relatively easier to unravel. The emergence in the 1860s of the anticlinal theory for petroleum accumulations by Sterry Hunt combined with tempting oil seeps associated with FTB anticlines led to these becoming a favorite target for worldwide hydrocarbon exploration. Today, many FTBs are covered by high- quality 2D and 3D reflection seismic profiles that together with extensive drilling have provided some of the best-documented subsurface images of many FTBs. However, for some of them (e.g., onshore California, the Zagros) good quality long regional seismic profiles have yet to be published. Overall ocean-verging FTBs are here called active margin fold and thrust belts (AMFTBs). They contrast with the continental, overall cratonward-verging FFTBs. Table 25.3 classifies many FTBs and cites numerous examples. The 30þ classes and subclasses listed here are testimony to the relative complexity of FTBs, which is principally due to the following factors: (1) Variations among basic plate tectonic orogenic settings due to differing directions of far field stresses (e.g., Plate 25.11 and Table 25.2) (2) Subducting oceanic crust (Plate 25.27) as opposed to subduction/under- thrusting by a continental foreland basement. In both oceanic and continen- tal “forelands,” basement trends and also overlying stratigraphic trends often strike discordantly and at a considerable angle to the strike of the FTB (see Plates 25.20 and 25.27). Thus, re-activation of former structural trends is bound to be highly selective and needs careful documentation. (3) Complex “mechanical stratigraphies,” that is, the number and types of de´collement/detachment zones may change either gradually or occasionally abruptly along the strike of an FTB leading in turn to lateral changes in struc- tural styles. In the case of continental basements, the nature and mechanical properties of the basement layers are still important. 1047 Table 25.3 A classification of fold and thrust belts (FTBs) with examples 1. Active margin fold and thrust belts (AMFTBs), directly associated with active margins and intra-oceanic subduction 1.1. Accretionary wedges, overlying subducting oceanic crust 1.1.1. Involving thick sediments of deep sea fans For example, Arakan/Cachar/Tripura, Bangladesh, Eastern India; onshore/offshore Makran (Iran, Pakistan); Barbados ridge; Offshore Magdalena (Colombia) 1.1.2. Involving relatively thinner sediments dominated by deep-sea mudstones and turbidites, that is, the majority of all accretionary wedges For example, S. Java–SW. Sumatra; offshore Washington/Oregon; Vancouver island offshore; Nankai (Japan) 1.1.3. In addition to the above, also involving oceanic crust For example, Andaman onshore/offshore (India) Notes: (1) All of the above are not easily differentiated on small-scale global maps (2) This list excludes spectacular examples of exhumed/outcropping relatively pristine accretionary wedges such as the Oman mountains because they are no longer active and they are assigned to the FFTB class (see below) 1.2. Compressional/transpressional folds and thrusts in forearc basins For example, Sacramento/San Joaquin Basins of California; Cook Inlet/Alaska; Sumatra Forearc Basins Note: Smaller forearc basins of this type are not easily differentiated on a small-scale map 1.3. Transpressional fold and thrusts in backarc domains 1.3.1. Dominated by de´collement tectonics For example, Taiwan 1.3.2. Basement-involved inversion of earlier extensional systems, that is, compressional/transpressional inversions of extensional grabens systems, associated with active margin–related episutural basins For example, Natuna area, Indonesia; Bali basin—Indonesia 1.4. Folded belts associated with transtensional/transpressional strike-slip systems associated with active margin-related basins For example, Los Angeles, Ventura, Santa Maria Basins of California Note: Small basins of this type are often not easily differentiated on small-scale maps 1.5. Interior AMFTBs: Folds associated with former active margin basins in the interior of older folded belts; deformation of episutural basins in the interior of folded belts 1.5.1. Dominantly compressional and transpressional For example, Mesozoic, Bowser basin, British Columbia; Upper Paleozoic Gulf of St Lawrence; Upper Paleozoic/Mesozoic/ Cenozoic Sverdrup Basin 1.5.2. Dominantly extensional/transtensional, following compressional phase For example, US Basin and Range Note: AMFTBs overlap with the episutural basins of the basin classification scheme. Subsidence and infilling of these basins are intimately related to structures. Compressional/transpressional and extensional/transtensional structures are common. 2. Foreland fold and thrust belts (FFTBs) associated with now mostly inactive Mz/Cz and Pz continental subduction boundaries dominated by de´collement tectonics 2.1. FFTBs dominated by sediment de´collement over monoclinal basement ramp 2.1.1. Dominated by flexural slip folding involving one basal and several subsidiary de´collement levels 2.1.1.1. Buried/covered and discordantly overlain by linked extensional/compressional gravitational growth faults system For example, Campeche–Reforma (Mexico); offshore Trinidad/Columbus Channel 2.1.1.2. Partially buried under onlapping foreland basin deposits For example, Mesopotamian Zagros foreland 2.1.1.3. Uplifted/eroded For example, Jura Mountains; Mackenzie Mountains (NWT); Melville Island, Canadian Arctic Islands; onshore Zagros; Sichuan basin (China); Appalachian valley and Ridge FFTB 2.1.2. Dominated by imbricates associated with one basal and/or several subsidiary de´collement levels 2.1.2.1. Buried, covered, or partially covered For example, Southern and Northern Po plain in Italy; Faja Sepultada/Zongolica fold belt in Mexico 2.1.2.2. Uplifted/eroded For example, Alberta Montana FFTB; Wyoming FFTB; Sub-Andean FFTB (Bolivia); Naga–Schuppenbelt (Assam) 2.1.3. Dominated by imbricates associated with multiple de´collement levels 2.1.3.1. Buried/covered For example, Faja Sepultada of W. Veracruz basin 2.1.3.2. Uplifted/eroded For example, East Venezuela, Maturin/Furial FFTB 1048 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Table 25.3 A classification of fold and thrust belts (FTBs) with examples—cont’d 2.1.4. Dominated by frontal triangle zones 2.1.4.1. Mostly buried/or mildly uplifted For example, Alberta foothills, North Caucasus FFTB 2.1.4.2. Uplifted partially or completely eroded For example, The Sub-Alpine Molasse of Switzerland and Bavaria; hypothetical in some FFTBs, for example, Foothills of Western Canada, B.C. Note: The term “Triangle Zone” originated in the mid-1950s as an informal term for a boundary zone with the foreland basin of Alberta, which on seismic profiles had a triangular appearance, that is, a special case of wedging of conjugate reverse faults that is common in many types of FFTBs. 2.1.5. Dominated by extensive de´collement nappes and thrust fault systems For example, Helvetic Nappes of Switzerland Notes: (1) FFTBs are also commonly subdivided into shale-based (e.g., Alberta FFTB), evaporite-based (e.g., Northern Apennines, Melville Island, Zagros FFTBs), Jura Mountains, or other de´collement levels. (2) In addition, it may be useful to differentiate de´collement levels associated with high pore pressures from those that do not appear to be overpressured. The differentiation may be complex because often in FFTBs high pore pressures are plausibly hypothesized but no longer preserved and observed in wells. (3) In many FFTBs, there is vigorous debate as to whether “thick-skinned” (i.e., basement-involved) or “thin-skinned” (i.e., de´collement of sediments) is the dominant structural theme. Both tectonic styles are known to occur side-by-side within the same FFTB (e.g., South Pyrenean FFTB of Spain; Kucha FFTB ofWestern China). Ultimately, the resolution of the problem is a matter of reflection seismic resolution and drilling. “Thin-skinned” de´collement tectonic interpretations often allow for a larger number of structural prospects, while basement-involved tectonics tends to limit the number of prospects to a few larger structures. 2.2. Dominated by basement-involved structures and uplifts 2.2.1. Basement-involved structures and uplifts in the foreland For example, Wyoming/Colorado Rocky Mountains; Merida Andes; Bachu uplift, W. Tarim basin; Junggar basin 2.2.2. Dominated by basement-involved structures and uplifts in the nearby hinterland For example, Alpine “Autochthonous”/Central massifs, Southern Alps 2.2.3. Basement-involved compressional structures associated with major strike-slip fault systems For example, Chaidam basin (Western China) 2.2.4. Far-traveled allochthonous basement thrusts acting as “Backstops” for FFTBs For example, Eastern Alpine Thrust sheets; Daciannappes of theNorthern and Eastern Carpathians; Blue Ridge of the Appalachians Note: Many authors consider far-traveled allochthonous basement thrust sheets to be a part of the internal folded belts (Internides). However, different types of metamorphic core complexes that rise from the lower crust domain of the orogen to the surface are also part of the “Internides.” Metamorphic core complexes often are associated in an overall context of late orogenic extensional/transtensional tectonics or else in a purely extensional rifting context that affects a mature, already peneplaned, folded belt. While the hydrocarbon potential of the inner orogenic zones appears to be limited, the specifics of each case do matter; for example, sediments overlying the upper Eastern Alpine Nappes often appear to be less metamorphosed and often contain viable source beds, for example, Triassic source beds of the Southern Alps and possible Mesozoic source beds underlying the Tertiary Fill of the Pannonian and Transylvanian basins. In simple words: the higher units (“backstops” ) of FFTBs “float” on top of the FFTBwhile the frontal units get progressively more buried as they are “tucked” under the backstop. 3. Basement-involved inversion of earlier extensional systems (i.e., compressional/transpressional inversion of extensional graben systems) 3.1. Located close to front of de´collement FFTBs For example, Central and Eastern High and Middle Atlas (Morocco): Uinta Mountains of Utah; Richardson Mountains of N. Yukon 3.2. Associated with distal foreland of FFTBs For example, Central Europe/North Sea Inversions 1049 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Active margin fold and thrust belts 1050 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps (AMFTBs) AMFTBs are best subdivided, into accretionary wedge systems (i.e., 1.1 on Table 25.3), as well as FTBs that deform adjacent forearc and backarc basement and their overlying sediments (i.e., 1.2 on Table 25.3). Common strike-slip related tectonics and associated transpressional and transtensional, as well as inversion, tectonics are included in this group. AMFTBs and their associated Mesozoic/Tertiary basins are limited to the Circum- Pacific and South East Asia active margins and range from intra-oceanic arcs to continental volcanic arcs systems. Forearc basins are located on the ocean- ward side of the often still active arcs or else the deactivated batholithic roots of former arcs. Backarc basins are located behind the arcs. Both basin types are typically underlain by peneplaned, sometimes rifted, basements that merge with outcropping igneous rocks of the arc or else underlie them. Both forearc and backarc basins are members of the episutural basin class which is discussed in Chapter 4, Vol. 1A (Roberts and Bally) and Tables 25.2 and 25.3 therein. Most AMFTBs differ greatly from FFTBs because they involve relatively “young basements” formed during a continuing Mesozoic/Cenozoic orogenic process. Associated episutural basins and their tectono-stratigraphic megasequences also have sediment provenances differing drastically from their continent-verging FFTBs counterparts. AMTFBs often involve the continental forearc or else backarc basements, that is, peneplaned earlier continental arcs systems, for example, the Sierran or Franciscan basement of some Californian basins. Part of this older arc- related basement may also outcrop in adjacent mountains. Igneous continental arcs may intrude earlier, that is, much older basement. Following our respective definitions of orogens and basements, older reworked basement would be dated by the latest regional igneous andmetamorphic reworking episode (e.g., Plate 20 Continental basement map). AMFTBs by definition coincide with Cenozoic/ Mesozoic active margins. For some years Von Huene and Scholl (1991) and Scholl and von Huene (2007) have differentiated accreting margins from non-accreting (eroding) margins whereby accretionary wedges are representative of FTBs overriding the subducting slab adjacent to deep sea trenches. These appear as wedges of folds and/or faulted, imbricated oceanic sediments that in a broad sense may be viewed as “conjugate” to continental FFTB wedges (Plates 25.17 and 25.18). Seismic atlases that include many of examples from active margins around the world were published by Bally (1983) and VonHuene (1987). In recent years, many more detailed studies have been published. With the important exception of ocean-verging accretionary wedges, AMFTBs typically correspond to deformed forearc-, backarc-, and also strike-slip-related basin margins that in the literature Cenozoic/Mesozoic) and grey (for Paleozoic) horizontal lines. The term “thick Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps skinned” has traditionally been used for basement-involved compressional structures. Table 25.3 differentiates various types of FFTBs. In reality, however, no two FFTBs are alike and any attempt to pigeonhole FFTBs on a table is bound to raise eyebrows. The fundamentalist “thin-skinned versus-thick skinned FFTB” debate has are commonly viewed as sedimentary basins and not as folded belts. Thus, the many compressional/transpressional fold and thrust systems associated with the basins of California are commonly presented in the overall context of the evolution of their sedimentary basins. For a more detailed list of AMFTB examples, see Table 25.3 and also page 1090 etc: Vol. 1C : Basins locatedwithin orogenic belts (episutural basins), Vol. 1C. Foreland fold and thrust belts (FFTBs) FFTBs (Plates 25.17–25.19) are here defined as continental foreland-verging fold and thrust systems. This continental “cratonic” foreland is characterized by a com- plex basement evolution that includes a peneplaned collage of various earlier oro- gens ranging from the Archean to the Upper Paleozoic. FFTBs involve the deformation of wide areas that were initially part of the adjacent foreland basin and its underlying “cratonic platform cover.” Frequently that “cratonic platform” cover was formed in the proximal shallow water shelf domains of a former passive margin. The distal/oceanward sediments of these passive margins are now exposed in the adjacent FFTBs. Away and updip from the outer FFTB boundary (i.e., the “continental subduction boundary” of Plates 25.17–25.19), the Precambrian or Paleozoic basement of the FFTB, the foreland basin with its underlying platform sequence may eventually outcrop at the updip termination of the foreland basin. Thus a Precambrian or else a Paleozoic basement becomes a Precambrian shield (e.g., the Canadian shield) or else a Paleozoic shield (e.g., the Bohemian craton). Seen in reverse order, the grad- ual progression from a stable shield to a less stable foreland basement to the increasingly more deformed Internides of an orogen that overlie an attenuated foreland basement wedge marks a continuous but steady destabilization of the craton. Where to draw the FFTB–Foreland basin boundary often reflects local practice. Thus, in some regions relatively minor folds are included as part of the fore- land basin while in other regions similar folds are viewed as part of the foreland folded belts. Precambrian basement (e.g., the Colorado/Wyoming Rocky Mountains) or else former Paleozoic basements (e.g., the Tien Shan, the Kunlun Shan of Central Asia, or the Sierra Pampeana of Argentina) are involved in “thick skinned” uplifts in the foreland. On our maps, these uplifts are differentiated by patterns of blue (for 1051 run its course. Many FFTBs do indeed display both thin-skinned (i.e., no basement 1052 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps involvement) as well as thick-skinned (i.e., basement-involved) structures occurring side by side. For some years now, high quality regional seismic reflection profiles have allowed a differentiation of detachment structures that involve only sediments from basement-involved structures. In the absence of adequate seismic images of the basement top and without a clear geological and geophysical definition of the basement, speculation that the basement may be involved will tend to discour- age further seismic exploration for hydrocarbons, whereas interpretations that suggest a number of de´collement levels and the possibility of duplex structures with multiple objectives may encourage additional seismic surveys. The reduction of structural terminology to a number of deformation styles, for example, detachment folds, fault bend, fault-propagation folds, and various types of triangle/wedge structures greatly facilitates a basic understanding of the struc- tural themes in FFTBs (and also AMTBs). The outstanding Atlas by Shaw et al. (2005) serves as guide for the seismic interpretation of contractional fault-related folds pioneered by Suppe (1983, 1985). To illustrate the regional context of FFTBs, long regional profiles that are based on the best possible seismic data will remain an important key. The reduction of a seis- mic profile registered in two-way time to a depth-converted and not-exaggerated profile that is properly tied to FFTB outcrops is a complex process. As with all regional seismic transects, the principle challenge is to reduce long regional pro- files to realistic, hopefully appealing graphics that permit the reader to easily sep- arate uninterpreted seismic profiles from interpretations. Given the best possible seismic information, properly “balanced” regional cross needs to be firmly anchored in the undeformed, “fixed” foreland. Ideally, one should strive for 3D balancing, which would also account for large- scale strain partitioning associated with strike slip and/or extensional/transten- sional faulting thatmay ormay not be coeval with the deformation of the fold belt. These and many more demands are not easily met. Given all the uncertainties associated with interpretation of seismic profiles as well as the graphic limitations due to the reduction of profiles that are hundreds of kilometers in length to aman- ageable format for publications, surface geological maps will provide at least part of the third dimension to complement 2D seismic transects. Published balanced cross-sections are far from being unique interpretations and often are one of several possible interpretations. Different “balanced cross section across the same area reflect the views of individuals, schools and often also of graphic preferences. Significant weaknesses often are as follows: (1) The absence of published supporting seismic data (2) Given seismic reflection profiles in time, approximate seismic interval veloci- ties used for depth conversions (3) The absence of a “fixed anchor” in the foreland (4) The absence of a clear definition of the underlying basement Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps (5) The lack of a lithostratigraphic column subdivided into tectono-stratigraphic megasequences, their thickness ranges, and a column that shows (and comments on) the mechanical stratigraphy and associated de´collement levels (6) The use of unsupported constant thicknesses for stratigraphic intervals and no indication of thickness ranges Instead of unique solutions, there is merit in proposing alternative interpretations that explore different options. This throwback to the oldmultiple working hypoth- esis often helps in planning for the acquisition of information in key areas of an FFTB. Examples of multiple interpretations that are “balanced” are discussed by Ghisetti et al. (1993) and Hung (2005). Based on observations made in a tunnel project, Buxtorf (1916) first proposed evaporite based de´collement FTBs. Sommaruga (1999) provides a modern inter- pretation of the Central Jura fold belt based on reflection seismic profiles. The Jura is a few tens of kilometers wide and about 200 km. long. This contrasts with the Zagros, which is 200 km wide and >2000 km long or else the Parry Islands Fold Belt of the Canadian Arctic, which is about 150 km wide and >350 km long. In Table 25.3, these three detachment fold belts are lumped in one group, together with the Appalachian Valley and Ridge FFTB and the folds of the Sichuan basin. Of all these detachment fold belts, the Melville Island profiles of the Parry Island Fold Belt offer by far the best published seismically documented example that is also supported by a superb surface geological map (see Fox, 1983; Harrison and Brent, 2005; Harrison, 1995; see Chapter 24, Vol. 1C (Harrison)). There are notable differences between the three above-listed fold belts; for exam- ple, the basement underlying the Jura is reasonably expected to be Paleozoic, the basement under the Zagros is likely to be Precambrian, but not mappable on seis- mic, and the basement underlying the Melville Island fold belt is separated from the base of the evaporitic de´collement level by over 15 km of partially rifted Upper Proterozoic sediments, that are revealed on seismic reflection profiles. Also parts of the Zagros are characterized by many salt diapirs while the absence of diapirs in the Jura and on the Melville island fold belt may in part be explained by the pet- rologic composition of the evaporitic de´collement layer. However, when com- pared with the Zagros, it is obviously also due to the overall smaller dimensions of their evaporite basins. The purpose of simple classifications such as Table 25.3 is not only to compare but also to contrast seemingly similar fold belts. The comparison of three relatively simple FTBs suggests that even when broad folds dominate, there is a need for regional state of the art seismic profiles down to depths that can image basement and hopefully even deeper reflectors. Only quality regional seismic profiles will allow reduction of the number of interpretative options in most, if not all FFTBs. Needless to say, by their very nature FTBs will always present significant 1053 challenges to the acquisition logistics, processing, and interpretations of reflection seismic profiles. and thrust belts (FFTBs) FFTBs are dominated by compressional and transpressional structures during 1054 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps most of their evolution. Table 25.4 is a reminder that extensional structures are also associated with FFTBs and lists some of the more obvious examples of preorogenic rifts and their selective inversion within the FFTB or else in the foreland of the FFTB. Our examples are limited to inversions that are still judged to fall within the boundaries of the FFTB and exclude more distal inversions such as the Alpine inversions so well mapped and documented by Cooper and Williams (1989), Ziegler (1990), Ziegler et al. (1995, 2002), and many other authors. This arbitrary exclusion is matter of personal judgment and convention. It simply wouldnotmakemuch sense to includemostofNorthernEuropeaspart of theAlpine systemeven though the impact of theAlpine collision goeswell beyond theAlps and their immediate foreland. Also excluded from the list are extensional and transtensional tectonics associated with metamorphic core complexes, that is, low angle detachment faults systems overlying metamorphic core complexes. The presence of low angle detachment faults within the oceanic crust on the distal offshore parts of passivemargins and in many different rifts systems requires a better overview over all these systems and yet another global map. A list of seismic atlases with useful FTB profiles includes Bally (1983), Gries and Dyer (1985), Shaw et al. (2004), and the Seismic Atlas of SE-Asian Basins (2011). Volumes that review examples of a great variety of FFTBs include Letouzey (1990), McClay (1992a,b), Mitra and Fisher (1992), Nemcock et al. (2005), Ries et al. (2007), McClay (2003), Mazzoli and Butler (2006), Lacombe et al. (2007), Lallemand and Funicello (2009), and Roeder (2009). Updates on FFTB hydrocarbon prospectivity include Goffey et al. (2010) and Roeder (2010). Cooper (2007) emphasizes that hydrocarbon prospectivity in FFTBs is not related to structure alone. Equally important are the presence of source beds, the timing and degree of their maturation with respect to the timing of deformation in the fold belt, and whether the folds are buried or else uplifted and partially eroded. It was not the presence of anticlines alone that attracted early explorers to FFTBs. Instead it was the presence of surface seeps combined with structures that attracted early exploration in FTBs and also anticlines/domal structures outside FTBs, for example, salt domes in the Gulf of Mexico coastal plain. Normal faulting in foreland fold Table 25.4 Normal faulting in foreland fold and thrust belts (FFTBs) Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps 1. Pre-orogenic normal faulting associated with rifting event prior to deposition of the platform/proximal passive margin TSMs 1.1. Preserved and not inverted in the foreland of the FFTB For example, Devonian, Alaska North Slope 1.2. Involved and often inverted in FFTB For example, High and Middle Atlas of Morocco; ? Jurassic of Caucasus; Lower Paleozoic of Wichita Mountains; Uinta Mountains of U.S. Rocky Mountains; Richardson Mountains, NW Territories Canada; Uppermost Triassic/Jurassic Southern Alps 1.3 Rift fill selectively detached from underlying rift For example, Rides Pre-Rifaines (Northern Morocco); Dnepr-Donets basin (Ukraine) 2. Preorogenic, passive Margin, extensional salt structures, growth faults For example, Reforma/Campeche area (SE Mexico); (?) Central Apennines of Umbria. 3. Syn-orogenic normal faulting associated with flexure of “subducted” continental foreland. For example, NW—Australia/Timor; Molasse Basin of Bavaria; Outer Carpathians (Poland, Ukraine); Ouachita foreland/Arkoma basin (U.S.); St. Lawrence Lowlands (Quebec) References to 25.7Cenozoic/Mesozoic and Paleozoic Orogenic Systems and Their Fold and Thrust Belts (FTBs) Orogeny Versus Epeirogeny; Subduction, Sutures, and Orogens (Plates 25.17 and 25.18); Active Margin Fold and Thrust Belts (AMFTBs); Foreland Fold and Thrust Belts (FFTB’s) (Plates 25.17–25.18); Normal Faulting in Foreland Fold and Thrust Belts (FFTBs) Bally, A.W. (Ed.), 1983. Seismic Expression of Structural Styles—A Picture and Work Atlas, vol. III. Tectonics of Compressional Provinces/Strike-Slip Tectonics American Association Petroleum Geologists. Studies in Geology Series 15, vol. 3. p. 329. 4. Minor normal faults associated with anticlinal crests (“extra dos” normal faults of French authors) For example, Associated with stretching along the crest of individual anticlines 5. Late orogenic to post-orogenic extensional/listric or transtensional normal faults 5.1. Late orogenic to post-orogenic extensional or transtensional normal fault systems associated with uplift of orogenic system and/or the rise of associated metamorphic core complexes For example, Rocky Mountains of S.E. B.C., Canada; Basin and Range of Western U.S. and Northern Mexico; Gulf of Paria (Venezuela/Trinidad; Vienna Basin; Central and southern Apennines (Italy); Triassic rifts of Appalachians). 5.2. Late orogenic to post-orogenic linked extensional-compressional gravitational systems located on top of (FFTBs) 5.2.1. Overall parallel (i.e., concordant with overall strike of FFTB) For example, Mexican Ridges, offshore Mexico 5.2.2. Trending at substantial to right angle to the underlying FFTB For example, Reforma/Offshore Campeche FFTB, Southern Trinidad Offshore 5.3. Associated with continental backarc basins For example, Transylvanian Basin, Gulf of Lyon (offshore S. France and Spain) 5.4. Associated with gravitational collapse of the accretionary wedge (extensional wedge-top basins) For example, Nappe Pre-rifaine, Onshore and Offshore Morocco; ?Onshore Makran (Iran) 1055 Bally, A.W., Snelson, S., 1980. Realms of subsidence. In: Miall, A. (Ed.), Facts and Principles of World PetroleumOccurrence, Canadian Society of PetroleumGeologistsMemoir 6, pp. 9–94. Brown, M., 2010. Metamorphic patterns in orogenic systems and the Geological record. In: 1056 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Cawood, P.A., Kro¨ner, A. (Eds.), Earth Accretionary Systems in Space and Time, Geological Society London Special Publication 318, 37–74. Burke, K.B., Gunnell, Y., 2008. The African erosion surface: a continental-scale synthesis of geo- morphology, tectonics and environmental change over the past 180 million years. Geol. Soc. Am. Memo. 201, 66. Buxtorf, A., 1916. Prognoses und Befunde beim Hauensteinbasis-und Grenchenburg-tunnel und die Bedeutung der letzteren fu¨r die Geologie des Juragebirges. Naturforschung Gesellschaft Basel Verhandlungen 27, 184–254. Cawood, P.A., Kro¨ner, A., Collins,W.J., Kusky, T.M.,Mooney,W.D., Windley, B.F., 2009. Accretion- ary orogens through Earth history. In: Cawood, P.A., Kro¨ner, A. (Eds.), Earth Accretionary Systems in Space and Time, Geological Society London Special Publication 318, pp. 1–36. Condie, K.C., 2007. Accretionary Orogens in space and time. In: Hatcher Jr., R.D., Carlson, M.P., McBride, J.H., Martı´nez Catala´n, J.R. (Eds.), 4-D Framework of Continental Crust, Geological Society of America Memoir 200, pp. 145–158. Coney, P.J., Jones, D.L.,Monger, J.W.H., 1980. Cordilleran suspect terranes. Nature 288, 329–333. doi:10.1038/288329a0. Cook, F.A., Simony, P.S., Coffin, K.C., Green, A.G., Milkereit, B., Price, R.A., et al., 1987. Lithop- robe southern Canadian Transect; Rocky Mountain thrust belt to Valhalla gneiss complex. Geophys. J. Roy. Astron. Soc. 89, 91–98. Cook, F.A., Green, A.G., Simony, P.S., Price, R.A., Parrish, R., Milkereit, B., et al., 1988. Lithoprobe seismic reflection structure of the southeastern Canadian Cordillera: Initial results. Tectonics 7, 157–180. Cook, F.A., Vadsudevan, K., 2006. Reprocessing and enhanced interpretations of the initial COCORP Southern Appalachians traverse. Tectonophysics 420, 161–174. Cooper, M., 2007. Structural style and hydrocarbon prospectivity in fold and thrust belts: a global review. In: Ries, A.C., Butler, R.W., Graham, R.H. (Eds.), Deformation of Continental Crust, Geological Society London Special Publication 272, 447–472. Cooper, M.A., Williams, G.D. (Eds.), 1989. Inversion Tectonics. Geol. Soc. Spec. Pub., 44. 375 pp. Daukeev, S.Zh. et al., (Ed.), 2002. Atlas of the Lithology-Paleogeographical, Structural, Palinspas- tic and Geo-environmental maps of Central Asia. YUGEO, Scientific Research Institute of Nat- ural Resources, Almaty, Kazakhstan, 37 maps and text. Dennis, J.G., Atwater, T., 1974. Terminology of geodynamics. Am. Assoc. Petrol. Geol. Bull. 58, 1031–1032. Dewey, J.F., 2007. The secular evolution of plate tectonics and the continental crust, an outline. In: Hatcher Jr., R.D., Carlson, M.P., McBride, J.H., Martı´nezCatala´n, J.R. (Eds.), 4-D Framework of Continental Crust, Geological Society of America Memoir 2001–8. Ernst, W.G., 2001. Subduction, ultrahigh pressure metamorphism and regurgitation of buoyant crustal slices-implications for arcs and continental growth. Phys. Earth Planetary Interiors, 127, 253–275. Ernst, W.G., Maryuama, S., Wallis, S.R., 1997. Byoyancy-driven, rapid exhumation of ultrahigh- pressure metamorphosed continental crust. Proc. Natl. Acad. Sci. USA, 94, 9532–9537. Fox, F.G., 1983. Structure sections across Parry Islands Fold Belt and Vesey Hamilton Salt Wall, Arc- tic Archipelago Canada. In: Bally, A.W. (Ed.), Seismic Expression of Structural Styles, American Association Petroleum Geologists 3. 3.4.1-54 to 3.4.1-62. Ghisetti, F., Barchi, M., Bally, A.W., Moretti, I., Vezzani, L., 1993. Conflicting balanced structural sections across the Central Apennines (Italy): Problems and implications. In: Spencer, A.M. (Ed.), Generation, Accumulation and Production of Europe’s Hydrocarbons III, Special publi- cation of the European Association of Petroleum Geoscientists No. 3. pp. 219–231. Gilbert,G.K., 1890. LakeBonneville.U.S.Geological Survey,WashingtonDC,Monograph, 1, p. 338. Gries, R.R., Dyer, R.C. (Eds.), 1985. Seismic Exploration of the Rocky Mountain Region. Rocky Mountain Association of Geologists and the Denver Geophysical Society, p. 300. Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Goffey, G.P., Craig, J., Needham, T., Scott, R., 2010. Fold thrust belts: overlooked provinces or justifiably avoided? In: Goffey, G.P., Craig, J., Needham, T., Scott, R. (Eds.), Hydrocarbons in Contractional Belts. Geological Society London Special Publication 348, pp. 1–6. Hammer, P.T.C., Clowes, R.M., Cook, F.A., Vadsudevan, K., van der Velden, A.J., 2011. The big picture: A lithospheric cross section of the North American continent. GSA Today 21, 4–10. Harrison, J.C., 1995. Melville Island’s salt-based fold belt, Arctic Canada. Geol. Surv. Canada Bull. 472, 331. Harrison, J.C., Brent, T.A., 2005. Basins and Fold Belts of Prince Patrick Island and adjacent areas, Canadian Arctic Islands. Geol. Surv. Canada Bull. 560, 197. Hatcher Jr., R.D., Bream, B.R., Merschat, A.J., 2007a. Tectonic map of the southern and Central Appalachians: a tale of three orogens and a complete Wilson cycle. In: Hatcher Jr., R.D., Carlson, M.P., McBride, J.H., Martı´nez Catala´n, J.R. (Eds.), 4-D Framework of Continental Crust, Geological Society of America Memoir 200, pp. 595–632. Hatcher Jr., R.D., Carlson, M.P., McBride, J.H., Catala´n, M.J.R. (Eds.), 2007b. 4-D Framework of Continental Crust, Geological Society of America Memoir 200, pp. 641. Holdsworth, R.E., Handa, M., Miller, J.A., Buick, C.K., 2001. Continental reactivation and reworking: an introduction. In: Miller, J.A., Holdsworth, R.E., Buick, I.S., Hand, M. (Eds.), Continental reactivation and reworking, Geological Society of London Special Publication 184, pp. 1–12. Hung, E.J., 2005. Thrust belt interpretation of the Serranı´a del Interior andMaturı´n subbasin, east- ern Venezuela. In: Ave´ Lallemant, H.G., Sisson, V.B. (Eds.), Caribbean-South American Plate Interactions, Venezuela, Geological Society of America Special Paper 394, 251–270. Lacombe, O., Lave´, J., Roure, F., Verge´s, G. (Eds.), 2007. Thrust Belts and Foreland Basins—From Fold Kinematics to Hydrocarbon Systems. Heidelberg Springer, Berlin, 491. Lallemand, S., 1999. La subduction oce´anique. Gordon and Breach Science Publishers, p. 194. Lallemand, S., Funicello, F. (Eds.), 2009. Subduction Zone Geodynamics. Springer, p. 276. Lallemand, S., Heuret, A., Boutellier, D., 2005. On the relationships between slab dip, back-arc stress, upper plate absolute motion, and crustal nature in subduction zones. Geochem. Geo- phys. Geosyst. 6, 18 pp, Q09006. doi:10.1029/2005GC000917. Lallemand, S., Heuret, A., Facenna, C., Funicello, F., 2008. Subduction dynamics as revealed by trench migration. Tectonics 27, TC3014. doi:10.1029/2007TC002212, 2008 PDF. Laubscher, H.P., 1983. Detachment, shear and compression in the Central Alps. Geol. Soc. Am. Bull. 96, 710–718. Letouzey, J. (Ed.), 1990. Petroleum and Tectonics in Mobile Belts 4th IFP Exploration and Produc- tion Research Conference, Bordeaux. E´ditions Technip, 224 pp. LITHOPROBE, 2010a. Seismic Atlas of Canada. SC Southern Cordillera Transect. www.litho. ucalgary.ca/atlas/atlas2.html. LITHOPROBE, 2011a. Seismic Atlas of Canada. SNORCLE Northern Cordillera. Transect www .litho.ucalgary.ca/atlas/atlas2.html. LITHOPROBE, 2011b. Seismic Atlas of Canada. AG Abiti-Grenville Transect. www.litho.ucalgary .ca/atlas/atlas2.html. Macqueen, R.W., Leckie, D.A. (Eds.), 1992. Foreland basins and Folded belts, AAPG Memoir 55, 460 pp. Mazzoli, S., Butler, R.W.H. (Eds.), 2006. Styles of Continental Contraction, Geological Society of America Special Paper 414, 184 pp. McClay, K.R. (Ed.), 1992a. Thrust Tectonics. Chapman and Hall, London, p. 447. 1057 McClay, K.R., 1992b. Glossary of thrust tectonic terms. In: McClay, K.R. (Ed.), Thrust Tectonics. Chapman and Hall, pp. 419–433. McClay, K.R. (Ed.), 2003. Thrust Tectonics and Hydrocarbon Systems, In: American Association of 1058 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Petroleum Geologists Memoir 82, 667 pp. Miller, R.B., Snoke, A.W., 2009. The utility of crustal cross sections in the analysis of orogenic processes in contrasting tectonic settings. In: Miller, R.B., Snoke, A.W. (Eds.), Crustal Cross Sections from the Western North American Cordillera and Elsewhere: Implications for Tec- tonic and Petrologic Processes, Geological Society of America Special Paper 456, pp. 1–38. Mitra, S., Fisher, G.W. (Eds.), 1992. Structural Geology of Fold and Thrust Belts. Johns Hopkins University Press, p. 254. Monger, J.W.H., Souther, J.G., Gabrielse, H., 1992. Evolution of the Canadian Cordillera: a plate tectonic model. Am. J. Sci. 272, 577–602. Murphy, J.B., Keppie, J.D., Hynes, A.J. (Eds.), 2009. Ancient Orogens and Modern Analogues, Geological Society London Special Publication 327, 481 pp. Nemcock, M., Schamel, S., Gayer, R., 2005. Thrust Belts; Structural Architecture, Thermal Regimes and Petroleum Systems. Cambridge University Press, p. 541. Oldow, J.S., Ave´ Lallemant, H.G., Bally, A.W., 1991. Transpression, orogenic float, and lithospheric balance. Geology 18, 991–994. Pubellier, M., 2008. Structural Map of Eastern Eurasia. 1:12 500 000, 1 Sheet. CGMW, Paris (Com- mission of the Geological Map of the World)/UNESCO. Ries, A.C., Butler, R.W.H., Graham, R.H., 2007. Deformation of continental crust. Geological Soci- ety London Special Publication 272, 595 pp. Roeder, D., 1973. Subduction and orogeny. J. Geophys. Res. 78, 5005–5024. Roeder, D., 2009. American and tethyan fold-thrust belts. In: Bender, F., Jacobshagen, V., Lu¨ttig, G. (Eds.), Beitra¨ge zur Regionalen Geologie der Erde. Borntra¨ger, p. 168. Roeder, D., 2010. Fold-thrust belts at peak oil. In: Goffey, G.P., Craig, J., Needham, T., Scott, R. (Eds.), Hydrocarbons inContractionalBelts,GeologicalSocietyLondonSpecialPublication348,pp.7–31. Schermer, E.R., Howell, D.G., Jones, D.L., 1984. The origin of allochthonous terranes. Annu. Rev. Earth Planet. Sci. 12, 107–131. Sengo¨r, A.M.C., Natalin, B.A., 1996. Paleotectonics of Asia: fragments of a synthesis. In: Yin, A., Harrison, T.M. (Eds.), The Tectonics of Asia. CambridgeUniversity Press, NewYork, pp. 486–640. Scholl, D.W., von Huene, R., 2007. Crustal re-cycling at modern subduction zones applied to the past—Issues, of growth and preservation of continental basement crust, mantle geochemis- try, and supercontinent reconstruction. In: Hatcher Jr., R.D., Carlson, M.P., McBride, J.H., Martı´nez-Catala´n, J.R. (Eds.), 4-D Framework of Continental Crust, Geological Society of America Memoir 200, pp. 9–32. Seismic Atlas of SE-Asian Basins, 2011. Compilation of seismic images of geological features in SE- Asia basins, related to hydrocarbon potential of this region. Countries covered; Indonesia, Malaysia, Brunei, Philippines, Thailand and Vietnam. http://geoseismic-seasia.blogspot. com/search?updated-min=20. Sengo¨r, A.M.C., 2003. The LargeWave-Wavelength Deformations of the Lithosphere: Material for a History of the Evolution of Thought from the Earliest Times to Plate Tectonics. Geological Society of America Memoir 196, Boulder, CO, 365 pp. Sengo¨r, A.M.C., Natalin, B.A., 1996. Paleotectonics of Asia: fragments of a synthesis. In: Yin, A., Harrison, T.M. (Eds.), The Tectonics of Asia. Cambridge University Press, pp. 486–640. Shaw, J.H., Connors, C., Suppe, J. (Eds.), Seismic Interpretation of contractional fault-related folds. 2004. An AAPG Seismic Atlas, Studies in Geology 531, 66 pp. Sommaruga, A., 1999. De´collement tectonics in the Jura foreland fold and thrust belt. Mar. Petrol. Geol. 16, 111–134. Suppe, J., 1983. Geometry and kinematics of fault-bend bend folding. Am. J. Sci. 283, 648–721. Suppe, J., 1985. Principles of Structural Geology. Prentice Hall, Englewood, p. 537. Von Huene, R. (Ed.), 1987. Seismic Images of Modern Convergent Margin Tectonic Structure, AAPG Studies in Geology 26, 60 pp. Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Von Huene, R., Scholl, D.W., 1991. Observations at convergent margins concerning sediment sub- duction, subductionerosionandthegrowthofcontinental crust.Rev.Geophys.29(3),279–316. Yin, A., Harrison, T.M., 2000. Geologic evolution of the Himalayan-Tibetan orogen. Annu. Rev. Earth Planet. Sci. 28, 1–280. Ziegler, P.A., 1990. Geological Atlas of Western and Central Europe. second ed. Shell Int. Petrol Mij. B. BV., distributed by Geol. Pub. House, Bath, p. 239, 56 enclosures. Ziegler, P.A., Cloetingh, S., vanWees, J.D., 1995. Dynamics of intra-plate compressional deforma- tion: the Alpine Foreland and other examples. Tectonophysics 252, 7–59. doi:10.1016/0040- 1951(95)00102-6. Ziegler, P.A., Bertotti, G., Cloetingh, S., 2002. Dynamic processes controlling foreland development— the role of mechanical (de)coupling of orogenic wedges and forelands. In: Bertotti, G.K., Schulmann, K., Cloetingh, S. (Eds.), Continental Collision and the Tectono-Sedimentary Evolution of Forelands. EuropeanGeophysical Society. StephanMu¨ller Special Publication Series 1, pp. 17–56. 2255..88 Age of Continental basement Introduction to basements, that is, the “residual” peneplaned former fold belts Deformed igneous, metamorphic complexes and sediments are commonly lumped under the term “basement.” The earlier discussion on the topographic expression of evolving fold belts as illustrated by the digital elevationmap (Plate 1) suggested that today’s orogens are destined to become the basements of the future. Conversely, the basement of today’s sedimentary basins are the product of “the rise and fall” of a complex orogenic evolution that ends with their peneplanation. We define the basement as dominantly crystalline igneous/meta- morphic and also, on occasion, as the partly metamorphosed to nonmeta- morphic, deformed remnant of a folded belt that was peneplaned. The top basement unconformity that underlies most sedimentary basins on continental crust is typically a flat to low-relief surface. Initially, the erosion of an orogen is directly related to the interaction of varying climates with tectonic uplifts that are directly related to the evolution of the orogen. Eventually, other processes will terminate the peneplanation, for example, pre- and postrift uplifts that are not directly associated with the earlier orogeny, or else glacial episodes and global eustatic sea-level changes. For example, the African Erosion Surface as presented by Burke and Gunnell (2008) is related to a combination of the opening of the S. Atlantic and Indian Oceans, the movement of Africa over an anomaly at the core-mantle boundary, the emplacement of LIPs over the past 200 Ma, and the creation of two basin and swell episodes associated with intracontinental rifts. 1059 In most sedimentary basins, “wedges of multiple unconformities” branch in a downdip direction, that is, from higher areas toward the structurally deeper 1060 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps domains of most basins. Each unconformity in the wedge reveals the episodic nature of varying regional erosional events. Erasing the record of a complex polyphase erosional history, the top basement unconformity commonly repre- sents the largest hiatus within a basin. Simply stated, the larger the stratigraphic hiatus across an unconformity, the less we are able to isolate succeeding factors that ultimately lead to the formation of the top basement unconformity that underlies all sedimentary basins. Paleogeological/subcrop maps (French: e´corche´) define in map form the result of the tectonic processes responsible for supraregional unconformities. Levorsen was known to refer to this in his “Studies in Paleogeology (1961)” as “The keenest tool in geologist’s kit” for dealing with what Philip King (1959) has described as “the science of gently dipping strata”. These assessments made by two authors of true geological clas- sics are still relevant today! Archean, Early, Mid, and Late Proterozoic from Paleozoic and Mesozoic/Cenozoic basements are here simply differentiated for convenience. In this context, still evolving Mesozoic–Cenozoic (Mz/Cz) folded belts will only be defined as base- ment when overlain by a sedimentary basin (e.g., the sedimentary basins of Indonesia and California) (Plate 25.36). Requiring the overlying unconformity to be the key part of the basement definition leads to the exclusion of structurally reworked older basement fragments that are subsequently involved in younger folded belts. For example, the Grenville age basement of the Blue Ridge of the Appalachians is an integral part of this Paleozoic fold belt and is therefore included as a part of a Paleozoic orogen. Also the Paleozoic of the Eastern Alpine thrust sheets or else the “para-authochtonous” basement highs that underlie but deform the overlying Helvetic nappes of the Swiss Alps are all part of the Ceno- zoic/Mesozoic fold belts of Plates 25.15–25.18. This simplication is useful and provides a well-defined interface between basin studies and the, at least, equally important studies that dissect the anatomy of accretionary folded belts (e.g., Cawood and Kro¨ner, 2010). The formation of most basements involves thermotectonic “re-working” of pre-existing continental basements. Such re-working is also associated with con- siderably younger igneous intrusions and contact metamorphism, as well as regional, that is, orogenic metamorphism that also includes both syn- and late-postorogenic episodes. There remains some ambiguity when, despite being associated with significant younger thermal events, the older basement age of continental rift systems is retained because it forms base of the sedimentary rift fill of grabens. The USGS (2010) prefers to differentiate rift systems and orogenic sys- tems as subsets of their thermotectonic map. Instead, we prefer to separately include rifts as part of basin maps. Such differing preferences are immaterial as long as they are clearly explained. Other kinds of “basement” are often mentioned. The term “magnetic basement” has on occasion led to basement-depth predictions that differ seriously from near- Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps basement tops imaged on good quality reflection seismic profiles and also cali- brated by drilling. It is often not clear whether the source of a magnetic anomaly is from the top of the basement or within the basement. To complicate things, occasionallymagnetic anomalies are due tomagnetite-rich sediment layers or else arise from allochthonous igneous units (e.g., ophiolites) that are involved in youn- ger folded belts, which are underlain by a deeper basement. In this context, cri- teria used to define both nature and the evidence matter, when used to support models/interpretations that in foreland folded belts prefer basement-involved over de´collement/thin-skinned duplex interpretations. Basement-involved inter- pretations tend to discourage further exploration for hydrocarbons while the pos- sibility of mappable duplex interpretations may encourage acquiring additional data in support of an interpretation that may be economically more appealing to the explorer. Petroleum geologists often refer to “economic basement.” The term is not espe- cially useful as there are a number of instances of economic hydrocarbon produc- tion from altered and fractured basement rocks. Likewise, “volcanics” are also often regarded as unattractive reservoirs for hydrocarbon production despite occasional, yet significant, hydrocarbon production from volcanic reservoirs. For practical reasons, hydrocarbon explorationists and researchers interested in the evolution of sedimentary basins more often focus on the latest age of thermal reworking of the underlying basement. Thus, various Precambrian basements are overlain by late Precambrian, Paleozoic as well as Mesozoic sediments, that is, “objective sections,” while a Paleozoic basement will be sometimes overlain by late Paleozoic and more by Mesozoic and Tertiary “objective sections.” On the Tectonic Map (Plate 25.15), older basement units that are mechanically emplaced as uplifts and allochthonous thrust sheets are not differentiated because they were emplaced during a much later orogenic event. Exceptions are Paleozoic and Mesozoic uplifts in the foreland that may extend well beyond the continental subduction boundary of the orogen as shown on Plate 25.15. Some of these uplifts involve only the crystalline basement and overlying sedi- ments while other others correspond to inversions of former extensional/rift systems. Such inversion systems are best documented when the presence of a syn-rift sequence is established by seismic and/or surface data. Merging the global tectonic map with a Precambrian basement map Plate 25.20 merges elements of the Tectonic map (Plate 25.15) with subdivisions of Precambrian outcropping shields and their peneplaned subsurface equiva- lents. Archean (say 2500–4000þ Ma), Early Proterozoic (1600–2500 Ma), 1061 Mid-Proterozoic (1000–1600 Ma), and Late Proterozoic (553–1000 Ma) tectonic 1062 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps provinces are differentiated. However, note that on Plates 25.15 and 25.20 the Paleozoic folded belts and/or basements also include selected Proterozoic/ Vendian folded belts such as the Timanides, and Baikalides of Russia or the Adelaidian folded belt of Australia, because these were viewed as early precursors of the complex assembly of Paleozoic folded belts. Strictly speaking, these folded belts are also part of the Panafrican family of folded belts and lumping them here together with Paleozoic folded belt, though convenient, is not quite correct. Plates 25.15 and 25.20 were principally designed to serve as background for the rift map (Plate 25.29) and the basin maps (Plates 25.31 and 25.32). These plates merge outcropping folded belts (orogens) or else shields with their peneplaned subsurface equivalents, that is, the basement of sedimentary basins. Our tectonic and basin maps emphasize only the last significant thermomechanical/orogenic event prior to peneplanation. This simplification contrasts with many more detailed and precise Precambrian maps that are commonly used to get a best-fitting match of reconstructed past continents and their margins, as well as tectonic models of ancient supercontinents. To display a more complete record of the evolution on conjugate plate margins, more detailed maps often include the mapping of prebreakup dike swarms likely to constrain even better fitting plate tectonic reconstructions. There is an overall consensus that plate tectonics as we know it today may be confidently tracked back in time to about 1000 Ma. However, there are varying perspectives regarding the degree to which plate tectonic principles may apply to the Mid- and Early Proterozoic and Archean record. Hamilton (2003, 2007) and Stern (2008) both caution against excessive uniformitarian zeal in applying modern plate tectonic principles to Archean tectonics and outline models that differ substantially from conventional plate tectonic models. Hamilton (2007) suggests the formation of a “global metabasaltic protocrust” around 4450 Ma from which tonalites-trondjhemites-granodiorites (TTG) were extracted by par- tial melting. Rigid lithospheric plates may have formed locally only a billion years later. Stern (2008) suggests for an unknown part of the Archean, and the Neoarchean that stagnant tectonics alternated with episodes of proto-plate plate tectonics during 2500–2700 Ma and 1.8–2.0 Ma. Condie and Kro¨ner (2008) list petrotectonic and other criteria that characterize plate tectonics and conclude that plate tectonics initiated locally in the early Archean and became more widespread in the late Archean. Extensive reviews addressing the inception of plate tectonics on earth are assembled in Condie and Pease (2008). The uncertainties associated with defining the inception of global plate tectonics, the wide range of crustal growth estimates, for example, the compilation in Rino et al. (2004) and the recognition of evolving plate tectonics, suggests that the following episodes are best viewed as immediate precursors of the Phanerozoic plate tectonics, that is, plate tectonics s.s of Dewey (2007): Later Paleozoic Super Continents include the following: (4) Laurasia (360–300 Ma) separated by the Tethys from Gondwana, that is, the Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps merged East and West Gondwana continents (540–300 Ma) (5) Pangea (300–200 Ma), that is, the merging of Gondwana with Laurussia (Lucas, 2005) Note that the northern continental cratons all derived from Laurasia, which is characterized by dominant Archean to Mid-Proterozoic Grenvillian assemblages while the southern continents were also greatly affected by the Late Proterozoic Panafrican events. Thewell-preserved Tertiary geologic record of 65 Maand a relatively solid record for the combined Mesozoic and Paleozoic for 480 Ma contrast with the progressively more incomplete record of the Precambrian (i.e., the Neoproterozoic (440 Ma), the Mesoproterozoic, the Paleo-Proterozoic (900 Ma), and the Archean (over 2100 Ma). However, it is remarkable how Precambrian studies have progressed rapidly since the outstanding Precambrian Synthesis of Goodwin (1996). Recent global overviews include Condie (1989, 2002) and Eriksson et al. (2004). References: Age of Continental Basement NOTE: A number of the following references to this Age of Continental Basement subchapter are not specifically cited in the text. However they were consulted in the compilation Plate 20. In the following these references are printed in Italics. Alkmim, F.F., Martins-Neto, M.A., 2005. Brazil. In: Selley, R.C., Cocks, L.R.M., Plimer, I.R. (Eds.), Ency- clopedia of Geology, vol. 1. Elsevier, Amsterdam, pp. 306–311. Artemieva, I.M., Mooney, W.D., 2002. On the relations between cratonic lithosphere thickness, plate motions, and basal drag. Tectonophysics 358, 211–231. Blecker, W., 2005. North America. Precambrian Continental Nucleus. In: Selley, R.C., Cocks, L.R. M., Plimer, I.R. (Eds.), Encyclopedia of Geology, vol. 4. Elsevier, Amsterdam, pp. 8–21. Mid- to Late Proterozoic Super-continents setting the stage, include the following: (1) TheGrenvillian assemblyand subsequentdispersal of Rodinia (1100–750 Ma) (for recent reviews, see Condie (2003) Meert and Torsvik (2003) and Li et al. (2008) (2) A postulated late Precambrian assembly of Pannotia (Dalziel, 1997), that is, the combined Laurentia/Siberia and Baltica content and Gondwana (600–540 Ma), and their subsequent dispersal (3) A Panafrican assembly of East andWest Gondwana (540–320 Ma); see recent review by Veevers (2005) 1063 Brown, M., 2009. Metamorphic pattern in orogenic systems and the geological record. In: Cawood, P. A., Kro¨ner, A. (Eds.), Accretionary Systems in Space and Time, Geological Society London Special 1064 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Publication 318, pp 37–74. Bogdanova, S.V., Gorbatschev, R., Garetsky, R.G., 2005. East European Craton. In: Selley, R.C., Cocks, L.R.M., Plimer, I.R. (Eds.), Encyclopedia of Geology, vol. 2. Elsevier, Amsterdam, pp. 34–49. Burke, K.B., Gunnell, Y., 2008. The African Erosion Surface: A Continental-Scale Synthesis of Geo- morphology, Tectonics and Environmental Change over the Past 180 Million Years. Geologi- cal Society of America Memoir 201, 66 pp. Cawood, P.A., Kro¨ner, A. (Eds.), 2009. Accretionary Systems in Space and Time, Geological Society London Special Publication 318, 415 pp. Cawood, P.A., Kro¨ner, A., Collins, W.J., Kusky, T.M., Mooney, W.D., Windley, B.F., 2009. Accretionary orogens through earth history. In: Cawood, P.A., Kro¨ner, A. (Eds.), Accretionary Systems in Space and Time, Geological Society London Special Publication 318, pp 1–37. Condie, K.C., 1989. Plate Tectonics and Crustal Evolution, third ed. Pergamon Press. Condie, K.C., 2000. Episodic growth models; afterthoughts and extensions. Tectonophysics 322, 153–162. Condie, K.A., Des Marais, D.J., Abbott, D., 2001. Precambrian superplumes and superconti- nents: a record in black shales, carbon isotopes and paleoclimates? Precambrian Res., 106, 239–260. Condie, K.A., 2002. Breakup of the paleoproterozoic supercontinent. Gondwana Res., 5, 41–43. Condie, K.C., 2003. Supercontinents, superplumes and continental growth; the Neoproterozoic record. In: Yoshida,M.,Windley, B.F., Dasgupta, S. (Eds.), Proterozoic East Gondwana Assembly and Breakup, Preface Geological Society London Special Publication 206, pp. 1–21. Condie, K.C., Kro¨ner, A., 2008. When did plate tectonics begin? Evidence from the geologic record. In: Condie, K.C., Pease, V. (Eds.), When did Plate Tectonics Begin on Planet Earth? Geological Society of America Special Paper 440, pp. 281–294. Condie, K.C., Pease, V. (Eds.), 2008. When did Plate Tectonics Begin on Planet Earth? Geological Society of America Special Paper 440, 294 pp. Cordani, U.G., Milani, M.J., Thomaz Filho, A., Campos, D.A. (Eds.), 2000a. Tectonic Evolution of South America, 31st Int. Geological Congress, Rio de Janeiro, 856 pp. Cordani, U.G., Sato, K., Texeira, W., Tassinari, C.C.G., Basei, M.A.S., 2000b. Crustal evolution of the South American platform. In: Cordani, U.G., Milani, E.J., Thomas Filho, A., Campos, D.A. (Eds.), Tectonic Evolution of South America, 31st Int. Geological Congress, Rio de Janeiro, pp.19–40. Dalziel, I.W.D., 1997. Neoproterozoic geography and tectonics: review, hypothesis and environ- mental speculation. Geol. Soc. Amer. Bull., 109, 16–42. Dewey, J.F., 2005. Caledonides of Britain and Ireland. In: Selley, R.C., Cocks, L.R.M., Plimer, I.R. (Eds.), Encyclopedia of Geology, vol. 2. Elsevier, Amsterdam, pp. 56–63. Dewey, J.F., 2007. The secular evolution of plate tectonics and the continental crust, an outline. In: Hatcher Jr., R.D., Carlson, M.P., McBride, J.H., Martı´nez Catala´n, J.R. (Eds.), 4-D Framework of Continental Crust, Geological Society of America Memoir 200, pp. 1–8. Eriksson, P.J., Altermann, W., Nelson, D.R., Mueller, W.U., Catuneanu, O. (Eds.), 2004. The Pre- cambrian Earth Tempos and Events. In: Condie, K.C. (Ed.), Developments in Precambrian Geology, vol. 12. Elsevier B.V, Amsterdam, p. 941. Franke, W., Matte, M., Tait, J., 2005. Variscan orogeny. In: Selley, R.C., Cocks, L.R.M., Plimer, I.R. (Eds.), Encyclopedia of Geology, vol 2. Elsevier, Amsterdam, pp. 75–86. Gee, D.G., 2005a. Timanides of Northern Russia. In: Selley, R.C., Cocks, L.R.M., Plimer, I.R. (Eds.), Encyclopedia of Geology, vol 2. Elsevier, Amsterdam, pp. 49–55. Gee, D.G., 2005. Scandinavian Caledonides (with Greenland). In: Selley, R.C., Cocks, L.R.M., Plimer, I. R. (Eds.), Encyclopedia of Geology, vol 2. Elsevier, Amsterdam, pp. 64–74. Goodwin, A.M., 1996. Principles of Precambrian Geology. Academic Press, p. 327. Hamilton, W.B., 2003. An alternative earth. GSA Today 13, 4–12. doi:10.1130/1052-5173(2002) 0132.0.Co:2. Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Hamilton, W.B., 2007. Earth’s first 2 billion years—The era of internally mobile crust. In: Hatch- er Jr., R.D., Carlson, M.P., McBride, J.H., Martı´nez Catala´n, J.R. (Eds.), 4-D Framework of Con- tinental Crust, Geological Society of America Memoir 200, pp. 233–296. Hatcher Jr., R.D., Carlson, M.P., McBride, J.H., Martı´nez Catala´n, J.R. (Eds.), 2007. 4-D Framework of Continental Crust, Geological Society of America Memoir 200, 641 pp. King, P.B., 1959. The Evolution ofNorth America. PrincetonUniversity Press, Princeton,N.J., 189pp. Kro¨ner, A., 2010. The role of geochronology in understanding continental evolution. In: Kusky, T.M., Zhai, M.G., Xiao, W. (Eds.), The Evolving Continents: Understanding Processes of Continental Growth. Geological Society of London Special Publication 338, pp. 179–196. Kro¨ner, A., Stern, R.J., 2005. Pan-African orogeny. In: Selley, R.C., Cocks, L.R.M., Plimer, I.R. (Eds.), Encyclopedia of Geology, vol. 1. Elsevier, Amsterdam, pp. 1–12. Kusky, T.M., Zhai, M.G., Xiao, W., 2010. The Evolving Continents: Understanding Processes of Conti- nental Growth—Introduction. In: Kusky, T.M., Zhai, M.G., Xiao, W. (Eds.), The Evolving Conti- nents: Understanding Processes of Continental Growth, Geological Society London Special Publication 338, pp. 1–6. Levorsen, A.I., 1961. Paleogeologic Maps. Freeman, W.H, p. 178. Li, Z.C., Bogdanova, S.V., Collins, A.S., Davidson, A., DeWeale, B., Ernst, R.E., et al., 2008. Assem- bly configuration andbreak-up history of Rodinia. A synthesis. Precambrian Res. i60, 179–210 and printed: Geodynamic map of Rodinia. See also 3 Appendices. doi:10.1016/ jprecamres.200.04.021. Lucas, S.G., 2005. Pangea. In: Selley, R.C., Cocks, L.R.M., Plimer, I.R. (Eds.), Encyclopedia of Geol- ogy, vol. 4. pp. 225–228. Ma, X.Y., Bai, J., 1998. Precambrian Crustal Evolution of China. Springer, Berlin, 321 pp. Meert, J.G., Torsvik, T.H., 2003. Themaking and unmaking of a continent; Rodinia re-visited. Tec- tonophysics 375, 261–268. Pysarevski, S.A., Murphy, J.B., Cawood, P.A., Collins, A.S., 2008. Late Neoproterozoic and early Cam- brian Paleogeography. Geological Society London Special Publication 241, pp. 9–31. Pysarevski, S.A., Wingate, M.T.D., Powell, C.M.C.A., Johnson, S., Evans, D.A., 2003. Models of Rodinia assembly and fragmentation. In: Yoshida, M., Windley, B.E., Dasgupta, S. (Eds.), Proterozoic East Gondwana: Supercontinent and Assembly and Breakup, Geological Society London Special Publi- cation 206, 1–21. Rino, S., Komiya, T., Windley, B.F., Katayama, I., Motoki, A., Hirata, T., 2004. Major episodic increases in crustal growth determined from zircon ages of river sands; implications or mantle overturn in the Early Precambrian. Phys. Earth Planet. Inter. 146, 369–394. Santosh, M., Mauryama, S., Komiya, T., Yamamoto, S., 2010. Orogens in the evolving Earth: from surface continents to lost continents at the core mantle boundary. In: Kusky, T.M., Zhai, M.G., Xiao, W. (Eds.), The Evolving Continents: Understanding Processes of Continental Growth, Geo- logical Society London Special Publication 338, pp. 77–117. Stein, M., Ben-Avraham, Z., 2007. Mechanisms of continental crust growth. In: Schubert, G. (Ed.), Treatise of Geophysics, Stephenson, D. (Ed.), Evolution of the Earth, vol. 9. 171–195. Stern, R.J., 2008. Modern style plate tectonics began in Neoproterozoic time: An alternative inter- pretation of Earth’s tectonic history. In: Condie, K.C., Pease, V. (Eds.), When Did Plate Tectonics Begin on Planet Earth?, Geological Society of America Special Paper 440, pp. 265–280. Stern, R.J., 2010. The anatomy and ontogeny of modern Intraoceanic systems. In: Kusky, T.M., Zhai, M.G., Xiao, W. (Eds.), The Evolving Continents: Understanding Processes of Continental Growth, Geological Society London Special Publication 338, pp. 7–34. Stevenson, D.J., 2007. Earth formation and evolution. In: Schubert, G. (Ed.), Treatise of Geophysics, Stephenson, D. (Ed.), Evolution of the Earth, vol. 9. 1–12. Santosh, M., Mauryama, S., Komiya, T., Yamamoto, S., 2010. Orogens in the evolving Earth; from surface continents to lost continents at the core mantle boundary. In: Kusky, T.M., Zhai, M.G., 1065 W, Xiao (Eds.), 2010. The Evolving Continents: Understanding Processes of Continental Growth, 1066 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps 2255..99 Hot spots, linear island chains, large igneous provinces (LIPs), and radiating dike swarms; active volcanoes Introduction Phanerozoic volcanism (as discussed by Kerr and Menzies in Chapter 2, Vol. 1A) is an important key to understanding regional plate tectonic settings. Plates 25.21– 25.27 display a selection of globally/regionally important igneous provinces. Wilson (1963) proposed that plates moving over “fixed” deep mantle plumes leave a hot spot trail such the Hawaiian-Emperor chain of volcanic islands. Morgan (1971, 1972a, 1972b) suggested that groups of hot spots may be fixed with ref- erence to each other while others (e.g., Turcotte and Oxburgh, 1973) suggest extensional faulting may be influenced by the location of hotspots. Plate 25.21 was re-drafted on the basis of overviews compiled by Norton (2000) and Windley, B.F., 1995. The Evolving Continents. John Wiley, 526 pp. Windley, B.F., 2002. Continental growth in the Proterozoic: a global perspective. In: Yoshida, et al., (Ed.), Geological Society London Special Publication 206, vol. 437. pp 23–33. Yoshida, M., Windley, B.F., Dasgupta, S. (Eds.), 2002. Proterozoic East Gondwana Assembly and Breakup, Geological Society Special Publication 206, 472 pp. Xiao, W., Han, C., Chao, Y., Sun, M., Zhao, G., Shan, Y., 2010. Transitions among Mariana-, Japan-, Cordillera-, and Alaska-type systems and their final juxtapositions leading to accretionary and col- lisional orogenesis. In: Kusky, T.M., Zhai, M.G., Xiao, W. (Eds.), The Evolving Continents: Under- standing Processes of Continental Growth, Geological Society London 338, pp. 35–53. Geological Society London Special Publication. 338, pp 77–117. Tollo, R.P., 2005. Grenvillian Orogeny. In Selley, R.C., Cocks, L.R.M., Plimer, I.R. (Eds.), Encyclopedia of Geology, Vol. 1. pp 1–25. Torsvik, T.H., 2003. The Rodinia jigsaw puzzle. Science 30, 1379–1381. Trompette, R., (translated by A.V.Carozzi) 1994. Geology ofWesternGondwana (2000–500Ma); Pan- African-Brasilian Aggregation of South America and Africa. A.A. Balkema, Rotterdam, 350 pp. Tyler, I.M., 2005. Australia. In: Selley, R.C., Cocks, L.R.M., Plimer, I.R. (Eds.), Encyclopedia of Geology, vol. 1. Elsevier-Academic Press, Amsterdam, pp. 208–251. USGS, Retrieved 2010. Geologic Province and Thermo-Tectonic Age Map Earthquake Hazards Program. http://earthquake.usgs.gov/research/structure/crust/map. Van Hunen, J., van der Berg, A., Vlaar, N.J., 2002. On the role of oceanic subducting plateaus in the devel- opmentofshallowslabsubduction. In:Proceedingsofan InternationalWorkshop;RoleofSuperplumes inthe Earth system. Tokyo Institute of Technology, Tokyo, Japan: Abstract volume 120–123. Veevers, J.J., 2005. Gondwanaland and Gondwana. In: Selley, R.C., Cocks, L.R.M., Plimer, I.R. (Eds.), Encyclopedia of Geology, vol. 3. Elsevier-Academic Press, Amsterdam, pp. 128–154. Courtillot et al. (2003). Norton (2000) differentiated three oceanic hot-spot Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps families, that is, a Pacific, an Indo-Atlantic, and a North Atlantic Icelandic family and concluded that the large relative motion between these hot-spot families “makes the notion of a fixed hot-spot reference frame untenable.” Aside from Burke andWilson’s (1976) 117 hotspots, in recent years the number of hotspots identified by different authors ranges from a few to Norton’s 27 oceanic hotspots to others proposing over 40 or 50 hotspots including seven on the continents. Plates 25.21 and 25.22 list and show locations from Norton (2000), Steinberger (2000), and Wo¨lbern (2007). Suetsugu et al. (2005) provide a short overview of the relationship betweenmantle plumes and hotspots. Foulger (2003) has maintained a well-organized website that provides easily accessible additional information on mantle plumes. Anderson (2004) and Anderson and Natland (2005) provide an easily readable history of the plume hypothesis. An “inter- mezzo” within their review describes the semantic chaos associated with hot- spot/plume terminology and is followed by a definition of a plume as “a narrow buoyant, active upwelling that is continuous from a deep mantle boundary that is connected to a surface hotspot.” This definition explicitly excludes passive upwel- lings and upwellings associated with plate tectonics or large-scale mantle convec- tion. Accordingly, positive plume identification needs to include credible arguments supporting the upwelling process from the deepmantle to the surface. Anderson and Schramm (2005), Foulger et al. (2005), and Foulger and Jurdy’s (2007) books give deep insights into the views of defenders of the plume hypoth- esis as well as the various perspectives of protagonists that propose alternative hypotheses. Yet another, lengthy, but most readable book by Foulger (2010) though controversial (see review by Arndt 2012), updates the status of the plume, hot-spot, and melting anomalies debate and suggests that melting is associated with anomalies that are rooted in the shallow mantle. Courtillot et al. (2003), with a list of 49 hotspots and a number of ranking criteria (listed in Chapter 2, Vol.1A (Kerr and Menzies)), isolate only seven “primary” hotspots that indicate a deep origin at the core mantle boundary, as shown in bold letters on Plate 25.21. According to these authors, the primary hotspots may well serve as a “quasi-fixed” frame of reference for plate motions during the last 80–100 Ma. Based on shear wave tomography, Courtillot et al. (2003) also differentiated two “super-plumes” where hot, less dense material rises from the mantle core boundary from where a number of shallow subsidiary hotspots may arise (i.e., one under Namibia on- and offshore, and another centered around the Tuamotu Archipelago of Polynesia). On the other hand, Anderson (2005) concludes that the seven listed primary hotspots that scored well with Courtillot et al.’s subjective plume criteria score poorly using criteria “appropri- ate for deep or thermal processes.” Anderson (2005) specifically mentions that the Iceland, Easter, Afar, Tristan, and Yellowstone hotspots have not been con- firmed by tomography. At least for Iceland, Foulger et al. (2001), Foulger (2005), Foulger and Anderson (2005), and Foulger et al. (2005) have discussed 1067 the tomographic evidence favoring a melting anomaly limited and confined to to the original hotspot concept. Foulger (2007), following Anderson (2001), 1068 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps reviews many of the observations and arguments that favor a new “plate model” invokingmelting anomalies that are the consequence of and due to shallow stress- controlled processes. The proximity of melting anomalies (hotspots) close to spreading midocean ridges and to transform faults (Plate 25.23) suggests a struc- tural control and that the emplacement of a hotspot is essentially top-driven (Anderson, 2001; King and Anderson, 1995). The reality as well the nature and origin of hotspots is likely to remain a matter of controversy for years to come. The suitability of hotspots as a fixed frame of refer- ence for plate tectonic reconstructions is also questioned and many authors ques- tion the deep-mantle origin of hotspots and prefer a shallower origin. All these add further doubts about the original hotspot concept. Formore, see Kerr andMenzies (Chapter 2 this volume, and Foulger [2010]. Plates 25.11–25.26 show the distribution of oceanic and continental hotspots, oce- anic linear island chains, LIPs (large igneous provinces), that is, volcanic plateaus in the oceans, on their margins, and on continents, and giant radiating dike swarms dated as far back as the early Proterozoic. Accurate dating and the thermal scope of some of these igneous events are likely to guide subsidencemodels for sedimentary basins that are believed to originate with hotspot-related rifting. Large igneous provinces (LIPs) (Plate 25.24) Coffin and Eldholm (1992) compiled basic information that catalyzed all subsequent and vigorous debates about large igneous provinces (LIPs). Their map included a long list of names of LIPs classified by their tectonic setting. Since then, their map has been updated. The relation of LIPs tomantle plumes remains a subject of ongoing discussion.Mahoney andCoffin (1997) review key examples of LIPs, and Coffin and Eldholm (2001a,b, 2005) give concise overviews on the sub- ject. They specifically exclude products of mid-ocean magmatism (i.e., much of the upper mantle. Campbell (2001, 2006) still strongly supports traditional plume concepts and visualizes large plume heads (say, 1000 km diameter) connected by narrow feeder tails to the core mantle boundary. He also postulates large domal uplifts (say, 1000 m or less) associated with rising plume heads. Konter et al. (2008) also sug- gest that “retiring mantle plumes may be premature.” Volumes edited by Foulger et al. (2005) and Foulger and Jurdy (2007) include evi- dence that early hotspot concepts, though appealing, now need to be differen- tiated and substantially modified. The deep-mantle origin of hotspots is now questioned as a number of authors prefer a shallower origin, adding further doubt the oceanic crust) and arc magmatism. Bryan et al. (2002) identified a significant Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps number of silicic LIPs. Some of these are associated with passive margins (e.g., South America and East Australia and others with backarc basins). These silicic LIPs are not shown on Plate 25.25. Bryan and Ernst (2008) recently defined LIPs as “Large Igneous Provinces with life spans of about 50 Ma, placed in intra-plate tectonic settings and characterized by short igneous pulses (about 1–5 Ma) that produce > 75% of the total emplaced igneous volume.” On the other hand, Coffin et al. (2002) report in some detail intermittent magmatic activity of the Kerguelen hotspot for about 130 Ma that links the Rajamahal traps of India, the Broken ridge, and the Bunbury basalts of Western Australia as well as Antarctic Dike systems to a Kerguelen Plateau hotspot. Scenarios involving either multiple or else dismembered plume sources have been proposed. Bryan and Ernst (2008) further add to their definition, “LIPs are dominantly mafic but may have significant ultramafic and silicic components and in some case are dominated by silicic magmatism. Giant continental dike swarms, sills and mafic- ultramafic-dominated intrusions, silicic LIPs and possibly Archean tholeites and komatitites.” To summarize, LIPs include large continental flood basalts, volcanic passive margins, ocean basin flood basalts, oceanic plateaus, submarine ridges, ocean islands, and seamount chains and flood basalts. LIPs occur on the continental crust, on the oceanic crust, and on the continent ocean boundary; they also form in syn- to late-orogenic settings (e.g., Yellowstone), on nearly peneplaned old orogens (e.g., Central Atlantic/CAMP event), or on the surface of Precambrian cratons and their neighboring deeply buried Paleozoic basement cover. Some extensive basalt covers have been traced under passive margins or else some sedimentary basins (e.g., the basalt plateau of East Siberia and the adjacent West Siberia basin). Specifically excluded from this LIP definition are regular spreading generated oce- anic basements, as well as anomalous sea floor crust (not shown on Plate 24), sea- mounts seamount chains, and submarine ridges (both included on Plate 25.24). The principal reason for including the latter on Plate 25.24 will become evident later from the discussion of plate 24 (subchapter 25.10) that emphasizes the over- all likelihood that most oceanic LIPs will eventually end up being subducted together with the oceanic lithosphere that underlies them. Only relatively minor fragments may be obducted and/or accreted to an orogen as relatively small allochthonous ophiolitic or “quasi-ophiolitic” remnants to eventually end up as a minor part of deeply eroded and peneplaned basement. LIPs also specifically exclude subduction-related arc magmatism. Although shown on separate maps (Plate 25.25), most continental Paleozoic and Proterozoic LIPs were exhumed and eroded and now are “recognized by their exposed plumbing systems of giant dike swarms, sill provinces and layered intru- sions” (Ernst et al., 2005). 1069 While some LIPs are associatedwith hotspots of one kind or another, documentation of any links to a deepmantle plume is often lacking. The LIP Commissionof the Inter- nationalAssociationofVolcanologyandGeochemistry (LIPCommission,2004)main- tains an exemplary web site that among others includes an “LIP Record” that ranks and rates them in terms of confidence in their linkage tomantle plumes. The highest (A) rating is given to areas of >100,000 sq. km., to mostly mafic magmatism of short duration (10 Ma), linkage to present day hotspots, and also to radiating dike swarms. Table 25.5 lists (A)-rated LIPs that are presumably tied to hotspots. The “LIP of the month” list (LIP Commission, 2004) includes an informative series of succinct and up-to-date summaries onmany of the LIPs of the world and related issues. Whatever their origin and their possible relation to hotspots, LIPs are very real. When present, LIPs are an integral part of regional studies, as they record a significant thermal event affecting very large areas. Some LIPs are thought tomark short-lived eruptions that last only a fewmillion years. Supposedly they would not leave any hotspot trails. However, other LIPs are associated with hotspot trails that formed during the last 80–100 Ma. A number of hotspot studies have focused on the origin, and identification of these trails assuming that they were created by a deep-seated mantle plume. The LIPs shown on Plate 25.24 were redrawn from Coffin and Eldholm (l992, 1994, 2001a,b) and Eldholm and Coffin (2000) and modified slightly on the basis Table 25.5 List of top (A) ranked LIPs tied to hotspots based on LIP record 1070 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps (LIP commission, 2004) LIP Hotspot/mantle plume Columbia River Yellowstone Afar Afar The North Atlantic Province (NAVP) Iceland Deccan Reunion Madagascar Marion Alpha Ridge Queen Elisabeth Rajmahal Kerguelen, Ontong Java Louisville/Manihiki Mid Cretaceous Superplume Parana/Etendeka Japan, Sakhalin? Sorachi Karroo-Ferrar Bouvet Central Atlantic Magmatic Province (CAMP) Cape Verde or Fernando de Noronha of CGMW (2000). In addition to the LIPs themselves, Plate 25.24 also shows sea- Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps mount chains in the Pacific as well as volcanic ridges and swells in the Atlantic Ocean. LIPs associated with passive margins were also differentiated from Conti- nental LIPs and, in the case of West Siberia, their subsurface continuation. On the basis of both published and unpublished sources, volcanic passive margin bound- aries were added to Plate 25.25. Recently, Bryan et al. (2010) published an extensive, updated overview of LIPs with emphasis on late Paleozoic to Tertiary continental LIPs. Giant radiating dike swarms (maps b-6 and b-7) Ernst and Buchan (1997, 2001a,b) have compiled and maintained a large data bank and maps on giant (i.e., over 300 km long) dikes. These authors differen- tiate dikes that may be linear, often coast-parallel (associated with rifting), from arcuate dike swarms (due to regional stress changes or later deformation) or radial dike swarms. Of over 150 giant dike swarms, only 25 qualify as radiating swarms that suggest a central magma source possibly associated with a rising mantle plume. Giant radiating dike swarms are thought to be the “the roots of an LIP,” that is, they expose the plumbing system and are an inseparable aspect of the emplacement of LIPs. Combined with selected basin maps (e.g., Plate 25.34), the information provided by Ernst and Buchan (1997, 2001a,b) was simplified, redrafted, and printed over a greatly simplified age of basement map. Presumably, the emplacement of these dike swarms may reset the age of the lithospheric mantle underlying the area intruded by them. LIPs and associated dike swarms are particularly useful in the context of reconstructions of ancient continents (Ernst et al., 2009). Is there a “canonical progression of tectonic themes” preceding and/or following the emergence of a plume? Ideally, it would be useful to demonstrate a “canonical progression” of events beginning with a plume rising from deep mantle causing a hotspot-related large domal uplift, associatedwith rifting and erosion of the rift shoulders, then followed by the outpouring of LIPs associated with underlying radiating dike swarms, and later by the outpouring of additional volcanics associated with seaward diverging reflectors and their coast-parallel dike swarms, followed by the emplacement of volcanics along passive margins, and finally ending up with a spreading mid- ocean ridge. However, aside from the controversies swirling around the deep versus shallowmantle origin of hot-spots and LIPs, their questionable fixed nature, 1071 and their geochemical and thermal signature, the evidence supporting for any 1072 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps systematic, that is, “canonical progression” is often lacking. Pre-eruptive uplifts were thought to be associated with most LIPs (e.g., Sengo¨r, 2001, and many others). Rainbird and Ernst (2001) illustrate some convincing examples for such uplifts (e.g., Middle Jurassic North Sea, Neoproterozoic of the Coppermine District). However, evidence supporting a pre-eruptive uplift also often appears to be speculative and is challenged by many authors on the basis of compelling field observations, for example, the Indian Deccan Traps (Sheth, 2007) or the Permian Emeishan basalts of SE China (Ali et al., 2010; Uktins Peate and Bryan 2008). The case for the basal Triassic Siberian (Tunguska) Traps is particularly vexing. Following Sharma (1997), Czamanske et al. (1998) and Kamo et al. (2003) show compelling evidence for the absence of a pre-eruptive uplift. Instead, surface geological data show a posteruptive uplift for parts of theNorilsk area. Furthermore, Reichow et al. (2004) on the basis of age-dated samples obtained by drilling in the West Siberian basin indicate the wide subsurface distribution of flood basalts under the West Siberian basin followed with the subsidence of this very large basin. Based on a grid of several regional seismic profiles, this was mapped extensively and con- firmedbyVyssotski et al. (2006) (see alsoChapter 21, Vol. 1C (Vyssotski et al)). Again we are confronted with a situation, whereby there is credible evidence supporting the absence of a significant pretrap uplift. Instead, we observe posttrap-uplift of the basalt to form the present Tunguska plateau that is roughly coeval with themas- sive Post-Triassic mostly Jurassic/Cretaceous subsidence associated with the West Siberia basin where there are some of the largest hydrocarbon accumulations of the world! To complicatematters, according to Vyssotski et al. (2006), rifts mapped on the base reflection seismic profiles appear mostly to be limited to a few narrow “pre-trap” rifts that were subsequently aborted. Clearly, no ocean spreading occurred following the emplacement of the Siberia basalts anywhere near the loca- tion of the center of the radial dikes shown by Ernst and Buchan (1997) and trans- ferred to Plate 25.25. In stark contrast, evidence for syn- and/or post-eruptive Neogene uplifts is obvious from a digital elevation map of East Africa that shows the morphology associated with the Afar triple junction hotspot. Earlier, Cloos (1939) made some compelling diagrams of the Nubia–Arabian part of this uplift thus showing the nexus among syn-rift shoulders, rifting, and volcanism. To conclude, more often than not, there is really no compelling evidence for a plume-associated uplift preceding the emplacement of a LIP; instead, there is evidence for both uplift and subsidence during and following the eruptive event. On the other hand, there appears to be a broad consensus that LIPs associated with radiating dike swarms were emplaced within just a few million years. Collier et al. (2008) tabulated for eight passive margins the delay between peak LIP volcanism and the often vaguely timed inception of ocean spreading. The delay times listed range from35 to 2 Ma for the time available to emplace thick volcanic wedges asso- Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps ciated with seaward-dipping reflectors (SDRs). Ernst (2004), and Bleeker and Ernst (2006) state that short-livedmantle generatedmagmatic events, often last less than 1 Ma but they may be part of an overall longer series of events (say, 10–30 million years) consisting of several discrete short lived magma pulses. Thus, it could also be inferred fromthis that followinga short time for theemplacementof a continental LIP a longer “gestation” period is required to initiate the spreading along volcanic margins. The emplacement of SDR/volcanic wedges may fill that transition time. Thedelay timesmay vary greatly amongpassivemargins thatwere previouslyweak- ened by one or several rifting events that preceded the last/most important rifting event that led to the opening of an ocean and other, stronger margins that under- went only a single rifting episode prior to spreading. In some cases, rifting may be immediately preceded by the emplacement of the radiating dike swarm, or else be much older but also be re-activated later. Often, it is challenging to decide whether the location of LIPs is related to regional stresses, or whether the rifting is consequence of plume-related emplacement pro- cesses associated with LIPs. For example, the Central Atlantic (CAMP) volcanics (Hames et al., 2003) are associated with rifts mostly as part of a late syn-rift phase. However, occasionally CAMP volcanics appear to have a slightly later post-rift age, that is, they are less clearly related to the rift system. Schlische (2003) emphasizes that some rifting preceding the emplacement of LIPs is diachronous along a pas- sive margin as is the emplacement of volcanics on passive margins and ultimately the opening of the oceans (see also Chapter 13 Vol. 1B (Withjack et al); Chapter 7, Vol 1B, Ebinger, Hafid et al., 2000, 2008). Coast-parallel rifts serving as the commonprologueof the evolutionof passivemar- gins (Beutel, 2009a, 2009b) suggest that stresses associated with the creation of newplate boundariesmaywell be the cause for the emplacement of some LIPs that are first emplaced on continental crust. To sum up, the Phanerozoic record shows that rifting may occur prior to, and/or during and after the emplacement of LIPs. There is no published systematic review of long supra-regional seismic strike profiles that connect well-documented SDRs that so far are only shown on selected dip pro- files. Thus, we cannot systematically document whether the SDRs also dip at a lesser angle along strike, in harmony with the notion of an ocean-opening process that may be similar to a propagating crack. Strike profiles also would give a better per- spective on faults, on the “growth” from a continental transfer fault to an oceanic transform fault and its possible relationship to changing igneous processes. To conclude, there is inadequate evidence to support, with the desirable precision, an idealized/clean progression that would systematically and with some accuracy relate the time of emplacement of dike swarms to rifting, to the emplacement of igneous plateaus, or else to volcanicmargins and followed the inception of ocean spreading. Needless to say, numerous rifts were also aborted, that is, they never led to ocean spreading and associated rifted or volcanic passive margins (Plate 25.29). 1073 In a regional context, short lived LIP episodes may influence the early thermal evo- lution of a sedimentary basin. Therefore, the thermal modeling of any given basin 1074 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps based on events that are thought to dominate its inception has to be grounded in a careful analysis of all the evidence and in context of ordered tectono- stratigraphic megasequences that include well-dated igneous episodes. Ernstetal. (2005)reviewpast researchandlist frontiers inLIPresearchtotestalternative hypotheses for the origins of LIPs and to further characterize them in detail. The distri- bution in time and reconstructed space needs to be refined in particular. The episodic characterof LIPs iswell illustratedby their “barcode.”Particularly, forPrecambrianepi- sodes it is not clear howmuch thebar code record reflects randompreservation of the LIPs. The role of LIP clusters in the initial breakup of supercontinents is emphasized by these authors.Coevalmagmatismandassociateddike swarmshelp tomatch formerly adjacent continent margins when reconstructing supercontinents (For more, see Bleeker and Ernst, 2006; Buchan and Ernst, 2006; Ernst et al., 2009). The distribution of active volcanoes (Plate 25.26) Active Volcanoes are yet another neotectonic expression of plate tectonics. Plate 25.26 complements Plates 25.11 and 25.30. These plates are centered on the Pacific and Indian oceans, that is, the main theatre of ongoing plate tectonism. The Smithsonian (Siebert and Simkin, 2002) maintains a very informative global volcanism program that catalogs presently active volcanoes, as well as Holocene and Pleistocene eruptions. The relation of active volcanism to earthquakes and to divergent, transform, and convergent plate boundaries is displayed on the back- ground of the global topography and bathymetry. All this information is shown on an interactivemap that allows the user to zoom in from the global scale to regional and local scales. A Global wall map (Simpkin et al., 2006) is supplemented by an illustrated catalogue of Holocene volcanoes of the world. Siebert and Simkin (2002) and Sigurdsson et al. (2000) published a comprehensive Encyclopedia of Volcanoes. Francis and Oppenheimer (2004), Schminke (2004), and Lockwood and Hazlett (2010) authored informative textbooks on Volcanism. References to 25.9 Hot Spots, Linear Island Chains, Large Igneous Provinces (LIP’s) and Radiating Dyke Swarms; Active Volcanoes NOTE: A number of the following references to this Age of Continental Basement subchapter are not specifically cited in the text. However they were consulted in the compilation Plate 20. In the following these references are printed in Italics. Ali, J.R., Fitton, J.G., Herzberg, C., 2010. Emeishan large igneous province (SW China) and the mantle plume updoming hypothesis. J. Geol. Soc. Lond. 167, 953–959. Anderson, D.L., 2001. Top-down tectonics. Science 293, 2016–2018. Anderson, D.L., 2004. What is a plume?; http://www.mantleplumes.org/PlumeDLA.html. Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Anderson, D.l., 2005a. Scoring hotspots: the plume and plate paradigms. In: Foulger, G.R., Natland, J.H., Presnall, D.C., Anderson, D.l. (Eds.), Plates, Plumes and Paradigms, Geological Society of America Special Paper 388, pp. 31–54. Anderson, D.L., Natland, J.H., 2005. A brief history of the plume hypothesis and its competitors: concept and controversy. In: Foulger, G.R., Natland, J.H., Presnall, D.C., Anderson, D.l. (Eds.), Plates, Plumes and Paradigms, Geological Society of America Special Paper 388, pp. 119–145. Anderson, D.L., Schramm, K.A., 2005. Global hotspot maps. In: Foulger, G.R., Natland, J.H., Presnall, D.C., Anderson, D.l. (Eds.), Plates, Plumes and Paradigms, Geological Society of America Special Paper 388, pp. 19–29. Arndt, N.T., 2012. Review of G.R. Foulger. Plates vs. plumes a geological controversy. Lithos 128, 148–149. Beutel, E.K., 2009a. Magmatic rifting of Pangea linked to the onset of South American Plate Motion. Tectonophysics 468, 149–157. Beutel, E.K., 2009b. Non-radial dikes of the central atlantic magmatic province reveal tectonic source and evolution of the break-up of Pangea. http://www.mantleplumes.org/PangeaDikes .html. Bleeker, W., Ernst, R.E., 2006. Short-lived mantle generated magmatic events and their dyke swarms: the key to unlocking Earth’s paleogeographic record back to 2.6 Ga. In: Handski, E., Mertanen, S., Ramo¨, T., Vuollo, J. (Eds.), Dyke Swarms as timemarkers of Crustal Evolution, Tay- lor and Francis/Balkema. pp. 3–26 http://www.largeigneousprovinces.org/06may. Bryan, S.E., Ernst, R.E., 2008. Revised definition of Large Igneous Provinces (LIP’s). Earth Sci. Rev. 86, 175–202. Bryan, S.E., Riley, T.R., Jerram, D.A., Stephens, C.J., Leat, P.T., 2002. Silicic volcanism: an under- valued component of large igneous provinces and volcanic rifted margins. In: Menzies, M.A., Klemperer, S.L., Ebinger, C., Stephens, J., Baker, J. (Eds.), Volcanic Rifted Margins. Special Paper 362. Geological Society of America, pp. 99–120. Bryan, S., Uktins, P.I., Self, S., Peate, D., Jerram, D., Mawby, M., et al., 2010. Large igneous pro- vinces: sites of the largest volcanic eruptions in earth history. LIP Commission: LIP of the Month 1–31. http://www.largeigneousprovinces.org/LOM.html. Buchan, K.L., Ernst, R.E., 2006. Giant dyke swarms and the reconstruction of Canadian Arctic islands, Greenland, Svalbard and Franz-Josef Land. In: Handski, E., Mertanen, S., Ramo¨, T., Vuollo, J. (Eds.), Dyke Swarms time markers of Crustal Evolution. Taylor and Francis/Balkema, pp. 27–49. http://www.largeigneousprovinces.org/06may. Burke, K., Wilson, J.T., 1976. Hot spots on the Earth’s surface. J. Geophys. Res. 93, 7690–7708. Campbell, I.H., 2001. Identification of ancient mantle plumes. In: Ernst, R.E., Buchan, K.L. (Eds.), Mantle Plumes: Their Identification Through Time. Geological Society of America Special Paper 352. pp. 5–21. Campbell, I.H., 2006. Large igneous provinces and the mantle plume hypothesis. Elements 1, 265–269. CGMW, 2000. Geological map of the World 1:25 000000 in 3 sheets, third ed. (P. Bouysse, compiler). Commission of the Geological Map of the World (in Mercator and Polar Stereo- graphic Projections UNESCO, Paris. Reprinted in 2010). Cloos, H., 1939. Hebung; Spaltung; Vulkanismus; Elemente einer geometrischen Analyse irdischer Grossformen. Geologische Rundschau 30, 405–427. Coffin,M.F., Eldholm, O., 1992. Volcanism and continental breakup: a global compilation of large igneous provinces. In: Storey, B.C., Alabaster, T., Pankhurst, R.J. (Eds.), Magmatism and the Causes of Continental Breakup. Geological Society London Special Publication 68. pp. 21–34. Coffin, M.F., Eldholm, O., 1994. Large igneous provinces: crustal structures, dimensions and external consequences. Rev. Geophys. 32, 1–36. 1075 Coffin,M.F., Eldholm,O., 2001a. Large igneous provinces. In: Selley, R.C., Cocks, L.M.R., Plimer, I. R. (Eds.), Encyclopedia of Geology, vol. 3. Elsevier, Oxford, pp. 315–323. 1076 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Coffin, M.F., Eldholm, O., 2001b. Large igneous provinces. In: Steele, J.H., Thorpe, S.A., Turekian,K.K. (Eds.), EncyclopediaofOceanSciences. Academic Press, London,pp. 1290–1298. Coffin, M.F., Pringle, M.S., Duncan, R.A., Gladczenko, T.P., Storey, M., Mu¨ller, R.D., et al., 2002. Kerguelen Hotspot Magma Output since 130 Ma. J. Petrol. 43, 1121–1137. Coffin, M.F., Eldholm, O., 2005. Large igneous provinces. In: Selley, R.C., Cocks, R., Plimer, I.R. (Eds.), Encyclopedia of Geology. Elsevier, Oxford, 315–323. Collier, J.S., Sansom, V., Ishizuca, O., Taylor, R.N., Minshull, T.A., Whitmarsh, R.B., 2008. Age of India Seychelles breakup. Earth Planet. Sci. Lett. 272, 264–277. See also: http://www .manteplumes.org/Seychelles.html. Courtillot, V., Davaille, A., Besse, J., Stock, J., 2003. Three distinct types of Hot Spots in the earth. Earth Planet. Sci. Lett. 205, 295–308. Czamanske, G.K., Gurevich, A.B., Fedorenko, V., Simonov, O., 1998. Demise of the siberian plume: paleogeographic and paleotectonic reconstruction from the pre-volcanic and volcanic records, North Siberia. Int. Geol. Rev. 40, 95–115. Eldholm, O., Coffin, M.F., 2000. Large igneous provinces and plate tectonics. In: Richards, M.A., Gordon, R.G., van der Hilst, R.D. (Eds.), The History and Dynamics of Global Plate Motions, American Geophysical Monographs 121, pp. 309–326. Ernst, R.E., 2004. Characteristics and Origin of Giant Radiating Dyke Swarms. 7 pp. http://www .mantleplumes.org, GiantRadDykeSwarms.html. Ernst, R.E., Bleecker, W., Hamilton, M.A., So¨derlund, U., 2009. Reconstructing ancient continents using the large igneousprovince record: implications formineral,hydrocarbonandearth systems. LIP Commission: Lip of the Month 10 pp http://www.largeigneousprovinces.org/LOM.html. Ernst, R.E., Buchan, K.L., 1997. Giant radiating swarms: their use in identifying pre-mesozoic large provinces and mantle plumes. In: Mahoney, J.J., Coffin, M.F. (Eds.), Large Igneous Pro- vinces, Continental, Oceanic and Planetary Flood volcanism. Geophysical Monographs, 100 American Geophysical Union, Washington D.C, pp. 297–334. Ernst, R.E., Buchan, K.L., 2001a. The use of mafic dyke swarms in identifying and locating mantle plumes. In: Ernst, R.E., Buchan, K.L. (Eds.), Mantle Plumes, Their Identification Through Time. Geological Society of America Special Paper 352, pp. 247–265. Ernst, R.E., Buchan, K.L., 2001b. Large Mafic events through time and links to mantle plume heads. In: Ernst, R.E., Buchan, K.L. (Eds.), Mantle Plumes, Their Identification Through Time. Geological Society of America Special Paper 352, pp. 483–575. Ernst, R.E., Buchan, K.L., Campbell, I.H., 2005. Frontiers in large igneous province research. Lithos 79, 271–297. Foulger, G.R., 2003. www.MantlePlumes.org. Foulger, G.R., 2005. Iceland and the North Atlantic Igneous province. www.MantlePlumes.org. Foulger, G.R., 2007. The “plate” model for the genesis of melting anomalies. In: Foulger, G.R., Jurdy, D.M. (Eds.), Plates, Plumes and Planetary Processes. Geological Society of America Special Paper 430, pp. 1–28. Foulger, G.R., 2010. Plates Versus Plumes: A Geological Controversy. John Wiley, Oxford, 364 pp. Foulger, G.R., Anderson, D.L., 2005. A source for Icelandic magmas in re-melted Oceanic crust. J. Volcanol. Geoth. Res. 141, 23–44. Foulger, G.R., Jurdy, D.M. (Eds.), 2007. Plates, Plumes and Planetary Processes, In: Geological Society of America Special Paper 430, pp. 997. Foulger, G.R., Natland, J.H., Presnall, D.C., Anderson, D.L., 2005. Plates, Plumes and Paradigms. Geol. Soc. Am. Spec. Pap. 388, 861. Foulger, G.R., Pritchard, M.J., Julian, B.R., Evans, J.R., Allen, R.M., Nolet, G., et al., 2001. Seismic tomography shows that upwelling beneath Iceland is confined to the uppermantle. Geophys. J. Int. 146, 504–530. Francis, P., Oppenheimer, C., 2004. Volcanoes, second ed. Oxford University Press, 521 pp. Gernigon, L., Planke, S., Ringenbach, J.C., Le Gall, B., 2005. Tectonic and deep crustal structures along the Norwegian Passive Margin; Implication for the “Mantle Plume or Not ?” debate www.man- Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps tleplumes.org/VM-Norvay.html. Hafid,M., Ait Salem, A., Bally, A.W., 2000. TheWestern Termination of the Atlas system—Offshore Essaouira Basin. Mar. Petrol. Geol. 17, 431–443. Hafid, M., Tari, G., Bouhadioui, D., El Moussaid, I., Echarfui, H., Aı¨t Salem, A., et al., 2008. Atlantic basins. In: Continental Evolution: The Geology of Morocco. Lecture Notes in the Earth Sciences 116, pp. 303–329. Hames, W.E., McHone, J.G., Renne, P.R., Ruppel, C. (Eds.), 2003. The Central Atlantic Magmatic Province. Geophysical Monograph 13, American Geophysical Union, p. 267. Kamo, S.L., Czamanske, G.K., Amelin, Y., Fedorenko, A., Davis, D.W., Trofimov, R., 2003. Rapid eruption of Siberian Flood volcanics and evidence for coincidence with the Permian Triassic boundary. Earth Planet. Sci. Lett. 214, 75–91. Kerr, A.C., 2003. Oceanic Plateaus. Treatise of Geochemistry Vol. 3. pp. 537–565. King, S.D., Anderson, D.L., 1995. An alternative mechanism to flood basalt formation. Earth Planet. Sci. Lett. 136, 269–279. King, S.D., Anderson, D.L., 1998. Edge-driven convection. Earth Planet. Sci. Lett. 160, 289–296. Konter, J.G., Hanan, B.B., Blichert-Toft, J., Koppers, A.A.P., Plank, A.A.P., Staudigel, H., 2008. One hundredmillion years of mantle geochemical history suggests the retiring of mantle plumes is premature, manuscript in preparation for. Earth Planet. Sci. Lett. 275, 285–295. LIP Commission, 2004. LIP of the Month http://www.largeigneousprovinces.org/LOM.html. Lockwood, J.P., Hazlett, R.W., 2010. Volcanoes: Global Perspectives. Wiley-Blackwell, 552 pp. Lundin, E., Dore´, T., 2005. The Iceland “Anomaly”—An Outcome of Plate Tectonics. 14 pp. www .MamtlePlumes.org. Mahoney, J.J.,Coffin,M.F. (Eds.),1997. Large IgneousProvinces:Continental,OceanicandPlanetary Flood Volcanism, American Geophysical Union Geophysical Monograph, 100, 438 pp. McHone, J.G., 2002. Volatile emissions of central Atlantic magmatic province basalts: mass assump- tions and environmental consequences. In: Hames, W.E., McHone, J.G., Renne, P.R., Ruppel, C.R. (Eds.), The Central Atlantic Magmatic Province. American Geophysical Union Geophysical Mono- graph, 136, 241–254. McHone, J.G., Anderson, D.L., Beutel, E.R., Fialko, Y.A., 2005. Giant dikes, rifts, flood basalts, and plate tectonics: a contention of mantle models. In: Foulger, G.R., Natland, J.H., Presnall, D.C., Anderson, D.L. (Eds.), Plates, Plumes and Paradigms, Geological Society of America Special Paper 388, pp. 401–420. McHone, J.G., Puffer, J., 2003. Flood Basalt provinces of the African Pangean. Rift: Regional extent and significance. In: LeTourneau, P.M., Olsen, P.E. (Eds.), The Great Rift Valleys of Pangea in Eastern North America, vol. 1. Columbia University Press, pp. 141–154. Morgan, W.J., 1971. Convective plumes in the lower mantle. Nature 230, 42–43. Morgan, W.J., 1972a. Plate motions and deep mantle convection. In: Shagam, R. et al., (Ed.), Studies in Earth and Space Sciences: A Volume in Honor of Harry Hammond Hess, Geological Society of America Memoir132, Boulder, CO, pp. 7–22. Morgan,W.J., 1972. Deepmantle, convection plumes and platemotions. Am. Assoc. Petrol. Geol. Bull. 56, 203–213. Morgan, W.J., 1981. Hotspot tracks and the opening of the Atlantic. In: Emiliani, C. (Ed.), The Sea, vol. 7. Wiley Interscience, pp. 443–487. Mu¨ller, R.D., Roest, W.R., Royer, J.Y., Gahagan, L.M., Sclater, J.G., 1997. Digital Isochrons of the world’s ocean floor. J. Geophys. Res. 102, 3211–3214. Mu¨ller, R.D., Royer, J.Y., Lawver, L.A., 1993. Revised plate motions relative to the hotspots from com- bined Atlantic and Indian ocean hotspot tracks. Geology 21, 275–278. Mu¨ller, R.D., Sdrolias, M.S., Gaina, C., Roest, W.R., 2008. Age, spreading rates and spreading sym- metry of the worlds ocean crust. Geochem. Geophys. Geosyst 9, 19. QO4006. doi: 1029/ 2007GCOO1743. 1077 Norton, I.O., 2000. Global hotspot reference frame and plate motions. In: Richards, M.A., Gordon, R.G., Hilst, R.D. (Eds.), The History and Dynamics of Global Plate Motions, American 1078 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Geophysical Union Geophysical Monograph, 121, 339–357. Parfit, L., Wilson, L., 2008. Fundamentals of Physical Volcanology. Wiley-Blackwell, 265pp. Rainbird, H., Ernst, R.E., 2001. The sedimentary record of mantle plume uplift. In: Ernst, R.E., Buchan, K.L. (Eds.), Mantle Plumes: Their Identification Through Time, Geological Society of America Special Paper352, Boulder CO, pp. 227–245. Reichow, M.K., Saunders, A.D., Ivanov, A.V., Puchkov, V.N., 2004. The Siberian large igneous province. Event 36 of LIP database. LIP of the Month 7. Saunders, A.D., Fitton, J.G., Kerr, A.C., Norry, M.J., Kent, R.W., 1997. The North Atlantic Igneous province. In: Mahoney, J.J., Coffin, M.F. (Eds.), Large Igneous Provinces: Continental, Oceanic and Planetary Flood Volcanism, American Geophysical Union Geophysical Monograph, 100, pp. 45–94. Schlische, R.W., 2003. Progress in understanding the structural geology, basin evolution, and tectonic history of the eastern North American rift system, In: Le Tourneau, P.M., Olsen, P.E. (Eds.), The great Rift Valleys of Pangea, vol. 1. Columbia University Press, New York, pp. 21–64. Sengo¨r, A.M.C., 2001. Elevation as indicator of mantle-plume activity. In: Ernst, R.E., Buchan, K.L. (Eds.),Mantle Plumes: Their Identification Trough Time, Geological Society of America Special Paper 352, pp. 183–225. Schminke, H.U., 2004. Volcanism. Springer, Berlin, Heidelberg, 320. Sharma, M., 1997. Siberian traps. In: Mahoney, J.J., Coffin, M.F. (Eds.), Large Igneous Provinces: Continental, Oceanic and Planetary Flood Volcanism. American Geophysical Union Geophys- ical Monograph, 100, pp. 273–295. Sheth, H.C., 2007. Large Igneous provinces (LIP’s): definition, recommended terminology and a hierarchical classification. Earth Sci. Rev. 85, 117–124. Siebert, L., Simkin, T., 2002. Volcanoes of theWorld: an Illustrated Catalog of Holocene Volcanoes and their Eruptions. Smithsonian Institution, Global Volcanism Program Digital Information Series, GVP-3, http://www.volcano.si.edu/world/. Sigurdsson, H., Houghton, B., McNutt, S., Rymer, H., Stix, J., (with foreword Ballard R. D. 2000. Encyclopedia of Volcanoes. Academic Press, San Diego, 1417pp. Simpkin, T., Tilling, R.I., Vogt, P.R., Kirby, S.H., Kimberly, P., Stewart, D.B., 2006. This Dynamic Planet: World Map of Volcanoes, Earthquakes, Impact Craters, and plate Tectonics. U.S. Geo- logical Survey Geologic Investigations Series Map I-2800, 1 two-side sheet, scale 1:30 000 00. http://mineralsciences.si.edu/tdpmap/. Steinberger, B., 2000. Plumes in a convectingmantle: models and observations for individual hot- spots. J. Geophys. Res. 105(B5), 11127–11152. Suetsugu, D., Steinberger, B., Kogso, T., 2005. Mantle plumes and Hot Spots. Selley, R.C., Cocks, L.R.M., Plimer, I.R. (Eds.), Encyclopedia of Geology. Elsevier, Amsterdam Vol. 3. pp. 335–343. Turcotte, D.L., Oxburgh, E.R., 1973. Midplate tectonics. Nature 244, 337–339. Ukstins Peate, I., Bryan, S.E., 2008. Re-evaluating plume-induced uplift in the Emeishan large igneous province. Nat. Geosci. 1, 625–629. Vyssotski, A.V., Vyssotski, V.N., Nezhdanov, A., 2006. Evolution of the West Siberian Basin. Mar. Petrol. Geol. 23, 93–126. Wilson, J.T., 1963. A possible origin of the Hawaiian islands. Can. J. Phys. 41, 863–870. Wo¨lbern, I., 2007. File: Hotspots.jpg. http://commoms.wikimedia.org/File:Hotspots.jpg. Zhao, D., 2004. Global tomographic images of mantle plumes and subducting slabs; Insights into deep earth dynamics. Phys. Earth Planet. Int. 146, 3–34. tectonic processes involved repeated recycling of oceanic crust as well significant Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps portions of the transitional active/passive margin crust. Plate 25.27 presents a “menu” of say up to 195 Ma old “oceanic subductables” “contrasted with the dispersed left-overs”—accretionary continental orogens that survived 4.5 Ga of recycling (see Section 25.7/Plates 25.17–19 and Section 25.8/Plate 25.20). In the following, we will proceed with examples of subducted oceanic plateaus, followed by examples of accreted oceanic plateaus; the “saga of the ophiolites” will be reviewed and finally followed by examples of exhumed continental crust/mantle transitions. Table 25.6 contrasts “Pristine” and the “Allochthonous” tectonic settings of mafic and ultramafics. Dilek and Ernst 2008a and b explore the links between Ophiolites and LIPs. 2255..1100 Tectonicsettingsofmafic/ultramafic oceanicand intra-oceanic arcsystem crust, LIPs, rifted and volcanic passive margins, tectonic setting and discussion of equivalent allochthonous “ophiolitic” fragments in orogens Introduction The oldest preserved oceanic crust may be Early Jurassic in age (say, 195 Ma). Clearly, much of the oceanic crust/lithosphere as well as oceanic LIPs, sea mounts, and ridges that formed during the Triassic, Paleozoic, andmuch of the Proterozoic has been almost entirely subducted. Based on lithospheric buoyancy considerations, Cloos (1993) differentiated the following: (1) “Inherently subductable oceanic lithosphere” including oceanic lithosphere older than 10 Ma and oceanic plateaus capped by �17-km basaltic crust that, because of subduction metamorphism may become even less buoyant than oceanic lithosphere. (2) Lithosphere that “resists subduction” ranging from, say, 100-km-thick con- tinental lithosphere with a� 30-km continental/granitic crust, to continental crusts and over 20 Ma oceanic island arc crusts that are> 15-km-thick basal- tic plateaus that have crustal thicknesses > say, 30 km also will collide with continents rather than subduct. From this, it follows that, at least for the Phanerozoic and the Proterozoic, plate 1079 Table 25.6 Tectonic settings of mafic/ultramafic oceanic and introceanic arc systems, LIPs, and passive margins and example 1. Pristine tectonic settin 1. 1. Oceanic crust and exposed on the s (Mid-ocean ridge 1.1.1. Fast-spread 1.1.2. Slow-sprea 1.1.3. Ultraslow-s 1.2. Active margin sup 1.2.1. Arc Forearc 1.2.2. Oceanic Ba 1.2.3. Volcanic Ar Note that in the Ba are suggested for 1.3.Oceanic Plateaus, Ridges, and LIPs (e.g., see Plate 25.24) (Plume (P) type of Dilek and Furness, 2011) 1.3.1. Atlantic-typ 1.3.2. Atlantic-typ 1.4. Oceanic Seamoun 1.4.1. Atlantic-typ Pacific Type seam 1.5. Nonvolcanic, hyp 1.6. Volcanic passive m 2. Allochthonous oceani lithosphere/crust inco 2.1. Formed at Mid-O 2.2. Formed as regular Caribbean plateau 2.3. Formed as Ocean 2.4. Formed in the co 2.5. Formed on transit 3. Allochthonous contine Alps, Lac the Lherz of 1080 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps e passive margin plateaus associated with volcanic margins (e.g., Voring Plateau) e rise and plateaus (e.g., Bermuda Rise) t chains e seamount chains (e.g., New England seamounts) ount chains (e.g., Hawaii-Emperor, Line Islands, or Louisville seamounts) s of allochthonous fragments in orogens gs of mafic/ultramafic oceanic and introceanic arc systems, LIPs, and passive margins (Plate 25.27) upper mantle formed at Mid-Oceanic Ridges (MOR). Upper mantle (abyssal) peridotites may be exhumed and ea floor in the context of extensional or transtensional core complexes often associated with fracture zones (MOR-type) of Dilek and Furness, 2011) ing MOR (e.g., East Pacific Rise) ding MOR (e.g., Mid-Atlantic Ridge, Iceland, S. Atlantic Ridge) preading ridges (e.g., SW Indian Ocean Ridge, Arctic Ocean Gakkel Ridge) ra-subduction setting (Supra-subduction-zone (SSZ-type) of Dilek and Furness, 2011) systems (e.g., Intra-oceanic Izu Bonin-Mariana Forearc-arc backarc system) ckarc Basin systems (e.g., Sandwich Arc-Basin) c Systems (e.g., Indonesia) (Volcanic arc (VA) type of Dilek and Furness (2011)); sin classification Tables 25.2 and 25.3, in Bally and Roberts, of Chapter 4 of this volume four different settings backarc basins, ranging from intra-oceanic, to ocean continent to intracontinental backarc settings. In a recent publication, Brown and Ryan (2011) reviewed the role of arc- continent collisions in the context of the assembly of an orogen and Afonso and Zlotnick (2011) discuss in substantial detail key theoretical aspects of continental subduction. Subducted oceanic plateaus Plate 25.27 shows that a number of seamount chains and oceanic plateaus are currently being subducted. The following is a selection of a few examples of the subduction of oceanic plateaus that have been studied in some detail: (1) Mann and Taira (2004a, 2004b), and Mann and Gahagan (2004) review the subduction of the largest and thickest plateau during the Neogene conver- gence of theOntong Java Plateau and the Solomon Islands and conclude that 80% of the crustal thickness of the plateau has been subducted while only the uppermost basalts and sediments (say, 7 km) are accreted and preserved er extended, distal passive margin and exhumed transitional mantle (e.g., Iberia and Newfoundland) argins (see Plate 25.24) c (ophiolithic) fragments/terranes found in continental orogenic settings (i.e., remnants of former oceanic rporated in a continental mostly accretionary orogen) cean Ridge (MOR) (e.g., Franciscan Ophiolites of California) oceanic plateaus (LIPs) (e.g., Basaltic volcanics of theW. Cordillera of Colombia, representing fragment of the ) ic plateaus overlying island arc systems (Wrangellia terranes of NW America) ntext of Intra-oceanic Active Margin Supra-subduction (Troodos (Crete), and Semail (Oman) Ophiolites) ional, hyper-extended, rifted Passive Margins (e.g., Alpine ophiolites of the original “Steinmann Trinity”) ntal crust–mantle transition fragments of outcropping in orogenic folded belts (e.g., Ivrea Zone of southern the Pyrenees, Ronda of The Betic Cordillera-Spain, Beni Bouchera of the Inner Rif Mountains-Morocco). as part of the Solomon Islands arc. The authors note that this is consistent Shatsky (see Plate 25.24) and Hess Plateaus of the western Pacific were sub- et al. (2008a, 2008b), this Triassic terrane is generally characterized by Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps subaerial basaltic flows, but in Vancouver Island >3-km-thick pillowed and submarine basaltic flows are overlain by 400–1500 m of hyaloclastites and pillow breccias. This Triassic terrane accreted to North America in the Late Jurassic–Early Cretaceous. Remarkably, the Triassic plateau basalts uncon- formably overlie arc-derived upper Paleozoic volcanics. Allochthonous fragments, oceanic and intra-oceanic arc systems, and lower crust and uppermost mantle of hyper-extended passive margins Aside from accreted oceanic plateaus, relatively smaller fragments of oceanic and transitional crust were added, exhumed and partially preserved as allochthonous ducted under the Eastern US and Mexico Cordillera in a flat-slab subduction mode that was related to the Laramide orogeny. On the basis of a tomo- graphic image, today, following relative continuing eastward drift of N. America, the Shatsky/Hess conjugate remnant appears to rest at depth below the Great Lakes region. Allochthonous accreted oceanic plateaus and intra-oceanic island arc terranes The preceding segment cited examples of LIPs that where subducted. However, there are also examples of relatively large fragments of oceanic LIPs that accreted to orogens, to become part of an orogenic continental crust. These include the following: (1) Fragments of the Cretaceous Caribbean plateau that accreted to the western Andes orogen of Ecuador (e.g., Jaillard et al., 2009) andColombia (Kerr et al., 1998; Kerr et al., 2009). (2) The Wrangellia Terrane of the N. American Cordillera. According to Greene with fragments of plateaus and seamount chains that are preserved in Precambrian and Phanerozoic orogens. (2) FollowingGutscher et al. (1999), Ramos and Folguera (2009) list criteria for the flat-slabsubductionof the Incaplateau followedbyslabsteepeningundera thick continental crust that leads to delamination and basaltic underplating, melting of the lower crust extension, and rhyolitic volcanism. Instead, slab steepening under thin crust would lead to extension and widespread basaltic lava flows. (3) Liu et al. (2010) indicate that plateaus that were the conjugates of the 1081 units in orogens. These fragments are frequently referred to as “ophiolites,” that is, 1082 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps key elements of “ophiolitic sutures” that often separate terranes within accretion- ary orogens. They include the following: (1) Fragments of oceanic crust, that is, “ophiolites” and overlying sediments. (2) Accretionary wedges (prisms) involving oceanic sediments as well as ophiolites. (3) Supra-subduction intra-oceanic forearc, arc, and backarc systems. (4) Lower crust and uppermost mantle of nonvolcanic, hyper-extended passive margins Notably absent from this list of allochthonous units embedded in orogens are fragments of volcanic passivemargins (i.e., the SDRs). It remains an open question whether this is due to a lack of compelling evidence or else because transitional volcanic passive margin crust always is destined to be subducted in its entirety. While Callot et al. (2002) address the rheological development associatedwith the formation of volcanic margins, Afonso and Zlotnick (2011) explain some of the processes that may prevent the accretion incorporation of volcanic passive margin crust in accretionary orogens Each of the above-listed fragments has characteristic, petrologic/geochemical, as well as stratigraphic signatures. Island-arc and active margin-related volcanic arc terranes are also modified by igneous intrusions and extrusions that involve mag- mas that are generated either from the subducting ocean slab or else in varying degrees by the backarc mantle and its overlying continental crust. Magmatic underplating is also often inferred in this context. Table 25.6 under (1) lists the diverse “pristine” oceanic and continental margin set- tings that are currently available for future subduction. Regular oceanic crust is formed on spreadingmid-ocean ridges (MOR) or in, perhapsmore ephemeral, oce- anic backarc basins. On the other hand, LIPs that include oceanic plateaus, volcanic island chains, etc., are emplaced on oceanic crust or else on highly attenuated rem- nants of continental crust. Also shownonPlate 25.27 are intra-oceanic island arc sys- tems and volcanic passive margins. In this perspective, an acceptable definition of the term “ophiolite” remains elusive. Informally used, the scope of the term is likely to continue evolvingwith time until superseded by better defined terms. In general, ophiolites are often relatively small allochthonous fragments of more or less com- plete sequences of oceanic crust and upper mantle and/or oceanic LIPs or else the transitional oceanic/continental crust of passive margins. When incorporated in an orogenic belt, these uprooted and often widely traveled fragments help to delineate suture zones that separate various accreted terranes within an orogen. The term “ophiolitic suture” implied that ophiolites were remnants of an ocean that separated originally widely separated terranes. The study of ophiolites includes their origin, as well as subsequent petrologic/geo- chemical modifications that these rocks acquired in their travel across the oceans, and following their final incorporation in continental and intra-oceanic orogens. Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Subsequent structural modifications of ophiolites are also associated with the progress and completion of the orogeny. Many “ophiolitic sutures” have been affected by post emplacement strike-slip, or else by normal or thrust faulting that occurred much later than their initial incorporation in an orogen. Robinson et al. (2003), Dilek and Robinson (2003), Bernoulli et al. (2003), Dilek (2003), and Dewey (2003) all present historical perspectives on the evolution of the ophiolite concept. Much of the following was gleaned from these papers. Dilek and Furness (2011) have recently published a comprehensive review on ophiolite genesis. The origin of the current debate on ophiolites goes back to Steinman who, in the early 1900s in the Arosa area of the eastern Swiss Alps, noted an association of ser- pentinites, pillow basalts, and radiolarites. This association became widely known as the “Steinmann Trinity” andwas originally interpreted as igneous intrusions and volcanics in the axial part of a “geosyncline” (Steinmann, 1927 and translation in Steinmann et al., 2003). Going back to the impact of the 1872–1876 Challenger expedition, Steinmann emphasized the oceanic/pelagic nature of the overlying radiolarites. Bernoulli and Jenkyns (2009a) note the historic differences in perspec- tives between the European versus the American view of the eugeosyncline (see also Table 25.1). Alpine authors saw ophiolites as intrusions in embryonic nappes overlain by oceanic/pelagic geosynclinal sediments while American authors thought their “eugeosynclines” to be more closely associated with volcanic island arcs. Bernoulli and Jenkyns (2009b) document our progressive understanding of the association of pelagic sediments associated deep-water continental and/or oceanic crust. An oceanic lithospheric origin was proposed byMoores and Vine (1971) for the rel- atively complete ophiolitic sequences of the Troodos (Cyprus) (for more see Robin- son et al., 2003) and the Semail (Oman) complex (Glennie et al. 1973, 1974, 1990). Both were thought to be analogous to regular oceanic lithosphere that was formed on mid-ocean ridges. For more, see Chapter 27, Vol. 1A (Nicholas) DuringaPenrose conference (Anonymous, 1972)ophioliteswere re-definedasadis- tinct sequence of ultramafic rocks that included the following from bottom to top: (1) ultramafics with varying parts of harzburgite, lherzholite, and dunite that were often overprinted by a more or less serpentinized metamorphic fabric, overlain by (2) gabbros that were often less deformed, followed by (3) sheeted dikes and eventually by (4) pillow lavas The definition also included the associated overlying sediments (e.g., “ribbon cherts”/radiolarites, thin shale interbeds, and minor limestones) and chromite pods associated with dunite and sodic felsic intrusive and extrusive rocks. It was 1083 also recognized that fault contacts between these units were common. Dilek and 1084 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Picardo (2008) remind us that the inverse (i.e., from top to bottom) sequence was correlated with seismically defined crustal layers, that is, Layer 1: Oceanic sedi- ments; Layer 2 extrusives/pillow lavas and sheeted dikes; Layer 3: Gabbros; and Layer 4: upper mantle ultramafics. Initially, many workers thought that ophiolites were representative of oceanic crust that formed on mid-ocean ridges. The Pen- rose conference definition was based on more complete ophiolite complexes that differed substantially from the Alpine Steinmann Trinity, which appeared to be incomplete mostly because sheeted dike complexes were absent or rare. Originally deeply steeped in geosynclinal lore, the epic of the original Steinmann Trinity and its ophiolites eventually led to a new vista of its type locality as an exhumed hyperextended rifted passive margin. Manatschal (1995) presented a carefully documented reconstruction of an allochthonous ocean continentmargin transition of the highest Penninic nappe and the overlying Err nappe of eastern Switzerland. Today ophiolites of the Steinmann trinity are compared in great detail with the well-documented hyper-extended rifted passive margin prototype offshore Galicia (Boillot and Froitzheim, 2001; Desmurs et al., 2001; Manatschal and Bernoulli, 1999; Manatschal, 2004, Pe´ron–Pinvidich and Manatschal (2010)). A careful petrologic geochemical evaluation of ultramafic andmafic rocks exposed at the seafloor of the Galicia margin by Chazot et al. (2005) traces the evolution of a possibly pre-/syn-Hercynian subcontinental lithospheric mantle. Wilson et al. (2001), Manatschal et al. (2007) and Tucholke et al. (2007) place the breakup of the Galicia/Iberian and Newfoundland passive margins into a broader context. (For more, see Chapter 9, Vol. 1C (Whitmarsh and Manatschal); Tucholke and Whitmarsh Chapter 10, Vol. 1C)). As noted by Dilek (2003), soon after the 1972 Penrose conference when it was commonly understood that ophiolites originated at mid-ocean ridges, there was already a suspicion, based on petrologic/geochemical work, that some ophio- lites might not be fragments of crust and upper mantle generated at mid-ocean ridges. New concepts (Dilek and Flower, 2003; Dilek and Robinson, 2003; Flower and Dilek, 2003) included the generation of “ophiolitic magmas” in supra- subduction zones of embryonic arcs and their fore and backarc settings of intra-oceanic arcs (e.g., Marianas, Izu-Bonin arcs of the W. Pacific). Stern (2010) details the anatomy and ontogeny of modern intra-oceanic arc systems. Recently Dilek and Furness (2011) updated earlier reviews on ophiolites. In another example, Petterson (2010) reviews and interprets the Kohistan arc of north Pakistan as an accreted early to late Cretaceous Intra-oceanic Island arc fragment that evolved into an Andean-type arc prior to the Paleogene collision with India. The debate on mid-oceanic ridge versus supra-subduction ophiolites (e.g., back- arc, forearc, and intra-arc rift ophiolites) resulted in an extended California Coast Range controversy that was recently reviewed by Hopson (2007) and Hopson et al. (2008). These authors carefully retrace the long voyage from a Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Mid-Jurassic–Early Cretaceous paleo-equatorial spreading ridge, followed by additional spreading that transported oceanic crust below the CCD, to regions where the coincidence of upwelling and fertilization by volcanic ashes led to blooming of radiolarians and deposition of radiolarian oozes. A change from a subduction to a dextral transform boundary regime in the latest Jurassic coincides with Late Jurassic oceanic magmatism and the deposition of volcaniclastic sedi- ments that in turn was followed by turbidite deposition of the Great Valley group of California. The compressional plate boundary regime was resumed during the early to mid- Cretaceous with thrusts forming along the early transform faults and followed by westward growth of the Franciscan accretionary wedge and continuing sedimen- tation in the Great Valley forearc basin. To encapsulate the complex evolution of the California ophiolites in a short paragraph is a challenge. However, Hopson et al. (2008) do show why understanding ophiolites requires the close interaction of stratigraphers/paleontologists with petrologists/geochemists aswell as geophy- sicists and, of course, careful evaluation of outcrops. Allochthonous, exhumed continental crust-mantle transitions and the Ivrea-Verbano zone The outcrops of the Lac de Lherz area of the Pyrenees, the Ronda area of the Betic Cordillera, and the Beni Bouchera area of the Inner Rif Mountains ofMorocco have long attracted a great deal of attention as they offer a glimpse of the nature of the continental crust mantle transition. However, the Ivrea-Verbano outcrops of southern Alps have drawn most attention because they are easily accessible and because, during the Tertiary a, now exhumed, complete crustal profile was rotated into an upright/subvertical position. In a first approximation, the Ivrea-Verbano outcrops expose the now allochtho- nous mantle of the Eastern Alpine basement-involved thrust sheet, which was the northern part of the Adriatic plate/platform. The Ivrea-Verbano crustal profile includes the crust mantle transition of a pre-/Late Carboniferous basement. The exposed mantle rocks are commonly described as lherzolites. They are overlain by granulite-grade metapelites of the lower crust and amphibolite-grade metape- lites of an attenuated transitional middle crust, and greenschist facies of the upper crust that was intruded by Permian granites. With a focus on the Ivrea-Verbano outcrops, additional insights are gained from outcrops of the upper crust farther east, that is, the Lake Lugano and Lake Como areas (Strona-Ceneri zone/Scisti dei Laghi Zone) and the Southern Alps to the east. In past few decades, numerous papers were published on all these areas. The fol- lowing much abbreviated, “calendar of tectonic events” is mostly based on 1085 overviews by Schmid et al. (1987), Handy and Zingg (1991), Handy et al. (1999), and Siegesmund et al. (2008). 1086 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Following Boriani et al. (1990), the upper crust exposed in the Lake zone to the east of the Ivrea-Verbano zone consists of deformed early Paleozoic metasedi- ments and paragneisses intruded by Ordovician granitoids and pegmatites. Iso- clinal folding was followed by uplift. Boriani et al. (1990) contrast this Paleozoic convergent margin setting with the subsequent extensional setting of the Late Permian and Late Triassic/Early Jurassic. Permian transtensional shear zones and/or low angle normal faulting affect the lower crust. Also, during the Permian, granites intrude the upper crust. Still further east, low angle brittle detachment faults separate the basement from overlying Permian sediments (Froitzheim et al. 2008). During the Late Triassic and Early Jurassic extension, rifting and low angle normal faults (Bertotti et al., 1993) in part overprinted earlier Permian low-angle faults in the lower crust. Rutter et al. (1999, 2003) emphasize the role of lower crustal magmatic underplating in association with Mid-Permian crustal extension and laccolithic intrusions. During much of the Mesozoic, the Southern Alps were part of a passive margin with sediments deposited on the, now allochthonous, basement, Eastern Alpine basement-involved thrust sheets. All these units were originally part of much larger Adriatic platform. Finally, during the Ter- tiary/Neoalpine phase, the Ivrea zone was tilted into a subvertical position and exhumed exposing an “exhumedMoho” and a complete cross-section of continen- tal crust first attenuated during the Permian and, later, during the Late Triassic/Early Jurassic. Fountain (1976) measured, under appropriate confining pressures, crustal veloci- ties of rocks from the Ivrea-Verbano zone at 6.45 km/s for amphibolite facies schist and gneiss of the upper level, to 7.1 km/s for mafic gneiss and paragneiss of amphibolite-granulite facies, and to 7.5 km/s for intermediate andmafic granulitic facies rocks at the base of the crustal profile followed by an abrupt change to 8.5 km/s for the ultramafic complexes of the Ivrea-Verbano zone. Rutter et al. (1999, 2003), based on P-wave velocities, produced a 76-m-long synthetic seis- mic reflection profile and concluded that the image corresponds closely to present-day crustal seismic profiles. The synthetic profiles would not reproduce the important recumbent folds that were observed on outcrops nor the relation- ship of intrusives to the surrounding rocks (see also Khazanehdari et al. 2000). The seismically layered/“laminated” lower crust of the Ivrea-Verbano zone is likely to be associated with some of the subhorizontal shear zones associated with Permian and upper Triassic to lower Jurassic extension. Conclusion This very limited review of allochthonous mafics and ultramafic rocks reveals a continuum of primary, that is, pristine oceanic, island arc active and passive margin settings (see Plate 25.27). In the past, allochthonous occurrences of these rocks were generically lumped under the term “ophiolites.” However, following Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps insights gained in recent decades, today we have to differentiate not only the orig- inal settings but also the various tectonic processes that modified the textures of these rocks acquired during their voyage prior to being incorporated into orogens. Of particular interest are our examples of the crust–mantle transition of former passive margins that reveal a continuum ranging from an attenuated continental crust (e.g., Ivrea-Verbano/Strona Ceneri basement) all the way to the hyperex- tended crust of the Eastern Alps of Switzerland. In this context, the comparison with present-day distal passive margin settings has turned out to be particularly fruitful. The advancing deep-water frontier exploration for hydrocarbons also keeps adding new insights that are relevant to an understanding of their allochtho- nous remnants in folded belts (e.g., Flinch et al., 2009; Mohriak et al., 2010; Unternehr et al., 2010). The outcrops of the Ivrea Zone have been converted into seismic P-wave velo- cities and into synthetic seismic profiles of an attenuated continental crust and its basal transition into the mantle. It is now tempting to relate this contin- uum from an attenuated to a hyper-extended reconstructed Alpine passive margin and compare it to cratonic and orogenic crustal profiles of the type assembled by Hammer et al. (2011). Turning to the idealized cartoon of Plate 25.17, there now remains the question whether the basement wedge underly- ing the “Internides” was attenuated during the pro-orogenic rifted passive mar- gin phase or else during syn-orogenic or late compressional phases. The latter would be associated either with the emplacement of accreting terranes and/ or else the emplacement of basement-involved compressional structures. In addition, the basement wedge could also be attenuated during late to postoro- genic extension. The answer to these questions will determine the true age of the Moho and the underlying lithospheric mantle. All these ages may differ greatly from the ages of accreted terranes, that is, the inception and/or the completion of the final accretion of an orogen. The ages of orogens as shown on Plates 25.15 and 25.20 denote large age brackets implicit in our definition of basement that only focuses on last age of its reworking and the, obviously polyphase, peneplaned unconformity that underlies a sedimentary basin. In this context, the re-working of a lower crust mantle transition basement by late to post- orogenic extension is an important issue that, whenever possible, needs documentation. With regard to extensional shear planes in the lower crust–upper mantle transi- tion, it is worth noting that inclined reflectors observed on crustal profiles are com- monly interpreted as compressional reverse faults or shear zones. Given the experience of the Ivrea zone, it may be challenging to determine whether inclined reflectors on crustal profiles are imaging segments of compressional or extensional faults. 1087 References to 25.10Tectonic settings 1088 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps of mafic/ultramafic oceanic and intraoceanic arc system crust, LIPs, rifted and volcanic passive margins, tectonic setting and discussion of equivalent allochthonous “ophiolitic” fragments in orogens Afonso, J.C., Zlotnick, S., 2001. The subductability of continental lithosphere: the before and after story. In: Brown, D.R., Ryan, P.D. (Eds.), Arc-Continent Collision. Frontiers in Earth Sciences, Springer, Berlin, pp. 53–86. Anonymous, 1972. Penrose field conference on ophiolites. Geotimes 17, 24–25. Bernoulli, D., Jenkyns, H.C., 2009a. Ancient oceans and continental margins of the Alpine- Mediterranean Tethys: deciphering clues from Mesozoic pelagic sediments and ophiolites. Sedimentology 56, 149–190. Bernoulli, D., Jenkyns, H.C., 2009b. From the Steinmann Trinity to sea floor spreading. Compt. Rend. Geosci. 341, 363–381. Bernoulli, D., Manatschal, G., Desmures, L., Mu¨ntener, O., 2003. Where did Gustav Steinmann see the Trinity? Back to the roots of the Alpine ophiolite concept. In: Dilek, Y., Newcombe, S. (Eds.), Ophiolite Concept and the Evolution of Geologic Thought. Geological Society of America. Spec. Paper 373, pp. 93–100. Bertotti, G., Stiletto, G.B., Spalla, M.I., 1993. Deformation and metamorphism associated with crustal rifting: the Permian to Liassic evolution of the Lake Lugano-Lake Como area (Southern Alps). Tectonophysics 280, 185–197. Boillot, G., Froitzheim, N., 2001. Non-volcanic rifted margins continental breakup and the onset of seafloor spreading: some outstanding questions. In: Wilson, R.C.L., Whitmarsh, R.B., Taylor, B., Froitzheim, N. (Eds.), Non volcanic Rifting of Continental Margins. Geological Soci- ety London Special Publication, 187, 586pp. Boriani, A., Giobbi-Oregoni, E., Borghi, A., Caironi, V., 1990. The evolution of the “Serie dei Laghi” (Strona-Ceneri and Scisti dei Laghi): the upper component of the Ivrea Verbano crustal sec- tion; Southern Alps, North Italy and Ticino, Switzerland. Tectonophysics 182, 103–118. Brown, D.R., Ryan, P.D., et al., 2011. Arc continent collision: the making of an Orogen. In: Brown, D.R., Ryan, P.D. (Eds.), Arc-Continent Collision. Frontiers in Earth Sciences. Springer, Berlin, pp. 477–493. Callot, J.P., Laurent, G., Brun, J.P., 2002. Development of volcanic passive margins: Three- dimensional laboratory model. Tectonics 21, 13. Chazot, G., Charpentier, S., Kornprobst, J., Vannucci, R., Luais, B., 2005. Lithospheric mantle evo- lution during continental break-up: theWest Iberia Non-Volcanic PassiveMargin. J. Petrol. 46, 2527–2568. Cloos, M., 1993. Lithospheric buoyancy and collisional orogenesis: subduction of oceanic plateaus, continental margins, island arcs, spreading ridges, and seamounts. Geol. Soc. Am. 105, 715–737. Desmurs, L., Manatschal, G.R., Bernoulli, D., 2001. The Steinmann Trinity revisited: mantle exhu- mation along an ocean-continent transition; the Platta nappe, eastern Switzerland. In: Wilson, R.C.L., Whitmarsh, R.B., Taylor, B., Froitzheim, N. (Eds.), Non volcanic Rifting of Con- tinental Margins. Geological Society London Special Publication, 187, pp. 235–266. Dewey, J.F., 2003. Ophiolites and lost oceans: rifts, arcs and/or scrapings? In: Dilek, Y., Newcomb, S. (Eds.), Ophiolite Concept and the Evolution of Geologic Thought. Geological Society of America Special Paper 373, pp. 153–158. Dilek, Y., 2003. Ophiolite concept and its evolution. In: Dilek, Y., Newcomb, S. (Eds.), Ophiolite Concept and the Evolution of Geological Thought. Special Paper 373, Geological Society of Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps America, pp. 1–16. Dilek, Y., Flower, M.F.J., 2003. Arch-trench rollback and forearc accretion: a model template for ophiolites in Albania, Cyprus and Oman. In: Dilek, Y., Robinson, J. (Eds.), Ophiolites in Earth History, Geological Society London Special Publication, 218, pp. 43–68. Dilek, Y., Furness, H., 2011. Ophiolite genesis and global tectonics: geochemical and tectonic fin- gerprinting of ancient oceanic lithosphere. Bull. Geol. Soc. Am. 1234, 387–411. Dilek, Y., Newcomb, S. (Eds.), 2003. Ophiolite Concept and the Evolution of Geological Thought, Geological Society of America Special Paper 373, 170 pp. Dilek, Y., Robinson, P.T., 2003. Introduction. In: Dilek, Y., Robinson, P.T. (Eds.), Ophiolites in Earth History, Geological Society London Special Publication, 218, pp. 1–8. Dilek, Y., Picardo, G.B., 2008. Alpine Definition of Ophiolites and Oceanic Lithosphere. Joint Meeting of the Geological Society of America with Gulf Coast section of SEPM, Paper 245– 6 (Abstract). Dilek, Y., Ernst, W.G. (Eds.), 2008a. Links between ophiolites and Large Igneous Provinces (LIPs) in earth history. Lithos, 100, 345 p. Dilek, Y., Ernst, W.G., 2008b. Links between ophiolites and Large Igneous Provinces (LIPs) in earth history; Introduction. In: Dilek, Y., Ernst, G.R. (Eds.), Lithos, 100, 1–13. Flinch, J.F., et al., 2009. The Sierra Leone–Liberia Emerging Deepwater Province. #20224. AAPG Search and Discovery. Flower, M.F.J., Dilek, Y., 2003. Ophiolite pulses, mantle plumes and orogeny. In: Dilek, Y., Robinson, J. (Eds.), Ophiolites in Earth History, Geological Society London Special Publication, 218, pp. 1–19. Fountain, D.M., 1976. The Ivrea Verbano and Strona Ceneri zones northern Italy: a cross section across the continental crust-new evidence from seismic velocities. Tectonophysics 33 (145), 166. Froitzheim, N., Derks, J.F., Walter, J.M., Sciunnach, D., 2008. Evolution of an extensional brittle detachment from syn-intrusive, mylonitic flow to brittle faulting (Grassi Detachment fault), Orobic anticline, southern Alps, Italy. In: Siegesmund, S., Fu¨genschuh, B., Froitzheim, N. (Eds.), Tectonic Aspects of the Alpine Dinaride-Carpathian system, Geological Society London Special Publication, 298, pp. 69–82. Glennie, K.W., Beouf, M.G., Hughes-Clarke, M.H.W., Moody Stuart, M., Pilaar, W.F., Reinhardt, B.M., 1973. Late Cretaceous nappes in the Oman Mountains and their geologic evolution. Am. Assoc. Petrol. Geol. Bull. 57, 5–27. Glennie, K.W., Boeuf, M.G., Hughes-Clark, M.H.W., Moody-Stuart, M., Pilaar, W.F., Reinhardt, B.M., 1974. Geology of the Oman Mountains. Verhandelingen Koninklijk Nederlands Geologisch Mijbouwkundig Genootschap 31, 423 pp. Glennie, K.W., Hughes-Clarke, M.H.W., Boeuf, M.G., Moody Stuart, M., Pilaar, W.F., Reinhardt, B.M., 1990. Inter-relationship of Makran-Omanmountain belts of convergence. In: Robertson, A.H.F., Searle, M.P., Ries, A.C. (Eds.), Geological Society, London Special Publication, 48, pp. 773–786. Greene, A.R., Scoates, J.S.,Weiss, D., 2008a. The Accreted Late TriassicWrangelliaOceanic Plateau in Alaska, Yukon and British Columbia. December 2008 LIP of the Month1–14. http://www .largeigneousprovinces.org/08decc. Greene, A.R., Scoates, J.S., Weiss, D., 2008b. The architecture of oceanic plateaus revealed by the volcanic stratigraphy of the accreted Wrangellia oceanic plateau. Geosphere 6, 47–73. Gutscher, M.A., Olivet, J.L., Aslanian, D., Eissen, J.P., Maury, R., 1999. The lost Inca Plateau: Cause of flat subduction beneath Peru? Earth Planet. Sci. Lett. 171, 335–341. Handy,M.R., Franz, L., Heller, F., Janott, B., Zurbriggen, R., 1999.Multistage accretions and exhu- mation of continental crust (Ivrea crustal section, Italy and Switzerland). Tectonics 18, 1154–1177. 1089 Handy, M.R., Zingg, A., 1991. The tectonic and rheological evolution of an attenuated cross sec- tion of the continental crust: Ivrea crustal section, southern Alps. northwestern Italy and southern Switzerland. Geol. Soc. Am. Bull. 103, 236–253. 1090 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Hammer, P.T.C., Clowes, R.M., Cook, F.A., Vadsudevan, K., van der Velden, A.J., 2011. The big picture: a lithospheric cross section of the North American continent. GSA Today 21, 4–10. Hopson, C.A., 2007. Subvolcanic sheeted sills and non-sheeted dikes in ophiolites: occurrence, origin and tectonic significance for oceanic crust generation. In: Cloos, M., Carlson, W.D., Gilbert, M.C., Liou, J.G., Sorensen, S.S. (Eds.), Convergent Margin Terranes and Associated Regions. A Tribute to W.G. Ernst, Geological Society of America Special Paper 419, pp. 225–254. Hopson, C.A., Mattinson, J.M., Passagno, E.A., Luyendyk, B.P., 2008. California coast range Ophiolite: composite middle to late Jurassic oceanic lithosphere. In: Wright, J.E., Shervais, J.W. (Eds.), Ophiolites, Arcs and Batholiths: A Tribute to Cliff Hobson. Geological Society of America Special Paper 438, pp. 1–101. Jaillard, E., Lapierre, H., Ordonez, M., Toro-A´lava, J., Amo´rtegui, A., Van Melle, J., 2009. Accreted terranes in Ecuador: southern edge of the Caribbean Plate? In: James, K.H., Lorente, M.A., Pindell, J.L. (Eds.), The Origin and Evolution of the Caribbean Plate. Geological Society London Special Publication, 328, pp. 469–485. Kerr, A.C., Tarney, J., Nivia, A., Marriner, G.F., Saunders, A.D., 1998. The internal structures of oce- anic plateaus: inferences from obducted Cretaceous terranes in western Colombia and the Caribbean. Tectonophysics 219, 173–188. Kerr, A.C., Pearson, D.G., Nowell, G.M., 2009. Magma source evolution beneath the Caribbean Oceanic plateau: new insights from elemental and Sr-Nd-Hf isotopic studied of ODP Leg 165 Site 1001 basalts. In: James, K.H., Lorente, M.A., Pindell, J.L. (Eds.), The Origin and Evolution of the Caribbean Plate. Special Publication, 328, pp. 809–827. Khazanehdari, J., Rutter, E.H., Brodie, K.H., 2000. High-pressure-high-temperature seismic veloc- ity structure of the midcrustal and lower crustal rocks of the Ivrea-Verbano zone and Serie dei Laghi. NW Italy. J. Geophys. Res. Solid Earth 105, 13843–13858. Liu, L., Gurnis, M., Seton, M., Saleeby, J., Mu¨ller, R.D., Jackson, J.M., 2010. The role of ocean pla- teau subduction in the Laramide orogeny. Nat. Geosci. 3, 353–357. Manatschal, G., 1995. Jurassic rifting and Formation of a passive continental margin (Platta and Err nappes, Eastern Switzerland). Geometry, kinematics and geochemistry of fault rocks and comparison with Galicia margin. Dissertation E.T.H. Nr. 11188, Zu¨rich. Manatschal, G., 2004.Newmodels for evolution ofmagma-poor riftedmargins based on a review of data and concepts from West Iberia and the Alps. Int. J. Earth Sci. (Geol. Rundsch.) 93, 432–466. Manatschal, G., Bernoulli, D., 1999. Architecture and tectonic evolution of nonvolcanic margins: present-dayGalicia and ancient Adria. Tectonics 18, 1099–1119, doi:10.1029/1999TC900041. Manatschal, G., Mu¨ntener, O., Lavier, L.L., Minshull, T.A., Pe´ron-Pinvidic, G., 2007. Observations from the Alpine Tethys and Iberia–Newfoundland margins pertinent to the interpretation of continental breakup. In: Karner, G.D., Manatschal, G., Pinheiro, L.M. (Eds.), Imaging Mapping, Modelling Continental Lithosphere Extension and Breakup, Geological Society London Special Publication, 282, pp. 291–324. Mann, P., Gahagan, L., 2004a. Plate Animation Showing the Interaction Between the Northern Melanesian Arc System and the Ontong Java Plateau 20-0 Ma. Frames 3–26. Institute of Geo- physics University of Texas, Austin. Mann, P., Taira, A., 2004b. Global tectonic significance of the Solomon Islands and the Ontong Java Plateau. Tectonophysics 389, 137–190. Mann, P., Taira, A. (Eds.), 2004. Tectonics, Seismicity, and Crustal Structure of the Ontong-Java Plateau-Solomon Islands Convergent Zone, Southwest Pacific Ocean, Special Issue of Tecto- nophysics, 389, pp. 125–307. Mohriak, W.U., No´brega, M., Odegard, M.E., Gomes, B.S., Dickson, W.G., 2010. Petrol. Geosci. 16, 231–245. Moores, E.M., Vine, F.J., 1971. The Troodos massif. Cyprus and other ophiolites as oceanic crust: evaluations and implications. Phil. Trans. R. Soc. Lond. A 268, 433–466. Pe´ron-Pinvidic, G., Manatschal, G., 2010. From microcontinents to extensional allochthons: wit- nesses of how continents rift and break apart? Petrol. Geosci. 16, 189–197. Petterson, M.G., 2010. A review of the tectonics of the Kohistan Island Arc, north Pakistan. In: Kusky, T.M., Zhai, M.G., Xiao, W. (Eds.), The Evolving Continents: Understanding Processes of Continental Growth, Geological Society London, 338, pp. 278–327. Ramos, V.A., Folguera, A., 2009. Andean flat-slab subduction through time. In: Murphy, J.B., Keppie, J.D., Hynes, A.J. (Eds.), Ancient Orogens and Modern Analogues, Geological Society London Special Publication, 327, pp. 31–54. Robinson, P.T., Malpas, J., Xenophontos, C., 2003. The Troodos Massif of Cyprus: its role in the Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps evolution of the ophiolite concept. In: Dilek, Y., Newcomb, S. (Eds.), Ophiolite Concepts and the Evolution of Geological Thought, Geological Society of America Special Paper 373, pp. 295–308. Rutter, E.H., Khazanehdari, J., Brodie, K.H., Blundell, D.J., Waltham, D.A., 1999. Synthetic seismic reflection profile through the Ivrea zone-Series dei Laghi continental crustal section, north- western Italy. Geology 27, 78–82. Rutter, E.H., Brodie, K., James, T., Blundell, D.J., Waltham, D.A., 2003. Seismic Modeling of the lower andmid-crustal structure as exemplified by theMassiccio dei Laghi (Ivrea zone and Serie dei Laghi) crustal section. In: Goff, A.J., Holliger, K. (Eds.), Heterogeneity in the Crust and Upper Mantle: Scaling and Seismic Properties. Kluwer Academic/Plenum Press, pp. 67–97. Schmid, S.M., Zingg, A., Handy, M., 1987. The kinematics of movement along the Insubric Line and the emplacement of the Ivrea zone. Tectonophysics 135, 47–66. Siegesmund, S., Layer, P., Dunkl, I., Vollbrecht, A., Steeken, A., Wemmer, K., et al., 2008. Exhu- mation and deformation history of the lower crustal section of the Valstrona di Omegna in the Ivrea Zone southern Alps. In: Siegesmund, S., Fu¨genschuh, B., Froitzheim, N. (Eds.), Tectonic Aspects of the Alpine-Dinaride-Carpathian system, Geological Society London Special Publi- cation, 298, pp. 45–68. Steinmann, G., Bernoulli, D., Friedman, G.M., 2003. Die ophiolitischen Zonen in den mediterra- nen Kettengebirgen (English translation). Geol. Soc. Am. Special Paper 373, pp. 77–91. Stern, R.J., 2010. The anatomy and ontogeny ofmodern intra-oceanic arc systems. In: Kusky, T.M., Zhai, M.G., Xiao, W. (Eds.), 2010. The Evolving Continents: Understanding Processes of Con- tinental Growth, Geological Society London Special Publication, 338, pp. 7–34. Tucholke, B.E., Sawyer, D.S., Sibuet, J.C., 2007. Breakup of the Newfoundland Iberia rift. Geol. Soc. Lond. Special Publication 282, 9–46. Unternehr, P., Pe´ron-Pinvidic, G., Manatschal, G., Sutra, E., 2010. Hyper-extended crust in the South Atlantic in search of a model. Petrol. Geosci. 16, 207–215. Wilson, R.C.L., Whitmarsh, R.B., Taylor, B., Froitzheim, N. (Eds.), 2001. Non-volcanic rifting of continental margins. Geological Society London Special Publication, 187, 586 pp. 1091 PLATES FOR TECTONICS, OROGENIC SYSTEMS, HOT SPOTS, LIPS, VOLCANOES: Segments 25.6–25.10 (For online version of the plates/figures cited in this chapter, the reader is referred to http://www.elsevierdirect.com/ companion.jsp? ISBN=9780444563576). Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Continents Folded (Orogenic) Belts Ages of Deformation Oceanic Basins “Foreland” Folded Belts Oceanic (B) Subduction Boundary; On Plate 18 only: Narrowing or non accreting (erosional) margin Oceanic Subduction Boundary; On Plate 18 only: Growing or accreting margin Continental (A) Subduction Boundary Major Strike Slip Faults Basement - Involved Compressional Uplifts and /or Inversions Cenozoic / Mesozoic Vendian to Paleozoic Precambrian Basement Precambrian Shield Passive Margins: Ocean / Continent Boundary Mid Ocean Ridge Oceanic Transform Zones Oceanic Back Arc Basins Ages of Oceanic Crust Neogene Paleogene Cretaceous Jurassic Rifts and Basaltic Plateaus Neogene Rifts a cb a. undifferentiated b. poorly defined c. with volcanic seaward-dipping reflectors (SDRS) a. Cenozoic/ Mesozoic b. Paleozoic Basaltic Plateaus a. Outcrops b. Subsurface a. a. b. b. Deccan Passive Margins with extensive volcanic cover Foreland Fold and Thrust Belts (FFTB’s) Retro -wedge or occasional pro- wedge in the case of continental collisions Active Margin Fold and Thrust Belts (AMFTB’s) “Pro-wedge” A.W. Ballyc Plate 25.14 Legend for Plates, 15, 16a & b, 18. 1 0 9 3 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s Credits: Simplified from CGMW (1967 and 2000) Exxon (1985), and Haghipour (2001) 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W A.W. Ballyc Plate 25.15 Tectonic Map of the World. 1 0 9 4 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s E W W E Credits for Arctic Map: CGMW (1967), Harrison (2005), Reed 2005, 2006 and others. Credits for Antarctic Map: Choubert and Faure Muret CGMW (1967), and Anderson J. (1999 A&B) A.W. Ballyc A.W. Ballyc Plate 25.16 A. Tectonic Map of the Arctic and B. Tectonic Map of the Antarctic. 1 0 9 5 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s I II IVIII ACCRETIONARY PRISM FORE-ARC BASIN I-IV TERRANES ARC FORMER ACCRETIONARY PRISM (INCLUDING OPHIOLITIC SUTURES) METAMORPHIC CORE COMPLEXES BASEMENT THRUST FORELAND FOLD AND THRUST BELTS THRUSTED FORELAND BASEMENT UPLIFTS OCEAN - VERGING EXTERNIDES CONTINENT - VERGING EXTERNIDESINTERNIDES O R O G E N I C F L O A T A.W. Ballyc Plate 25.17 Orogenic Float. 1 0 9 6 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W A.W. Ballyc Plate 25.18 Cenozoic/Mesozoic Orogens, Active Margin Fold and Thrust Belts (AMFTBs), Foreland Fold and Thrust belts (FFTBs), Basement Uplifts and Cenozoic Rifts. 1 0 9 7 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S Plate 25.19 Cenozoic/Mesozoic and Paleozoic Foreland Fold and Thrust belts (FFTBs) and age of Continental Basement. 1 0 9 8 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s ? ? Archean (2500 - 4000) Paleozoic (251 - 553 Ma) (incl. some Vendian) Paleozoic and Precambrian orogenic assemblages reworked during the Mesozoic / Cenozoic (only for China and Mongolia) Cenozoic - Mesozoic (0 - 251 Ma) Early Proterozoic (1600 - 2500 Ma) Mid Proterozoic (1000 - 1600 Ma) Late Proterozoic (553 Ma - 1000 Ma) Credits: Goodwin (1966), Trompette (1994) , Artemieva and Mooney (2002), Kröner and Stern (2005), Cordani et al (2000) and many others (see List of References) 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S Plate 25.20 Age of Continental Basement. 1 0 9 9 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s List of Hotspot Names North America Pacific Ocean Indian Ocean Basaltic Plateaus “Hotspots” Sea Mounts, Islands and Plateaus Volcanic Plateaus/Ridges covering attenuated passive margin crust BA Balleny BO Bowie CA Caroline CO Cobb EI Easter Island GA Galapagos GL Guadalupe HA Hawaii JF Juan Fernandez LH Lord Howe LV Louisville MA Marquesas MC MacDonald MG Marshall - Gilbert Islands MU Musicians Hotspot PI Pitcairn SA Samoa SF San Felix SC Socorro (Revillagedos) SG Sala y Gomez TA Tahiti (Society Islands) TS Tasman CO Comores CR Crozet EA East Australia KE Kerguelen MA Marion RE Reunion SP St. Paul A Anahim RA Raton YE Yellowstone Europe EI Eiffel Africa A Afar CA Cameroon HO Hoggar Mtns DA Darfur (Jbel Marra) TI Tibesti Atlantic Ocean AS Ascension AZ Azores BE Bermuda BO Bouvet IC Iceland JM Jan Mayen CA Canary CV Cap Verde FE Fernando de Noronha GO Gough MD Madeira ME Great Meteor SH St. Helena TC Tristan da Cunha TR Trinidade (Martin Vaz) VE Vema Dike Swarms of inferred earlier Plume Heads Convergence Point of Dike Swarm Age Tertiary Mesozoic Paleozoic Late Proterozoic Mid Proterozoic Early Proterozoic a. Outcrops b. Subsurface a. b. Deccan Shatski Passive Margins with extensive volcanic cover Primary Hotspots listed by Courtillot et al (2005) and Foulger (2007) On Oceanic Crust: On Continents & Passive Margins: CM EI Easter Island Source at Core/Mantle BoundaryNeogene Rifts Ocean / Continent Boundary a cb a. undifferentiated b. poorly defined c. with volcanic seaward-dipping reflectors (SDRS) Plate 25.21 Legend for Plates 22-25; LIPs, Hot Spots, Volcanic passive margins and Tertiary rifts. 1 1 0 0 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s LV MC SA TA MA MU PI EI SG SF GA SC GL RA BE AZ ME MD CA CV FE TR TC GODI BO VE MA CR KE SP BA TA EA LH CA RE CO AFDA TI HO EI IC JM CM AS SH YE CO BO AN HA JF Credits: Norton (2000), Steinberger (2000) and Wolbern (2007): 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S Plate 25.22 Hot Spots with abbreviated names. 1 1 0 1 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s LV MC SA TA MA MU PI EI SG SF GA SC GL RA BE AZ ME MD CA CV FE TR TC GODI BO VE MA CR KE SP BA TA EA LH CA RE CO AFDA TI HO EI IC JM CM AS SH YE CO BO AN HA JF 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S Credits: Steinberger (2000) and Wolbern (2007): Plate 25.23 Hot Spots and Tectonics. 1 1 0 2 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s LV MC SA TA MA MU PI EI SG SF GA SC GL RA BE AZ ME MD CA CV FE TR TC GODI BO VE MA CR KE SP BA TA EA LH CA RE CO AF DA TI HO EI IC JM CM AS SH YECO BO AN HA JF Deccan Tunguska West Siberia Shatski Emperor Seamounts Ontong - Java Kerguelen Etendeka Karoo Parana Manihiki Broken Ridge Rajmahal Naturaliste Credits: Coffin and Eldholm (1992 & 1994), Eldholm and Coffin (2000) and CGMW (2000) 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S Plate 25.24 Large Igneous Provinces (LIP’s), Volcanic island chains and Hot Spots. 1 1 0 3 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s West Australia Widgiemootha Willouran Rajmahal Sri Lanka X Deccan Afar Madagascar Karoo Parana Caribbean Central Atlantic (Camp) Etendeka Gannakuriep Central Japetus (Camp) Mistassini Matachewan Columbia River Abitibi Anakim Mackenzie Franklin Alpha Ridge Ungava Jutland Kola-Onega Siberia YakutskNorth Atlantic Credits: Simplified from Ernst and Buchanan (1997 & 2001) combined with Bally (2010) Plate 15 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S Plate 25.25 Dyke swarms by age (with names) and Age of Continental Basement. 1 1 0 4 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s Credits: Smithsonian Institution - Global Volcanism Program : http:www//volcano.si.edu 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° 30° E30° E 60° E 90° E 120° E 150° E 180° 150° W 120° W 90° W 60° W 30° W 0°30° E 30° E60° E 90° E 120° E 150° E 180° Submarine Subglacial Hydrothermal Subaerial Plate 25.26 Active Volcanoes, Plate Boundaries and Tertiary Rift systems. 1 1 0 5 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s Credits: Simplified from CGMW (1967 and 2000) Exxon (1985), and Haghipour (2001) 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W A.W. Ballyc Plate 25.27 Cenozoic, Mesozoic, Paleozoic Orogenic systems; Continental basement; Foreland Fold and thrust belts (FFTB’s); Active Margin Fold and Thrust Belts (AMFTB’s); Age of Oceanic Basement formed onMid-Oceanic Ridges (MOR) and also in oceanic backarc basins and Large Igneous Provinces (LIPs). 1 1 0 6 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s 2255..1111 Sedimentary basins and rifts (including Rifts)* Table 25.7 Simplified basin classification and list of plates Legend for sedimentary basin maps (including Arctic and Antarctic) Rifts by age and age of continental basement Sedimentary basins of the world Arctic basins and Antarctic basins 1. Basins on relatively stable/rigid lithosphere 1.1. Rift systems Rifts by age and age of continental basement Tertiary rift systems and recent plate motions 1.1.1. Continental rift systems 1.1.2. Oceanic rift systems 1.2. Passive margins Passive margins, age of basement, and Cenozoic/Mesozoic rifts Passive margins. Cenozoic/Mesozoic rifts and radiating dike swarms 1.2.1. Passive margins or divergent/atlantic-type margins 1.2.2. Transform passive margins 1.2.2.1. Transtensional passive margins 1.2.2.2. Transpressional passive margins Cratonic and foreland basins 1.3. Cratonic basins 2. Basins on the periphery of orogenic belts Deep sea trenches foreland basins, foreland fold and thrust belts (FFTBs), and episutural basins 2.1. Deep-sea trenches 2.2. Remnant ocean basins 2.3. Foreland and associated basins 2.3.1. Simple Foreland basin s.s. or foredeeps 2.3.1.1. Pro-foreland basins 2.3.1.2. Retro-foreland basins 1108 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps 2.3.1.3. Wedge-top basin associated with foreland fold and thrust belts (FFTBs)a 2.3.2. Composite foredeep/foreland basins segmented by basement upliftsb 2.3.2.1. US. Rocky Mtn.-type foreland basins segmented by compressional/transpressional basement uplifts or inversions 2.3.2.2. Central asia-type foreland basins segmented by compressional/transpressional basement uplifts Table 25.7 Simplified basin classification and list of plates—cont’d Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps 3. Basins located within orogenic belts (episutural basins of Bally and Snelson, 1980)c Foreland basins, foreland fold and thrust belts (FFTBs), and episutural basins 3.1. Basins associated with intra-oceanic subduction 3.1.1. Intra-oceanic forearc basin 3.1.2. Intra-oceanic backarc basin 3.2. Basins associated with oceanic/continent (OC) subduction related island arcs 3.2.1. Forearc basins 3.2.2. Backarc basins 3.3. Mostly continental backarcs basins associated with oceanic subduction 3.3.1. Collision-associated backarcs basins with continental crust 3.3.2. Collision-associated backarc basins on highly extended continental crust and partly oceanic crust Introduction Plates 25.31 and 25.32A,B show the world’s sedimentary basins on Mercator and Polar projections with an accompanying legend on Plate 25.28. Plate 25.29 (discussed in more detail in Section 3.10.2) may be considered as part of this set of mapsas itdepicts theglobaldistributionof rifts,manyofwhichhostprolificpetroleum systems. On Plates 25.27–25.36, the outlines of sedimentary basins aremostly based on the Exxon Tectonic Map (Exxon Production Research, 1985). It should be noted that generalized basin outlinemaps are also published by IHS (2010), Fugro-Robertson (2009), St. John et al. (1984), and St. John (2010). These maps all are designed 4. Active margin basins associated with major transform and/or strike-slip fault systems 4.1. Mostly transtensional active margin basinsd 4.2. Mostly transpressional active margin basinsd 5. Oceanic basins formed by spreading mid-ocean ridgese Topography and bathymetry maps Arctic topography and bathymetry Antarctic topography and bathymetry Cenozoic, Mesozoic, Paleozoic Orogenic systems; Continental basement; Foreland Fold and thrust belts (FFTB’s); ActiveMargin Fold and Thrust Belts (AMFTB’s); Age of Oceanic Basement formed onMid-Oceanic Ridges (MOR ) and also in oceanic backarc basins and Large Igneous Provinces (LIPs) aNot differentiated on maps as they are too small to show on world maps of this scale. bMay be recognized, by their association with the basement uplift/inversion symbol. cOn small-scale world maps, this basin class is not easily differentiated into “sub-types” on world maps of this scale, because they are too small to show up. They could be easily differentiated on world maps that would be more than 2.5� larger (cf. the maps of Fugro-Robertson, 2009; HIS, 2007, and of course the Exxon Tectonic Map). dThese basins are difficult to differentiate on small-scale world maps. eOceanic basin provinces are not differentiated here. 1109 to be printed at larger scales than our plates but they do not include the tectonic 1110 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps context depicted on plates 25.31 and 25.32A,B and explained on the legend (Plate 25.28). A review byMann et al. (2003) illustrates the tectonic setting of sedimentary basin close-ups extracted from the Exxon (1985) map. The USGS has published and maintains a basin map that serves as background for their ongoing studies of Global Reserve Estimates derived from Petroleum Systems analysis (Ahlbrandt et al., 2005). For additional background on petroleum sys- tems, see Chapter 23 and 24 Vol. 1A, as well as Tissot and Welte (1978), Bois et al. (1979), Klemme (1986), Klemme and Ulmishek (1991), Perrodon (1983), Tissot et al. (1987), Magoon and Dow (1994), and Biju-Duval (1999). The eco- nomic relevance of all these studies cannot be emphasized enough. Note that, for good reasons, in the publications here cited the basin outlines are often more detailed than ours, as they include economically important sub-basins, rather than the entire offshore or onshore basin. In some areas, the USGS basin outlines coin- cide with political/administrative boundaries that better fit a statistical database that changes from one country or state to another (Ahlbrandt, 2001). This refer- ence also includes a particularly useful table that assigns USGS provinces to iden- tified petroleum systems. Any basin classification, including ours, could easily be combined with that table, thus providing a first-order overview of the tectonic setting for any given hydrocarbon system. Because rifts underlie or overlap so many sedimentary basins, it will always remain a matter judgment whether a rift should be isolated as the principle “identifier” of a basin or else be viewed as a sub- basin of a larger basin. Much of this book is dedicated to Phanerozoic rifts and sedimentary basins;, there- fore, the text accompanying the Global Rift and Basin map (Plate 25.29) is limited to brief introductory comments that relate themaps to the basin classification that is discussed in Chapter 4, Vol. 1A and also Vol. 1C (Roberts and Bally) (including its Tables 25.2 and 25.3) of this volume. Segment 3.9.4 of this chapter summarizes the relations of rifting to various igneous events commonly associated with the inception of rifting. The role of rifting in the formation of passive margins is the subject of Chapter 1, Vol. 1B (and also Vol. 1C Roberts and Bally). A sedimentary basin is here defined as a depression that is filled with more than 1 km of sediments. Most, if not all basins, have had a polyphase evolution that is best analyzed by differentiating tectono-stratigraphic megasequences (TSMs) (for more, see Chapter 4, Vol. 1A (Roberts and Bally)). The sum of all TSMs is responsible for the present outline and shape of the basin. Academic scholars often model a given basin type as a function of its origin. In reality, changing far-field tectonic forces often overwhelm the initial conditions of the earliest evolution of a basin. An exclusive perspective on the origin of a basin is likely to be off-target as it often bypasses the impact of subsequent tectonic and climatic vicissitudes of the basin. It follows that the polyphase evolution of most sedimentary basins is the sum of region on Plate 25.1). Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Another example is the highly asymmetric Paris basin where the uplifted, broad eastern flank exposes the complete basin fill together with the underlying Paleo- zoic basement exposed in the Vosges massif, a horst that is associated with the Tertiary Rhine-Rhone graben system. Needless to say, where applicable, this late uplift still leaves most of the partially eroded basin to the west reasonably intact but still is an important element of Paris basin analysis. Table 25.7 is the Simplified Basin Classification used for all Basin maps with the most relevant Plates/Maps listed under its title. Examples of typical Tectono- Stratigraphic Megasequences characterizing various basin types are listed in Chapter 4, Vol. 1A (Roberts and Bally). References to 25.11 Sedimentary basins (incl. Rifts); Introduction NOTE: The references of 25.11 have been listed in 5 separate topical reference lists. Ahlbrandt, T.S., 2001. The Sirte-Zelten Total Petroleum System. US Geological Survey Bulletin, 2202-F, 29. Ahlbrandt, T.S., Charpentier, R.R., Klett, T.R., Schmoker, J.W., Schenk, C.J., Ulmishek, G.F., 2005. Global Resource Estimates from Total Petroleum Systems. AAPG Memoir 86, 324 pp. Biju-Duval, B., 1999. Ge´ologie Sedimentaire: Basins, Environnements des de´pots. Formation du Pe´trole. Collections des cours de l’ENSPM, E´ditions Technip, 736. Bois, C.P., Bouche, P., Pelte, R., 1979. Histoire Ge´ologique et Repartition des reserves dans le Monde, vol. 53, Revue Institut Franc¸ais du Pe´trole et Annnales des Combustible Liquides, Paris pp. 275–298. Exxon Production Research, 1985. Tectonic Map of the World, World Mapping Project, Scale 1:5.000 000, 20 panels, published by Am. Assoc. Petrol. Geol. HIS, 2007. Worldwide Sedimentary Basin Maps Scale 1:20 000 000. http://Ihsenergy.com. the interactions in a sequence of events, many of which have little to do with its early origins and may have been caused by far-field plate tectonic stresses. This is a matter of concern for hydrocarbon explorers who deal with the end product of a long and complex basin evolution that affects the deposition and later maturation of source beds, the distribution and diagenesis of both clastic and carbonate reservoirs, as well as changing hydrodynamic conditions. However, to keep the definition of a basin manageable, broad relatively late “epeirogenic” uplifts and tilting of entire basins and their surrounding areas are neglected in our classification. For instance, the Po foreland basin of Northern Italy mainly corresponds in outline to the low-lying relief of the Po Plain and its buried southern FFTB while the northern flank of the basin has been uplifted, tilted as a whole and now forms parts of the outcropping southern Alps. The southern Po Basin contrasts sharply with the Denver “foreland basin” which has been elevated to the “Mile High City” of Denver as part of late to postorogenic uplift (see also the topography of the uplifted 1111 IHS, 2007. Sedimentary Basins of the World, third ed. 1:20 000000. IHS, Englewood, CO. Tissot, B.P., Welte, D.H., 1978. Petroleum Formation andOccurrence: A New Approach to Oil and Gas Exploration. Springer-Verlag, Berlin, 521 pp. 1112 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Rift systems on relatively stable/rigid lithosphere Plate 25.29: Rifts by age and age of continental basement Plate 25.30: Tertiary Rift Systems and Recent Plate Motions Volume 1B is mostly dedicated to Rifts. Chapter 1, Vol. 1B and also in Volume 1C (Roberts and Bally) provides an overview of the progression from rifts to different passive margin types and a simple classification scheme with some key references, which is not repeated here. In a general sense, mid-ocean ridges can be considered as “oceanic rift systems” and are the main loci of new seafloor generation although minor spreading cen- ters occur in backarc basins such as theWoodlark basinwhere the tip of the spread- ing center is actively extending continental lithosphere (Taylor et al., 1994). In another case, the Gakkel ridge of the Arctic Ocean is actively propagating south- ward into the Laptev Sea (Chapter 11, Vol. 1B (Franke and Hinz)) and perhaps onshore southward to Lake Baikal (Chapter 37, Vol. 1B (Hus et al.)). Plate 25.29 depicts the age of rifts in terms of epochs and the age of the base- ment. Precambrian rifts seem to be confined mainly to North America, the East Klemme, H.D., 1980. Petroleum basins—classifications and characteristics. J. Petrol. Geol., 3, 187–207. Klemme, H.D., Ulmishek, G., 1991. Effective petroleum source beds of the world: stratigraphic dis- tribution and controlling depositional factors. Am. Assoc. Petrol. Geol. Bull., 75, 1809–1851. Reprinted and modified for online presentation as AAPG Search and Discovery Article # 30003 (1999). Klemme, H.D., 1986. Field size distribution related to Basin Characteristics. In: Rice, D. (Ed.), Oil and Gas Assessment-Methods and Applications, AAPG Studies in Geology, 21, pp. 85–89. Fugro-Robertson, 2009. Fugro-Tellus Sedimentary Basins of the World Map. Scale 1:25 000 000. Available from AAPG Datapages. Magoon, L.B., Dow,W.G. (Eds.), 1994. The Petroleum System from Source to Trap. AAPGMemoir 60, 655 pp. Mann, P., Gahagan, L., Gordon, M.B., 2003. Tectonic setting of the world oil and gas fields. In: Halbouty, M.T. (Ed.), Giant Oil and Gas Fields of the Decade 1990–1999. American Associa- tion of Petroleum Geologists Memoir 78, pp. 15–107. Perrodon, A., 1983.Dynamics ofOil andGasAccumulations. Elf Aquitaine/Editions Technip, 368pp. St. John, W. (Ed.), 2010. Sedimentary Provinces of the World: Hydrocarbon Productive and Non- productive, second ed (2010) AAPG/Datapages. ESRI ArcGIS ArcMap (mxd, shp and accom- panying files). St. John, W., Bally, A.W., Klemme, H.D., 1984. Sedimentary Provinces of the World— Hydrocarbon Productive and Non productive. In: American Association of Petroleum Geologists, 35 pp. Tissot, B.P., Pelet, R., Ungerer, P., 1987. Thermal history of sedimentary basins,maturation indices, and kinetics of oil and gas generation. Am. Assoc. Petrol. Geol. Bull. 71, 1445–1466. Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps European craton (Chapter 20, Vol. 1C (Eyer)), and East Siberia (Nikishin et al., 2010). Paleozoic rifts in northern South America and North Africa may be the probable remnants of the rifts that initiated the Paleozoic passive margins of North Africa and Arabia in the Early Paleozoic. Stephenson and Stovba (Chapter 15, Vol. 1B (Stephenson and Stoba)), provide a review of the Dniepr- Donets basin of Central Europe and its linkage to the Tornquist line. Important but now inverted Paleozoic rifts occur in Northwest Australia. Paleozoic rifting also controlled the initiation of the Pre-Caspian basin and Timan-Pechora basin, pre- cursors to the Uralide foreland basin. The Late Paleozoic Karroo rifts of Southern Africa and Madagascar may herald the opening of the ocean basin between East Africa and Madagascar. The Triassic rifts of Eastern North America nucleated on Appalachian-Acadian struc- tures and formed a rifted onshore/offshore Nova Scotia that created accommoda- tion space for accumulation of up to 4 km of salt (Chapter 12, Vol. 1B (Withjack et al.)). The Mesozoic rifts of Northwest Europe are composite in time and space and collectively comprise the prolific petroleum systems of the North Sea and off- shore Norway. In this area, regional Triassic rifting was succeeded by a passive ther- mal subsidence phase followed by renewed rifting in the Late Jurassic. The latter rift had propagated south from the Arctic cutting across the cratonized Caledonian orogen and the Hercynian foreland in the northern and southernNorth Sea, respec- tively (Roberts et al., 1999). This rift aborted at the end of the Jurassic and was then obliquely cut by a northward propagating, hyperextended Early Cretaceous rift associated with the opening of the North Atlantic. The course of this rift is defined by the Rockall Trough, More, and Voring Basins. The fabric of both the Jurassic and Cretaceous rifts was strongly influenced by underlying Caledonian structures (Mosar et al., 2002; Osmundsen et al., 2005; Roberts et al., 1999). The Late Jurassic rift was subject to important inversion in the southern North Sea during the Late Cretaceous and the Early Cretaceous rift and to later inversion associated with the early Tertiary onset of spreading in the Norwegian-Greenland Sea (Tsikalas et al., Chapter 5, Vol. 1C (Tsikalas et al.); Mosar et al., 2002). Early Cretaceous rifts crosscut most of Central and Northern Africa and are asso- ciated with major nonmarine petroleum systems in Brazil (Szatmari, Chapter 14, Vol. 1B (Szamatari); Mohriak and Feinstein, Chapter 7, Vol. 1C (Mohriak and Feinstein)), Angola, Gabon, Chad, Cameroon, Sudan, Egypt, and parts of Libya (Hemsted, 2003) and more recent discoveries in Uganda. The Early Cretaceous rifts of Egypt were later inverted during the Late Cretaceous and Early Tertiary. These Early Cretaceous rifts predate spreading in the South Atlantic and may be part of a northward propagating rift system that heralded the breakup of the South Atlantic. It seems likely that the overall pattern of these rifts was controlled by the structure of the underlying Pan African basement, much as it seems to have controlled the Tertiary rifts of East Africa (Daly et al., 1989). Plate 25.30 depicts the global and complex pattern of Tertiary rifts some of which are active at present. The map demonstrates the occurrence of Tertiary rifts on 1113 Pre-Cambrian cratonic crust, in Phanerozoic orogenic regions associated with slab 1114 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps roll back (e.g., China Chapter 35, Vol. 1B (L: Desheng) and Chapter 36, Vol. 1B (Hsiao and Graham); the Tyrrhenian Sea (Scrocca et al., Chapter 13, Vol. 1C (Scrocca et al.)) and in areas associated with backarc-extension and strike-slip fault systems such as Indonesia and the Andaman Sea. Tertiary rifts also occur in asso- ciation with Cordilleran strike-slip systems, notably the San Andreas Fault and Rocky Mountain trench systems here including the Basin and Range province of the Western United States, and rifts associated with Tertiary continental collision. The above brief discussion of Phanerozoic rift systems shows that simple classifica- tions such as the one shown in Chapter 1, Vol. 1B and also Vol. 1 (Roberts and Bally) conceal the great variety of tectonic settings that encompass modern and ancient rifts. It should be noted that the majority of rifts abort and do not lead to ocean spreading. Notable exceptions include the South Atlantic and the Central North Atlantic, and some of the backarc basins of Southeast Asia (e.g., the South China Sea, Woodlark Basin, and Andaman Sea) and the western Mediterranean (e.g., the Balearic Basin and the Tyrrhenian Sea). References to Rift Systems Daly, M.C., Chorowicz, J., Fairhead, J.D., 1989. Rift basin evolution in Africa: the influence of reac- tivated steep basement shear zones. Geol. Soc. Spec. Publ. 44, 309–334. Hemsted, T., 2003. Second and third millennium reserves development in African basins. Geol. Soc. Spec. Publ. 207, 9–20. Mosar, J., Eide, E.A., Osmundsen, P.T., Sommaruga, A., Torsvik, T.H., 2002. Greenland–Norway separation: a geodynamic model for the North Atlantic. Norw. J. Geol. 82, 282–301. Nikishin, A.M., Sobornov, K.O., Prokopiev, A.V., Frolov, S.V., 2010. Tectonic evolution during the Vendian and Phanerozoic. Moscow Univ. Geol. Bull Springer Link 65, 1–16. Osmundsen, P.T., Braathen, A., Sommaruga, A., Skilbrei, J.R., Nordgulen, O., Roberts, D., et al., 2005. Metamorphic Core Complexes and Gneiss Cored Culminations along the Mid Norwe- gian Margin: An Overview and Some Current Ideas. Norwegian Petroleum Society Special Publication 12, pp. 29–41. Roberts, D.G., Thompson, M., Mitchener, B., Hossack, J., Carmichael, S., Bjornseth, H.M., 1999. Paleozoic to Tertiary rift basin dynamics: mid Norway to the Bay of Biscay—a new context for hydrocarbon prospectivity in the deep water frontier. In: Fleet, A.J., Boldy, S.A.R. (Eds.), Petro- leum Geology of Northwest Europe: Proceedings of the 5th Conference. Geological Society, London, pp. 7–43. Taylor, B., Goodliffe, A., Martinez, F., Hey, R., 1994. Continental rifting and initial sea floor spread- ing in the Woodlark Basin. Nature 374, 534–537. Passive margins on relatively stable/ rigid lithosphere Plate 25.33: Passive margins, Age of Basement and Cenozoic/Mesozoic Rifts Plate 25.34: Passive margins, Cenozoic/Mesozoic Rifts and Radiating Dike Swarms A very basic subdivision of passive margins is listed on Table 25.7 of subchapter 25.11. A more elaborate subdivision of passive margin types and their associated tectonostratigraphic megasequences is given on Table 4.3 of Chapter 4, Vol. 1A (Roberts and Bally). The evolution of passive margins from rifting stages to a variety of passive margins types is discussed in some detail and Chapter 1 Vol. 1B (also reprinted as Chapter 1, of Vol. 1C). That chapter also has references to several books on passive margins that were published in previous years. Much of Volume C contains summaries of selected passive margins. The term passive (i.e. Atlantic-type) margin was originally introduced to empha size the relative tectonic stability of these continental margins in stark contrast to the oceanic subduction related, earthquake-prone active continental margins However, the accepted designation “passive” may be somewhat deceiving because so many passive margins are affected by extraordinary gravitationa tectonics. Areas most affected by gravitational tectonics are not flagged on Plates 33 and 34. In the following, some issues associated with salt and/or shale-based linked extensional-compressional tectonics will be briefly reviewed These margins are particularly important for petroleum geologists as they contain some of the largest hydrocarbon reserves of the world. They are also likely to Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps remain of continued exploration interest for many years to come. Table 25.8 dif- ferentiates a variety of settings and also for comparison, similar tectonics asso- ciated with oceanic and continental backarc basins, that is, an orogenic setting. For an introduction to salt and shale tectonics, we refer to Hudec and Jackson in Chapter 2, Vol. 1C andWoods in Chapter 3 of Vol. 1C. Additional recent overviews include salt tectonics with global maps published by Hudec and Jackson (2007) and Morley et al. (2011), who thoughtfully compared, contrasted, and classified Table 25.8 Gravitational, linked extensional, and compressional systems on passive margins contrasted with similar systems in late orogenic settings 1. Associated with deltaic depocenters systems, mostly on passive margins 1.1. Mostly shale based For example, Namibia, Niger Delta 1.2. Associated with late/post orogenic uplift of FFTBs For example, Mexican Ridges (Eastern Mexico offshore) 1.3. Mostly salt based For example, Eastern U.S. Gulf of Mexico, Kwanza Basin, Angola 1.4. Mixed salt and shale based For example, Central and Western U.S. Gulf of Mexico 2. Associated with orogenic settings (for comparison only) 2.1. Evaporites associated with marine backarc basins For example, Gulf of Lyons 2.2. Discordantly overlying, and in part coeval with foreland fold and thrust belts For example, Reforma-Sonda Campeche area of southern Mexico 2.3. Associated with “backarc basins” overlying the internal zones of an orogen For example, Transylvania, salt-based FTB 1115 , - . l . deep-water fold thrust belts. Many papers include various examples of linked 1116 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps extensional/compressional/gravitational systems on passive margins and also dif- ferentiated salt- from shale-based systems. However, Morley et al. (2011) also chose to include a selection of examples from subduction- and collision-related fold belts associated with active margins. Because the deep-water compressional folded belts are comparable with selected profiles of decollement foreland folds, they have attracted the interest of structural geologists. Obviously, the tectonic setting of these applies to active margins and so they are excluded here. Mitra (2002) modeled faulted detachment folds associated with deep-water fold belts and FFTBs and notes that these structures should not be confused with fault prop- agation folds. The continental slope of the Orange Basin (South/Africa, Namibia) with its imbricated “Namibia thrust belt” is perhaps one of the simplest examples of a shale-based linked gravitational system. de Vera et al. (2009) noted the discrepancy between updip extension (24 km) and downdip shortening (16 km). Butler and Paton (2010) suggest that the missing compressional strain may be distributed in the form of lateral compaction in the partially lithified strata. Burke and Gunnell (2008) review the uplift and erosional record of the “African erosion surface” and note that the coastal area of Namibia over much of the timewas relatively low lying andmost of the coastal uplift occurred during the last 30 Ma. Note that the location of much of the Namibia thrust belt itself is mostly on oceanic crust. The relatively so much more complex Niger delta is the best documented shale- based “linked extensional/compressional gravitational prototype” (e.g., Ajakaye and Bally, 2002; Briggs et al., 2009; Doust and Omatsola, 1990; Maloney et al. 2010; Reijers et al., 2007). The deep-water fold belt together with much of the adjacent extensional Niger Delta is emplaced on oceanic and transitional crust. Any balancing of cross-sections will likely remain problematic until better 3D seis- mic resolution is available to reveal possible structures within the thick underlying overpressured shales. Contrast the Orange Basin linked system with the Mexican ridges, a simple shale- based linked post Laramide system that appears to be coeval with the subsequent Neogene uplift of the Mexican plateau and its underlying continental basement (Buffler, 1983; Rodriguez and Mann, 2011). Le Roy et al. (2008) observe a deep seated reverse basement fault zone that flattens toward the Moho and interpret this as transpressional fault. Similar ill-defined faults were also noted under the west flank of the Burgos basin to the north (Perez-Cruz, 1993) where the uplift of the adjacent Mexican plateau is post-Oligocene and clearly the main cause for the eastward-dipping west flank of the Burgos basin. Based on paleothermal data, Alzaga-Ruiz et al. (2009) suggest that up to 4 km of Late Cretaceous and Eocene clastics were eroded, prior to the deposition of the Late Paleogene Neo- gene sediments that are involved in the extensional-compressional structures of the Mexican ridges. Because there is no major depocenter immediately to the west, the sediments involved in this system were probably carried in by coast- Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps parallel currents from the deltaic areas farther north. Duval et al. (1992), Cramez and Jackson (2000), Marton et al. (2000), Tari et al. (2003), and Hudec and Jackson (2004, 2006) all review salt tectonics offshore Gabon and Angola. The offshore Kwanza basin serves as a good example of a “dominantly salt based prototype.” A cross-section of the Kwanza basin reveals updip Mid- to Late-Cretaceous extensional rafts, superimposed by a later Eocene to Miocene rafting episode that is associated with an over 3-km-thick and an over 300-km-wide folded and thrusted salt allochthon that occupies the distal conti- nental slope. Following Duval et al. (1992), Jackson et al. (2005) evaluate the contribution of onshore uplifts to gravitational salt tectonics in the Kwanza basin. Based on apatite fission track data from onshore samples, three thermal events are differentiated (i.e., Jurassic 150 Ma, Cretaceous 100–70 Ma associated with a first Upper Creta- ceous extensional rafting episode, and Tertiary 20–10 Ma associatedwith the third rafting episode). Overall the Tertiary uplift, located about 100 km inland culmi- nates in an arch-like uplift that separates the passive margin from the adjacent inland depressions. Burke and Gunnell (2008) specifically address the nature of these Tertiary arches and suggest that they may be best explained as “dynamic uplifts” that are associated with a new 30 Ma or less shallow mantle convection system and leave the question open whether the pace of the uplift was even or else accelerated rapidly. The northern Gulf of Mexico basin of the United States and its coastal areas is a well-explored “Mixed salt- and shale-based gravitational prototype.” Selected references include overviews by Worrall and Snelson (1989), Salvador (1991), and Galloway (2008) and additional structurally oriented papers by Diegel et al. (1995), Peel et al. (1995), Rowan (1995), Trudgill et al. (1999), Fiduk et al. (1999), and Rowan et al. (2004). The Gulf of Mexico basin outline roughly coincides with the distribution of the Mid-Jurassic Louann salt that overlies both the continental and oceanic crust separated by a poorly defined transitional zone. The northern edge of the salt coincides roughly with surface normal fault zones (e.g., the Balcones/Mexia/Talco faults system of Texas and their continua- tion into Arkansas Mississippi and Alabama). Gravitational salt tectonics, in the form of small extensional rafts occur a few kilometers from this margin of the salt basin. Moving south, a large number of classical salt diapirs rise through the Mesozoic and Tertiary section. A lower Cretaceous carbonate shelf margin sepa- rates the interior Gulf coast basin from the coastal plain and the offshore northern Gulf of Mexico. A higher shale-based de´collement level was triggered along this Lower Cretaceous “barrier” reef slope that rimmed much of the Texas/Louisiana Gulf of Mexico. The shale-based growth faulting was mostly triggered by the massive increased influx of delta associated siliciclastics from the late Cretaceous and is continuing today. 1117 The combined impact of the lower salt-based and the upper shale-based de´colle- 1118 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps ment systems resulted in the creation of a huge allochthonous salt mass of Alpine dimensions. Today, the allochthonous salt has a distinct topographic expression that is dramatically displayed on high resolution digital bathymetricmaps. The dis- tal offshore Sigsbee Escarpment marks the front of the allochthonous salt massif, which along strike terminates with “lateral ramps” when merging westward into the growth fault systems of the onshore–offshore Burgos basin and eventually with the sole faults of theMexican ridges. The eastern termination of the allochthonous salt mass and its relation to deeper salt tectonic structures were documented in some detail by Wu et al. (1990) with reconstructions by Moretti et al. (1990). The salt allochthon is by far the most spectacular feature of the Gulf of Mexico basin. However, going landward and also in the coastal zones of the Gulf of Mexico, extensional growth fault systems dominate. While reconstructions and models of individual structures are instructive, “balanced” reconstructions of long regional profiles remain elusive. Over large areas, there are few seismic constraints for the top basement surface. Reconstructing an approximate primary salt thickness by “volume balancing” would need to be done in 3D and necessarily make allowances for salt thatmay have been dissolved. Three-dimensional seismic across growth fault systems are also known to reveal coherentmappable reflectors within the overpressured shales that suggest deeper/earlier growth fault systems within the overpressured shales. These older structures discordantly underlie the much better known higher/younger growth fault systems (for good documenta- tion of older vintage growth fault seismic profiles, see Bally, 1983). Beyond the distal edge of allochthonous salt, and partly underlying it, are the salt-based Perdido and Mississippi Fan fold belts. Both fold belts rest on oceanic basement. Assuming a connection under the overhanging allochthonous salt, the fold belt may be over 200 km wide and over 600 km long, that is, covering an area that involves deformation of, say, 3-km thickness of Mesozoic and Tertiary sediments, ranging up to 5 km in thickness. Superficially, the Gulf of Mexico deep water fold belt has dimensions comparable to the Paleozoic Melville Island Fold Belt of the Canadian Arctic (see also Mitra, 2002). However, the tectonic setting, the basic stratigraphy and therefore, the structural style are fundamentally differ- ent. The obvious conclusion is that superficial structural similarities, while intriguing, can only be sensibly evaluated in a more complete regional tectonic and stratigraphic context. For petroleum geologists, the understanding of the whole hydrocarbon system and, particularly, the presence and maturity of source beds is essential. While not directly related to passive margins, it is tempting to compare the above- listed gravitational systemswith similar systems that are superimposed on orogens often during the late stages of their deformation. A regional overview by Garcia- Molina (1994) describes the Reforma/Sonda de Campeche basin of SE. Mexico as follows: the basin is underlain by a Paleozoic basement that farther offshore merges with an oceanic basement. Evidence supporting Triassic rifting is limited. The basement is overlain by a salt-based fold belt. Gravitational salt tectonics pre- Alzaga-Ruiz, H., Granjeon, D., Lopez, M., Seranne, M., Roure, F., 2009. Gravitational collapse and Neogene transfer across the western Margin of the Gulf of Mexico: insights from numerical Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps models. Tectonophysics 470, 21–41. Bally, A.W. (Ed.), 1983. Seismic Expression of Structural Styles—A Picture and Work Atlas, “Tec- tonics of Extensional Provinces”, American Association PetroleumGeologists: Studies in Geol- ogy Series # 15. V-2. 329 pp. Briggs, S.E., Cartwright, J., Davies, R.J., 2009. Crustal structure of the deepwater West Niger Delta passivemargin fromthe interpretationof seismic reflectiondata.Mar. Petrol.Geol. 26, 936–950. Buffler, R.T., 1983. Structure of theMexican ridges fold belt, southwest Gulf ofMexico. In: Bally, A. W. (Ed.), Seismic Expression of Structural Styles: A Picture and Work Atlas, vol. 2. American Association of Petroleum Geologists, Studies in Geology #27, pp. 16–21. Burke, K.B., Gunnell, Y., 2008. The African Erosion Surface: A Continental-Scale Synthesis of Geo- morphology, Tectonics and Environmental Change over the Past 180 Million Years. Geologi- cal Society of America Memoir 201, 66 pp. Butler, R.W.H., Paton, K., 2010. Evaluating lateral compaction in deepwater fold and thrust belts: how much are we missing from Nature’s Sandbox? GSA Today 20, 4–10. Cramez, C., Jackson, M.P.A., 2000. Superposed deformation straddling the continent—oceanic transition in deep water Angola. Mar. Petrol. Geol. 17, 1095–1109. de Vera, J., Granado, P., McClay, K.R., 2009. Structural evolution of the Orange basin gravity driven system, offshore Namibia. Mar. Petrol. Geol. 27, 223–237. Diegel, F.A., Karlo, J.R., Schuster, D.C., Shoup, R.C., Tauvers, P.R., 1995. Cenozoic structural evo- lution and tectonostratigraphic framework 1995 of the Northern Gulf coast continental ceded much of the Tertiary folding and ranges from salt rafts to diapirs. Salt rafts overlain bymostly pelagic sediments deposits formed at themost distal bottom of the slope of the Cretaceous Campeche carbonate platform. In turn, the Tertiary fold and thrust belt is discordantly overlain at a considerable angle by a Neogene extensional growth-fault system, characterized by sole faults occasionally asso- ciated with salt pods derived from underlying diapirs. Moving farther north in the offshore and separated by a conjugate landward dipping listric normal fault and/or weld, there is yet another large allochthonous salt mass. Finally, seismic profiles across the Transylvanian basin of Romania illustrate a salt- based linked extensional-compressional system that glides from the rising East Carpathians toward the center of the basin (Kreszek and Bally, 2006). To conclude, gravitational extensional-compressional systems are best known from passive margins where many of them have been aggressively explored for hydrocarbons. The seismic documentation of these includes 2D and 3D surveys. Similarly, gravitational systems that are associated with a variety of late orogenic settings are also well documented. References to Passive margins on relatively stable/rigid lithosphere Ajakaye, D., Bally, A.W., 2002. Course Manual and Atlas of Structural Styles on Reflection Profiles from the Niger Delta. Continuing Education Course Note Series 41, American Association of Petroleum Geologists, p. 100 37 plates. 1119 margin. In: Jackson,M.P., Roberts, D.G., Snelson, S. (Eds.), Salt Tectonics AGlobal Perspective, American Association of Petroleum Geologists Memoir, 56, pp. 109–152. 1120 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Doust, H., Omatsola, E., 1990. Niger delta. In: Edwards, J., Santagrossi, P. (Eds.), Divergent Passive Margins. American Association of Petroleum Geologists Memoir 48, pp. 389–404. Duval, B., Cramez, C., Jackson,M.P.A., 1992. Raft tectonics in the Kwanza basin.Mar. Petrol. Geol. 9, 389–404. Fiduk, J.C., Weimer, P., Trudgill, B.D., Rowan, M.G., Gale, P.G., Gafford, W.T., et al., 1999. The perdido fault belt, northwestern deep Gulf of Mexico: part 2 stratigraphy and petroleum sys- tems. Am. Assoc. Petrol. Geol. Bull. 83, 578–610. Galloway, W.E., 2008. Depositional evolution of the Gulf of Mexico sedimentary basin. In: Miall, A.D. (Ed.), The Sedimentary Basins of the United States and Canada. Vol. 5. Sedimen- tary Basins of the WorldElsevier, Amsterdam, pp. 505–549. Garcia-Molina, G., 1994. Structural Evolution of SE Mexico (Chiapas-Tabasco-Campeche) Off- shore and Onshore. : PhD thesis Rice University, Houston, TX, 161 pp. Hudec, M.P., Jackson, M.P.A., 2004. Regional Restoration across the Kwanza Basin, Angola: salt tectonics triggered by repeated uplift of ametastable passivemargin. Am. Assoc. Petrol. Geol. Bull. 8, 971–990. (see also Bureau of Economic Geology: 18 Angola_Uplift.ppt. Hudec, M.R., Jackson, M.P.A., 2006. Advance of allochthonous salt sheets in passive margins and orogens. Am. Assoc. Petrol. Geol. Bull. 90, 1535–1564. Hudec, M.R., Jackson, M.P.A., 2007. Terra infirma: understanding of salt tectonics. Earth Sci. Rev. 82, 1–28. Jackson, M.P.A., Hudec, M.R., Hegarty, K.A., 2005. The Great West African coastal uplift: Fact or fiction? A perspective from the Angolan divergent margin. Tectonics 24, TC6014. doi:101029/2005TC001836. Kreszek, C.S., Bally, A.W., 2006. The Transylvanian Basin (Romania) and its relation to the Carpathian Fold and thrust belt; insights into gravitational salt tectonics. Mar. Petrol. Geol. 4, 405–442. Le Roy, C., Rangin, C., Le Pichon, X., Ngoc, N.T., Andreani, L., Aranda-Garcia, M., 2008. Neo- gene crustal shear zone along the western Gulf of Mexico and implications for gravity sliding processes. Evidence from 2D and 3D multichannel seismic data. Bull. Soc. Ge´ol. Fr. 179, 175–193. Maloney, D., Davis, R., Imber, J., Higgins, S., King, S., 2010. New insights into the deformation mechanism in the gravitationally driven Niger Delta deep-water fold and thrust belt. Am. Assoc. Petrol. Geol. Bull. 94, 1401–1424. Marton, I.G., Tari, G.C., Lehmann, C.T., 2000. Evolution of the Angolan passive margin, West Africa, with emphasis on post salt structure. In: Mohriak, W., Talwani, M. (Eds.), Atlantic Rifts and Continental Margins, American Geophysical Union Monograph, 115, pp.129–149. Mitra, S., 2002. Structural models of faulted detachment folds. Am. Assoc. Petrol. Geol. Bull. 86, 1663–1694. Moretti, I., Wu, S., Bally, A.W., 1990. Balanced cross section (LOCACE) to reconstruct allochtho- nous salt sheet, offshore Louisiana. Mar. Petrol. Geol. 7, 371–377. Morley, C.K., King, R., Hillis, R., Tingay, M., Backe, G., 2011. Deepwater fold and thrust belt clas- sification, tectonics, structure and hydrocarbon prospectivity. Earth Sci. Rev. 104, 41–91. Peel, F.J., Travis, C.J., Hossack, J.R., 1995. Genetic structural provinces and salt tectonics of the Cenozoic offshore Gulf of Mexico. A preliminary analysis. In: Jackson, M.P., Roberts, D.G., Snelson, S. (Eds.), Salt Tectonics: A Global Perspective, American Association of Petroleum Geologists Memoir 56, pp. 153–152. Perez-Cruz, 1993. Geologic evolution of the Burgos Basin, northeasternMexico. Ph.D. Thesis Rice University, 375 pp. http://hdl.handle.net/1911/16657. Reijers, T.J.A., Petters, S.W., Nwadjide, C.S., 2007. Chapter 7: The Niger Delta basin. The Niger Delta basin. In: Selley, R.C. (Ed.), African Basin. In: Hsu¨, K.J. (Ed.), Sedimentary Basins of the World, vol. 3. Elsevier, pp. 151–172. Rodriguez, A.B., Mann, P., 2011. Origin of the Mexican Ridges Passive Margin Fold Belt Based on Seismic andWell Integration of the Shelf Slope Basin and Structural Restoration. AAPG Search Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps and Discovery Article #90124# 2011. Rowan, M.G., 1995. Structural styles and evolution of allochthonous salt, Central Louisiana outer shelf and outer slope. In: Jackson, M.P.A., Roberts, D.G., Snelson, S. (Eds.), Salt Tectonics a Global Perspective, American Association of Petroleum Geologists Memoir 56, pp. 199–222. Rowan, M.G., Peel, F.J., Vendeville, B., 2004. Gravity driven fold belts on passive margins. In: McClay, K.R. (Ed.), Thrust Tectonics and Hydrocarbon Systems, American Association of Petroleum Geologists Memoir 82, pp. 157–182. Salvador, A. (Ed.), 1991. The Gulf of Mexico Basin. Geological Society of America. The Geology of North America (DNAG).V.J., 568 pp. 6 Plates. Tari, G., Molnar, J., Ashton, P., 2003. Examples of salt tectonics from West Africa: a comparative approach. In: Arthur, T.J.M., Cameron, N.D. (Eds.), Petroleum Geology of Africa, Geological Society of London Special Publication, 207, pp. 85–104. Trudgill, B.D., Rowan,M.G., Fiduk, J.C.,Weimer, P., Gale, P.E., Korn, B.E., et al., 1999. The Perdido Fold belt, Northwestern Deep Gulf of Mexico: part 1 structural geometry, evolution and regional implications. Am. Assoc. Petrol. Geol. Bull. 83, 88–113. Worrall, D.M., Snelson, S., 1989. Evolution of the northern Gulf of Mexico, with emphasis on Cenozoic growth faulting and the role of salt. In: Bally, A.W., Palmer, A.R. (Eds.), The Geology of North America: An Overview. Geological Society of North America. Decade of North Amer- ican Geology, V.A, pp. 97–138. Wu, S., Bally, A.W., Cramez, C., 1990. Allochthonous salt, structure and stratigraphy of the north- eastern Gulf of Mexico: Part II structure. Mar. Petrol. Geol. 7, 334–370. Cratonic basins on relatively stable/ rigid lithosphere Plate 25.35: Cratonic and Foreland Basins Cratonic and foreland basins are both located on continental cratons. When adja- cent to each other, the separating boundary is arbitrary and often on amonocline, for example, Western Canada or the Arabian peninsula, or else on an arch that separates varying cratonic subsidence domains. Following Kober (1928), Stille (1936a and b) introduced the term “Kraton” for a stable consolidated older platform. According to Dennis (1976) Kay (1951) angli- cized the term to “craton” and recognized bounding flexures that define the inte- rior of a craton. However, the boundaries of a craton remain difficult to define (Leighton and Kolata, 1990). Sloss (1988) defined a craton as an extensive region of thick continental crust that lies at a constant position relative to sea level of many tens of millions of years. Long periods of stability characterize cratons though there may be episodic deformation of varying degrees (e.g., Nikishin et al., 1996, 2010). The term craton includes the shields of basement rocks exposed over large areas and the platforms, that is, regions of flat-lying or gently deformed sediments underlain either by basement rocks or rocks deformed during earlier deformation phases. Accreted terranes that now behave as cratons are thus 1121 included. Following sub-chapter 25.8 and 3.8.2 and the basement Map (Plate 1122 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps 25.20), it is useful to differentiate Precambrian from Paleozoic cratonic basements. Cenozoic/Mesozoic basements here are not classified as cratonic (e.g., Franciscan/ Sierran basement of California, and the backarc basement of Indonesia). Note that in the older literature of Indonesia, there are references to a “Sunda craton”—an area which here is not classified as cratonic. Instead, its basins are grouped as “epi- sutural” backarc basins on a Mesozoic continental basement/crust. To sum up, cratonic basins, by definition, form on Precambrian or Paleozoic basement that in places may have been rifted. The setting of cratonic basins varies from “pericratonic” foreland basins that toward the adjacent shieldmay expose the earlier cratonic basin platform to basins within cratonic interior that are separated from the foreland basin by a cratonic arch. Note that rifts or aulacogens may be viewed by some as a part of a cratonic basin. Leighton (1990) notes two families of sedimentary basins in cratonic settings: (1) basins located on rigid lithosphere and not associated with megasutures, that is, our cratonic basins. Note that on Plates 25.31 and 25.35 these basins all are marked by dark blue centers when cratonic basin fill exceeds 2 km. (2) perisutural basins located on rigid lithosphere associated with and flanking cratonic basins, that is, our foreland basins (light green on Plates 25.31 and 25.35) Leighton also notes that cratonic basins are defined as “simple” and are located on pre-Mesozoic lithosphere. They can be further differentiated either by the position of earlier rifts or on former basins of another type, for example, West Siberia (see Vyssotski et al. Chapter 21, Vol. 1C). Bally (1989) notes that the deceptively simple craton “saucer” and circular shapes of many cratonic basins remain unexplained. This is particularly vexing when realizing that many arches adjacent to cratonic basins separate distinctively different yet coeval subsidence regimes on their flanks. An initial extensional origin for cratonic basins is often taken to be implicit but is not always clear. It is therefore important to evaluate whether a rift underlying a sedimentary basin can be properly related to the subsidence of that basin. Com- bining Plate 25.29 (Rifts) and Plate 25.35 (Basins) in one plate results in a “messy” map (not included here) that shows the great majority of all sedimentary basins to be somewhere underlain by or else associated with a later rift system. It remains important to ask in each case how relevant underlying rifts are for the ensuing polyphase evolution of a basin. In some cases, for example, the Michigan basin, the underlying Precambrian rift is too old to explain the subsidence of the sau- cer-shaped Paleozoic basin (Burgess, 2008). Insight into the shape of cratonic basins may come from the recognition that subsi- dence of plate interiors can occur as a result of lithospheric flexure accompanying the loading of adjacent foreland basins, which may account for the regional arches and subcrop patterns around the margins of cratonic basins such as the Williston and Illinois basins (Buschbach and Colata, 1990). For continentwide subcrop maps Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps of cratons, see also Cook and Bally (1975) and Bally (1989). The importance of dynamic topography is reviewed by Burgess (2008). Nikishin et al. (1996) have invoked the transmittal of horizontal in plate stress to account for episodic deforma- tion, for example, in the East European craton. In the case of East Siberia cratonic basin (Nikishin et al., 2010), the history seems to have been much more complex than the pattern of subsidence and peripheral uplift observed in Central North America. Nikishin et al. (2010) describe a complex history with fold belt accretion on Proterozoic basement in Riphean-Vendian time, and the associatedVendian fore- land basin (Gladkochub et al., 2009) followed by earliest-Paleozoic rifting (?) and passive margin (?) formation, Mid-Paleozoic compression, and Mid–Late Devonian rifting. This phase was followed by fold belt deformation along the west margin during the Hercynian, on the northern margin during the late Hercynian/Triassic (Taymir), and to the east during the Cretaceous (Verkhoyansk). Added to this complex evolution was the emplacement of one LIP near the SE margin of Siberian basement around 1750 Ma, that is, prior to the consolidation of the east Siberian craton (Gladkochub et al., 2009) and more importantly another LIP around 350 Ma near Yakutsk near the SE margin of the basin associated with the rifting phase, and finally, around 250 Ma, by the giant LIP associated with the widely distributed Siberian traps and its exposed dike swarm center in the vicinity of Norilsk in the northwest corner of the E. Siberia basin (see Plate 25.25). The area occupied by all the North American Cratonic/Foreland basins is compa- rable in magnitude to that of the East Siberian craton. Seen in this perspective, the following features reveal that the North American craton also underwent a com- plex tectonic evolution including the following: – Meso-Proterozoic Rift systems in the West, (Uinta Mountains, Belt Purcell Rift systems.) In the Great Lakes area, (Mid-Continent Rift system) – Neo-Proterozoic rifts in the south and southeast (Proto-Reelfoot/Rough creek system) – Cambrian Rift systems (Rome trough Kentucky; Oklahoma aulacogen; Richardson Mountain aulacogen (N. Yukon) – Lower Paleozoic Cratonic cover merging into passive margins surrounding most of N. America – Mid-Ordovician to Mid-Devonian emplacement of Appalachian/Ouachita Foreland basin – Northerly trending Siluro-Devonian uplifts of the Boothia arch in Northern Canada, and in the center of the Hudson Bay – Pre- and Middle Devonian Transcontinental and Peace River Arches – Upper Paleozoic foreland basins adjacent to the margins of the N. American craton—Permo-Pennsylvanian Wichita/Ancestral Rocky Mountain Uplifts and associated basins 1123 – Northerly striking uplifts, perhaps related to earlier Mid-Proterozoic fault sys- tems, formed the Upper Paleozoic of the Nemaha ridge system of Oklahoma 1124 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps and Kansas. They are bounded by faults that still are active today earthquakes – Mesozoic Foreland basins associated with the Western Cordillera and asso- ciated arches that separate them from the Interior Cratonic basins of the continent – Paleogene Laramide Uplifts and associated basins – Post-Paleogene Uplift of the Foreland of the Western Cordillera Cook and Bally (1975) compiled a systematic stratigraphic atlas of North America. Blakey (2011) presents an easily accessible, selection of 40 superb Paleo-Geo/ Physiographic reconstructions of North America. Miall (2008) contains a number of chapters that deal with “cratonic North America”. An updated overview of the Phanerozoic evolution of the north American Craton by Burgess (2008) suggests that variations in dynamic topography due to subducting lithospheric slabs may help explain the variety among cratonic basins of North America. A further complexity that relates to the evolution of the North American craton is that particularly along theWestern Cordillera of North America, theMid- to Late- Proterozoic and Paleozoic cover of the “craton” was involved and deformed in the adjacentMesozoic/Cenozoic FTBs of the U S and Canadian RockyMountains. This deformation involves thick Proterozoic sediments of the Lewis overthrust that straddles the US/Canada boundary in the Selkirk Mountains to the west. Crustal profiles show the westward continuation of the basement monocline. Thus, no allochthonous Proterozoic sediments were decoupled from the under- lying basement. Furthermore, Sears (2007a,b) presents a carefully reconstructed interpretation of the allochthonous Meso-Proterozoic US and Canadian Rocky Mountains as a “complex intracontinental rift system” that may have been connected to Siberia prior to the supposed Late Proterozoic separations of North America from Siberia. The suggestion is that the western margin of North Amer- ican craton and its underlying pre-Proterozoic basement were seriously “de- cratonized” by several rifting events and the following emplacement of a Paleozoic passive margin prior to formation long before its cover was stripped off to form a Cenozoic–Mesozoic FFTB. In a different example showing the importance of plate tectonics reconstructions of presumably stable cratons, Tankard et al. (1995/1998) restore South America and South Africa as part of Southwestern Gondwana, prior to the separation of the two continents. Veevers (2004, 2006) also shows the breakup of East Gond- wana, that is, the mostly Precambrian part of the Antarctic craton from the Pre- cambrian cratons of Australian and India. The examples of Eastern Siberia, North America, and South America, and the examples of the breakup of the E. Gondwana craton emphasize the problems associated with isolating unique stable elements, for example, shields or well-circumscribed basins to define a craton. Occasionally, one might or should ask whether the renowned relative “stability” of cratons is merely in the eyes of Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps a beholder who carefully isolates surviving cratonic fragments that best support a preferred model. Or else “Is the survival of a craton merely fortuitous and due to changing Paleotectonic stress regimes?” Be that as it may, the relative stability of cratons with respect to orogens remains very real, but so are impor- tant “nuances” among cratonic and progressively de-cratonized areas that range from continent-wide dimensions to seriously reduced mini-cratons such as the North China-Korea, the Yangtse, and Tarim “cratons” of China. References to Cratonic Basins Note:The following references list some publications that are not specifically cited in text, but were deemed to be important general references that pertain to Cratonic Basins. An asterisk, marks selected Stratigraphic Atlas publications that illustrate large cratonic areas or parts of them in con- siderable detail. These are not specifically cited in this abbreviated overview. *Baldschuhn, R., Bindt, P., Plug, S., Kokell, T., 2001. Geotektonischer Atlas von Nordwestdeutschland und dem deutschen NW Sector. Geol. Jahrbuch Reihe A 153, 3–93. Bally, A.W., 1989. Phanerozoic basins of North America. In: Bally, A.W., Palmer, A.R. (Eds.), The Geology of North America—An Overview. Geological Society of America, VA, pp. 397–446. Blakey, R., 2011. Paleogeography and Geologic Evolution of North America. 40 images of Phan- erozoic reconstructions of N, America with text and references http://www2.nau.edu/rcb7/ nam.html retrieved in 2011. Burgess, P.M., 2008. Phanerozoic evolution of the sedimentary cover of the North American Cra- ton. In: Miall, A.D. (Ed.), The Sedimentary Basins of the United States and Canada, vol. 5. Else- vier, Sedimentary Basins of the World, pp. 31–61. Buschbach, T. C., Kolata, D.R. (Eds.), 1990. Regional setting of the Illinois Basin. In: Leighton, M.W., Kolata, D.R., Oltz, D.F., Eidel, J.J., (Eds.), 1990. Interior Cratonic Basins. AAPG Memoir 51, 29–55. *Cook, T.D., Bally, A.W. (Eds.), 1975. Stratigraphic Atlas of North and Central America. Princeton University Press, p. 272. *Debrand-Passard, S., Courbouleix, S. (Eds.), 1984a. Synthe`se Ge´ologique du Sud-Est de la France, Stratigraphie et Pale´ogeographie. Me´moires du Bureau de Recherches Ge´ologiques et Minie`res N.126, vol. 1. 615 pp. *Debrand-Passard, S., Courbouleix, S. (Eds.), 1984b. Synthe`se Ge´ologique du Sud-Est de la France, Atlas. Me´moires du Bureau de Recherches Ge´ologiques et minie`res N., vol. 2. 126 57 Maps. *Daukeev, S.Z.h. et al., (Ed.), 2002. Atlas of the Lithology-Paleogeographical, Structural, Palinspastic and Geo-environmental Maps of Central Asia. YUGEO, Scientific Research Institute of Natural Resources, Almaty, Kazakhstan. 37 maps and text. Dennis, J.C., 1997. International tectonic dictionary. English terminology. Amer. Assoc. Petr. Geol. Memoir 7, 196 pp. Gladkochub, D.P., Pisarevsky, E.R., Donskaya, T.V., So¨derlund, U., Mazukabzov, A.M., Hames, J., 2009. Large igneous province of about 1750 Ma in the Siberian Craton. Doklady Earth Sci. 430 (2), 168–171. doi:10.1134/S1028334X100020042. Kay, M., 1951. North American Geosynclines. Geological Society of America. Memoir 48, 143 pp. Leighton, M.W., Kolata, D.R., Oltz, D.T., Eidel, J.J. (Eds.), 1951. Interior Cratonic Basins. AAPGMemoir 51, p. 819. Leighton, M.W., Kolata, D.C., 1990. Selected Interior Basins and their place in the scheme of global tectonics: a synthesis. In: Leighton, M.W., Kolata, D.R., Oltz, D.T., Eidel, J.J. (Eds.), Inte- rior Cratonic Basins. American Association of Petroleum Geologists Memoir 51, pp. 729–797. 1125 Leighton, M.W., 1990. Introduction to Cratonic Basins. In: Leighton, M.W., Kolata, D.R., Oltz, D. 1126 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps T., Eidel, J.J. (Eds.), Interior Cratonic Basins, American Association of Petroleum Geologist Memoir 51, pp. 1–24. Miall, A.D., 2008. The Paleozoic Western CratonMargin. In: Miall, A.D. (Ed.), The Sedimentary Basins of the United States and Canada, vol. 5. Sedimentary Basins of the World, Elsevier, pp. 181–210 (Chapter 5). Miall, A.D., Blakey, R.C., 2008. The phanerozoic tectonic and sedimentary evolution of North America. In: Miall, A.D. (Ed.), The Sedimentary Basins of the United States and Canada, vol. 5. Sedimentary Basins of the World, Elsevier, pp. 31–61 (Chapter 2.). McCann, T., Mader, H.M., Coles, S.G. (Eds.), 2008. The Geology of Central Europe, vol.1. Precambrian and Paleozoic. Geological Society London, 800 pp. McCann, T. (Ed.), 2008. The Geology of Central Europe, vol 2. Mesozoic and Cenozoic. Geological Society London, 752 pp. *Me´gnien, C. (Ed.), 1980a. Synthe`se du Bassin de Paris, vol. 1. Stratigraphie et Pale´ogeographie, Me´moires du Bureau de Recherches Ge´ologiques et Minie`res N0.103, 466 pp. *Me´gnien, C. (Ed.), 1980b. Synthe`se du Bassin de Paris, vol. 2. Me´moires du Bureau de recherches ge´ologique et Minie`res No.103. Atlas. 55 Maps. *Me´gnien, C. (Ed.), 1980c. Synthe`se du Bassin de Paris. Me´moires du Bureau de recherches ge´ologi- ques et Minie`res No.103., vol. 3. Lexique des Noms de Formations, p. 477. *MillenniumAtlas Co. Ltd., BritishGeological Survey ProjectManagers—ExplorationGeosciences-Tech- nical Editors, 2002. Millennium Atlas. Petroleum Geology of the Central and Northern North Sea. Geological Society London. Miscellaneous Titles. 390 pp. Now available as CD. Nikishin, A.M., Sobornov, K., Prokopiev, A.V., Frolov, S.V., 2010. Tectonic evolution of the Siberian Platform during the Vendian and Phanerozoic. Moscow Univ. Geol. Bull. 5 (1), 1–16. Nikishin, A.M., Ziegler, P.A., Stephenson, R.A., Cloetingh, S.A.P.L., et al., 1996. Late Pre-Cambrian to Triassic history of the East European craton: dynamics of sedimentary basins. Tectonophy- sics 268, 23–63. Miall, A.D. (Ed.), 2008. The Sedimentary Basins of the United States and Canada. In: Hsu¨, K.J. (Ed.), Sedimentary Basins of the World, Elsevier, p. 610. Mossop, G.D., Shetsen, I. (Eds.), 1994. Geological Atlas of the Western Canada Sedimentary Basin. Canadian Society of Petroleum Geologists and Alberta Research Council, Special Report 4, URL http://www.ags.gov.ab.ca/publications/wcsb_atlas/atlas.html, March 15, 2011. Poulton, T.P., Christopher, J.E., Hayes, B.J.R., Losert, J., Tittemore, J., Gilchrist, R.D., 1949. Jurassic and lowermost Cretaceous strata of the Western Canada Sedimentary Basin. In: Mossop, G.D., Shetsen, I. (Eds.), Geological atlas of Western Canada Sedimentary Basin. Canadian Society of Petroleum Geologists and Alberta Research Council, Special Report 4, pp. 297–316. Sears, J.W., 2007a. Rift destabilization of a Proterozoic epicontinental pediment: a model for the Belt-Purcell Basin. In: Link, P.K., Lewis, R.S. (Eds.), Proterozoic Geology of Western North America and Siberia. SEPM Special Publication, 86, pp. 55–68. Sears, J.W., 2007b. Belt-Purcell Basin; Keystone of the Rocky Mountain fold and thrust belt. In: Sears, J.W., Harms, T.A., Evenchick, C.A. (Eds.),Whence theMountains? Inquiries into the Evo- lution of Orogenic Systems. A volume in honor of R.A. Price. Geological Society of America. Special Papers 433, pp. 147–166. Selley, R.C. (Ed.), 1997. African Sedimentary Basins. In: Hsu¨, K.J. (Ed.), Basins of the World Series, Else- vier, p. 394. Sloss, L.L., 1988. Tectonic evolution of the craton in Phanerozoic time. In: Sloss, L.L. (Ed.), Sedi- mentary Cover-North American Craton, U.S. The Geology of North America, vol. D. Boulder, Geological Society of America. pp. 25–52. Sloss, L.L., 1988a. Sloss, L.L. (Ed.), Sedimentary Cover—North American Craton. The Geology of North America, v. D-2, Geological Society of America, p. 506. Sloss, L.L., 1988b. Tectonic evolution of the craton in Phanerozoic times. In: Sloss, L.L. (Ed.), Sedimen- tary cover—North American craton. The Geology of North America, v. D-2, Geological Society of America, pp. 25–51. Sovetov, J.K., 2002. Vendian foreland basin of the Siberian cratonic margin: Paleopangean accretion- atlas.html. March 15, 2011. *Ziegler, P.A., 1990. Geological Atlas of Western and Central Europe. second ed. Shell Int. Petrol. Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Mij. B. BV., distributed by Geol. Pub. House, Bath, 239 pp, 56. Ziegler, P.A., Cloetingh, S., van Wees, J.D., 1995. Dynamics of intra-plate compressional deforma- tion: the alpine foreland and other examples. Tectonophysics 252, 7–59. doi:10.1016/0040- 1951(95)00102-6. Ziegler, P.A., van Wees, J.D., Cloetingh, S., 1998. Mechanical controls on collision-related compres- sional intraplate deformation. Tectonophysics 300, 103–129. doi:10.1016/S0040-1951(98), 00236-4. Basins on the periphery of orogens: Deep sea trenches and foreland basins Plate 25. 35: Cratonic and Foreland Basins Plate 25.36: Deep Sea Trenches Foreland basins, Foreland Fold and Thrust Belts (FFTBs), and Episutural basins Deep sea trenches and foreland basins are homologous. The former are the loci of subducting oceanic slabs and associated active margin basins. Foreland basins are associated with FFTBs that are “under-thrusted” by cratonic continental forelands and their underlying cratonic basement/crust/lithosphere. While continental sub- duction is now accepted in principle, the area/volume and the process of varying types of continental lithosphere subduction remain uncertain. Details of typical tectono-stratigraphic sequences in foreland settings and additional comments are given in Chapter 4, Vol. 1A (Roberts and Bally). The Western Canada Atlas (Mossop and Shetsen, 1994; Wright et al., 1994) and the Rocky Mountain ary phases. Russ. J, Earth Sci. 4. Stille, H., 1936a. Tektonische Beziehungen zwichen North Amerika und Europa; 16th Interna- tional Geological Congress Report, Washington 1933, pp. 829–838 (fide; Dennis 1979). Stille, H., 1936b. Wege und Ergebnisse der geologisch-tektonischen Forschung. 25 Jahre. Kaiser Wilhelm Gesellschaft 2. 84–85 (fide Dennis 1979). Stille, H., 1941. Einfu¨hrung in den Bau Amerikas. Borntraeger, Berlin, 717 pp. Tankard, A.J., et al., 1995/1998. Structural and tectonic control of basin evolution in south-west- ern Gondwana during the Phanerozoic. In: Tankard, A.J., Suarez-Soruco, R., Welsink, A.J., Welsink, A.J. (Eds.), Petroleum Basins of South America, AAPG Memoir 62, pp. 5–52. *Veevers, J.J., 2001. Atlas of Billion-Year Earth History of Australia and Neighbours in Gondwanaland. GEMOC Press, North Ryde, 80 pp. Veevers, J.J., 2004. Gondwanaland from 600–570 Ma assembly through 320 Ma merger in Pan- gea to 180–100Ma breakup: supercontinental tectonics via stratigraphy and radiometric dat- ing. Earth Sci. Rev. 68, 1–132. Veevers, J.J., 2006. Updated Gondwana (Permian–Cretaceous) earth history of Australia. Gond- wana Res. 9, 231–260. Wright, G.N., McMechan, M.E., Potter, D.E.G., Holter, M.E., 1994. Structure and architecture of the Western Canada Sedimentary Basin. In: Geological Atlas of the Western Canada Sedimentary Basin. G.D. Mossop and I. Shetsen (comp.), Canadian Society of PetroleumGeologists and Alberta Research Council, Special Report 4, URL, http://www.ags.gov.ab.ca/publications/wcsb_atlas/ 1127 Geological Atlas (Rocky Mountain Association of Geologists, 1972) are examples 1128 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps of comprehensive documentations of two foreland basin provinces. Note that the old term “Foredeep” has become restricted to basins on continental crust, that is, today’s foreland basins. While this may be appropriate, it should not be forgotten that remnant ocean basins such as the Timor Trough above may rep- resent incipient foreland basins. It is also useful to think of the term “foredeep” as referring to the ephemeral underfilled part of a foreland basin typically associated with “flysch” (an old Swiss word) deposition. Foreland basins can be conveniently divided into “pre-foreland/platformmegase- quences” and one or more “foreland basin megasequences” that are separated from the former by a basal foreland basin unconformity. The down dip end of this unconformity merges into a correlative conformity marking the inception of the main collision with the foreland. The erosion of the down dip termination of that unconformity, whichmay eventuallymerge to a conformity,may bemostly due to erosion by deep oceanic turbidity and/or contour currents. Thus, only a minor, if any, hiatus characterizes the downdip submarine part of the basal foreland basin unconformity. However, on the updip part of the basal foredeep unconformity, shallow water and/or subaerial erosion truncates underlying platform strata, which are overlapped by the overlying foreland strata. The fill of most foreland basins often reveals pronounced longitudinal drainage as shown by oblique progradation and axial turbidite deposition. Foreland basin megasequences have an overall geometry of thickening into the adjoining fold belt (see Chapter 6, Vol. 1A (Roeder)), which is attributed to contemporane- ous deposition in accommodation space created by the flexural loading imposed by the fold belt. Foreland basins are typically associated with an updip migrating low relief flexural bulge whose trace is the basal foreland basin megasequence unconformity. It is also important to note that foreland basins are sometimes further subdivided (e.g., Catuneanu, 2004; Garzanti, 2007; Jordan, 1995) as follows: (1) Retro-foreland basin: Regional monocline that dips in the opposite sense to the associated oceanic subduction, for example, Western Canada basin, Venezuela, and Andean foreland basins. (2) Proforeland basin: Regional monocline that dips in the same sense as the associated subduction, for example, the Alpine Molasse Basin, the Po Plain/Adriatic foreland basins, the Arabian Gulf-Mesopotamian foreland basins, and the Pakistan, India foreland basins. The floor of preforeland/platform basin megasequences is defined by the top basement unconformity overlying a dominantly crystalline basement, that is, the top of a Precambrian or Paleozoic lithosphere. All foreland basins are by definition located on continental/cratonic crust in contrast to remnant ocean basin and deep sea trenches, which are located on oceanic crust. Composite pre-foreland basin megasequences reveal a complex polyphase rift to passive Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps margin evolution already, which precedes the onset of the foreland basin formation. Combined rift/flexural basins located on Mesozoic/Cenozoic basements are here classified as parts of the “episutural” forearc or backarc basins associated with oceanic subduction that in some cases (e.g., Taiwan) may be overlain by flexural foreland basin megasequences. Most commonly, when preserved, preforeland basin megasequences can be defined and divided into pre- and syn- and post-rift megasequences. If not, the use of the term cratonic platform megasequence may be more appropriate to describe the preforeland basin megasequence. In the case of many prefore- land basin megasequences, one only sees the thin landward or proximal post- rift (or steers head) part of a passive margin, which is typically more rigid than the outboard transitional passive margin. In this case, the distinction between proximal postrift passive margin and cratonic platform obviously becomes tenuous. However, a formal definition would state that in the case of foreland basins one often deals only with the thin landward portion of a passive margin, which is commonly flexurally more rigid than the outboard transitional passive margin. All this fits with observations of foreland basin development. The foreland basin preserved today is the end point of a continuum, which involves early collision, for example, Timor and the ultimate complete collision and overthrusting of the thin end of the passive margin steer’s head where one is unlikely to find evidence of major extension. Anymajor rift-related extensionwill lie in the internal zones unless dislocated elsewhere by strike-slip and inversions associated with complex strain partitioning. In common with rifts and passive margins, it is useful to further subdivide foreland basins according to nature of the sedimentary infill: (1) Predominantly marine coastal clastic foreland basin fill, for example, Western Canada, Alaska North Slope, Alpine/Carpathian, and Andean basins. (2) Lacustrinemostly clastic foreland basin fill, for example, . . . Paleogenemega- sequences of Laramide Uinta Green River basins, Jurassic-Cretaceous infill of the Sichuan, the Permian–Mesozoic, and Tertiary infill of the Junggar and Tarim basins of China. (3) Predominantly carbonate foreland basin fill, for example, Fort Worth, Delaware (U.S.), Gulf of Arabia, and Mesopotamian basins. (4) Mixed clastic carbonate/clastic foreland basin fill, for example, onshore/ offshore Papua-New Guinea basin. Simple foreland basins contrast with tectonically composite foreland basins that are segmented by compressional/transpressional basement uplifts (Bally and 1129 Snelson, 1980) (see also Chapter 4, Vol. 1A (Roberts and Bally).). These composite 1130 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps basins may be further differentiated as follows: (1) U.S. Rocky Mountain-Type Foreland basins segmented by compressional or transpressional basement uplifts, for example, the Wind River, Green River, Bighorn, and Uinta basins of the U.S Rocky Mountains (Lawton, 2008; Rocky Mountain Association of Geologists, 1972). These basins are superim- posed on an earlier, simpler, and much wider simple foreland basin that is segmented by later compressional/transpressional basement-involved uplifts or occasionally by inverted half-grabens. Some of them may be recognized on Plates 19, 22, and 23 from their association with the basement uplift/ inversion symbol. Note that on Plate 25.31 the Maracaibo basin of Venezuela is shown as an episu- tural basin. However, with equal or even better justification, it could also be listed as a “Rocky Mountain-type” foreland basin because it is surrounded by Paleozoic basement-involved uplifts of Tertiary age. Beginning as a Cretaceous foreland basin in the west, its complex polyphase evolution involves, during the Paleogene and Neogene massive infill from the rising Merida Andes to the south and also the Sierra de Perija mountains to the West (Castillo andMann, 2006; Lugo andMann, 1995/1998; Mann and Escalona, 2006; Mascle et al., 1998; Miall, 2008; Parnaud et al., 1995). (2) Central Asia-Type Foreland basins segmented by compressional/transpres- sional basement uplifts (Chinese basins of Bally and Snelson, 1980) These basins are associated with long lasting orogenic systems associated with sequential accretion of multiple terranes. The Tarim basin of Western China is located on Precambrian crust but has been directly affected by pre-Silurian foreland fold and thrust faulting of lower Paleozoic sediments and or basement-involved deformation. Extensive left-lateral SW/NE trending strike-slip fault systems affect the southeast part of the Tarim basin. A large, mostly buried, Tertiary basement involved uplift dominates parts of the Western Tarim basin. The Tarim basin is particularly well-documented by both 2D and 3D seismic profiles and numerous wells (Chengzao et al., 2003; Liangshu et al., 2003). Other basins of Central Asia are underlain by Paleozoic basement that outcrops in the adjacent mountain ranges, for example, the Junggar (Lawrence, 1990) and Turpan basins of Western China. An issue of particular interest is the foreland basin record of accretionary episodes in the adjacent orogen (e.g., Ettenson, 2008; Jordan et al., 1988; Ricketts, 2008; Stockmal et al., 1992). Typically, the inception of accretion to an orogen is marked by a basal foreland basin unconformity to be followed by later unconformities that mark the inception of additional accretions. The sequences between these uncon- formities are typical for clastic wedges that are derived during the accretion to the adjacent orogen. However, in a number of cases, amore limited amount of clastics As already indicated in Chapter 4, Vol. 1A (Roberts and Bally), the basal foreland Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps basin unconformity commonly separates an underlying platform sequence from the overlying foreland basin megasequences s.s. Most commonly these are domi- nated by clastics but some, economically important, foreland basinmegasequences are dominated by carbonates. Prominent examples include the basins of Alberta, West Texas (e.g. Dorobeck and Ross,1992) and the Middle East (e.g. Beydoun 1991, Sharland et al., 2001, Ziegler, 2001 and Bordenave and Hegre, 2010). Foreland basin models have been proposed and discussed by Beaumont (1981), Beaumont et al. (1987), DeCelles and Giles (1996), Jordan (1981, 1995), and Jordan et al. (1988). References to Basins on the Periphery of Orogens: Deep Sea Trenches and Foreland Basins Note: In italics are some selected reference that are not specifically cited in the text but were used in the compilations of the maps discussed in the above segment. Italics with an asterisk *mark selected Stratigraphic Atlas Publications that illustrate large foreland areas or parts of them in considerable detail. Allen, P.A., Homewood, P. (Eds.), 1986. Foreland Basins. Special Publications International Asso- ciation of Sedimentologists. Special Publication 8. Blackwell Scientific Publications, p. 453. Bally, A.W., Snelson, S., 1980. Realms of subsidence. In: Miall, A.D. (Ed.), Facts and Principles of World PetroleumOccurrence, Canadian Society of PetroleumGeologistsMemoir 6. pp. 9–94. Beaumont, C., 1981. Foreland basins. Geophys. J. R. Astr. Soc. 65, 291–329. Beaumont, C., Quinlan, G.M., Hamilton, J., 1987. The AlleghenianOrogeny and its relation to the evolution of the Eastern Interior, North America. In: Beaumont, C., Tankard, A.J. (Eds.), Sedi- mentary Basin and Basin-Forming Mechanisms, Canadian Society of Petroleum Geologists Memoir 12, pp. 425–445. may also be derived from adjacent distal cratonic shields and their cover. Orogen- derived sands tend to be lithic, while craton/shield-derived sands are cleaner and more quartzose. Significant amounts of the world’s petroleum resources are hosted in foreland basins of Paleozoic to Tertiary age and their adjacent FFTBs. The prolific petroleum systems in these basins and the adjacent FFTBs typically reflect a polyphase history, which has resulted in deposition of multiple source beds, reservoirs, and seals in both the preforeland/platform basin megasequence and the successor foreland basin megasequence. Books and papers that focus on foreland basins with more detailed descriptions include Allen and Homewood (1986), Katz (1990), Dorobek and Ross (1994), Lacombe et al. (2007), Macqueen and Leckie (1992), Mann (1999), Mascle et al. (1998), Miall (2008), Mossop and Shetsen (1994), Tankard et al. (1995), and Van Wagoner and Bertram (1995). 1131 Beydoun, Z.R., 1991. Arabian plate hydrocarbon geology and potential—a plate tectonic approach. Am. Assoc. Petr. Geol. Stud. Geol., 33, 77pp. 1132 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Bordenave, M.L., Hegre, J.A., 2010. A current distribution of oil and gas fields in the Zagros fold belt of Iran and contiguous offshore as the result of petroleum system. In: Leturmy, P., Robin, C. (Eds.), Tectonic and stratigraphic evolution of the Zagros and Makran during the Mesozoic Cenozoic. Geol. Soc. Spec. Publ. 330, 291–353. Castillo, M., Mann, P., 2006. Cretaceous to Holocene structural and stratigraphic development of south Lake Maracaibo, Venezuela, inferred from well and 3D seismic data. Am. Assoc. Petrol. Geol. Bull. 90, 529–565. Catuneanu, O., 2004. Retroarc foreland systems—evolution through time. J. Afr. Earth Sci. 38, 225–242. Chengzao, J., Guoqi, W., et al., 2003. Mesozoic and Cenozoic evolution of the Tarim basin. In: Tarim Basin—A Twelve Volume Multi-author Synthesis, vol. 2. Petroleum Industry Press, Beijing, p. 229. Dorobeck, S.L., Ross, G.M., (Eds.), 1995. Stratigraphic evolution of Foreland Basins. SEPM Spec. Publ., 52, 310. Ettensohn, F.R., 2008. Appalachian foreland basin in Eastern United States. In: Miall, A.D. (Ed.), The Sedimentary Basins of the United States and Canada. In: Hsu¨, K.J. (Ed.), Sedimentary Basins of the World, vol. 5. Elsevier, Amsterdam, pp. 105–179. DeCelles, P.G., Giles, K.A., 1996. Foreland basin systems. Basin Res. 8, 105–123. Jordan, T.E., 1981. Thrust loads and foreland basin evolution. Cretaceous, western United States. Am. Assoc. Petrol. Geol. Bull. 65, 2506–2520. Jordan, T.E., 1995. Retroarc Foreland and related basins. In: Busby, C.J., Ingersoll, R.V. (Eds.), Tec- tonics of Sedimentary Basins. Blackwell Science, Oxford, pp. 331–362. Jordan, T.E., Fleming, P.B., Beer, J.A., 1988. Dating thrust-fault activity by use of foreland-basin strata. In: Kleinspehn, K.L., Paola, C. (Eds.), New Perspectives in Basin Analysis. Springer- Verlag, New York-Berlin-Heidelberg, pp. 307–330. Garzanti, G., 2007. Orogenic belts and orogenic sediment provenance. J. Geol. 115 (3), 315–334. Katz, B.J., 1990. Lacustrine basin exploration: case studies andmodern analogs. Am. Assoc. Petrol. Geol. Bull. 50, 340. Lacombe, O., Lave´, J., Roure, F., Verge´s, J. (Eds.), 2007. Thrust Belts and Foreland Basins—From Fold Kinematics to Hydrocarbon Systems. Springer, Berlin, Heidelberg, 492 pp. Liangshu, W., Liangshu, W., et al., 2003. Tectonic evolution of Tarim basin. In: Tarim Basin—A Twelve Volume Multi-author Synthesis, vol. 1. Petroleum Industry Press, Beijing, p. 202. Lawrence, S.R., 1990. Aspects of the petroleum geology of the Junggar basin, Northwest China. In: Brooks, J. (Ed.), Classic Petroleum Provinces, Geological Society of London Special Publi- cation, 50, pp. 545–557. Lawton, T.F., 2008. Laramide sedimentary basins. In: Miall, A.D. (Ed.), The Sedimentary Basins of the United States and Canada. In: Hsu¨, K.J. (Ed.), Sedimentary Basins of theWorld, vol. 5. Else- vier, pp. 429–450 (Chapter 12). Lugo, J., Mann, P., 1995/1998. Jurassic-Eocene Evolution of Maracaibo basin. In: Tankard, A.J., Suarez-Soruco, R., Welsink, A.J. (Eds.), Petroleum Basins of South America, American Associa- tion of Petroleum Geologist Memoir 62, pp. 699–726. Macqueen, R.W., Leckie, D.A., 1992. Foreland Basins and Folded Belts. AAPG Memoir 55, 460. Mann, P. (Ed.), 1999. Caribbean Basins. In: Hsu¨, K.J. (Ed.), 1999. Sedimentary Basins of theWorld. Elsevier, p. 736. Mann, P., Escalona, A. (Eds.), 2006. 3-D. Anatomy of a supergiant, Maracaibo Basin, Venezuela. American Association of Petroleum Geologists Bulletin, vol. 90, pp. 443–696. Mascle, A., Puigdefabregas, C., Fernandez, M., Luterbacher, H.P. (Eds.), 1998. Cenozoic Foreland Basins of Europe. Geological Society London Special Publication, 134, pp. 400. Miall, A.D. (Ed.), 2008. The Sedimentary Basins of the United States and Canada. In: Hsu¨, K.J. (Ed.), 2008. Sedimentary Basins of the World, vol. 5. Elsevier, p. 610. *Mossop, G.D., Shetsen, I. (Eds.), 1994. Geological Atlas of the Western Canada Sedimentary Basin. Canadian Society of Petroleum Geologists and Alberta Research Council, Special Report 4, URL Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps http://www.ags.gov.ab.ca/publications/wcsb_atlas/atlas.html. March 15, 2011. Parnaud,F.,Gou,Y., Pascual, J.C., Truskovski, I.,Gallango,O.,Passalaqua,H.,1995.Stratigraphic syn- thesis of Western Venezuela. In: Tankard, A.J., Suarez-Soruco, R., Welsink, A.J. (Eds.), Petroleum Basins of South America, American Association of Petroleum Geologist Memoir 62, 681–699. Ricketts, B.D., 2008. Cordilleran sedimentary basin of Western Canada record 180million years of terrane accretion. In: Miall, A.D. (Ed.), The Sedimentary Basins of the United States and Canada. In: Hsu¨, K.J. (Ed.), Sedimentary Basins of the World, vol. 5. Elsevier, pp. 363�394 (Chapter 2). *Rocky Mountain Association of Geologists, 1972. Geologic Atlas of the Rocky Mountain Region. In: Rocky Mountain Association of Geologists, Denver, CO, USA, p. 331, CD-ROM. Sharland, P.R., Archer, R., Casey, D.M., Davies, R.B., Hall, S.H., Heward, A.P., Horbury, A.D., Simmons, M.D., 2001. Arabian plate sequence stratigraphy. GeoArabia Spec. Publ., 2, 369 pp. Stockmal, G.S., Cant, D.J., Bell, J.S., 1992. Relationship of the stratigraphy of the Western Canada foreland basin to Cordilleran tectonics: insights from geodynamic models. In: Macqueen, R. W., Leckie, D.A. (Eds.), Foreland Basin and Thrust Belts, American, Association. Petroleum Geologists Memoir 55, pp. 107–124. Tankard, A.J., Suarez-Soruco, R., Welsink, A.J. (Eds.), 1995. Petroleum Basins of South America, In: American Association of Petroleum Geologist Memoir 62, pp. 800. VanWagoner, J.C., Bertram,G.T., 1995. Sequence stratigraphy of foreland basin deposits: outcrop and subsurface deposits. Am. Assoc. Petrol. Geol. Bull. 64, 490. Wright, G.N., McMechan, M.E., Potter, D.E.G., Holter, M.E., 1994. Structure and architecture of the Western Canada sedimentary basin. In: Mossop, G.D., Shetsen, I. (Eds.), Geological Atlas of the Western Canada Sedimentary Basin. Canadian Society of Petroleum Geologists and Alberta Research Council, Special Report 4, URL http://www.ags.gov.ab.ca/publications/ wcsb_atlas/atlas.html. March 15, 2011. Ziegler, M.A., 2001. Late Permian to Holocene Paleofacies distribution. GeoArabia, 1, 445–503. Basins located within orogens (Episutural basins of Bally and Snelson, 1980, or epi-eugeosynclines and successor basins of earlier authors) Plate 25.9:Diffuse Plate Boundaries, Cenozoic/Mesozoic Folded Belts, and Ceno- zoic Rifts Plate 25.17: Orogenic Float Plate 25.18: Cenozoic/Mesozoic Orogens, Active Margin Fold and Thrust Belts (AMFTBs), Foreland Fold and Thrust belts (FFTBs), Basement Uplifts and Cenozoic Rifts Plate 25.30: Tertiary Rifts and Plate Motions Plate 25.36: Foreland basins, Foreland Fold and Thrust Belts and Episutural basins 1133 Episutural basins are shown in orange on Plate 25.36. Traditionally, many Episu- 1134 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps tural basins are also often referred to as “Successor Basins.” These basins also roughly correspond to the “Epi-eugeosynclinal Basins” of the now obsolete geo- synclinal classification. Bally and Snelson (1980) introduced episutural basins. However, the term “megasuture” was never widely accepted and is now dropped. On the other hand, episutural basins (i.e., basins located on “megasutures”) were more readily accepted and are here retained for lack of a better term. Today, it appears implausible to retain “successor basin” as a useful term because most if not all basins represent a succession of many, mostly tectonically, controlled megasequences that are modulated by eustatic sea-level changes. Reasons for specifically segregating episutural basins are as follows: (1) Subduction-related orogens are well-circumscribed and commonly subdivided into accretionary and collisional orogens that are both subduction related. A majority of oceanic subduction zones are subjected to subduction erosion (e.g., Scholl and von Huene, 2007). However, some ocean subduction bound- aries are characterized by ocean-verging accretionary wedges. The latter contrast with and appear antithetic to FFTBs that verge toward the adjacent subducting/underthrustingcontinental/cratonic foreland (Plates25.17–25.18). (2) Cenozoic-Mesozoic orogens are best preserved and often located between ocean-verging and continent-verging FTBs (Plates 25.17 and 25.36) or else, in the case of collisions, that is, between two conjugate continent-verging FTBs. The intervening complex orogen (i.e., the Internides of Plate 25.17) is entirely allochthonous and represents the product of processes that occurred within former and present diffuse plate boundaries. Smaller plates that are now well defined by Bird (2003) within and excised from Cenozoic/ Mesozoic orogens may be relatively more ephemeral “miniplates” over lon- ger periods. Typical episutural basins were deformed by extension/transtension, compression/ transpression, or a combination of the two. For this reason, several Tertiary episu- tural basins coincide with AMFTBs (Plate 25.18), while others are characterized by complex, often diffuse rift systems. The relation of present plate motions to Tertiary strike-slip fault systems and Ter- tiary rifts that are superposed on the Cenozoic/Mesozoic orogens shown on Plate 25.30 suggests that many rift systems may in fact be incipient epi-sutural basins that are associated with strike-slip fault systems.Most of these systems are due to a combination of varying degrees of oblique plate convergence and slab-rollback as well as varying convergence rates. Jolivet et al. (1999) in a timely and concise review list the principal components that best explain the kinematics of oceanic backarc basins as (1) the slab-pull force that causes slab retreat; that is, slab or trench rollback is due to density contrasts between the subducting oceanic or continental slab and asthenosphere, (2) lateral density contrasts within the crust associated with crustal thickening that leads to crustal spreading, and (3) far-field stresses due to intracontinental deformation. In this context, Jolivet et al. (1999) Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps also note the importance of strike-slip faults systems and also compare and contrast the Sea of Japan backarc basin with the Northern Tyrrhenian Sea and the Aegean Sea. These authors address the formation of oceanic and/or partially continental backarc basins (e.g., horizontal line symbols on Plates 25.31 and 25.36). Many of their arguments also apply to most continental backarc/episu- tural basins. Furthermore, a significant number of strike-slip-related episutural basins on continents and their active margins underwent mainly compressional or else alternating extensional and compressional episodes. Plate 25.30 is a simplification of Chamot-Rooke and Rabaute (2006) combined with some of the basic elements of this global map series. It shows today’s nearly completed Africa–Europe collision, the impact of the counter-clockwise rotation of the Indian plate opposed by the overall clockwise rotation of the Pacific plate. Table 25.2 of this chapter lists and classifies various types of orogenic settings shown on Plates 25.11 and 25.30. Both plates also show schematically (with blue lines) the direction of Cenozoic/Mesozoic continental strike-slip faults that are often due to far-field stresses within these orogenic systems. Thus, larger, often diffuse, continental rift systems within the Cenozoic/Mesozoic orogen are tied to supra-regional strike-slip fault systems, For example, the Tarim basin of Western China is underlain by a remnant Precambrian basement/“cratonic” core that is surrounded by compressional and transpressional Tertiary “Himalayan” FTBs, that involve Paleozoic orogens of varying age. Significantly, the Tarim basin and the adjacent Tibetan plateau are also located at the apex of intersecting northwest- striking West-Central Asian, northeast-striking Mongolian/NE China and a south- east striking SE-Asian strike-slip fault systems. Many of these fault systems were prominent in the discussions that followed the seminal India–Eurasia collisional/ indentation and escape ideas originally proposed by Molnar and Tapponier (1975, 1978) and Tapponier and Molnar (1979). Their concepts visualized that the above listed strike-slip systems were due to the “indentation” of India into Eur- asia and the consequent lateral escape along strike-slip faults systems in Central Asia and beyond. The distribution of Tertiary extensional and transtensional rifts system in Eastern Asia is closely tied to these supra-regional strike-slip systems, while the western and central Asian compressional/transpressional mountains are associated with the West-Central Asia strike-slip system. Converted to our basin terminology and shown on Plate 25.31, collision-related basins mostly associated with north-west striking transpressional faults are recog- nized as foreland basins (including the Tarim basin and the adjacent Chaidam basin), that is, the “Chinese Basins” of Bally and Snelson (1980). On the other hand, the Jurassic/Lower Cretaceous transtensional rift systems that characterize the basins of Mongolia and NE-China, as well as the Tertiary rifts of Eastern China onshore and offshore, show the influence of Pacific Ocean Slab retreat combined with the far-field stresses associated early on with the Mesozoic collisions of Tibetan Plateau terranes and, culminating with the India–Eurasia collision! 1135 The Mesozoic and Tertiary rifted/episutural basins of Eastern China and Mongo- 1136 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps lia, and their association with slab rollback, contrast greatly with the complex strike-slip related episutural basins of California and also New Zealand. To ensure legible graphics, the rifts shown on Plate 25.30 over-emphasize their association with active margins at the expense of compressional and transpressional episu- tural and also foreland basins. In comparing Plate v18 with Plate 25.36, it becomes evident that our episutural basins also coincide with well-defined com- pressional to transpressional active margins that are all lumped in our map unit “Active Margin Fold and Thrust belts” (Plate 25.18). In fact, many episutural basins are characterized by alternating extensional/transtensional and com- pressional/transpressional regimes. In the final analysis, all these episodes are ultimately a function of varying plate convergence rates, varying degrees of convergence obliquity, and the nature of oceanic and/or continental forelands. Table 4.2 of Chapter 4, Vol. 1A (Roberts and Bally) lists idealized, succeeding tec- tono-stratigraphic megasequences of eight different types of episutural basins. Note that all the examples listed for the different types are from Tertiary basins that are best understood in a “neo-plate tectonic” context (Plate 25.30). The scale of our global maps does not allow depiction of the many well-defined smaller episutural basins or sub-basins. However, most episutural basins are easily identifiable on the 1:5,000,000 Exxon Production Research (1985) Tectonic Map. As with all basins, there is a voluminous, but dispersed, literature on episu- tural basins. While forearc and backarc basins are widely accepted episutural basin types, episutural basins that are frequently tied to strike-slip systems and/ or slab-rollback-generated extension are frequently simply listed as rifted basins. For example, the basins of Eastern China described in Chapters 35, Vol. 1B (Li Desheng), 36, Vol. 1B (Hsiao and Graham), 37 Vol. 1B (Graham et al.) and 44, Vol. 1B (Li Desheng) are included in segments concerning rifts and schemati- cally shown on Plate 25.29, yet on Plate 25.31 these rifts are also assigned to the episutural basin class. Crowell’s work on the Ridge basin of California was instrumental for the under- standing of the role of strike-slip tectonics in the formation of basins (Crowell, 1974, 1982, 2003). Mann et al. (1983) and Christie-Blick and Biddle (1985) fur- ther elaborated the important role of pull-apart basins. Cunningham and Mann (2007) and Mann (2007) reviewed the relationship of restraining bends in strike-slip fault systems and their relationship to pull-apart basins. Biddle (1991) and Wright (1991) published papers that focus on the Los Angeles basin which, considering its relatively small area, is often listed as one of the richest hydrocarbon- producing basins of the world. Ingersoll and Rumelhart (1999) aptly summed up the polyphase history of that basin. Most, if not all, of California’s basins have their own complex polyphase history. As one aspect of the complex Cordilleran subduction-related sedimentary basins, Ingersoll (2008) reviews the setting of California’s post-Nevadan forearc basins. Some of the complex terranes of theWestern Canadian Cordillera are overlain by a Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps number of various types of episutural basins (Ricketts, 2008).They are only sche- matically shown on Plate 25.36. A compendium of North American offshore “epi- sutural” basins that provides a first order introductionwas compiled by Scholl et al. (1987). The basins of the Bering Sea were reviewed byWorrall (1991), Miller et al. (2002), and Klemperer et al. (2002). In contrast to Tertiary episutural basins, preserved Paleozoic episutural basins can only be placed in the context of reconstructions that often are more debatable. Following Gibling et al. (2008), the Paleozoic Maritime/Gulf of St Lawrence basin of Eastern Canada was initiated by Mid- and Upper-Devonian rifting, followed by the deposition of evaporites and carbonates in an overlying Mississippian sag basin. Another separate, late Mississippian/Pennsylvanian rifting episode followed prior to the development of the wide Pennsylvanian sag basins. Finally, Triassic rift- ing southeast of the basin heralded the onset of ocean spreading that separated the basin from similar, coeval basins in Europe. The formation of the Maritime basin was terminated prior to opening of the N. Atlantic. On plate tectonic reconstructions, equivalent Paleozoic episutural basins that originally were next to and on strike with the Maritime basin are now buried under the North Sea and in Northern Germany (Gibling et al. 2008). What was classified as an episutural basin in Eastern Canada morphed into the subsidiary and later tectono-stratigraphicmegasequences of theNorth Sea and Northern Germany rift systems and their overlying Mesozoic/Tertiary infill (e.g., Baldschuhn et al., 2001; Ziegler, 1990). On Plate 25.31, these basins were assigned to the more general class of Cratonic basins that overlie a complex Paleozoic cra- tonic basement. Note that the change from an episutural basin in theMaritimes to become one of several megasequences of a cratonic basin is the consequence of the necessity to define basin outlines by their youngest/topmost preserved tec- tono-stratigraphic megasequence, and not by its origin. This approach simply accentuates the polyphase evolution of the great majority of all basins. The Sverdrup basin of the Canadian Arctic is a fine example of another Paleo- zoic episutural basin. The outline of this basin is seriously distorted on the Mer- cator projection of Plate 25.31; however, Plate 25.32A shows a more realistic outline of the Sverdrup basin. A good overview of the Sverdrup basin is given by Embry and Beauchamp (2008). Following these authors, the Sverdrup basin straddles the deformed and partly metamorphosed Lower- to Mid-Paleozoic fold belt to the north and, to the southeast and the south, the Ellesmere and Parry Island FFTBs. Carboniferous to lower Permian rift systems mark the incep- tion of the Sverdrup basin. The overlying Permo-Mesozoic sag basin is filled with Mesozoic clastics. An arch-like uplift “the Sverdrup Rim” separates the Sverdrup basin from the Amerasian passive margin to the north. Early to Late Cretaceous volcanics and sills relate to the emplacement of the Alpha ridge (Plates 25.16A and 25.25). Far field stress related to the late Paleogene opening 1137 of Baffin Bay and associated Nares Strait transform zone resulted in folding and 1138 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps diapirism in the Sverdrup basin. Subsequently, the basin was also uplifted in the east, that is, tilted westward. The evolution of the Sverdrup basin contrasts greatly with the simpler “peri- cratonic” evolution of theMelville Island fold and thrust belt, located to the south- west (Harrison, 1995; and see Chapter 24B, Vol. 1C. (Harrison and Brent)). Yet, on central and eastern Melville Island, a Mid-Ordovician evaporite/salt layer facili- tated the formation of a salt-based foreland fold belt that involved the upper parts of a former Mid- Ordovician-Lower Devonian passive margin sequence, as well as an overlying Mid- to Upper-Devonian foreland basin sequence sourced from the Franklinian fold orogenic belt to the north. The Melville FFTB is rightly included in the deformed Paleozoic orogen on Plates 25.16A and 25.31A. However, below the base of the detachment level, seismic profiles crossing Melville Island reveals a gently northward dipping and only mildly deformed, about 5-km-thick sedimen- tary sequence that is the relict of a former passive margin. The underlying low-angle unconformity separates this sequence from about 9–10-km-thick well-bedded Proterozoic that is laced with Mid- and Late-Proterozoic sills. These in turnmay be related to late Proterozoic rifting events. Finally, the lowest reflector separates the overlying sediments from a presumably crystalline Precambrian basement. The depth of the Moho in this area is not well defined (Darbyshire, 2003; Kanasevich and Berkes, 1990), but assuming a depth of say 35–40 km and given an average thickness of say 20 km of Proterozoic/Cambro–Devonian sediments, the thickness of the underlying cratonic basement may be in the order of 15–20 km suggesting a basement wedge that may thin and dip in a northerly direction toward the Sverdrup basin to north. Note the Carboniferous to Lower Permian rifts of the Sverdrup basins “trespass” and overlie part of the northern Melville Island fold and thrust belt. To conclude the evolution of the Sverdrup basin is overlapped to the north by the evolution of the Mesozoic passive margin and also includes the dike system asso- ciated with Alpha Ridge LIP of the Arctic. The evolving Sverdrup episutural basin, in turn overlaps the Melville Island fold and thrust belt that deforms the upper layers of a passive margin sequence as well as an overlying foreland basin sequence. However, the FFTB also masks the thick lower part of an Ordovi- cian/Late Proterozoic passive margin (or else cratonic) system that appears to be rifted in some areas. Underlying this sequence is a wedge of a, perhaps atte- nuated, cratonic Precambrian shield to the south. The comparison of different successively over-lapping tectonic and basinal provinces underlines the limita- tions of any classifications. Instead it emphasizes the importance of thoughtful, carefully documented separation of tectono-stratigraphic megasequences (see Chapter 4, Vol. 1A (Roberts and Bally)). Given that these generally are tectoni- cally controlled 1st to 3rd order sequences, it is also desirable to supplement the description of the separating unconformities with subcrops (Paleogeologic maps) that integrate all relevant subsurface data. References to Basins located within orogens. (Episutural basins ofBally and Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Snelson, 1980 or epi-eugeosynclines and successor basins of earlier authors) Baldschuhn, R., Bindt, P., Plug, S., Kokell, T., 2001. Geotektonischer Atlas von Nordwestdeutsch- land und dem deutschen NW Sector. Geol. Jahrbuch Reihe A 153, 3–93. Bally, A.W., Snelson, S., 1980. Realms of subsidence. In: Miall, A.D. (Ed.), Facts and Principles of World Petroleum Occurrence, Canadian Society of Petroleum Geologists Memoir 6, pp. 9–94. Biddle, K., 1991. ActiveMargin Basins, American Association of PetroleumGeologists Memoir 52, pp. 324. Bird, P., 2003. An updated digital model of plate boundaries. Geochem. Geophys. Geosyst. 4, 3. An Electronic Journal of the Geosciences 102. Chamot-Rooke, N., Rabaute, A., 2006. Plate Tectonics from Space. 1:50 000 000. CGMW, Paris. (Comm, 1985). Christie-Blick, N., Biddle, K.T., 1985. Deformation and basin formation along strike slip faults. In: Biddle, K.T., Christie-Blick, N. (Eds.), Strike-Slip Deformation, Basin Formation and Sedimen- tation, vol. 37. Special Publication, Society of Economic Paleontologists and Mineralogists, pp. 1–34. Crowell, J.C., 1974. Sedimentation along the San Andreas fault. In: Dott, R.H., Shaver, R.H. (Eds.), Modern and Ancient Geosynclinal Sedimentation, vol. 19. Special Publication, Society of Eco- nomic Paleontologists and Mineralogists, pp. 292–303. Crowell, J.C., 1982. The tectonics of the Ridge basin Southern California. In: Crowell, J.C., Link, M.H. (Eds.), Geologic History of Ridge Basin. Southern California, Pacific Section of Economic Paleontologists and Mineralogists, pp. 25–42. Crowell, J., 2003. Tectonics of Ridge Basin Region, southern California. In: Crowell, J.C. (Ed.), Ridge basin: An Interplay of Sedimentation and Tectonics, Geological Society of America Spe- cial Paper 367, pp. 157–203. Cunningham, W.D., Mann, P., 2007. Tectonics of strike-slip restraining and releasing bends. In: Cunningham,W.D., Mann, P. (Eds.), Tectonics of Restraining and Releasing Bends. Geological Society London Special Publication, 290, pp. 1–12. Darbyshire, F.A., 2003. Crustal Structure across the Canadian High Arctic Region from teleseismic receiver function analysis. Geophys. J. Int. 152, 372–391. Embry, A., Beauchamp, B., 2008. Sverdrup Basin. In: Miall, A.D. (Ed.), The Sedimentary Basins of the United States and Canada. In: Hsu¨, K.J. (Ed.), Sedimentary Basins of theWorld, vol. 5. Else- vier, pp. 451–472. Exxon Production Research. (1985). Tectonic Map of the World, World Mapping Project, Scale 1: 5.000 000, 20 panels, published by Am. Assoc. Petrol. Geol. Gibling, M.R., Culshaw, N., Rigel, M.C., Pascucci, V., 2008. The Maritimes Basin of Atlantic Canada: basin creation and destruction in the Collisional Zone of Pangea. In: Miall, A.D. (Ed.), The Sedimentary Basins of the United States and Canada. In: Hsu¨, K.J. (Ed.), Sedimen- tary Basins of the World, vol. 5. Elsevier, pp. 211–243. Harrison, J.C., 1995. Melville Islands Salt-Based Fold Belt, Arctic Canada. Geological Survey of Canada Bulletin 472, 331 pp. Ingersoll, R.V., 2008. Subduction-related sedimentary basins of the U.S.A. Cordillera. In: Miall, A.D. (Ed.), The Sedimentary Basins of the United States and Canada. In: Hsu¨, K.J. (Ed.), Sedimentary Basins of the World, vol. 5. Elsevier. 1139 Ingersoll, R.V., Rumelhart, P.E., 1999. Three stage evolution of Los Angeles basin Southern California. Geology 27, 593–597. 1140 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Jolivet, L., Facenna, C., d’Agostino, N., Fournier, M., Worrall, D., 1999. The kinematics of backarc basins, examples from the Tyrrhenian, Aegean and Japan Seas. In: MacNiocaill, C. (Ed.), Continental Tectonics, Geological Society London Special Publication, 164, pp. 21–53. Kanasevich, E.R., Berkes, Z., 1990. Seismic structure of the Proterozoic on Melville Island, Cana- dian Arctic Archipelago. Marine Geology 93, 421–448. Kanasewich, E.R., Berkes, Z., 1990. Seismic structure of the Proterozoic on Melville Island, Cana- dian Arctic Archipelago. Mar. Geol. 93, 421–448. Klemperer, S.L., Miller, E.L., Grantz, A., Scholl, D.W., the Bering Chuckchi Working Group, 2002. Crustal Structure of the Bering and Chukchi shelves; Deep seismic reflection profiles across North American continent between Alaska and Russia. In: Miller, E.L., Grantz, A., Klemperer, S. (Eds.), Crustal Structure of the Bering Sea: Tectonic Evolution of the Bering shelf-Chukchi Sea-Arctic Margin and Adjacent Landmasses, Geological Society of America Special Paper 360, pp. 1–24. Mann, P., 2007. Global catalogue, classification and tectonic origins of restraining and releasing bends on active and ancient strike-slip fault systems. In: Cunningham, W.D., Mann, P. (Eds.), Tectonics of Restraining and Releasing Bends, Geological Society London Special Publication, 290, pp. 13–142. Mann, P., Hempton, M.R., Bradley, C.D., Burke, K., 1983. Development of Pull-apart basins. J. Geol. 91, 529–554. Miall, A.D. (Ed.), 2008. The Sedimentary Basins of the United States and Canada. In: Hsu¨, K.J. (Ed.), Sedimentary Basins of the World, vol. 5. Elsevier, p. 610. Miller, E.L., Grantz, A., Klemperer, S. (Eds.), 2002. Crustal Structure of the Bering Sea. Tectonic Evolution of the Bering shelf-Chukchi Sea-Arctic Margin and Adjacent Land masses. Geologi- cal Society of America. Special Paper 360, 393 pp. Molnar, P., Tapponier, P., 1975. Cenozoic Tectonics of Asia: effects of a continental collision. Science 189, 419–426. Molnar, P., Tapponier, P., 1978. Active tectonics in Tibet. J. Geophys. Res. 83, 5361–5364. Ricketts, B.R., 2008. Cordilleran basins of Western Canada. Record 180 million years of terrane accretion. In: Miall, A.D. (Ed.), The Sedimentary Basins of the United States and Canada. In: Hsu¨, K.J. (Ed.), Sedimentary Basins of the World, vol. 5. Elsevier, pp. 363–395. Scholl, D.W., Grantz, A., Vedder, J.G., 1987. Geology and resource potential of the continental margin of western N. America and adjacent ocean basins-Beaufort Sea to Baja California. Circum-Pacific Council for Energy and Mineral Resources. Earth Sci. Ser. 6, 799. Scholl, D.W., von Huene, R., 2007. Crustal re-cycling at modern subduction zones applied to the past—Issues of growth and preservation of continental basement crust, mantle geochemistry, and supercontinent reconstruction. In: Hatcher Jr., R.D., Carlson, M.P., McBride, J.H., Martı´- nez-Catala´n, J.R. (Eds.), 4-D Framework of Continental Crust, Geological Society of America Memoir 200, pp. 9–32. Tapponier, P., Molnar, P., 1979. Active faulting and Cenozoic tectonics of the Tienshan, Mongolia and Baikal regions. J. Geophys. Res. 84, 3425–3459. Worrall, D.M., 1991. Tectonic History of the Bering Sea and the Evolution of Tertiary Strike-Slip Basin on the Bering Shelf. Geological Society of America Special Paper Vol. 257. 120. Wright, K.T., 1991. Structural geology and Tectonic evolution of the Los Angeles basin. In: Biddle, K. (Ed.), Active margin basins, American Association of Petroleum Geologist Memoir 52, pp. 35–134. Ziegler, P.A., 1990. Geological Atlas of Western and Central Europe, second ed. Shell Int. Petrol Mij. B. BV., distributed by Geol. Pub. House, Bath 239 pp, 56 enclosures. Oceanic basins formed by spreading ridges (incl. oceanic back-arc basins) Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Plate 25.1: Topography and Bathymetry Map Plate 25.2A: Arctic Topography and Bathymetry Map—Ice surface Plate 25.2B: Antarctic Topography and Bathymetry Map—Bedrock surface Plate 25.27: Cenozoic, Mesozoic, Paleozoic Orogenic systems; Continental basement; Foreland Fold and thrust belts (FFTB’s); Active Margin Fold and Thrust Belts (AMFTB’s ); Age of Oceanic Basement formed onMid-Oceanic Ridges (MOR ) and also in oceanic backarc basins and Large Igneous Provinces (LIPs) Plate 25.33: Passivemargins, Age of Basement andCenozoic/Mesozoic Rifts. The Oceans form by far the largest sedimentary basin of the world, yet the basin clas- sification here presented does not include an effort to list and order various regions of the oceans. Most, if not all, published basin classifications address sedimentary basins on continents including their active and passive margins, whereby particu- larly the offshore boundary of a passive margin basin is not always clearly defined (see Plate 25.33). The above-listed plates do highlight selected tectonic aspects of all the oceans. Unlike the remainder of the maps presented in this chapter and discussed in this accompanying text, no attempt has beenmade to thematically classify ocean basins or provinces. The above listed plates provide an overview of the linkages between global bathymetry, the age of the oceanic crust in the different ocean basins, and the timing of the onset of spreading along passive margins. Plates 25.1 and 25.27 serve as overall background for the different tectonic settings of the oceanic crust and are also well introduced in selected themes of the following chapters of Vol.1A: – Ocean Floor tectonics: Chapter 26 Vol.1A (Fowler) – Ocean floor volcanism in: Chapter 3 Vol.1A (Kerr and Menzies) – Hotspots, rifts and reefs viewed from space in Chapter 9 Vol.1A (Dickerson) – Wide-angle seismic on ocean and their margins in Chapter 11, Vol.1A (White) – Pelagic sedimentation and stratigraphy in Chapter 19, Vol. 1A (Winterer) – The formation of organic-matter-rich rocks along Ocean continent margins in Chapter 23 Vol.1A (Bohacs et al.) – Tectonostratigraphic megasequences on Ocean/Continent margins in: Chap- ter 4 Vol.1A (Roberts and Bally) – For an overview of the geology of the oceans, we refer to Steele et al. (2001). General reference Steele, J.H., Turekian, K.K., Thorpe, S.A. (Eds.), 2001. Marine Geology and Geophysics: A Deriva- tive of the Encyclopedia of Ocean Sciences (Paperback). Academic Press, Elsevier, p. 640. 1141 SEDIMENTARY BASINS AND RIFTs: Segment 25.11 (For online version of theplates/figures cited in this chapter, the reader is referred to http://www. elsevierdirect.com/companion.jsp? ISBN=9780444563576). 1142 Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps Passive Margins: Ocean / Continent Boundary a c ba. undifferentiated b. poorly defined c. with volcanic seaward-dipping reflectors (SDRS) Folded (Orogenic) Belts Ages of Deformation/ Basement Sedimentary Basins on Continental and Transitional Crust Oceanic Basins Oceanic (B) Subduction Boundary Tertiary Mesozoic Paleozoic Proterozoic Continental (A) Subduction Boundary Major Strike Slip Faults Basement Involved Uplifts and /or Inversions Basaltic Plateaus Cenozoic / Mesozoic Vendian to Paleozoic Precambrian Basement Precambrian Shield Passive Margin Cratonic Platform / Shallow Basins Deep (> 2 km) Cratonic Basin Cratonic Arches Foreland Basins Deep Sea Trenches Episutural BasinsForeland Fold & Thrust Belts (FFTB’s) Active Margin Fold and Thrust Belts (AMFTB’s) a. Outcrops b. Subsurface a. Palaeozoic b. Cenozoic/ Mesozoic a. b. a. b. Oceans Oceanic Back-Arc Basins * * On Plates 31, 32 a & b, and 35 only Modified from Sengor & Natalin 2001 and other sources .. Rifts Passive Margins with extensive volcanic cover Plate 25.28 Legend for sedimentary basin maps. 1 1 4 3 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s Credits: Based on and modified from Sengor and Natalin 2001 and other sources. Combined with parts of Bally (2010) Plates 9 & 18 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W A.W. Ballyc Plate 25.29 Rifts, age of rifting and age of basement. 1 1 4 4 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s Credits: Simplified and modified from Chamot Rooke and Rabaute (2006) combined with Bally (2010) Plates 9 & 18 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 150° W 120° W 90° W 60° W 30° W 0° W 30° E30° E 60° E 90° E 120° E 150° E 180° 30° E 60° E 90° E 120° E 150° E 180° 100 80 60 40 20 10 Graphic Scale: mm / yr FIXED EU RA SIA FIXED A.W. Ballyc Plate 25.30 Recent Plate Motions, Cenozoic Oceanic Basement, Back-arc Basins and Tertiary Rifts. 1 1 4 5 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s Credits: Basin Outlines based on but modified Exxon (1985), Ahlbrandt et al (2005), and other sources 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W A.W. Ballyc Plate 25.31 Sedimentary Basins of the world. 1 1 4 6 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s E W 70 S 70 S W E Credits: CGMW (1967), Harrison (2005) Reed (2005 & 2006) and other sources Credits: CGMW (1967), Harrison (2005) and Reed (2003) Plate 25.32 A. Arctic basins and B. Antarctic basins. 1 1 4 7 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S Plate 25.33 Passive margins, Age of Basement and Cenozoic/Mesozoic Rifts Plate. 1 1 4 8 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S Credits: Simplified from Ernst and Buchan (1997 & 2001) combined with parts of Plates 29 and 33 Plate 25.34 Passive margins, Cenozoic/Mesozoic Rifts and Radiating Dyke Swarms. 1 1 4 9 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s Credits: Based on but modified from Exxon (1985), Ahlbrandt et al (2005), and other sources. 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S Plate 25.35 Cratonic and Foreland Basins. 1 1 5 0 P h a n e ro z o ic P a s siv e M a rg in s, C ra to n ic B a sin s a n d G lo b a l T e c to n ic M a p s 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 150° W 120° W 90° W 60° W 30° W 0° W 30° E 60° E 90° E 120° E 150° E 180° E180° W 80° N 70° N 60° N 50° N 40° N 30° N 20° N 10° N 0° N 10° S 20° S 30° S 40° S 50° S 60° S Credits: Based on but modified from Exxon (1985), Abraham et al (2005) and other sources Plate 25.36 Foreland basins, Foreland fold and Thrust belts and Episutural basins. 1 1 5 1 P h a n e r o z o ic P a ss iv e M a rg in s, C ra t o n ic B a sin s a n d G lo b a l T e c to n ic M a p s Tectonic and Basin maps of the world Global geological maps: Introduction Global geological cartography: Selected milestones Stratigraphic nomenclature and the geological time scale References to 25.1. Global Topography and Plate Tectonics Tectonic and Basin maps of the world Introduction Global relief models: Onshore and submarine morphology and plate tectonic regimes (Plates 25.1 and 25.2A,B) References to 25.2. Global Topography Tectonic and Basin maps of the world Neotectonics: Introduction Global earthquake distribution Well-defined versus diffuse plate boundaries Neotectonic plate motions: Their relation to a fixed Eurasia and to Cenozoic/Mesozoic fold belts (Plate 25.11) References 25.3. Neotectonics: earthquakes and rigid versus diffuse plate boundaries Tectonic and Basin maps of the world References 25.4. Stress Maps and Paleostress Studies Tectonic and Basin maps of the world Introduction The continental lithosphere The continental crust Crustal layers, rheological models, and conclusions References: 25.5 The Continental Lithosphere and The Continental Crust PLATES FOR GLOBAL TOPOGRAPHY, NEOTECTONICS, THE CONTINENTAL LITHOSPHERE AND CRUST: Segments 25.1-25.3 and 25.5 (For online vers Tectonic and Basin maps of the world Introduction to tectonic maps Recent advances in alpine tectonics: An example of the scope of larger scale tectonic maps Simplified tectonic maps of the world About Phanerozoic plate tectonic reconstructions References 25.6. Tectonic Maps of the World Tectonic and Basin maps of the world Polar tectonic maps: Introduction Arctic tectonic map Antarctic tectonic map References to Polar Tectonic Maps Tectonic and Basin maps of the world Orogeny versus epeirogeny Subduction, sutures, and orogens Active margin fold and thrust belts (AMFTBs) Foreland fold and thrust belts (FFTBs) Normal faulting in foreland fold and thrust belts (FFTBs) References to 25.7 Cenozoic/Mesozoic and Paleozoic Orogenic Systems and Their Fold and Thrust Belts (FTBs) Tectonic and Basin maps of the world Introduction to basements, that is, the ``residual´´ peneplaned former fold belts Merging the global tectonic map with a Precambrian basement map References: Age of Continental Basement Tectonic and Basin maps of the world Introduction Large igneous provinces (LIPs) Giant radiating dike swarms (maps b-6 and b-7) Is there a ``canonical progression of tectonic themes´´ preceding and/or following the emergence of a plume? The distribution of active volcanoes References to 25.9 Hot Spots, Linear Island Chains, Large Igneous Provinces (LIP's) and Radiating Dyke Swarms; Active Volcanoe Tectonic and Basin maps of the world Introduction Subducted oceanic plateaus Allochthonous accreted oceanic plateaus and intra-oceanic island arc terranes Allochthonous fragments, oceanic and intra-oceanic arc systems, and lower crust and uppermost mantle of hyper-extended passive Allochthonous, exhumed continental crust-mantle transitions and the Ivrea-Verbano zone Conclusion References to 25.10 Tectonic settings of mafic/ultramafic oceanic and intraoceanic arc system crust, LIPs, rifted and volcanic PLATES FOR TECTONICS, OROGENIC SYSTEMS, HOT SPOTS, LIPS, VOLCANOES: Segments 25.6-25.10... Tectonic and Basin maps of the world Introduction References to 25.11 Sedimentary basins (incl. Rifts); Introduction Tectonic and Basin maps of the world Rift systems on relatively stable/rigid lithosphere References to Rift Systems Tectonic and Basin maps of the world Passive margins on relatively stable/rigid lithosphere References to Passive margins on relatively stable/rigid lithosphere Tectonic and Basin maps of the world Cratonic basins on relatively stable/rigid lithosphere References to Cratonic Basins Tectonic and Basin maps of the world Basins on the periphery of orogens: Deep sea trenches and foreland basins References to Basins on the Periphery of Orogens: Deep Sea Trenches and Foreland Basins Tectonic and Basin maps of the world Basins located within orogens (Episutural basins of Bally and Snelson, 1980, or epi-eugeosynclines and successor basins of earl References to Basins located within orogens. (Episutural basins of Bally and Snelson, 1980 or epi-eugeosynclines and successor Tectonic and Basin maps of the world Oceanic basins formed by spreading ridges (incl. oceanic back-arc basins) General reference SEDIMENTARY BASINS AND RIFTs: Segment 25.11 (For online version of the plates/figures cited in this chapter, the reader is refe