National Academies Press: OpenBook

Active Tectonics: Impact on Society (1986)

Chapter: 6 Coastal Tectonics

« Previous: 5 Alluvial River Response to Active Tectonics
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 95
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 96
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 97
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 98
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 99
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 100
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 101
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 102
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 103
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 104
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 105
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 106
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 107
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 108
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 109
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 110
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 111
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 112
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 113
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 114
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 115
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 116
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 117
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 118
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 119
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 120
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 121
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 122
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 123
Suggested Citation:"6 Coastal Tectonics." National Research Council. 1986. Active Tectonics: Impact on Society. Washington, DC: The National Academies Press. doi: 10.17226/624.
×
Page 124

Below is the uncorrected machine-read text of this chapter, intended to provide our own search engines and external engines with highly rich, chapter-representative searchable text of each book. Because it is UNCORRECTED material, please consider the following text as a useful but insufficient proxy for the authoritative book pages.

COASTAL TECTONICS 95 6 Coastal Tectonics KENNETH R.LAJOIE U.S. Geological Survey, Menlo Park INTRODUCTION Between one-third and one-half of the Earth's marine coastlines lie along or near tectonically and seismically active plate boundaries (Inman and Nordstrom, 1971). Many of the world's major population centers lie along these active coastlines and, therefore, are vulnerable to adverse impacts of large earthquakes. Fortunately, there has been remarkable progress over the past two decades in the field of coastal tectonics, the study of recent crustal deformation and paleoseismicity in coastal regions. Of particular importance is progress in establishing earthquake recurrence, dating recent fault activity, and measuring rates of recent crustal deformation, all of which are crucial in determining earthquake potential and assessing seismic hazards and risk. Much of the recent progress in coastal tectonics stems directly from the development of new dating techniques that have led to a better understanding of Quaternary [past 2 million years (m.y.)] sea-level fluctuations and their relationship to marine strandlines (abandoned or relict marine shorelines), which are among the most numerous and widespread tectonic markers in the Quaternary geomorphic record (Figures 6.1, 6.2, and 6.3). Sea level is the common and unifying element of coastal tectonics. Present sea level, the universal datum for measuring elevation, is the most convenient reference for detecting ongoing vertical crustal movement and assessing short-term tectonic stability in coastal areas. Accordingly, coastal tectonics often includes analysis of tide-gauge data (Figure 6.4) and other historical and archeological information that might reveal apparent sea-level changes. Past sea levels derived from the geologic record comprise a composite datum for measuring long-term crustal movement. Consequently, coastal tectonics also includes mapping and dating marine strandlines and separating the vertical crustal movements from the past sea-level fluctuations they si FIGURE 6.1 Emergent marine strandlines form steplike terraces on the Palos Verdes Peninsula of southern California. These and similar strandlines on technically active coastlines record both crustal uplift and major glacio-eustatic sea-level fluctuations (see Figure 6.6). The lowest strandline terrace is about 100 ka and the highest is about 1000 ka. The highest point on the peninsula is about 450 m above sea level. Modified from Davidson (1889).

COASTAL TECTONICS 96 multaneously record. Because marine strandlines are the physical records of past sea levels, the study of sea-level history is an integral part of coastal tectonics. Time-transgressive sequences of displaced and deformed Pleistocene [2 m.y. to 10,000 yr (10 ka) ago] marine strandlines document patterns of and, where dated, yield rates of continual, long-term crustal deformation (Figures 6.1 and 6.2). Sequences of emergent Holocene (past 10 ka) strandlines (Figure 6.3), which occur only along the most rapidly uplifting coastlines, commonly record abrupt coseismic uplift events of 1–15 m (see Figures 6.25–6.28 below) that cumulatively comprise the long-term deformation recorded by Pleistocene strandlines. Consequently, sequences of Holocene strandlines often record past earthquakes and, where dated, yield earthquake periodicity and provide a means of forecasting future seismic events. In many areas, apparent sea-level changes documented by tide-gauge records (Figure 6.4) or subtle shifts in the location of the modern shoreline reflect ongoing vertical crustal movement. Frequently, this movement is opposite in sense to long-term trends and, therefore, may represent postearthquake crustal relaxation or pre- earthquake strain accumulation. Coastal tectonics includes the study of both onshore (emergent) and offshore (submergent) marine strandlines and structural features. However, offshore coastal tectonics is a highly specialized field and is beyond the scope of this brief review, which focuses mainly on the formation and deformation of emergent marine strandlines and stresses their importance in determining the style and measuring the rates of recent crustal deformation, especially in highly active coastal regions. Most examples of strandline displacement and deformation cited in this review reflect sustained tectonic processes, but a few, included mainly for comparative purposes, reflect transitory volcanic and glacio-isostatic crustal deformation. COASTAL MORPHOLOGY AND TECTONIC SETTING On global and regional scales coastal morphology correlates closely with tectonic setting (Inman and Nordstrom, 1971). The greatest geomorphic contrast is between the subdued coastlines along passive continental margins and the rugged coastlines along convergent plate boundaries. Along most coastlines modern (active) coastal landforms are similar to their Pleistocene counterparts, which suggests that current tectonic and coastal processes have been fairly uniform over considerable periods of time. Most exposed coastlines along passive continental FIGURE 6.2 Cross-sectional profiles of emergent Pleistocene strandline terraces. A, Erosional strandlines near Santa Cruz, California. Solid line is existing topographic profile. Dashed line represents original profile of each strandline terrace, which consists of a relict wave-cut platform backed by a relict sea cliff. The intersection of the platform and the sea cliff is the shoreline angle, which closely approximates the paleoshoreline. Seaward thinning wedges of alluvial sediment derived from the degrading sea cliffs overlie the wave-cut platforms. A 120-ka strandline was removed from this section when the next lower (104 ka) platform was cut; the 120-ka strandline occurs a few kilometers north of this site and is projected onto this cross section. The three lowest strandlines were dated by paleontological, amino acid, and geomorphic techniques. The average uplift rate (0.35 m/ka) was derived from the vertical displacement (28 m) of the 82-ka strandline. The three highest strandlines were dated by extrapolation of this uplift rate and by numerical analysis of the progressively gentler slopes of successively higher (older) relict sea cliffs. Modified from Hanks et al. (1984). B, Depositional strandline terraces on the Huon Peninsula, Papua New Guinea. Each platform is a relict coral reef. U-series dates on fossil corals from these strandlines yield a history of glacio-eustatic fluctuations that serves as a tectonic datum for measuring vertical tectonic movements on other coastlines throughout the world (see Figure 6.6). The maximum average uplift rate of 4 m/ka on the Huon Peninsula was derived from the maximum vertical displacement (500 m) of the 120-ka strandline. Note that the three lowest strandlines on the Santa Cruz coastline correlate with the three highest strandlines in this sequence. Modified from Chappell (1974a).

COASTAL TECTONICS 97 margins consist of broad coastal plains bordered offshore by wide continental shelves and gentle continental slopes. These relatively stable coastlines are characterized by depositional landforms such as broad sandy beaches and offshore barrier bars and at low latitudes by broad coral reefs. The southeast coast of North America and the northeast coast of Australia are typical passive-margin coastlines with low-to-moderate topographic relief and subdued depositional landforms. Much of the former is bordered offshore by barrier bars (Oaks and Du Bar, 1974), and most of the latter by extensive coral reefs (Hopley, 1983). FIGURE 6.3 Cross-sectional profiles of Holocene strandlines. A, Erosional terraces on the Boso Peninsula of Honshu, Japan. These strandlines, unlike their Pleistocene counterparts, represent periodic coseismic uplift events, not sea-level fluctuations (see Figures 6.7, 6.25, 6.26, 6.27, and 6.28). The lowest platform (IV) records uplift that accompanied two major earthquakes in 1703 and 1923, and FIGURE 6.4 Tide-gauge records. the three highest strandlines were dated by radiocarbon A, San Francisco, California—the secular drift of this analysis of fossil shells. The average uplift rate is 3.9 m/ka record (~2 mm/yr) is similar to that found in other parts of (see Figure 6.25). Modified from Matsuda et al. (1978). the world and, therefore, probably represents eustatic rise B, Depositional strandlines (beach ridges) at Te Araroa on in sea level. North Island, New Zealand. Each ridge most likely B, Juneau, Alaska—the extreme apparent drop in sea-level represents a storm event not a coseismic uplift event. If represents crustal uplift due to either tectonic uplift or uplift events are recorded in this sequence of strandlines, residual glacio-eustatic rebound. they are indistinguishable from storm events. The average C, Mississippi delta—the apparent rise in sea-level uplift rate derived from the highest beach ridge (6 ka) is 1 represents subsidence due to sediment compaction and m/ka. Modified from Garrick (1979). isostatic adjustments of the crust to the sediment load of the Mississippi delta. The eustatic rise (2 mm/yr) was subtracted from the relative rates to obtain the uplift and subsidence rates. Modified from Hicks and Crosby (1974). Long-term tectonic stability along most passive-margin coastlines is expressed stratigraphically by undeformed continental and marine sediments that underlie flat coastal plains and continental shelves and geomorphically by broad accretionary strandline terraces that consist of subdued beach ridges separated by abandoned tidal flats (Oaks and Du Bar, 1974). However, rapid sediment accumulation, such as at the mouth of a large river, may isostatically depress an otherwise stable passive-margin coastline (Figure 6.4C) (Fisk and McFarlan, 1955; Hicks and Crosby, 1974). Also, occasional large earthquakes, such as the 1886 Charleston, South Carolina, seismic event on the Atlantic coast of North America (Hays and Gori, 1983), indicate that passive continental margins are not completely aseismic, even though there is little stratigraphic or geomorphic evidence in these areas of recent, near-surface crustal deformation. In contrast to the subdued coastlines along passive plate boundaries, most coastlines along or near active continental margins consist of coastal hills or mountains bordered offshore by narrow continental shelves and

COASTAL TECTONICS 98 deep submarine trenches, particularly in the circum-Pacific region. These tectonically active coastlines are characterized by rugged erosional landforms such as steep sea cliffs and rocky headlands, islands and sea stacks. However, at low latitudes coral reefs commonly form narrow depositional platforms even along rugged coastlines (Figure 6.2B)(Chappell, 1974a). The mountainous west coast of South America is an active-tectonic coastline that geomorphically reflects regional crustal uplift and subsidence related to rapid plate convergence and resultant subduction along the offshore Peru-Chile Trench (Plafker, 1972). The hilly-to-mountainous coast of California in western North America is an active-tectonic coastline that geomorphically reflects slower regional uplift and local basin subsidence related to oblique plate convergence and resultant large-scale, right-lateral displacement across the San Andreas Fault system. Long-term crustal instability along most active-tectonic coastlines is expressed stratigraphically by folded and faulted marine sediments that fill youthful structural basins and geomorphically by narrow uplifted and deformed Pleistocene strandline terraces that notch steep coastal slopes (Figures 6.1 and 6.2). Short-term instability along extremely active coastlines is expressed geomorphically by emergent Holocene strandline terraces (Figure 6.3) and by dramatic changes in the location and configuration of modern shorelines that result from rapid vertical crustal movements (Figure 6.4). Most of the Earth's seismicity occurs along or near active-tectonic coastlines (Tarr, 1974), and many large earthquakes in these areas are recorded geomorphically by emergent Holocene strandline terraces (Figure 6.3A; also see Figures 6.25–6.28 below). The subdued coastlines bordering the shallow epicontinental seas in North America (Hudson Bay) and Fennoscandia (Bay of Bothnia) are exceptions to the general rule that coastal regions undergoing rapid crustal deformation are characterized by marked topographic relief. However, the rapid crustal uplift recorded so dramatically by sequences of highly emergent Holocene strandlines in these recently deglaciated areas reflects transitory isostatic rebound (see Figure 6.10 below) not sustained tectonic deformation. MARINE STRANDLINES Marine strandlines are the geological and historical records of former sea levels. In the geologic record marine strandlines are the depositional and erosional remains of abandoned marine shorelines (Figures 6.1, 6.2 and 6.3), and in the historical record they are most commonly tide-gauge measurements (Figure 6.4) or high-water marks on man-made coastal structures. A strandline, or a sequence of strandlines, is a relative sea-level record that potentially represents both real and apparent sea-level changes (Figure 6.5): (6.1) Real sea-level changes are absolute vertical movements of the ocean's surface and may be local to worldwide in extent; if worldwide, they are called eustatic changes. FIGURE 6.5 Relative, apparent, and real sea-level changes. A, Late Pleistocene. B, Holocene. All sea-level records (strandlines or tide-gauge measurements) are relative, which means they potentially represent both apparent and real sea-level changes (relative=real +apparent). Apparent sea-level changes are the inverse of the vertical crustal (or ground) movements that produce them and, therefore, are the focus of all coastal tectonic studies. The apparent sea- level history is obtained by algebraically or graphically subtracting the real sea-level history from the relative record of marine strandlines. In effect, the real sea-level history is a composite tectonic datum. The real sea-level history is obtained by subtracting apparent sea level changes from a relative sea-level record. In this case, the uplifting coastline is a moving sea- level datum. E is the present elevation of a strandline; e is the original elevation of a strandline; D is the vertical displacement (D=Eíe); A is the age of a strandline; R is the crustal displacement rate (R=D/A).

COASTAL TECTONICS 99 Quaternary sea-level history was characterized by periodic eustatic fluctuations of 100–150 m caused by the advance and retreat of continental glaciers. Apparent sea-level changes are not real but result from and are the inverse of vertical ground movements. Consequently, apparent sea-level changes are only local or regional in extent. FIGURE 6.6 Pleistocene sea-level fluctuations and origin of emergent Pleistocene strandlines. Emergent strandlines simultaneously record tectonic uplift and major sea-level highstands. The rising coastline is a moving strip chart on which sea- level highstands are recorded sequentially as strandlines whose ages increase with elevation. The slope (R) of the diagonal line connecting each highstand to the elevation of its strandline is the average uplift rate. If the uplift rate was constant, the uplift lines for all strandlines are parallel. Strandlines formed during lowstands are usually destroyed by subsequent sea-level fluctuations and rarely appear in the emergent geologic record. Strandlines younger than 60 ka appear above sea level only where the uplift rate is greater than 1 m/ka. The sea-level fluctuation curve was derived from a sequence of U-series dated coral-reef strandlines on the Huon Peninsula, Papua New Guinea (Figure 6.2B) by subtracting tectonic uplift from the relative strandline record (Figure 6.5A). Sea-level curve modified from Chappell (1983); oxygen-isotope stages (1–9) from deep-sea cores (Shackleton and Opdyke, 1973). The primary task in coastal tectonics is separating the apparent component from the real component of a relative sea- level record. This is done graphically or algebraically by subtracting the real sea-level history from a relative sea-level record: (6.2) In effect, the real sea-level history is a fluctuating tectonic datum to which each strandline (tectonic marker) must be correlated by age (Figures 6.5, 6.6, and 6.7). The real sea-level history is obtained by subtracting apparent sea-level changes (the inverse of vertical crustal movements) from a detailed and well-dated relative sea-level record (Figure 6.5): (6.3) In this case, the tectonic history is an absolute sea-level datum and can be either moving or stationary. A rapidly and steadily rising coastline is the best datum for measuring major, long-term sea-level fluctuations (Figures 6.5A and 6.6) (Chappell, 1983), whereas a stable coastline is the best datum for measuring minor, short-term sea-level changes (Figures 6.5B and 6.7) (Bloom, 1970; Scholl et al., 1970). Generally, the greatest uncertainties in most long-term sea-level histories derived from strandline data stem from the simplifying assumption of constant uplift. However, on a tilted coastline where Pleistocene strandlines converge, constant long-term uplift at a particular locality can be demonstrated empirically (Chappell, 1983). The greatest uncertainties in short-term sea-level histories stem from the assumption of coastal stability. Where short-term stability at a coastal locality cannot be demonstrated by independent geodetic data, subtle sea-level changes can be expressed in only relative terms. An additional complication is FIGURE 6.7 Holocene sea-level changes and origin of emergent Holocene strandlines. In contrast to Pleistocene strandlines, emergent Holocene strandlines represent discrete uplift events or storm events, not sea-level fluctuations. The highest strandline correlates with the tangent point at the sea-level inflection between 7 ka and 5 ka BP. All lower strandlines represent two paleoshorelines, one occupied during the transgressive phase and one during the regressive phase (see Figure 6.5B for relative sea-level changes). The sea-level curve is general and may vary slightly from region to region owing to minor geoidal distortions.

COASTAL TECTONICS 100 that different measurement techniques (e.g., geodetic, tide gauge) occasionally yield conflicting results for short-term sea- level changes (Brown, 1978). Pleistocene Strandlines Along steep coastal slopes in uplifting areas Pleistocene strandlines occur most commonly as narrow (<1 km), steplike terraces within a few hundred meters above present sea level. Consequently, a vertical sequence of uplifted strandline terraces resembles a flight of stairs (Figures 6.1 and 6.2). Each strandline terrace consists of a virtually horizontal erosional or depositional platform backed by a steep sea cliff along its landward margin. The shoreline angle, the intersection of the relict platform and sea cliff, closely approximates the location and elevation of the abandoned marine shoreline and is the linear structural marker depicted on most longitudinal profiles of deformed strandlines (for examples see Figures 6.8A, 6.13, 6.17, 6.20, and 6.21). The shoreline platform, also referred to as the terrace platform, is the planar tectonic marker depicted on most cross-sectional profiles and detailed isobase maps of deformed strandlines (Figures 6.2 and 6.3A; also see Figures 6.19 and 6.21 below). Commonly, a thin (1–3 m) veneer of shallow-water sand, gravel, and cobbles, which locally contains fossil marine shells used for dating, overlies the terrace platform. Where subaerial slope degradation is rapid, a seaward-thinning prism of alluvium derived from upland streams and abandoned sea cliffs overlies the platform and its associated marine sediments and buries the shoreline angle (Figure 6.2A). In these areas the geomorphic expression of successively higher (older) Pleistocene strandline terraces is progressively subdued, and the position and configuration of both the shoreline angle and the terrace platform must be derived from borehole or shallow seismic data (see Figure 6.21 below) (Bradley and Griggs, 1976; Lajoie et al., 1979b). However, in some areas the progressively greater degradation of successively older relict sea cliffs provides a means of dating emergent strandlines (Figure 6.2A) (Hanks et al., 1984). Along gentle coastal slopes in slowly uplifting areas emergent strandlines commonly consist of broad (1–10 km) terrace platforms backed by low sea cliffs obscured by relict beach ridges and dune fields (Hoyt and Hails, 1974; Lajoie et al., 1979a). A general consensus has developed over the past two decades that a flight of emergent Pleistocene strandlines is the geologic record of periodic glacio-eustatic sea-level highstands superimposed on a rising coastline (Figure 6.6) (Broecker et al., 1968; Mesolella et al., 1969; Matthews, 1973). In this model, a rising coastal land-mass is a moving strip chart on which brief sea-level highstands were successively recorded as depositional or erosional strandlines. Strandlines also formed during sea-level lowstands (Emery, 1958; Lewis, 1971a,b, 1974; Ridlon, 1972), but along uplifting coastlines most of these strandlines were destroyed by wave erosion during subsequent sea-level fluctuations and, therefore, rarely appear in the emergent marine record (Figure 6.6). An important exception is found in the deeply incised sequence of emergent strandline terraces on the Huon Peninsula of Papua New Guinea, where sea-level lowstands are recorded as deltaic accumulations preserved beneath a protective cover of coral reefs deposited during subsequent sea-level highstands (Chappell, 1974a, 1983). Along many subsiding coastlines, strandlines formed during sea-level lowstands probably dominate the submergent geomorphic record but are difficult to distinguish from strandlines formed during highstands (Moore and Fornari, 1984). The most detailed tectonic datum for deriving uplift from emergent Pleistocene strandlines on coastlines throughout the world is the paleosea-level curve obtained by subtracting well-documented constant tectonic uplift from the relative sea-level record of emergent coral-reef strandlines on the Huon Peninsula of Papua New Guinea (Figures 6.2B and 6.6) (Veeh and Chappell, 1970; Bloom et al., 1974; Chappell, 1983). There, uranium-series dates on fossil corals yield a paleosea-level curve back to about 340 ka before present (BP) that agrees with longer, but less detailed, sea-level curves derived from emergent coral-reef strandlines on the island of Barbados in the West Indies (Broecker et al., 1968; Mesolella et al., 1969; Matthews, 1973; Bloom et al., 1974; Bender et al., 1979) and from oxygen-isotope data from deep-sea cores (Shackleton and Opdyke, 1973). The uplift rate (maximum 4.0 m/ka) of the Huon Peninsula was derived from the elevation of the prominent 120-ka strandline that formed 2–10 m above present sea level during the last major interglacial sea-level highstand (Thurber et al., 1965; Veeh, 1966; Ku et al., 1974; Neumann and Moore, 1975; Marshall and Thom, 1976; Stearns, 1976; Harmon et al., 1978; Shubert and Szabo, 1978; Szabo et al., 1978; Szabo, 1979). The most important features of the New Guinea paleosea-level curve as a tectonic datum are the periodic interglacial highstands at approximately 100 ka intervals and the successively lower interstadial highstands at approximately 20 ka intervals over the past 120 ka. Virtually all Pleistocene strandlines along emergent coastlines throughout the world were formed during these brief paleosea-level highstands. The longer paleosea-level records on Barbados and in the deep-sea cores indicate that the main 100-ka cycle of interglacial high

COASTAL TECTONICS 101 stands extends back to at least 700 ka (Shackleton and Opdyke, 1973; Bender et al., 1979). Morner (1976, 1983) argued that no locally derived Pleistocene sea-level curve should be used as a worldwide tectonic datum because the sea surface, an approximate geoid or equipotential gravitational surface, is grossly distorted by gravitational fluctuations on a time scale of 10–100 ka. If correct, this supposition implies that major, glacially induced changes in ocean volume were expressed by neither synchronous nor uniform sea-level fluctuations. However, similar absolute ages of Pleistocene strandlines on emergent coastlines throughout the world and, more importantly, similar elevations of sea-level highstands on several independently derived late Pleistocene sea-level curves demonstrate that, within reasonable margins of uncertainty (±5 ka and ±10 m; Stearns, 1976, 1984; Harmon et al., 1979; Chappell, 1983), major sea-level fluctuations were synchronous and relatively uniform over at least the past 300 ka, and most likely muc h longer (Veeh and Valentine, 1967; Bloom et al., 1974; Konishi et al., 1974; Ku and Kern, 1974; Moore and Samayajulu, 1974; Chappell and Veeh, 1978; Harmon et al., 1978; Marshall and Launay, 1978; Bender et al., 1979; Dodge et al., 1983; Ward, 1985; also see references on elevation of 120-ka highstand in preceding paragraph). Also, reasonable graphical correlations of undated strandline sequences in Japan (see Figure 6.14 below) (Miyoshi, 1983), New Zealand (W.Bull, University of Arizona, personal communication, 1983), and California (Hanks et al., 1984) with 120-ka to 40-ka highstands on the New Guinea paleosea-level curve indicate that strandlines from all four areas were synchronous and had a common datum. An apparent exception is found along the southeast Atlantic coast of North America, where the ages, but not the relative heights, of late Pleistocene highstands agree with paleosea-level data from New Guinea and other parts of the world (Belknap, 1979; Cronin et al., 1981). However, the discrepancies of 20–30 m found along this passive continental margin may represent subtle tectonic or isostatic crustal movements not regional sea-level variations. In any event, although rapid geoidal distortion is a potential problem that requires further investigation, most existing data indicate that geoidal distortion could not have been significant over late Pleistocene time. Consequently, there is no compelling reason not to use the New Guinea sea-level curve worldwide as a first-order tectonic datum, at least over the past 120 ka. Because virtually all emergent Pleistocene strandlines were formed during sea-level highstands (Figure 6.6), the task of dating a particular strandline is reduced to correlating it with a specific peak on the New Guinea paleosea-level curve using one or more absolute (isotopic) or relative (geomorphic, paleontologic, and chemical) dating techniques (Chappell and Veeh, 1978; Kennedy et al., 1982; Sutherland, 1983; Hanks et al., 1984). Uranium-series analyses of fossil corals yield the most reliable absolute dates for Pleistocene strandlines (Broecker and Bender, 1973; Harmon et al., 1979), but because fossil corals are virtually restricted to tropical latitudes, less-reliable dating techniques must be used at higher latitudes. Frequently, only one or two Pleistocene strandlines in an emergent sequence can be dated independently, but the ages of other strandlines can be inferred from the resultant uplift rate (Chappell and Veeh, 1978; Hanks et al., 1984; Ward, 1985). Occasionally, an entire sequence of undated late Pleistocene strandlines can be correlated with the paleosea-level curve by trial and error if different, but constant, uplift rates are assumed; the best graphical correlation of strandline elevations and paleosea-level highstands simultaneously yields the most reasonable uplift rate and strandline ages (see Figure 6.6). In effect, the uplift rate itself is a useful dating tool, even if not independently derived. However, in some complexly deformed coastal areas no strandlines can be dated confidently using this graphical technique (Weber, 1983), which suggests that the uplift rate varied with time. Generally, for strandlines dated between 700 ka and 200 ka BP by either absolute or relative techniques the uncertainty in correlation with the Pleistocene paleosea-level curve is at least one major glacioeustatic cycle (±100 ka), and for strandline dated between 120 ka and 30 ka the uncertainty is commonly one minor cycle (±20 ka) (Harmon et al., 1979; Stearns, 1984). The best-preserved and most reliably dated Pleistocene strandlines along most emergent coastlines correlate with the 120-ka, 104-ka, and 82-ka sea-level highstands (see Figure 6.11B below). These three highstands correlate with deep-sea oxygen-isotope substages 5e, 5c, and 5a, respectively (Shackleton and Opdyke, 1973; Chappell, 1983), and their strandlines are often referred to by these alphanumeric designations. Strandlines that correlate with younger (lower) highstands appear in the emergent geologic record only where uplift rates are sufficiently high (>0.3–1.0 m/ka) to elevate them above present sea level (Figure 6.6). Along most erosional coastlines some emergent Pleistocene strandlines are missing or are laterally discontinuous owing to sea-cliff retreat during the formation of younger strandlines (see Figures 6.17B and 6.20A below). Generally, strandlines older than 400 ka are poorly preserved owing to subaerial slope degradation and stream incision; exceptions are found on stable or slowly rising coastlines where strandlines as old as 2.4 m.y. are commonly preserved

COASTAL TECTONICS 102 (Ward, 1985). In areas of rapid uplift (greater than 4 m/ka), subaerial erosional processes are greatly accelerated and strandlines older than 120 ka are rarely preserved. However, in areas of rapid uplift the fine structure of late Pleistocene sea-level history is most clearly and completely recorded by strandlines younger than 120 ka (Chappell, 1983). Holocene Strandlines Radiocarbon dates on fossil wood, peat, and shell from sedimentary deposits on continental shelves and in shallow coastal embayments throughout the world indicate that sea level stood about 100–150 m below its present position during the last glacial maximum between 20 ka and 15 ka BP, and then rose rapidly to within 4–6 m of its present position by about 7–5 ka BP as the major ice sheets retreated at the end of Pleistocene time (Bloom, 1977). Although sea-level changes over the past 5 ka may have varied regionally owing to minor geoidal distortion and hydro-isostatic crustal loading (Clark et al., 1978), the most reliable data indicate that sea level has not fluctuated significantly (±2 m), nor has it been higher than its present position, over this period of time (Figure 6.7) (Thom et al., 1969; Bloom, 1970; Scholl et al., 1970; Omoto, 1979; Faure et al., 1980; Thom and Roy, 1983; Sneh and Klein, 1984; Gibb, in press). Consequently, Holocene strandlines, unlike Pleistocene strandlines, do not represent eustatic sea-level fluctuations but rather episodic crustal movements or major storm events. However, the highest strandlines in most Holocene strandline sequences usually represent the sudden and drastic decrease in postglacial sea-level rise between 7 ka and 5 ka BP (Figure 6.7). Along erosional coastlines emergent Holocene strandlines form and survive only where crustal uplift is sufficiently rapid (>1–2 m/ka) to offset the destructive effects of subsequent wave erosion. However, along prograding coastlines depositional strandlines survive even where there is little or no crustal uplift (Figure 6.3B) (Schofield, 1973, Garrick, 1979). Most emergent Holocene strandlines of both erosional and depositional origin are similar in configuration to their Pleistocene counterparts but usually are smaller in scale, finer in geomorphic expression, and better preserved (Figure 6.3). Also, because of their relatively youthful age, most Holocene strandlines lie within tens rather than hundreds of meters above present sea level, even along the most rapidly uplifting coastlines. In protected embayments and near the mouths of large streams and rivers, Holocene strandlines commonly consist of bouldery to sandy beach ridges that are relatively imprecise paleosea-level indicators (Figure 6.3B) (Schofield, 1973; Garrick, 1979). However, along many wave-resistant coastlines with small tidal ranges (less than 1 m) emergent Holocene strandlines consist of horizontal solution notches and faint waterlines on rocks and sea cliffs that are extremely precise paleosea-level indicators (Machida et al., 1976; Pirazzoli et al., 1982). Along some recently uplifted coastlines, strandlines consist of horizontal bands of fossil intertidal sessile organisms such as barnacles, which are also precise paleosea-level indicators (Sawamura, 1953) even in areas with large tidal ranges (Plafker, 1965). In the historical record, emergent Holocene strandlines most commonly consist of points on tide-gauge records (Figures 6.4 and 6.9B) (Hicks and Crosby, 1974; Berrino et al., in press), but in a few areas they consist of waterlines and burrows of intertidal organisms on recently uplifted or down-dropped man-made coastal structures (see Figure 6.9A below) (Grant, 1970; Flemming, 1972; Berrino et al., in press). Along many rapidly uplifting coastlines, prehistoric Holocene strandlines are similar to strandlines produced by abrupt vertical crustal movements associated with major historical earthquakes (Sugimura and Naruse, 1954; Plafker, 1965; Wellman, 1969; Matsuda et al., 1978). This similarity suggests that all emergent Holocene strandlines are of coseismic origin. However, along stable or steadily uplifting coastlines, both erosional and depositional strandlines probably form during infrequent major storms (Schofield, 1973; Garrick 1979; Hillaire-Marcel, 1980). A major uncertainty in deriving a history of paleoseismicity from a sequence of emergent Holocene strandlines is that coseismic and storm-produced strandlines may be indistinguishable. Consequently, if coseismic and storm-produced strandlines occur together (Sandweiss and Rollins, 1981; Lajoie et al., 1982a), the number of past earthquakes could be overestimated from the geologic record. On the other hand, wave erosion may remove coseismic strandlines from the geologic record (see Figure 6.27B below) (Plafker et al., 1981), in which case the number of earthquakes represented by a sequence of emergent Holocene strandlines could be underestimated. Because of minor regional or subregional geoidal distortions, local hydro-isostatic adjustments, variable oceanographic conditions (currents, water temperature, salinity), and climatic fluctuations, there is probably no locally derived late Holocene sea-level curve that can be used as a precise universal tectonic datum for measuring minor (less than 2 m) vertical crustal movements (Clark et al., 1978; Newman et al., 1978). At best, a locally derived curve is a regional tectonic datum. However, data from coastlines throughout the world indicate that relative sea-level changes greater than 2–4 m over the past 6 ka are either apparent (caused by tectonic and isostatic crustal movements or sediment

COASTAL TECTONICS 103 compaction) or spurious (caused by reworking or contamination of dated fossil materials), not real. Therefore, present sea- level or a generalized sea-level curve (Figure 6.7) is often used as an approximate tectonic datum for measuring late Holocene crustal movements, especially where rates of deformation are very high (see Figures 6.25, 6.26, and 6.27 below) (Matsuda et al., 1978; Plafker and Rubin, 1978; Plafker et al., 1981). Most Holocene strandlines in the geologic record are dated directly or indirectly by radiocarbon techniques. However, the sudden and drastic decrease in the sea-level rise between 7 ka and 5 ka BP occasionally leads to incorrect or misleading radiocarbon dates on fossil materials (particularly marine shells) from emergent Holocene strandlines. This potential problem exists because all but the highest strandline in a sequence of emergent Holocene strandlines represent two different paleoshorelines—a transgressive shoreline older than about 6 ka and a regressive shoreline younger than 6 ka (Figure 6.7). Consequently, marine shells deposited on the older, transgressive shoreline could be reworked into sediments deposited on the younger, regressive shoreline. On the other hand, the two-part configuration of the Holocene sea-level curve provides an independent means of approximating the ages of emergent Holocene strandlines and deriving minimum uplift rates. On a plot of strandline elevation versus sea-level history (Figure 6.7), the straight line connecting the elevation of the highest strandline on the vertical axis and the tangent point at the break on the sea-level curve represents the average uplift since that strandline formed. The coordinates of the tangent point are the tentative age and original elevation of the highest strandline, and the slope of the line is the tentative average uplift rate. If constant uplift is assumed, interpolation of the uplift rate yields age estimates for any lower strandlines (see Figure 6.28 below) (Wellman, 1969). Because the highest Holocene strandline on most coastlines represents the break in sea-level rise, its age is usually between 7 ka and 5 ka, depending on the uplift rate; because the break is rounded, the age of the highest strandline increases from 5 ka to 7 ka as the uplift rate increases. However, if subsequent cliff erosion destroyed the 7–5-ka strandline, the highest surviving strandline would be somewhat younger. Obviously, other dating techniques are needed to test for this possibility. STRANDLINE DISPLACEMENT AND DEFORMATION Regional patterns of vertical strandline displacement (uplift or subsidence) and deformation (folding and faulting) commonly reflect primary tectonic processes (subduction and rifting) related directly to horizontal plate motions. Important exceptions are found in the northernmost continental areas of North America and Europe where regional patterns of latest Pleistocene and Holocene strandline displacement reflect major but transitory isostatic responses of the crust and mantle to geologically recent glacial unloading. Local patterns of strandline displacement and deformation generally involve minor tectonic structures (folds, faults, horsts, and grabens) that reflect secondary tectonic processes related to subregional stress patterns. Available data indicate that most secondary and some primary tectonic deformation recorded by marine strandlines extends no farther than a few tens of kilometers inland from the coastline. In many areas, vertical ground displacements on a local scale also reflect nontectonic processes such as volcanic tumescence (Kaizuka et al., 1983; Berrino et al., in press), isostatic adjustments due to sediment accumulation (Fisk and McFarlan, 1955), and sediment compaction due to natural processes (Atwater et al., 1977) or fluid extraction (Poland, 1971; Buchanan-Banks et al., 1975). In many cases, especially those involving minor vertical displacements, it is difficult, if not impossible, to distinguish tectonic from nontectonic apparent sea-level changes. It is noteworthy that on both local and regional scales sequences of marine (and lacustrine) strandlines constitute the longest and most detailed and areally extensive records of late Quaternary crustal deformation. Indeed, the magnitudes and rates of tectonic, isostatic, and volcanic processes derived from deformed and displaced strandlines are commonly used as references for comparing and interpreting tectonic data from less complete or shorter records in other geological environments. Vertical Displacement Because marine strandlines define horizontal lines and planes (where curved or closed), they record vertical crustal displacements (uplift or subsidence) more clearly than horizontal displacements. Net vertical displacement (D) is the difference between the present elevation (E) and the original elevation (e) of a strandline and is also the product of the strandline's age (A) and the average displacement rate (R) (Figure 6.5): (6.4) The average displacement rate (R) is the displacement (D; Eíe) of a strandline divided by its age (A): (6.5) An isobase is a locus of points on a planar tectonic marker (a marine platform or a plane defined by a curved or closed strandline) along which the vertical

COASTAL TECTONICS 104 displacement rate (R) is equal (see Figures 6.12, 6.15A, 6.16A, 6.18, and 6.21 below). The history of vertical crustal displacements at a coastal locality is derived graphically by plotting the displacement (D) of each independently dated strandline as a function of its age (A) (Figure 6.5). The resultant locus of points is the apparent sea-level history, which, as stated previously, is the inverse of the displacement history. If this locus of points defines a straight line, the displacement rate (R), which is the inverse of the slope of the line, was constant. If it does not define a straight line, the displacement rate was variable (see Figures 6.9A and 6.10B below). Commonly, only one strandline in a sequence can be dated independently, and, therefore, only an average displacement rate can be derived from a relative sea-level record. A plot of strandline elevation (E) as a function of age (A) yields the relative sea-level history, which is the inverse of the approximate displacement history where vertical displacements are large compared to sea-level changes. Frequently, uplift histories based on highly emergent Holocene strandlines (see Figures 6.25, 6.26, and 6.27 below) or very old Pleistocene strandlines (Bender et al., 1979) are approximated by relative sea-level curves. In other words, present sea level can be used as an approximate datum. The presence of emergent marine strandlines along most active-tectonic coastlines suggests that crustal uplift is more common than subsidence. However, subsidence is probably underestimated merely because most geologic evidence for downward crustal movements is buried in sedimentary basins (Atwater et al., 1977) or is submerged offshore (Lewis, 1974). Many studies of long-term crustal movements in coastal areas focus on uplift mainly because the emergent strandline record is better exposed and easier to interpret than the submergent record. Most long-term crustal movements recorded by marine strandlines reflect sustained tectonic deformation along active plate boundaries, but the most rapid known rates of vertical crustal displacement reflect transitory volcanic tumescence and glacio-isostatic rebound. Both processes are noteworthy primarily for comparative purposes, but also because they are, themselves, related to tectonic processes in various ways. Volcanic Displacements The highest known rates of sustained vertical ground displacement exceed 100 mm/ yr and are recorded by marine strandlines on the flanks of active insular and coastal volcanoes. For example, on the island of Iwo Jima, the tip of a large volcano in the western Pacific Ocean, radiometrically and historically dated emergent strandlines yield an average uplift rate of 200 mm/yr over the past 800 yr (Figure 6.8) (Kaizuka et al., 1983). However, tide-gauge data suggest that uplift rates on the island probably fluctuated between 100 and 800 mm/yr over this period of time (Figure 6.8B), during which only minor phreatic eruptions are known to have occurred (Corwin and Foster, 1959). FIGURE 6.8 Recent volcanic uplift of Iwo Jima, a volcanic island 1200 km south of Tokyo, Japan. A, Longitudinal strandline profiles. The strandline labeled 1779 was the active shoreline mapped in 1779 by the crew of the English explorer, Captain Cook. The age and maximum elevation of this strandline yield an average uplift rate of 200 mm/yr, which is similar to the rate of 170–240 mm/yr derived from radiocarbon dates of 0.5–0.7 ka on the 110-m strandline. The well-defined strandlines were formed by wave action during brief pauses in uplift or during periodic storms. B, Average uplift rates derived from 110-m (0.5–0.7 ka) and 40-m (0.2 ka) strandlines. Tide-gauge data record variable uplift that reached 800 mm/yr over the past 80 yr. Both modified from Kaizuka et al. (1983). An even longer record of vertical ground displacements related to volcanic processes is found in the Phlegraean Fields caldera on the Mediterranean coast near Naples, Italy, where historically dated strandlines on man-made structures document alternating subsidence and uplift that averaged 12 mm/yr over the past 2 ka (Figure 6.9) (Berrino et al., in press). However, uplift and subsidence exceeded 150 mm/yr for a few decades before and after a minor eruption in A.D. 1538. Tide-

COASTAL TECTONICS 105 gauge records at Pozzuoli, a coastal suburb of Naples, yield uplift rates of 350 and 500 mm/yr for two brief periods between 1970 and 1983 (Figure 6.9B), which suggests that another eruption is imminent. However, the data from Iwo Jima indicate that rapid uplift and even short surges of extremely rapid uplift are not by themselves definite indicators of imminent eruption. Glacio-Isostatic Displacements Highly emergent latest Pleistocene and Holocene marine (and lacustrine) strandlines in the deglaciated areas of northern North America and Europe record the highest known rates of regional vertical crustal displacement (Figure 6.10). Around Hudson Bay in Canada (Farrand and Gajda, 1962; Andrews, 1970; Peltier and Andrews, 1983) and the Gulf of Bothnia in Fennoscandia (Lundqvist, 1965; Donner, 1965, 1980), which were the regions of thickest ice accumulation during the last major glacial advance, rates of postglacial isostatic rebound reached 100 m/ka (100 mm/yr) between 13 ka and 7 ka BP and decreased quasi-exponentially to the present rates of 6–9 mm/yr (Barnett, 1966; Balling, 1980). Even these rates of residual glacio-isostatic rebound are extremely high compared with most rates of vertical tectonic displacement. Along the margins of these deglaciated areas, Holocene strandlines and tide-gauge data record complex histories of rapid subsidence as well as uplift, which probably reflect the collapse and inward migration of a crustal forebulge peripheral to the retreating ice front (Grant, 1980; Morner, 1980; Clague et al., 1982). FIGURE 6.10 Postglacial isostatic rebound curves derived FIGURE 6.9 Recent volcanic displacements at Pozzuoli, a from dated marine and lacustrine strandlines. coastal suburb of Naples, Italy, in the Phlegraean Fields A, Relative sea-level changes (minimum apparent sea- caldera. level changes), Gulf of Bothnia, Scandinavia. Modified A, Vertical displacement based on historical data such as from Lundqvist (1965). water marks on Roman buildings. Mount Nuovo, a small B, Apparent sea-level changes, Hudson Bay, North cinder cone nearby, erupted in A.D. 1538, about 40 yr America. Modified from Farrand and Gajda (1962). after the uplift rate increased from 10 mm/yr to about 100 In both areas crustal movement is the inverse of the mm/yr. The general subsidence that followed this small curves. Regional crustal uplift exceeded 100 m/ka shortly eruption has now reversed. after deglaciation in both areas. B, Tide-gauge data at Pozzuoli Harbor from 1970 to 1983. During two brief periods (1970–1973 and 1982–1983) uplift reached 800 mm/ yr. Because of this rapid uplift, increased fumerole activity, and ground cracking, parts of Pozzuoli have been evacuated in anticipation of another eruption. Modified from Berrino et al. (in press).

COASTAL TECTONICS 106 In some deglaciated areas rapid vertical crustal movements recorded by strandlines probably reflect tectonic as well as isostatic processes (Grant, 1970; Mathews et al., 1970; Clague et al., 1982; Anderson et al., 1984), and in a few areas the two processes apparently interact. For example, in western Scotland successively older emergent Holocene strandlines are offset progressively greater amounts across a series of local vertical faults (Sissons and Cornish, 1982), which indicates that the rate of glacio-isostatic rebound differed on either side of each fault. Also, in both North America and Europe, concentrations of recent seismic activity are closely associated with residual glacio-isostatic adjustments (Oliver et al., 1970; Morner, 1978, 1980; Thompson et al., 1983; Anderson et al., 1984). Apparently, even residual isostatic movements trigger earthquakes along zones of crustal weakness, even in areas of relative tectonic stability. Rapid glacio-isostatic displacements are also important in a general tectonic context because they yield valuable information on the strength of the Earth's crust and the visco-elastic properties of the underlying mantle (Walcott, 1970; Peltier and Andrews, 1983). Tectonic Displacements In marked contrast to the extremely rapid rates of volcanic and glacio-isostatic crustal displacement (>100 m/ka), the highest known rates of sustained vertical tectonic displacement recorded by marine strandlines are only 4 to 10 m/ka (Figure 6.11). Furthermore, these rapid displacement rates are recorded only by Holocene and latest Pleistocene strandlines, which suggests that they are of fairly limited duration. Also, virtually all rapid tectonic displacements recorded by marine strandlines occur along the axes of youthful anticlines above thrust-type faults in compressional tectonic regimes. Consequently, rapid vertical displacements of tectonic origin mainly reflect secondary deformation and are of limited importance in studies of regional crustal movements. However, because these rapid displacements are usually produced by episodic coseismic uplift events that are sufficiently large (1–15 m) to be expressed in the geologic record as emergent Holocene strandlines, they are extremely important in studies of paleoseismicity (see Figures 6.25–6.28 below). The episodic nature of rapid displacements recorded by marine strandlines suggests that maybe all tectonic deformation occurs in discrete increments. If this model is correct, these increments are too small or too infrequent at low deformation rates to be clearly expressed or preserved in the long-term geologic record. Data from emergent Pleistocene strandlines throughout the world indicate that most rates of long-term tectonic uplift are less than 2 m/ka (Figure 6.11B), even along rapidly converging plate boundaries. Evidently, FIGURE 6.11 Rates of vertical crustal displacement (uplift only). A, Holocene: IJ, Iwo Jima, volcanic uplift included for comparative purposes (Kaizuka et al., 1983); CR, Crete (Angelier, 1979); Z, Zagros (Vita-Finzi, 1979); B, Bahrein (Ridley and Seeley, 1979); MI, Middleton Island (Plafker and Rubin, 1978); IC, Icy Cape (Plafker et al., 1981); M, Makran (Page et al., 1979); T, Turakirae (Wellman, 1969); BP, Boso Peninsula (Matsuda et al., 1978; MF, Miller Flat; OH, Ocean House; V, Ventura; (Lajoie et al., 1982a, 1983); O, Osado, K, Kosado (Tamura, 1979); MP, Muroto Peninsula (Kanaya, 1978); KJ, Kikai Jima (Nakata et al., 1978). B, Pleistocene: NZ, New Zealand (Wellman, 1979); V, Ventura (Lajoie et al., 1982b); A, Arauco (Kaizuka et al., 1973); HP, Huon Peninsula (Chappell, 1983); W, Wairarapa (Gahni, 1978); KJ, Kikai Jima (Konishi et al., 1974); MV, Muroto (Yoshikawa et al., 1964); O, Osado; K, Kosado (Tamura, 1979); HMB, Half Moon Bay (Lajoie et al., 1979b); SC, Santa Cruz (Hanks et al., 1984); RI, Royalty Islands (Marshall and Launay, 1978); TA, Taaanaki (Pillans, 1983); T, Timor; AT, Atauro (Chappell and Veeh, 1978); H, Haiti (Dodge et al., 1983); B, Barbados (Bender et al., 1979); AH, Arroyo Hondo; CO, Cojo; Al, Algeria; IV, Isla Vista; SM, Santa Monica (Lajoie and Sarna-Wojcicki, 1982); PL, Point Loma (Ku and Kern, 1974); J, Jamaica (Moore and Samayajulu, 1974); BA, Bahamas (Neumann and Moore, 1975); Y, Yucatan (Szabo et al., 1978); BE, Bermuda (Harmon et al., 1979, 1981); Q, Queensland (Ward, 1985).

COASTAL TECTONICS 107 in most areas isostatic adjustments are sufficiently rapid to compensate for (and thus prevent) higher rates of vertical crustal movement on a regional scale. Although most short-term, vertical tectonic displacements recorded by Holocene marine strandlines are episodic or otherwise variable (see Figures 6.25–6.28 below) (Matsuda et al., 1978; Plafker and Rubin, 1978; Plafker et al., 1981), most average, long-term displacements recorded by Pleistocene strandlines appear to be relatively constant, at least over the past 100–500 ka (Bloom et al., 1974; Konishi et al., 1974; Moore and Samayajulu, 1974; Chappell and Veeh, 1978; Bender et al., 1979; Harmon et al., 1981; Dodge et al., 1983; Chappell, 1983; Hanks et al., 1984). Along some slowly uplifting coastlines, strandline data indicate that vertical displacement was fairly constant over the past 2–2.5 m.y. (Ward, 1985). Constant uplift can be demonstrated by comparing actual strandline elevations with those predicted by assuming constant uplift and using the New Guinea sea-level curve as a tectonic datum. Close agreement supports both the assumption of constant uplift and, of course, the assumption that the sea-level curve was a common tectonic datum. Along a few coastlines, strandline data indicate that long-term tectonic uplift was clearly variable. For example, in the Christchurch area on the island of Barbados radiometrically dated Pleistocene strandlines yield an average uplift rate of 0.5 m/ka between 300 ka and 200 ka BP, and a lower rate of 0.3 m/ka after 200 ka BP (Bender et al., 1979). Interestingly, in the nearby Saint George's Valley area uplift was relatively constant and averaged 0.3 m/ka over the past 640 ka. Where data from marine strandlines and other tectonic markers are sufficiently abundant, regional compilations of vertical crustal movements provide valuable insights into both local and regional tectonic processes. Regional data are most conveniently and clearly expressed planimetrically by isobases (Figure 6.12) (Research Group for Quaternary Tectonics Map of Japan, 1969; Dambara, 1971; Wellman, 1979). Not surprisingly, most long-term regional isobases closely mimic general topographic contours, which simply means that the highest uplift rates occur in mountainous areas and the lowest rates or subsidence occur in low-lying regions. Usually, short-term isobases derived from historical information (tide-gauge and geodetic data) agree with longer-term isobases derived from geologic information, but locally short-term and long-term displacement rates differ drastically or the sense of displacement is reversed. For example, along the coastlines of northern California, Oregon, and Washington, net (long-term) displacement rates derived from Pleistocene strandlines are very low (0–0.5 m/ka; derived from dates given in Kennedy et al., 1982), but short-term rates derived from tide-gauge data locally reach 3 mm/yr (3 m/ ka) (Hicks and Crosby, 1974). Some differences and reversals between short- term and long-term displacement rates probably represent pre-earthquake strain accumulation (Matsuda, 1976) or postearthquake crustal relaxation and, therefore, are important in determining earthquake potential (Thatcher, 1984). On the Muroto Peninsula of Shikoku, Japan, the 120-ka and 6-ka strandlines yield similar long-term uplift rates of 1.7 and 2.0 ka, respectively (Yoshikawa et al., 1974) (see Figure 6.16A below), whereas episodic uplift associated with earthquakes over the past 300 yr averaged 12.5 mm/yr (see Figure 6.26A below). Geodetic and tide-gauge data show that the peninsula actually subsided between these coseismic uplift events (see Figure 6.16B below). FIGURE 6.12 Isobase map of South Island, New Zealand, summarizing uplift data from marine strandlines and other vertically displaced geomorphic features. An isobase is a locus of points along which the uplift rate was equal. Isobases of South Island generally mimic topographic contours. Modified from Wellman (1979). Tilt Crustal tilt (T) is the differential vertical displacement (DIíDII) of a horizontal tectonic marker (strandline or platform) divided by the horizontal distance (d) between any two observation points (I and II)

COASTAL TECTONICS 108 along a coastline. The resultant dimensionless ratio is, of course, the tangent of the tilt angle (ș): (6.6) The tilt rate (R) is the tilt (T) divided by the age (A) of the marker: (6.7) Even minor crustal tilt is clearly expressed in the longitudinal and cross-sectional profiles of marine strandlines. If the profiles of two or more tilted strandlines are parallel (Figure 6.13A), the differential vertical displacement took place at the time or after the youngest strandline formed. If the profiles converge (Figures 6.13B and 6.13C), which is the most common case, the differential displacement was progressive and took place over the period of time during which the strandlines formed. If the tilt rate is constant, the geometric relationships of converging strandline profiles (Figures 6.13B and 6.13C) provide a useful means of correlating Pleistocene strandlines from widely separated areas, even where no independent age data are available. The algebraic expression relating the present elevation (E1) of one strandline to the present elevation (E2) of a second strandline at the same coastal locality is (6.8) where the first constant (A 1/A2) is the ratio of the strandline ages, and the second constant [(A1/A2) e2í e1] is the difference between the original strandline elevations (e1 and e2). This equation defines a straight line generated by graphically plotting the elevation of the first strandline (E1) as a function of the elevation of the second (E 2) at two or more localities along a tilted coastline or from different coastlines with different uplift rates (Figure 6.14). The first constant (A1/A2) is the slope of the line, and the second constant [(A1/A 2) e2íe1] is its intercept on the vertical axis. The numerical values of these two constants are unique for any pair of Pleistocene strandlines and remain fixed regardless of the differential uplift rates. Therefore, the graphically derived values for these constants can be used to correlate undated strandline pairs from widely separated areas, such as from one island to another (Figure 6.14) (Miyoshi, 1983). These graphical correlations are possible only if the strandline pairs had common datums and were synchronous, and furthermore, only if the tilt rate at each locality was constant (but not necessarily equal). Consequently, Holocene strandlines cannot be correlated using this graphical technique because they are not necessarily synchronous along different coastlines. Along straight coastlines, emergent strandlines are FIGURE 6.13 Tilted longitudinal profiles of emergent marine strandlines. A, Parallel Holocene strandlines on Ogi Peninsula of Sado Island, Japan. The lowest strandline was formed by differential uplift associated with a large earthquake in 1802. The 6-ka strandline is parallel to this historical strandline, which indicates that no other tilt event has occurred in the past 6 ka. The 2 m separating these strandlines probably represents either gradual or abrupt uplift after 6 ka BP but before 1802. See Figure 6.24 for location. From data in Ota et al. (1976). B, Converging Pleistocene strandlines on the Huon Peninsula, Papua New Guinea (see Figure 6.2B for cross-sectional profile). These converging strandlines record continual tilt over the past 120 ka. Ages are U-series dates on fossil corals. Minor irregularities in vertical spacing represent local variations in tilt rate. Modified from Chappell (1974b). C, Converging Pleistocene strandlines near Point Año Nuevo on the central California coastline. Modified from Lajoie et al. (1979b).

COASTAL TECTONICS 109 linear tectonic markers that record components of crustal tilt parallel to the shoreline. One of the best examples of tilt along a straight coastline is found on the Huon Peninsula of Papua New Guinea where the converging longitudinal profiles of six coral-reef strandlines record continual northwestward tilt of a large structural block over late Pleistocene time (Figure 6.13B) (Chappell, 1974b). There, as in other areas (Figure 6.13C) (Lajoie et al., 1979b), minor irregularities in the converging profiles of Pleistocene strandlines reflect slight variations in the patterns and rates of long-term differential uplift. FIGURE 6.14 Correlation of Pleistocene strandlines by graphical techniques. The elevation of one strandline (E1 ) is related to the elevation of a second strandline (E2) at the same locality by the general equation E1=(A1 /A2)E2í[(A1/A2) e2íe1]. The constants (A1/A2 ) and [(A1/A2) e2íe1] are unique for any pair of late Pleistocene strandlines and, therefore, can be used as a means of correlating strandline pairs from widely separated areas. The general equation (here simplified to Y=Ax+B) defines a straight line generated by plotting the elevation of one strandline (E1 ) as a function of the elevation a second (E2); the strandline pairs must be from localities with different but constant uplift rates, such as along a tilted coastline (Figure 6.13B), or on widely separated islands. A, A plot of elevations of dated strandlines (60 ka, 82 ka, 104 ka, and 120 ka) from different parts of the world illustrate the relationship. B, Undated strandline pairs (I and S; II and S) from throughout the Japanese archipelago. The constants for the two plots indicate that the S, I, and II strandlines correlate with the 120-ka, 80-ka, and 60-ka strandlines, respectively. Modified from Miyoshi (1983). Along straight, curved, and irregular coastlines, marine platforms are planar and virtually horizontal structural markers that record the actual direction of local tilt. In some areas progressively greater slopes of successively higher (older) marine platforms clearly record continual crustal tilt in both seaward and landward directions (Bradley and Griggs, 1976; Sarna- Wojcicki et al., 1976; Gahni, 1978; Pillans, 1983). Along irregular coastlines, curved or closed strandlines around small coastal embayments, peninsulas, or islands define broad planes that also record crustal tilt on a local, but slightly larger, scale (Figures 6.15 and 6.16A) (Ota, 1964, 1975; Tamura, 1979). A particularly instructive example of tilt is found on the Muroto Peninsula of southeast Shikoku, Japan, where structural contours on the planes defined by 120-ka and 6-ka strandlines yield similar rates of landward (northward) tilt of 6.1×10í5/ka and 5.5×10í5/ka, respectively (Figure 6.16) (Yoshikawa et al., 1964; Kanaya, 1978). Significant landward tilt of the peninsula accompanied the 1947 Nankai earthquake, which suggests that long-term tilt in that area is episodic, not continuous. Gradual short-term seaward tilt documented by geodetic and tide-gauge data prior to the 1947 seismic event (Figure 6.16B) (Yoshikawa et al., 1964) probably reflected pre-earthquake strain accumulation (1984). Synchronous strandlines on widely separated islands or on distant parts of deeply embayed coastlines define areally extensive planes that record crustal tilt on a regional scale. For example, two planes defined by the 120-ka and 6-ka strandlines on numerous islands of the Ryukyu archipelago in southern Japan may record general eastward tilt over an area of 100,000 km2 (Konishi et al., 1974; Ota and Hori, 1980). Folds Folds are merely compound tilts and, therefore, are similarly expressed by emergent marine strandlines in the coastal geologic record (Figures 6.17 and 6.18). In some areas folds expressed by deformed strandlines occur

COASTAL TECTONICS 110 FIGURE 6.16 Tilted strandlines on the Muroto Peninsula, Japan. A, Isobases on 6-ka strandline reveal short-term northward (landward) tilt of peninsula at the rate of 5.5×10í5/ka, which is similar to the long-term rate derived from the 120-ka strandline but much lower than the 5.3×10í4 /ka rate preduced by historical earthquakes. Note—data points occur only along coastlines; isobases interpolated across inland areas. B, Geodetic data show that significant landward tilt accompanied the 1947 Nankai earthquake. This coseismic FIGURE 6.15 Tilted strandlines on Sado Island, Japan. movement indicates that tilt in this area is episodic not A, Isobases defined by 120-ka and 6-ka strandlines reveal continuous. Gradual southward (seaward) tilt documented continuous southeastward tilt of two structural blocks by geodetic data prior to the 1947 earthquake probably (Osado and Kosado) that constitute the island. Note—data reflected pre-earthquake strain accumulation. See points occur only along coastlines; isobases interpolated Figure 6.24 for location. Modified from Yoshikawa et al. across inland areas. (1964) and Kanaya (1978). B, Longitudinal profiles of Pleistocene and Holocene strandlines also reveal continual tilt. The strandline data indicate that the tilt rates of the two blocks have been constant but significantly different. If the fault separating the two parts of the island is vertical, the 120-ka strandline is offset about 75 m and the 6-ka strandline is offset about 4 m, which yield similar average slip rates of 0.6 m/ka and 0.7 m/ka, respectively. See Figure 6.24 for location. Modified from Tamura (1979). above and closely mimic tighter bedrock structures, which indicates that strandline deformation represents the most recent increment of long-term crustal movement (Figures 6.17A and 6.18B) (Wellman, 1971a,b; Gahni, 1978; Lajoie et al., 1982b). In many areas, however, folded strandlines are not associated with obvious bedrock structures and, therefore, are the only clear evidence for recent crustal folding (Figure 6.17B; also see Figure 6.21 below) (Plafker, 1972; Kaizuka et al., 1973; Lajoie et al., 1979a). In either case, progressively tighter folds in successively older strandlines reflect continual differential crustal movement (Figure 6.17). Along

COASTAL TECTONICS 111 some highly active coastlines coseismic strandlines are warped into progressively tighter folds (Figure 6.17A) (Wellman, 1971a,b; Kaizuka et al., 1973; Gahni, 1978), which suggests that tectonic folding, like uplift and tilt, is incremental, not continuous. FIGURE 6.17 Crustal folds recorded by strandlines and illustrated by profiles. A, Longitudinal profiles of Holocene strandlines on Wairarapa coast of North Island, New Zealand, reveal tight folds clearly expressed in underlying Tertiary rocks. Each strandline is probably of coseismic origin, which indicates that folding was incremental not continuous. Highest (6 ka) strandline dated by graphical techniques. Modified from Wellman (1971b). B, Longitudinal profiles of Pleistocene strandlines along the Arauco coast of Chile reveal a broad warp not expressed in bedrock units but generally expressed in the generalized skyline profile of the local mountain range. Age of lowest strandline estimated to be 120 ka. Modified from Kaizuka et al. (1973). C, Profiles of historical coseismic strandlines reveal broad crustal warps in the Gulf of Alaska (1964) and on the Chilean coast (1960). These profiles are not continuous but are defined by numerous points on highly irregular coastlines. See Figure 6.18A for isobase map of warp in Gulf of Alaska. Modified from Plafker (1972). On local to regional scales, vertical displacement data from deformed marine strandlines are conveniently summarized as structural contours and isobases that express folds planimetrically (Figure 6.18; also see Figure 6.21 below). Structural contours and standard isobases depict deformation of a single tectonic marker (Figure 6.18A; also see Figure 6.21A below), whereas integrated isobases depict deformation normalized from two or more tectonic markers of different ages (Figure 6.18B) (Gahni, 1978). Isobases on the surface defined by the 1964 strandline in the Gulf of Alaska (Figure 6.18A) reveal broad, gentle warps produced by coseismic crustal deformation over an area of 200,000 km2 (Plafker, 1965). These isobases indicate that during the great 1964 earthquake an area of 60,000 km2 was uplifted an average of 1–2 m and was tilted northward; maximum uplift was 3–10 m above local, north-dipping secondary faults (Figure 6.17C) (Plafker and Rubin, 1978). The isobases also show that an area of 110,000 km2 north of the uplifted zone subsided a maximum of 2 m during the earthquake. The pattern of coseismic crustal warping associated with the earthquake is similar to the general pattern of long-term deformation documented by older Holocene strandlines. However, drowned vegetation in the uplifted area south of the epicenteral region documents subsidence during the 1.4 ka prior to the earthquake (Plafker and Rubin, 1978). This subsidence was probably pre-earthquake strain accumulation above the interplate megathrust on which the slip occurred. The general pattern of coseismic uplift and subsidence that accompanied the 1964 earthquake in the Gulf of Alaska is similar to the pattern of crustal defor

COASTAL TECTONICS 112 mation that accompanied the great 1960 megathrust earthquake in coastal Chile (Figure 6.17C) (Plafker and Savage, 1970; Plafker, 1972). These two well-documented examples of coseismic crustal deformation suggest that broad warping (subsidence as well as uplift) in the overthrust plate is characteristic of great megathrust earthquakes. It is noteworthy that the subtle coseismic crustal deformation expressed by vertically displaced strandlines in both Alaska and Chile would not have been recorded in such great detail nor over such large areas in noncoastal regions, nor even along coastlines without deep embayments and offshore islands. An interesting contrast to the broad, regional warps depicted by isobases on the coseismic strandline in the Gulf of Alaska is found in the tight, local folds and related faults depicted by integrated isobases on Pleistocene and Holocene strandlines along the Wairarapa Coast of North Island, New Zealand (Figure 6.18B). The former reflect primary crustal deformation near the leading edge of a major overthrust plate (Plafker, 1972), and the latter reflect intense secondary deformation due to rapid crustal shortening within an accretionary prism in a similar tectonic setting (Gahni, 1978; Cole and Lewis, 1981). An interesting aspect of crustal folds, which commonly is ignored or misunderstood in tectonic studies, is that the rate of vertical displacement along the axis of a fold and the rate of tilt along its limbs both decrease as the fold matures. If this apparent decrease in crustal deformation is documented by successively older strandlines, it could be misinterpreted as a reduction in local or regional tectonic strain. Vertical Fault Movement Because marine platforms and strandlines are virtually horizontal structural markers, they record vertical fault movement more clearly than lateral movement. Often, vertical discontinuities in cross-sectional profiles of platforms (Figure 6.19) (Ota, 1975; Lajoie et al., 1979b; Dames and Moore Consultants, 1981; Sarna-Wojcicki et al., 1986) and in longitudinal profiles of strandlines (Figures 6.15B and 6.20) (Yoshikawa et al., 1964; Ota, 1975; Lajoie et al., 1979a,b; Tamura, 1979; Sissons and Cornish, 1982; Pillans, 1983) express even minor vertical fault movements, even where there is no other evidence of fault activity. Jogs in isobases and structural contours on marine platforms also express vertical fault movements and, where aligned, define local fault traces (Figures 6.12, 6.18B, and 6.21A) (Gahni, 1978; Lajoie et al., 1979b). However, not all vertical offsets of marine strandlines and platforms doc FIGURE 6.18 Crustal folds recorded by strandlines and illustrated by isobases. An isobase is a locus of points along which the rate of vertical crustal movement is equal. A, 1964 strandline records subtle crustal warps associated with large megathrust earthquake in Gulf of Alaska. See Figure 6.17C for profile of warped surface. Modified from Plafker (1972). B, Integrated isobases on 100-ka surface of the Wairarapa coast of North Island, New Zealand. Elevation data from 120-ka and 80-ka strandlines normalized to 100-ka surface. Complex folds and faults revealed by these isobases mimic structures in Tertiary bedrock. Intense folding is due to crustal shortening in sedimentary prism at leading edge of overthrust plate. Modified from Gahni (1978).

COASTAL TECTONICS 113 ument fault movement. Some, especially along steep coastal slopes, represent large, block-type slope failures. FIGURE 6.20 Vertical component of fault displacement recorded by offsets in longitudinal profiles of marine strandlines. FIGURE 6.19 Vertical component of fault displacement A, Progressively greater vertical offset of successively recorded by offsets in cross-sectional profiles of marine older marine strandlines documents continual fault activity strandlines. over at least the past 700 ka on the Taranaki coast of North A, Wave-cut platform and overlying marine (Qtm) and Island, New Zealand. Strandline ages (81–596 ka) derived alluvial (Qal) sediment offset about 1 m across a minor from tephrochronology, amino-acid data, and uplift rates. bedding-plane fault near Point Conception in southern Modified from Pillans (1983). California. Platform dated at 82 ka by graphical and B, Vertical offset of 40-ka strandline across Red Mountain amino-acid techniques. Overlying terrace cover may be as thrust fault (RMTF) near Ventura, California. Strandline young as latest Pleistocene or early Holocene. Modified age based on amino-acid data from fossil shells. Modified from Dames and Moore consultants (1981) and Lajoie and from Lajoie et al. (1982b). Sarna-Wojcicki (1982). B, 40-ka wave-cut platform and overlying marine sediment offset 45 m across Javon Canyon fault (JCF) near Ventura, California, yields a long-term displacement rate of 1.1 m/ka. A stream-cut platform graded to a 3.5-ka marine strandline is offset 4 m across the same fault, which yields a similar short-term displacement rate of 1.2 m/ka. See Figure 6.22A for coseismic slip events on this fault. Modified from Sarna-Wojcicki et al. (1986). Along many coastlines marine strandlines and platforms provide relative age control for recent fault activity. If a wave-cut strandline truncates a bedrock fault, the latest movement on the fault is, of course, older than the strandline. Conversely, if a strandline is offset across a fault, the latest movement is younger (Figures 6.15 through 6.21). Both the U.S Nuclear Regulatory Commission and the State of California Public Utilities Commission use relative strandline-fault relationships indirectly to define active faults in their assessment of seismic hazards to critical engineered structures in coastal regions. Both agencies consider a fault potentially active if it has moved in the past 100 ka. This somewhat arbitrary age is used specifically because the 120-ka to 82-ka strandlines are the most prominent and most accurately dated tectonic markers along most coastlines. Where dated, faulted strandlines also yield rates of fault movement. The offset of a strandline or platform divided by its age is, of course, the average slip rate since the strandline formed. In some areas, progressively greater offsets of successively older strandlines document continual fault activity and yield long-term average slip rates (Figures 6.15B, 6.19B, and 6.20) (Sissons and Cornish, 1982; Pillans, 1983; Sarna-Wojcicki et al., 1986). For example, on Sado Island off the west coast of Honshu, Japan, vertical discontinuities in the profiles of five Pleistocene and Holocene strandlines record continuous differential movement across the steeply dipping fault that separates the two mountainous parts of the

COASTAL TECTONICS 114 island (Figure 6.15) (Tamura, 1979). If this fault is vertical, the 120-ka strandline is offset about 75 m, which yields a long-term slip rate of 0.6 m/ka. The 6-ka strandline is offset about 4 m, which yields a similar short-term slip rate of 0.7 m/ka. Another excellent example of continuous fault slip recorded by progressively greater offset of successively older marine strandlines is found near the town of Ventura on the southern California coastline (Figure 6.19B) (Sarna-Wojcicki et al., 1986). There, the wave-cut platform of the 40-ka strandline (Kennedy et al., 1982) is vertically offset about 45 m across a high-angle reverse fault, which yields an average slip rate of 1.1 m/ka. Nearby, a stream-cut platform graded to an emergent 3.5-ka strandline is offset 4 m across the same fault, which yields a similar shorter-term slip rate of 1.2 m/ka. FIGURE 6.21 Closely related folding and faulting recorded by deformed 82-ka strandline at Half Moon Bay, California. Structural contours on wave-cut platform depict drag folds that plunge obliquely into and away from the vertical fault plane. Orientation of folds indicates predominant right-lateral fault movement, which is consistent with regional fault movements. Longitudinal profiles of strandline and faulted platform record gentle warps and vertical offset across fault. Shoreline angle of strandline does not intersect fault plane so local slip vector and slip rate are not known. Vertical separation of wave- cut platform across fault revealed by platform profiles on fault plane (dotted lines; E, east; W, west). Structural contours derived mainly from shallow borehole and FIGURE 6.22 Coseismic slip events on faults recorded by seismic refraction data. Modified from Lajoie et al. scarp-derived talus. (1979b). A, Stream-cut platform graded to 3.5-ka marine strandline offset 4 m across Javon Canyon fault near Ventura, California (see Figure 6.19B). At least five discrete slip events (0.5–1.4 m) are recorded by lenses of scarp-derived talus preserved below the upthrown block. If these slip events are coseismic, the average earthquake recurrence interval for this fault is about 700 yr. Modified from Sarna-Wojcicki et al. (1986). B, Vertical component of fault displacement recorded by offset of 104-ka wave-cut platform at Point Año Nuevo, California. At least nine discrete slip events are recorded by lenses of scarp-derived talus and faulted sediment preserved on the wave-cut platform beneath the upthrown block. If slip events are coseismic, the average earthquake recurrence interval is about 12 ka. It should be stressed, however, that average recurrence intervals have little meaning if total fault movement occurred during a short period of time (dotted line). Independently derived ages for each event are needed to resolve this potentially serious problem. Modified from Weber and Cotton (1981). If the size of an average coseismic slip event is assumed or is known from a historical seismic event, it can be divided into the slip rate to yield the average earthquake recurrence interval. However, where discrete slip events are recorded geologically, earthquake recurrence can be determined more directly. For example, lenses of scarp-derived Holocene talus preserved beneath the upthrown block of the high-angle reverse fault near Ventura, California (Figure 6.19B), record at least five discrete slip events during the past 3.5 ka (Figure 6.22A) (Sarna-Wojcicki et al., 1986). If these slip events were of coseismic origin, the average vertical displacement was 0.8 m and the average earthquake recurrence interval was 700 yr. An even longer history of discrete slip events is recorded where a 104-ka marine platform is offset across a strand of a complex fault zone at Point Año Nuevo about 100 km south of San Francisco on the central California coastline (Lajoie et al., 1979b). This wave-cut platform

COASTAL TECTONICS 115 is offset 5.5 m (dip-slip component), which yields an average long-term slip rate of 0.05 m/ka. More importantly, however, scarp-derived talus that accumulated on the marine platform below the periodically rejuvinated fault scarp records at least nine slip events over the past 104 ka (Figure 6.22B) (Weber and Cotton, 1981). If coseismic, these slip events, which averaged 0.6 m, yield an average earthquake recurrence interval of about 12 ka. It should be stressed, however, that if total fault movement was restricted to relatively short periods of time (Figure 6.22B), average recurrence intervals are of little use in predicting the time of the next event. Each event must be dated independently to make meaningful predictions of future events. Lateral Fault Movement As indicated previously, lateral fault movement is rarely recorded by marine strandlines, even where strandlines clearly cross active strike-slip faults. Usually, the amount of offset is ambiguous because of irregularities in or poor preservation of the paleoshoreline. For example, at Point Año Nuevo on the central California coastline, two Pleistocene strandlines that cross a major right-lateral fault system appear to be offset across several fault strands (Weber and Lajoie, 1977; Weber and Cotton, 1981). However, because these strandlines must be projected considerable distances (up to 1.0 km) at low angles across some of the fault strands, the individual and cumulative offsets of each strandline are too uncertain to yield meaningful rates of lateral fault movement. In some areas, the sense but not the amount of lateral fault movement is expressed by the orientation of drag folds recorded by marine strandlines. For example, the 82-ka strandline at Half Moon Bay about 60 km northwest of Point Año Nuevo on the central California coastline records broad crustal warps on both sides of a major fault strand within the coastal fault system (Figure 6.21) (Lajoie et al., 1979b; Kennedy et al., 1981). Structural contours on the warped wave-cut platform define a broad syncline that plunges obliquely into the vertical fault plane from the east and a broad anticline that plunges away from the fault plane to the west. These structural relationships suggest that the syncline and anticline are drag folds related to right-lateral fault movement, which is consistent with regional fault movements. Unfortunately, the strandline of this deformed marine platform does not intersect the fault plane and, therefore, neither the local slip vector nor the slip rate is known. However, where latest Pleistocene stream courses cross the warped marine terrace northeast of the fault they are deflected toward the axis of the syncline (Figure 6.21), which suggests that fault movement and related crustal warping were continual over the past 82 ka. COSEISMIC UPLIFT AND EARTHQUAKE RECURRENCE In several seismically active coastal areas historical coseismic uplift has produced conspicuous emergent marine strandlines 1–15 m above sea level. The best documented examples of historical coseismic strandlines are found in Japan (Sugimura and Naruse, 1954; Nakamura et al., 1965; Ota et al., 1976; Matsuda et al., 1978; Shimazaki and Nakata, 1980), New Zealand (Wellman, 1969; Stevens, 1973), Alaska (Plafker, 1965, 1972), Chile (Plafker and Rubin, 1967; Plafker and Savage, 1970; Plafker, 1972), and Iran (Page et al., 1979). The historical coseismic strandlines in these areas suggest that similar higher (older) Holocene strandlines in these and other active-tectonic areas are also of earthquake origin. If this interpretation is correct, a sequence of emergent Holocene strandlines is a physical record of past earthquakes from which it should be possible to predict the size and date of the next seismic event. If the average uplift rate is constant, coseismic uplift events may follow a time-predictable or displacement-predictable pattern (Figure 6.23) (Shimazaki and Nakata, 1980). In a time-predictable pattern the time between events is proportional to the size of the preceding event, and, therefore, the date, but not the size, of the next event can be predicted. In a displacement-predictable pattern the time between events is proportional to the size of the succeeding event, and the size of the next event can be predicted for any future date, but that date cannot be predicted. Of course, if the period and size of uplift events are regular, both can be predicted for the next event. In practice, however, strandline elevations and dates are usually too variable or imprecise to fit any pattern exactly, and, therefore, only average displacements and recurrence intervals can be derived from sequences of emergent Holocene strandlines. In some cases, the uplift rate is so variable that no reasonable predictions can be made. A few examples of coseismic uplift from Japan, Alaska, and New Zealand illustrate some of the possibilities and limitations in deriving detailed seismic histories from sequences of emergent Holocene strandlines. Japan The Japanese archipelago, which lies along the leading, overthrust margin of the Eurasian tectonic plate in the western Pacific Ocean, is one of the most seismically

COASTAL TECTONICS 116 active areas on Earth (Tarr, 1974). Great intraplate megathrust earthquakes occur along the major subduction zone that forms the eastern boundary of the archipelago, and smaller, but still potentially destructive, interplate earthquakes occur along its complexly faulted eastern boundary. Particularly good examples of coseismic strandlines are found on the Boso Peninsula, the Muroto Peninsula, and Kikai Island along the overthrust Pacific Coast and on Awashima Island and the Ogi Peninsula of Sado Island along the block-faulted Japan Sea Coast (Figure 6.24). FIGURE 6.24 Generalized tectonic map of Japan showing locations of emergent marine strandlines produced by historical coseismic uplift. O, Ogi Peninsula of Sado Island, 1802 earthquake (see Figure 6.13A for strandline profiles and Figures 6.15A and 6.24 for location); A, FIGURE 6.23 Theoretical patterns of coseismic uplift (or Awashima Island, 1964 Niigata earthquake; B, Boso fault movement). Peninsula, 1703 and 1923 Kanto earthquakes (see A, If average uplift rate is constant, the incremental Figure 6.25 for uplift history); M, Muroto Peninsula, 1707, coseismic uplift events may follow either a time- 1855, and 1947 Nankaido earthquakes (see Figure 6.26A predictable or a displacement-predictable pattern. In a for uplift history; see Figure 6.16 for strandline and time-predictable pattern the time between events (T) is geodetic data); K, Kikai Jima, no historical coseismic proportional to the size of the preceding displacement strandlines, but see Figure 6.26B for Holocene uplift (PD). In this case, the time, but not the size, of the next history. event can be predicted. In a displacement-predictable pattern the time between events is proportional to the size of the succeeding displacement (SD). In this case, the size of the next event can be predicted for any future date, which cannot be predicted. In practice, strandline elevations and ages are usually too uncertain to reveal precise patterns of uplift. Modified from Shimazaki and Nakata (1980). B, If a displaced marker (e.g., wave-cut platform) is dated but individual events are not dated independently, their intervals and age ranges can be estimated from the average displacement rate. However, if total displacement is highly irregular (dotted line), average displacement intervals, rates, and age values are meaningless. Independent ages of events are needed to establish displacement pattern. Four emergent Holocene strandlines (Numa I-IV) form prominent terraces around the southern tip of the Boso Peninsula about 80 km south of Tokyo (Figures 6.3A, 6.24, and 6.25) (Sugimura and Naruse, 1954; Matsuda et al., 1978). Isobases on the planes defined by these curved strandlines reveal progressive uplift and northward tilt of the southernmost part of the peninsula. Radiocarbon dates on fossil shells from the three highest strandlines (Numa I-III terraces) yield an average uplift rate of 3.9 m/ka (Figure 6.25A). The lowest major strandline (Numa IV or Genroku terrace) at 5 m and a minor strandline at 1.0 m were formed by uplift during the great Kanto earthquakes of 1703 and 1923, respectively. The ages and vertical distances between successively higher pairs of dated strandlines yield an average uplift per event of about 6 m and an average recurrence interval of about 1.5 ka. More importantly, however, the uplift and age data appear to follow a

COASTAL TECTONICS 117 time-predictable pattern. After being corrected for probable postearthquake subsidence (currently 1–2 mm/yr), the strandline data from the Boso Peninsula indicate that the next uplift event similar to the combined 1703 and 1923 events will occur in 0.8–1.3 ka (Figure 6.25B). Other workers (Nakata et al., 1980; and Shimazaki and Nakata, 1980; Yonekura and Shimazaki, 1980) also studied the emergent Holocene strandlines on the Boso Peninsula and reached similar conclusions. However, some of these workers (Yonekura and Shimazaki, 1980) suggested that the largest uplift events may have accompanied moderate-sized earthquakes on steep, secondary faults within the overthrust plate of the Sagami subduction zone and that the smallest uplift events may have accompanied larger earthquakes on the main but more gently dipping thrust fault. If this interpretation is correct, the 1703 Kanto earthquake may have been smaller than the 1923 event, even though the relative coseismic uplift would suggest otherwise. Interestingly, at Oiso, a densely populated suburb of Tokyo, about 70 km northwest of the Boso Peninsula, the sum of the uplifts that accompanied the 1703 and 1923 earthquakes is less than the uplift required to maintain the long-term uplift rate derived from older strandlines. This relationship suggests that the next uplift event in the Oiso area is overdue. Consequently, the Oiso area may currently be at greater seismic risk than the southern Boso Peninsula (Matsuda et al., 1978). FIGURE 6.26 Coseismic uplift on Muroto Peninsula and Kikai Jima, Japan. A, Historical coseismic strandlines at Murotsu Harbor. Uplift events appear to follow a time-predictable pattern and yield an average short-term uplift rate of 12.5 mm/yr, which is about six times higher than the long-term rate of 1.7–2.0 m/ka derived from 6-ka and 120-ka strandlines. If the short-term rate prevails, the next coseismic event should occur in the year 2040. If the long-term rate prevails, the next event (or sequence of closely spaced events) will not occur for another 3 ka. (See Figure 6.16 for more uplift data.) B, Radiocarbon-dated strandlines on Kikai Island, Japan, appear to follow a time-predictable pattern. If correct, a FIGURE 6.25 Coseismic uplift, Boso Peninsula, Japan. large coseismic uplift event should have occurred between A, Average uplift appears to be constant, and coseismic 1400 and 1600. There is no geologic or historical evidence uplift events appear to follow a time-predictable pattern. for such an event, so the area may be overdue for a large B, Uplift pattern indicates that the next event (or sequence earthquake. Modified from Shimazaki and Nakata (1980). of events) will occur in A.D. 3000. Interearthquake See Figure 6.24 for locations. subsidence (strain accumulation) from geodetic data is incorporated into prediction model. See Figure 6.24 for location and Figure 6.3A for profile of strandlines. Modified from Matsuda et al. (1978). At Murotsu harbor on the southern tip of the Muroto Peninsula on the island of Shikoku (Figure 6.24), the uplifts associated with the great Nankaido earthquakes of 1707 (1.8 m), 1855 (1.2 m), and 1947 (1.2 m) were determined by primitive but reasonably accurate harbor soundings and by a vertically displaced strandline defined by barnacles (Figures 6.16 and 6.26A) (Sawamura, 1953; Shimazaki and Nakata, 1980). These historical data yield a constant short-term uplift rate of 12.5 mm/yr, which is about six times higher than the

COASTAL TECTONICS 118 long-term rate of 1.7–2.0 m/ka derived from the 6-ka and 120-ka strandlines (Yoshikawa et al., 1964; Kanaya, 1978). If the exceptionally high uplift rate produced by historical coseismic events is maintained, the next large earthquake should occur in 2040 (Shimazaki and Nakata, 1980). However, if the long-term rate prevails, which seems more likely, the next event, or a series of closely spaced events, should not occur for another 3 ka. Tide-gauge and geodetic data from central Japan, which reveal the spatial and temporal patterns of interearthquake crustal deformation between the 1855 and 1947 events, indicate that because of the nonlinearity of strain buildup and the significant permanent deformation, simple recurrence calculations from strandline data may overestimate the true interval between major earthquakes by a factor of 2–3 in the Nankai area (Thatcher, 1984). On the other hand, the three closely spaced historical uplift events on the Muroto Peninsula and the two closely spaced events on the Boso Peninsula suggest that the large vertical spacing between emergent Holocene strandlines in these and possibly other areas may represent multiple, not single, earthquakes. If this interpretation is correct, earthquakes at a particular locality on a convergent plate boundary may occur in relatively tight clusters separated by considerable periods of time (1–3 ka), in which case the date of the next event would be difficult if not impossible to predict. On Kikai Island, the easternmost island in the Ryukyu archipelago of southern Japan, radiocarbon dates of fossil corals from four Holocene strandlines yield a uniform uplift rate of 1.8 m/ka over the past 6 ka and indicate that coseismic uplift followed a time-predictable pattern (Figures 6.24 and 6.26B) (Nakata et al., 1978, 1979; Shimazaki and Nakata, 1980). If the time-predictable model is correct, there should have been a large earthquake between A.D. 1400 and 1600. The lack of historical or geological evidence for an earthquake of this age suggests that Kikai Island, like the Oiso area near Tokyo, is overdue for a major earthquake (Shimazaki and Nakata, 1980). Two very different earthquake recurrence intervals are recorded by emergent Holocene strandlines on Sado and Awashima Islands in the intensely faulted area off the west coast of central Honshu, the largest Japanese island (Figure 6.24). Seven emergent Pleistocene and Holocene strandlines occur on the Ogi Peninsula, the southernmost tip of Sado Island (Figure 6.15A) (Ota et al., 1976). The lowest (2 m) strandline was formed by uplift during a major earthquake in 1802 (Figure 6.13A). This strandline and the next higher (4 m) 6-ka strandline are both tilted northward about 2.5×10í2, which indicates that the 1802 event was the only tilt event in the past 6 ka. The uniform difference of 2 m between these two strandlines is ascribed to a 2-m highstand of sea level at about 6 ka BP (Ota et al., 1976). A more likely explanation is that the area was uplifted uniformly after 6 ka BP but before the 1802 tilt event. In any case, dividing the coseismic tilt produced in 1802 into the long-term tilt rate derived from the 120-ka strandline yields a recurrence interval for seismic events of 5–9 ka, which is consistent with the >6-ka interval documented by the two lowest strandlines. At Awashima, which lies about 70 km northeast of Sado Island (Figure 6.24), tilted strandlines yield a much shorter recurrence interval (Nakamura et al., 1965). Awashima was uplifted a maximum of 1.8 m and tilted northward about 2.4×10í4 during the 1964 Niigata earthquake. The long-term tilt rate derived from the 82-ka strandline at 50–70 m above sea level is 4.3×10 í4/ka. Dividing this tilt rate by the 1964 tilt yields an average recurrence interval of about 1.5 ka. The different earthquake recurrence intervals for the faults near the Ogi Peninsula and Awashima Island demonstrate that similar faults within the same tectonic province can have very different displacement histories. Alaska The narrow zone comprising the Aleutian archipelago and the south coast of mainland Alaska lies along the overthrust margin of the North American tectonic plate in the north Pacific Ocean. At numerous localities in this seismically active area, great interplate megathrust earthquakes are recorded by emergent Holocene and historical strandlines. On Middleton Island (Figure 6.18A) emergent Holocene strandlines record six coseismic uplift events over the past 4.5 ka (Figure 6.27A) (Plafker and Rubin, 1967, 1978). The lowest strandline at 3.5 m above sea level was produced by uplift associated with the great 1964 earthquake in southern Alaska. Radiocarbon dates on peat and wood from the wave-cut platforms of the five highest strandlines indicate that the time interval between uplift events increased gradually from about 0.5 ka to 1.4 ka and that the uplift rate decreased from 14 m/ka to 5.6 m/ka. If the island subsided slightly between major earthquakes, the uplift that accompanied each earthquake was greater than the vertical spacing between strandlines. It is not clear if the uplift events, which ranged from 7–12 m and averaged about 9 m, followed a displacement-predictable or a time-predictable pattern, but it is obvious that the 3.5-m uplift associated with the 1964 earthquake was not sufficiently large to maintain even the diminishing uplift rate (Figure 6.27A). In effect, at least half of the stress accumulated since the formation of the second lowest strandline at

COASTAL TECTONICS 119 about 1.4 ka BP has yet to be released. Consequently, if the Holocene trend continues, coseismic uplift at least as large as the 1964 event should occur on Middleton Island sometime in the near future (Plafker and Rubin, 1978). FIGURE 6.27 Coseismic uplift, Gulf of Alaska. A, Radiocarbon-dated and historical coseismic strandlines on Middleton Island, Alaska. Average uplift rate is decreasing and interval between events is increasing while the size of each event has remained fairly constant. Average uplift is 0.95 m/ka. The 1964 coseismic uplift did not bring the island up to the long-term rate, which suggests that another event is due in the near future. Present sea level used as tectonic datum. Modified from Plafker and Rubin (1978). B, Radiocarbon-dated coseismic strandlines at Icy Cape (about 300 km east of Middleton Island on the Alaska mainland) yield a fairly constant uplift of about 1 m/ka. Uplift events appear to follow a time-predictable pattern. If this pattern continues, the next event is due in the near future. Modified from Plafker et al. (1981). See Figures 6.17C and 6.18A for 1964 coseismic strandline deformation. The focal region of the 1964 Alaskan earthquake lies north and west of Middleton Island (Figure 6.18A), and a seismic gap lies to the east. Consequently, the next uplift event to affect the island may accompany a megathrust earthquake in the area of the seismic gap, possibly on the submarine Pamplona Fault (Plafker and Rubin, 1978). This interpretation is supported by data from Icy Cape, a broad sedimentary headland about 300 km east of Middleton Island on the Alaska mainland and within the area of the seismic gap. There, depositional strandlines dated at 4.9 ka, 2.4 ka, and 1.3 ka BP by radiocarbon techniques record at least three uplift events that followed a time-predictable pattern over the past 5 ka (Figure 6.27B) (Plafker et al., 1981). The lowest (1.3 ka) strandline may have been formed during the same seismic event that produced the second lowest strandline dated at 1.35 ka on Middleton Island (Figure 6.27A). If the 1.2–1.3-ka time span between strandlines at Icy Cape is the recurrence interval for the area of the present seismic gap, the next uplift event is overdue. Presumably, this event would make up some or all of the apparent deficiency in uplift on Middleton Island. An interesting aspect of the strandline sequence at Icy Cape is that the vertical distance between the 4.9-ka and 2.4-ka strandlines is 27 m (Figure 6.27B), which seems too large for a single uplift event. If uplift averaged 10–15 m per event, a 14-m uplift at about 3.8 ka BP may not have been recorded or its strandline may have been destroyed by subsequent wave erosion (Plafker et al., 1981). New Ze aland The large southwest Pacific islands of New Zealand straddle the convergent boundary between the Australia-India tectonic plate to the west and the Pacific plate to the east. Northeast of the islands, the Australia-India plate is being thrust eastward over the Pacific plate, and southwest of the islands the relationship is reversed. The islands themselves are a manifestation of the uplift along the transition between the two opposing subduction zones on the same interplate boundary (Walcott, 1984). Within this complex structural area both interplate and intraplate earthquakes are common. At several localities, especially along the southeast coast of North Island, large earthquakes are recorded by emergent Holocene strandlines. At least five coseismic uplift events are recorded by six

COASTAL TECTONICS 120 bouldery beach ridges up to 27 m above sea level on the rapidly uplifting coastline at Turakirae Head, the southernmost tip of North Island (Figure 6.28) (Wellman, 1969; Stevens, 1973). The two lowest strandlines at 7 and 3 m above sea level, which were formed during historical earthquakes in 1460 and 1855, respectively, appear to record regional northwestward tilt across the entire southern tip of the island, whereas the four highest strandlines appear to record intense crustal warping between two local faults parallel to the offshore subduction zone (Wellman, 1969). However, all six strandlines may reflect regional tilt (Stevens, 1973). In either case, the five coseismic uplift events represented by these strandlines range from 2.5 to 9 m, and average 5.4 m (Figure 6.28). An assumption of constant uplift yields tentative ages for the four highest strandlines and also yields an average uplift rate of about 4 m/ka (see Figure 6.7) (Wellman, 1969). These graphically derived dates appear to indicate that uplift events follow a time-predictable pattern, which would indicate that the next event should occur in about 500 yr (Wellman, 1969). However, the pattern of uplift is inherent in the assumption of constant uplift—it is not independently demonstrated. Consequently, no firm estimate for the date of the next coseismic uplift event can be made. FIGURE 6.28 Coseismic Holocene strandlines at Turakirae Head, North Island, New Zealand. Highest (6 ka) strandline dated by graphical techniques (see Figures 6.5B and 6.7) assuming constant uplift rate (4 m/ka). Rate interpolated to estimate ages of lower strandlines. If uplift follows a time-predictable pattern, the next event should occur in 0.5 ka. It should be stressed, however, that these strandlines must be dated independently to demonstrate constant uplift and the time-predictable pattern. Modified from Wellman (1969). The often disparate and inconclusive estimates of past and future seismic events derived from sequences of emergent Holocene strandlines in the three areas discussed here (Japan, Alaska, and New Zealand) illustrate many of the difficulties encountered in interpreting even excellent historical and geological strandline data. However, in spite of these difficulties sequences of coseismic strandlines are among the most complete records of past earthquakes and in many coastal areas provide extremely valuable insights into the distribution of earthquakes in both space and time. FUTURE RESEARCH This brief review of coastal tectonics illustrates the breadth and scope of neotectonic deformation and seismic history that can be derived from the study of marine strandlines, which so conspicuously record the dynamic interaction between the fluctuating sea level and mobile tectonic plates along many of the world's active coastlines. Without sea-level changes, discrete strandlines would not be produced and the long-term record of Pleistocene crustal movements in coastal areas would be extremely difficult if not impossible to extract from a relative sea-level record. Conversely, without vertical crustal movements, a detailed history of sea-level fluctuations would be virtually impossible to reconstruct. As our understanding of this complex interaction increases we discover new problems that require the reinterpretation of existing data and the acquisition of new, more precise information. It should be stressed, however, that even at our present level of understanding the uncertainties inherent in extracting past crustal movements and seismic histories from relative sea-level records are far outweighed by the wealth of useful information obtained. Following is a list of some of the most pressing problems and needs that must be addressed if we are to progress in the new and rapidly evolving field of coastal tectonics. • Theoretical models relating sea-level changes and vertical crustal movements should be developed. General relationships are expressed by Eqs. (6.1) and (6.8) and Figures 6.5, 6.6, and 6.7. Theoretical models will provide a framework in which tectonic and sea-level data can be interpreted and evaluated. They will also provide graphical and mathematical means of correlating and dating strandlines. Models will also provide a coherent framework for compiling and comparing strandline data from different areas. • The resolution of existing radiometric, chemical, and paleontologic dating techniques should be im

COASTAL TECTONICS 121 proved and the ranges of these techniques extended. New dating techniques should be developed, especially for the period between 100–1000 ka BP. • Both Pleistocene and Holocene (including historical) strandlines on many more coastlines throughout the world should be mapped and dated. Particular attention should be paid to precise ages and elevations. These data are needed to resolve potentially serious problems of geoidal distortion and regional hydro-isostatic crustal movements. If geoidal distortion is significant on a time scale of 10–100 ka, regional sea-level curves should be developed as tectonic datums. If geoidal distortion is relatively insignificant, its magnitude will help to establish uncertainty limits for rates of crustal deformation. In short, both Pleistocene and Holocene sea-level curves should be improved to provide more accurate estimates of tectonic deformation. • More effective tide-gauge stations and geodetic networks to monitor ongoing crustal movements and regional sea-level changes should be developed. Particular care should be taken in comparing data from these two sources and from other historical and geological sources. • Detailed studies of coseismic strandlines on rapidly deforming coastlines should be undertaken to establish patterns of paleoseismicity. Particular attention should be placed on precise dating. • Existing data on the age and elevation of coastlines throughout the world should be compiled to establish regional patterns of crustal deformation and sea-level change. • Submerged strandlines should be mapped and dated. New techniques for conducting these difficult and expensive investigations should be developed. REFERENCES Anderson, W.A., J.T.Kelley, W.B.Thompson, H.W.Borns, D. Sanger, D.C.Smith, D.A.Tyler, R.S.Anderson, A.E.Bridges, K. J.Crossen, J.W.Ladd, B.G.Anderson, and F.T.Lee (1984). Crustal warping in coastal Maine, Geology 12, 677–680. Andrews, J.T. (1970). Present and postglacial rates of uplift for glaciated northern and eastern North America derived from postglacial uplift curves, Can. J. Earth Sci. 7, 703–715. Angelier, J. (1979). Recent Quaternary tectonics in the Hellenic Arc: examples of geologic observations on land, Tectonophysics 52, 267–275. Atwater, B.F., C.W.Hedel, and E.J.Helley (1977). Late Quaternary depositional history, Holocene sea- level changes, and vertical crustal movement, southern San Francisco Bay, California, U.S. Geol. Surv. Prof. Paper 1014, 15 pp. Balling, N. (1980). The land uplift in Fennoscandia, gravity field anomalies and isostasy, in Earth Rheology, Isostasy and Eustasy, N.-A.Morner, ed., John Wiley & Sons, New York, pp. 297–310. Barnett, D.M. (1966). A re- examination and re-interpretation of tide- gauge data for Churchill, Manitoba, Can. J. Earth Sci. 3, 77–88. Belknap, D.F. (1979). Application of amino- acid geochronology to stratigraphy of late Cenozoic marine units of the Atlantic Coastal Plain, unpublished Ph.D. dissertation, Univ. Del. Bender, M.L., R.G.Fairbanks, F.W.Taylor, R.K.Matthews, J.G. Goddard, and W.S.Broecker (1979). Uranium-series dating of the Pleistocene Indies, Geol. Soc. Am. Bull. 90, 577–594. Berrino, G., G.Corrado, G.Luongo, and B.Toro (in press). Surface deformation and gravity changes accompanying the 1982 Pozzuoli uplift, Bull. Volcan. Bloom, A.L. (1970). Holocene submergence in Micronesia as the standard for eustatic sea-level changes, Geol. Soc. Am. Bull. 81, 145–154. Bloom, A.L. (1977). Atlas of sea- level curves, IGCP Project 61 Sea Level Project, UNESCO, Paris. Bloom, A.L.. W.S.Broecker, J.Chappell, R.K.Matthews, and K.J. Mesolella (1974). Quaternary sea-level fluctuations on a tectonic coast: New 230Th/ 234U dates from the Huon Peninsula, New Guinea, Quat. Res. 4, 185–205. Bradley, W.C., and G.B.Griggs (1976). Form, genesis and deformation of central California wave- cut platforms, Geol. Soc. Am. Bull. 87, 433–449. Broecker, W.S., and M.L.Bender (1973). Age determination on marine strandlines, in Calibration of Homonoid Evolution, W.W. Bishop and J.A.Miller, eds., Scottish Academic Press, pp. 19–35. Broecker, W.S., D.L.Thurber, J.Goddard, T.-L.Ku, R.K.Matthews, and K.J.Mesolella (1968). Milankovitch hypothesis supported by precise dating of coral reefs and deep-sea sediments, Science 159, 297–300. Brown, L.D. (1978). Recent vertical crustal movement along the east coast of the United States, Tectonophysics 44, 205–231. Buchanan-Banks, J.M., R.O.Castle, and J.I.Ziony (1975). Elevation changes in the central Transverse Ranges near Ventura, California, Tectonophysics 29, 113–125. Chappell, J.M. (1974a). Geology of coral terraces, Huon Peninsula, New Guinea: A study of Quaternary tectonic movements and sea- level changes, Geol. Soc. Am. Bull. 85, 553–570. Chappell, J.M. (1974b). Upper mantle rheology in a tectonic region: Evidence from New Guinea, J. Geophys. Res. 79, 390–398. Chappell, J.M. (1983). A revised sea-level record for the last 300,000 years from Papua New Guinea, Search 14, 99–101. Chappell, J.M., and H.H.Veeh (1978). Late Quaternary tectonic movements and sea-level changes at Timor and Atauro Island, Geol. Soc. Am. Bull. 89, 356–368. Clague, J., J.R.Harper, R.J.Hebda, and D.E.Howes (1982). Late Quaternary sea levels and crustal movements, coastal British Columbia, Can. J. Earth Sci. 19, 597–618. Clark, J.A., W.E.Farrell, and W.R.Peltier (1978). Global changes in postglacial sea level: A numerical calculation, Quat. Res. 9, 265–287. Cole, J.W., and K.B.Lewis (1981). Evolution of the Taupo- Hikurangi subduction system, Tectonophysics 72, 1–21. Corwin, G., and H.L.Foster (1959). The 1957 explosive eruption on Iwo Jima, Volcano Islands. Am. J. Sci. 257, 161–171. Cronin, T.M., B.J.Szabo, T.A.Ager, J.E.Hazel, and J.P.Owens (1981). Quaternary climates and sea levels of the U.S. Atlantic coastal plain, Science 211, 233–239. Dambara, T. (1971). Synthetic vertical movements in Japan during recent 70, J. Geod. Soc. Jpn. 17, 100–108. Dames and Moore Consultants (1981). Marine terrace studies and age dating: Final geoseismic investigation for proposed LNG terminal at little Cojo Bay, California, unpublished consultant's report prepared for Western LNG Terminal Associates, Los Angeles, Calif. Davidson, G. (1889). Coast Pilot of California, Oregon and Washington, 4th ed., U.S. Coast and Geodetic Survey.

COASTAL TECTONICS 122 Dodge, R.E., R.G.Fairbanks, L.K.Benninger, and F.Maurrasse (1983). Pleistocene sea levels from raised coral reefs of Haiti, Science 219, 1423–1425. Donner, J.J. (1965). The Quaternary of Finland, in The Quaternary 1, K.Rankama ed., Interscience, New York, pp. 221–272. Donner, J.J. (1980). The determination and dating of synchronous late Quaternary shorelines in Fennoscandia, in Earth Rheology, Isostasy and Eustasy, N.-A.Morner ed., John Wiley & Sons, New York, pp. 285–293. Emery, K.O. (1958). Shallow submerged marine terraces of southern California, Geol. Soc. Am. Bull. 69, 39–60. Farrand, W.R., and R.T.Gajda (1962). Isobases on the Wisconsin marine limit in Canada, Geogr. Bull. 17, 5–22. Faure, H., J.C.Fontes, L.Hebrard, J.Moteillet, and P.A.Piazzoli (1980). Geoidal changes and shore- level tilt along Holocene estuaries: Senegal River area, West Africa, Science 210, 421–423. Fisk, H.N., and E.McFarlan, Jr. (1955). Late Quaternary deltaic deposits of the Mississippi River, in Crust of the Earth, A.Poldervaart, ed., Geol. Soc. Am. Spec. Paper 62, pp. 269–302. Flemming, N.C. (1972). Eustatic and tectonic factors in the relative vertical displacements of the Aegean Coast, in The Mediterranean Sea: A Natural Sedimentary Laboratory, D.J.Stanley, ed., Dowden, Hutchinson and Ross, Inc., Stroudsburg, Pa. Gahni, M.A. (1978). Late Cenozoic vertical crustal movements in the southern North Island, New Zealand, N.Z. J. Geol. Geophys. 21, 117–125. Garrick, R.A. (1979). Late Holocene uplift at Te Araroa, East Cape, North Island, New Zealand, N.Z. J. Geol. Geophys. 22, 131–139. Gibb, J.G. (in press). A New Zealand regional Holocene eustatic sea-level curve and its application to vertical tectonic movements, in Proceedings Volume for 1984 Recent Crustal Movements Conference, Wellington, N.Z. Grant, D.R. (1970). Recent coastal submergence of the Maritime Provinces, Canada, Can. J. Earth Sci. 7, 676–689. Grant, D.R. (1980). Quaternary sea-level changes in Atlantic Canada as an indication of crustal deleveling, in Earth Rheology, Isostasy and Eustasy, N.- A.Morner, ed., John Wiley & Sons, New York, pp. 201–214. Hanks, T.C., R.C.Bucknam, K.R.Lajoie, and R.E.Wallace (1984). Modification of wave- cut and faulting- controlled landforms, J. Geophys. Res. 89, 5771–5790. Harmon, R.S., H.P.Schwarcz, and D.C.Ford (1978). Late Pleistocene sea- level history of Bermuda, Quat. Res. 9, 205–218. Harmon, R.S., T.-L.Ku, R.K.Matthews, and P.L.Smart (1979). Limits of U-series analysis: Phase 1 results of the uranium-series intercomparison project, Geology 7, 405–409. Harmon, R.S., S.L.Lynton, R.M.Mitterer, P.Garrett, H.P. Schwarcz, and J.L.Larson (1981). Bermuda sea level during the last interglacial, Nature 289, 481–483. Hays, W.W., and P.L.Gori (1983). The 1886 Charleston, South Carolina, earthquake and its implications for today, U.S. Geol. Surv. Open-File Rep. 83– 843, 508. Hicks, S.D., and J.E.Crosby (1974). Trends and variability of yearly mean sea level 1893–1972, U.S. Natl. Ocean. Atmos. Admin. Tech. Mem. NOS 13, 1–14. Hillaire-Marcel, C. (1980). Multiple component postglacial emergence eastern Hudson Bay, Canada, in Earth Rheology, Isostasy and Eustasy, N.- A.Morner, ed., John Wiley & Sons, New York, pp. 215–230. Hopley, D. (1983). Deformation of the North Queensland continental shelf in the late Quaternary, in Shorelines and Isostasy, D.E.Smith and A.G.Dawson, eds., Academic Press, New York, pp. 347–368. Hoyt, J.H., and J.R.Hails (1974). Pleistocene stratigraphy of southeastern Georgia, in Post-Miocene Stratigraphy, Central and Southern Atlantic Coastal Plain, R.Q.Oaks, ed., Utah State University Press, pp. 191–205. Inman, D.L., and C.E.Nordstrom (1971). The tectonic and morphologic classification of coasts, J. Geol. 79, 1–21. Kaizuka, S., T.Matsuda, M.Nogami, and N.Yonekura (1973). Quaternary tectonic and recent seismic crustal movements in the Arauco Peninsula and its environs, central Chile, Geogr. Rep. Tokyo Met. Univ. 8, 1–49. Kaizuka, S., T.Miyauchi, and S.Nagaoka (1983). Marine terraces, active faults and tectonic history of Iwo Jima, Ogasawara Res. Comm. of Tokyo Met. Univ. 9, 13–45. Kanaya, A. (1978). Holocene terraces and tectonic movements of Muroto Peninsula, southern Shikoku, Japan, Geogr. Rev. Jpn. 51, 451–463. Kennedy, G.L., K.R.Lajoie, S.A.Mathieson, and D.J.Blunt (1981). Half Moon Bay terrace, California, and the age of its Pleistocene invertebrate faunas, S. Calif. Assoc. Malacol. Proc. 1981. Kennedy, G.L., K.R.Lajoie, and J.F.Wehmiller (1982). Aminostratigraphy and faunal correlations of late Quaternary marine terraces, Pacific Coast, U.S.A., Nature 299, 545–547. Konishi, K., A.Omura, and O.Nakamichi (1974). Radiometric coral ages and sea level records from the late Quaternary reef complexes of the Ryukyu Islands, in Proceedings Second International Coral Reef Symposium, pp. 595–613. Ku, T.-L., and J.P.Kern (1974). Uranium-series age of the upper Pleistocene Nestor Terrace, San Diego, California, Geol. Soc. Am. Bull. 85, 1713–1716. Ku, T.-L., M.A.Kimmel, W.H.Eastan, and T.J.O'Neil (1974). Eustatic sea level 120,000 years ago on Oahu, Hawaii, Science 6, 959–962. Lajoie, K.R., and A.M.Sarna-Wojcicki (1982). Late Quaternary coastal tectonics in the central and western Transverse Ranges, southern California, Geol. Soc. Am. Abstr. Programs 14–4, 179. Lajoie, K.R., J.P.Kern, J.F.Wehmiller, G.L.Kennedy, S.A. Mathieson, A.M.Sarna-Wojcicki, R.F.Yerkes, and P.F.McCrory (1979a). Quaternary marine shorelines and crustal deformation, San Diego to Santa Barbara, California, in Geological Excursions Southern California Area, P.L.Abbott, ed., San Diego State University, pp. 3–15. Lajoie, K.R., G.E.Weber, S.A.Mathieson, and J.Wallace (1979b). Quaternary tectonics of coastal Santa Cruz and San Mateo Counties, California, as indicated by deformed marine terraces and alluvial deposits, in Coastal Tectonics and Coastal Geologic Hazards in Santa Cruz and San Mateo Counties, California, G.E.Weber, K.R.Lajoie, and G.B.Griggs, eds., Geol. Soc. Am. Guidebook. Lajoie, K.R., A.M.Sarna-Wojcicki, and Y.Ota (1982a). Emergent Holocene marine terraces at Ventura and Cape Mendocino, California—indicators of high tectonic uplift rates, Geol. Soc. Am. Abstr. Programs 14–4, 178. Lajoie, K.R., A.M.Sarna-Wojcicki, and R.F.Yerkes (1982b). Quaternary chronology and rates of crustal deformation in the Ventura area, California in Neotectonics of Southern California, J.D. Cooper, comp., Geol. Soc. Am. Guidebook, pp. 43–51. Lajoie, K.R., G.L.Kennedy, S.A.Mathieson, A.M.Sarna-Wojcicki, A.A.Morrison, and M.K.Tobish (1983). Emergent Holocene marine terraces at Cape Mendocino and Ventura, California, U.S.A., in Proceedings International Symposium on Development of Holocene Shorelines, Tokyo, Japan. Lewis, K.B. (1971a). Erosion and deposition on a tilting continental shelf during Quaternary oscillations of sea level, N.Z. J. Geol. Geophys. 2, 281–301. Lewis, K.B. (1971b). Growth rate of folds using tilted wave- planed

COASTAL TECTONICS 123 surfaces: Coast and continental shelf, Hawke's Bay, R. Soc. N.Z. Bull. 9, 225–231. Lewis, K.B. (1974). The continental terrace, Earth Sci. Rev. 10, 37–71. Lundqvist, J. (1965). The Quaternary of Sweden, in The Quaternary, K.Rankama, ed., Interscience, New York, pp. 139–198. Machida, H., H.Nakagawa, and P.Pirazzoli (1976). Preliminary study on Holocene sea levels in the central Ryukyu Islands, Rev. Geomorph. Dynamique 25, 49–62. Marshall, J.F., and J.Launay (1978). Uplift rates of the Loyalty Islands as determined by 230Th/ 234U dating of raised coral terraces, Quat. Res. 9, 186–192. Marshall, J.F., and B.G.Thom (1976). The sea level in the last interglacial, Nature 263, 120–121. Mathews, W.H., J.G.Fyles, and H.W.Nasmith (1970). Postglacial crustal movements in southwestern British Columbia and adjacent Washington State, Can. J. Earth Sci. 7, 690–702. Matsuda, T. (1976). Empirical rules on sense and rate of recent crustal movements, J. Geod. Soc. Jpn. 22, 252–263. Matsuda, T., Y.Ota, M.Ando, and N.Yonekura (1978). Fault mechanism and recurrence time of major earthquakes in southern Kanto district, Japan, as deduced from coastal terrace data, Geol. Soc. Am. Bull. 89, 1610–1618. Matthews, R.K. (1973). Relative elevation of late Pleistocene high sea- level stands: Barbados uplift rates and their implications, Quat. Res. 3, 147–153. Mesolella, K.J., R.K.Mathews, W.S.Broecker, and D.L.Thurber (1969). The astronomical theory of climatic change: Barbados data, J. Geol. 77, 250–274. Miyoshi, M. (1983). Estimated ages of late Pleistocene marine terraces in Japan, deduced from uplift rate, Geogr. Rev. Jpn. 56, 819–834. Moore, J.E., and D.J.Fornari (1984). Drowned reefs as indicators of subsidence of the island of Hawaii, J. Geol. 92, 752–759. Moore, W.S., and B.L.K.Samayajulu (1974). Age determinations of fossil corals using 230Th/ 234Th and 230Th/ 227Th, J. Geophys. Res. 79, 5065–5068. Morner, N.-A. (1976). Eustasy and geoid changes, J. Geol. 84, 123–151. Morner, N.-A. (1978). Faulting, fracturing and seismic activity as a function of glacial-isostasy in Fennoscandia, Geology 6, 41–45. Morner, N-.A. (1980). The Fennoscandian uplift: Geological data and their geodynamical implication, in Earth Rheology, Isostasy and Eustasy , N.- A.Morner ed., John Wiley & Sons, New York, pp. 251–283. Morner, N.-A. (1983). Comment on tectonic uplift of a middle Wisconsin marine platform near Mendocino triple junction, California, Geology 11, 621. Nakamura, K., K.Kasahara, and T.Matsuda (1965). Tilting and uplift of an Island, Awashima, near the epicentre of the Niigata earthquake in 1964, J. Geod. Soc. Jpn. 10, 111–214. Nakata, T., T.Takahashi, and M.Koba (1978). Holocene-emerged coral reefs and sea-level changes in the Ryukyu Islands, Geogr. Rev. Jpn. 51, 87–108. Nakata, T., M.Koba, W.Jo, T.Imaizumi, H.Matsumoto, and T. Suganuma (1979). Holocene marine terraces and seismic crustal movement, Sci. Rep. Tohoku Univ., 7th Ser. (Geogr.) 29, 195–204. Nakata, T., M.Koba, T.Imaizumi, W.Jo, H.Matsumoto, and T. Suganuma (1980). Holocene marine terraces and seismic crustal movements in the southern part of Boso peninsula, Kanto, Japan, Geogr. Rev. Jpn. 53, 29–44. Neumann, A.C., and W.S.Moore (1975). Sea-level events and Pleistocene coral ages in the northern Bahamas, Quat. Res. 5, 3215–224. Newman, W.W., L.J.Cinquemani, R.R.Pardi, and L.L.Marcus (1978). Holocene deleveling of the United States' east coast, in Earth Rheology, Isostasy and Eustasy, N.-A.Morner, ed., John Wiley & Sons, New York, pp. 449–463. Oaks, R.Q., Jr., and J.R.Du Bar, eds. (1974). Post Miocene Stratigraphy Central and Southern Atlantic Coastal Plain, Utah State University Press, 275 pp. Oliver, J., T.Johnson, and J.Dorman (1970). Postglacial faulting and seismicity in New York and Quebec, Can. J. Earth Sci. 7, 579–590. Omoto, K. (1979). Holocene sea- level change: A critical review, Sci. Rep. Tohoku Univ., 7th Ser. 29, 205–222. Ota, Y. (1964). Coastal terraces of the Sado Island, Japan, Geogr. Rev. Jpn. 37, 226–242. Ota, Y. (1975). Late Quaternary vertical movement in Japan estimated from deformed shorelines, in Quaternary Studies, R.P.Suggate and M.M.Creswell, eds., Royal Society of New Zealand, pp. 231–239. Ota, Y., and N.Hori (1980). Late quaternary tectonic movement of the Ryukyu Islands, Japan, Quat. Res. 18, 221–240. Ota, Y., T.Matsuda, and K.Naganuma (1976). Tilted marine terraces of the Ogi Peninsula, Sado Island, central Japan, related to the Ogi Earthquake of 1802, Seismol. J. Soc. Jpn. 11–29, 55–70. Page, W.D., J.N.Alt, L.S.Cluff, and G.Plafker (1979). Evidence for the recurrence of large- magnitude earthquakes along the Makran coast of Iran and Pakistan, Tectonophysics 52, 533–547. Peltier, W.R., and J.T.Andrews (1983). Glacial geology and glacial isostasy of the Hudson Bay region, in Shorelines and Isostasy, D.E. Smith and A.G.Dawson, eds., Academic Press, New York, pp. 285–320. Pillans, B. (1983). Upper Quaternary marine terrace chronology and deformation, South Taranaki, New Zealand, Geology 11, 292–297. Pirazzoli, P.A., J.Thommeret, Y.Thommeret, J.Laboreal, and L.F. Montaggioni (1982). Crustal block movements from Holocene shorelines: Crete and Atikythira (Greece), Tectonophysics 86, 27–43. Plafker, G. (1965). Tectonic deformation associated with the 1964 Alaska earthquake, Science 148, 1675–1687. Plafker, G. (1972). Alaskan earthquake of 1964 and Chilean earthquake of 1960: Implications for arc tectonics, J. Geophys. Res. 77, 901–924. Plafker, G., and M.Rubin (1967). Vertical tectonic displacements in south- central Alaska during and prior to the great 1964 earthquake, J. Geosci. Osaka City Univ. 10, 53–66. Plafker, G., and M.Rubin (1978). Uplift history and earthquake recurrence as deduced from marine terraces on Middleton Island, Alaska, in Proceedings of Conference VI Methods and Identification of Seismic Gaps, pp. 687–721. Plafker, G., and J.C.Savage (1970). Mechanism of the earthquakes of May 21 and 22, 1960, Geol Soc. Am. Bull. 81, 1001–1030. Plafker, F., T.Hudson, M.Rubin, and K.Dixon (1981). Holocene marine terraces and uplift history in the Yakataga seismic gap near Icy Cape, Alaska, U.S. Geol. Surv. Cir. 844, 111–115. Poland, J.F. (1971). Land subsidence in the Santa Clara Valley, Alameda, San Mateo and Santa Clara Counties, California, U.S. Geol. Surv.MF-336, 1:125,000. Research Group for Quaternary Tectonic Map (1969). Quaternary tectonic map of Japan, 1:2,000,000, National Research Center for Disaster Prevention, 6 sheets. Ridley, A.P., and Seeley, M.W. (1979). Evidence for recent coastal uplift near Al Jubail, Saudia Arabia, Tectonophysics 52, 319–327. Ridlon, J.B. (1972). Pleistocene-Holocene deformation of the San Clemente Island crustal block, California, Geol. Soc. Am. Bull. 83, 1831–1844. Sandweiss, D.H., and H.B.Rollins (1981). A single large magnitude

COASTAL TECTONICS 124 uplift in the Holocene record of the Peruvian north coast, Geol. Soc. Am. Abstr. Programs 13, 545. Sarna-Wojcicki, A.M., K.M.Williams, and R.F.Yerkes (1976). Geology of the Ventura fault, Ventura County, California, U.S. Geol. Surv. Map MF-781, 1:6,000. Sarna-Wojcicki, A.M., K.R.Lajoie, and R.F.Yerkes (1986). Recurrent Holocene displacement on the Javon Canyon fault: A comparison of fault movement history with calculated average recurrence intervals, U.S. Geol. Surv. Prof. Paper 1339. Sawamura, T. (1953). Relation between the activities of the outer earthquake zone in southwestern Japan and the geologic structures and crustal movements of Shikoku and its vicinity, Res. Rep. Kochi Univ. 2(15), 1–46. Schofield, J.E. (1973). Post- glacial sea levels of Northland and Auckland, N.Z. J. Geol. Geophys. 16, 359–366. Scholl, D.W., F.C.Craighead, and M.Stuiver (1970). Florida curve revised: Its relation to coastal sedimentation rates, Science 163, 562–564. Shackleton, N.J., and N.D.Opdyke (1973). Oxygen isotope and paleomagnetic stratigraphy of equatorial Pacific core V28–238, Quat. Res. 3, 39–55. Shimazaki, K., and T.Nakata (1980). Time- predictable recurrence model for large earthquakes, Geophys. Res. Lett. 7, 279–282. Shubert, C., and B.J.Szabo (1978). Uranium- series ages of Pleistocene marine deposits on the islands of Curacao and La Blanquilla, Caribbean Sea, in Caribbean Geol. Conf. 8th Proc., H.J.MacGillavry and D.J.Beets, eds., Geol. Mijmbouw 57, 325–332. Sissons, J.B., and R.Cornish (1982). Rapid localized glacio- isostatic uplift at Glen Roy, Scotland, Nature 297, 213–214. Sneh, Y. and M.Klein (1984). Holocene sea-level changes at the coast of Dor, southeast Mediterranean, Science 226, 831–832. Stearns, C.E. (1976). Estimates of the position of sea level between 140,000 and 75,000 years ago, Quat. Res. 6, 445–429. Stearns, C.E. (1984). Uranium- series dating and the history of sea level, in Quaternary Dating Methods, W.C.Mahaney, ed., Elsevier, New York, pp. 53– 66. Stevens, G.R. (1973). Late Holocene marine features adjacent to Port Nicholson, Wellington, New Zealand, N.Z. Geol. Surv. 16, 455–484. Sugimura, A., and Y.Naruse (1954). Changes in sea level, Seismic upheavals and coastal terraces in the southern Kanto Region, Japan, Jpn. J. Geol. Geogr. 24, 101–113. Sutherland, D.G. (1983). The dating of former shorelines; in Shorelines and Isostasy, D.E.Smith, and A.G.Dawson, eds., Inst. Brit. Geogr. Spec. Publ. 16, pp. 129–160. Szabo, B.J. (1979). Uranium-series age of coral reef growth on Rottuest Island, western Australia, Mar. Geol. 29, M11-M15. Szabo, B.J., W.C.Ward, A.E.Weidie, and M.J.Brady (1978). Age and magnitude of the late Pleistocene sea- level rise on the eastern Yucatan Peninsula, Geology 6, 713–715. Tamura, A. (1979). Holocene marine terraces and crustal movements of Sado Island, central Japan, Geogr. Rev. Jpn. 52, 339–355. Tarr, A.C. (1974). World Seismicity map, U.S. Geological Survey. Thatcher, W. (1984). The earthquake deformation cycle at the Nankai Trough, southwest, Japan, J. Geophys. Res. 89, 3087–3101. Thom, B.G., and P.S.Roy (1983). Sea- level changes in New South Wales over the past 15,000 years, Geogr. Dept. James Cook Univ. Monogram Series, Occasional Paper 3, 64–84. Thom, G.G., R.H.Hails, and A.R.H.Martin (1969). Radiocarbon evidence against higher postglacial sea levels in eastern Australia, Mar. Geol. 7, 161– 168. Thompson, W.B., K.J.Borns, H.W.Borns, Jr., and B.G.Anderson (1983). Glacial-marine deltas and late Pleistocene-Holocene crustal movements in southern Maine, in New England Seismotectonic Study Activities in Maine during Fiscal Year 1982, W.B.Thompson and J.T.Kelley, eds., Maine Geol. Surv. Rep., U.S. Reg. Comm., pp. 153–171. Thurber, D.L., W.S.Broecker, W.S.Blanchard, and H.A.Protratz (1965). Uranium- series ages of Pacific atoll coral, Science 149, 55–58. Veeh, H.H. (1966). 230Th/ 238U and 234U/238U ages of Pleistocene high sea-level stand, J. Geophys. Res. 71, 3379–3386. Veeh, H.H., and J.Chappell (1970). Astronomical theory of climate change: Support from New Guinea, Science 167, 862–865. Veeh, H.H., and J.W.Valentine (1967). Radiometric ages of Pleistocene fossils from Cayucos, California, Geol. Soc. Am. Bull. 78, 547–550. Vita-Finzi, C. (1979). Rates of Holocene folding in the coastal Zagros near Bandar Abbas, Iran, Nature 278, 632–634. Walcott, R.I. (1970). Isostatic response to loading of the crust in Canada, Can. J. Earth Sci. 7, 716–734. Walcott, R.I. (1984). The major structural elements of New Zealand, in An Introduction to the Recent Crustal Movements of New Zealand, R.I.Walcott comp., R. Soc. N.Z. Misc. Ser. 7, 1–6. Ward, W.T. (1985). Correlation of east Australian Pleistocene shorelines with deep-sea core stages: A basis for a coastal chronology, Geol. Soc. Am. Bull. 96, 1156–1166. Weber, G.E. (1983). Geologic investigation of the marine terraces of the San Simeon region and Pleistocene activity of the San Simeon fault zone, San Luis Obispo County, California, U.S. Geol. Surv. Tech. Rep. 66. Weber, G.E., and Cotton, W.R. (1981). Geologic investigation of recurrence intervals and recency of faulting along the San Gregorio fault zone, San Mateo County, California, U.S. Geol. Surv. Open-File Rep. 81–263. Weber, G.E., and K.R.Lajoie (1977). Late Pleistocene and Holocene tectonics of the San Gregorio fault zone between Moss Beach and Point Año Nuevo, San Mateo County, California, Geol. Soc. Am. Abstr. Programs 9, 524. Wellman, H.W. (1969). Tilted marine beach ridges at Cape Turakirae, N.Z., Tuatara 17, 82–93. Wellman, H.W. (1971a). Holocene tilting and uplift of the Glenburn coast, Wairarapa, New Zealand, R. Soc. N.Z. Bull. 9, 221–223. Wellman, H.W. (1971b). Holocene tilting and uplift on the White Rocks coast, Wairarapa, New Zealand, R. Soc. N.Z. Bull. 9, 211–215. Wellman, H.W. (1979). An uplift map for the South Island of New Zealand, and a model for uplift of the Southern Alps, R. Soc. N.Z. Bull. 18, 13–21. Yonekura, N., and K.Shimazaki (1980). Uplifted marine terraces and seismic crustal deformation in arc- trench systems: A role of imbricated thrust faulting, EOS 61, 1111. Yoshikawa, T., S.Kaizuka, and Y.Ota (1964), Mode of crustal movement in the late Quaternary of the southeast coast of Shikoku, southwestern Japan, Geogr. Rev. Jpn. 37, 627–648.

Next: 7 Tectonic Geomorphology of Escarpments and Mountain Fronts »
Active Tectonics: Impact on Society Get This Book
×
Buy Paperback | $75.00
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

Over 250,000 people were killed in the Tangshan, China earthquake of 1976, and other less active tectonic processes can disrupt river channels or have a grave impact on repositories of radioactive wastes. Since tectonic processes can be critical to many human activities, the Geophysics Study Committee Panel on Active Tectonics has presented an evaluation of the current state of knowledge about tectonic events, which include not only earthquakes but volcanic eruptions and similar events. This book addresses three main topics: the tectonic processes and their rates, methods of identifying and evaluating active tectonics, and the effects of active tectonics on society.

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  6. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  7. ×

    View our suggested citation for this chapter.

    « Back Next »
  8. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!