Geography

views updated May 21 2018

Geography

I. The FieldRichard Hartshorne

BIBLIOGRAPHY

II. Political GeographyHarold H. Sprout

BIBLIOGRAPHY

III. Economic GeographyRichard S. Thoman

BIBLIOGRAPHY

IV. Cultural GeographyEdward T. Price

BIBLIOGRAPHY

V. Social GeographyAnne Buttimer

BIBLIOGRAPHY

VI. Statistical GeographyBrian J. L. Berry

BIBLIOGRAPHY

The articles under this heading describe the main fields of contemporary geography and the field of statistical geography, which is an approach to geography rather than a discrete field. Other material of direct or related interest to geography may be found under Area; Cartography; Central place; City; Conservation; Culture area; Diffusion, article onthe diffusion of innovations; Ecology; Enclaves and exclaves; Environment; Environmentalism; Industrial concentration; Land; Landscape; Location theory; Planning, social; Population; Rank-size relations; Region; Regional science; Water resources. Biographical articles of relevance to geography include Bowman; Brown; Fleure; Hettner; Humboldt; Huntington; KjellÉn; Mackinder; Marsh; Ratzel; Rltter; Sauer; Teleki; Vldal de la Blache.

I THE FIELD

Geography is neither a purely natural nor a purely social science. From its early development as an organized field of knowledge in classical Greece, geography has included animate as well as inanimate things, man and his works as well as nature. This was of little concern as long as man was regarded as an integral part of nature. But geography, although a very old subject, did not become established as a university discipline with an organized academic profession until after the natural and social sciences had become divided into separate faculties. Regular university departments of geography were first established in German-speaking countries in the 1870s and 1880s; in France a little later; and in Great Britain and the United States generally in the present century. In each country the first generation of professors of geography had been trained in other fields, in most cases the natural sciences. Self-taught in geography, few of them outside Germany were fa miliar with its past development.

One consequence was that geography tended to separate into two parts: one natural, more commonly called physical geography, and one human or social, sometimes called economic geography. (In different countries and in different institutions in the same country practice varies as to whether the subject is part of the natural science faculty, part of the social science faculty, or split between the two.) For many students, on the other hand, it was the connections between the two parts, the man-land relationships, that constituted the distinctive subject matter of geography. During the period of its initial establishment as a university subject in the English-speaking countries, geography was commonly defined as the study of the relationships, in whichever direction, between the natural environment and man. This “environmen talist” concept of the field seems to have been first formulated in Germany, and then in France, late in the nineteenth century, but it was very soon discarded, both in theory and practice by German geographers and, to a large degree, in practice by those of France; in both countries geography returned to its historic focus of interest in the study of the distinctive character of the areas of the earth. This concept and its historical background were made familiar to American and British geographers only in the 1920s and 1930s. And although today few geographers would assert the environmentalist concept, many of its aspects, notably the emphasis on a man-nature dichotomy, still color much of their writing. [seeEnvironmentalism.]

Modern geography, like the geography of past centuries, studies the earth as the space in which man lives—his habitat, milieu, or environment. This includes not just part of the environment, the physical or natural part, but the total environment; in any inhabited area the environment of today has been in part produced by man, and the existing population constitutes a living factor in the present environment. Geography, of course, is not alone in studying man’s environment. Many fields in the natural and social sciences study a particular category of phenomena, not excluding its distribution and variations over the earth. What geography, and geography alone, studies is the areal character of the earth in which man lives—the form, the content, and the function of each areal part, region, or place and the pattern of and interconnections between the areal parts.

If the total diversity of places and their interrelations were simply the sum of areal variations and connections of physical, biological, and social phe nomena, the subject could readily be divided into distinct fields: physical geography, biogeography, and human, or social, geography; or possibly two parts, the geography of nature and the geography of man. In reality, however, the phenomena of these several abstract categories are in many cases very closely interrelated in their areal variations and connections from place to place. Indeed, what the geographer observes as individual features—i.e., a soil, river water, a farm, a transport route—are element complexes in which factors of physical, animate, and social origin are so intricately interwoven as to require study within a single field.

Places or areas, large or small, may be studied either specifically or generically. Human interest in individual places is indicated by the practice from earliest times of giving each area a proper name—“Hudson River,” “Pennsylvania,” or “the South.” Geography, like history, is ultimately concerned with attaining maximum comprehension of individual cases. An essential step in the description as well as the understanding of the individual area is the determination of its generic characteristics. When we speak of places as “deserts,” “can yons,” “cities,” “farms,” or “culture areas,” we limit the criteria in each case to a few closely interrelated features, overlooking aspects in which places of the same type may be radically different. Comparative study of the characteristics of places by kind may reveal indications of significant correspondence, leading to hypotheses of generic relationships. [The use of modern statistical methods to discover and determine such correlations is discussed below in Geography, article onstatistical geography.]

Among the social sciences geography, like history, overlaps the fields which study a particular category—economic, political, or sociological. In geography, as in history, it is the integrated combinations of diverse elements, in their complex inter relationships, that form the direct subjects of study. While the ultimate objective in geography is comprehension of the full integration of areas, analysis requires focusing successively on partial integrations. Comparative study of areas, to establish generic concepts and relationships, must be limited to partial integrations over many areas or over the whole world. Such studies may be confined to a very narrow topic, such as the relation of crop yield to rainfall, or may cover the whole group of features which form the economy of areas. [Divisions of human geography based on common groupings of social features are treated in the articles that follow.]

Richard Hartshorne

BIBLIOGRAPHY

An excellent introduction by a professional geographer is Broek 1965. A much more exhaustive treatment, with lengthy bibliographies, may be found in Hartshorne 1939; 1959; two collections of essays, James & Jones 1954 and Taylor 1951, should also be consulted. The classic work in German geography is Hettner 1927. The French literature is reviewed in Claval 1964. The bibligraphies of the articles that follow should also be consulted.

Broek, Jan O. M. 1965 Geography: Its Scope and Spirit. Columbus, Ohio: Merrill.

Claval, Paul 1964 Essai sur I’évolution de la géographic humaine. Cahiers de géographie de Besançon, No. 12. Paris; Les Belles-Lettres.

Crone, Gerald R. 1951 Modern Geographers: An Outline of Progress in Geography Since 1800 A.D. London: Royal Geographical Society.

Freeman, Thomas W. (1961) 1963 A Hundred Years of Geography. Chicago: Aldine.

Hartshorne, Richard (1939) 1964 The Nature of Geography: A Critical Survey of Current Thought in the Light of the Past. Lancaster, Pa.: Association of American Geographers.

Hartshorne, Richard 1959 Perspective on the Nature of Geography. Association of American Geographers, Monograph Series, No. 1. Chicago: Rand McNally. → A restatement and, in part, an extensive revision of Hartshorne (1939).

Hettner, Alfred 1927 Die Geographie: Ihre Geschichte, ihr Wesen und ihre Methoden. Breslau (then Ger many): Hirt.

James, Preston E.; and Jones, Clarence F. (editors) (1954) 1964 American Geography: Inventory and Prospect. Syracuse Univ. Press.

National Council for the Social Studies 1959 Year book, 29th: New Viewpoints in Geography. Edited by James E. Preston. Washington: The Council.

Taylor, Thomas Griffith (editor) (1951) 1958 Geography in the Twentieth Century: A Study of Growth, Fields, Techniques, Aims and Trends. 3d ed., enl. New York: Philosophical Library.

II POLITICAL GEOGRAPHY

Political geography may be defined from the disciplinary perspective of either geography or political science. From the former perspective, political geography appears as “the study of political phenomena in their areal context” (Jackson 1964, p. 1). This is amplified in the statement political geography is “the study of areal differences and similarities in political character as an interrelated part of the total complex of areal differences and similarities” (Hartshorne et al. 1954, p. 178). Attention to areal dimensions and patterns—sug gested by such terms as location, distance, space, distribution, configuration—is also evident in research under the rubric of political science; the same holds for research in political history and politically oriented research in sociology and other disciplines.

Concepts and techniques

Classical writings on political subjects contain many speculations regarding man’s relations to the earth (Thomas 1925). Several contemporary political scientists have given special attention to areal aspects of political institutions, processes, relationships, and policies (e.g., Spykman 1938; 1942; 1944; Spykman & Rollins 1939; Deutsch 1953; Smuckler 1953; Herz 1957; Sprout 1931; 1963; Sprout & Sprout 1946; 1960; 1962; 1965). But with very few exceptions (e.g., Van Dyke 1960, p. 128), commentaries on the political science discipline deal with the areal focus trivially or not at all.

Geographers have given more attention to areal aspects of political phenomena. Although its antecedents go back to the nineteenth century and earlier (Hartshorne 1935), the modern field of political geography dates in America from World War I (e.g., Bowman 1921) and in Europe from somewhat earlier (Ratzel 1897; George 1901; Mackinder 1902; 1904; 1911-1923; Fairgrieve 1915; and others). The American college catalogues of 1930 announced few courses in political geography. Thirty years later the number exceeded 300, a growth accompanied by proliferation of teaching materials and buttressed by theoretical and substantive research. Contributors to theory include Whittlesey (1939), Hartshorne (1935; 1950), Gottmann (1952), and Jones (1954a; 1954b).

In the idiom of modern geography, geographic quality attaches to any phenomena, human as well as nonhuman, intangible as well as tangible, exhibiting areal dimensions and associations that “give character to particular places” (James & Jones 1954, p. 4). To anticipate a point that will be stressed later, areal patterns of behavior and other intangible human phenomena are becoming increasingly central to research in political geog raphy.

Similarly, the “political” in political science refers to more than the formal apparatus of government; political quality attaches to any aspect of power and influence in society. A community organized on the basis of power is by definition a political community. Every political community (though not every political organization) has a territorial base. Country, often used as a synonym for state, denotes the territorial aspect of a state. Province, city, village, school district, port authority, and other subdivisions of a state all carry territorial connotations. The same holds for empire, political bloc, coalition, and other terms that identify units and combinations of units in imperial and international relationships.

Political areas

In the idiom of geography, any section of the earth’s surface delineated by refer ence to political criteria is a political area. These criteria include de jure jurisdiction and authoritative decision making. Political areas so delineated include nation-states, their formally constituted subdivisions, and empires. These are unquestion ably significant political areas, and a great deal of political geography has been written in terms of them. As geographers have emphasized (e.g., Whit tlesey 1935), the “impress” of political authority changes both the physical and social aspects of landscapes: it affects, for example, inspection stations and other boundary structures; transportation grids that conform to political requirements (e.g., Wolfe 1963); movements of goods and peo ple within a frame of migration and commercial laws; and linguistic and other cultural homoge neities imposed by political authority.

But delineating political areas by reference only to political authority and legal jurisdiction leaves many phenomena uncovered: for example, it does not account for frontier zones that exhibit political homogeneities of personal behavior at variance with de jure jurisdiction (Hartshorne 1950), and areal patterns of behavior that are within a recog nized territorial jurisdiction but not coterminous with its boundaries, such as the region of “isolation” in the United States (Smuckler 1953) or the Washington—Boston “megalopolis” (Gottmann 1961).

The criteria of jurisdiction and authority completely fail to delineate areas that exhibit patterns of political interaction but no overarching organization of authority. These are the characteristic patterns of international politics, whether of the society of nations as a global whole or of less-than-global areas such as the communist “realm,” the so-called free world, the Atlantic “community,” and many others.

Geographers have traditionally emphasized the more tangible aspects of political areas. This emphasis is evident in Sauer’s morphological conception of political geography as “the study of political landscapes”—i.e., “the administrative centers, the boundaries, and the defensive lines and positions” (1927, p. 208). With reference to political bound aries, Fischer (1949) noted that geographers have usually stressed the stabilizing influence of physio graphicfactors, often to the neglect of historical and other cultural processes. Stephen B. Jones (1959) reviewed the geographical literature on boundaries, analyzing the ways in which these have been conceived in different places and periods.

Geographers have given increasing attention to intangible factors and to processes of social change. This trend is evident in the writings noted above. In a plea for more “functional” political geography, Hartshorne (1950) emphasized the importance of “centripetal” and “centrifugal” ideas and social movements in the evolution of state areas. Gottmann (1951) introduced the concepts of circulation (a French term for which the nearest English equivalent is probably “movement”) and icono-graphie (symbols that foster loyalty, solidarity, and conformity) as organizing ideas for the analysis of change and resistance to change in political areas. Jones (1954a) brought these and other ideas together in a “unified field theory.”

Political potential

In a period of history, the results of political interaction, whether within a single national community or upon the broader stage of international politics, exhibit areal patterns of coercion–submission and influence-deference. Within nation-states and empires, these patterns evolve under processes of authoritative decision making, no matter how primitive or obscure these processes may be. The patterns of international politics, however, have evolved in the absence of legitimized overriding authority. In consequence, these patterns are derived in the main from less sharply delineated images of the distribution of power and influence among the interacting national communities.

There is no standard term to denote the aggre gate pressure, pull, attraction, or simply effect that one nation or coalition exerts on the behavior of others. The term “power,” with its strong military connotations, is too restrictive, since relations of influence–deference are derived from a much broader spectrum of behavior than coercion–sub mission—i.e., brute force and the threat of it. The term “political potential” has been suggested to denote this broader spectrum (Sprout & Sprout 1962, p. 158).

The concept of political potential has definite areal connotations. It expresses areal variations in the intensity and efficacy of a government’s demands on other nations, i.e., when and how it gets what where. In addition, political potential expresses the total, or aggregative, effect of a nation’s statecraft plus effects deriving from that nation’s sheer presence on the international landscape, which are evident in the behavior of other nations. Common sense, confirmed by observation, indicates that location, distance, space, the configuration of lands and seas, and the distributions of population, raw materials, technology, institutions, ideologies, and other phenomena may all have some bearing on the political potential of every nation and on the resulting patterns of political interaction and relationship [seeMilitary power potential].

Maps of the international potentials of both particular states and the major regions and the society of nations as a whole might somewhat resemble a topographic contour map. Such maps, however, do not exist; perhaps the available data are too amorphous and ambiguous to render trustworthy mapping possible in the present state of knowledge. But studies of political potential from a geographic perspective, utilizing geographic methods, should help to clarify the areal concepts implicit in the vocabulary of international politics—e.g., bipolar-ity, polycentricity, balance of power, sphere of influence, political orbit, and many others in common use [seeInternational politics].

Geographic areas and political systems

Although geographers and political scientists share an inter est in political phenomena, their disciplinary frame works are recognizably different. Political scientists exhibit interest in areal dimensions and patterns only to the extent that these seem to contribute to an understanding of institutions, processes, relationships, and issues of public concern. In conse quence, political analysts tend to view geographic dimensions and patterns from a predominantly ecological perspective, i.e., relations between “po litical actors” and their milieux. The ways in which students of politics generally frame problems tend to focus attention especially on the psychological linkages between actor and milieu (Sprout & Sprout 1965).

For most geographers (but there are many exceptions) interest in ecological relationships, although generally active, tends to be subsidiary to areal patterns per se. However, this contrast in perspective should not be exaggerated. When one considers the increasing attention that geographers are giving to intangible social patterns and their evolution through time and to the values and moti vations that underlie such patterns, it may be more nearly correct to say that “area” is simply the frame of reference within which geographers study polit ical behavior and its results.

In political theory the concept of system has come to occupy a position somewhat analogous to the concept of area in geography [seeSystems analysis]. These organizing ideas are interestingly relatable. What appears from the geographic perspective to be a political area—city, province, state, empire, major political region, etc.—may appear from the viewpoint of political science to be a system, i.e., a constellation of political units (in dividuals, groups, or organized communities) that interact and relate in describable patterns. Hartshorne and others have emphasized the comple mentarity of these perspectives (James & Jones 1954, p. 174).

This complementarity comes through strongly in Jones’s “unified field theory of political geography” (1954a). Jones’s model consists of five interconnecting categories—“political idea-decision-move ment-field-political area.” These are likened to a “chain of lakes or basins … at one level, so that whatever enters one will spread to all the others.” There are close counterparts to Jones’s “basins” in other vocabularies. His term “political idea” ap proximates the concepts of image and goal-orientation in behavioral theory. “Decision” is just what it is elsewhere. “Movement” and “field” seem to be more or less analogous to a course of action that changes, even as it is changed by, the encompas sing milieu. Finally, “political area” includes “any political organized area” that has “recognized limits, though not necessarily linear or permanent.” Thus, the communist international system, the Atlantic alliance, the British Commonwealth, or any other international system constitutes (with suitable change in perspective and terminology) a political area, just as does a state, a subdivision thereof, or any other areally expressible system of political interaction.

Flow from idea to area, in Jones’s model, is the process by which people control and alter their milieu. The idea of man as a geographic agent, refashioning the landscape along with the physical processes of nature, is an important concept of modern geography. It was given arresting expression in the mid-nineteenth century (see Glacken 1956, pp. 70 ff.) and, more recently, by Sauer (1925) and others. The increasing capacity in technically advanced societies to alter the milieu has immense political implications.

The concept of flow from idea to area, in Jones’s model—i.e., from image and purpose to operational result in the behavioral idiom—rests (more often implicitly than explicitly) on the general man–milieu hypothesis of “possibilism.” This is the proposition that the capacities of the actor and the properties of his milieu set limits to his accomplishment with reference to any given course of action and that these limitations may be operative irrespective of whether or how he perceives and takes them into account. One corollary is the hypothesis that the higher the level of technology, the greater becomes human capacity to control and modify nonhuman components of the milieu. Another corollary is that an operator’s ability to affect the behavior of human components of his milieu depends on his capabilities in relation to theirs at the place where their relative capabilities are tested.

The reverse flow, from area to idea, in Jones’s model, is the process by which the milieu is said to condition human behavior. This conditioning process has been a focus of controversy, largely because of the teleological imagery to which many writers (though not Jones) are addicted. The influences ascribed to nature or to other aspects of the milieu can be expressed, free from teleological overtones, by such psychological concepts as perception, cognition, recognition, stimulus, response, feedback, etc. (Sprout & Sprout 1965). Expressing the conditioning process in such terms emphasizes the complementarity of geography and behavioral science. With certain exceptions, usually unimportant in political contexts, the phenomena of psychological stimulus and response provide the only demonstrated path of influence from milieu to actor, from environment to environed organism, from “area” to “ideas.” The psychological nature of environmental conditioning of behavior (from which many areal patterns are derived) has long been understood, although not always clearly delineated (e.g., Mackinder 1919, p. 28; Febvre 1922, p. 171; James & Jones 1954, p. 13). At least one contemporary geographer has explicitly restated this process in the idiom of psychological theory (Kirk 1952).

Geographic techniques

A major contribution which geographers have made to the study of political phenomena is the development of graphic techniques. Most political areas are too large to be directly perceived in toto. The eye may not differentiate and relate various categories of phenomena distributed over the area, even when they are directly perceivable; hence the value of graphic modes of research, analysis, and presentation, by means of model globes, maps, cartograms, photographs, etc. (Bowman 1934, chapter 4; Boggs 1948). Maps delineating selected features of an area can be compared and superposed (e.g., Bowman [1921] 1928, pp. 146, 460). Large segments of the earth’s surface can be examined from different perspectives (Harrison 1944; Boggs 1945). High-altitude and low-altitude photographs and oblique-angle pictures add new dimensions and textures to the perception of smaller areas (e.g., Gutkind 1956, pp. 1 ff.). Maps and cartograms can deceive as well as inform and hence are powerful instruments of political propaganda (Boggs 1946). Maps and other graphic tools not only carry preconceived messages but, when studied, may also evoke new insights and hypotheses [seeCartography].

Research in political geography

The substantive literature divides roughly into two categories: (1) works that focus on political areas as such; and (2) works that utilize areal concepts and patterns to explain or to predict political events. This cleavage more or less follows disciplinary lines, but by no means consistently. Some of the more important theoretical works have been cited, and a few teaching books are included in the bibliography, along with works cited in this text. A more comprehensive bibliography is appended to the long essay on political geography in James and Jones (see Hartshorne et al. 1954).

With respect to research on particular political areas, one should distinguish between works that deal primarily with political phenomena in an areal context and those that merely utilize political boundaries as a frame of reference for a wider range of phenomena—e.g., works on areal distributions of agriculture, industry, communication grids, etc.

Works that are politically oriented in the stricter sense cluster around various focuses. One of these is the formation, expansion, and disintegration of political areas (e.g., Bowman 1921; Whittlesey 1939; Hartshorne 1950; Deutsch 1953). The following are of particular interest: Herz’s analysis of the formation of modern “large-area” states and of the advances in technology that are making these states progressively vulnerable to economic, psychological, and military penetration (1957); Vevier’s essay on geographical ideas in the territorial expansion of the United States (I960); and Hart’s hypotheses regarding the logistical requisites of political areas (1949). There are many studies of functioning political areas which reflect the varied research perspectives and methods of several disciplines. The Searchlight Series, edited by G. W. Hoffman and G. E. Pearcy, offers a continually growing list of short books of this type.

A second focus is on the analysis of political areas in the light of population growth, spreading communication grids, industrialization, and urban sprawl (e.g., Gottmann 1961; Wolfe 1963). Worthy of special mention is Mumford’s historical study of cities (1961).

A third focus is on areal patterns delineated not by political authority but by civic attitudes and preferences. This involves studies of integrative and disintegrative ideas and movements within state areas (e.g., Hartshorne 1950); attitudes toward the “national space” (e.g., Herman’s study of communist China, 1959); regional variations in civic postures toward public problems such as military defense and foreign policy (e.g., Beard 1934; Sprout & Sprout 1939; Smuckler 1953).

A fourth focus is on the role of political authority in the development, use, depletion, conservation, and renewal of resources. These questions are approached from different angles in International Symposium on Man’s Role …(1956) and also in Udall (1963) and Herber (1962). The effects of resource use and of regulations governing use constitute important exhibits of the “impress” of political authority upon the earth.

A fifth research focus is on political regions larger than nation-state areas. These include an cient and modern empires (e.g., Fawcett 1951; Fisher 1950) and international regions delineated in various ways (Jones 1955a). Boggs’s essay on the Western Hemisphere (1945) focused attention on the need for precise criteria for delineating ma jor political regions. This need is exemplified in some textbooks, in which political regions are variously delineated by physiographic, historical, broadly cultural, or other criteria, sometimes with out clear demonstration of political relevance.

The geographic dimensions and patterns of in ternational politics have been analyzed from various perspectives. Virtually every textbook on international politics gives attention to the uneven distribution of people and things among nations. Areal variations are recognized to be strategic in the estimation of state capabilities (e.g., Sprout & Sprout 1962; Jones 1954b). Such variations form the basis of hypotheses invented to explain or to predict patterns of interaction in the society of nations (Jones 1955a; 1955fc; Sprout 1963). Such hypotheses represent attempts to identify factors whose uneven distribution in space provides a plausible explanation of international patterns.

Geopolitics

Most attention has been given to hypotheses derived from the global and regional configuration of lands and seas. These include Mahan’s sea-power interpretation of history (1900; Sprout & Sprout 1962); Mackinder’s hypothesis of trend toward a world empire based in the “heartland” of Eurasia (1904; 1919), later modified considerably (see especially 1911-1923, vol. 2; 1943); and variants and critiques of Mahan’s and Mackinder’s theories too numerous to list here [seeMackinder; Mahan; see also, e.g., Fairgrieve 1915; Dorpalen 1942; Spykman 1944; East & Moodie 1956, chapter 18].

Climatic variations have inspired another set of geopolitical hypotheses and critiques (e.g., Huntington 1915; Wheeler 1946; Mills 1949; Missenard 1954). International political patterns have also been linked with the uneven distribution of the various raw-material requisites of modern industry. There is some disposition to regard areal differentials in technology as the critical variable (e.g., Brown 1956), a hypothesis that has been linked with demographic distribution to produce a prediction that international political patterns will ultimately be determined by the latter. The pre diction is based on the premise that technological primacy will vary with relative numbers of supe rior scientists and other gifted individuals, the incidence of such individuals varying, in the long run, with the size of population (Blount 1957; critique by Sprout 1963).

The adjective “geopolitical” requires some explanation. Political geography in general, and in ternational political geography in particular, is often confused with geopolitics. This word entered the English language as a loose translation of Geopolitik, which came, in the interwar period, 1919-1939, to denote mobilization of areal knowl edge for purposes of state—in short, geo-policy. Geopolitics was associated in particular with the Institut für Geopolitik in Munich, directed by Karl Haushofer, a general turned geographer and propagandist, who is widely believed (perhaps mistakenly) to have contributed significantly to Hitler’s strategy of conquest (e.g., Dorpalen 1942; Fifield 1945).

Because certain Germans exploited the concept of Lebensraum and other geopolitical ideas for aggressive purposes, many in America and else where illogically concluded that any mixing of geography and politics must be tainted with war and conquest. Geographers insisted that geopolitics was a part of political science. Political scientists tossed the pariah subject back to the geographers. Time has blurred the odious policy connotations of geopolitics, perhaps more so in America than in Europe. The term has even acquired some respect ability, especially in the context of military-defense analysis.

The adjective “geopolitical,” never as value laden as the noun “geopolitics,” was employed sparingly in the 1930s (e.g., Whittlesey 1939), and increasingly in more recent years, to denote the areal aspect of any political pattern and, in particular, hypotheses that purport to explain or to predict areal distributions and patterns of political potential in the society of nations. All such hypotheses represent assessments of opportunities and limitations implicit in the properties of the interacting political communities and of the milieu in which they operate. Such assessments (in the idiom of ecological theory) are essentially possibilistic, even though they may be expressed in deterministic or near deterministic rhetoric (Sprout & Sprout 1965).

Harold Sprout

[See alsoEcology; International Politics. Other relevant material may be found underEnclaves and Exclaves; International Relations.]

BIBLIOGRAPHY

Alexander, Lewis M. (1957) 1963 World Political Pat terns. 2d ed. Chicago: Rand McNally. → A college text organized regionally.

Beard, Charles A. 1934 The Idea of National Interest: An Analytical Study in American Foreign Policy. New York: Macmillan.

Blount, B. K. 1957 Science Will Change the Balance of Power. New Scientist 2, no. 32:8–9.

Boggs, S. W. 1945 This Hemisphere. U.S. Department of State, Bulletin 12:845–850.

Boggs, S. W. 1946 Cartohypnosis. U.S. Department of State, Bulletin 15:1119–1125.

Boggs, S. W. 1948 Geographic and Other Scientific Techniques for Political Science. American Political Science Review 42:223–248.

Bowman, Isaiah (1921) 1928 The New World: Prob lems in Political Geography. 4th ed. New York: World. → The first edition is important as an early comprehensive American work on political geography; the later editions, considerably revised, have more enduring value.

Bowman, Isaiah 1934 Geography in Relation to the Social Sciences. New York: Scribner.

Brown, Harrison 1956 Technological Denudation. Pages 1023-1032 in International Symposium on Man’s Role in Changing the Face of the Earth, Prince ton, N.J., 1955, Man’s Role in Changing the Face of the Earth. Edited by William L. Thomas et al. Univ. of Chicago Press.

Dasmann, Raymond F. 1963 The Last Horizon. New York: Macmillan.

Deutsch, Karl W. 1953 The Growth of Nations: Some Recurrent Patterns of Political and Social Integration. World Politics 5:168–195.

Dorpalen, Andreas 1942 The World of General Haus hofer. New York: Farrar.

East, William G.; and Moodie, A. E. (editors) 1956 The Changing World: Studies in Political Geography. New York: World. → A symposium textbook in the tradition of Isaiah Bowman’s The New World.

Fairgrieve, James (1915) 1941 Geography and World Power. 8th ed., rev. New York: Dutton. → After 1915 a new Chapter 18 was included that considerably altered the main thesis of the book.

Fawcett, Charles B. (1951) 1957 Geography and Em pire. Pages 418-432 in Thomas Griffith Taylor (edi tor), Geography in the Tvientieth Century. 2d ed., rev. New York: Philosophical Library.

Febvre, Lucien (1922) 1925 A Geographical Introduction to History. New York: Knopf. → First published as La terre et revolution humaine.

Fifield, Russell H. 1945 Geopolitics at Munich. U.S. Department of State, Bulletin 12:1152–1162.

Fischer, Eric 1949 On Boundaries. World Politics 1: 196–222.

Fisher, Charles A. 1950 The Expansion of Japan: A Study in Oriental Geopolitics. Geographical Journal 115:1-19, 179–193.

Fisher, Charles A. 1964 Southeast Asia: A Social, Economic, and Political Geography. New York: Dutton.

George, Hereford B. (1901) 1924 The Relations of Geography and History. 5th ed., rev. & enl. Oxford: Clarendon.

Glacken, Clarence J. 1956 Changing Ideas of the Habitable World. Pages 70-92 in International Sym posium on Man’s Role in Changing the Face of the Earth, Princeton, N.J., 1955, Man’s Role in Changing the Face of the Earth. Edited by William L. Thomas et al. Univ. of Chicago Press.

Goblet, Yann M. 1955 Political Geography and the World Map. New York: Praeger.

Gottmann, Jean 1951 Geography and International Re lations. World Politics 3:153–173.

Gottmann, Jean 1952 La politique des etats et leur geographic Paris: Colin.

Gottmann, Jean (1961) 1964 Megalopolis: The Urbanized Northeastern Seaboard of the United States. Cam bridge, Mass.: M.I.T. Press. → An urban-economic evaluation of the nature of a continuously urbanized section of the eastern United States.

Gutkind, E. A. 1956 Our World From the Air: Conflict and Adaptation. Pages 1-44 in International Sym posium on Man’s Role in Changing the Face of the Earth, Princeton, N.J., 1955, Man’s Role in Changing the Face of the Earth. Edited by William L. Thomas et al. Univ. of Chicago Press.

Harrison, Richard E. 1944 Look at the World: The Fortune Atlas for World Strategy. New York: Knopf.

Hart, Hornell 1949 Technology and the Growth of Political Areas. Pages 28-57 in William F. Ogburn (editor), Technology and International Relations. Univ. of Chicago Press.

Hartshorne, Richard 1935 Recent Developments in Political Geography. American Political Science Review 29:785-804, 943–966.

Hartshorne, Richard 1950 The Functional Approach in Political Geography. Association of American Geographers, Annals 40:95–130.

Hartshorne, Richard 1959 Perspective on the Nature of Geography. Association of American Geographers, Monograph Series, No. 1. Chicago: Rand McNally. → A restatement and, in part, an extensive revision of “The Nature of Geography: A Critical Survey of Cur rent Thought in the Light of the Past,” published in 1939.

Hartshorne, Richard et al. 1954 Political Geography. Pages 167-225 in Preston E. James and Clarence F. Jones (editors), American Geography: Inventory and Prospect. Syracuse Univ. Press. → An excellent bibli ography appears on pages 222–225.

Herber, Lewis 1962 Our Synthetic Environment. New York: Knopf.

Herman, Theodore 1959 Group Values Toward the National Space: The Case of China. Geographical Review 49:164–182.

Herz, John H. 1957 Rise and Demise of the Territorial State. World Politics 9:473–493.

Huntington, Ellsworth (1915) 1924 Civilization and Climate. 3d ed., rev. New Haven: Yale Univ. Press.

International Symposium on Man’s Role in Changing the Face of the Earth, Princeton, N.J., 1955 1956 Man’s Role in Changing the Face of the Earth. Edited by William L. Thomas et al. Univ. of Chicago Press.

Jackson, W. A. Douglas 1958 Whither Political Geog raphy? Association of American Geographers, Annals 48:178–183.

Jackson, W. A. douglas (editor) 1964 Politics and Geographic Relationships. Englewood Cliffs, N.J.: Prentice-Hall. → A well-chosen collection of teaching materials drawn from several disciplines.

James, Preston E.; and Jones, Clarence F. (editors) 1954 American Geography: Inventory and Prospect. Univ. of Syracuse Press.

Jones, Stephen B. (1954a) 1964 A Unified Field The ory of Political Geography. Pages 101-109 in W. A. Douglas Jackson (editor), Politics and Geographic Re lationships. Englewood Cliffs, N.J.: Prentice-Hall.

Jones, Stephen B. (1954b) 1964 The Power Inventory and National Strategy. Pages 318-338 in W. A. Doug las Jackson (editor), Politics and Geographic Relation ships. Englewood Cliffs, N.J.: Prentice-Hall.

Jones, Stephen B. 1955a Views of the Political World. Geographical Review 45:309–326.

Jones, Stephen B. 1955b Global Strategic Views. Geo graphical Review 45:492–508.

Jones, Stephen B. 1959 Boundary Concepts in the Set ting of Place and Time. Association of American Geographers, Annals 49:241–255.

Kirk, William 1952 Historical Geography and the Con cept of the Behavioural Environment. Indian Geo graphical Journal [1952]: 152–160.

Mackinder, Halford (1902) 1930 Britain and the British Seas. 2d ed. Oxford: Clarendon.

Mackinder, Halford 1904 The Geographical Pivot of History. Geographical Journal 23:421–444.

Mackinder, Halford 1911-1923 Nations of the Mod ern World. 2 vols. London: Philip.

Mackinder, Halford (1919) 1942 Democratic Ideals and Reality: A Study in the Politics of Reconstruction. London: Constable: New York: Holt.

Mackinder, Halford 1943 The Round World and the Winning of the Peace. Foreign Affairs 21:595–605.

Mahan, Alfred Thayer (1900) 1905 The Problem of Asia and Its Effect Upon International Policies. Bos ton: Little.

Maull, Otto (1925) 1956 Politische Geographic Ber lin: Safari Verlag.

Mills, Clarence A. 1949 Temperature Dominance Over Human Life. Science 110:267–271.

Missenard, AndrÉ 1954 À la recherche de Vhomme. Paris: Librairie Istra. → See especially Part 3, “Climat et milieu physique.”

Mumford, Lewis 1961 The City in History: Its Origins, Its Transformations, and Its Prospects. New York: Harcourt.

Pearcy, George E. et al. (1948) 1957 World Political Geography. New York: Crowell. → A regionally organ ized symposium textbook.

Pounds, Norman J. G. 1963 Political Geography. New York: McGraw-Hill. → A book designed for college teaching.

Ratzel, Friedrich (1897) 1923 Politische Geographie 3d ed. Edited by Eugen Oberhummer. Munich and Berlin: Oldenbourg.

Sauer, Carl O. (1925) 1963 The Morphology of Land scape. Pages 315-350 in Carl O. Sauer, Land and Life: A Selection From the Writings of Carl Ortwin Sauer. Berkeley: Univ. of California Press.

Sauer, Carl O. 1927 Recent Developments in Cultural Geography. Pages 154-212 in E. C. Hayes (editor), Recent Developments in the Social Sciences. Philadelphia: Lippincott.

SchÖller, Peter 1958 Das Ende einer politischen Geo graphie ohne sozialgeographische Bindung. Erdkunde: Archiv für wissenschaftliche Geographie 12:313–316.

Smuckler, Ralph H. 1953 The Region of Isolationism. American Political Science Review 47:386–401.

Sprout, Harold 1931 Political Geography as a Political Science Field. American Political Science Review 25: 439–442.

Sprout, Harold 1963 Geopolitical Hypotheses in Tech nological Perspective. World Politics 15:187–212.

Sprout, Harold; and Sprout, Margaret 1939 The Rise of American Naval Power. Princeton Univ. Press.

Sprout, Harold; and Sprout, Margaret (1946) 1960 Atlantic, Command of. Volume 2, pages 628-632 in Encyclopaedia Britannica. 14th ed. Chicago: Benton.

Sprout, Harold; and Sprout, Margaret (1960) 1964 Geography and International Politics in an Era of Revolutionary Change. Pages 34-51 in W. A. Douglas Jackson (editor), Politics and Geographic Relation ships. Englewood Cliffs, N.J.: Prentice-Hall.

Sprout, Harold; and Sprout, Margaret 1962 Foundations of International Politics. Princeton, N.J.: Van Nostrand.

Sprout, Harold; and Sprout, Margaret 1965 The Eco logical Perspective on Human Affairs, With Special Reference to International Politics. Princeton Univ. Press.

Spykman, Nicholas J. 1938 Geography and Foreign Policy. American Political Science Review 32:28-50, 213–236.

Spykman, Nicholas J. 1942 America’s Strategy in World Politics. New York: Harcourt.

Spykman, Nicholas J. 1944 The Geography of the Peace. New York: Harcourt.

Spykman, Nicholas J.; and Rollins, Abbie A. 1939 Geographical Objectives in Foreign Policy. American Political Science Review 33:391-412, 591–614.

Taylor, Thomas G. (editor) (1951) 1957 Geography in the Twentieth Century. 3d ed. New York: Philosophi cal Library.

Thomas, Franklin 1925 The Environmental Basis of Society. New York: Century.

Udall, Stewart L. 1963 The Quiet Crisis. New York: Holt.

Van Dyke, Vernon 1960 Political Science: A Philosophical Analysis. Stanford Univ. Press.

Van Valkenburg, Samuel; and Stotz, Carl L. (1954) 1955 Elements of Political Geography. Englewood Cliffs, N.J.: Prentice Hall. → Topically organized text.

Vevier, Charles 1960 American Continentalism: An Idea of Expansion, 1845–1910. American Historical Review 65:323–335.

Wagner, Philip 1960 The Human Use of the Earth. Glencoe, III.: Free Press.

Weigert, Hans W.; and Stefansson, Vilhjalmur (edi tors) 1944 Compass of the World. New York: Mac-millan.

Weigert, Hans W.; and Stefansson, Vilhjalmur (edi tors) 1949 New Compass of the World. New York: Macmillan.

Weigert, Hans W. et al. 1957 Principles of Political Geography. New York: Appleton. → Wide-ranging text by several authors, with strong emphasis on economic, demographic, and broadly cultural factors.

Wheeler, Raymond H. 1946 Climate and Human Behavior. Pages 78-87 in Philip Harriman (editor), Encyclopedia of Psychology. New York: Philosophical Library.

Whittlesey, Derwent 1935 The Impress of Effective Central Authority Upon the Landscape. Association of American Geographers, Annals 25:85–97.

Whittlesey, Derwent (1939) 1944 The Earth and the State. New York: Holt.

Wolfe, Roy I. 1963 Transportation and Politics. Princeton, N.J.: Van Nostrand.

Wolfe, Roy I. 1964 Perspective on Outdoor Recreation: A Bibliographical Survey. Geographical Review 54: 203–238.

III ECONOMIC GEOGRAPHY

The subject matter of economic geography is related substantively and historically to both disciplines from which the field receives its name. It obtains from geography an emphasis upon similarities and differences from area to area, large and small, on the earth’s surface, and upon linkages or circulations between areas. It acquires from economics an interest in the production, distribution, exchange, and consumption of goods and services. Economic geography therefore may be denned as an inquiry into similarities, differences, and link ages within and between areas in the production, exchange, transfer, and consumption of goods and services. Particular attention is given to the location of economic activity, considered both in theoretical and practical terms [seeSpatial economics].

Relationship to geography and economics. Economic geography is so intimately a part of the whole of geography that separating it from the general field is difficult. Because gaining a livelihood not only is essential to human existence but also involves a wide range of cultural and natural (physical and biological) features and interrelationships between those features, most matters of concern within economic geography are of concern within the over-all discipline, and vice versa. This is especially true if natural and noneconomic cultural features are considered in terms of their positive and negative implications for human use of earth space in gaining a livelihood. However, the emphasis upon livelihood in economic geogra phy does mean a corresponding reduction of attention to those cultural or natural conditions which may be only loosely related to spatial aspects of livelihood. Thus, for example, neither the cultural origin of a religious belief nor the process of land-form development is of direct interest to the economic geographer unless applicable in some way— as an advantage or constraint—to the location and interrelationships of economic activities.

As in the whole of geography, the spectrum of the totality of interrelationships in economic geography may be horizontal (involving different areas or different points within a given area), vertical (involving a morphological column of cultural and natural features at a specific point on the earth’s surface), or both horizontal and vertical. Hori zontal relationships are stressed in most work now being carried on.

We may consider economic geography, therefore, as emphasizing the livelihood aspects of the whole of geography, rather than as a compartment of the parent discipline. The field has a direct tie to economics and, by way of the whole of geog raphy, indirect ties to other disciplines in the social and natural sciences. So considered, the field by definition is sufficiently broad in scope to antici pate any methodological changes from time to time and place to place, although specific approaches, concepts, points of view and emphasis, immediate objectives, and methods have ranged rather widely with change of either time or place.

Trends. However, among economic geographers there is lack of agreement as to specific direction, particularly in the United States and Canada. For a time after its emergence as a separate field in the United States, during the early years of the twentieth century, economic geography relied primarily upon the inductive approach, with individual scholars aggregating ideas and data from field and library work into descriptions, classifications, and qualitative interpretations, utilizing numerical evidence when possible. Partially because of limitations on the amount of information obtainable in these ways, emphasis was placed on the unique or at least the distinctive features and interrelationships in both systematic and regional work; generalizations were made when possible. Although this approach continues to be utilized profitably by many economic geographers, the past decade has marked the emergence of a new school of thought, with immediate roots going as far back as the 1930s and indirect roots into the early portion of the nineteenth century. This school has chosen an explicitly theoretical approach, emphasizing nomothetic research and depending ap preciably upon mathematical abstraction. In the early 1960s the suggestion was made that geography basically is concerned with systems analysis (Ackerman 1963) and that the overriding problem of geography is understanding the man-land system of the earth. This giant system in turn is considered by Ackerman to comprise a large number of hierarchically arranged subsystems and processes. Such a concept places economic geography in the position of searching for laws involving livelihood within a context of systems analysis applied to earth space. The degree to which the concept has been generally accepted is not yet certain.

Two basic rationales. Like the over-all discipline, economic geography in the United States and Canada can be visualized in terms of two funda mental rationales—the topical, or systematic (not to be confused with system in systems analysis), and the regional. In economic geography both rationales focus on primary activities (here, as generally in economics, considered to be agriculture, grazing, forest-products industries, mining, fishing, and hunting), secondary activities (manufacturing and construction), and tertiary activities (all other occupations). The two rationales differ espe cially as to initial starting position. The topical, or systematic, rationale begins with structural aspects and works toward their earth-space expression and relationships, whereas the regional begins with space and works toward structure. The respective starting positions are usually reflected in various emphases in the completed works. Both rationales can be divided into subjects which themselves have become research interests for certain economic geographers. In practice these subjects are usually called systematic or topical specialties. Agricultural geography, manufacturing geography, and marketing geography have been of long-standing interest. Transportation geography and theoretical approaches to certain tertiary activities and domestic trading patterns have been developed actively within the past decade [seeCentral place]. In addition, increasing attention is being given to recreation geography and to aspects of tertiary ac tivities not yet accorded full consideration, to geographical aspects of international trade, and to the economic geography of primary activities other than agriculture. Work has begun on the geogra phy of consumption and the geography of price.

Because economic activities are earth based and may be clustered, as to location and/or function, into patterns of differing kinds and intensity, regions are important to economic geography. As in the whole of geography, regions in economic geography may be either homogeneous or nodal. The homogeneous region, sometimes called the formal region, is more or less an inventory of static features and relationships within an area that has been delimited on the basis of prevailing homo geneity of at least one feature. Both cultural and natural features and relationships may be so classified: the spring wheat belt, the manufacturing belt between Chicago and New York, and the Appa lachian Mountains are three examples in North America of this classification. Each example has been delimited on the basis of a single criterion, but multiple-criterion regions are possible at higher levels of generalization.

The nodal, or functional, region classifies hu man organization of earth space. This region has a point of focus (such as a city or town), an organ ized area of mutual interdependence with respect to that point and associated territory (such as the trading area of a city or town), and connecting lines to the territory (such as transportation and communication routes) providing the linkage (such as commuter and freight traffic and communications flows) between the point and the area.

Both the homogeneous and the nodal regions are here envisaged as means of classification and not as objective entities to be discovered in a scien tific sense. Either type of region can be considered at various levels of observation and detail. When several levels are superimposed, a pyramidlike framework with hierarchical tendencies may be recognizable.

Realms of interdisciplinary contact. Besides its subfields, economic geography extends into realms of interest shared with other disciplines. The uti lization and conservation of natural and human re sources is of long-standing interest to some economic geographers. More recently, the development of regional science is of definite interest to some economic geographers, many of whom participated in that development. Especially in the 1960s some economic geographers have become very interested in regional inequalities of economic development, whether within a country or at a continental or global level of observation.

Theoretical and practical implications. The theoretical importance and practical significance of economic geography are inextricably intertwined. Theories seek optimum circumstances and efficien cy in human utilization of earth space: what are the most desirable size, spacing, and intermix of specific economic units of production, exchange, transfer, and consumption within selected typolo gies of cultural and natural conditions? On the other hand, evaluations of historical and current practices indicate the degree to which theoretical models are actually applicable, especially in view of the cultural institutions, personalities and wills of key individuals, and specific natural conditions of a given area. Such evaluations also provide a degree and type of qualitative insight not obtainable through hypothesis alone.

Once a satisfactory relationship between theory and practice has been established, a logical step is application to planning procedures. In economic geography the value of both theoretical and prag matic study to planning is clear, whether the view point of the planner is the broad outlook of the regional analyst, concerned with an intermix of varied economic activities and resources, or the more restricted and highly specialized view of the expert in finding locations for individual units of economic activity.

Methods of study. Inasmuch as developments and trends in the United States from 1904 to 1954, the first 50 years of geography’s existence as a uni versity discipline, have been evaluated elsewhere (James & Jones 1954), the events of succeeding years will be emphasized here. A survey of the literature indicates that research since 1954 can be divided into three categories, on the basis of approach and method: qualitative interpretation, usually with substantial numerical evidence and sometimes making use of the case-study technique; quantitative classification, in a more or less de scriptive sense, with qualitative elaboration and explanation and involving a specific procedure applicable to different areas and time periods; and formulation and testing of specific hypotheses and models. These approaches have been applied, with varying degrees of intensity, to most facets of economic geography, but they will be discussed here with respect to agriculture, manufacturing, trade, and transportation.

Agricultural geography. All three approaches are utilized in agricultural geography, which is still a subject of keen interest. Books and articles in volving qualitative interpretation are diverse as to specific subject matter, but most can be classified under certain broad headings. Approximately one-third of the articles on agriculture appearing in four professional journals of the United States and Canada—Annals (Association of American Geog raphers), Canadian Geographer, Economic Geogra phy, Geographical Review—treated a specific agri cultural activity in a definite area, evaluating such aspects as type and size of enterprises, natural environmental advantages or constraints, combinations of selected crops and livestock with other crops and livestock, allocation of land, general farming practices, and trends. One-fourth empha sized over-all use of agricultural land in a specified area, considering other aspects of agricultural geography in a subsidiary way. Nearly one-fifth were concerned primarily with reclamation of agri cultural land—irrigation, drainage, erosion control, etc. A final one-fourth involved miscellaneous in terests, such as land redistribution in a given area, agricultural colonies of a minority group in a given area, land tenure in a specified place, the economic development of a country that is heavily dependent on agriculture, and the historical geography of agri cultural change in a selected location, etc. These four categories of qualitative interpretation, which cover the most numerous writings on agricultural geography in literature published in the United States and Canada, provide valuable insight into conditions evaluated by each author but have almost no common denominator.

Descriptive and analytic interpretation of a quantitative nature involves a classification based on numerical information, preferably official data continually available. The classification reveals pattern when plotted on a map and hence may be used to construct generic regions (typologies of regions) based on quantitative criteria. Once the criteria become standardized on a world-wide basis, the classifications will become standardized and applicable to all parts of the world. In principle both the homogeneous and the functional region may be so constructed, but in practice the homo geneous region has received the most attention to date in agricultural work. Prior to 1954 most studies were based on inadequate quantitative evi dence, rather highly generalized, and presented at continental or even global levels of observation. Weaver (1954, especially pp. 175-184) applied a greater measure of objectivity to an area of inter mediate size, aggregating data from county units to compute, for the Middle West of the United States, areal differences in crop combinations on the basis of degree of variance from a theoretical curve of optimal combinations. Subsequent work by other geographers includes a quantitative sampling approach to agricultural regionalization, the application of multiple correlation and regression analysis to rural farm population densities in the Great Plains, and the statistical association of cash grain farming in the Middle West with landforms.

The application of hypotheses to agricultural geography has been based on a rediscovery of im plications of Thiinen’s pioneer work treating the effect of transport cost and market price on crop and livestock combinations (1826–1863). Cur rent research suggests that, while the influence of distance to market on agricultural land use does not result today in patterns so simple as those set down by Thtinen, the influence of market price and transport cost does exist and can be analyzed mathematically [seeRent].

Manufacturing geography. Qualitative interpre tation of manufacturing geography, as indicated in publications of the United States and Canada, has been applied especially to areas other than Anglo-America, usually where numerical information is not fully available. Such an article or monograph may be an appraisal of a specific industry or group of industries or may involve a general examination of all manufacturing. The method of the historical geographer, providing the time di mension, may or may not be utilized. The case study is seldom used. Work involving only qualitative interpretation constitutes a relatively small percentage of all studies in manufacturing geog raphy, largely because numerical data are available to a greater degree than, for example, in agri culture.

Quantitative classifications are numerous in manufacturing geography. Labor force and value added are usually the criteria of measurement, although the list of possibilities is long and includes value of product, wages paid, amount of energy consumed, area of ground space, area of floor space, and land value. One such classification, with labor force used as the measuring criterion, has revealed structural and spatial changes in the manufacturing of the United States. Another has been based on magnitude (number of employees, wages paid, and value added) and intensity (ratios of labor force and value added to selected national totals), the classification being applied, with allow ance for kind and amount of data available, to the United States, Japan, and the Soviet Union. In still another classification based on labor force and value added, the author developed a technique for showing change over time on a single map and applied his technique not only to conditions within specific regions but also to differential rates of regional growth, as measured by national totals. Another study classified cities by labor force on the basis of prevailing industries, after first removing from consideration the city-serving functions of the industries. Several indices of industrial di versification also have been developed, with labor force the prime criterion of measurement. A typology of manufacturing flows relies, on the one hand, upon the orientation of manufacturing to raw materials or agglomerations of input factors and, on the other, upon access to both national and regional markets.

Increased attention is also being accorded theo retical approaches to manufacturing geography. Here, especially, work is being shared—by economic geography and regional science [seeRegional science]. Among initial geographical mod els was a construction by Harris (1954) showing the importance of market potential to industrial localization. Other geographers have used models to measure association tendencies in manufactur ing, concluding that agglomeration is a very important force. Models have been used to examine tendencies of high-value-added manufacturing to concentrate in certain areas, and they have also been used to associate circular and cumulative causation with the growth of manufacturing and associated urbanization. [SeeEconomies of scaleandExternal economies and diseconomies.]

Geography of trade. Attention to trade in economic geography has focused especially on domes tic trade. Theoretical work is being pursued vigor ously under the stimulus of central-place theory. Pragmatic work, whether qualitative interpretation or quantitative classification, also has been of keen interest. This pragmatic work is usually called mar keting geography, although some studies in domes tic commodity flow would not necessarily be in cluded under such a heading. Early research of this kind was associated especially with urban geogra phy, which developed as a field in the first half of the twentieth century and is becoming increasingly important. This early research has been a valuable antecedent to both central-place theory and mar keting geography. From numerous studies of indi vidual urban units and, subsequently, from use of census information three significant contributions were made: the idea of the cityregion, a functional region of interdependent units, including an urban unit; the classification of all cities in a country by relatively dominant functions, as measured by nationwide norms; and the concept of an economic base comprised, on the one hand, of trading and related activities or portions of activities which provide financial support for a specified area and, on the other, of activities or portions of activities which merely provide local interaction and have no influence outside that area (James & Jones [1954] 1964, pp. 142-166, 245-251). Subsequently, Murphy and Vance (1954) developed a technique for delimiting the central business district (CBD), based on land use involving the retailing of goods and services and the provision of office space. Ullman (1957) developed a series of maps in terpreting commodity flow within the United States, emphasizing principles of complementarity, transferability, and intervening opportunity. Other writers have been interested in the most appropriate location for shopping centers, especially with respect to market areas shared by competing firms. Questions have been raised as to the necessity of formal theories in marketing geography, but as yet no definite answers have emerged. Meanwhile, in a related area efforts are being made to map and evaluate the spatial distribution of finance as an economic activity and to associate this distribution with trading areas of cities.

Most geographical work in international trade has been pragmatic. One study has associated broad regional patterns (countries or groups of countries) and degree of dependence upon categories of exports and imports, and another has classified countries by degree of dependence upon exports. Still another has examined the free port, concluding that it may have outlived its usefulness in technically advanced, highly industrialized coun tries. Thoman and Conkling (1967) appraised national and bloc trends in international trade between 1938 and 1963, dependence of individual economies upon exports and export specialties, and logistics and mechanics of such trade.

Transportation geography. Although transpor tation long has been treated in the literature of economic geography, its full significance is only coming to be realized as the functional region, which depends for interpretation largely upon the flow of commodities and people to and from central places, comes to the forefront of attention. This is especially true of linkage studies involving transportation and trade—the carrier and route, plus shipping costs, plus direction and composition of commodity and passenger movement, plus alternative opportunities for such movement. Such link age studies reveal the dynamic aspects of an area, whether for a specified time or with change over time. Again, all three categories of approach are to be seen, with qualitative interpretation utilized particularly to present an unusual idea or to appraise general conditions over a wide area where adequate numerical information is lacking. Some geographers, however, have produced excellent qualitative interpretations concerning areas, small or large, that are covered rather fully by census and comparable data. In addition, historical geographers have provided insight into the development of somewhat analogous transportation routes that have evolved at different times and for different purposes.

Classifications by density of route and by general direction and function of route, again usually at continental or global levels of observation, are not new in transportation geography; but these have been augmented, usually at national or regional levels, by more-detailed and more-meaningful studies. Models have also been used to indicate the impact of highways on geographic change, to anticipate the development of transportation networks, and to explain differential rates of growth in passenger traffic between leading airports.

Regional economic geography. Several recent books and monographs have applied regional approaches to economic geography. Gottmann (1961) evaluated the urban-economic aspects of Megalopolis, an urbanized section of the Atlantic seaboard of the United States stretching from Boston south ward beyond Washington, D.C. Hance (1964) presented a regional examination of Africa, based on many years of study there. Camu and his associates (1964) produced a regional-systematic survey of the economic geography of Canada, introducing a multiple-criterion regional construct and several new ideas in regionalization, including the areal distribution of the total amount and types of capital investment. These and similar works have carried forward the continuing aspects of traditional economic geography as expressed regionally. In addition, as has been shown, many economic geographers have contributed to regional science and to regional economic development.

Other viewpoints. Economic geography is widely accepted outside the United States and Canada and generally is defined in terms already stated. The field is especially comprehensive in the Soviet Union, by definition virtually replacing human geography (Geograficheskoe Obshchestvo SSSR [1961] 1962, especially pp. 31-44). Under the stimuli of dialectical materialism and national development, economic and physical geographers of the Soviet Union have devoted close attention to pragmatic aspects of their respective subdivisions of the over-all discipline. A keen interest also exists in the Soviet Union in planning, particularly in the roles of theory and measurement. An emerging school of thought there holds that geography should not be compartmentalized but be considered as a unit—a view somewhat similar to Ackerman’s concept of the discipline as a man-land system.

Europe and the United Kingdom have continued a long-standing interest in economic geography. German geographers have built on the work of Thünen and Alfred Weber, fusing these studies with evaluations of management practices, enterprises, and land use to produce a well-founded inductive-deductive concept of the field (Otremba 1953). The term “applied geography” has come into use, particularly in France but also in other parts of Europe and, recently, in the International Geographical Union (Phlipponneau 1960). Although not limited to economic geography, applied geography stresses maximum efficiency in man’s use of earth space. In the United Kingdom economic geographers have bridged the gap rather smoothly between the deductive and inductive approaches, providing valuable and well-written reviews that take cognizance of historical development (Estall & Buchanan 1961; Chisholm 1962). Planning is of long-standing interest to British economic geographers, and recent work evinces growth of theoretical work (Haggett 1965). In Scandinavia, notably Sweden, theoretical approaches have been utilized actively for a long time, although pragmatic work continues.

There are many examples in other countries of application of the approaches already mentioned. Qualitative interpretation frequently is a detailed inventory of available resources under specified conditions and cutoff limits. Classifications are becoming more numerous, and there is some experimentation with hypotheses.

No one of the three approaches in economic geography necessarily is superior to the others, and more work is urgently needed in all. As additional data become available and increasingly standardized internationally, classifications and theories probably will become more numerous and will be produced mainly by committee or team efforts. Qualitative interpretations by individuals, however, always will be necessary—not only to provide unusual stimuli and insights but also to bring together cogently the threads of complex ideas, a result that cannot be expected from anthologies.

Richard S. Thoman

[See alsoCentral Place; Conservation; Regional Science; Transportation.]

BIBLIOGRAPHY

Ackerman, Edward A. 1963 Where Is a Research Frontier? Association of American Geographers, Annals 53:429–440. → A logical argument for the proc ess-system concept of geography as a nomothetic science.

Alexandersson, Gunnar; and NÕrstrÕm, GÕran 1963 World Shipping: An Economic Geography of Ports and Seaborne Trade. New York: Wiley. → A thorough assessment of ocean commerce and ports on a worldwide and regional basis.

Camu, Pierre; Weeks, E. P.; and Sametz, Z. W. 1964 Economic Geography of Canada. Toronto: Macmillan. → An intriguing systematic and regional survey, using a 68-region classification developed over a ten-year period.

Chatterjee, Shiba Prasad 1964 Fifty Years of Science in India: Progress of Geography. Calcutta: Indian Science Congress Association. → Contains a detailed bibliography.

Chisholm, Michael 1962 Rural Settlement and Land Use: An Essay in Location. London: Hutchinson’s University Library. → An able presentation of theoretical approaches to agricultural geography, viewed at different levels of observation and in cognizance of technical change.

Estall, R. C; and Buchanan, R. O. 1961 Industrial Activity and Economic Geography. London: Hutchinson’s University Library. → An excellent survey of selected theoretical and pragmatic considerations, in cluding government policy, in the location of industry.

Garrison, William L. 1959-1960 Spatial Structure of the Economy. Association of American Geographers, Annals 49:232-239, 471-482; 50:357–373. → An excellent review of trends and methods in theoretical economic geography.

Geograficheskoe Obshchestvo SSSR (1961) 1962 Soviet Geography: Accomplishments and Tasks. New York: American Geographical Society. → First published in Russian. A methodological statement by 56 leading Soviet geographers.

Ginsburg, Norton S. (editor) 1961 Atlas of Economic Development. Univ. of Chicago Press. A careful interdisciplinary effort to map economies by specified criteria and, through factor analysis, by combinations of those indices.

Gottmann, Jean (1961) 1964 Megalopolis: The Urbanized Northeastern Seaboard of the United States. Cambridge, Mass.: M.I.T. Press → An urban-economic evaluation of the nature of a continuous urbanized section of the eastern United States.

GrÖtewald, A. 1959 Von Thünen in Retrospect. English Geography 35:346–355.

Haggett, Peter (1965) 1966 Locational Analysis in Human Geography. New York: St. Martins.

Hance, William 1964 The Geography of Modern Africa. New York: Columbia Univ. Press.

Harris, Chauncy D. 1954 The Market as a Factor in the Localization of Industry in the United States. Association of American Geographers, Annals 44: 315–348. → An evaluation of the role of market potential in industrial location, especially as expressed in numbers of people and associated sales less shipping charges from specified central places.

James, Preston E.; and Jones, Clarence F. (editors) (1954) 1964 American Geography: Inventory and Prospect. Syracuse Univ. Press. → A survey of trends in geography during its first 50 years in the United States and of the status of geography at mid-century. See especially pages 3-68, 142-166, 240-332 and references in bibliographies to monographs by Richard Hartshorne.

Johnson, Hildegard B. 1962 A Note on Thünen’s Circles. Association of American Geographers, Annals 52:213–220.

Murphy, Raymond E.; and Vance, J. E. Jr. 1954 Delimiting the CBD. Economic Geography 30:189–222.

Otremba, Erich 1953 Allgemeine Agrar- und Industriegeographie. Stuttgart (Germany): Franckh’sche Verlagshandlung. → Provides a thorough review of the development and mid-1950 status of economic geography in Germany, especially the German Federal Republic.

Otremba, Erich 1957 Allgemeine Geographie des Welt-handels und des Weltverkehrs. Stuttgart (Germany): Franckh’sche Verlagshandlung.

Phlipponneau, Michel 1960 Géographie et action: Introduction à la géographie appliquée. Paris: Colin. → Emphasis on the need for study of the functional region, with attention to planning.

Taaffe, Edward J. 1962 The Urban Hierarchy: An Air Passenger Definition. Economic Geography 38:1–14. → An experiment in the use of a model to predict trends in air passenger traffic between major cities of the United States.

Thoman, Richard S.; and Conkling, Edgar C. 1967 Geography of International Trade. Englewood Cliffs, N.J.: Prentice-Hall. → A survey of characteristics and trends in world trade by global, regional, and national patterns, and of the logistics and mechanics involved in such trade.

ThÜnen, Johann H. von (1826–1863) 1930 Der iso-lierte Staat in Beziehung auf Landwirtschaft und Nationalökonomie. 3 vols. Jena (Germany): Fischer.

Ullman, Edward L. 1957 American Commodity Flow: A Geographical Interpretation of Rail and Water Traffic Based on Principles of Spatial Interchange. Seattle: Univ. of Washington Press.

Weaver, John C. 1954 Crop-combination Regions in the Middle West. Geographical Review 44:175–200. → A pioneer experiment in close measurement of crop combinations in terms of relative percentages of harvested land.

IV CULTURAL GEOGRAPHY

Cultural geography as treated here is peculiar to American geography and can be understood as a complement to certain of the trends in American geography in the early part of this century. Cultural geography in a broader sense deals with any part of man’s culture in the same way that plant geography deals with the distribution of plant species and vegetation or that economic geography is concerned with the production and distribution of goods and services. Cultural geography in the narrower sense used here is also characterized by certain cultural topics with which it deals, although its unifying thread is its manner of using the anthropological idea of culture to give meaning to its material. The tracing of continuity in space and time can help account for cultures and culture traits whose presence may not seem satisfactorily explained by their function in meeting overt ends. The subject matter of cultural geography has been winnowed by its need of such probing into origins. Cultural geography is not a self-sufficient field of study that produces all its own data and examines them as part of a closed system; it is rather an exchange in which data and interpretations from many sources are examined from one general point of view.

Development. Physiography was emphasized in the formative period of American geography, which occurred around the turn of the century. The first human geography then admitted inquiry into selected relationships between man and his physical environment. Later, concern with the productive capacities of the land came to be coupled with the growth of an economic geography that concentrated on production and trade. Economic geog raphers either assumed or looked for a functionalism or an adherence to economic laws that assured efficiency.

Carl Sauer (1925; 1931) outlined a new cultural geography dealing with those elements of material culture that give character to area through being “inscribed into the earth’s surface.” The focus was to be on those works of man rather than on man himself. The study was to be empirical and historical, without preference for environmental or any other selected class of explanation. The elements studied were to be broadly economic as well as material, although Sauer (1941) was later to expand their scope.

Sauer’s proposals produced, first, a general change in the direction of American geography and, later, the more special cultural geography, whose early growth was chiefly through his own students. This cultural geography has increasingly occupied the territory its name and rationale have staked out; its content has been limited chiefly by what other fields have previously claimed.

European geography contributes a great portion of the material of cultural geography, especially in dealing with Europe’s own rich heritage, but mostly under the heading of a general human geography or a less inclusive social geography. The models for Sauer’s program were heavily German, especially Friedrich Ratzel’s work on culture spread, Eduard Hahn’s work on agricultural development, and regional studies focused on settlement history. Kulturgeographie continues as a broad division of geography in Germany, where the modern idea of culture developed, but it does not parallel the American cultural geography as a specific hub for the swapping of ideas (National Research Council …1965). The absence of a recognized cultural geography in Britain and France is hardly surprising, for the culture concept is less used in those countries (the genre de vie of Vidal de la Blache is similar in use but much less inclusive).

Content. Most studies in cultural geography develop one or more of the following subjects:

The growth of man’s exploitation of his habitat. The study of man’s use of his environment includes such topics as the early use of tools and implements, domestication of plants and animals, and the various economies of food production. Human development and human invention both have geographic dimensions in that they are composed of specific events occurring in specific places. The geographer’s concern with both the spatial arrangement and the qualities of habitats qualifies him for taking part in the reconstruction of man’s past as well as in the understanding of the present. The cultural geographer’s interest in the past begins with the beginning of man and follows his wanderings with ever more cultural equipment into new surroundings (Sauer 1952).

Archeology, history, physical geography, and field observation provide much of the raw material for work in this field. Archeology and history, each limited by its sources of data, must remain incom plete records. Cultural geography often asks questions they are least likely to have answered. Thus, many early chapters in man’s growth can at best be speculative theories that may remain unverified.

Physical change induced on the surface of the earth by man. Man’s material advance leaves its mark on the earth he works. Physical change may be an inadvertent product of man’s use of the land: soil erosion induced by cultivation, soil enrichment around human habitations, and vegetation change induced by grazing or burning. It may be a deliberate change, such as the clearing of woods or terracing of hillsides for farming (Spencer & Hale 1961). The actions of man and nature have been of such duration and of such intermixture as to be often no longer separable.

American concern with man’s part in the processes of physical geography dates from George P. Marsh’s writing, in 1864, of Man and Nature. A more recent study, Man’s Role in Changing the Face of the Earth (International Symposium …1956), explores the earth as the imprint of man’s way of life, as a record of his past, and as a material resource for the present and future. The proc esses that explain the past may serve as guides for the future.

Settlement formsrural and urban. Settlement forms make up a large part of the features of the man-made landscape. Study of rural settlement has dealt mostly with house types, the arrangement of houses and other structures in relation to each other and to road networks, and the arrangement of fields. Although the settlement of the United States is recent, the history of its house forms is already partly lost, and tracing them is difficult (Kniffen 1965). The settlement patterns, more easily reconstructed, have converged on variations of the isolated farmstead. For most of Anglo-America the field pattern had to fit the rectangular survey; older areas show the confusion of metes and bounds, but traditional patterns survive, particularly in the old French riverine settlements. Reconstruction of the history of American settlement forms has not proceeded nearly so far as the inventorylike studies of rural patterns in Germany or of rural houses in Italy. Even where inventories are thorough, the forms have resisted satisfactory explanatory generalization.

American geographers have done relatively little with urban settlement forms. Enticing opportunities for study may be found in regionally dominant town and city plans, the pattern of street and lot layout, and cross-city comparison of house types, building materials, and architectural styles.

Nonmaterial culture, such as language and religion. Languages are considered the most reliable ethnic tracers, since the arbitrary choosing of words from a very large pool makes accidental repetition most improbable. Further, language, as a means and a mark of intercommunication, is a maker and a product of group cohesion and, hence, of cultures. Analysis of languages and their distributions has usually been the work of the specialist, but the results are widely used in geography. Toponyms are components of language that are very close to geography because of their fixture on the land and their frequent reference to its qualities.

Religion, also a conservative marker of peoples, is a social institution with significant spatial structure and a molder of the cultural landscape. Geography deals with religion in ways that range from the distribution of specific religions to the expression of a primal sense of order in the landscape—for example, orientation of streets or property lines with the compass (Isaac 1965).

Origin and spread of cultures and civilizations. The spread of culture is most simply studied through particular culture traits, but often an entire complex of culture traits may be welded to gether by a powerful or influential people and spread over a wide area in relatively uniform fashion. A way of living or a civilization may be traced from its inception to its expansion into a greater territorial base, until finally it reaches its limits and is absorbed or replaced by another expanding culture. The growth of potamic civilizations from their home in Mesopotamia, the growth of the Chinese nation from its culture hearth in the Wei Valley, or the grafting of the Marxist politico-economic complex onto a variety of cultural trunks in the past fifty years would be suitable subjects for such analysis.

Geographers have treated cultures in a variety of ways, such as dividing the world into major culture regions (Russell & Kniffen 1951), examining the development of cultures and subcultures, mapping the core and fringe areas of cultures (Meinig 1965), and studying culture islands.

Cultural evaluation of the environment. The favorite theories of geography are often themselves generalized evaluations of environment. Their change with the passage of time is evidence that they too, however reasoned, are part and parcel of changing culture. What man does with his natural resources depends on his technology, on his perception of his natural resources and of his place among them, and on a complex of values concerning the present and future. Clarence Glacken (1956) has studied man’s place in nature from the viewpoint of changes in Western ideas about the habitable world. Non-Western conceptions, as shaped by cosmologies, modified by experience, and revealed in language, are also essential to interpretation of living patterns (Lowenthal 1961).

The environment may be graded in aesthetic terms. Modification of the landscape, including the productive portion of the landscape, may then be guided by aesthetic senses and axioms. Yet the perception of the environment and the responses to environmental stimuli are probably not purely cultural. Sonnenfeld (1965) has asserted the need for isolating the noncultural parts of the behavior that relate to the environment and effect the shaping of the landscape. To do so would better sort the variables with which the cultural geographer deals and would clarify his ideas of causality.

Purposes. Many studies are undertaken to provide factual answers to specific questions. Some studies are primarily in the geographical tradition of exploration; these include many of the regional inquiries in cultural geography (e.g., Wagner 1958) as well as more specific data-collecting trips (e.g., Zelinsky 1958). Other studies seek links in the solution of specified larger problems, be they geographic in nature or otherwise. A major theme of Kniffen’s work has been the use of a particular culture trait as an index to migration and diffusion and the culture regions they shape.

H. C. Brookfield (1964), in a thoughtful critique of American cultural geography, noted its frequent reluctance to compare, generalize, or explain, especially if doing so meant going into social organization or social attitudes. American cultural geog raphers have often preferred to focus their immediate interest on filling selected gaps in knowledge rather than on using the material gained primarily as a means to further the abstract concepts of the field. There are advantages in developing a reliable body of elaborated description relatively free of the data selection that would be suitable for testing prechosen abstractions. However, cultural geography should also work ahead with generalization as fast as the data and advance of theory permit. Some division of labor may optimize the exchange between description and abstraction.

Purposes of scholarly study are often indeterminate. Idle curiosity may impel the investigator, even while specific hope of practical application accounts for his financial support. Many of the world’s problems are not abstract generalizations but are stresses that arise from unique local conditions. A geography that views every place as sui generis is a likely source of help. The theme of man’s use of his environment, enmeshed in most of cultural geography, gives geography its widest views and most likely application on a world scale. Certainly the cultural geographer is not much drawn to his study by the hope of immediate application, even though it is with past and present application that he constantly deals. Rather he is lured on by the prospect of a better articulated view of man’s work and works in their terrestrial frame.

Methods. Studies of small areas are most likely to depend for their data on field observation. Those on a world-wide scale necessarily depend on secondary sources. Spatial and chronological analysis of the data often form the logical core of the study. The former seeks simplification through the demonstration of spatial order or pattern; the latter emphasizes change in historical depth or the more detailed sequence of change known as process.

Spatial arrangement of data is an essential characteristic of geography. The map is the visual means of arranging the data in its spatial order. (To spread out or to unfold in this manner approximates the literal meanings of explanare and explicare, bases for the verb explain in English and the modern Latin languages). Mapping the distribution of a culture trait or complex does not con stitute an explanation, but it leads to hypotheses as to how the culture trait developed. One common method of map interpretation is to seek correlative distributions that may also be causal associations. A second method is to treat the distribution as a changing product of diffusion and extinction and to seek the conditions that have governed the changes (Zelinsky 1958; Spencer & Hale 1961). An isochronic map is useful in combining both temporal and spatial analysis.

Chronological arrangement of data is essential to the study of culture. The conservative nature of culture, which follows earlier models even in the process of change, encourages one to study it by tracing its continuity. If a device is not demonstrably functional in every aspect of its design, it may be explained genetically by tracing its origins and movements. John Leighly emphasized this type of cultural explanation in his proposal for a study of the tangible works of man in the landscape in the terms of art history, stressing “the essential time-bond of culture rather than its looser place-bond” (1937, p. 135). The culture concept provides a means of giving intelligibility to what, at least to people removed from the particular cultural context, seems irrational.

Plumbing the basic reasons for cultural choices may provide a functional explanation of what was once considered irrational. Cultural geography would become less dependent on genetic explanation as its mainstay to the extent that culture change—or lack of it—could be explained in terms of human satisfactions. Predictions of the direction of culture change will lead to more application of the findings of cultural geography. Any theory of culture change could have a corresponding theory of cultural geography as the spatial expression of the change. Uses of the culture concept in regional study by geographers and anthropologists have been considered by Thomas (1957).

A battery of general methods is summed up by Wagner and Mikesell (1962, p. 24): “Who? Where? What? When? and How? The themes of culture, culture area, cultural landscape, culture history, and cultural ecology respond to these queries.” The themes imply cross sections of investigation, most of them with conceptual dimensions rather than geometrical ones as in the case of the map. Since cultural geography deals with such a great range of phenomena, its methods must also be varied.

Persistent questions. The objects of study in cultural geography may be seen as forms—abstract or concrete, single or complex. Many of the general questions of cultural geography hinge on the derivation of these forms, which is variously sought in environment, function, ideology, technology, ornamentation, previous forms, and the chances of invention and accident. The derivation of one form does not necessarily serve as a model for the derivation of others.

The role of the physical environment. Cultural geography can deal with man’s relations with his physical environment at any depth of understanding of this environment. Sauer’s work with early man leaned heavily on assessments of environmental change and of environmental opportunity to supply gaps in the a posteriori record. On the other hand, the continuity of culture can be traced over the earth without attention to the physical environments through which the culture spread, although to do so is to ignore a part of the possible explanation.

The reaction against environmental determinism (the doctrine that the physical environment determines the way in which man lives in an area) has united with a culturally oriented geography to produce a new geographic etiology that stresses cultural determinism (a doctrine that emphasizes the role of culture as opposed to that of the environment). One statement of cultural determinism is that culture determines what the environment means to man. An even stronger position holds that man’s perception of the environment is all that matters about the physical environment; since perception is culturally controlled, explanation of human behavior is then deemed to be cultural. Cultural determinism presents useful views of the continuity of culture growth, but is less successful as an analysis of cause (a precursor without which the result would have been different). One must remember, first, that perception, by its definition, is not merely hallucination (i.e., independent of external stimulus) and, second, that the environment may not respond to man’s management in just the way his perception of it orders.

The difficulties of the man-nature and culture-environment dichotomies are sometimes dissolved in union. Man is often conveniently viewed as a part of nature. Man, culture, and environment in a land have been treated as one in sketches of regional “personality.”

Form versus function. What part does function play in cultural design and what part do previous forms play? How fast do forms change to conform to changing needs? Are forms also molded by aesthetic considerations? A barn is built for certain purposes, and a given design is likely to be retained only so long as it fulfills its function reasonably well. But the differences between two barns performing the same functions must be explained on some other basis. Even if the differing features mirror cultural traditions, they may still have originated in functions of other times and other places. Purely decorative features, too, sometimes originate in function. Relic forms may continue to be built or may survive in structures that have out lived their original functions. Whether function or form is stressed in any particular comparison may depend on whether the similarities or the differences are sought.

Form and ideals. A simple feature in the landscape, say a post, serves its function and reflects some particular post-making tradition. It probably has no particular relation to the ethos or set of ideals of the people who use it. A more complicated form with a more complicated function, say the physical form of the village or community built to facilitate a way of living, is more likely to reflect the distinctiveness of that way of living or of the ideals that lie behind it. An ascetic people, not given to social intercourse, probably would not provide their town with plazas and esplanades or decorate it in bright colors. The few geographic studies that deal specifically with form in relation to ideology suggest that rather different peoples may adopt the same readily available forms for the same overt purposes. An account of a Dutch Reformed settlement in Michigan (Bjorklund 1964) tells how immigrants largely gave up their old forms in favor of the common American forms, but it also shows how the American forms still conformed to the old ideology. If identical forms conform to different ideologies, we must assume that the differing peo ples see the forms as fulfilling different inner functions and gain different satisfactions from them. In another study Philip Wagner concludes: “Nicoya suggests that thoroughgoing transformations in the social sphere may sometimes produce only moderate variations in technology and landscape, and thus that two or more very dissimilar societies may differ little in the way they conceive and utilize a given habitat” (1958, p. 248). Conversely, peoples with similar ideologies are likely to evolve different community forms for their living. Inquiry into the regulation of form by ideals should be carried further.

Forms of distributions. The forms of distributions are also subject to systematization. The significance of continuous and discontinuous distributions, the relation of area of origin to area of greatest intensity, or the persistence of the core area of a culture region are all aspects of the structure or form of culture areas.

Relation to other fields. Historical geography is the division of geography most closely related to cultural geography. The two complement each other in a cultural-historical geography, in which history provides the explanation of culture and culture provides the organizing concept for the subjects of most geographic interest in history. Historical geography can stand alone, however, for it is often not organized around the culture concept. On the other hand, culture is always dependent on the past for its explanation.

Economic geography and cultural geography could go separate ways as long as one depended primarily on economic laws for its explanation while the other depended on the patterns of the past. But an entire system of economic laws may be found to be a culture complex that has evolved somewhat accidentally and not in an inescapable mesh of cosmic law. Even in our market economy the prices bid for goods may be but expressions of cultural preference (for example, the American preference for corn-fed ham versus the European for barley-fed). And within a changing economy new institutions, products, and types of enterprise originate and diffuse in a fashion suggestive of the culture traits that they are. On the other hand, the relevant economic laws are also necessary parts of the cultural explanation. Thus, economic and cultural geography join in any broad view of the two fields.

Political geography, like economic geography, is supported by an independent discipline that has its own laws (although less precise), and its systems fulfill stated functions. A nation represents an idea that had an origin and has been spread with some show of power to the borders of the land. The generic idea of nation is again a cultural concept that had an origin and a diffusion that just now seems about to complete the circuit of the earth. Each nation depends on a community of interests, often including such culture traits as a national language and religion. The behavior of its voters, too, reflects persistent patterns in its regional culture.

Cities have both economic and political dimensions, although the economic has loomed larger in urban geographical studies. City planning and urban sociology are cognate disciplines that help provide the rationale for an independent urban geography. At the same time, cities are concentrations of culture that are very sensitive to cultural differences. Internally, the distribution of subcultures within a complex city and the city landscape are equally concerns of cultural geography. Externally, urban functions and rural-urban attitudes vary from culture to culture.

The mathematical formulations now popular in economic and urban studies will be most successful if they can describe the regularities and implications of cultural behavior. Swedish studies have originated quantitative analysis of culture transmission and have attempted to simulate both its rational and random qualities.

The idea of culture provides a frame into which man-made functional systems fit. The narrower cultural geography and the fields that deal with the functional systems that are also a part of culture can fuse into a broader cultural geography in which the culture concept constantly insists on a proper relativity in time and space. Grasp of the concepts of culture and cultures is perhaps the quickest route to a viewpoint not entirely bound by one’s own culture and from which one can even see one’s own culture in some perspective. Culture is a fit mediator in a study whose point of departure is the comparison of different peoples and lands.

Edward T. Price

[Other relevant material may be found inCartography; Culture; Environmentalism; Landscape; and in the biographies ofMarsh; Ratzel; Sauer; Vldal de la Blache.]

BIBLIOGRAPHY

Bjorklund, Elaine M. 1964 Ideology and Culture Exemplified in Southwestern Michigan. Association of American Geographers, Annals 54:227–241.

Brookfield, H. C. 1964 Questions on the Human Frontiers of Geography. Economic Geography 40:283–303.

Glacken, Clarence J. 1956 Changing Ideas of the Habitable World. Pages 70-92 in International Symposium on Man’s Role in Changing the Face of the Earth, Princeton, N.J., 1955, Man’s Role in Changing the Face of the Earth. Edited by William L. Thomas et al. Univ. of Chicago Press.

International Symposium on Man’s Role in Changing the Face of the Earth, Princeton, N.J., 1955 1956 Man’s Role in Changing the Face of the Earth. Edited by William L. Thomas et al. Univ. of Chicago Press.

Isaac, Erich 1965 Religious Geography and the Geography of Religion. Pages 1-14 in Man and Earth. Series in Earth Sciences, No. 3. Boulder: Univ. of Colorado.

Kniffen, Fred 1965 Folk Housing: Key to Diffusion. Association of American Geographers, Annals 55: 549–577.

Leighly, John B. 1937 Some Comments on Contemporary Geographic Method. Association of American Geographers, Annals 27:125–141.

Lowenthal, David 1961 Geography, Experience, and Imagination: Towards a Geographical Epistemology. Association of American Geographers, Annals 51:241–260.

Marsh, George P. (1864) 1965 Man and Nature: Or, Physical Geography as Modified by Human Action. Edited by David Lowenthal. Cambridge, Mass.: Harvard Univ. Press.

Meinig, D. W. 1965 The Mormon Culture Region: Strategies and Patterns in the Geography of the American West, 1847–1964. Association of American Geog raphers, Annals 55:191–220.

National Research Council, Ad Hoc Committee on Geography 1965 The Science of Geography: Report. National Research Council Publication No. 1277. Washington: National Academy of Sciences-National Research Council.

Russell, Richard; and Kniffen, Fred B. 1951 Culture Worlds. New York: Macmillan.

Sauer, Carl O. (1915–1962) 1963 Land and Life: A Selection From the Writings of Carl Ortwin Sauer. Edited by John Leighly. Berkeley: Univ. of California Press.

Sauer, Carl O. (1925) 1963 The Morphology of Landscape. Pages 315-350 in Carl O. Sauer, Land and Life: A Selection From the Writings of Carl Ortwin Sauer. Berkeley: Univ. of California Press.

Sauer, Carl O. (1931) 1962 Cultural Geography. Pages 30-34 in Philip L. Wagner and Marvin W. Mikesell (editors), Readings in Cultural Geography. Univ. of Chicago Press. → First published in Volume 6 of the Encyclopaedia of the Social Sciences.

Sauer, Carl O. (1941) 1963 Foreword to Historical Geography. Pages 351-379 in Carl O. Sauer, Land and Life: A Selection From the Writings of Carl Ortwin Sauer. Berkeley: Univ. of California Press. → First published in Volume 31 of the Association of American Geographers, Annals.

Sauer, Carl O. 1952 Agricultural Origins and Dispersals. New York: American Geographical Society.

Sonnenfeld, Joseph 1965 A Behavioral Approach to Cultural Geography. Pages 10-18 in Discussion Papers in Cultural Geography. Unpublished manuscript. → Prepared for the 61st annual meeting of the Association of American Geographers, Columbus, Ohio.

Spencer, J. E.; and Hale, G. A. 1961 The Origin, Nature, and Distribution of Agricultural Terracing. Pacific Viewpoint 2:1–40.

Thomas, William L. Jr. 1957 Land, Man and Culture in Mainland Southeast Asia. Glen Rock, N.J.: Privately published.

Wagner, Philip L. 1958 Nicoya: A Cultural Geography. University of California Publications in Geography, Vol. 12. Berkeley: Univ. of California Press.

Wagner, Philip L.; and Mikesell, Marvin W. (editors) 1962 Readings in Cultural Geography. Univ. of Chicago Press.

Zelinsky, Wilbur 1958 The New England Connecting Barn. Geographical Review 48:540–553.

V SOCIAL GEOGRAPHY

No generally accepted definition of social geography exists. The variety of literature which has appeared under the title of social geography is astounding; even within particular schools there are wide disparities of approach and definition. With some notable exceptions, for example, in Sweden and Holland, social geography can be considered a field created and cultivated by a number of individual scholars rather than an academic tradition built up within particular schools. Further more, for many people the term “social geography” itself is in disfavor because of its past association with various forms of determinism that postulated a causal connection between society and the geographical environment.

Perhaps, therefore, the best way to examine social geography is to establish a general theoretical outline of the field and, on this basis, to review the existing literature. Naturally, many of the works relevant to what is here called social geography will have been written as contributions to some other discipline.

The argument that social geography is a necessary discipline can be made in at least two ways. One is by analogy with other, better established branches of geography. A widely accepted definition of “human geography” is that it deals with mankind in the context of his total geographical milieu. For the purposes of analysis this milieu has been subdivided into separate categories corre sponding to various orders of human activity, for example, the economic, the political, and the cultural. Therefore, one could postulate that social geography is the subdivision of geography that deals specifically with the social order, or that it is the systematic study of the social dimension in areal differentiation.

An alternative way is to begin with the definition of geography as the study of similarities and contrasts between places on the face of the earth. Society, that is, social organization and values, patterns of social movement and interaction, and social dynamics and change, plays such an important role in producing similarities and contrasts between places on the earth that it justifies syste matic consideration within the discipline.

The question immediately arises as to how to isolate this social dimension for independent study. In fact, since human activities characteristically are group activities, how can human geography be anything else but social? The virtually interchangeable use of the terms “human” and “social” by several geographers in the British and Dutch schools serves to emphasize the logical (and etymological) basis for this question. Yet, although in the evolution of human geography emphasis has been placed in varying degrees on purely social elements—and although languages, races, and religions have rarely been excluded from consideration—the function of these social elements in the total conceptual frame work has not been very clear. In fact, the idea that such social elements could be systematized into a general framework for geographical analysis has been only recently proposed (Bobek 1959; van Paassen 1965).

There are two primary questions social geography must answer: How do mankind’s social characteristics vary through space? How do these characteristics affect (or reflect) man’s adaptation to and adaptation of his total geographical milieu? Since such questions touch every aspect of human geography, it is difficult to conceive of social geography as a separate field. Its distinctive feature would thus appear to lie more in its focus and objectives than in any clearly delineated subject matter. In practical terms, the traditional twofold method of geography can be applied to these central questions in the following way: by the exam ination of spatial variations in the distribution and interaction of social groups within their total geographical milieus and by the examination of differential patterns of society’s use of the earth, as indicated in settlement forms, livelihoods, circulation networks, and land use patterns. While the first method implies a morphological or formal study of world social patterns, the second method implies a functional interpretation of such patterns in terms of their underlying social processes.

Having thus outlined, in broad terms, the place and function of social geography, let us now see how these fundamental questions have been studied in the past. From such a general and necessarily eclectic survey we may discern some of the major conceptual and technical ingredients from which a definition of social geography can be formulated.

The development of social geography

Studies explicitly or implicitly directed toward the exploration of social geography can be considered under two broad headings: first, the his torical precedents, which fall roughly into three major stages, each one characterized by a different approach; and second, the works of twentieth-century geographers.

Historical precedents

Descriptive reports written by explorers and men of letters during classical times, for example, the writings of Herodotus, Thucydides, Strabo, and others, provide the first written recognition of world social differences. Such encyclopedic descriptions continued to appear intermittently in the Occident up to the seventeenth century, for example, the accounts of Marco Polo and the lettres edifiantes of Jesuit missionaries. The twofold implication of these works was that social life takes various forms in different parts of the world and that these differences are caused by, or at least are associated with, differences in the physical—particularly climatic—environment.

A second phase consisted of the various philosophical reflections on these and later geographical discoveries. On the one hand, speculative thinkers sought normative principles for an ideal social order from natural law, and, on the other hand, the positivists insisted that such principles should be sought in the existing and empirically observable conditions of society. The essential message of this second phase was that there is a rational order in world society and that this order can be discovered deductively (speculative approach) or inductively (positivistic approach).

A third and far more significant phase began in the nineteenth century, accompanied in France by the emergence of the idea of democracy, in Ger many by the rise of national consciousness, and elsewhere by the slow yet effective permeation of a “scientific” approach to knowledge. Ethnographers and historians were among the first to study world social variations in a systematic way. As early as 1725 Giambattista Vico suggested that human development followed an identical series of stages and that the actual variations in world society at any particular time were due to their differential positions within that series. Later in the eighteenth century, Johann Gottfried Herder in Germany and Condorcet in France expressed similar ideas. The geographer Johann Georg Kohl examined the social function and significance of various settlement types; later, his colleague E. Hahn (1896) studied the evolution of livelihoods and demonstrated the religious and social origins of some economic prac tices. Yet this “scientific” approach to the study of mankind’s social differences was also associated with exaggerated single-factor explanations, for example, the biological interpretation first expounded by A. Schäffle (1875–1878) and the psychological interpretation, which found its fullest expression in the Durkheimian school in France. Friedrich Ratzel’s Anthropogeographie (1882–1891) incorpo rated both these elements: the ecological view of society within its natural environment and the role of human intelligence (the “idea”) in enabling man to overcome physical barriers (1901). Unfortunately, the latter perspective did not emerge too clearly in his monumental work—on which the whole tradition of anthropogeography has been pat terned—and so his name has been linked with the idea of society being determined by the physical environment. His Politische Geographie (1897) and some articles (1876; 1901) in fact contained hypotheses that were far more relevant to social geography than the Anthropogeographie.

One of the most significant precedents to social geography in the nineteenth century was the work of Frederic Le Play. Disdainful of the various a priori explanations of society prevalent in his day, he set out to study the actual social conditions of worker families in France. His famous mono graph technique produced an encyclopedic inventory of social facts, and from a great number of studies he deduced certain basic types, which then served as bases for comparison. Traces of Le Play’s analytical formula lieu-travail-famille, later adapt ed by Geddes (1915) into the formula “place-work-folk,” can be found in the writings of such early British geographers as H. J. Fleure (1918). French geographers inherited important elements from Le Play, for example, the monograph technique in empirical field studies, but the most important legacy of lieu-travail-famille was the social survey movement, which flourished in Britain and America during the early part of the century.

Many geographers, such as Ritter, von Humboldt, Hassinger, Ruhl, and Hettner in Germany, Reclus in France, George Perkins Marsh in America, and H. J. Mackinder in Britain, deserve recog nition as pioneers of social geography. However, the three major channels of thought that contained the most useful concepts were those initiated by Le Play (the social survey movement), Ratzel (anthropogeography), and Durkheim (social mor phology).

Twentieth-century social geographers

The mutual relations of society and environment was a subject that aroused great speculation and interest at the turn of the century. Yet there was no discipline equipped to embrace the entire question. Ratzel had made an abortive attempt to do so, and his environmentalist disciples exaggerated rather than corrected the deterministic premise of an thropogeography. Many scholars, particularly the Durkheimian sociologists, remained unconvinced that geography had any right to entertain such a monumental task.

At this juncture came one of geography’s great est entrepreneurs, Paul Vidal de la Blache. Society for Vidal (1896; 1902) and his school could not be explained entirely in terms of biological, psychological, or environmental interpretations. It was rather an intricate network of ideas and bonds that provided stability and orientation to human life within particular geographical milieus. In his classical studies of the Mediterranean world and of monsoon Asia (1917–1918), Vidal demonstrated the complex, yet harmoniously balanced, interplay between human institutions and particular natural settings. Genres de vie (literally, patterns of living) were the concrete expressions of a society’s ongoing contact with nature: sets of techniques, cemented through tradition, whereby human groups secured the material necessities of life within a functional social order (Vidal 1911; Sorre 1948). Repeated experiences in meeting life’s common problems within a particular geographical milieu occasioned the development of community consciousness, which made a genre de vie truly an ecological system. Variations of this basic concept appear in the literature of other disciplines, for example, social anthropology (Kroeber & Kluckhohn 1952; Redfield 1955), American human ecology (McKenzie 1934) and urban sociology (Park & Burgess 1921). By means of genre de vie and other concepts, the French school of human geography replaced the exaggerated Ratzelian notions of environmental de terminism by the more elastic concepts of possi-bilism and dismissed the charges made in the Annee sociologique between 1890 and 1910 more by substantive works than by theoretical argu ments. “La géographie humaine,” thus formulated, was a social geography in the broad, integral sense: all other dimensions of the human milieu were studied from the vantage point of society. Many British and American human geographers followed almost identical lines, while the Dutch “sociale geografie” was the direct equivalent of the French “géographie humaine.” The kernel of this orientation, namely, society as the source and framework for all human activity, reappears in the work of Hans Bobek (1959) in Vienna. Lucien Febvre’s famous apologia (1922) articulated the philosophical and historical raison d’etre of such a discipline.

To Vidal’s essentially ecological approach, his disciple Jean Brunhes added the important dimension of group psychology, asking, for example, why similar environments were used in entirely different ways at different periods in history. He defined social geography as the third level of complexity in human geography’s fourfold structure. The fourfold structure included the primary groups of family, kin, and culture; the secondary groupings of livelihood and special interest; the various forms of spatial interaction within and among these groups; and, finally, the legal systems which institutionalize a society’s subdivision and access to land and property ([1910] 1924, pp. 36-46). This definition, admirably suited to the study of European—particu larly French—rural society of the early twentieth century, remained the basic framework for social geography among British, French, and Dutch scholars up to World War II. Most of the early studies in social geography were regional in character, and their excellence consisted more in their artistic cohesion and integrative descriptions than in their analytical or theoretical expertise. The empirical conditions which favored the use of the regional framework by French scholars did not exist to the same extent elsewhere; this partly explains the divergence of orientation and method which developed among the various schools of human geog raphy.

During the 1930s, British social geographers were involved in methodological controversy. Does social geography consist in merely mapping mankind’s social characteristics, or must it also analyze the processes involved in relating a society to its geographical environment? What is the relation between social geography and human ecology? Why not replace the term “human” by “social” as the generic term to signify all the nonphysical aspects of geography?

The fundamental dichotomy between a formal and functional approach expressed in this British debate reiterated the duality that had developed in Holland since the 1920s. While at Utrecht the study of social groups within their territorial framework (de Vooys 1950) was being pursued along the lines of the French school, at Amsterdam Steinmetz’ “sociography” was used to study the entire social content of space as a system in itself —aside from any considerations of a group’s relation to its natural environment. The birth of sociology in Holland—particularly rural sociology as a separate discipline—has no doubt modified the original disciplinary orientations of these two schools (van Paassen 1965).

Prior to World War II little attempt was made to systematize the elements of social geography. In general, the important associations evident in the spatial organization of society—particularly in the United States—appeared in the literature of human ecology (Theodorson 1961) and urban sociology (Park & Burgess 1921). One major exception, of course, was the work of the environmentalists in examining connections between human behavior and the geographical environment (Thomas 1925).

Pierre George and Maximilien Sorre (1943–1953) were the first great systematizers of social geography. In George’s works a close link is maintained between social and economic aspects of hu man behavior, the social being one facet of the economic (1946, p. 1). For Sorre (1948, pp. 13-16, 66-122) society represented a system of techniques—family and kinship systems, livelihoods (genres de vie), languages, and religions, each one having a specific influence on the spatial organization of mankind and his work. Sorre’s schema does not make clear, however, whether social geography consists of a series of systematic subfields based on these various kinds of techniques, or whether a distinction is to be made between the “social” and “political” techniques. In his work all forms of organization from family and kin groups to giant political blocs form a continuum (1961, pp. 211-264). Gourou’s more comprehensive concept of civilisation (1964) comprises both material techniques (modes of production) and spiritual techniques (ideas, values). These three approaches at generalization are important because they try to maintain the integral and holistic character of social geography at the same time that they establish some order and a basis for comparative work. Bobek has made a similar attempt to construct a spatiotemporal framework for world society (1959). His work is a fertile synthesis of French and German traditions: his systematic framework is based on a holistic approach involving types of societies defined in terms of their actual use of their geographical environment (1961).

Several other attempts to formulate the problem of society in geography in terms of a particular systematic framework have appeared: for example, those of Wagner (1960), Ackerman (1963), and van Paassen (1965). Yet more characteristic of postwar work is the development of individual systematic lines of enquiry, for example, geography of rural and urban life, population studies, and geography of religions and political behavior. Associated with this is a more lively va-et-vient between geographers and scholars in other disciplines, particularly concerning questions of rural and urban life (Friedmann 1953) and regional planning (Phlipponneau 1960). Studies are still being made within a regional framework, but the focus has changed. Juillard in Alsace (1953) studied particular social problems from a regional perspective, while Rochefort in Sicily (1961) studied regional life from the perspective of the social processes at work. Such reorientations have, of course, raised new methodological problems and prospects. Chatelain (1947; 1953), for example, postulates a duality between the geography of social classes (a kind of social morphology) and the geography of social life (a sociological geography). Claval (1964) envisions the latter as the most feasible future direction for the discipline, citing the work of W. Hartke at Munich as an example. It is difficult, however, to see how these two aspects of the field can be separated.

To label the research being done at Munich as sociological geography may be misleading. Certainly the perspective is social: social-geographic differentiation (sozialgeographische Differenzierung) implies that social values—as expressed in the occupational structure—are the primary agents of landscape differentiation. Thus, maps of socioprofessional structure (Sozialkartierung) for a series of periods are collated with a corresponding series of land-use patterns (Nutzfiachenkartie-rung), and significant associations are sought. This basic formula has been applied successfully both in rural and urban contexts. Geipel’s study (1952) of one German region, for example, demonstrated that the sources of regional unity—which varied at different periods—are found essentially in the collective decision-making mechanism of the regional community. This is quite a contrast to the sources of regional unity commonly sought in the natural (physiographic) or economic (agricultural) landscape. Hartke (1956) demonstrated that regions where this phenomenon existed had similar geographic (regional) characteristics. Associations found in urban studies are even more interesting. Hartke’s intraurban corridors (Passagen) suggest some qualifications to the traditional concentric zone and sector theories of urban structure, while his study of urban expansion patterns provides new bases for the classification of cities (1961).

In marked contrast to the inductive, empirical, and microscopic approach of the Munich social geographers is the more highly developed theoretical and deductive approach found in Sweden. Torsten Hägerstrand (1952) and Sven Godlund (1956) have applied refined mathematical techniques to the study of migration, rural-urban interaction, circulation, and other dynamic aspects of the field. One of the most interesting developments has been the use of simulation models for the analysis and prediction of spatial movement.

This approach has been adopted and modified in the postwar period by a number of American geographers. At Iowa spatial models have been used to study the distribution patterns of schools, churches, and settlements, often with a view to spatial planning. Morrill’s study of Swedish towns (1963) exemplifies this approach. Yet, in general, social geography in the United States is not a unified field: on the one hand there are holistic regional studies, for example, Platt’s Saarland study (1961) and Broek’s southeast Asian study (1944), and, on the other hand, there are a growing number of systematic studies in racial, linguistic, religious, and other spheres. Some interesting associations have been elaborated, for example, between religion, land use, and livelihood (Isaac 1959), between cultural pluralism and political integration (Lowenthal 1961), and between migration and political behavior within ethnic groups (Lewis 1965). However, the exciting developments in the actual social geography of America have been treated mainly by foreigners (Gottmann 1961) or by scholars in other disciplines.

Résumé of contemporary social geography

In general, the empirical record would seem to characterize social geography as a multifaceted perspective on the spatial organization of mankind. The implication is that some important sources of areal differentiation emanate from society, thus reversing the premises of anthropogeography and other deterministic explanations of social differentiation. Analysis of this social dimension in human geography has involved two basic approaches: the examination of the formal distributions of social phenomena as indices of areal differentiation and the interpretation of these distributions in the light of their underlying social processes. A recent development, particularly in northwestern Europe, is the involvement of social geographers in interdisciplinary research and regional planning.

Nevertheless, the social dimension is one of the least studied aspects of human geography. Social geography lacks definite boundaries and has neither a central unifying concept nor even an agreed content. Instead, there are scattered individual efforts to analyze the changing social patterns of the modern world. Generalizations regarding the nature and potential function of the field, therefore, can only be proffered as suggestions, based on the substantive research directions and ideas of contemporary experts in the field and on the current trends and technical possibilities in other social science disciplines.

The future of social geography

The challenge

Social geography faces a set of challenges that are unprecedented. Revolutionary changes in world social patterns have rendered past analytical techniques obsolete, while philosophical and cultural currents within modern social life tend to increase the propensity to change of both reality itself and its social-psychological significance. Thus, while technological, economic, and commercial evolution tends to produce a certain degree of standardization in society’s spatial order, there is a universal tendency to emphasize social, that is, ethnic, religious, or linguistic, differentiation. The philosophical problems of intersubjectivity and coexistence are ubiquitously discussed. “The home of contemporary man,” wrote Plattel ([I960] 1965, pp. 1-2) “does not lie primarily in a localized environment, but in his fellow-man.” The traditional methods and objectives of social science are being fundamentally challenged. Analysis must some how be broadened so as to arrive at a more holistic vision of social reality: the classical Cartesian premises underlying accepted research method ology led to the discovery of systems, but mechanics and structures of systems constitute only a partial view of reality. Today both subjective (internal) aspects of reality and objective (external) aspects of reality must be analyzed. Modern psychology and sociology have endeavored to meet this challenge by forging new analytical techniques, and many other social science disciplines have adopted a decidedly behavioristic approach in recent years.

In the light of these developments, the spatial patterns of world society assume a new significance; the immediate challenge for social geographers would seem to be the collaboration with other scholars in the monumental task of describing world society within its geographical setting. For such an endeavor, social geography needs a unifying theme, a conceptual framework that will enable it to contribute toward and benefit from the research efforts of scholars in related social science disciplines. Such a unified framework seems to be emerging from the work of some contemporary social geographers. Some of its characteristics are described below.

Social space as central theme

Claval’s critique of contemporary social geography concludes that “to understand the geography of a place means to understand the social organization of those who inhabit it, their mentality, their beliefs, their ‘representations’” (1964, p. 123). Watson’s study of Hamilton demonstrates how “The spatial pattern is, in the last analysis, a reflection of the moral order” ([1951] 1965, p. 476). In this article I have postulated that the raison d’etre of social geography rests on the fact that the social order is distinct from (even if closely interrelated with) the other orders of human activity in space. In order to describe adequately this social dimension or order, contemporary thought would seem to demand the use of both internal and external perspectives. Is this possible?

Sociologists, for example, Chombart de Lauwe (1956) and Gaston Bardet (1951), and human ecologists, for example, Firey (1960), have demonstrated the technical possibility of exploring a society’s perception of its geographical milieu. Geographers, for example, Rochefort (1961), Burton and Kates (1964), and Pataki (1965), have also shown that space has different meanings for different societies, and thus distance and spatial movement can no longer be considered in traditional geodesic terms but must be considered in terms of those dimensions perceived by their human occupants. For example, groups of Italians, Poles, Pakistanis, and Negroes may live side by side in one section of a city. Yet each group, because of economic, historical, cultural, or other reasons, may possess an entirely different conception of space. Some groups may have a social horizon that scarcely transcends the block in which they live or the set of stores in which they work or shop, while others may have social contacts with relatives thousands of miles away. Whether contact with distant relatives is frequent or rare does not influence the fact that a bond is perceived which ignores the barriers of space and time. The social geography of urban neighborhoods cannot ignore these differential attitudes toward space.

This illustration, which challenges traditional notions of space, may lead to the impression that only the social-psychological conception of space matters. Rochefort (1963), in discussing this problem, strongly emphasized that the real dimensions of geographic space must always be kept in mind. Therefore, the central conceptual problem in social geography is to define space in such a way that both subjective and objective dimensions are included.

Sorre’s response (1957) to this challenge was the concept of social space: the synthesis of real and perceived dimensions of space. The subjective component of social space in his view is embodied in the distribution of fundamental social groups, while the objective component consists of their concrete geographic setting.

Bobek’s concept of social landscape already expressed the main idea that a unit of social space is a region or place in which one or several groups live and have a common set of ideas of their environment (1943; 1948). The fundamental merit of this concept, as a central theme for social geography, is that it incorporates the traditional elements of groups and environment, while redefining them in terms which are relevant to the examination of modern society. Let us see how the methodology of contemporary social geography could be organized around such a central theme.

Subjective componentsocial groups. Sociology has shown how the dimensions and meaning of space are colored by the beliefs and group affiliations of its human occupants. Sociologists speak of ethnic space, religious space, and other spaces, and social morphology maps the distribution of groups on the premise that their formal spatial configurations imply the values held by the group (Halbwachs 1938). Social geography must go further: these groups, the subjective component of social space, must be studied not only as morphological patterns on the earth but also as formative influences in molding a society’s perception of its environment. The relevant groups include those which determine or condition the spatial distribution and interaction of people, for example, language and ethnic groups; those which influence a society’s use of space, for example, religious and kin groups; and, most significantly, those which develop as a result of society’s mode of material subsistence, namely, the genres de vie or livelihood groups. The bonds and values engendered by participation in these groups are not directly observable on the earth’s surface, but they are essential to the understanding of the spatial movements and distribution of people on the earth. Classical French geography used such formal categories of relevant groups, but profound transformations in social structure have occurred since the analytical framework of Brunhes, or even Sorre, was first formulated. Even though the choice of relevant grouping will demand close cooperation with sociologists and others, the social geographer does not have to abandon entirely the analytical techniques of his predecessors. Rather, such traditional con cepts should be re-examined in the light of the new analytical possibilities which appear in many social science disciplines. One example which might merit re-examination, for example, is the Vidalian notion of genre de vie. Settlements forms, land use, social interaction, and even political integration have been explained by geographers in terms of genres de vie. Many feel that the concept has lost its applicability to modern social life (George 1951; Le Lannou 1949), but others argue that it can be reformulated (Sorre 1948; Varagnac 1948). By discounting the various modifications which have accrued through the years and by re-examining the original notion in the light of contemporary developments both in world society and in social science, guidelines for a reformulation may become apparent. A genre de vie, in Vidal’s opinion (1911), implies more than a means of material subsistence; its geographical significance stems in large part from its spiritual component, the structures mentales which persist even after the external modalities of livelihood change. The important point is that both material and spiritual elements are harmoniously integrated in the genre de vie community within a particular milieu. Such a conception closely resembles the notion of “com munity” in rural sociology (Hillery 1950).

Without changing the concept at all, there are some applications in the modern world. Witness the adaptation problems of immigrants from rural to urban areas, the psychological problems involved in the retraining of unemployed miners, the social consequences of colonialism and economic restructuring within the “third world.” In the urban industrial world, however, livelihood is a less compelling basis for community consciousness than other similarities, for example, a common racial, professional, or linguistic background or similar consumption habits (Fourastié 1963). But whatever the source, if a recognizable consistency in a group’s perception and consequent use of its environment are associated with a common structure mentale, why not consider this pattern as a genre de vie, for example, that of travel agents, of salesmen, of truck drivers, of commuting students, of social scientists? Chombart de Lauwe (Chombart de Lauwe et al. 1952, p. 243) showed how a deep social rift could prevail in a small dormitory village because the inhabitants belonged to two different genres de vie. The same could be said of immigrant ethnic groups in some urban centers (Taeuber & Taeuber 1965). Ideally, within either an urban or a rural region, one could thus identify the component genres de vie and see if there is a hierarchy of importance among them, the dominant one giving a character to the place, as in pilgrimage, market, or university towns. Many other possibilities exist, but much more substantive work, preferably in conjunction with other disciplines, is needed before any formal categories of modern genres de vie can be made. Until this is done, the existing formal groupings of language, religion, race, etc., may serve to constitute the subjective component of social space; however, if these sociological categories can somehow be integrated into the more geographical concept of genre de vie, the result would be an ideal subjective ingredient for social geography.

Objective componentthe social environment. The term “social environment” is used here to connote all the socially significant aspects of the total geographic milieu. Traditionally, geographers have tended to exaggerate the distinction between the natural (physical-biotic) environment and the artificial network of human establishments created by society. This dualistic conception tends to ignore the fact that mankind’s environment-creating ap paratus has by no means entirely destroyed the natural framework and that the interplay of natural and artificial assumes very different forms throughout the world. The social environment, as objective component of social space, includes more than these two levels. It includes, for example, the relation of social attitudes and traditions to nature, resource use, and the ethics of group relations.

Social geographers are far from a satisfactory definition of the social environment; they lack substantive studies which would provide the raw material for such a definition. What is the social significance, for example, of purely physical elements, such as humidity, temperature, or altitude? Geographers have added very little to the “findings” of the Huntingtonian environmentalists. Yet the behavioral sciences are interested in knowing the connections, real or perceived, between society and its natural environment. The research challenges proposed in Sorre’s Géographie psychologique (1954) remain virtually untouched. In addition, little is known about the “synthetic environment” (Herber 1962): the various consequences of atmospheric and oceanic pollution, or the consumption of medicated foods, stimulants, and sedatives. What are the physiological and pathological consequences of changes in the environment, for example, housing, communication, and diet?

Recently some geographers have viewed the environment as an amalgam of systems (Wagner & Mikesell 1962; Ackerman 1963; van Paassen 1965). This approach is satisfactory from the theoretical and technical points of view, but does it admit of nonsystematic (dysfunctional) elements which often play such an important part in social life? The social geographer must be sensitive to the local exceptions which give special char acter to individual places, such as Rochefort (1961) demonstrated in her study of Sicilian social environment.

Approaches to the study of social space

We have seen that the study of social groups within their territorial (environmental) framework has constituted the basic traditional methodology of the Dutch, British, and some French social geographers. In theory, this has involved a combination of a morphological approach (mapping of social groups) and an ecological approach (relations of groups to environment). Today, however, the latter (vertical) dimension is perhaps less significant than the horizontal one, namely, the spatial patterns of interaction between social groups, such as Lowenthal’s Caribbean study illustrated (Lowen-thal 1961). A psychological approach to group attitudes, such as one finds in the Revue de psychologic des peuples, may provide clues to the origins of some spatial discontinuities in social interaction.

In terms of the notions of group and environment, as redefined above, let us see what analytical methods can be used in the study of social space. Two of the many possible approaches are (1) to consider social space as a mosaic of social areas defined in terms of the occupant groups, for example, genres de vie or ethnic groups; and (2) to view social space as nodally organized, that is, as a network of spatial relations radiating around certain centers (Sorre’s points privilégiés) and permeated by the arteries of circulation.

Formal approachsocial areas. Initially, the formal approach examines the spatial patterns and characteristics of social groups in virtually the same fashion as that used by the disciples of Stein-metz in Amsterdam. On the basis of these distributions a series of regions, homogeneous in terms of individual characteristics, can be compared and associations can be sought. Such associations, however, must then be examined in terms of the social environment in which these social characteristics occur, that is, an ecological approach must supplement the more formal “sociographical” stage of analysis. In addition to these two steps, the geographer must endeavor to see how all these elements combine to form the social whole within a particu larregion and must seek explanations for the vari ations through space in the incidence and functional character of these social wholes, Jones sees “social regions” within the city of Belfast (1960), for example, as a product of historical and religious forces, while the “social area analysis” tradition in American human ecology (Theodorson 1961) has demonstrated the use of various other indices in the establishment of intraurban social regions.

Functional approach. A more dynamic and increasingly popular approach is to consider social space in terms of its nodal organization. The orbit of group activities and the related horizons of social consciousness can be examined (cartographically) in terms of their use of these nodes, for example, markets, cinemas, and schools (Chombart de Lauwe et al. 1952). The hinterland of each of these nodes varies in scale and significance, and these variations provide crucial insights into the social character of particular places. The study of nodal regions and of circulation are two examples of a functional approach to the study of social space.

Sorre (1961) suggests that settlement units— towns, cities, metropolises—provide a primary set of nodes on a world scale. Within each of these nodes is an internal system of centers (schools, churches, cinemas) whose social significance can also be examined cartographically. Here again the social geographer can collaborate with and utilize some of the existing principles of central place theorists and perhaps somewhat qualify definitions of centrality currently based on commercial and industrial criteria. Edgar Kant’s Umland studies (Kant et al. 1951), J. Labasse’s circulation studies (1955), and Pierre George’s studies of the urban fringe (1962) provide orientation for this kind of study. As world society becomes more urbanized, social geographers will concern themselves more with the urban field, and, here, collaboration with other scholars will be imperative.

The essential clue to the internal dynamism of social space can be found in its circulation system. Circulation here includes all kinds of movement of goods, services, people, and ideas—any kind of spatial movement which occasions social communication. As the Paris study (Chombart de Lauwe et al. 1952) demonstrated, the actual and potential use of a circulation system indicates the concrete social horizons of the group it serves; changes within it may indicate or produce changes in the relation between groups and between a group and its social environment.

A vast number of research questions emanate from this dimension of social space, particularly now that the processes of social differentiation and cultural standardization are so closely tied with large-scale, mass-produced goods and services. Interregional traffic, the currents of the tourist world, pilgrimages, daily and seasonal commuting —these are only a few samples of the many activities the student of circulation could investigate.

To summarize, social geography can be defined as the study of the areal (spatial) patterns and functional relations of social groups in the context of their social environment; the internal structure and external relations of the nodes of social activity; and the articulation of various channels of social communication.

Although the discussion has distinguished between various elements and approaches to social geography, it must be emphasized that one of the fundamental characteristics of the field has been, and must remain, its integral, holistic character. Like social history, it must endeavor to maintain the holistic view, that is, to show how the individual parts and their functional connections integrate to give a specific character to the social whole. French geographers have supplied ample precedent for this kind of holism; so, indeed, have the social anthropologists of the Anglo-American world. The more the field becomes theoretically systematized, the greater will the challenge of integration become.

For social geography to fulfill its potential, the various approaches to the field need to be coordinated into a systematic conceptual framework. Sorre’s concept of social space could provide a central theme for such a framework. Its ingredient elements could be considered as bases for systematic subdivisions, for example, geography of language, of religions, and of diet, each of which contributes a valuable perspective on society’s spatial order. The definition given here seems to incorporate the various elements which have belonged to the field of social geography and which, given the trends in contemporary social science, could constitute fruitful future directions for the discipline.

Anne Buttimer

BIBLIOGRAPHY

Ackerman, Edward A. 1963 Where Is a Research Frontier? Association of American Geographers, Annals 53:429–440.

Bardet, Gaston 1951 Social Topography: An Analytico-Synthetic Understanding of the Urban Texture. Town Planning Review 22:237–260.

Bobek, Hans 1943 Der Orient als Soziallandschaft. Unpublished manuscript.

Bobek, Hans 1948 Stellung und Bedeutung der Sozial-geographie. Erdkunde 2, no. 1/3:118–125.

Bobek, Hans 1959 Die Hauptstufen der Gesellschafts-und Wirtschaftsentfaltung in geographischer Sicht. Die Erde: Zeitschrift der Gesellschaft für Erdkunde zu Berlin 90:259–298.

Bobek, Hans 1961 Sozialgeographie: Neue Wege der Kultur- und Bevölkerungsgeographie. Deutsche Gesell schaft für Bevölkerungswissenschaft, Mitteilungen 3:62–67.

Broek, Jan O. M. 1932 The Santa Clara Valley, California: A Study in Landscape Changes. Utrecht (Netherlands): Oosthoek.

Broek, Jan O. M. 1944 Diversity and Unity in South east Asia. Geographical Review 34:175–195.

Brookfield, H. C. 1961 The Highland Peoples of New Guinea: A Study of Distribution and Localization. Geographical Journal 127:436–448.

Brunhes, Jean (1910) 1924 Human Geography. Lon don: Harrap. → First published in French. A fourth French edition was published in 1934 by Alcan.

Burton, Ian; and Kates, Robert W. 1964 The Perception of Natural Hazards in Resource Management. Natural Resources Journal 3:412–441.

Chatelain, M. Abel 1947 Les fondements d’une geographie sociale de la bourgeoisie francaise. Annales de géographie 56:455–462.

Chatelain, M. Abel 1953 Horizons de la géographie sociologique. Revue de géographie de Lyon 28:225–228.

Chombart de Lauwe, Paul H. 1956 La vie quotidienne des families ouvriéres. Paris: Centre National de la Recherche Scientiflque.

Chombart de Lauwe, Paul H. et al. 1952 Paris et I’agglomération parisienne. Volume 1: L’espace social dans une grande cité. Paris: Presses Universitaires de France.

Claval, Paul 1964 Essai sur I’évolution de la géographie humaine. Cahiers de geographie de Besançon No. 12. Paris: Les Belles-Lettres. → Translation of extract in text provided by Anne Buttimer.

Dickinson, Robert E. (1947) 1956 City Region and Regionalism: A Geographical Contribution to Human Ecology. London: Routledge.

Durkheim, Émile (1893) 1960 The Division of Labor in Society. Glencoe, III.: Free Press. → First published as De la division du travail social.

Durkheim, Emile (1895) 1958 The Rules of Sociological Method. 8th ed. Edited by George E. G. Catlin. Glencoe, III.: Free Press. → First published as Les règies de la méthode sociologique.

Evans, Emyr E. 1942 Irish Heritage: The Landscape, the People and Their Work. Dundalk (Ireland): Tempest.

Evans, Emyr E. 1957 Irish Folk Ways. New York: Devin-Adair.

Febvre, Lucien (1922) 1925 A Geographical Introduction to History. New York: Knopf. → First published as La terre et l“évolution humaine.

Firey, Walter I. 1960 Man, Mind and Land: A Theory of Resource Use. Glencoe, III.: Free Press.

Fleure, Herbert J. 1918 Human Geography in Western Europe: A Study in Appreciation. London: Williams & Norgate.

Fleure, Herbert J. 1947 Some Problems of Society and Environment. London: Institute of British Geographers.

Forde, Daryll (1934) 1963 Habitat, Economy and Society: A Geographical Introduction to Ethnology. 5th ed. London: Methuen.

FourastiÉ, Jean 1963 Le grand espoir du XXesiècle. Paris: Gallimard.

Friedmann, Georges (editor) 1953 Villes et campagnes: Civilisation urbaine et civilisation rurale en France. Paris: Colin.

Geddes, Patrick (1915) 1950 Cities in Evolution. New ed., rev. Oxford Univ. Press.

Geipel, Robert 1952 Soziale Struktur und Einheits-bewusstein als Grundlagen geographischer Gliederung. Rhein-Mainische Forschungen, No. 38. Frankfurt ahi Main (Germany): Kramer.

George, Pierre (1946) 1956 Géographie sociale du monde. 4th ed. Paris: Presses Universitaires de France.

George, Pierre 1951 Introduction à I’étude géographique de la population du monde. France, Institut National d’Études Demographiques, Travaux et Documents, Cahier No. 14. Paris: Presses Universitaires de France.

George, Pierre 1956 La campagne: Le fait rural à travers le monde. Paris: Presses Universitaires de France.

George, Pierre 1962 Pre’cis de géographie urbaine. Paris: Presses Universitaires de France.

Godlund, Sven 1956 The Function and Growth of Bus Traffic Within the Sphere of Urban Influence. Lund Studies in Geography Series B, No. 18.

Gottmann, Jean (1961) 1964 Megalopolis: The Urbanized Northeastern Seaboard of the United States. Cam bridge, Mass.: M.I.T. Press.

Gourou, Pierre 1964 Changes in Civilisation and Their Influence on Landscapes. Impact of Science on Society 14:57–71.

Groenman, S. joerd (1950) 1953 Methoden der sociografie: Een inleiding tot de practijk van het sociale onderzoek in Nederland. Assen (Netherlands): Van Gorcum.

HÄgerstrand, Torsten 1952 The Propagation of Innovation Waves. Lund Studies in Geography Series B, No. 4.

Hahn, Eduard 1896 Die Haustiere und ihre Beziehungen zur Wirtschaft des Mensches. Leipzig: Duncker & Humblot.

Halbwachs, Maurice (1938) 1960 Population and Society: Introduction to Social Morphology. Glencoe, III.: Free Press. → First published as Morphologie sociale.

Hartke, Wolfgang 1956 Die Hütekinder im Hohen Vogelsberg: Der geographische Charakter eines Sozial-problems. Munchner Geographische Hefte, No. 11. Regensburg (Germany): Lassleben.

Hartke, Wolfgang 1959 Gedanken über die Bestim-mung von Räumen gleichen sozialgeographischen Verhaltens. Erdkunde 13:426–436.

Hartke, Wolfgang 1961 Die sozialgeographische Differenzierung der Gemarkungen ländlicher Kleinstädte. Geografiska annaler 43:105–113.

Hartke, Wolfgang 1963 Der Weg zur Sozialgeographie: Der wissenschaftliche Lebensweg von Professor Dr. Hans Bobek. Österreichische Geographische Gesell schaft, Mitteilungen 105:5–22.

Hartmann, F. 1958 Volkach am Main: Kulturgeographische Studien über eine unterfrankische Kleinstadt. Unpublished manuscript.

Hartshorne, Richard (1939) 1964 The Nature of Geography: A Critical Survey of Current Thought in the Light of the Past. Lancaster, Pa.: Association of American Geographers.

Hartshorne, Richard 1950 The Functional Approach in Political Geography. Association of American Geog raphers, Annals 40:95–130.

Herber, Lewis 1962 Our Synthetic Environment. New York: Knopf.

Hillery, George A. Jr. 1950 Definitions of Community: Areas of Agreement. Rural Sociology 20:111–123.

International Symposium on Man’s Role in Changing the Face of the Earth, Princeton, N.J., 1955 1956 Man’s Role in Changing the Face of the Earth. Edited by William L. Thomas et al. Univ. of Chicago Press.

Isaac, Erich 1959 Influence of Religion on the Spread of Citrus. Science 129:179–186.

Jones, Emrys A. 1960 A Social Geography of Belfast. Oxford Univ. Press.

Juillard, Étienne 1953 La vie rurale dans la plaine de Basse-Alsace: Essai de géographic sociale. Strasbourg, Université, Institut des Hautes Études Alsaciennes, Vol. 9. Strasbourg: Le Roux.

Kant, Edgar et al. 1951 Studies in Rural-Urban Interaction. Lund Studies in Geography Series B, No. 4.

Kroeber, A. L.; and Kluckhohn, Clyde 1952 Culture: A Critical Review of Concepts and Definitions. Harvard University, Peabody Museum of American Archaeology and Ethnology Papers, Vol. 47, No. 1. Cambridge, Mass.: The Museum. → A paperback edition was published in 1963 by Vintage Books.

Labasse, Jean 1955 Les capitaux et la région, étude géographique: Essai sur le commerce et la circulation des capitaux dans la région lyonnaise. Paris: Colin.

Le Lannou, Maurice 1949 La géographie humaine. Paris: Flammarion.

Le Play, FrÉdÉric (1855) 1877-1879 Les ouvriers européens, 2d ed. 6 vols. Tours (France): Mame.

Le Play, FrÉdÉric (1871) 1907 L’organisation de la famille: Selon le vrai modèle signalé par I’histoire de toutes les races et de tous les temps. 5th ed. Tours (France): Mame.

Lewis, Peirce F. 1965 Impact of Negro Migration on the Electoral Geography of Flint, Michigan, 1932-1962: A Cartographic Analysis. Association of American Geographers, Annals 55:1–25.

Lowenthal, David 1961 The West Indies Federation: Perspectives on a New Nation. American Geographical Society, Research Series, No. 23. New York: Columbia Univ. Press.

Mckenzie, R. D. 1931 Ecology, Human. Volume 5, pages 314-315 in Encyclopaedia of the Social Sciences. New York: Macmillan.

Mckenzie, R. D. 1934 The Field and Problems of De mography, Human Geography and Human Ecology. Pages 52-66 in Luther L. Bernard (editor), The Fields and Methods of Sociology. New York: Farrar.

Mackinder, Halford (1919) 1942 Democratic Ideals and Reality: A Study in the Politics of Reconstruction. London: Constable; New York: Holt.

Morrill, Richard L. 1963 The Development of Spatial Distributions of Towns in Sweden: A Historical-Pre dictive Approach. Association of American Geogra phers, Annals 53:1–14.

Paassen, C. van 1965 Over vormverandering in de sociale geografie. Grbningen (Netherlands): Wolters.

Park, Robert E.; and Burgess, Ernest W. (1921) 1929 Introduction to the Science of Sociology. 2d ed. Univ. of Chicago Press.

Pataki, K. J. 1965 Shifting Population and Environment Among the Auyuna: Some Considerations on Phenomena and Schema. MA. thesis, Univ. of Wash ington.

Pelletier, Jean 1959 Alger; 1955: Essai d’une géo graphic sociale. Cahiers de géographie de Besançon, No. 6. Paris: Belles Lettres.

Phlipponneau, Michel 1960 Géographie et action: In troduction à la géographie appliquée. Paris: Colin.

Planhol, Xavier de 1957 Le monde islamique: Essai de géographie religieuse. Paris: Presses Universitaires de France.

Platt, Robert S. 1961 The Saarland; an International Borderland: Social Geography From Field Study of Nine Border Villages. Erdkunde 15:54–68.

Plattel, Martin G. (1960) 1965 Social Philosophy. Duquesne Studies, Philosophical Series, No. 18. Pitts burgh, Pa.: Duquesne Univ. Press. → First published in Dutch.

Ratzel, Friedrich 1876 Stddte und Culturbilder aus Nordamerika. Leipzig: Brockhaus.

Ratzel, Friedrich (1882–1891) 1921-1922 Anthropogeographie. 2 vols. Stuttgart (Germany): Engelhorn. → Volume 1: Grundziige der Anwendung der Erd kunde auf die Geschichte, 4th ed. Volume 2: Die geo-graphische Verbreitung des Menschen, 3d ed.

Ratzel, Friedrich (1897) 1923 Politische Geographie. 3d ed. Edited by Eugen Oberhummer. Munich and Berlin: Oldenbourg.

Ratzel, Friedrich (1901) 1902 Man as a Life Phe nomenon on the Earth. Volume 1, pages 61-106 in Hans F. Helmolt (editor), The History of the World: A Survey of Man’s Record. New York: Dodd. → First published in German.

Redfield, Robert 1953 The Primitive World and Its Transformations. Ithaca, N.Y.: Cornell Univ. Press. → A paperback edition was published in 1957.

Redfield, Robert 1955 The Little Community: View points for the Study of a Human Whole. Univ. of Chicago Press. → A paperback edition, bound together with Peasant Society and Culture, was published in 1961 by Cambridge Univ. Press.

Redfield, Robert; and Villa Rojas, Alfonso (1934) 1962 Chan Kom: A Maya Village. Carnegie Institution of Washington, Publication No. 448. Univ. of Chicago Press.

Recherches sociographiques. → Published since 1960 by Laval University, Faculty of Social Sciences, Quebec.

Revue de psychologie des peuples. → Published since 1946 by the Institute Havrais de Sociologie Économique et de Psychologie des Peuples, Le Havre, France.

Rochefort, RenÉe 1961 Le travail en Sicile: Étude de géographie sociale. Paris: Presses Universitaires de France.

Rochefort, RenÉe 1963 Géographie sociale et sciences humaines. Association de Géographes Francais, Bul letin No. 314-315:18–32.

Ruppert, Karl 1958 Spalt: Ein methodischer Beitrag zum Studium der Agrarlandschaft mit Hilfe der kleinräumlichen Nutzflächen und Sozialkartierung und zur Geographie des Hopfenbaus. Münchner Geographische Hefte, No. 14. Regensburg (Germany): Lassleben.

Sauer, C. O. 1931 Geography, Cultural. Volume 6, pages 621-624 in Encyclopaedia of the Social Sciences. New York: Macmillan.

SchÄffle, Allen (1875–1878) 1896 Bau und Leben des sozialen Körpers. 2d ed. 4 vols. Tübingen (Ger many): Laupp.

Schnore, Leo F. 1961 Geography and Human Ecology. Economic Geography 37:207–217.

Sorre, Maximilien 1943-1953 Les fondements de la géographie humaine. 3 vols. Paris: Colin.

Sorre, Maximilien 1948 La notion de genre de vie et sa valeur actuelle. Annales de géographie 57:97-108, 193–204.

Sorre, Maximilien 1954 La géographie psychologique: L’adaptation au milieu climatique et biosocial. Traité de psychologie appliquée, Vol. 3, part 3. Paris: Presses Universitaires de France.

Sorre, Maximilien 1957 Rencontres de la géographie et de la sociologie. Paris: Rivière.

Sorre, Maximilien 1961 L’homme sur la terre. Paris: Hachette.

Steward, Julian H. 1950 Area Research: Theory and Practice. Bulletin No. 63. New York: Social Science Research Council.

Taeuber, K. E.; and Taeuber, A. F. 1965 Negroes in Cities: Residential Segregation and Neighborhood Change. Chicago: Aldine.

Theodorson, G. A. (editor) 1961 Studies in Human Ecology. Evanston, III.: Row, Peterson.

Thomas, Franklin 1925 The Environmental Basis of Society: A Study in the History of Sociological Theory. New York: Century.

Tijdschrift vcor economische en sociale geografie. → published since 1910 by the Nederlandse Vereniging voor Economische en Sociale Geografie.

Varagnac, Andre 1948 Civilisation traditionnelle et genres de vie. Paris: Michel.

Vidal de la Blache, Paul 1896 Le principe de la géographie generate. Annales de géographie 5:129–142.

Vidal de la Blache, Paul 1902 Les conditions géographiques des faits sociaux. Annales de géographie 11:13–23.

Vidal de la Blache, Paul 1911 Les genres de vie dans la géographie humaine. Annales de geographie 20: 193-212, 289–304.

Vidal de la Blache, Paul 1917-1918 Les grandes agglomérations humaines. Annales de géographie 26: 401-422; 27:92-101, 174–187.

Vooys, Adriaan C. de 1950 De ontwikkelung van de sociale geografie in Nederland. Inaugural lecture at the University of Utrecht. Groningen (Netherlands): No publisher given.

Wagner, Philip 1960 The Human Use of the Earth. Glencoe, III.: Free Press.

Wagner, Philip L.; and Mikesell, Marvin W. (editors) 1962 Readings in Cultural Geography. Univ. of Chi cago Press.

Watson, James W. (1951) 1965 The Sociological As pects of Geography. Pages 463-499 in Thomas Griffith Taylor (editor), Geography in the Twentieth Century: A Study of Growth, Fields, Techniques, Aims and Trends. New York: Philosophical Library.

VI STATISTICAL GEOGRAPHY

Statistical geography is to geography what econometrics is to economics, sociometrics to sociology, psychometrics to psychology, or even jurimetrics to jurisprudence—an approach to the field rather than a subdivision of that field. Like these other approaches, it is of recent development—the mani festation within geography of the trend to a more quantitative approach that has characterized all the social sciences since the end of World War II. As is also the case with these other approaches, many of the pioneering contributions to statistical geography have come from workers in other fields. The parallel term “geometries” is not used to describe the work of the statistical geographer, how ever, not so much because it provides an inappro priate picture of the statistical geographer’s attempts to identify and measure regularities observable in spatial distributions as because the branch of mathematics that originated in Greek attempts to measure the earth has a two-thousand-year priority in its right to the name. In addition, mathematical geography, concerned as it is with map projections, is the branch of geography that today relies most heavily and directly upon geometry.

The syndrome that characterizes statistical geography today includes a more formal theoretical orientation than was true of geography in the past, a reliance upon statistical inference and numerical analysis in empirical research, the use of mathematical programming and simulation procedures in applied research, a basic concern with model construction, and an involvement with high-speed computers, mass-data banks, and automated map ping devices.

Yet this assemblage of interests is not mono lithic. There are differences in emphasis, corresponding to each of the four main traditions of geographic research: spatial, area studies, man-land, and earth science (Pattison 1964; National Research Council 1965). Statistical geographyorig inated within the spatial tradition, with its em phasis upon analysis of spatial distributions and associations, and initially relied heavily upon exer cises in distribution fitting and upon regression and correlation analysis. It spread to area studies when the spatial tradition began using multivariate analysis, then to the man-land tradition via be havioral studies of environmental perception and individual decision making, and to the earth-science tradition as spatial studies and related systematic sciences began using systems analysis. To understand these several facets of statistical geography, with their differences in use of theory, choices, and timing of applications of statistical methods, and their contacts with the rest of science, therefore requires some understanding of the four research traditions. These, in turn, can best be viewed within a formal overview of approaches to regional analysis (Berry 1964a).

Approaches to regional analysis

Virtually any sort of regional analysis may be considered as starting from information about one or more places. For each place, one or more prop erties, or characteristics, may be measured, and the measurements may be made at one or more points in time. It is often convenient to think in terms of a three-dimensional array X of order v × p × t. Cell xijk of this array (or matrix) re cords the value of variable i at place j in time k. Rows of the array—vectors of array values for fixed i, k—record the distribution of variables over places at some point in time. Column vectors inventory the properties of places at time t. Time vectors report on the t states of variable i at place j.

Operational specification of variables, places, and times varies with the interests of the particular analyst, but a matrix such as X is (albeit usually implicitly) the object of all forms of geographic study. A row vector of X is a spatial distribution that can be mapped. A column vector is a locational inventory. Such row and column vectors are the bases of systematic (topical) and regional geography, respectively. Time vectors report changes in spatial distributions and at locations and are basic to historical geography.

The data. Cell xijk may contain a variety of different records, depending upon specification of i, j, and k. Columns, for example, may be denned as places with area, such as countries, states, counties, census tracts, or quarter-mile-square cells of a half-mile grid. Alternatively, they may be di-mensionless points; three such examples are tri-angulation points used for measuring altitude, weather stations, and soil-sample-core locations (Kao 1963). Rows may be defined as properties of the places (scalar quantities, such as population residing in each of a set of census tracts or altitude at each of a set of triangulation points), or they may refer to connections between places (vector quantities, such as flow of coal from southern Illinois to each of the Midwestern counties). Rows also might be airline traffic flow between cities; in this case columns would represent pairs of cities. In addition, any level of measurement may appear, whether nominal, ordinal, interval, or ratio [seeStatistics, Descriptive, article onlocation and dispersion, for a discussion of levels of meas urement].

Spatial tradition

Geography’s spatial tradition is founded upon cartographic portrayal and subsequent study of spatial distributions, thus having as its base row-wise analysis. A few examples of the kinds of spatial distributions studied follow:

(1) A land-use map—scalar, nominal, of areas, an areal distribution.

(2) A map showing the location of major cities by dots—scalar, nominal, at points, a point distribution.

(3) An airline-route map—vector, nominal, joining points.

(4) A map of soil quality—scalar, ordinal, of areas.

(5) A map showing average annual temperature —scalar, interval. If the map shows averages by states or districts, it is also discontinuous and areal (a choropleth map, if different shadings are ap plied to the units), but if it shows the temperature varying over the country as a surface, it is a con tinuous generalization. The generalization is of point observations (usually isopleth, since the sur face will be depicted by contours) if, for example, weather-station data are used, or of areal data if interpolations were made, for example, with re spect to the district averages treated as points central to each of the districts.

(6) A highway map, with routes classified by quality—vector, ordinal, joining points.

(7) A map of city-to-city air-passenger move ments—vector, ratio, joining points.

Two dominant themes emerge in spatial analysis of such maps: evaluation of the pattern of scalar distributions and of similarities in pattern over a number of such distributions and evaluation of the connectivity evidenced by vector distributions and of similarities in the connectivity of several such distributions. Apparently, the funda mental properties of pattern include absolute location (position), relative location (geometry), and scale, with a family of interesting derived prop erties, including density and density gradients, spacing, directional orientation, and the like. Simi larly, accessibility is central to the study of con nectivity, and from it are derived such properties as centrality itself, relative dominance, degree of interdependence, etc. The two themes merge in spatial-systems analysis, where pattern and connectivity are examined in their association. For example, urban land values decline with increasing distance from the city center, and type of farming varies with distance from market. Such examples are readily generalized to the dynamic, that is, time-dependent, case.

Area-studies (chorographic) tradition

Just as row-wise analysis is the basis of the spatial tradition, so columnwise analysis provides the base for geography’s area-studies tradition. The essential problems of this tradition are those of regional intelligence: the characterization of place in terms of the associations between characteristics localized in that place. This approach is often restricted to those features of place that are directly observable as landscape but, especially among the French school of human geographers, is also extended to an evaluation of both tangible and intangible aspects of “regional character” and of the differentiation of places (the study of “areal differentiation”) [seeLandscape].

An appetite for information, a penchant for the peculiar, emphasis upon field work, attachment to the people and language of a particular part of the world, a strong literary bent, a companionship with history and great reliance upon historical modes of explanation (in contrast to the functional, de terministic, and probabilistic modes of the other geographic traditions)—all these serve to identify the work of the student in the area-studies tradition, whatever the areas examined: countries or continents, regions or culture areas.

Man-land tradition

In the tradition of medieval philosophy, classical geographers distinguished between two major sets of variables: the physical (inorganic plus biotic) and the cultural. The as sociated methodological argument was that these provided the bases of the two major segments of the field: physical geography and cultural geography. Finer groupings of variables within these categories led to the variety of systematic branches of the subject, such as the geographies of landforms, of plants, of industry, of cities, or of language (the many topical fields and subfields might thus be identified as nested subsets of rows of the matrix X ).

The classical modes of thought, however, also led to a particular tradition in geography, the man-land tradition, in which the relationships between physical variables and human characteristics and activities were examined. In combination with the social Darwinism of the late nineteenth century, simple one-way studies were made of the effects of environment on man. These were later complemented by studies of the effects of man on environment, from which emerged much of the original thinking in the field of conservation. More recently the ancient dichotomy has been relaxed with, for example, studies of the effects of environmental perception on resource evaluation and decision making in resources management (Kates 1962) and with the adoption of a systems-analysis frame (Ackerman 1963).

Earth-science tradition

During the eighteenth century, geography was an integral and substantial part of natural philosophy. At that time geographical study embraced all aspects of the earth, air, and waters. Since then, however, most aspects of these studies have branched off as separate systematic sciences, and geography’s earth-science tradition has been left with such concerns as the study of landforms and their evolution (geomorphology), descriptive climatology, and certain as pects of the geographies of soils, plants, and animals, together with the attempt to achieve some spatial synthesis of these in order to identify “natu ral environmental complexes.” It is to this latter end that systems-analysis procedures have recently been employed in this research tradition (Chorley 1962).

Antecedents and stimuli

Prior to World War ii few papers of a statistical nature had been published by geographers. Perhaps the only contributions worthy of note were Matui’s (1932) fitting of the Poisson distribution to quadrat counts of settlements in a portion of Japan and Wright’s (1937) discussion of Lorenz measures of concentration in the spatial case. However, two general antecedents can be distin guished, in addition to the pioneering contributions of workers in other fields: centrography and social physics. From the former came the idea of developing a special family of descriptive statistics for spatial distributions, and from the latter the recognition of certain classes of regularities in such distributions.

Centrography

During the early part of the twentieth century there was a lively debate among statisticians concerning such measures as the center of population. Part of the debate stemmed from publication by the U.S. Bureau of the Census of a piece entitled Center of Population and Median Lines … (U.S. Bureau of the Census 1923). Many articles were published, notably in the Jour nal of the American Statistical Association and in Metron, concerning the relative advantages of al ternative centers, the center of gravity, the spatial median, and the center of minimum aggregate travel. (The U.S. Census Bureau’s geography branch still reports on the center of gravity of the United States’ population after each census.) This debate gradually subsided in the United States, but in the Soviet Union centrography flourished during the 1920s and 1930s. A centrographic laboratory was founded at Leningrad, under the auspices of the Russian Geographic Society, in 1925, and its director, E. E. Sviatlovsky, pursued studies of the “actual” and “proper” centers and the distributions of all manner of phenomena. However, the set of “proper” centers of economic activities prepared for the Gosplan of 1929 was at odds with the second Five-Year Plan, and as a result, the laboratory was finally disbanded. Porter (1963) provides a fairly complete bibliography of the relevant lit erature on centrography. Recently, the Israeli stat istician Bachi (1963) has attempted to revive cen trography, with the development of a variety of measures of dispersion and association of spatial distributions. Current interest within geography is slight, however, except as embraced by the social physicists (Stewart & Warntz 1958).

Social physics

The attempt to describe human phenomena in terms of physical laws has a long history in every social science [seeRank-size relations].

In geographical studies this has been expressed in two major ways: (1) by use of “gravity models” to describe spatial interaction; and (2) by use of “potential models” as general summaries of interdependency between all places in large areas. Such models are said by their advocates to summarize a wide variety of social and economic distributions in economically advanced societies. Gravity models were first used in a relatively formal way by E. G. Ravenstein, in his seminal study “The Laws of Migration” (1885; 1889). Thereafter, these models found wide application, for example, in marketing geography—Reilly’s “law of retail gravitation” (1931)—and in urban transportation studies describing interzonal travel. Carrothers (1956) reviews this work and the basic postulate that interactions or movements between places are proportional to the product of the masses and inversely proportional to some exponent of distance: that is, Iij α Mi. Such gravity ana logs were generalized by the astronomer J. Q. Stewart (1947) to the case of the potential surface, which simultaneously describes the interactions of each place and every other. Thus, the potential at any point i is given by

The surface is interpolated from such measures for a sample set of points. There is still considerable interest within geography in social physics (Stewart & Warntz 1958), and new applications are continually being developed. Mackay (1959), for example, used gravity models to translate the depressing effects upon telephone communications of the French-English language boundary in Canada and of the United States-Canadian political boundary into their physical-distance equivalents, thus showing how social space can be transformed into the metric of physical space. Similar applications are to be found in all branches of cultural geography today.

Pioneering contributions from elsewhere

Workers in other fields provided several significant examples of the application of statistical methods to geographic problems, identified the major statisti cal problems of regional analysis, and prepared the first text on statistical geography, thereby doing much to set the pace and tone of statistical geography today. A statistician, M. G. Kendall (1939), for example, showed how principal-components analysis could be used to develop a multivariate index that would portray the geographical distribution of crop productivity in England. M. D. Hagood (1943), an agricultural statistician, used multiplefactor analysis to define multivariable uniform regions. An economist, C. Clark (1951), showed that the negative exponential distribution fitted population density patterns within cities. G. K. Zipf (1949), a philologist, developed the rank-size distribution of cities. A mathematical social scientist, H. A. Simon (1955), showed the bases of this distribution in simple stochastic processes, and sociologists found that a repetitive three-factor structure characterized the social geography of cities (Berry 1964fo; 1965). G. U. Yule and M. G. Kendall ([1911] 1939) identified the problem of modifiable units: if data are of areas rather than at points, results of any analysis will be in part de pendent upon the nature of the areal units of observation utilized. W. S. Robinson (1950) pro vided the relevant relationship between individual and ecological (areal, set-type) correlations, subsequently extended by Goodman (1959) in the context of ecological regression. A second problem, that of contiguity, or spatial autocorrelation, has been examined by Moran (1948) in the nominal case and by Geary (1954) more generally. Geary also applied his measures of contiguity to evaluating lack of independence of residuals from regression in studies of spatial association. These studies are reviewed in Statistical Geography (Duncan et al. 1961), the first general book in the field, written by sociologists. Further examination of autocorrelation in spatial series is to be found in “Spatial Variation” (Matern 1960), a major con tribution to areal sampling by a mathematician. Spatial analysis remains basic to quantitative plant ecology and to epidemiology, and from these fields have come many of the ideas used today in the study of pattern in point distributions and of spatial diffusion processes.

Statistical geography—1950 to 1965

Centrography, social physics, and external stim uli, facilitated by developments in computer technology that for the first time enabled the mass data of geographic problems to be handled conveniently, combined to stimulate workers in geography’s spa tial tradition to work quantitatively. The older forms of cartographic analysis provided firm bases for this development, and many of the early studies were simply quantitative extensions of analyses cartographically conceived and executed. Arthur H. Robinson (1962), a cartographer, for example, utilized correlation and regression analysis to im prove the ways by which he could map spatial associations. McCarty and his associates (1956) used similar procedures to replace older carto graphic means of comparison. Thomas (1960) showed the various ways in which residuals from regression could be treated cartographically so as to draw upon traditional geographical means of map analysis in model reformulation and refine ment. King (1962) applied the “nearest-neighbor” methods of the quantitative plant ecologist to the study of pattern in point distributions, with, like the 1932 Matui study, expectations derived from the Poisson distribution. These represent but a few examples of the spatial studies concerned with distribution fitting as a means of studying spatial pattern or with uses of correlation and regression in studies of spatial association. Many other exam ples are to be found in uses of regression to fit gravity models and obtain the distance exponents for different phenomena (Carrothers 1956) or to fit negative exponential distributions to urban population densities and the like (Berry 1965).

These kinds of studies represent the beginnings, from which statistical geography has grown rap idly. Dacey and Tung (1962) have made major advances in point-pattern analysis, for example, by transforming the distribution-fitting exercise into an explicit hypothesis-testing frame, with relevant expectations derived from settlement the ory. Curry (1964) views many urban phenomena as the outcome of known-probability mechanisms. The Swedish geographer Hägerstrand (1953) was the first to show that many spatial patterns might be considered as the outcome of diffusion processes that could be simulated, using Monte Carlo methods, and his work led to a burst of similar simulation studies in the United States (Morrill 1963).

New approaches to spatial analysis have also been developed. Most of the examples outlined above use scalar data. Garrison (1960) showed that the mathematical theory of graphs provided an excellent base from which to examine vector distributions, and Nystuen and Dacey (1961) extended his argument to the case of organizational regions, using graph-theoretic measures of accessi bility of places to communications networks to define relatively independent subsets of relatively interdependent places. Tobler (1963) showed how a generalization of map projections, traditionally studied by the mathematical geographer, could be used as the basis for mapping social, economic, cultural, or political space into physical space, as a further means for merging geographical applications of various statistical and mathematical meth ods with the more traditional means of geographical analysis. Finally, in addition to developments of the descriptive kind, statistical geography has extended its work to embrace investigations of a prescriptive nature. Garrison and Morrill (I960), for example, applied the techniques of spatial priceequilibrium analysis to determining what should be the patterns of interregional trade in wheat and flour in the United States. Other research workers are now much concerned with the procedures of spatial programming. Haggett (1965) has provided an excellent review of the substance of the first decade of quantitative work in the spatial tradition.

With the use of multivariate analysis, statistical geography has spread from the spatial tradition to that of area studies. A traditional geographic problem in this latter tradition is that of regionalization —the attempt to derive areas relatively uniform in terms of a complex of associated characteristics and also relatively different from other areas in terms of that complex. Such problems, involving mass-data analysis, were traditionally handled by overlaying maps. This earlier procedure has been replaced, however, by the use of the modern computer, applying such multivariate procedures as factor analysis to reduce many variables to a few factors representing “complexes” of associated characteristics, and the application of numerical taxonomy to get optimal classification (minimizing within-group variance) of observations into regions on the basis of the distances between observations in the factor space (Berry & Ray 1966). Output from the entire procedure of data analysis and reduction includes the complexes of characteristics that define “regional character,” measures of the similarity of the observations, and the regions [seeClustering].

Statistical work also characterizes the man-land tradition, largely by virtue of either simple correlation and regression studies that include physical variables, on one side, and cultural variables, on the other (for example, correlations of annual precipitation and population densities in the high plains), or through uses of probability theory. It is the latter, indeterministic type of study that repre sents new departures. Curry (1962a), for example, shows how livestock management in the intensive grassland-farming areas of New Zealand is related to probabilities of fodder availability, which in turn are derived from probabilities of requisite climatic conditions. Much of the basic research goes into establishment of the relevant probabilities, in this case, of the probabilities of repetitive events that play a central role in farm management. Kates (1962), on the other hand, examined relations of management of flood-plain property to flood haz ards, rare events. He found management practices to be conditioned, not by reasonably precise evaluations of the situation, as in the case of the New Zealand farmers, but by a widely varying set of preconceptions, at variance with the actual probability mechanisms.

Work in the earth-science tradition of geography has, also, become statistical and ranges from the attempt to reformulate the geography of landforms generated by fluvial mechanisms in the framework of general-systems theory (Chorley 1962) through studies of climatic change as a random series (Curry 1962b) to the analysis of precipitation climatology using harmonic methods (Sabbagh & Bryson 1962) or to the development of linear mod els predictive of some characteristic through priormultivariate analysis, so as to satisfy the assumptions of the model ultimately to be produced (Wong 1963). There is today perhaps more work of a statistical kind in the earth-science tradition than in either the area-studies or man-land tradition.

Statistical geography—analysis of both the statistical and the mathematical kind—is to be found in all branches of geography today. However, in the methods utilized, certain differences between geog raphy’s four main research traditions are to be noted. In the spatial tradition, distribution fitting, correlation and regression analysis, uses of such methods as the mathematical theory of graphs, and prescriptive uses of spatial programming dominate, along with uses of probability mechanisms to study diffusion processes. The area-studies tradition re lies upon multivariate analysis, particularly factor analysis, and upon numerical taxonomy, to facili tate mass-data analysis. In the man-land tradition, a neat contrast is to be noted between those of traditional deterministic outlook, who use regression methods, and those concerned with decision making in resources management, who focus upon probabilities of the a priori and a posteriori kinds. Finally, in the earth-science tradition those pro cedures that facilitate systems analysis have been those most rapidly adopted and used. At the end of World War II geography was nonquantitative. Statistical geography has played an integral, even critical, part in the transformation of geography into a modern social science in the postwar years.

Brian J. L. Berry

[See alsoCartography; Central Place; Regional Science. Other relevant material may be found inClustering; Factor analysis; Multivariate Analysis.]

BIBLIOGRAPHY

Ackerman, Edward A. 1963 Where Is a Research Frontier? Association of American Geographers, Annals 53:429–440.

Bachi, Roberto 1963 Standard Distance Measures and Related Methods for Spatial Analysis. Regional Science Association, Papers 10:83–132.

Berry, Brian J. L. 1964a Approaches to Regional Anal ysis: A Synthesis. Association of American Geog raphers, Annals 54:2–11.

Berry, Brian J. L. 1964b Cities as Systems Within Systems of Cities. Pages 116-137 in John Friedmann and William Alonso (editors), Regional Development and Planning: A Reader. Cambridge, Mass.: M.I.T. Press.

Berry, Brian J. L. 1965 Research Frontiers in Urban Geography. Pages 403-430 in Philip M. Hauser and L. F. Schnore (editors), The Study of Urbanization. New York: Wiley.

Berry, Brian J. L.; and Ray, Michael 1966 Multivari ate Socio-economic Regionalization: A Pilot Study in Central Canada. Unpublished manuscript.

Carrothers, Gerald A. P. 1956 An Historical Review of the Gravity and Potential Concepts of Human Inter action. Journal of the American Institute of Planners 22:94–102.

Chorley, Richard J. 1962 Geomorphology and General Systems Theory. U.S. Geological Survey, Professional Paper 500B.

Chorley, Richard J. 1963 Geography and Analogue Theory. Association of American Geographers, Annals 54:127–137.

Clark, Colin 1951 Urban Population Densities. Journal of the Royal Statistical Society Series A 114:490–496.

Curry, Leslie 1962a The Climatic Resources of Intensive Grassland Farming: The Waikato, New Zealand. Geographical Review 52:174–194.

Curry, Leslie 1962b Climatic Change as a Random Series. Association of American Geographers, Annals 52:21–31.

Curry, Leslie 1964 The Random Spatial Economy: An Exploration in Settlement Theory. Association of American Geographers, Annals 54:138–146.

Dacey, Michael F.; and Tung, tse-hsiung 1962 The Identification of Randomness in Point Patterns. Jour nal of Regional Science 4:83–96.

Duncan, Otis Dudley; Cuzzort, Ray P.; and Duncan, Beverly 1961 Statistical Geography: Problems in Analyzing Areal Data. New York: Free Press.

Garrison, William L. 1960 Connectivity of the Interstate Highway System. Regional Science Association, Papers 6:121–137.

Garrison, William L.; and Morrill, Richard L. 1960 Projections of Interregional Patterns of Trade in Wheat and Flour. Economic Geography 36:116–126.

Geary, R. C. 1954 The Contiguity Ratio and Statistical Mapping. Incorporated Statistician 5:115–145. → In cludes four pages of discussion.

Goodman, Leo A. 1959 Some Alternatives to Ecological Correlation. American Journal of Sociology 64:610–625.

HÄgerstrand, Torsten 1953 Innovationsförloppet ur horologisk synpunkt. Lund (Sweden): Gleerupska Universitetsbokhandeln.

Haggett, Peter (1965) 1966 Locational Analysis in Human Geography. New York: St. Martins..

Hagood, Margaret D. 1943 Statistical Methods for De lineation of Regions Applied to Data on Agriculture and Population. Social Forces 21:287–297.

Kao, Richard C. 1963 The Use of Computers in the Processing and Analysis of Geographic Information. Geographical Review 53:530–547.

Kates, Robert W. 1962 Hazard and Choice Perception in Flood Plain Management. Department of Geography Research Paper No. 78. Univ. of Chicago Press.

Kendall, M. G. 1939 The Geographical Distribution of Crop Productivity in England. Journal of the Royal Statistical Society Series A 102:21–62.

King, Leslie J. 1962 A Quantitative Expression of Pat terns of Urban Settlements in Selected Areas of the United States. Tijdschrift voor economische en sociale geografie 53:1–7.

McCarty, Harold H.; Hook, J. C; and Knos, D. S. 1956 The Measurement of Association in Industrial Geog raphy. Iowa City: State Univ. of Iowa, Department of Geography.

Mackay, J. Ross 1959 The Interactance Hypothesis and Political Boundaries in Canada: A Preliminary Study. Canadian Geographer 11:1–8.

MatÉrn, Bertil 1960 Spatial Variation: Stochastic Models and Their Applications to Some Problems in Forest Surveys and Other Sampling Investigations. Sweden, Statens Skogsforskningsinstitut, Meddelanden 49, no. 5.

Matui, Isamu 1932 Statistical Study of the Distribution of Scattered Villages in Two Regions of the Tonami Plain, Toyama Prefecture. Japanese Journal of Geol ogy and Geography 9:251–256.

Moran, P. A. P. 1948 The Interpretation of Statistical Maps. Journal of the Royal Statistical Society Series B 10:245–251.

Morrill, Richard L. 1963 The Development of Spatial Distributions of Towns in Sweden: A Historical-Pre dictive Approach. Association of American Geog raphers, Annals 53:1–14.

Moser, Claus A.; and Scott, Wolf 1961 British Towns: A Statistical Study of Their Economic and Social Differences. Edinburgh: Oliver.

National Research Council, Ad Hoc Committee on Geography 1965 The Science of Geography: Re port. National Research Council Publication No. 1277. Washington: National Academy of Sciences-National Research Council.

Nystuen, John D.; and Dacey, Michael F. 1961 A Graph Theory Interpretation of Nodal Regions. Regional Science Association, Papers 7:29–42.

Pattison, William D. 1964 The Four Traditions of Geography. Journal of Geography 63:211–216.

Porter, P. W. 1963 What Is the Point of Minimum Aggregate Travel? Association of American Geog raphers, Annals 53:224–232.

Ravenstein, E. G. 1885 The Laws of Migration. Jour nal of the Royal Statistical Society Series A 48:167–235. → Includes seven pages of discussion.

Ravenstein, E. G. 1889 The Laws of Migration: Second Paper. Journal of the Royal Statistical Society Series A 52:241–305. → Includes three pages of discussion.

Reilly, William J. 1931 The Law of Retail Gravitation. New York: Putnam.

Robinson, Arthur H. 1962 Mapping the Correspondence of Isarithmic Maps. Association of American Geographers, Annals 52:414–429.

Robinson, W. S. 1950 Ecological Correlations and the Behavior of Individuals. American Sociological Review 15:351–357.

Sabbagh, Michael A.; and Bryson, Reid A. 1962 As pects of the Precipitation Climatology of Canada In vestigated by the Method of Harmonic Analysis. Asso ciation of American Geographers, Annals 52:426–440.

Simon, Herbert A. 1955 On a Class of Skew Distribution Functions. Biometrika 42:425–440.

Stewart, John Q. 1947 Empirical Mathematical Rules Concerning the Distribution and Equilibrium of Popu lation. Geographical Review 37:461–485.

Stewart, John Q.; and Warntz, William 1958 Physics of Population Distribution. Journal of Regional Sci ence 1:99–123.

Thomas, Edwin N. 1960 Maps of Residuals From Re gression: Their Characteristics and Uses in Geographic Research. Iowa City: State Univ. of Iowa, Department of Geography.

Tobler, Waldo R. 1963 Geographical Area and Map Projections. Geographical Review 53:59–78.

U.S. Bureau of the Census 1923 Fourteenth Census of the United States; 1920: Center of Population and Median Lines and Centers of Area, Agriculture, Manu factures, and Cotton. Washington: Government Printing Office.

Wong, Shue Tuck 1963 A Multivariate Statistical Model for Predicting Mean Annual Flood in New England. Association of American Geographers, Annals 53:298–311.

Wright, John K. 1937 Some Measures of Distributions. Association of American Geographers, Annals 27: 177–211.

Yule, G. Udny; and Kendall, M. G. (1911) 1958 An Introduction to the Theory of Statistics. 14th ed., rev. & enl. London: Griffin. → M. G. Kendall has been joint author since the eleventh edition (1937) and revised the 1958 edition. See especially pages 310-325 on “Correlation and Regression: Some Practical Problems.”

Zipf, George K. 1949 Human Behavior and the Princi ple of Least Effort: An Introduction to Human Ecology. Reading, Mass.: Addison-Wesley.

Geography

views updated May 29 2018

GEOGRAPHY.

Only relatively recently accepted as a subject of study by universities, geography has been characterized as a Cinderella among the disciplines. It was not one of the traditional liberal arts, and it appeared in its modern form in the curriculum of universities in the nineteenth and twentieth centuries; it still remains a small component, and is sometimes not present at all, in institutions of higher learning. Part of the reason for this is that society, and even geographers themselves, are not sure of the nature of geography. Geographers are only rarely members of national academies of science, or of the humanities, falling between the stools with the social or so-called soft sciences.

The Nature of Geography

In his seminal studies on the methodology of the subject, Richard Hartshorne (18991992) proposed the following definition: "Geography is concerned to provide an accurate, orderly, and rational description of the variable character of the earth's surface" (Hartshorne, p. 21). Understandably this characterization has not been universally accepted, and others have suggested terms such as "areal differentiation," and "spatial interaction" as better expressing the core of geography. It has been seen as more akin to history than to the systematic sciences (physics, chemistry, biology, geology, meteorology, etc.) in that it has no body of material peculiar to itself, but rather adopts a point of view. But subjects studied by some geographers, such as map projections, are highly "scientific."

In France the alliance between geography and history"geohistory"extends from Jean Bodin to Montesquieu to Jules Michelet to the Annales school, especially Lucien Febvre, A Geographical Introduction to History, and Fernand Braudel and their followers. In Germany geography was an auxiliary science in the encyclopedia of history, or Historik, as taught in the universities from the eighteenth century; and there are parallels in other national traditions.

If geography is Cinderella, its Prince Charming is cartography and, by extension, remote sensing of the environment. Maps and related images of the Earth have a wide appeal to collectors and others and are used professionally in several disciplines. But preeminently, they are the tools of geographers so that their study is often confused with the reality of the Earth itself, as expressed in the old tag "Geography is about maps."

Maps may help in an understanding of the "reality" of geography, but are not "reality" themselves, consisting, as they do, of conventional symbols. Humankind, since prehistoric times, has been concerned with the local environment, as evidenced in maps made before the written record. The subject came into focus in the later classical period as exemplified by the Geography of Strabo (63 b.c.e.c. 24 c.e.), a verbal description of the then-known world, and the similarly titled Geographia of Ptolemy (second century c.e.), containing instructions for map-making, of essentially the same area of Eurasia and North Africa described earlier by Strabo. The Greeks from the time of Plato (427348 or 247 b.c.e.) appear to have accepted the idea of the Earth as a perfect sphere, which, apparently, was not a part of early Babylonian, Egyptian, or Chinese cosmography. Although Buddhism spread from India to China and Japan (after 400 b.c.e.), and following the establishment of Buddhism there, priests returned to India to seek their religious roots and wrote about their travels, this geographical lore did not enter the mainstream of thought in translation until comparatively recent times. The same is largely true of Islam following the death of Muhammad (570632 c.e.), in spite of close contact between this religion and Christianity in the Mediterranean and elsewhere over many centuries. Thus the travels of "Sinbad the Sailor" and more scientific geographies were available only in translation as relatively late additions to European literature and in this sense are considered to be "nonhistorical" in the West. Even the accounts of Marco Polo (12541324) of his travels from Venice to Cathay (China) and return were at first disbelieved.

This article need not go into detail concerning the remarkably accurate measurement of the circumference of the Earth by Eratosthenes (third century b.c.e.), or its rejection by others (including Ptolemy), until the later Renaissance and the scientific revolution in Europe, of the fifteenth to seventeenth centuries. At that time Ptolemy's Geographia was "re-discovered" and translated from Greek into Latin and formed the basis of much of the study of geography in this era. It was in turn criticized, improved upon, and superseded during the period of European ascendancy in science and global discovery when half the coasts of the world were "discovered" and charted. The dichotomy represented by the conceptions of the GreeksStrabo on the one hand and Ptolemy on the othercontinued into the Enlightenment period through the writings of, for example, Bernhard Varen (Varenius, 16221650) in regional geography, or chorography, and in the ideas of Edmond Halley (16561742), who, in addition to his work in astronomy, laid the foundations of physical, thematic mapping, with representations of winds, tides, and Earth magnetism with isogones (lines of equal magnetic variation) delineated on published charts. More than a century later, the polymath Alexander von Humboldt (17691859), well trained in the natural and physical sciences, attempted to give unity to geography, while still considering the Earth in relation to the cosmos (Kosmos is the title of his greatest work). It was Humboldt's contemporary Carl Ritter (17791859) who, similarly, emphasized the unity of the field, but with a person-centered (even teleological) approach to human/land relationships, following Immanuel Kant (17241804) and others. But the division between physical and human geography continued and increased in the nineteenth and in the first half of the twentieth century in France, Britain, the United States, and areas influenced by these countries. That this is still the case is evidenced by recent multiauthored volumes titled, respectively, Horizons in Physical Geography (1987) and Human Geography: An Essential Anthology (1996). Accordingly, it is necessary to recognize recent trends in these major, separate divisions of geography; this article will later cite attempts at reconciliation between these two disparate streams, and others.

Geographical Determinism

A concept that retarded the acceptance of geography as a serious academic endeavor until quite recently was geographical determinism. Although stemming from earlier work by the German geographer and ethnographer Friedrich Ratzel (18441904), with adherents in other European countries, the high priestess of this cult in the United States was Ellen C. Semple (18631932); another American espouser of "determinism" was Ellsworth Huntington (18761947). In its extreme expression the theory asserts that the work of humans is controlled or "determined" by geographical conditions: climate, landforms, and the like. This idea was opposed by the scientifically trained English scholar Eva G. R. Taylor (18791966) and others in Britain, France, and elsewhere. The debate continued throughout the twentieth century, but has few adherents in the early 2000s. An alternative to determinism was proposed, namely possibilism, which suggests that humans have a number of possibilities from which to select. Possibilism apparently owed its origin to the French geographer Paul Vidal de la Blache (18451918), who, with his followers, never accepted the concept of determinism. At this time most of the world, including North America, was influenced by European ideas so that traditional, indigenous geographies became subsumed under colonial and other European ideology. Thus India, under British rule, became one of the best studied and surveyed areas in the world. China, Japan, and Korea resisted this cultural hegemony, but eventually accepted it.

Military and Public Geography

If geography has had a mixed reception in research universities, its ideas and practitioners have been embraced by both the military and the public sectors. Thus Napoléon Bonaparte (17691821) not only developed strategy based on knowledge of geography but also sponsored a translation of Strabo's Geography. Following the Napoleonic Wars there was a great interest in geographical exploration worldwide, especially the interiors of the continents (little known at the time), and geographical societies were founded in major cities. Furthermore, instruction in geography has been part of the training in service academies ever since. There is, understandably, an increased interest in geographical intelligence during times of war in all of the military servicesnavy, army, and air forceswhich engage in so-called defense studies and mapping. Thus during World War II the British Royal Navy, Naval Intelligence Division, commissioned a series of handbooks on the geography of various areas that were later declassified and made available to general libraries. This was also true of maps made by the U.S. Army Map Service and of charts, the work of various hydrographic services together with coastal studies in the form of navigational pilot books. The role of air forces is well-known in not only providing the means for aerial reconnaissance but also in sponsoring aeronautical chart series at "geographical" scales. Thus map coverage of the Earth on the scale of 1:1,000,000, begun through the efforts of Albrecht Penck (18581945) as the International Map of the World (IMW), was completed at this scale by the maps of the U.S. Aeronautical Chart and Information Service during World War II. Ironically the United States had not officially cooperated with the IMW, but the private American Geographical Society of New York mapped all of Latin America at this scale. Furthermore, as a result of wartime experience, many returning veterans in several countries during this period made careers in applied or theoretical geography, some founding or working in geography departments, which were established in many colleges and universities in the 1940s and 1950s. These personnel are now mostly retired, or deceased, and later wars did not produce a similar great expansion in academic geography. In fact some of the then newly created departments, especially in the United States, were merged with other instructional units, renamed, or terminated. This is attributable to a number of causes, not least the abandonment of the geography department at Harvard University during the presidency of the scientist James Conant (18931978). The Harvard precedent was followed by other, even public, institutions that formerly had strong departments of geography. These universities often have splendid map collections, which find little use among students and faculty not geographically "literate."

Just as geography is essential to the military establishment, so it is valued in the public, civilian sector. Thus, that most fundamental of human geographical distributions, population itself, is of the greatest interest to census bureaus of various countries and internationally, with the United Nations having a vital concern with demography. Similarly, topographic and land use data of various scales is essential to the effective administration of urban and rural areas in the form of maps and reports. Thomas Jefferson (17431826), who also sponsored geographical exploration and wrote an important geographical treatise, well understood this, and he initiated the United States Public Land Survey, passed into law as the Land Ordinance of 1785 and first applied in Ohio. Subsequently recti-linear surveys expanded over three-quarters of the United States, which became the public domain, thereby transforming the American landscape and producing a torrent of cadastral maps, plat books, and county atlases in the nineteenth century, and beyond. Also, between World War I and World War II, the detailed Land Utilisation Survey of Britain was conducted under the direction of L. Dudley Stamp (18891966), and it has had a profound effect on the economic life of that area (despite its title, Scotland was not covered). The idea was to make a record of existing land uses and to plan for the future. By 1940 the survey was essentially complete and proved of enormous value to Britain as it expanded its agricultural production during World War II. The concept was adopted by other, especially densely settled, countries and gave rise to the formation of the Commission of the International Geographical Union (IGU) on land use.

The remarkable expansion and improvement of highways of all kinds during this period led to the production of road data, published by government highway departments, automobile clubs, or oil and tire companies in many countries, becoming perhaps the most commonly available geographical source material worldwide. To a lesser or greater degree all government departments from foreign offices to small municipalities require geographical data, and more personnel are needed to process this information than are trained in existing educational institutions. However, the great success of geography in these applied fields has not been matched by similar success in theoretical realms in recent years, which will be the subject of most of the remainder of this essay.

Geographical Theories

Both Kant in Königsberg and Isaac Newton (16421727) at Cambridge University in England taught what might be called geography today, but they are not remembered for that activity. Newton also postulated that the Earth is an oblate (polar flattened) spheroid before it was proved by geophysical methods. This and other findings were to be of practical use in the development of detailed topographic maps in the nineteenth and twentieth centuries and especially during the space age in the second half of the twentieth century and the first years of the twenty-first century. Among the greatest contributions to science have been understanding of the shape, size, and motions of the planet Earth and of its place in the Universe. Although other societies and cultures such as the Chinese and Indian probably recognized the "curved" surface of the Earth, a full realization of the figure, mass, and movements of this planet is essentially a triumph of Western thought. This has become almost universally accepted so that Eurocentrism, as well as Sino-, Indo-, and other "centrisms" are dead, or dying.

This article has stressed the duality of the subject between physical and human, and theoretical and applied aspects, and needs now to detail a further division, that between systematic and regional geography. Some scholars will take a physical entity, such as vegetation or soils, or a cultural feature, such as urbanization or transportation, and discuss it with little or no reference to other topics. Contrasting with this is regional geography, in which the worker attempts to characterize an assemblage of features such as landforms, rivers, roads, soils, human population, and settlements to demonstrate how they are related, or "interact." Of course, the choice of what factors are most significant in a given area is of critical importance. Some assert that it is easier to analyze than to synthesize, and that regional geography is "the highest form of the geographer's craft."

The Limits in Geography

What are the limits of the focus in the study of geography? It is usually assumed that geography is concerned with the surface or "shell" of the Earth, but workers do not specify how deep or high this sphere of interest to geographers extends. With some prescience Hartshorne wrote before 1966, "Man has for the first time projected his world of action beyond the [Earth's] atmosphere and may soon be expected to extend that range to the moon" (Hartshorne, p. 24). This prediction was shortly realized when the United States, through its Apollo 11 mission of July 1969, landed two astronauts on the Earth's natural satellite, and images were taken of the Earth from the Moon. Further, rocks from the lunar surface were collected for the study of which the term geology was employed. Subsequently, many artificial satellites, both with humans aboard and unmanned, have been launched so that we now have a much greater understanding of the "blue planet" Earth, from above. Considering that the first aerial photographs of parts of the Earth were taken from a balloon in 1858, and that the science of photogrammetrymaking maps from overlapping vertical air photographswas developed in the first half of the twentieth century, progress has been remarkable. This was largely accomplished in European and North American countries, which assisted other areas that benefited from this technology.

Another realm of the Earth that has been seriously investigated is the ocean depths, made possible by modern technologysonar or echo sounding. The oceans, the greater part of the surface of the Earth, have, of course, long interested maritime nations but until recently this interest had been confined to the surface, coastal limits, and other shallow areasor to speculation. The most remarkable example of the latter is the postulation by the German meteorologist and physicist Alfred Wegener (18801930), who proposed that at an earlier period the continent(s) consisted of a single or, at most, two major land masses that had subsequently drifted apart. At the time of his death, not enough evidence was forthcoming to prove Wegener's theory of "continental displacement," or "continental drift" as it was later termed. Sonic sounding now permits continuous traces, or profiles, to be made across the ocean floor by ships in progress, which, in aggregate, provide a true three-dimensional image, making possible the charting of the ocean basins. This process has revealed an assemblage of "forms" as varied as those on land areas above sea level, including profound depths greater than the highest mountains on Earth. Most important, it has given validity to Wegener's theory, through identification of mid-ocean ridges, from which apparently the continents spread for the most part laterally. Other forms of evidence support this fundamental, and now widely accepted but hitherto controversial, theory.

The two examples given above illustrate how the Earth's surface or limits have been vastly enlarged in the past half century, and they also suggest that geographers who are concerned with the Earth as the "home" of humankind must come to terms with this increased realm. Geography, however, remains a very divided subject searching for a core. As indicated above, a few women geographers in the past have made signal contributions to the subject. However, as in many other studies, often women geographers became editors, school teachers, and librarians, in spite of the tradition of the intrepid Victorian lady traveler. Until quite recently women were often "excused" from field work in geography departments, considered a necessary part of the curriculum for males. Now that they constitute about one-half of the enrollment in colleges and universities, and owing to changing mores, women are now making an impact on the subject at the research level. This is expected to continue and expand, since previously, half of the human in human (and physical) geography had been excluded. The closeness of the female to Mother Nature, it is speculated, gives women an advantage in geography that is now being realized, understood, and, to a greater extent than previously, appreciated.

Just as women and their special points of view have not been an important part of geography in the past, so, it is argued, have the interests and aspirations of the proletariat not been included. An attempt to address this lacuna has been through what has been called in its extreme form (and comparable to feminist geography) Marxist or, more acceptably, socialist geography. Notable proponents of feminist geography are Cindi Katz and Janice Monk; and of Marxist and Socialist geography, Massimo Quaino and David Harvey. Marxist and socialist practitioners assert, with considerable justification, that geography has been a white male, Eurocentric (even imperialistic) study, with the protection of the "establishment" in mind. This conservative view is being challenged by new graduates of "red-brick" universities, particularly, recalling their working-class roots. This concern also extends to various, often ethnic, minorities who were not previously part of the geographical equation. How far this development will redress past inequities and affect the future direction of the discipline may depend on further recruitment of university students from the underclasses and those concerned with their welfare.

The wide range of topics investigated by geographers will not be elaborated here, but it is suggested by the indexes of textbooks in physical and human geography where they are later treated in varying degrees of detail. Methods mentioned earlier, including field work, interpretation of aerial photographs, and individual images from space are of course still available, but other methods have been added recently: continuous surveillance and imaging of the Earth, since 1972, through the Earth Resources Technology Satellite (ERTS) and its successor, Landsat; and new computer programs, especially ARC/INFO, introduced in 1982.

Geographical Models

Before treating computer graphics and geographical information systems (GIS), which purists consider only tools, mention should be made of the criticism of academic geography as being mere description, with a lack of theories. Traditionalists would argue that all places on the Earth are different, that therefore description of these variations represents reality, and that geographers need only address themselves to the "real" world. These practitioners often reject models, two of the most successful of which in geography have been Central Place Theory in human geography, and the Köppen system of world climates in physical geography, which are examined, as examples, below.

Central Place Theory arose from studies of the distribution of settlements in the eighteenth century in Germany where it was postulated that places of varying size and function on a "uniform plain" would be arranged in hexagonal hierarchies. This has proved to be essentially the case, for example, on the delta of the Nile in Egypt, and elsewhere, as demonstrated by cartography and remote sensing. An elaboration of this is in the location of functional areas within cities, as exemplified by the case of Chicago. However, critics have observed that both concentric and sector patterns are evident in Chicago and that what obtains in a relatively modern city in the American Midwest does not necessarily apply to European and especially Asian cities. Chicago, with its location on the shores of one of the Great Lakes, is unique because of its special geographical setting and other factors.

More successful has been the Köppen system of climatic classification, which also had its origins in Europe, through the work of Wladimir Köppen (18461940) in the nineteenth and twentieth centuries. When long-term information was available from weather stations worldwide, it was observed that patterns were repeated in separate, but expected locations around the globe. Thus, to take one example, between approximately 30° and 40° latitude on the west coasts of all the continents, a Mediterranean climate was recognized. This climatic type is characterized by having high-sun (summer) drought and almost all of its precipitation in the low-sun (winter) months, the opposite of what would be expected. Thus in addition to the type example in Europe and the Middle East, it was found that California in the northern hemisphere, and Central Chile, Southwest Africa, and Southwest Australia in the southern, have "Mediterranean" climates. On the basis of climatic similarities, scholars have attempted to refine the Köppen system so that six major, and a total of sixteen, sub-types are now recognized globally. Being based to some extent on "native" vegetation, and not altogether on climate, critics consider the Köppen system to be a technical grouping rather than a true classification. Nevertheless, the Köppen system has proved to be a powerful teaching device that has not been superseded by other classifications. Although largely similar, each of the geographically separated Mediterranean areas have differences owing to local factors but, in comparison with other areas, they are more alike than different. This is also true of the other subtypes of the Köppen climatic classification. Critics assert that all places on Earth have some, if only minor, differences and thus cannot be classified.

The two examples of theoretical geography given above, with their origins in the past, have been fine-tuned and have become part of the curricula of geography departments, more than most other theories in a field that has been characterized as more empirical than theoretical. Since the mid-1970s, two developments have revolutionized geographythe computer and space explorationas much as any previous advances.

Computers and Space Exploration

Before the computer, proto-quantitative geography was developed with the aid of various calculating machines, but it was only after Herman Hollerith (18601929) combined punched cards with the then-recent electromagnetic inventions that it became possible to count and classify in a much shorter time and with greater detail and precision than by any previous methods. But through the 1950s the machines required to perform these operations remained expensive, large, and clumsy. A turning point occurred in 1982 with the introduction of ARC/INFO, a geographical information software package that combines traditional automated systems with advanced spatial data-based handling capability. This is accomplished by combining a series of layers each with a different theme: relief, roads, political boundaries, settlements, and so onthe desideratum of the regional geographer. Specifically, ARC/INFO uses both vector (line) and raster (tabular) storage; transformations can be undertaken and questions asked concerning numbers, distance, addresses, and so forth. The utility of such a system to those who are concerned with geographical distributions is enormous, as is time saved by these procedures. Maps can be made using the system and simulated, three-dimensional representations produced. It can also be animated to show, for example, population change through time. The machines on which these procedures can be accomplished have been incredibly reduced in size, price, and availability.

Equally as remarkable as the widespread utility of the computer to geography have been the space programs of various countries and consortia. As in the case of the computer, space technology did not arrive fully developed without a period of gestation, partly alluded to above in the references to aerial photography. A breakthrough similar to that of the computer was made when German rocket scientists joined the incipient United States space program, and that of the Soviet Union, following World War II. Prior to this, around 1910, the Germans had used rockets fitted with cameras to image small areas of the Earth. The range of these missiles was increased in World War II when, as Vergeltungswaffe 2 (V-2) rockets, they were used for military purposes. From 1960, the Television and Infra-Red Observation Satellite (TIROS), a series of unmanned satellites, was launched in the United States, and demonstrated the ability to gather weather data from above Earth's cloud layers, the first important use of the new technology. Meanwhile the Soviets launched the Synchronously Programmed User Terminal and Network Interface (Sputnik) in 1957, imaged the previously unseen side of the Moon in 1959, and put a human in space in 1961. The next year marked the first manned space flight by the United States, which soon began a series of missions imaging the Earth from spaceGemini (19651966), and Apollo (19681969), with hand-held cameras loaded with color and, later, color infrared (CIR) film, which had been perfected during World War II. As mentioned above, it was the Apollo 11 mission in 1969 that landed humans on the Moon. Subsequently, nonphotographic systems were also used so that the term "Remote Sensing of the Environment" was coined to replace air photo interpretation, which was included in the definition.

The next development was continuous, extensive surveillance of the planet Earth, first accomplished by the Earth Resources Technology Satellite (ERTS) in 1972. Another similar satellite was launched in 1975, and the program was renamed Land Remote Sensing Satellite (Landsat). Since that time, the surface of the Earth (except the polar regions that the system does not cover) has been scanned by Landsat every nine days. By international agreement Landsat imagery, which is telemetered to the Earth in at least four multispectral bands, is available to users in any part of the world. The French Satellite Pour l'Observation de la Terre (SPOT) and various Russian satellite programs produce very high quality images but, unlike Landsat, do not have continuous satellite coverage of the Earth. However, other countries and consortia (such as the European Space Center) make contributions to existing programs, as in the case of Britain and Australia. At the time of writing China has successfully launched an extraterrestrial satellite, recalling the early interest of the Chinese in gunpowder and rockets, and the United States has an operating imaging system on Mars.

Summary and Conclusion

An attempt to bring the subject together, after a long hiatus, is Geography: A Modern Synthesis by Peter Haggett. This is suggested by the titles of a selection of chapters of his book: "The Fertile Planet"; "Environment Risks and Uncertainties"; "Ecosystems and Environmental Regions"; "Resources and Conservation"; "Spatial Diffusion"; "Toward a Regional Convergence"; and "Outer Space, Inner Space." Analysis may give partial answers, but syntheses is essential to provide cohesion to the reality that is geography. The holistic view of the Earth as seen from space should be imperative for geography as a unified discipline, concerned with ecosystems on a fragile planet.

As indicated above, in recent decades the study of geography has suffered a decline in institutions of higher learning in the United States, and it is opined that there will only be improvement by the re-establishment of the subject in high schools. The abysmal ignorance of most Americans about the world, which may be a legacy of isolationism, contrasts with the situation in other (even so-called Third World) countries where geography is taught at all levels. In those colleges and universities in the United States where the subject continues to be taught it is often necessary to recruit faculty worldwide. If wars in the Middle East and elsewhere are not enough to goad Americans into giving more attention to the subject, then perhaps increasing dependence on resources from overseas will provide the impetus to achieve this end.

See also Demography ; Maps and the Ideas They Express .

bibliography

Agnew, John, David N. Livingstone, and Alisdair Rogers, eds. Human Geography: An Essential Anthology. Oxford and Cambridge, Mass.: Blackwell, 1996. See especially: David Harvey, "On the Present Condition of Geography"; J. B. Harley, "Deconstructing the Map"; Yi Fu Tuan, "Space and Place"; Torsten Hager-strand, "Diorama, Path and Project"; and Stan Openshaw, "A View of the GIS Crisis in Geography."

Clark, Michael J., Kenneth J. Gregory, and Angela M. Gurnell, eds. Horizons in Physical Geography. Basingstoke and London: Macmillan Education, 1987. See especially: Roger G. Barry, "Perspectives on the Atmosphere"; Keith M. Clayton, "Perspectives on the Geosphere"; and Richard J. Chorley, "Perspectives on the Hydrosphere."

Cosgrove, Denis. Apollo's Eye: A Cartographic Genealogy of the Earth in the Western Imagination. Baltimore and London: John Hopkins University Press, 2001.

Diamond, Jared. Guns, Germs, and Steel: The Fates of Human Societies. New York: Norton, 1999.

Entrikin, J. Nicholas, and Stanley D. Brunn, eds. Reflections on Richard Hartshorne's "The Nature of Geography." Washington, D.C.: Association of American Geographers, 1989.

Haggett, Peter. Geography: A Modern Synthesis. 2nd ed. New York: Harper and Row, 1975.

Hartshorne, Richard. Perspective on the Nature of Geography. Chicago: Published by Rand McNally and Company as Monograph No. 1, 1959.

. The Nature of Geography: A Critical Survey of Current Thought in the Light of the Past. Lancaster, Pa.: The Association, 1939. Reprint: Westport, Conn.: Greenwood Press, 1976.

Harvey, David. Spaces of the Capital: Towards a Critical Geography. Edinburgh: Edinburgh University Press, 2001.

Katz, Cindi, and Janis Monk, eds. Full Circles: Geographies of Women over the Life Course. London and New York: Routledge, 1993.

Lewis, Martin W., and Karen E. Wilgen. The Myth of Continents: A Critique of Metageography. Berkeley and Los Angeles: University of California Press, 1997.

Livingston, David N., and Charles W. J. Withers, eds. Geography and Enlightenment. Chicago: University of Chicago Press, 1999.

Quaino, Massimo. Geography and Marxism. Translated by Alan Braley. Totowa, N.J.: Barnes and Noble, 1982.

Romm, James S. The Edges of the Earth in Ancient Thought: Geography, Exploration, and Fiction. Princeton, N.J.: Princeton University Press, 1992.

Spate, O. H. K. "Toynbee and Huntington: A Study in Determinism." Geographical Journal 118 (1952): 406428.

Thrower, Norman J. W. Maps and Civilization: Cartography in Culture and Society. 2nd ed. Chicago and London: University of Chicago Press, 1999.

Norman J. W. Thrower

Geography

views updated May 17 2018

GEOGRAPHY

GEOGRAPHY Although the spatial connotations of the name "India" have varied substantially over time, this article will consider the components of the region included within the British Indian empire just prior to its partition in 1947, that is, the present republics of India, Pakistan, and Bangladesh, plus the small neighboring states of Ceylon (now Sri Lanka), the republic of the Maldives, and the former British protectorates of Nepal and Bhutan. Excluded from this purview will be Burma (now Myanmar), all or parts of which were included within the Indian empire between 1824 and 1937, and Afghanistan, a British quasi-protectorate from 1880 to 1919. The total area being considered comes to just over 2 million square miles (5.12 million sq. km). In latitude, the region extends from roughly 37° 5′ north, near the meeting point of the borders of Afghanistan, China, and the disputed state of Kashmir, south to 8° 3′ north, at Kanyakumari (Cape Comorin), the southern extremity of the Indian mainland, or 0° 40′ south at the southernmost Maldivian atoll. The region's longitudinal extent is from 60° 24′ east, at the westernmost point of the Pakistani province of Baluchistan, to 91° 27′ east near the point of convergence of the borders of the Indian state of Arunachal Pradesh, China, and Myanmar.

In respect to both physical and human geography, the area we shall be considering is exceedingly varied, more so than the whole of Europe. In this essay we shall limit our discussion to a consideration of its landforms and their associated land uses, other aspects of physical geography, economic geography, and settlement patterns. For discussions of historical, political and social geography, including such important topics as administrative structure, religion, language and caste, readers should refer to appropriate articles elsewhere in this work.

Terrain Regions and Associated Land Use

The region is conventionally divided into three main physical divisions: the northern mountain wall, the Indo-Gangetic Plain, and the so-called peninsular massif, including Sri Lanka and several island groups in the Indian Ocean. The northern mountain wall extends from the shores of the Arabian Sea northeastward through much of Pakistan to that country's northern border with Afghanistan; then in a southeasterly and easterly direction through northern India, Nepal, and Bhutan, along or in close proximity to the border of China; and finally, roughly southwestward along the border between India and Bangladesh on the west and Myanmar on the east, to the Bay of Bengal. Thus, it forms a barrier that largely isolates the region from the rest of Eurasia, giving rise to the concept of a distinct Indian subcontinent. This isolation gives the region a distinct climatic regime, as well as distinctive fauna and flora, and has served throughout history to limit interaction between its inhabitants and those of neighboring lands.

The northern mountain wall

The northern mountain wall consists of numerous mountain ranges of relatively recent geological origin and also includes several distinctive intermontane vales. The principal mountain range, the greater Himalayas, forms a great arc, roughly 1,500 miles (2,400 km) long, occupying the central portion of the border region. A number of its peaks exceed 25,000 feet (7,750 m). Among these, on Nepal's border with the Xizang (Tibetan) Autonomous Region of China, is Mount Everest, the world's highest summit, at 29,028 feet (8,848 m). Several ranges run parallel to and south of the greater Himalayas. These include the lesser Himalayas and the much lower Siwaliks, fronting on the Indo-Gangetic Plain. Toward their western limit in Kashmir, the Himalayas are roughly paralleled also by several ranges to their north. The most prominent of these, running along the border with the Xinjiang Uygur (Sinkiang) Autonomous Region of China, is the Karakoram Range, whose tallest peak, K2 (Mt. Godwin Austin) is among a cluster of closely grouped summits rivaling those of the Himalayas in elevation. In both the Karakorams and the Himalayas, glaciers are prominent. Those of the former range are the largest in the world outside the polar regions. The principal vales within the Himalayan system are those of Kashmir and Nepal, each exceedingly fertile and constituting the hearth of a distinctive and enduring culture. Natural vegetation and land use within the Himalayan region varies greatly. Deforestation is a major ecological problem. The density of the forest cover tends to decrease from east to west; windward slopes are much more heavily wooded than leeward slopes; and altitude exercises a major climatic control analogous to the influence of latitude at lower elevations. Terraced agriculture is common at elevations up to 8,000 feet (2,400 m), and pastoral activity occurs at many levels.

The mountains extending along Pakistan's boundary with Afghanistan and covering much the greater part of the province of Baluchistan are arrayed in numerous discontinuous chains. Wildly jagged and mainly barren in aspect, they include relatively few peaks in excess of 10,000 feet (3,000 m). Among the more prominent ranges (proceeding from north to south) are the Safed-i-Koh, the Sulaimans, the Kirthar Range, and, along the Arabian Sea coast, the Makran Range. Between the ranges, especially in the southwest, are numerous arid depressions, while cutting through the mountains farther north are several fabled passes, most notably the Khyber, Kurram, and Bolan, through which streams of invaders from Central Asia intermittently made their way into the subcontinent. Agriculture in this region is scant and dependent on several forms of irrigation, including karez, ingenious gravity-controlled tunnel systems. Pastoral activities also engage much of the largely tribal and nomadic population, many of whom—when political conditions permit—travel hundreds of miles between summer and winter pasturelands in Afghanistan and Pakistan.

The ranges forming the eastern segment of the northern mountain wall also consist of numerous discontinuous chains, few of which are much more than a hundred miles in length. Summits range from more than 15,000 feet (4,500 m) in the north to less than 5,000 feet in the south. While less jagged than the mountains of the west, their dense vegetative cover, torrential streams, endemic diseases, and, in former times, head-hunting inhabitants severely inhibited passage through the frontier region. The area forms the home of a host of tribal groups, who still mainly practice shifting slash-and-burn agriculture, supplemented by hunting, fishing, and gathering. Increasingly, however, these groups are turning toward more sedentary forms of cultivation.

The Indo-Gangetic Plain

Immediately south of the northern mountain wall lies the Indo-Gangetic Plain, the world's largest continuous expanse of alluvial (river-deposited) soil. This plain extends with no perceptible break from the Arabian Sea to the Bay of Bengal. In the west it is watered by the Indus and its five main tributaries, the Chenab, Jhelum, Ravi, Beas, and Sutlej (from west to east). These rivers all drain into Pakistan from areas held by India (largely from the disputed state of Jammu and Kashmir). When the 1947 partition interposed an international boundary within the Indus drainage basin, a major problem arose over water allocation. This issue was largely resolved, however, by the 1960 Indus Waters Treaty, whereby all the water of the three western rivers was allocated to Pakistan, while that of the Ravi, Sutlej, and Beas went to India. To make this division work, it was necessary to build new canals within Pakistan to link the western to the eastern rivers, thereby establishing the world's greatest integrated irrigation system.

Most of the Indian portion of the plain is watered by the Ganges (Ganga) and its tributaries, chief among which are the Yamuna, a right-bank tributary, and (from west to east) the Ghaghara, Gandak, and Kosi, all left-bank tributaries. These all emanate from the Himalayas and have relatively even year-round flows, whereas the Chambal, Ken, and Son rivers, flowing out of the peninsular massif, have very uneven regimes with maximum flows during the summer (wet monsoon) season. In the far east, the plain extends along the course of the mighty Brahmaputra, the flow of which far exceeds that of the Ganges. The Ganges and Brahmaputra systems converge in what is now Bangladesh to form an enormous delta, which is exceedingly prone to devastating floods from high river levels and surges of ocean water from tropical cyclones in the Bay of Bengal.

With maximum elevations lower than a thousand feet (300 m) above sea level, the average slope of the plain is less than one foot per mile. The inhabitants of the plain, especially in its upper reaches, distinguish between terraces of old alluvium (bhangar) and riverine tracts of new alluvium (khadar), often several tens of feet lower. The khadar tracts are more prone to flooding in the summer high-water season and tend to be more fertile than the bhangar, but, as modern canal irrigation is located mainly on the latter, the initial advantage enjoyed by the khadar has largely been reversed. Additionally, much irrigation comes from deep, mechanized tube wells tapping the vast reserves of groundwater found throughout the plain.

Virtually all of the Indo-Gangetic Plain is intensively cultivated. Over much of Bangladesh, rural population densities exceed 2,000 per square mile (roughly 800 per sq. km), higher than in any other predominantly rural region of the world, while densities of more than a thousand per square mile are found in many other plains localities. The plain thus supports nearly two-fifths the total population of India, more than four-fifths that of Pakistan, almost all that of Bangladesh, and roughly half that of the subcontinent as a whole. Two, or even three, crops per year are possible in well-watered regions. Little remains of the original vegetative cover, but villagers maintain groves of trees for fruit, shade, and firewood.

Straddling the Indo-Pakistani border is the Thar Desert, most of which, from a topographic perspective, may be considered an extension of the Indo-Gangetic Plain, despite the protrusion through the largely wind-deposited soils of numerous starkly dramatic hilly outliers of the peninsular massif. The area supports a mixed agricultural and pastoral economy and somehow sustains population densities that would be unimaginable in comparable environments in other parts of the world.

The peninsular massif

The peninsular massif, an area of great geological antiquity, is a remnant of the vast former southern continent of Gondwanaland. It comprises a portion of the Indo-Australian tectonic plate, the northern flank of which (now covered by the alluvium of the Indo-Gangetic Plain) collided with and subsided below the northern Eurasian plate to force the rising of the Himalayas and adjoining mountain ranges. The massif includes a highly variegated assemblage of plateaus, worn-down mountain ranges, escarpments, interior basins, and riverine plains, as well as a relatively narrow west coastal plain and a somewhat wider coastal plain to the east.

In its northern reaches, between the east-west trending Vindhya Range (actually a south-facing escarpment), the drainage of the massif is mainly to the northeast. Just south of the Vindhyas lie the west flowing Narmada and Tapti Rivers. Further south, virtually all rivers of any consequence—including (from north to south) the Godavari, Krishna, Tunghabhadra, Penner, and Kaveri—flow eastward from crests close to the Arabian Sea to the Bay of Bengal. Prominent among the highlands of the massif are the Western Ghats, a long, imposing, west-facing, fault-line escarpment terminating southward in the Nilgiri Hills. Just south of the Nilgiris lies the Palghat, the principal pass between the east and west coastal plains and, in the far south, the Palni and Cardamom Hills. These southern hill ranges contain summits surpassing 8,000 feet (2,400 m), the highest in peninsular India. Elsewhere, few elevations reach even half that altitude. The so-called Eastern Ghats (a toponymic invention of British geographers) are actually a discontinuous series of highlands more or less parallel to the coast of the Bay of Bengal. The northeastern portion of the massif, an area known as the Chhota Nagpur, is a complex hill and plateau region that is still largely forested and the home of numerous tribal peoples. As it happens also to be a highly mineralized region, containing the best and most abundant coal and iron ore resources in India, it is now undergoing rapid industrialization and has become an area of substantial immigration. Forming the northwest edge of the massif, bordering the Thar Desert, are the Aravalli Mountains, on whose craggy summits are situated many of India's most striking hill forts. Though now militarily useless, they provide a source of substantial tourist revenue.

Vegetative cover, soils, and agricultural land use in the peninsular massif are no less varied than the terrain. Forested tracts are, for the most part, small and largely degraded as a result of repeated clearing, often by fire, for agriculture and other purposes. Although soils are generally poor in interior locations, those derived from the basaltic rocks underlying a large western tract known as the Deccan Lava Plateau are relatively rich. So, too, are the alluvial soils along the courses of the major rivers and the largely deltaic soils of the two coastal plains. The intensity of agriculture and the resultant rural population densities depend mainly on soil fertility and rainfall, though irrigation by canals, tube wells, and tanks also plays a major role in determining agricultural productivity. The southern highlands are major areas of plantation agriculture.

The physical geography and land use of Sri Lanka largely mimic the patterns found in the southern peninsular portion of India. To the west of Sri Lanka lie the Maldives, a group of tropical atolls now comprising an independent republic. The Laccadive Islands, a union territory of India, form a similar island group to their north. Finally, in the Bay of Bengal, closer to Myanmar than to the Indian mainland, are the little developed, largely tribal Andaman and Nicobar Islands, which also comprise an Indian union territory.

Climate

No less than terrain, climate profoundly impacts the livelihood and lifestyles of the peoples of the region. Its dominant aspect, the monsoon regime, is more pronounced here than in any other part of the world. The temperature, wind, moisture, and rainfall patterns associated with this regime combine to produce three seasons: a hot, dry season from about March to mid-June; a hot, wet season from about mid-June to the end of September; and a relatively cool, dry season from early October to February. These are modal conditions; the actual duration of the three periods will vary not only from one part of the region to another, but also from year to year.

From March to mid-June, the temperature difference between the Indian Ocean and the more rapidly heating subcontinent becomes increasingly great, and a system of low atmospheric pressure becomes well established over the latter region. Increasingly steady landward winds bring considerable humidity, but little rain, to most of the region, creating a dusty period of steadily increasing discomfort and physiological stress.

Around late June the jet stream, a fast-moving current of air flowing eastward at altitudes of 25,000 to 35,000 feet (7,625 m–10,675 m), typically shifts from a position just south of the Himalayas to a path several hundred miles farther north, beyond the Tibetan Plateau. This shift usually occurs within a short time span and generates a rapid inflow of moisture-laden air, usually accompanied by torrential downpours. This phenomenon, known as the "burst of the monsoon," marks the onset of the so-called wet or southwest monsoon season. Despite its simplified directional name, the general flow of monsoon air is actually in two branches, one blowing eastward from the Arabian Sea across the west coast of India, and another northward from the Bay of Bengal. Monsoon rains are then produced in several ways, all involving the rising of air and its resultant cooling below the dew point, the temperature at which atmospheric saturation occurs: (a) when air streams are forced upward by intervening highlands such as the Western Ghats or the Himalayas, (b) when fronts form due to the confluence of the two main air streams, and (c) when the diurnal heating of the atmosphere causes sufficient upward convectional flow. Conversely, where monsoon winds cross highland barriers and descend their leeward slopes, the air masses heat up and increase their moisture-bearing capacity. This results in relatively dry "rain-shadow" conditions. The dry belt just east of the Western Ghats is the most prominent area exemplifying this phenomenon.

In the cool, dry season, that of the so-called northeast monsoon, the prevailing wind pattern is essentially reversed. As the Eurasian land mass cools, atmospheric pressures over the much warmer Indian Ocean become lower than those on the continent, and the flow of dry air is mainly from the northeast and from land to sea. Additionally, the desiccating effect is enhanced because of the descending flow of air from the Tibetan Plateau and the subcontinent's northern mountain rim onto the relatively low-lying terrain to the south.

Over most of the subcontinent, some two-thirds to four-fifths of the total rainfall is registered in the wet season. Total annual precipitation varies from around 450 inches (over 11 m) at Cherrapunji in the Khasi Hills (Shillong Plateau), lying directly in the path of the eastern branch of the monsoon in northeast India, to as little as 5 inches in the heart of the Thar Desert. But on the plains of India and Bangladesh, rainfall exceeds 100 inches (roughly 2.5 m) in only a few small areas.

Coastal areas experience monsoon rains as much as two months earlier than do deep interior locations and the length and intensity of their monsoon season is commensurately greater. Apart from this broad generalization, details of the pattern of rainfall vary greatly in relation to local topography, in respect to both elevation and the orientation of highlands and coasts relative to the paths of prevailing winds. Southeastern India and eastern Sri Lanka, for example, are exceptional in experiencing maximum rainfall during the period from October to December, because the initially dry winds of the northeast monsoon pick up moisture as they cross the Bay of Bengal and then lose it as they blow across these areas. Another area departing from the general seasonal regime is a belt along the southern flank of the northern mountain wall, which experiences the tail-end effects of winter storm systems moving eastward from the Mediterranean Basin.

The monsoons play a pivotal role in agriculture. Their substantial year-to-year variability in terms of timing and total precipitation makes for considerable uncertainty in crop yield. As a rule, the greater the average precipitation, the more dependable the rainfall, but few areas are totally free from the prospect of drought. Devastating tropical cyclones, which are most common along the northern shores of the Bay of Bengal, constitute yet another major climatic hazard.

Temperatures over most of the region generally peak in late May or early June, just prior to the cooling downpours of the southwest monsoon. A secondary maximum is often experienced in September or October, when precipitation wanes. In the northwest, where total annual rainfall is low, July is usually the warmest month. Temperature ranges tend to increase with distance from the coast and also with latitude. The difference between the mean temperature of the coolest and warmest months in Sri Lanka or near the southern tip of India, for example, fluctuates by no more than a few degrees around the annual mean of 80°F (27°C). In marked contrast to these marine conditions is the continental regime of Punjab. In Lahore, for example, the mean daily high temperature in late June is 104°F (40°C), while the mean daily low in January is only 39°F (4°C). Freezing temperatures, however, are rare on the northern plains and are virtually unknown on the peninsular massif. Temperatures are moderated substantially in highland locations, with many mountain towns, especially in the Himalayas, serving as "hill stations," where those who can afford to do so may escape the scorching heat of the plains.

Natural Vegetation and Fauna

Over the ages, some four-fifths of the forests that once blanketed the greater part of the region have been removed to make way for agriculture and for other uses. As in the past, however, the remaining natural vegetation reflects the distribution of rainfall. Areas of high precipitation (over 80 or so inches, or roughly 2 m) are marked by tropical, broad-leaved evergreen forests, a larger proportion of which survive than in forests in drier areas that are easier to clear. With increasingly dry conditions, one encounters scattered stands of mixed forests, partially evergreen and partially deciduous. Forests in nonmountainous areas with less than 40 inches (101 cm) of rainfall are entirely deciduous. Still drier climates are characterized successively by open scrub, or jangal, then grassland, and ultimately scrubby desert vegetation. Coniferous forests are confined to the northern mountain rim. Many of the region's 17,000 or so flowering species are found only in the subcontinent.

The subcontinent, along with most of Southeast Asia, forms a major part of the highly diverse Oriental or Indian zoogeographic province. Among its characteristic primates are the rhesus monkey and Hanuman langur. Its carnivores include tigers, leopards, and lions (confined to the Gir Forest in the Kathiawar Peninsula), mongooses, jackals, and foxes. Wild herds of elephants survive in several protected areas, as does the Indian rhinoceros. Other mega-fauna include several species of deer and wild bison. Among the more than 1,200 species of birds are herons, storks, ibis, flamingos, and cranes. Reptiles include crocodiles, gavials and cobras, pythons, and nearly 400 other species of snakes, many of which are poisonous. The variety of fish, insects, and other forms of animal life is exceedingly great.

Economic Geography and Settlement Pattern

Agriculture, forestry, and fishing

Despite sweeping economic changes since independence, agriculture, including animal husbandry, remains the predominant occupation in almost all parts of South Asia, accounting for about three-fourths of the labor force in Nepal, two-thirds in Bangladesh, three-fifths in India, half in Pakistan, and one-third in Sri Lanka. In all those countries its impress on the landscape is pervasive and profound. The scale of most operations, however, is quite small (except in parts of Pakistan and several plantation areas). Moreover, as per capita returns from agriculture are much less than in other sectors of the economy, the proportions of the gross national product derived from that sector are less than one-fourth the total in all of the countries named, except Nepal. The forms of agriculture and the combinations of crops and livestock vary enormously throughout the region, depending on climate, irrigation, soil type, proximity to urban markets and ports, culture and other factors. Staple crops, especially grains, account for a substantial majority of the total area sown. Other staples include a wide variety of pulses, such as gram (chickpeas), lentils, and peanuts. Potatoes are a major crop in the Himalayan region and, because of improved storage facilities, have become increasingly important on the Gangetic Plain as well.

Over most of the subcontinent, farmers recognize two growing seasons: the kharif season, in which crops mature over the wet monsoon months, and the rabi season, in which crops mature more slowly during the cool season and are harvested after the initial summer rains, or earlier if irrigation permits. In very wet areas, including parts of Bangladesh and Assam, three crops per year may be possible. Rice is almost always the principal staple crop where there is sufficient rainfall, in that it yields more calories per unit area than any other grain. Annual rainfall of 80 or more inches of rain will permit the cultivation of two crops of rice. With 40 to 80 inches (101–202 cm), farmers will typically seek to produce a summer crop of rice and a winter crop of wheat, pulses, or vegetables. Somewhat drier areas will grow mainly wheat, almost always as a rabi crop, with gram or vegetables in the kharif season. Generally, 15 to 20 inches (38–50 cm) of rain will suffice for a single crop of wheat. Where soils are poor, however, sorghum (jawar) and various millets (bajra, ragi, etc.) are likely to take the place of wheat. Maize is grown mainly in hilly and mountainous areas, especially in the northern mountain regions. Inter-tillage of grains, other than rice, with pulses is common.

Livestock occupy a very important place in the region's agriculture. Cattle (largely of the distinctive zebu breed) and buffalo are numerous and are utilized as draft animals, for milking, as providers of leather, and as sources of fertilizer. Goats and sheep are also multipurpose animals. Camels, horses, and donkeys serve mainly as beasts of burden, as do yaks and dzo (yak-cattle hybrids) at high elevations in the Himalayas. Per capita consumption of meat, however, is quite low. Other sources of animal protein include milk (production of which has increased enormously in recent decades), eggs, and fish. Fish are derived from both marine and freshwater sources and, increasingly, from aquaculture in flooded rice paddies.

Many areas of the subcontinent specialize in the cultivation of nonstaple crops. The foremost of these is cotton, the chief cash crop in much of both Indian and Pakistani Punjab, much of the interior of Maharashtra, and parts of Gujarat. Jute is a major fiber crop in much of Bangladesh and parts of West Bengal, Bihar, and Assam. Sugar is very widely cultivated in small quantities, and forms a major specialty crop over much of western Uttar Pradesh, Indian and Pakistani Punjab, and the leeward slopes of the Western Ghats. Tobacco is especially important in the delta regions of Andhra Pradesh.

Tree crops, which vary greatly throughout the region, include mangoes, oranges, and other citrus species. Coconuts, which provide food, vegetable oil, fiber, fuel, and building material, are of particular importance in the Indian state of Kerala and in Sri Lanka. Other useful palm species include the palmyra (sugar palm), areca (betel palm), and dates. Kashmir and other mountain regions are sources of apples and other temperate fruits.

Plantation cultivation is restricted to only a few relatively small areas. Unlike most agriculture in South Asia, it involves large-scale, highly capitalized corporate enterprises and employs large gangs of mainly immigrant labor. The chief plantation crop, by far, is tea. Neatly manicured tea plantations cover much of the highlands of Sri Lanka, which has surpassed India as the world's leading tea exporter. They also occupy similar terrain in the Indian states of Kerala and Tamil Nadu and are important along the lower flanks of mountains in West Bengal and Assam, as well as in northern Bangladesh. Coffee is grown in Sri Lanka, Kerala, and Karnataka, at elevations lower than those favored for tea, and rubber plantations are found at still lower elevations in Sri Lanka and Kerala.

Commercial and subsistence forestry are practiced in many scattered localities, mainly where rugged terrain makes cultivation impracticable. Among the many timber species, teak (sometimes grown on plantations) and sal are most prominent, while highly priced sandalwood and rosewood are also important. Pines and other conifers are extensively cut for timber in the western Himalayan region. Bamboo and a great variety of wild species are gathered as building materials and for fuel and other purposes.

Nonagricultural rural activities

Throughout recorded history, agriculture has been supplemented by a wide range of ancillary activities that, in the context of regional socioeconomic systems, have been carried out largely by members of occupationally specialized castes, such as weavers, potters, carpenters, well diggers, leather workers, traders, and so on. Muslims and followers of other faiths in rural India were also integrated within this hereditary system, in which services rendered were compensated by payments in grain or other goods, rather than by cash. Over the past century or more, with the increasing commercialization of agriculture and the supplanting of craft-made goods by the products of modern factories, this system has been greatly eroded and in some areas has virtually ceased to exist. Nevertheless, many traditional occupations continue to be pursued, often precariously, by particular castes, though increasingly within a market-based nexus.

A characteristic feature of rural India is the periodic market (hat) to which households resort to obtain goods that are typically not available in most villages: spices, cooking oils, brassware, bangles, saris, textiles, tools, and so forth. Most such markets occur on a particular day of the week and are served largely by itinerant merchants, craft specialists, and local hawkers of agricultural produce. There are also larger, more widely scattered seasonal markets, often associated with the observance of religious festivals (melas), at which a broader range of goods and services may be procured. Entertainers, astrologers, medical practitioners and, nowadays, government-sponsored purveyors of useful information (e.g., on family planning) are prominent actors at such venues. Apart from the periodic markets, there are thousands of small market towns (mandis), combining both wholesale and retail functions for the surrounding countryside. Their importance has grown greatly with the commercialization of agriculture in recent decades; and many such towns are also centers for the provision an expanding array of administrative, educational and medical services.

Rural settlement pattern

In contrast to the United States, almost all cultivators in the region reside in agglomerated settlements that vary in size from hamlets of only a dozen or so households to large villages with 5,000 or more inhabitants. A substantial majority, however, live in villages ranging in population from 500 to 2,500. Most villages are of an irregular shape, growing outward, with no predetermined plan, from an initial nuclear locality. Within almost all villages there is a high degree of residential segregation by both religion and caste. In general, households of the higher castes and those of the caste that initially settled the village will occupy the central portion of the village, while lower castes and (in India) Muslim groups will occupy peripheral locations. The degree of segregation of the lowranking castes, especially those formerly held to be untouchable, may be such as to consign them to separate tributary hamlets, which may be as much as a half-mile distant from the main village settlement.

Within the village, lanes and alleys between houses are typically narrow and winding, often terminating in dead ends. Most houses are built mainly of dried mud mixed with straw and smoothed on the exterior with a cow dung and mud paste; but those of more affluent households are typically made of brick or, in some regions, even of stone. In northeastern India and Bangladesh, bamboo and various types of reeds may replace mud as a construction material. Wood is sparingly used, largely for roofing beams and doorframes. Roofs are generally flat in relatively dry areas and peaked in wetter areas, with increasingly steep slopes in areas of greater precipitation. Houses of poor families typically have roofs of mud or straw (made from the stalks of rice and other grains, or from reeds), whereas those of wealthier households will most likely be of sun-baked tile. Detailed study will reveal great regional differences in vernacular forms of domestic architecture. Also varying greatly is the exterior décor of housing and the designs made on the floors of courtyards, the production of which is largely in the hands of women.

Houses typically accommodate both humans and livestock and include spaces for storing provisions for both. Granaries, however, are frequently built as separate, adjacent structures. Living space is generally scanty and furniture and other household furnishings are sparse. Many houses have a small area set aside for family worship.

While some settlements are wholly residential, most contain a variety of structures given over in whole or part to specialized functions: temples, mosques, schools, the offices of the panchayat (village council) and/or cooperative, shops, and small handicraft enterprises. Open spaces, facilitating social intercourse, are found mainly at or in close proximity to public buildings, the homes of major land owners, threshing floors, and large, widely shared wells. Village groves are also commonly situated near such wells.

In general, villages tend to be more compact in the north and west of the subcontinent than in the east and south, as well as more inward-oriented. That is, individual houses form parts of high-walled compounds, the inhabitants of which all belong to the same caste or other ethnic group and are likely to be related to one another by birth or marriage. Windows to the street are either altogether lacking or tend to be small and barred, while within the compound houses open onto interior courtyards. This arrangement facilitates the exclusion of women from public view. Toward the south and east, rural settlements tend to be more open. Spaces between houses are greater and the proportion of the population living in exterior hamlets increases.

Some regions display marked variations from the foregoing generalizations, In Tamil Nadu, some adjacent areas, and the Kathiawar Peninsula of Gujarat, for example, villages tend to be laid out on a grid pattern, which is thought by some scholars to provide a link to the supposed Dravidian founders of the ancient Indus civilization. Over most of Bangladesh and coastal Kerala rural dwellers live mainly in isolated households, rather than in village agglomerations; but, given the dense populations of those areas, few households there would be out of shouting distance from their nearest neighbors. In many tribal regions, especially in northeastern India, small hamlets, rather than proper villages, are the norm and houses are laid out on either side of a single village lane. Hamlets are also common in mountainous regions in which it is difficult to find large areas of terrain flat enough for the building of larger settlements.

Mineral resources, mining, hydroelectric power, and industry

India is well endowed with a wide range of mineral resources, whereas the other countries of South Asia are relatively mineral poor. Among India's resources are abundant and widespread reserves of coal, mostly bituminous; widespread, and largely high-grade, iron ore deposits; a wealth of minerals used as ferro-alloys, especially manganese; substantial quantities of bauxite, used in the manufacture of aluminum; modest supplies of copper, lead, zinc, antimony, silver, and gold; a variety of precious and semiprecious gemstones; and an abundance of such ordinary, low-value (per unit of weight) minerals as ceramic clays, refractories (for use in smelting furnaces), limestone, rock phosphate, and building stone. The chief area of concentration of mineral products is the Chhota Nagpur, but many other portions of the peninsular massif contain locales of significant mineral production.

India's coal reserves are among the greatest in the world, though the amount of coking coal, suitable for use in metallurgical industries, is limited. Thermal energy from coal-fired plants is the principal means of meeting the country's energy needs. The most important mines, those of the Jharia-Raniganj Field, near the northeastern edge of the Chhota Nagpur, provide coal of good coking quality. Their location, near major sources of iron ore, has led to a concentration of the iron and steel industry in this region. Uranium is also produced in the Chhota Nagpur and helps sustain the country's five nuclear reactors. Nuclear power, however, accounts for only a small fraction of India's total energy output.

India's most serious energy deficiency is that of petroleum and natural gas. Although production from fields in eastern Assam (the first to be developed), Gujarat, and the "Bombay High" (off the coast of Maharashtra) has grown rapidly, growth in demand has also been rapid. Hence, Indian petroleum production seldom exceeds 30 to 40 percent of its needs. Natural gas supplements petroleum as an energy source to only a modest degree.

Hydroelectric installations are very widespread in India and are largely associated with multipurpose river basin development plans that include irrigation, navigation, and recreation. In all, they provide approximately a sixth of all the electric power consumed in India.

Mineral resources in South Asian countries other than India include modest supplies of petroleum, natural gas, and coal, as well as a variety of minerals needed for chemical industries and construction in Pakistan; some natural gas and a little coal in Bangladesh; and various gemstones in Sri Lanka. Hydroelectricity is well developed in both Pakistan and Sri Lanka, to meet domestic needs, and in Nepal, mainly for export to India, the source of the engineering expertise for their construction. Modern developments notwithstanding, throughout South Asia firewood, cow dung, and other traditional organic sources continue to be major providers of energy in the rural economy.

Beginning with cotton milling in the mid-nineteenth century, the industrial revolution has largely transformed the traditional way of life for a substantial fraction of the population of South Asia and has significantly impacted all but the most remote of areas. Among modern manufacturing enterprises, the most ubiquitous are those related to the processing of agricultural goods and the weaving and processing of textiles. Given the differential availability and costs of the factors of production, other industries are prominent in only one or a few specialized production centers.

Industries that typically locate close to the source of raw materials are those that involve refining and loss of weight in the manufacturing process, for example, rice milling, sugar refining, cotton ginning and spinning, the processing of coir (coconut fiber), cigarette manufacture, the tanning of leather, the preparation and packaging of tea, and the smelting of minerals. Most such simple activities take place in relatively small urban centers or near plantations and mines. More sophisticated manufacturing, such as the weaving of cloth, the manufacture of shoes and wearing apparel, cement and chemical industries, light engineering industries, and so forth gravitate to larger cities. Few among the hundreds of cities in South Asia (i.e., places with populations of 100,000 or more) are without establishments devoted to several of these forms of manufacturing production.

Of particular importance in recent years has been the growth of foreign investment—mainly from the United States, but also from Japan and other developed countries—taking advantage of the abundance of cheap labor in assembling wearing apparel, footwear, small appliances, and so forth, much of which is produced for distribution through mass retail marketing corporations. Special industrial zones, with privileged production conditions and tax benefits for foreign investors, have been created in a number of ports and other cities to stimulate such development. Both Bangladesh and Sri Lanka have been important beneficiaries; in both countries clothing and accessories now account for more than half the total value of exports.

At a more technologically advanced level, metallurgical, heavy engineering, industrial chemical, pharmaceutical, and electronics industries are concentrated either in ports or locations in which bulky resources, such as iron ore and coal, can be combined with minimum transportation costs; or in other large cities (those with a million or more people), where capital and ancillary services (banking, insurance, advertising, etc.) are relatively abundant and where purchasing power is highest. Thus, one finds the manufacture of steel and steel products in Jamshedpur and other cities of the Chhota Nagpur, as well as in several port cities (including Karachi in Pakistan and Chittagong in Bangladesh, which utilize imported scrap metal and/or iron ore); automobile assembly plants in Chennai, Delhi, Pune, and Kolkata; high-tech electronics and aviation industries in Bangalore; pharmaceutical producers and petroleum refineries in Mumbai; ship building in Vishakhapatnam; and so forth. There are very few manufactured goods that India cannot produce. But, because it is not yet fully competitive with more advanced countries, some industries survive only because of high protective tariffs. As India moves increasingly toward a free trade regime and away from government-owned industries, the future of inefficient producers is likely to be bleak.

The global revolution in electronic communication has ushered in two major developments of which India has become a major beneficiary: first, the spectacular growth in the production and exporting of computer software; and subsequently, the outsourcing, mainly by American firms, of consumer services in such fields as banking, insurance, health maintenance industries, and the like. This has resulted in an enormous boom in the training and utilization of relatively cheap, but highly skilled, labor in such cities as Bangalore, Hyderabad, and Chennai. This has had a major favorable impact on India's balance of payment.

The globalization of labor markets has also led to significant flows of South Asian labor, both skilled and unskilled, to developed countries in North America and Europe as well as to the Middle East. Particularly important source areas in respect to the Middle East are Pakistan, Bangladesh, and Kerala. Remittances from overseas have been an important source of capital for both investment and consumption in many parts of South Asia.

A final aspect of globalization that has significantly impacted South Asia and that holds tremendous promise of future change is tourism. The number and variety of sites worthy of a visit defies the imagination: Mughal architectural masterpieces, such as the Taj Mahal, the nearby ruins of Fatehpur Sikri, or the Red Fort of Delhi; the great Hindu, Buddhist, and Jain temples (including cave temple complexes such as those at Ajanta and Ellora) found throughout India; the dramatic fortress cities of Rajasthan; the ruins of such ancient cities as Mohenjo Daro, Harappa, and Taxila in Pakistan; distinct cultural niches such as the Vale of Nepal or Manipur; areas of scenic splendor, such as Kashmir; national parks and other pristine areas reserved for nature study, trekking or serious mountaineering; and tropical beaches such as those of Kovalam, Goa, and the Maldives. Some of these have long been favorites of privileged elite classes, while others became meccas for more penurious globetrotters in and since the 1960s. Most sites await the development of modern hotels and improved travel facilities to reach their full potential. In relation to their populations, the Maldives, Nepal, and Sri Lanka have reaped the greatest gains from tourism.

Infrastructure

The fact that most of the countries of South Asia could develop to the extent that they have over the past century is attributable to no small degree to changes in their economic infrastructure, begun during the colonial period and, in many respects, accelerated in the period since independence. In the case of Nepal and Bhutan, however, the physical isolation resulting from their mountainous terrain creates a serious and persistent obstacle to development.

The building of railroads was of particular importance. Even before independence, India had the world's third-largest rail system, a major catalyst for subsequent development. While railroad mileage has not been greatly extended in South Asia since independence, the quality and volume of service has continued to expand. Today, Indian railroads log more passenger miles per year than in any other country in the world. Relatively speaking, however, more effort went into the building of roads. Although per capita ownership of automobiles remains very low by world standards, few areas are now out of reach of bus service, and the transportation of goods by truck has greatly multiplied. Few villages cannot be reached by road, at least during the dry season. While the rail system remains a government monopoly throughout South Asia, bus systems have become increasingly privatized. In the deltaic terrain of Bangladesh, the complex network of broad rivers greatly hampers overland transport, especially in the rainy season; hence, river traffic, largely in simple "country boats," carries a greater volume of goods than either road or rail. Finally, air services, both nationally and privately owned, are also expanding rapidly throughout South Asia.

Among the colonized areas of the world, India stood out for the quality of its institutions of higher learning. In addition to its indigenous traditions of learning, universities, along Western lines, were established in its major cities as early as 1857 and have proliferated in subsequent generations, especially so in the period since independence. In contrast to this emphasis on higher education, the reduction of illiteracy in South Asia, especially in the countryside and among females, has proceeded slowly. The most notable exceptions are in Sri Lanka and the Indian state of Kerala. Both attach great importance to education at all levels and have achieved virtually universal literacy. In general, Pakistan and Bangladesh lag behind India in their educational attainments, and the reliance of ordinary Pakistanis on education in traditional religious education in madrassah, given the paucity of government-sponsored alternatives, presents a significant potential obstacle in the way of economic development and the integration of Pakistan into the modern world.

Finally, one must note the remarkable extension, throughout South Asia, of government services in the form of community development schemes, public health clinics, government-aided cooperatives, and so forth. Nongovernmental organizations, including indigenous self-help programs such as those of the Grameen Bank in Bangladesh and SEWA (Self-Employed Women's Association) in India have also made enormous strides. All of these are engaged in one way or another in improving human capital. Therein may lie the best hope for the future of South Asia.

Urbanization and urban morphology

Despite the fact that South Asia is still among the least urbanized regions of the world, its total number of urban dwellers at the beginning of the twenty-first century was already in excess of 400 million. Of these, roughly 135 million lived in cities of a million or more inhabitants. The proportion of urban dwellers to the total population in 2003 was about 36 percent in Pakistan, 28 percent in India, 25 percent in Sri Lanka, 24 percent in Bangladesh, and only 13 percent in Nepal. Among the region's cities are some of the world's largest and most rapidly growing. The 2001 populations, to the nearest million, of India's largest metropolitan areas (suburbs included) were: Mumbai (Bombay), 16 million; Delhi, 13 million; Kolkata (Calcutta), 13 million; Chennai (Madras), 6 million; Bangalore, 6 million; Hyderabad, 6 million; Ahmedabad, 5 million; and Pune, 4 million. Another twenty-four cities had populations of from 1 million to 3 million each. The leading cities in the remainder of South Asia were: Dhaka, in Bangladesh, 12 million; Karachi and Lahore, in Pakistan, 9 million and 5 million; and seven others, with populations from 1 million to 2 million each.

As with villages, though to not quite the same degree, cities tend to be residentially segregated. Thus, many neighborhoods will be identifiable by their dominant religion, caste, language group (in that many cities have substantial populations of interstate or interprovincial migrants), and/or economic class. Within such neighborhoods, social clubs, places of worship and the purveyors of foods, other goods and services cater largely to the needs and preferences of the socially dominant group.

The spatial arrangement and architecture of cities also reflects the time of construction of particular areas and the functions they are intended to serve. There is, for example, usually a densely crowded old city core, dating from the pre-colonial period. Often walled (though many city walls have now been dismantled in whole or part), the old city is typically marked by a confusing maze of narrow streets and alleys and is likely to include the mansions of the original urban gentry, the principal places of worship, and traditional markets, often containing streets given over wholly to the vending of particular goods such as grains, sweets, saris, jewelry, and hardware. The burgeoning of urban population has, almost invariably, caused an overflow from the original core area to neighborhoods that are somewhat less crowded and slightly more regular in their layout. To one side of many an old city one often finds an adjacent area or areas known as "civil lines" or the "cantonment." As their designations imply, these were originally built mainly to accommodate the British civilian and military personnel and their families and household servants. Such areas are spaciously and regularly laid out and are marked by colonial-style bungalows, with broad verandas and adjacent gardens. Since the departure of the British, the civil lines and cantonments have been occupied mainly by Indians performing the same functions as those of the former colonial overloads.

In some industrial cities one finds squalid "coolie lines," built by factory owners to accommodate the mill hands, a large proportion of whom are likely to be immigrants from surrounding rural areas. Similar "lines" are found adjacent to many mines, plantations, and railway workshops. Since independence, there has been a proliferation of various types of housing "colonies," located mainly along and beyond the urban periphery. For relatively affluent families, these consist mainly of single-family dwellings, with small surrounding gardens; but more commonly they are comprised of drab multifamily apartment houses, many of which reflect the housing styles of the Soviet Union and Eastern Europe during the Stalinist period.

At an economic level below that of ordinary, regularly employed salaried laborers are masses of slum dwellers who, for the most part, survive from irregular and low-paid work, or who live from begging and other socially undesirable activities. This segment of the population occupies dilapidated tenements or inhabits flimsy, makeshift huts. One must also note that South Asian cities have more than their share of roofless pavement dwellers.

Finally, as in all major cities in an age of economic globalization, the megacities of South Asia are also witnessing the growth of central business districts, with skyscraper office buildings, major banking establishments, a wide range of other commercial services, shopping malls, luxury hotels, restaurants, and so forth. One may safely assume an acceleration of urbanization in the decades ahead and increasing concentration of economic activity and wealth-generation in South Asia's cities, widening income gaps between the economic elite and the masses, greater urban congestion and pollution, increased crime, and mounting challenges to urban governance.

Joseph Schwartzberg

BIBLIOGRAPHY

Ahmed, Aijazuddin, ed. Social Structure and Regional Development: A Social Geographic Perspective, Essays in Honour of Professor Moonis Raza. Jaipur and New Delhi: Rawat Publications, 1973.

Ahmed, Aijazuddin, and A. B. Mukerji, eds. India: Culture, Society and Economy: Essays in Honour of Asok Mitra. New Delhi: Inter-India Publications, 1985.

Breton, Roland J.-L. Atlas of the Languages and Ethnic Communities of South Asia. Walnut Creek, Calif., London, and New Delhi: Altamira Press: 1997.

Crane, Robert I., ed. Regions and Regionalism in South Asian Studies: An Exploratory Study. Monograph no. 5, Duke University Monograph and Occasional Paper Series. Raleigh, N.C.: Duke University, 1967.

Johnson, Gordon. Cultural Atlas of India: India, Pakistan, Nepal, Bhutan, Bangladesh & Sri Lanka. New York: Facts on File, 1996.

Maloney, Clarence. Peoples of South Asia. New York: Holt, Rinehart and Winston, 1974.

Misra, R. P., ed. Contributions to Indian Geography. 13 vols. New Delhi: Heritage, 1983–1994.

Mukerjee, Radhakamal. Man and His Habitation: A Study in Social Ecology. Mumbai: Popular Prakashan, 1968.

Schwartzberg, Joseph E. Occupational Structure and Level of Economic Development in India: A Regional Analysis. New Delhi: Office of the Registrar General, 1969.

Schwartzberg, Joseph E., ed. A Historical Atlas of South Asia. New York and Oxford: Oxford University Press, 1992.

A Social and Economic Atlas of India. Delhi and New York: Oxford University Press, 1987.

Sopher, David E., ed. An Exploration of India: Geographical Perspectives on Society and Culture. Ithaca, N.Y.: Cornell University Press, 1980.

Spate, O. H. K., and A. T. A. Learmonth. India and Pakistan: A General and Regional Geography. 3rd ed. London: Methuen, 1967.

Geography

views updated May 09 2018

GEOGRAPHY

GEOGRAPHY . A deeply rooted aspect of human behavior, the ordering of space is an activity that consists of establishing differences between places in terms of varied functions and degrees of meaning. Among peoples of diverse religious traditions, the most significant places are identified with special spiritual presences, qualities that set certain locales apart from ordinary, profane space. Charged with supernatural power, sacred places function as fixed points of reference and positions of orientation in the surrounding world. With the passage of time, sacred places become invested with accumulations of mythical and historical meanings in complex layers of cultural memory. When joined by paths, processional ways, or great routes of pilgrimage, sacred places form networks that may embrace local village or tribal lands, large nations, or vast regions of the globe occupied by major civilizations. These networks form sacred geographieswebs of religious meaning imposed upon the landwhere natural features and human-made symbols establish communication between the earthly and the spiritual, embodying collective values and shared norms of conduct. Sacred geographies form a unifying ground, a lasting source of remembrance and renewal for the most important aspects of individual and communal life in many cultural traditions.

The creation of sacred geographies is behavior partly anterior to the development of culture, for it stems from the marking, exploitation, and defense of territories that join humankind to the larger animal kingdom. But the articulation of landscapes with symbolic imagery and the way in which such landscapes are made to reflect layers of mythology and history also correspond to patterns of thought and complex ways of recording meaningful events that seem peculiar to humankind. The widely different ways in which sacred geographies have been organized show how humans have sought to grasp the perceived world and how they have explained their place within the cosmic schema. An examination of sacred geographies thus points to patterns of environmental cognition and ordering and to the wide range of spatial definitions that have evolved in response to different cultural needs, historical circumstances, and ecological possibilities.

This article focuses on four examples of sacred geography. They correspond to the symbolic landscapes of peoples of strikingly different social and cultural complexities who inhabit regions of varied ecologies. The first example is from the Australian Aborigines, whose gathering and hunting life in an austere desert environment was connected to systems of sacred places embedded within nature and unmarked by monumental art or architecture. These places, arranged in certain patterns within tribal territories, were thought to have been established by ancestral heroes in the Dreamingthe time of first creation. The second example is from the Maya Indian community of Zinacantan in southern Mexico, a farming people whose culture stems from an ancient native heritage. These Indians have evolved a pattern of centralization in their sacred geography that echoes similar structures among other peoples of sedentary farming life. The third example is from imperial China, where ancient beliefs concerning the worship of earth, water, and sky were expressed at great mountain shrines and in the sacred precincts of the imperial capital. In China is found the creation of a sacred geography closely tied to the concerns of a powerfully centralized state. The last example touches upon the sacred geography of medieval Europe. Though politically disunited, the peoples of Europe followed routes of pilgrimage to the periphery of Christendom and held Jerusalem, the Sacred City, to be the center of their world.

Aboriginal Australia

The Australian Aborigines are counted among the oldest human races. Their ancestors migrated from Southeast Asia perhaps thirty thousand years ago, when land bridges between New Guinea and Australia were almost certainly exposed. Throughout the millennia the Aborigines pursued an austere gathering and hunting life that was admirably adapted to their barren habitat. The Neolithic agricultural revolution never reached these isolated lands, where nomadic bands traveled within well-defined tribal territories, following seasonal rhythms in the unending search for food. But the inhabitants did not perceive the natural environment in economic terms alone; it was also seen as a storehouse of memory, replete with supernatural meanings. The flat, seemingly featureless terrain contained an invisible, magical domain in which hills, rocks, water holes, and groves were charged with sacred powers and mythical associations. Though apparently obeying the randomness of nature, such features were seen as well in terms of a specific order; people expressed their connection to them through pictographs, rock alignments, wooden sculptures, caches for totemic objects, and ceremonial places designed according to prescribed rules of organization.

The sacred sites of the Aborigines marked places where events in the Dreaming took place. This concept, which is central to Aboriginal cosmogony, concerns a time when heroes and heroines wandered over a land where there were no hills, water holes, or living things. The paths and camping places of these heroes are sacred places described in myths. The ancestral heroes also brought fire to the people as well as the laws by which people live; many such heroes eventually transformed themselves into trees, boulders, and other natural features, thus creating the landscape that exists at present. A symbolic order that related to the time of origins and the travels of creators, rather than the cardinal directions, was called into being. In these austere settings, no sharp divisions were made between animals, plants, inanimate objects, and humankind. The ancestral heroes were not distant entities but integral components of the land, and they were made part of the experience of daily life through religious reenactments of events of the Dreaming. Joined by a network of sacred places, the land itself became symbolic, affirming a coherence of the physical and mythological domains. Among the Aborigines, sacred geography was the source of authority as well as the source of tribal identity. The latter was based on a title to the land that went back to the time of first creation.

Zinacantan

Conquest, colonialism, and the advances of industrial civilization have often spelled destruction or major alteration for traditional religions. In North America, forcible removal of Indian populations from old homelands frequently meant social disintegration for those whose religious sense of belonging to a specific landscape had been destroyed. But colonial cultures have also often produced a range of creative adaptations, as subjected peoples evolved syncretistic religions in which ancient sacred geographies continued to play traditional functions. Latin America, among Indian communities in former Spanish possessions from the Rio Grande Valley in New Mexico to the Bolivian Andes of South America, is especially rich in such instances.

Among many examples that could be discussed, the community of Zinacantan in the high, forested mountain country of Chiapas, southern Mexico, is particularly well documented. The inhabitants are a Maya people, ultimately descended from those who built city-states in southern Mexico, Guatemala, and the Yucatán Peninsula during the first centuries ce. Today, these farming people live in hamlets dispersed throughout the hills, around a civic and religious center that consists of a church, school, and administrative buildings.

The visible sacred geography of Zinacantan incorporates mountains, caves, water holes, and human-made crosses erected at shrine sites at determined locations around the civic and religious center. Mountains have important economic meaning in the life of Zinacantan, but certain peaks are also considered to be homes of ancestral deities who live within. These ancestors control the mists and vapors that rise to form rain clouds on the peaks; they are able to direct the rain clouds over the community. Crosses, placed on pathways around the mountains surrounding the Zinacantan center, were borrowed by the Indians from the symbolic forms of Spanish Christianity. But the crosses are not seen in Christian terms; they are perceived as spiritual openings for communication with ancestral beings. Cave shrines are also places for communication with Yahval Balamil, the earth lord, who dwells beneath the surface of the land; he may also be reached through prayers in caves, at sinkholes, and at springs throughout the Zinacantan domain. Sacrificial offerings are regularly made, especially at the sites of cross shrines. They are most often performed by people walking on ceremonial circuits around the whole community. Shamans are the main ritualists, performing prayers and making offerings on behalf of patients. Processions of other worshipers may also follow ceremonial circuits, usually moving in counterclockwise direction. The village church and its Christian images are also frequently included in the processional itineraries. The circuits around Zinacantan are a way of establishing boundaries, a way of saying "This is our sacred center, through which the holy river flows and around which our ancestral gods are watching over us." Circuits are replicated on many levels around individual fields, houses, or other objects, symbolically establishing property rights as well as marking social spaces in the Zinacantan world. The community and land at Zinacantan are infused with the sense of being whole and sacred; sacred geography places the living community in an intimate religious bond with its natural setting and with the ancestors and gods that dwell within that setting.

Imperial China

The sacred geographies of tribes or small agrarian communities are usually encompassed by paths or roadways within relatively restricted zones. But the development of ancient empires embracing vast regions and diverse populations posed different problems of spatial symbolism. For rulers, the problem was to create symbolic orders that might tend to unify such disparate domains and polities. Great ritual centers were designed to communicate the religious and political concerns of state organizations. The symbolic structure of such places expressed the idea of a sacred geography in microcosm, through the use of monumental art and architecture.

In China, where continuity of religious themes has been maintained over millennia, imperial ritual was especially focused on two outstanding places. The first was a sacred mountain, Tai Shan, the central and most important of five sacred mountains associated with the cardinal directions and center. The worship of mountains has an ancient history in China, attested by early texts that describe peaks and the appropriate rites to be celebrated there. Some were local shrines affecting small areas, but others were majestic sovereigns that extended their influence over immense regions. These old beliefs surrounded Tai Shan, and the sacred place gradually was invested with imperial monuments throughout the centuries. In effect, the mountain became a symbol of the cosmos and the state. The mountain was given royal title during the Tang dynasty in 725, inaugurating a practice followed by successive emperors. These honorific names underlined the conception of the mountain as a producer of life forces. It was identified with rain clouds and fertility and figured as an object of worship in spring rites of planting and at the fall harvest season. It was also seen as a symbol of stability and permanence and as a preventer of droughts, floods, and earthquakes. Indeed, it was a divinity with a sacred force that could be touched by prayers and sacrifice.

Just as the mountain was a symbol of order in the natural environment, so did the emperor personify the social and moral order. A close relationship developed between ruler and mountain, for the emperor was the pivot between society and nature. But both the emperor and the mountain were subordinate to Heaven, for it was through the mandate of Heaven that all harmony and validation derived. These relationships were spelled out in the elaborate system of monuments with which Tai Shan was equipped. At the summit, an open circular platform was constructed for the Feng sacrifices, which consisted of burnt offerings to the heavens. Below, toward the base of the mountain, a polygonal open altar was constructed for the Shan sacrifices in honor of the earth. Between these key altars were a host of subsidiary temples dedicated to lesser nature divinities, ancestral heroes, and various miraculous saints and hermits. There were also commemorative monuments to various emperors, a school, a library, a Confucian temple, and many other sacred places scattered among the crags and groves. The whole site was simultaneously a symbol of the land, the empire, and the cosmosa great unifying topographic icon.

The symbolic order that governed Tai Shan also informed the cosmic imagery of ritual centers within the imperial capital. In Beijing, sacred enclosures featured open platform altars to the sky and earth, with subordinate temples to the sun, moon, and agriculture, each with its own complement of satellite monuments, altars, and secondary buildings. The Chinese love of order, hierarchy, and classification governed the orientation and symbolic ornament of the temples, with their gleaming white marble and carved imagery of mountains, clouds, water, and earthly or celestial dragons. These cosmic figures expressed the notion of a sacred geography in abstract form, becoming universal symbols. The magnificent altars, with their surrounding concourses and processional paths, formed the setting for imperial rites where the emperor offered sacrificial covenants to heaven and earth. The sacrifices expressed a complementary relationship to one another. The mysterious, limitless heights of heaven and the regular movements of the celestial bodies symbolized Heaven's regulative power to keep the universe in stable order and to produce the proper succession of seasons. This power was especially important to an agricultural people, and through sacrificial acts the emperors expressed the harmony of a well-ordered society within the universal schema.

Medieval Europe

The sacred geographies of the Far East and among the Indian peoples of the Americas tend to express a correspondence between humanity and nature and to be arranged in patterns of centralization. By contrast, the sacred geography of medieval Europe did not focus on a single European capital but was cast instead as a vast network that ran through many lands, leading to pilgrimage cities at the extremities of Christendom. From central and southern France, pilgrimage routes ran south over the Pyrenees, converging on a road that led across northern Spain to Santiago de Compostela, the tomb of the apostle James. Another route threaded down the Italian Peninsula to Rome. There, amid the ruins of antiquity, the pilgrim might meditate upon the early saints and martyrs and visit the old Lateranthe Mother Church and seat of the earthly Vicar of Christ. But the most perilous of routes led farther south to the port of Brindisi, and from there by ship across the Mediterranean, to the most distant, mysterious, and sacred of all goals, the city of Jerusalem.

The physical realities of this complex geography were translated in terms of a mythological hierarchy on thirteenth-century maps. These charts reveal a vision in which the horizontal surface of the earth was shown as a flat disc, with Jerusalem, the Holy City, marked clearly at the center. The outline of the Mediterranean Basin was summarily drawn, as were the features of Europe. Peripheral and unknown regions were shown to be the realms of bizarre or fabled races, sometimes held at bay behind great walled enclosures. On a vertical axis above Jerusalem, the Savior presided over the celestial sphere, while ferocious demons patrolled the infernal regions below. Such imagery was not primarily meant to illustrate an actual, material geography but rather a spiritual landscape whose central earthly icon, Jerusalem, could be interpreted on different levels. It was the Holy City of Palestine, the goal of pilgrims and crusaders; a symbol of the church; a metaphor for the Christian soul; and an analogy for the heavenly Jerusalem, the final Promised Land.

Such patterns were repeated in microcosm throughout Christendom in the art and architecture of cathedrals. In France, where these buildings reached their highest expression, sculpture, architecture, and stained glass formed a symbolic code to show the order of nature, to represent an abstract of history, and to summarize spiritual values. Within the soaring naves of the great cathedrals, at once mysterious and secure against the outside world, the assembled congregation saw the mirror of creation. Yet, like the maps, these buildings represented an essentially interior world, expressing the aims and aspirations of the innermost consciousness of the community. Such sacred places, joined in the larger network of routes to distant centers of faith, formed a sacred geography of a tradition that denied the physical world to emphasize instead theological, conceptual, and belief-oriented values that urged people to rise up, away from the earth, above animals, plants, and inanimate objects, toward a transcendent God.

Conclusion

This article summarizes the structure of four sacred geographies in societies that range from the level of tribal bands to complex civilizations. In all cases, sacred geographies have the functions of creating a sense of place and of creating a certain order in the world. Through the use of symbols, networks of meaning are imposed upon the land; such spatial orders clarify the difference between places by illustrating what is thought to be significant in the perceived world. These symbols may be natural features, such as mountains, lakes, or rivers; they may be pictographs or markings; or they may be elaborate works of art and architecture accompanied by writing. In the context of a landscape, such symbolic systems communicate the difference between sacred and profane space and answer the universal theme of establishing connections between a population and a time and place of origin.

Beyond such essential functions, sacred geographies are as varied as religions. These differences are the result of specific cultural and historical factors as well as geographical conditions. The circumstances of a wandering life in an isolated region; the need to form or unify a state organization; the pattern of an early chain of missions or military conquests; the lasting prestige and sacred quality of an ancient civic and religious centerthese and countless other factors may determine how sacred geographies are shaped. By incorporating the imagery of history and related information, sacred geographies make visible a cultural or ethnic domain and signal territorial possession.

The study of sacred geography is especially important in understanding the processes of cultural history, particularly among peoples whose traditions are not documented in writing or whose traditions may be recorded in partly deciphered hieroglyphic texts and figural imagery. In Africa, Oceania, and the Americas, where the native spiritual and intellectual heritage was largely transmitted orally, major archaeological sites must be decoded without textual sources, and early European reports of contact are colored by an Indo-European outlook. Among these peoples, the broad patterns of sacred geography provide indispensable insight on the role of religious thought and symbolism in the evolution of civilization.

See Also

Center of the World; Cosmology; Deserts; Gardens; Jerusalem; Lakes; Mountains; Oceans; Rivers; Sacred Space.

Bibliography

Bastien, Joseph W. Mountain of the Condor. Saint Paul, 1978.

Berndt, Ronald M., and E. S. Phillips, eds. The Australian Aboriginal Heritage. Sydney, 1973.

Campbell, Tony. Early Maps. New York, 1981.

Chavannes, Édouard. Le Tai Chan: Essai de monographie d'un culte chinois. Paris, 1910.

Combaz, Gisbert. Les temples impériaux de la Chine. Brussels, 1912.

Harrington, John Peabody. The Ethnogeography of the Tewa Indians. Twenty-ninth Annual Report of the Bureau of American Ethnology, 19071908. Washington, D.C., 1908.

Hiatt, L. R. "Local Organization among the Australian Aborigines." Oceania 32 (June 1962): 267286.

Hiatt, L. R. "Ownership and Use of Land among the Australian Aborigines." In Man the Hunter, edited by Richard B. Lee and Irven DeVore, pp. 99102. Chicago, 1968.

Sirén, Osvald. The Imperial Palaces of Peking. 3 vols. Paris, 1926.

Stanner, William E. H. "Aboriginal Territorial Organization: Estate, Range, Domain and Regime." Oceania 36 (September 1965): 126.

Townsend, Richard F. "Pyramid and Sacred Mountain." In Ethnoastronomy and Archaeoastronomy in the American Tropics, edited by Anthony F. Aveni and Gary Urton, pp. 3762. New York, 1982.

Turner, Victor, and Edith Turner. Image and Pilgrimage in Christian Culture. New York, 1978.

Vogt, Evon Z. Zinacantan: A Maya Community in the Highlands of Chiapas. Cambridge, Mass., 1969.

Wheatley, Paul. The Pivot of the Four Quarters: A Preliminary Enquiry into the Origins and Character of the Ancient Chinese City. Chicago, 1971.

New Sources

Gartner, William G. "Archaeoastronomy as a Sacred Geography." Wisconsin Archaeologist 77, no. 34 (1996): 128150.

MacDonald, Mary N. Experiences of Place. Cambridge, Massachusetts, 2003.

Park, Chris C. Sacred Worlds: An Introduction to Geography and Religion. London; New York, 1994.

Richer, Jean. Sacred Geography of the Ancient Greeks: Astrological Symbolism in Art, Architecture, and Landscape. Albany, 1994.

Scott, Jamie S., and Paul Simpson-Housley. Sacred Places and Profane Spaces: Essays in the Geographics of Judaism, Christianity, and Islam. New York, 1991.

Scott, Jamie S., and Paul Simpson-Housley. Mapping the Sacred: Religion, Geography and Postcolonial Literatures. Amsterdam; Atlanta, GA, 2001.

Stoddard, Robert H., and E. Alan Morinis. Sacred Places and Profane Spaces: The Geography of Pilgrimages. Baton Rouge, 1997.

Richard F. Townsend (1987)

Revised Bibliography

Geography

views updated Jun 11 2018

Geography

THE EMERGENCE OF THE DISCIPLINE OF GEOGRAPHY

GEOGRAPHY FROM 1900 ONWARD

BIBLIOGRAPHY

Geography is the study of the field of knowledge relating to the temporal and spatial dimensions of the processes that shape the Earths surface. Implicit within geography as a discipline is an emphasis on ways that diverse systems interact over time, producing particular landscapes in particular places. Geography is distinct from other disciplines (such as spatial economics and geology) in that it emphasizes the dual dimensions of time and space as causal, and because it highlights the importance of feedback between these different dimensions. A landscape can be understood as a place-specific configuration of interacting systems (at any scale), whose features influence the geographical processes operating in that particular case (this interaction produces feedback). The term landscape is often used to describe systems in which natural processes predominate, while place is used to describe systems dominated by human processes, though this distinction is not absolute. Indeed, in human geography the term landscape is used by many to describe anthropogenic systems, giving the effect that the phenomenon is distinct from the humans inhabiting it.

There are several ways of subdividing the discipline of geography. A common distinction is often made between physical geography as an environmental science and human geography as a social science. Human geography can further be divided according to the topic of study, creating subfields such as urban, political, economic, social, historical or cultural geography. More abstract dividing lines can also be drawn on the basis of methodologies, so that a distinction can be made between quantitative approaches, which have much in common with orthodox economics and cartography, and qualitative approaches, which can be similar to sociology, planning anthropology, and political science. Since the quantitative revolution of the 1960s, geographers have also distinguished themselves according to their epistemological perspectives on the nature of reality and the possibility of individuals adequately understanding and describing that reality.

THE EMERGENCE OF THE DISCIPLINE OF GEOGRAPHY

Geography as a discipline emerged in Germany in the eighteenth century in the context of attempts associated with Alexander von Humboldt at the University of Berlin to systematically harness the benefits that universities provided to society by creating applied disciplines linked to social needs. The ancient civilizations were interested in geographical topics such as cartography and cultural geography in the context of operating their military and trading empires. These proto-geographical ways of thinking were primarily concerned with describing the form of the earth, often in an attempt to better control a particular activity rather than produce pure knowledge of geographical systems. It is important to note that these applied uses of geography have created a continual stream of wealthy patrons willing to fund exploration to producing spatial knowledges, which have often served those patrons own power interests.

According to Richard Hartshornes The Nature of Geography, geography as a discipline emerged from a growing body of practical studies after 1500, but it was from 1750 onward that serious attempts were made to establish the two preconditions for a discipline, namely a distinctive subject area and a rigorous analytic methodology. The renowned philosopher Immanuel Kant (17241804) saw a close and necessary connection between understanding the nature of the world, and developing a more general (abstract) philosophical method. Hartshorne relates that Kants physical geography course became a staple of his philosophy curriculum, being repeated forty-eight times from 1756 to 1796 at the University of Köningsberg (now Kaliningrad). Kant developed the idea of Länderkunde, of studying particular distinctive regions in detail to understand their emergence. This approach was extended throughout the early nineteenth century by other early geographers, such as Von Humboldt and Carl Ritter, who viewed the human and physical elements of places as inseparable elements of a unitary whole (die Ganzheit ), of a region or landscape. Humboldt, in particular, tried to ensure that the focus of study was on classes of places (classified, perhaps by climate, geology, social form) rather than classes of objects (such as species).

Humboldts efforts to create a science of geography were ultimately responsible for a postmortem split between systematic geographers, studying physical processes and landforms, and regional geographers, studying societies in particular climatic and geological regions. This cleavage deepened in the course of the nineteenth century, as geographers sought to specialize and establish themselves in the many new and growing universities. However, this tension was never fully resolved, and as geography began to institutionalize, the newly formed learned societies reflected both halves of the discipline: systematic (physical) geography and regional (human) geography.

GEOGRAPHY FROM 1900 ONWARD

The story of geography since 1900 is one of an established discipline advancing rapidly, driven by and responding to the huge social and physical changes accompanying the rise of industrial society, alongside increasing technological opportunities for new forms of research and knowledge production. During the twentieth century, a number of currents of thinking emerged that were driven by ideological considerations. Environmental determinism, for example, provided an intellectual underpinning for the imperialistic Race for Africa in the late nineteenth century by arguing that climatic considerations meant that African civilizations were incapable of developing stable social institutions. Likewise, desire for Lebensraum (living space) of the National Socialists in Germany drew on geopolitical thinking that originated with the English geographer Halford Mackinders concept of geopolitics to rationalize these ideological desires within the school of Geopolitik. The collapse of both underpinning ideologies carried a collateral cost for their parent disciplines as a whole. It was this tendency for supposedly neutral geographical analyses to favor particular powerful groups that was the stimulus behind the radical turn in human geography that emerged in the 1970s.

In parallel with these changes, two important geographical movements emerged to reshape the discipline in the period from 1900 to 1970. The first of these was the systematization of regional (human) geography, which attempted to move beyond ideographic descriptions of societal phenomenon in particular places to more analytic expressions of places in terms of underlying processes. However, it is important to stress that national geographies remained highly distinctive at this time, and these new regional geographies reflected these national distinctions. British regional geography, particularly in the works of Henry Daysh, H. C. Darby, and Hilda Ormsby reflected national traditions of empiricism and pragmatism while emphasizing the importance of historical development. French writers, led by Paul Vidal de la Blache, produced the Annales school, named after its journal, which developed notions of environmental possibilism, in which social configurations (genres de vie ) were influenced both by physical environment and social and political decisions. German writers, notably Walter Christaller and Alfred Weber, used mathematical modeling to understand the spatial allocation of human activity such as settlements and industry. This represented a more analytic attempt to understand human activity. Although these diverse national approaches differed in the extent to which they emphasized theory over practical observation, they clearly all represented an attempt to systematize the production of geographical knowledge.

This consensus emphasizing systematization enabled the second great geographical movement of the twentieth century, the quantitative revolution. Early systematic approaches, inspired by the natural sciences, suggested that everything in nature was knowable, given sufficient data and analytic capacity. The increasing scientization of society in general, and the rise of computer power after 1947 in particular, seemed to bring this dream of total geographical knowledge within reach. The basis for quantitative geography lay in creating mathematical spatial models to account for the distribution and evolution of particular phenomenon over space. The quantitative revolution became associated in the United States with the Regional Science movement, pioneered by Walter Isard, then a professor at University of Pennsylvania, and his supporters. By infusing geography with mathematical and statistical tools, Regional Science superficially avoided the charges of selectivity and ideological determination that afflicted more ideographic geographical approaches. In part, the popularity of these quantitative approaches derived from the apparent certainty of knowledge thereby produced, which gave policymakers a robust evidence base for making decisions about land use and economic planning.

However, quantitative approaches often suffered from failing to adequately capture the independent variables, the factors responsible for causing particular phenomena. Quantitative practitioners were making assumptions and using convenient proxy variables rather than capturing complex social and cultural phenomena. The increasing urban segregation and deprivation in the 1960s and 1970s in western Europe and America provided a direct challenge to quantitative geographys capacity to explain away the emergence of ghettos, dereliction, and, frequently, rioting. Faced with the inability to measure causal variables, a number of geographers instead began to form theories to try to understand how large-scale social structures created micro- and meso-scale problems.

Arguably the most celebrated of these geographers is David Harvey, whose empiricist manifesto Explanation in Geography (1969) was supplanted by a Marxian commitment to understanding spatial process in terms of class struggle, beginning with his 1973 treatise Social Justice and the City. Harvey exemplified a radical geography in which class theory allowed hidden power relationships to be exposed and unexpected causalities to be identified. This radical approach fitted neatly with, and has since become associated with, a more general postmodern or relativistic turn in the social sciences. Harveys neo-Marxian writings reached what many regard as an apex of a theoretical research project with the publication of The Limits to Capital in 1982. This volume set out a sweeping vision for an ever-deepening form of capitalism that would eventually penetrate into every corner of the world, and it has proven to be an important foundation for radical critiques of globalization and neoliberalism.

Radical geography subsequently proceeded through a number of further turns, which have critiqued their predecessors in an attempt to better understand what drives the development of human societies. The cultural turn emphasized the importance of noneconomic interaction and transactions; the institutional turn noted the importance of durable organizations and social norms; and the relational turn highlighted the fact that people are influenced by what is close to them, and that proximity is not just physical but can be organizational, cultural, or virtual in nature. These turns were all led by movements comprising both established geographers who recanted or redacted their established positions and emerging geographers who became a new generation of intellectual leaders for geography. The cultural turn was signaled in 1989 by the parallel publications of David Harveys The Condition of Postmodernity and Edward Sojas Postmodern Geographies. Both books asked fundamental questions about the economic focus of much Marxian geography, and they laid the foundations for more reflexive understandings of economic activity. These turns have been contested by other geographers, who have expressed concerns that the intellectual efforts involved in reflexivity have detached them from geographers central task of understanding real places and spaces.

At the same time, quantitative approaches have continued to develop. Geographical information systems (GISs) are massive computer databases into which huge quantities of locational data is entered, allowing the performance of additional fieldwork capable of identifying regularities that are not immediately obvious. Although they are academic tools, much of the investment in them came from customers seeking to make sense of unknown and uncertain environments, notably the military and oil exploration companies. The continual investments these tools produce have generated a wide range of applications, and GISs are now commonplace across the spatial sciences beyond geography.

Quantitative geographers have also used novel mathematical techniques and increasing computing power to address traditional geographical problems of understanding how spatial distribution and irregularities in distribution cause geographical processes to behave in different ways in different places. These approaches are exemplified by the geographically weighted regression (GWR) technique. There has also been an increasing dialogue between physical and human geographers trying to deal with the complex environmental problems that have emerged with the increasing industrialization of society. This neatly emphasizes the ongoing indivisibility of the natural and the anthropogenic elements of geographical systems, as well as the mutual understanding necessary to understand the development of spatial systems.

SEE ALSO Cities; Cultural Landscape; Determinism, Environmental; Kant, Immanuel; Metropolis; Peasantry; Planning; Regions; Regions, Metropolitan; Segregation, Residential; Spatial Theory; Suburbs; Topography; Urbanization

BIBLIOGRAPHY

Hartshorne, Richard. 1939. The Nature of Geography: A Critical Survey of Current Thought in the Light of the Past. Lancaster, PA: Association of American Geographers.

Harvey, David. 1969. Explanation in Geography. London: Edward Arnold.

Harvey, David. 1973. Social Justice and the City. London: Edward Arnold.

Isard, Walter. 1956. Location and Space Economy. New York: John Wiley.

Scott, Alan. 2000. Economic Geography: The Great Half-Century. In The Oxford Handbook of Economic Geography, eds. Gordon Clark, Maryann Feldman, and Meric Gertler, 1844. Oxford: Oxford University Press.

Paul Benneworth

Geography

views updated Jun 11 2018

GEOGRAPHY

In the Bible

The geographic horizon in the early biblical period was the lu'aḥ ha-ammim, a table of 70 nations listed in Genesis 10. The identification of the names and the location of the countries are the subject of differences of opinion among scholars. It is clear however that included are all of Arabia, Syria, Asia Minor as far as the Caucasus, all the lands of the Tigris and Euphrates, the western part of the highlands of Iran, the regions of the middle and lower Nile including the desert extending to their west, and Greece and its islands (see The Seventy *Nations).

In the Talmud

Scattered throughout the Talmuds, the Targums, and the Midrashim are various geographic references connected with the halakhah and with expositions and homilies on the Bible and Midrash. Most of these references are associated with Ereẓ Israel: with laws about "commandments applying to Ereẓ Israel," which are to be observed only in Ereẓ Israel, with praise of the country, and with the identification of biblical place-names.

The mitzvot dependent on Ereẓ Israel have full application only within "the territories occupied by those who came back from Babylonia" (Ereẓ Israel); have partial application within the borders of those who came up from Egypt; and refer only marginally to that territory which lies within the wider borders promised to the patriarchs but outside the area of those who came up from Egypt – territory conquered by David on his own responsibility and known in the Talmud as Syria. Within the obligatory territories were exempted enclaves, such as Caesarea in the Sharon, Susita (Hippos) in the Golan, Ashkelon in the Judean coastal lowland, and within the exempted territories obligatory enclaves such as Kefar Ẓemaḥ on the southeastern shore of Lake Kinneret. The boundaries of these areas and also of the enclaves are laid down in the halakhah (Shev. 6:1; Tosef., Shev. 6:6–11; Tosef., Oho. 18:14; Sif. Deut. 51; tj, Shev. 6:1, 36b). In connection with the laws of usucapion, Ereẓ Israel was divided into three districts: Judea, Transjordan, and Galilee (bb 3:2). Concerning the laws for the removal of fruit from the house in the sabbatical year when they had stopped growing in the field, each of the three districts was subdivided into three regions: mountain, valley, and lowland. The phytogeographical features of these were: for mountains the Cillin pine, for valleys the palm, and for lowlands the sycamore (Ficus sycamora) (Tosef., Shev. 7:11; cf. Shev. 9:2). The area between Judea and Galilee was called "the country of the Cutheans" or contemptuously "the Cuthean Strip" (Matlit shel Kutim; Lam. R. 3:7). The question also arose as to whether the law applicable to levitically unclean heathen countries applied also to the country of the Cutheans. The sages decided that the law was applicable in those cities which had been surrounded by a wall since the time of Joshua and in which Megillat Esther is read on Adar 15th (Ar. 9:6; Ar. 32a; Meg. 4a; tj, Meg. 1:1, 70a).

Many identifications of geographic and ethnographic names in the Bible are in the nature of expositions. Onkelos contented himself with a few which he considered to be beyond doubt. Targum Pseudo-Jonathan and the Palestinian Targum frequently identified places solely on the basis of the similarity of names without regard to any geographic considerations. Among the identifications of the table of nations, given in the Midrashim and Targums, none includes all the nations and countries known to the sages. These identifications are frequently inconsistent and contradictory. The equation of Rome with biblical Edom which was apparently intended at first to allow for open criticism of the Roman authorities was later accepted as fact and hence the former and latter halves of the verse: "Behold, of the fat places of the earth shall be thy dwelling, and of the dew of heaven from above" (Gen. 27:39) were interpreted in the Midrash (Gen. R. 67:6) as referring respectively to Italy (Rashi, ad loc., adds "of Greece," i.e., Magna Graecia, southern Italy) and to Bet Guvrin. On the identification of Kenites, Kenizzites, and Kadmonites, who are mentioned in the covenant with Abraham (Gen. 15:19), and who were not conquered by those who came up from Egypt, there are divergent opinions: in a plausible interpretation R. Judah held that they were Arab tribes on the border of the land of the seven nations which the Israelites inherited, whereas R. Eliezer contended that they refer to Asia Minor, Thrace, and Carthage (Gen. R. 44:23, end; bb 56a). The identification of places in Ereẓ Israel, particularly in Galilee, is mostly realistic and is of aid in a scientific study of the historical topography of the country (tj, Meg. 1:1, 70a, b; tb, Meg 5b).

The sages thought that geographic and hydrologic factors exerted a great influence on man's physical and spiritual being. On Moses' instructions to the spies: "And see the land, what it is; and the people that dwelleth therein, whether they are strong or weak" (Num. 13:18), the Tanḥuma (Shelaḥ Lekha, 6) comments: "There is a country that raises strong men, and there is a country that raises weak men." A similar view is expressed in the midrashic statement: "Some springs raise strong, others weak men, some handsome, others ugly men, some modest, others dissolute men." The spring of Shittim (Num. 25:1), which was a place of licentiousness, watered Sodom (Num. R. 20:22).

From the statements of the sages one can reconstruct the geographic concept of the world current in talmudic times. The earth with its seas was seen as a circle ringed around by the ocean (Okyanos) with the center of the circle being the *even shetiyyah ("foundation stone") in the Holy of Holies, which was thought to be in the middle of the earth (tabbur ha-areẓ), not only in a geometrical sense. This was thought to be the beginning of creation. Around the center are concentric circles in order of importance: the Holy of Holies, the Temple, Jerusalem, Ereẓ Israel, and the world (Tanḥ. Kedoshim, 6); this particular idea was devised by a man who had never seen Jerusalem. The idea of the centricity of the Holy Land occurs first in the Apocrypha, influenced by the Greek concept of omphalos, which is that the center of Earth is at Delphi. The sages based the idea that the start of creation is with the even shetiyyah on biblical passages (Tosef., Yom ha-Kippurim 3:6; Yoma 54b), but not the centricity of Jerusalem, which was not of such great significance to Jews as to Christians who transferred the center to the cross of Jesus, a concept which the Church Fathers based on biblical verses (Ezek. 5:5; 38:12; Ps. 74:12). Thus the center of circular medieval maps is Jerusalem with the cross. The view that Ereẓ Israel is higher than all countries, Jerusalem than the whole of Ereẓ Israel, and the Temple Mount than all Jerusalem (Sif. Deut. 152 and 37; Sanh. 87a) is a literal homiletical interpretation of the verse: "Then shalt thou arise, and get thee up unto the place which the Lord thy God shall choose" (Deut. 17:8). The sages were however not unaware of the fact that the spring of Etam, from which water flowed to the Temple, was higher than the Temple Mount.

An estimate of the size of the "world" ranged between the extremes of 6,000 and 1,440,000 parasangs. But a still more exaggerated view held that the earth was only 1/12,960,000 part of Gehinnom (tj, Ber. 1:1, 2c; Pes. 94a). On the area of the inhabited world (οὶκουμένη) there were divergent opinions:

(1) a third is inhabited, the remaining two-thirds being sea and desert;

(2) the whole inhabited world is situated under one star;

(3) the inhabited world is located between the Wain and Scorpio, that is, about 80° from north to south (54° north of the equator and 26° south of it);

(4) it extends from east to west, a distance of one hour of the sun's course, that is 15° (Pes. 94a).

Even those sages who were aware that the earth is round did not deal with the problem of the date line. Alexander the Great during his campaigns is said to have risen upward until he saw the earth like a globe partially submerged in an enormous bowl of water, that is, the ocean (tj, Av. Zar. 3:1, 42c; Num. R. 13:14). The Zohar (Lev., s.v.ve-im zevaḥ shelamim (3:1), Soncino ed., 346) states that according to the Book of R. Hamnuna the Elder the earth is a revolving globe, that when it is day on one side, it is night on the other, that there is a place where there is no day and opposite it a place where there is no night. The comprehension of this is said to be the secret of the mystics and not of geographers. How this individual view came to be included in the Zohar is not clear.

The problem of the density of the earth occupied the aggadists. There was a widespread view that the circle of the earth is like a dish that floats on the face of the *deep, namely, the water, and that below the deep are mountains, so that the whole rests on a solid base. Another view holds that the earth rests on pillars which apparently reach down to those mountains. Views on the thickness of the earth range from a thousand cubits (about 500 m. = 547 yds.) to a 50-year journey. There was a generally accepted view that the "water of the deep" is close to the surface of the ground which accounts for the origin of springs and the moistening of the ground: to a handbreadth of rain the deep responds with two handbreadths (Ta'an. 25b). Some thought that these springs originated in the Euphrates. The four rivers that went out of the Garden of Eden were higher than all the rivers in the world, the highest of them being the Euphrates, and hence R. Judah in the name of Rav prohibited all the water in the world to anyone who took a vow not to drink from the Euphrates (Bek. 55a). Hot springs have their origin in the deep, and pass over the entrance to Gehinnom (Shab. 39a). "All the rivers run into the sea, yet the sea is not full … they return (to their source)" (Eccles. 1:7). How do they return? There are three views:

(1) through the channels of the deep;

(2) through vapors that rise from the sea and form clouds, the desalination of the seawater taking place in the deep or in the clouds;

(3) that river water disappears in the ocean because the latter has water which "absorbs water" even if brought up in a barrel on to dry land (a view which is apparently not an exposition of the passage in Ecclesiastes). The phenomenon of how such absorption takes place is not explained (Ta'an. 9b; Gen. R. 13:9; et al.).

The Jordan flows from the Dead Sea to the ocean below the earth (Bek. 55a). The idea that the ocean is higher than the land is apparently based on the homiletic interpretation of biblical verses (Jer. 5:22; Amos 9:6); the sand on the seashore prevents the flooding of the land, which happened twice, once in the generation of Enosh, when the flood reached Calabria, and once in the generation that witnessed the confusion of the tongues when the flood stretched as far as the ends of Barbaria (tj, Shek. 6:2, 50a; Gen. R. 23:7, end). In the sea there are river-like currents and waves whose height reaches 300 parasangs which is also the distance between one wave and another. Among the big waves there are small ones (bb 73a).

The sages distinguished between floral zones in Ereẓ Israel on the basis of differences in altitude and hence in temperature. But there are other universal reasons for such diversity, viz. the distinctive features of water and of soil. Koheleth-Solomon planted in his gardens and parks "trees … of all kinds of fruit" (Eccles. 2:5), which means, according to the aggadah, literally all the kinds in the world. That they might flourish he sent demons, over whom he had dominion, to irrigate each tree by bringing water from its country of origin. Another view held that arteries spread out from the center of the earth through the entire world, and Solomon, knowing how to distinguish them, planted on each artery the appropriate trees, even those from Africa and India (Eccles. R. 2:5, no. 1).

From the praise of Ereẓ Israel contained in the aggadah it is possible to put together an aggadic geography of the country before its destruction. The love of the Holy Land, the anguish at its impoverishment and at the depletion of its children, and the expectation of its future glory engendered exaggerations that are logically incomprehensible. Ereẓ Israel's situation in the center of the world and its altitude did not change even after the destruction of the Second Temple, nor did the weight of its stones, which was greater than those of the neighboring countries (Pdre 13). The aggadah is responsible for the extension of the western boundary up to the Atlantic Ocean, this being, for the aggadist, the interpretation of "the Great Sea" in the verse: "And for the western border, ye shall have the Great Sea for a border" (Num. 34:6). Extravagant conclusions were reached by Targum Pseudo-Jonathan. All the countries on the continent as well as the islands opposite Ereẓ Israel within the limits assigned to the patriarchs (from the Brook of Egypt to Taurus Amanus) up to the "primeval waters" at the furthermost extremity of the world and even the ships sailing the sea are all included in the Promised Land (ibid.). It was said that after the destruction of the Second Temple Ereẓ Israel "drew together," i.e., diminished inside. Alexander Yannai had 60 myriad "cities" in the King's Mountain and in each of them were 60 myriad people, except for three in which there were twice as many. To feed this population the country produced enormous crops of excellent quality. By the fourth century, the country had deteriorated to such an extent that it did not produce even a large number of reeds (tj., Meg. 1:1, 170a; tj., Ta'an. 4:8, 69a; Git. 57a). In the days of Simeon b. Shetaḥ rain fell at the right time, the grains of wheat were as large as kidneys, the grains of barley like olives, the lentils like golden denarii (Ta'an. 23a). Several species of trees, such as cinnamon, brought from distant lands in the time of Solomon, still grew in the Second Temple period, and Indian pepper continued to grow until the destruction of Bethar (Eccles. R. 2:8). In fulfillment of the biblical passage: "Thou shalt not lack anything in it" (Deut. 8:9), there were exiled with Israel to Babylonia through the channels of the deep 700 species of fish permissible as food and through the air 800 species of locusts permissible as food. The fish and the locusts returned with those who came back from Babylonia (Lam. R., Proem 34).

The fate of the Lost Ten Tribes has stirred the imagination of Jews from the days of the Second Temple to our times. A miraculous existence was invented for them in distant and unknown lands, the legend of the tribes being connected with those of the river *Sambatyon and the Mountains of Darkness. Thus the Ten Tribes were exiled across the Sambatyon, Σαββατείον, the Sabbath river, which rages and hurls stones on six days of the week but rests on the Sabbath, thus proving through nature the holiness of the Sabbath (Sanh. 65b; Gen R. 11:5, 73:6); Josephus describes it as a river in Syria which flows on one day and rests on six days of the week (Jos., Wars, 7:96–99); the origin of the legend being apparently to be found in rhythmically intermittent springs, such as Ein Farah in the Judean desert.

Medieval Jewish Geography

Knowledge of the spherical form of the earth, derived from observing the height of the stars in different latitudes, reached Jewish scholars in Islamic countries through Arab astronomy. The first Jew to consider the earth as a sphere was the Cordovan rabbi, Ḥasan b. Mar Ḥasan ha-Dayyan, in his book on intercalation (end of the tenth cent.). At approximately the same time in Baghdad *Sherira b. Ḥanina Gaon, followed by his son *Hai Gaon, rejected the opinion that the heavens are like a cap over a flat earth. Only fragments remain of the stories of Abraham b. Jacob who traveled in Germany and the Slavic countries in the 950s. The two letters from Joseph b. Aaron, king of the Khazars, to R. *Ḥisdai ibn Shaprut, which comprise not only historical, but also geographical material, were transmitted by Jewish merchants from Germany (about 950). The books of medieval travelers frequently contained material of geographic interest (see *Travelers).

By the 11th century the spherical form of the earth was accepted among Jewish scholars in Islamic countries, and from there the idea passed to Provence and Italy. Solomon ibn *Gabirol states in Keter Malkhut: "The terrestrial globe is divided into two, half is dry land and half water." The first work in Hebrew about the round shape of the earth and its division into climatic regions, together with a list of the countries in each region, was Sefer Ẓurat ha-Areẓ ("The Book of the Shapeof the Earth" (late 11th or beginning of the 12th century)), by *Abraham b. Ḥiyya. His system, like that of his Muslim teachers, is that of Ptolemy, the Alexandrian (c. 150 c.e.). According to Abraham b. Ḥiyya, the earth, with the seas upon it, is a globe. The western or lower half of the globe is entirely water. The eastern half is mostly dry land (except for seas such as the Mediterranean and the Red Sea), but there is no human settlement except in seven regions. North of latitude 66° there is no settlement because of the cold. In the far south (there are those who say from the equator to the south and those who say from a few degrees south of the equator) there is no populated area because of the heat, which increases as one progresses in a southerly direction. Ẓurat ha-Areẓ was published with a Latin translation by D. Schreckenfuchs and notes by Sebastian Muenster (Basle, 1546).

The discoveries at the end of the 15th and the beginning of the 16th century refuted the limitation of the earth's population to seven regions. Information regarding this refutation was conveyed to readers of Hebrew by Abraham b. Mordecai *Farissol in chapter 13 of his book Iggeret Orḥot Olam ("Epistle on the Ways of the World," 1525), but geographical ideas derived from legends or books are still to be found in homiletic and ḥasidic works, and they persisted in "scholarly" books until the 19th century. Still in 1550, Mattathias b. Solomon *Delacrut, in his short treatise Ẓel ha-Olam ("Shadow of the World"), based on a 13th-century French work, speaks of a quarter of the area of dry land which was not populated and where no human foot trod. As late as the end of the 18th century, Phinehas Elijah *Hurwitz of Vilna in Sefer ha-Berit [ha-Shalem] ("The [Complete] Book of the Covenant," 1797) maintains that most of the globe is water, either surface or underground, that the waters of the oceans are higher than the land, and that sand prevents their flooding the earth. It served as a basic text to those who wished to learn about nature but were apprehensive of the work of the new maskilim who belittled traditional literature. Geographic literature in Hebrew and the part played by Jews in systematic geographic research are slight compared with the Jewish contribution to other branches of science, such as astronomy, mathematics, and medicine.

Geography Textbooks

Abraham Farissol's Iggeret Orḥot Olam served as a Hebrew geography textbook until the 19th century. Like other 16th-century Jewish and Christian thinkers, Farissol believed in the existence of the Ten Tribes and the river Sambatyon, and devoted much space to them. Approximately 300 years later, Samson ha-Levi *Bloch, a maskil of the Galician school, published Shevilei Olam ("The Paths of the World": vol. 1, "Asia," 1822; vol. 2, "Africa," 1827), basing himself on German literature. The treatise is in the rhetorical and witty style of the times. Abraham Menaḥem Mendel *Mohr, still using only German sources, continued the work (1856) after Bloch's death. The information on Jewish communities and Jewish scholars, known to the two authors without having to do any special research, is their original contribution. In the 1780s with the establishment of schools that included secular instruction in the curriculum, special short textbooks began to appear. Reshit Limmudim ("The Beginning of Instruction," first ed. 1796; last ed. 1869), by Baruch Linda, the first such textbook in Hebrew, also has chapters on geography. A geography book, Ha-Kaddur ("The Globe," Prague, 1831), by Moses S. Neumann, was written partly in Hebrew and partly in German, though in Hebrew characters. Asher Radin's Ge'ografyah ha-Ketannah ("The Short Geography," Koenigsberg, 1860), is an abridgment of a German textbook. Two works on the principles of geography: Meẓukei Ereẓ ("The Foundation of the Earth," 1878), by Nahum *Sokolow, and Gelilot ha-Areẓ ("The Regions of the Earth," 1880), based on German literature, by Hillel Kahana, an experienced pedagogue who is one of the last of the Galician school, appeared about the same time. As was customary among writers who did not know any Western European language other than German, Kahana transcribed French and English names according to the German pronunciation. An innovation was a colored Hebrew map, and sketches and pictures with Hebrew captions. In this way he educated the Hebrew reader to map study and observation.

Writers of textbooks solved problems in Hebrew geographical terminology and paved the way for the teaching of geography in schools in Ereẓ Israel from the end of the 19th century.

[Abraham J. Brawer]

Modern Geography

In modern geography there has been development in the concentration on limited areas and specialization in particular fields of study. One of these limited areas is the city. Die Stadt Bonn, ihre Lage und raeumliche Entwicklung (1947), by Alfred *Philippson, a German geographer, is one of the most important works on urban geography. Another significant contribution was made by Norton Sidney *Ginsberg, a U.S. geographer, who at the invitation of the Japanese government studied Tokyo's urban problems and incorporated his findings in "Tokyo Memorandum" (Reports on Tokyo Metropolitan Planning, 1962). Another specialized field is economic geography. Julius *Bien, a U.S. cartographer, not only prepared atlases for a number of major cities but carried out a full-scale survey of intercontinental railways for the U.S. War Department. Saul Bernard *Cohen, who specialized in a number of geographic fields, wrote Store Location Research for the Food Industry (1961), considered a standard guide. In addition, in the sphere of political geography he wrote Geography and Politics in a World Divided (1963).

On physical geography Victor A. *Conrad wrote Fundamentals of Physical Climatology (1942) and Methods in Climatology (1944); the Israel meteorologist Dov *Ashbel published A Bio-Climatic Atlas of Israel (1950) and Climate of the Great Rift; Arava, Dead Sea, Jordan Valley (1966). Joseph Ḥefeẓ Gentilli (1912–2000), an Australian geographer, wrote Australian Climates and Resources (1947) and Geography of Climate (1958). In connection with the study of the geography of soils David *Amiran, an Israeli, edited for unesco "Land Use in Semi-Arid Mediterranean Climates" (in Arid Zones Research, vol. 26, 1964). Morton Joseph *Rubin, a U.S. meteorologist, did research in oceanography, meteorology, and in glaciology, particularly in connection with his studies on the Antarctic. Another specialized branch of modern geography is biogeography; a monumental work in this field is Studies in Medical Geography (7 vols., 1958–67), by Jacques Meyer May (1896–1976), a French-born American scientist. Nautical geography is another division which has drawn the interest of Jewish geographers, among them the Italian Carlo *Errera, who wrote the pamphlet L'italianità dell' Adriatico (1914). The modern period has also produced an increasing number of historians of geography. Gustavo Uzielli (1889–1911), an Italian, did extensive research on the explorations of Christopher Columbus, Toscanelli, and Amerigo Vespucci. His best known work is La vita e i tempi di P. Dal Pozzo Toscanelli (1894).

A number of geographers have turned their attention to the history of cartography. Roberto *Almagià, one of Italy's most distinguished geographers, edited Monumenta Italiae Geographica (1929) and Monumenta Cartographica Vaticana (4 vols., 1944–55). Erwin J. *Raisz, an American, wrote General Cartography (1938) and Principles of Cartography (1962).

bibliography:

Neubauer, Géogr; J.Z. Hirschensohn, Sheva Hokhmot (19122); A.J. Brawer, in: Yerushalayim, 10 (1914), 117–32; idem, Palaestina nach der Agada (1920); S. Klein, Zur Geographie Palaestinas in der Zeit der Mischna (1917); J. Obermeyer, Die Landschaft Babylonien im Zeitalter des Talmuds… (1929); J.M. Guttmann, Ereẓ Yisrael be-Midrash ve-Talmud (1929); M. Avi-Yonah, Atlas Karta li-Tekufat Bayit Sheni, ha-Mishnah ve-ha-Talmud (1966); F. Taeschner, in: zdmg, 77 (1923), 31–80; Zunz, Schr, 1 (1875), 146–216.

Geography

views updated May 23 2018

GEOGRAPHY

GEOGRAPHY. As the study of the earth's surface, geography is among the most concrete and accessible of all the sciences. Yet the very definition of geographical knowledge has been highly contested throughout the nineteenth and twentieth centuries. Geographers have disagreed over whether theirs is an analytic or a synthetic study, whether it deals primarily with the realm of nature or culture, and the degree to which it should be concerned with spatial relationships. Geography has also contended with a persistent reputation as simply descriptive inventory of the earth's surface, which has exacerbated its relationship with neighboring disciplines.

Institutional and Intellectual Origins

Through most of the nineteenth century geography was a broadly defined and practical field of knowledge utilized by scholars, explorers, bureaucrats, and politicians. Organizations such as the National Geographic Society and the American Geographical Society flourished in the nineteenth century as meeting grounds for men of science and government. The American Geographical Society, chartered in 1851, was devoted to the nation's growth and progress westward, especially the development of a transcontinental rail route. The organization welcomed not just geographers but also leaders in government, business, education, and science who shared their outlook. Through the society these members were exposed to the nation's exploration, surveying, and mapping efforts, primarily in the American West. Similarly, the National Geographic Society was founded in 1888 as a forum of exchange of information for the community of scientists and bureaucrats in Washington, D. C., involved in geological work. The society continued to facilitate geologically oriented research until the Spanish-American War, when it began a vigorous defense of the nation's mission abroad. In both these organizations, geographical knowledge served the state both concretely, through the supply of scientific expertise, and abstractly, in striking a nationalist posture.

Intellectually, American geography reflected a heavy European influence in the nineteenth century. Among the most influential and popular contemporary geographers were transplanted Europeans such as Karl Ritter and Alexander von Humboldt. Both elevated geography from the realm of description to that of science by considering the landscape as a unified entity to be studied as a whole, a process for which geography was uniquely suited in its stress on synthesis. Louis Agassiz, appointed at Harvard in 1848, was trained in the natural sciences and noted for his development of theories of glaciation and landforms. Arnold Guyot, appointed at the College of New Jersey (later Princeton University) in 1854, began to introduce a concept of geography not as a description of the earth's elements but rather as an observed interrelationship between land, oceans, atmosphere, and human life, all of which interacted harmoniously in a grand design. Though geography would gradually shed this teleological cast, Guyot had pushed geography from description to inter-pretation. George Perkins Marsh also explored this relationship in his Man and Nature (1864), though with a thoroughly theological bent. Into this basic framework of the relatively static view of the human and natural world, the work of Charles Darwin introduced the idea of evolution. As a result, geographers began to pay attention to the evolution of landforms over time, which eventually bolstered the study of physical geography.

By the late nineteenth century geography was no longer simply a tool of exploration, data gathering, and mapping. With the era of exploration waning, and with the coincident rise of American universities, geographers began to turn their attention toward reconceptualizing geography as an analytic, scientific body of knowledge. This was a difficult change for geographers, both intellectually and institutionally. Many worried that their field's reputation—as a broad field open to amateur armchair explorers as well as scientific experts—would taint its prospects in the newly professionalized university.

The unquestioned intellectual father of geography at this critical moment of late-century maturation was actually trained not in geography but geology, because doctoral programs in the former had yet to be developed. William Morris Davis was trained at Harvard as a geologist by Nathaniel Southgate Shaler and appointed professor of physical geography there in 1885. For Davis, the claims geographers made for their study as the "mother of all sciences" had to be halted if progress were to be made, for other scientists regarded this claim as the key indicator of geography's incoherence. Thus began a long tension within geography: What makes the field unique and worthy of its independence? How does a study that is essentially synthetic defend itself from the reach of neighboring sciences as diverse as geology, anthropology, and botany?

Together, Shaler and Davis initiated the first course of training in physical geography—the study of the surface features of the earth—and mentored the first generation of trained geographers in the United States. During the 1880s and 1890s Davis advanced an idea that applied Darwinian principles of evolution to the study of the physical landscape. The result was the science of geo-morphology, in which Davis argued that different elements of the environment worked to produce change on the landscape through dynamics such as soil erosion. This concept helped legitimate geography at the university level and in the process gave geographers a tremendous source of pride. At the same time, however, geomorphology reinforced geography's identity as a subfield of geology, thereby hampering its intellectual independence.

In the late 1870s modern geography began to appear as a field of study in American universities, usually found within departments of geology or "geology and geography." Only in 1898 was an independent department of geography established at the University of California. Davis was convinced that geography's weak reputation was in part attributable to organizations such as the American Geographical Society and the National Geographic Society—especially the latter, which became an increasingly popularized and middlebrow organization after the turn of the century. These groups were irritating to Davis because they reinforced in the mind of the academic and lay communities alike the sense that geography was the pastime of leisured travelers and curious amateurs. He actively dissociated himself from these organizations at the turn of the century, and at one point even attempted to take control of the National Geographic Society in order to return it to its serious, scientific roots. Thus Davis was enthusiastic about a new organization designed exclusively for professional geographers. The Association of American Geographers was founded in 1904, toward the end of the trend toward disciplinary organizations. While geologists were initially welcomed in order to solidify the new organization's membership base, within a few years their applications were deferred in the hope that disciplinary purity might be achieved.

The Advent of Human Geography

Davis was successful in training a number of young geographers at the turn of the century who began to return to the relationship between humans and their physical environment. More specifically, this generation found itself increasingly compelled to study the human response to the physical environment. This turn toward the "causal relationship" was in part a result of the imperative to strengthen geography's position among the disciplines. This new focus had the added benefit of distinguishing geography from geology. Physiography, which linked elements of the environment with one another, and ontography, which linked the environment with its human inhabitants, were the two main areas of disciplinary focus for geography just after the turn of the century. Most early geographers conceived of their discipline as having unique power to bridge the natural and human sciences. From the mid-1890s to World War I the prospect of uniting nature and culture through geography seemed both feasible and imminent at some of the most important centers of academic geography, including Pennsylvania, Chicago, Yale, and Harvard. But it was precisely this claim to breadth that neighboring sciences began to challenge, for in the new era of university science, disciplines were legitimated not by claims of breadth and inclusiveness but rather by narrowing their focus and delimiting their boundaries.

Because of their interest in the causal relationship, theories that united the realm of humans and their environment held special appeal for geographers. For instance, natural selection, though widely misinterpreted, was used to describe the relationship between the physical and the human environments as one of inorganic control and organic response. Evolutionary concepts became central to geography's effort to explain nature's influence upon human behavior, and geography focused increasingly on the question of why certain races, societies, or groups flourished while others languished. To be sure, geographers neglected the idea of random variation and exaggerated and accelerated the process of "struggle" in order to incorporate humans into the ecological world. Yet without this causal connection—the influence of environment on human behavior—the areas of study under geography could easily be divided up among other disciplines.

Even more important than Darwin's ideas were those of Jean-Baptiste de Monet de Lamarck, who suggested that characteristics acquired through the course of a lifetime could be passed biologically to off spring. Lamarck's ideas were well suited to the needs of the new social sciences at the turn of the century because they united the study of nature and humans by linking biology with environment. Though the rediscovery of Mendel's laws concerning genetic heredity in 1900 eroded the credibility of Lamarckian thought, geographers continued to invoke this model when describing the core of their study as the relationship between humans and their natural environment. In other words, Lamarck created for geographers a process to study, and this appeal was too strong to be easily dismissed. Furthermore, Lamarckian constructions meant that geographers were now studying the progress of civilization, which vastly expanded their field of inquiry. By focusing on one's adaptation to the physical environment, the random chance of Darwinian evolution could be replaced with the strength of an individual, a culture, a race, or a nation. These assumptions were not always conceived in deterministic ways. While some geographers invoked them as evidence of an intellectual and social hierarchy in order to justify American expansionism or European imperialism, others used them to open up possibilities for social change. This indeterminacy implicit in Lamarckism allowed it to shape geography long after it had been discredited in other behavioral sciences. In fact it was the range of interpretations possible in Lamarckian expositions that made it so attractive to geographers.

Geography and the State

One of the striking characteristics of geographical thought at the turn of the twentieth century was its implicit support of American expansionism, as demonstrated in the sharp turn that the fledgling National Geographic Society made toward an aggressive defense of America's position abroad during the Spanish-American War. Two Europeans, Halford Mackinder and Friedrich Ratzel, also exercised considerable influence over American geographical thought. Ratzel, trained as a zoologist, argued that a relationship existed between human history and physical geography, in some ways similar to Davis's idea of ontography. But while Davis was relatively tentative in his formulations, Ratzel painted in broad strokes by applying the idea of Darwinian struggle to human society in order to frame the state as an organism that was forced to expand in order to survive. Known by many as the father of geopolitical thought, Ratzel fit well with the contemporary expansionist posture of Josiah Strong, Alfred Thayer Mahan, and Theodore Roosevelt, each of whom was encouraging American expansion into world affairs. Much like the work of Frederick Jackson Turner, Ratzel's ideas allowed geographers to link nature and culture. Ratzel's well-regarded The Sea as a Source of the Greatness of a People (1900) argued that sea power was central to national survival in the twentieth century.

Similarly, Halford Mackinder emphasized environmental influence as a key to the disciplinary identity of the new profession of geographers. His "Geographic Pivot of History" (1904) gave him an extraordinarily solid reputation in the United States; in it he laid out the geopolitical dimension of international politics. For Mackinder, the age of exploration had given way to a new era where the manipulation of information would be critical. In Mackinder's mind the human experience of geography and space had changed in fundamental ways in the late nineteenth century. As Stephen Kern has noted, the rise of geopolitics owed much to the cultural and technological changes taking place around the turn of the twentieth century, including the arrival of standardized time, the advent of flight, the expansion of the railroads, and advances in communication and radio, all of which transformed the everyday experience of space and time. Ratzel and Mackinder used geopolitical ideas in order to come to terms with this changed sense of distance resulting from these innovations. Both emphasized the relationship between geographical influence and human response.

Among the first generation of university-trained geographers who inherited these ideas of Ratzel, Mackinder, and Davis were Ellen Semple, Ellsworth Huntington, and Isaiah Bowman. Semple, a student of Ratzel's, was especially taken with environmentalist models as a way to explain American history. In works such as American Historyand Its Geographic Conditions (1903), Semple argued that living organisms evolve from simple to more complex forms through adaptation to physical environment. The larger the state, race, or people, the more certain its chance of survival relative to others competing for the same resources. Similarly, Huntington posited that the primary influence over human history was climate, and even suggested that these effects could be biologically passed on through generations. Books such as his Civilization and Climate (1915) were tremendously popular with the general public in the early twentieth century, though roundly criticized within geography and other social sciences.

World War I had a substantial impact on American academic geography. Most obviously, the war demonstrated the flexible nature of geographical borders in Europe and the ephemeral nature of colonial associations worldwide. The faith in European civilization was now tempered by its unparalleled capacity for destruction. In the United States, the war demonstrated the utility of geographic knowledge to the public and also advanced the careers of professional geographers called to work for the government. The geographer who benefited most from the war was Isaiah Bowman, then director of the American Geographical Society. One of Bowman's goals had been to make the society more relevant to social and political problems, and by placing its resources at the disposal of the federal government, the society's vast reserve of maps became pivotal to the construction of postwar Europe. The war also led many geographers, especially Bowman, to admit the limits of the environment over human behavior and to stress human influence over the environment. After World War I, geographers devoted tremendous energy to searching for a new relationship to unite the disparate areas under their field, prove its worth in the university, and conform to modern social scientific wisdom, which had deemed environmentalism a false and damaging approach to the study of human affairs.

Geography since Midcentury

One response to the rejection of environmentalist frameworks as the basis for research was to narrow geography's field of inquiry. The clearest indication of this was Richard Hartshorne's The Nature of Geography (1939), a massive statement of the field's direction written on the eve of World War II. For Hartshorne, what had historically made geography unique was its attention to systematic description of areal variation, not speculation about change over time or causal relationships between humans and their environment. The hope among earlier generations to discover laws of human behavior was dismissed by Hartshorne in favor of a focus on concrete, discrete studies.

Carl Sauer, one of the century's most influential geographers, rejected Hartshorne's treatise—and the approach of the interwar geographers generally—and characterized this period as "the great retreat" when geographers studiously avoided causal relationships between humans and their environment. Sauer thought this unacceptable: geography now conceded physiography to geology and shied away from the social sciences for fear of repeating past sins of environmental determinism. One of Sauer's alternatives was to emphasize the influence of humans over their environment rather than the reverse. In his wake, many students adopted Sauer's new approach in delving into the particularities of place and paying close attention to the development of landscape. Yet despite Sauer's attempt to discredit environmentalism, many geographers continued to grant the physical environment influence over human behavior during the interwar period, an indication of the fractured nature of the discipline at midcentury. In 1947, Harvard made the decision to dissolve its department of geography, the original locus of academic geography in the United States. In subsequent years, Stanford, Yale, Michigan, and innumerable smaller institutions closed their geography departments. Yet the overall number of geography programs rose sharply in the postwar years, a reflection of the general growth of higher education.

Geographers themselves found renewed energy in the 1950s and 1960s by turning toward quantitative analyses as the basis for a redefinition of geography. The "quantitative revolution" did not constitute a change in goals so much as in method: geographers were still searching for locational patterns, but they began to adopt mathematical models, which in some cases led a return to a more abstract, general orientation and away from the idiographic focus on discrete regions. This school of geography drew heavily from economics. But by the late 1960s the quantitative revolution left many concerned that geography was bereft of any purposive, reformist content. Some argued that the quantitative model of geography essentially operated conservatively, in defense of the status quo, and contained little critical potential. A reaction to this—in part inspired by Thomas Kuhn's Structure of Scientific Revolutions (1962)—brought a resurgence of political concerns to the study of geography, but this time with a radical rather than a conservative thrust.

Postmodern, or radical, geography involves first and foremost a critique of the traditional relationship between notions of space and time. For geographers such as Neil Smith and Edward Soja, for instance, Western culture has been preoccupied since the nineteenth century with a historicist focus, and this has come at the expense of an explicitly spatial orientation. They argue that this temporal bent has obscured our awareness of just how deeply the dynamics of power—especially those created by capitalism—are inscribed in spatial relations. For both Smith and Soja, to remedy this requires a critique of historicism and a turn toward spatial concerns. This goal of a more activist, self-critical form of the discipline has continued from the late 1970s forward to the beginning of the twenty-first century, and has brought special attention to the relationship between power and capitalism in the study of urban space. It has infused geography with both theoretical concerns and concrete purpose. In recent years considerable research has also been undertaken in the field of feminist geography, which explores the way gender relations are reinforced by spatial arrangements of societies. The wide influence of these new, conceptually rich areas of research extends well beyond the disciplinary bounds of geography, which suggests the trend toward a more ambitious and socially relevant scope for the subject.

BIBLIOGRAPHY

Blouet, Brian, ed. The Origins of Academic Geography in the United States. Hamden, Conn. : Archon, 1981.

Driver, Felix. "Geography's Empire: Histories of Geographical Knowledge." Environment and Planning D: Society and Space 10 (1992): 23–40.

Godlewska, Anne, and Neil Smith, eds. Geography and Empire. Oxford: Blackwell, 1994.

Kern, Stephen. The Culture of Time and Space, 1880–1918. Cambridge, Mass. : Harvard University Press, 1983.

Kirby, Andrew. "The Great Desert of the American Mind: Concepts of Space and Time and Their Historiographic Implications." In The Estate of Social Knowledge. Edited by Jo Anne Brown and David K. van Keuren. Baltimore, Md. : Johns Hopkins University Press, 1991.

Livingstone, David N. The Geographical Tradition: Episodes in the History of a Contested Enterprise. Oxford: Blackwell, 1992.

Martin, Geoffrey J., and Preston E. James. All Possible Worlds: A History of Geographical Ideas. New York: Wiley and Sons, 1993.

Rose, Gillian. Feminism and Geography: The Limits of Geographical Knowledge. Cambridge, Mass. : Polity Press, 1993.

Schulten, Susan. The Geographical Imagination in America, 1880– 1950. Chicago: University of Chicago Press, 2001.

Smith, Neil. Uneven Development: Nature, Capital, and the Production of Space. Oxford: Blackwell, 1984.

Soja, Edward W. Postmodern Geographies: The Reassertion of Space in Critical Social Theory. London and New York: Verso, 1989.

Stoddart, D. R. On Geography and its History. Oxford: Blackwell, 1986.

SusanSchulten

See alsoEvolutionism ; Geology ; Maps and Mapmaking .

Geography

views updated Jun 11 2018

GEOGRAPHY

The topography and geography of the Middle East are closely related to the geology and climate of the region.

A zone of mountainous terrain in the north in combination with higher latitudes, lower temperatures, and increased precipitation gives a distinctive character to Turkey, Iran, and parts of the Levantine (eastern Mediterranean) coast. To the south, in North Africa and the Arabian Peninsula, tilted fault-block mountains and volcanoes provide intermittent physical relief to an area largely consisting of plateaus and plains. Unremitting aridity and high temperatures typify the desert that dominates this southern part of the region.

The geology of the Middle East is determined by the movement of continental plates in a north-westerly direction. This movement, in turn, deforms masses of sedimentary strata deposited in Paleozoic times in the ancient Tethyan Sea, which once separated Eurasia and Africa. The African plate is the largest and consists of ancient igneous materials overlain, in part, by a relatively thin layer of more recent sedimentary rocks. With the exception of the folded strata that make up the Atlas and Anti-Atlas Mountains in the west of the Maghrib (North Africa), the Ahaggar and Tibesti Mountains of Algeria and Libya, as well as the highlands of the Ethiopian plateau, are volcanic in nature. The Arabian plate to the east consists of tilted Mesozoic sedimentary strata dipping beneath the Persian/Arabian Gulf: These strata overlie pre-Cambrian igneous basement rock exposed by erosion in the Asir Mountains along the western shores of the peninsula. The Red Sea, which the uptilted edge of the Asir overlooks, is a continuation of the East African rift valley system and is formed by the moving apart of the African and Arabian blocks. This rift system continues north through the Gulf of Aqaba and forms the valley of the Dead Sea and the Jordan River. It eventually disappears in the down-folded strata of the Biqa (Bekaa) Valley of Lebanon.

The heavily folded Zagros Mountains bordering the Gulf on its eastern side result from the collision and subduction of the Arabian plate under the Iranian plate. The Persian/Arabian Gulf, which is an inlet of the Indian Ocean along the axis of the subduction zone, has accumulated huge quantities of sediments from the Tigris and Euphrates rivers, the Karun River, and numerous intermittent streams draining the lands on either side. Within these Tertiary sedimentary strata are found the largest petroleum fields in the world, with deposits in Saudi Arabia, Kuwait, Iran, Iraq, the United Arab Emirates, Qatar, and Oman, in descending order of importance. To the northwest, the Turkish
plate is sliding westward along a transform fault and colliding with the Aegean plate. These areas of movement create major fault zones subject to severe earthquakes. In Turkey, the Erzincan earthquake of 1992 was typical. Faulting and recent volcanism terminate the northward extension of the rich petroleum fields of the Gulf beyond a few poor deposits near Batman, Turkey.

The northern part of the Middle East is a mountainous extension of the Alpine orogeny. The Pontic Mountains paralleling Turkey's Black Sea coast merge with the eastern highlands notable for volcanic Mount Ararat of biblical fame. The Taurus Mountains along Turkey's south shore extend eastward as the Anti-Taurus Mountains, joining the Zagros Mountains running southeast between Iran and Iraq. Another extension forms the Elburz Mountains bordering the Caspian Sea in Iran. Mount Damavand (18,934 ft [5775 m]), the highest peak in the Middle East and North Africa, is part of this range. Still farther east, the Kopet Mountains merge with those in Afghanistan and the Hindu Kush.

Great rivers have played their part in the history and development of the Middle East. The WhiteNile, which rises in equatorial Africa, is joined at Khartoum in Sudan by the Blue Nile, flowing from the highlands of Ethiopia. No precipitation sufficient for human survival occurs from that juncture north to the Mediterranean Sea, and all life in Egypt depends on the use of the combined waters of the two Niles. The Euphrates River and its companion the Tigris both rise in Turkey and join in southern Iraq to form the Shatt al-Arab, which empties into the Persian/Arabian Gulf. The area between the two streams, ancient Mesopotamia, was the site of the earliest civilization, Sumer (3500 b.c.e.), and other ancient civilizations based on irrigation farming.


The Mediterranean Sea also has influenced many of the cultural and geographical characteristics of the Middle East. It has served as a major link between Europe, Africa, and southwest Asia since ancient times. The Turkish Straits, composed from north to south of the Bosporus, the Sea of Marmara, and the Dardanelles, are an important waterway joining the Black Sea to the Aegean Sea and the Mediterranean. Bronze Age ships plied these straits and sailed along the coast of Turkey as well as among the Aegean Islands. Early Phoenician traders established sea routes leading to the Straits of Gibraltar and beyond. Ancient Greek ships traveled through the Bosporus to bring grain from the shores of the Black Sea, and Roman triremes linked Italy and Africa. During the Middle Ages, some Arab navigational skills were conveyed to Europeans as Islam was spread. In the nineteenth century, the Mediterranean route was enhanced when the French and Egyptians completed the Suez Canal, joining the Mediterranean Sea to the Red Sea and thus reducing the trip from Europe to India (originally by way of the Cape of Good Hope) by thousands of miles.

The Middle East is composed of four environments, expressed by climate, vegetation, and traditional lifestyle. Well-watered humid and subhumid lands border the Black Sea in Turkey and extend along the Caspian shore of Iran. In these well-populated places, maize (corn), tea, hazelnuts, and rice are important crops.

Mountainous terrain, with remnant forests of pine, cedar, and juniper, rims the Anatolian plateau of Turkey and extends southward into coastal Syria and Lebanon. Similar but drier environments are found in the Zagros and Elburz Mountains of Iran. These areas once supported dense growths of mature trees, but with the exception of more remote places in the Taurus Mountains, the logger's ax and the charcoal burners' ovens have depleted the forests while nomads' goats have prevented regrowth through overgrazing. The result is either disturbed and impoverished woodlands (French, maquis ) or barren and rocky ground supporting low herbaceous shrubs (French, garrigue ).

The interior plateau of Turkey, the foothills of the Anti-Taurus and Zagros Mountains, and the northern portions of Jordan and Israel are semiarid, with grazing or grain production depending on the amount of each year's precipitation. The variance of rainfall on the drier margins of these areas makes permanent rain-fed agriculture difficult. As a result, ancient peoples developed pastoral nomadism as a lifestyle that met this challenge. Herds and flocks were moved seasonally to new pastures to avoid over-grazing of sparse vegetation as well as to seek out water sources. Once an important means of livelihood, nomadism has largely been abandoned.

The semiarid steppes merge gradually into true deserts, which dominate southern Israel, Jordan, and Iraq as well as the Arabian Peninsula and North Africa west to the Atlas Mountains. Saharan conditions extend to the Mediterranean shore of North Africa from Gaza to Sfax in Tunisia, the only exception being a small outlier of Mediterranean climate on the Jabal al-Akhdar (Green Mountain) of Libya. Under desert conditions, agriculture is possible only in scattered oases and along the banks of rivers like the Euphrates in Iraq and the Nile in Egypt.

The narrow rim of Mediterranean climate, which extends north from Gaza through Israel, along the coasts of Lebanon, Syria, and Turkey, is the fourth environment. This same climate is also found from Tunis west to the Atlantic shores of Morocco. The Mediterranean environment is typified by winter rains and hot dry summers, which allow for the production of irrigated vegetables and citrus fruits, as well as various winter grains.

See also Aegean Sea; Aqaba, Gulf of; Arabian Peninsula; Ararat, Mount; Atlas Mountains; Biqa Valley; Black Sea; Climate; Dead Sea; Hindu Kush Mountains; Levantine; Maghrib; Mediterranean Sea; Nile River; Persian (Arabian) Gulf; Petroleum, Oil, and Natural Gas; Shatt al-Arab; Straits, Turkish; Suez Canal; Taurus Mountains; Tigris and Euphrates Rivers.


Bibliography

Beaumont, Peter; Blake, Gerald H.; and Wagstaff, J. Malcolm. The Middle East: A Geographical Study, 2d edition. New York: Halsted Press, 1988.

Blake, Gerald; Dewdney, John; and Mitchell, Jonathan. The Cambridge Atlas of the Middle East and North Africa. Cambridge, U.K., and New York: Cambridge University Press, 1987.

Held, Colbert C. Middle East Patterns: Places, Peoples, and Politics, 3d edition. Boulder, CO: Westview Press, 2000.

Longrigg, Stephen H. The Middle East: A Social Geography, 2d edition. Chicago: Aldine, 1970.

John F. Kolars

Geography

views updated May 17 2018

Geography


Geography is the study of the physical and geopolitical aspects of the surface of Earth. Physical geography describes the different surface and climatic conditions around the world. Political geography is concerned with the division of the world into various levels of government, human activity, and production. Geography is not confined to merely describing Earth as it is now, but also understanding how it has evolved and how it may change in the future.

The problems that faced humankind at the dawn of history stimulated both geography and mathematics. In fact, much of early mathematics was concerned with making measurements of the land; so much so that a whole branch of mathematics became known as Earth-measurement, which in Greek is geometry. Geometry as a mathematical study is less concerned with its practical roots, but for geographers, geometry and trigonometry are invaluable tools.

Outside of the activities associated with mapmaking, geography was for many years mainly descriptive. The information collected about the physical and social characteristics of the world was reported in a narrative form with little attempt to analyze the data that had been collected. In the late 1940s and early 1950s, a revolution took place in geography when the acceptance of any theory within the science became subject to mathematical analysis. As with other sciences, geography began to use mathematics as the language to describe relationships in the discipline.

The Geographic Matrix

An observation made by a geographer has two main attributesa location and a physical attribute associated with that location. Each place may have more than one characteristic and each characteristic may be found at more than one location. These data can be recorded in a matrix with rows representing characteristics and columns the places where the

observations have been taken. The organization of data into a matrix greatly aids geographers with the mathematical analysis of the information they gather.

Before geographers collect data, they must select the locations within the region where they will measure the characteristics of interest. This requires an understanding of the sampling techniques found in mathematical statistics. From statistical sampling theories the geographer calculates how many locations will be required, which will consequently reveal the number of columns in the matrix. In establishing the appropriate number of locations, it is necessary to ensure that there is sufficient data so that the samples are representative of the whole region.

Three main types of sampling systems are used in selecting locations. The first type is a totally random sample, in which each location in the study is selected at random from all possible points in the region. The second type of sample is a systematic sample, in which an initial point is chosen at random and all other points are determined by fixed intervals from the randomly chosen point. The third type of sample is a stratified sample, in which the region is subdivided into subregions. Within the subregions, points are chosen by either using a totally random sample, or a stratified sample, or by dividing into further subdivisions. This process can continue until the degree of accuracy required matches the number of sampling points. For example, in studying a country, a geographer may first break the country into regions. Then the regions may subdivide by using political divisions such as a state, and this may go further by using counties, and at this level there may be a random selection of sampling points.

Ultimately, the selection of sampling locations should permit a rapid, accurate, and economical amount of calculation in order to analyze the data. The selection process should also be such that final analysis is comparable to data collected in other regions so that regional comparisons may be made. In addition, consideration needs to be given to national and international standards, and to enabling comparisons with data collected over time.

Analysis of the Geographical Matrix

When the collected data have been placed in a geographic matrix, an analysis of a region can proceed in many ways. One common method of analysis is an examination of how a characteristic is distributed over a region by examining the row of the matrix for that characteristic. For example, attention may be focused on the way in which the rural population is distributed over an area such as the Great Plains of the United States.

Secondly, a geographer may try to get an understanding of the complexity of a location by identifying its characteristics. In other words, the column for that location may be analyzed. For instance, interest may be in the rainfall, soil type, or most successful crop production at a location in order to make recommendations for other places with similar characteristics. If the location is an urban area, a geographer might try to connect transportation access data, raw material availability, and expert labor supply, in order to explain why a particular industry is successful at that location.

A third way to make comparisons is between rows. This enables an understanding of which characteristics are found together or separately, or to what degree they might mix. For example, looking at common characteristics for two economically successful locations can show why they contribute to the locations' success.

A fourth option for a method of analysis is to make a comparison of columns. This allows the geographer to describe which locations are similar and which are very different. For instance, by analyzing locations where the weather data are the same, a geographer can classify climates that are similar. All of these analyses require the statistical techniques of correlation and regression (defining and characterizing relationships among data).

Optimization Problems Solved by Geographers

Although mathematics has become an essential tool of modern geography, it was also present in the geography of the nineteenth century. In 1826, Von Thünen collected data on land values in agricultural communities; he also collected data on how farmers used land. His data were centered on a town that was the main market for a region.

Von Thünen found that for each particular type of crop the costs of getting the produce to market was a product of the distance from town, r, the volume of the crop produced in a unit area of land, v, and the cost of transportation per unit of distance, c. If the crop sells at a price of p and the fixed costs of producing the crop are a, then the net profit is expressed as R = (p a )v rcv.

Von Thünen constructed graphs of profit R plotted against the distance from town, r for various crops. The figure above shows the graphs for three crops. Crop 1 produces the highest profit as long as it is inside a distance of r 1 of the market. Between a distance of r 1 and r 2 crop 2 is the most profitable, and between r 2 and r 3 crop 3 is the most profitable. At r 3 all three crops become unprofitable. Von Thünen suggested that the land around a market be used to reflect these rings, and that there should be no cultivation of these crops beyond r 3 at all, as there was no profit in farming at this distance (see below).

The modern equivalent of this geographical distribution model is the understanding of why the location of a shopping center or a factory affects each one's success or failure. From the geographical matrix the locations of various resources that are required by a manufacturing plant can be established. Given the locations of different raw materials, labor resources, and transportation of raw materials to the factory, the location for the optimum manufacturing plant can be calculated and compared to an existing plant.

A geographer's data can also be used to support or refute the location of a manufacturing plant at a particular location. However, this is only part of the solution, for once the goods have been manufactured they have to be distributed to market centers, and this has an associated cost that can affect the decision concerning a manufacturing plant's location. By weighting distances with regard to cost, an optimum location can be found by finding the equivalent of the center of gravity of the system.

In the location of a factory, one of the problems that has to be tackled is the distribution of the product to market. Here another branch of mathematics aids the analysis. Graphs and trees deal with the analysis of networks, and can be employed in finding solutions to this part of the problem. One of the classic problems of networks, the travelling salesman problem, is concerned with the most efficient route for a travelling salesman to take in order to cover all the customers. This is also the route that the supply trucks will be interested in following. The full understanding of this problem is still the object of mathematical research.

Calculus in Geography

A major problem of geography is the modeling of population change. Change in populations implies that geographers are interested in data that are time dependent. Therefore any data collection has to be repeated at various intervals, the most familiar way being the United States census that is required every 10 years by law.

The census gives data over long periods of time, but annual sampling is necessary in order to monitor more detailed changes. The data can then be matched to a mathematical model. The most common model for population growth results in the construction of a differential equation in which the change in population, with respect to time, varies directly with time and can be solved through calculus.

Calculus also helps model the way in which the profile of a hill develops. Another application of calculus gives a mathematical model of the freezing of water in a lake. If air above a lake maintains temperatures below the freezing point of water for a prolonged period of time, the thickness of the ice will continue to increase. The rate of advance of the ice depends on the rate at which heat can be carried away from the surface by convection currents in the water below the ice surface. The model leads to a differential equation .

Fractals and River Watersheds

The way in which rivers begin their life as a collection of small springs or gullies that collect rain and spill into brooks or streams has been better understood in recent times by the analysis offered by mathematics through fractals . In river systems, fractal scaling can be seen in the organization of the river network at various levels of observation; that is, they conform to the fractals first described by Benoit B. Mandelbrot. Research around 1990 described the scaling properties of the geometry of several river systems and a calculation was made of their fractal dimension.

Probability and the Layout of Villages

Probability leads to an understanding of the way villages develop over a long period of time, given that there has been no deliberate planning. The model requires the description of two objects, a closed cell with an entrance, and an open cell (see figure below).

These cells are joined together to form a doublet so that the entrance always faces onto an open cellcorresponding to a house opening out onto the public space. In the modeling process, the doublets are allowed to accumulate with the condition that each new doublet that joins the village does so with its open cell having at least one edge common with another open cell. Which open cell a new doublet joins is chosen at random. This modeling process has been successful in describing a number of old villages in which town planning did not influence the layout.

see also Cartographer; Fractals; Global Positioning System; Mandelbrot, Benoit B.; Maps and Mapmaking; Probability, Theoretical.

Phillip Nissen

Bibliography

Haining, Robert. Spatial Data Analysis in the Social and Environmental Sciences. Cambridge, U.K.: Cambridge University Press, 1993.

Hillier, B., and J. Hanson. The Social Logic of Space. Cambridge, U.K.: Cambridge University Press, 1984

Rodriguez-Iturbe, Ignacio, and Andrea Rinaldo. Fractal River Basins: Chance and Self-Organization. Cambridge, U.K.: Cambridge University Press, 1997.

Wilson, A.G., and M. J. Kirby. Mathematics for Geographers and Planners. Oxford, U.K.: Clarendon Press, 1975.

Geography

views updated May 14 2018

Geography

Sources

Theoretical Aspects. While mathematical astronomy was mostly a matter for Greeks writing in Greek, the Romans were more receptive to geography. As the Romans extended their domain, they were curious about foreign lands, peoples, and cultures. In the second century b.c.e. Polybius, who visited and entertained the Romans with learned discourse, stressed geography as an aid to understanding politics and history. Even more to the point, the ever-expanding empire called for knowledge of the world beyond the borders of Roman control. Although the Romans’ contributions to geography were considerable when it came to places, peoples, and distances, nonetheless, the Greeks continued to provide the theoretical aspects. Nowhere can this case be more clearly seen than in the contrast between Eratosthenes (whose brilliant method resulted is a fairly accurate calculation of the circumference of the earth) and Marcus Agrippa. Eratosthenes wanted to calculate the circumference employing geometry and mathematics. When Agrippa wanted to know about Rome’s world, he simply added up the distances between the milestones along the roads to map the region by determining distances.

Circumference of the Earth. A librarian and geographer at Alexandria devised an ingenious calculation for the circumference of the earth that challenged and astounded the Romans. Eratosthenes (circa 276-circa 195 b.c.e.) wrote a three-book Geography that was brilliant in conception. Julius Caesar read it for information on the German regions and Strabo, a later geographer in the Roman world, based much of his work on Eratosthenes, although he disliked his mathematical theory. Through the application of plane geometry to a well-conceived experiment, Eratosthenes computed the circumference of the earth to the equivalent of 250,000 stades, or 29,000 English miles, and, by another interpretation, to 25,000—the actual circumference is 24,902.45 miles. Pliny the Elder wrote that this figure is “an audacious venture, but achieved by such subtle reasoning that one is ashamed to be skeptical” (Natural History. 2.112.247). Acting on his hypothesis, Eratosthenes calculated the parallel of Rhodes (36° N) to be under 200,000 stades in circumference. By simple arithmetic each degree of latitude is around 541 stades (195,000 ÷ 360 = 541.66 stades). This calculation permitted him to map the globe with fairly accurate coordinates for the Mediterranean world that they knew, including much of north Africa, Europe (including Britain and Ireland, but excepting northern and Scandinavian regions), and Asia as far as India and the Himalayan mountains.

Scipionic Circle. Unlike the theoretical aspects of astronomy, geography was valuable to the Romans because it directly related to the ever-expanding empire. In the second century c.e. the so-called Scipionic Circle, a coalition of thinkers who were intellectually and internationally minded, were sold on the ideas of Polybius, a visiting Greek, as to the importance of knowledge of both geography and politics. Following the destruction of Carthage in 146 b.c.e., Polybius explored the western Mediterranean and adjacent Atlantic Ocean regions, including northwestern Africa, and his accounts stimulated the Roman imagination as well as their practicality.

Fanciful Notions. Roman politicians read Homer and, through Crates, they celebrated Homer’s geography. According to Crates, Homer (although more aptly the Pythagoreans) conceived of the earth as consisting of four inhabited landmasses, transversely and diagonally opposite to one another, separated in the Western and Eastern hemispheres by two oceans running longitudinally from the northern and southern poles. The turbulence caused by the clash of ocean streams at the poles resulted in the ebb and rise of tides. Crates’s fanciful notions of Homer’s universe influenced Posidonius and Cicero. Later, Macrobius and Martianus Capella borrowed from Crates, and their works inspired Western medieval thought, especially regarding the landmasses in the Western hemispheres.

Roman Maps and Cartography. As early as 174 b.c.e. a Roman politician reportedly erected a map of Sardinia in a temple to visualize his family’s military exploits. The tradition was established to display regional maps in temples and public places. Varro (116-27 b.c.e.) described a painted map where political leaders spoke to the public. In 44 b.c.e. Julius Caesar commissioned four Greeks, each of whom was assigned a region, to survey and produce a map of Asia, Europe, and Africa, but the project was never brought to fruition. Based on Caesar’s earlier project, Augustus assigned Agrippa to produce a new map of the inhabited world. The undertaking was a major project, one that Agrippa did not live to conclude, but Augustus had it completed. It was erected on a wall of a portico named after Agrippa. Some have speculated that the map was circular, but, judging by its location, that seems unlikely. Although it does not survive, Pliny the Elder employed it in detail for the sections on geography, but Pliny’s account ranged from criticism to admiration. Pliny’s lengthy descriptions of geography, especially his focus area of northern Europe and Germany, were typically Roman in outcome; he rejected theoretical geography (ratio, “reason”) and emphasized the practical and empirical (usus, “habit,” and observatio, “observation”). A reduced version of Pliny’s description of the three continents was incorporated into Martianus Capella’s textbook and served the early Middle Ages for much of its geographical knowledge.

Far and Near Lands Described. In the familiar pattern, a Greek, Strabo of Amasia (64/63 b.c.e. -25 c.e.), completed the best work on Roman geography in sixteen books, called Geographia. Strabo traveled widely from Armenia and the Black Sea to Ethiopia along the Nile and from Italy to Syria; he combined what he knew and saw with a wealth of information drawn from earlier writers, now lost, and, we presume, personal accounts of travelers. Even though his interests were more along the lines of human geography, his work was influenced by the scientific geography of Eratosthenes, Eudoxus, and Hipparchus (mathematics and cartography). He employed Posidonius (information on Spain and Gaul), Caesar (Gaul and the Germanies), Artemidorus (Asia Minor and Egypt), and Apollodorus of Athens (Greece). He was, however, unfamiliar with Agrippa’s map and survey. Like Eratosthenes, he believed the three continents to be sur-rounded by a single ocean. He accepted the 360 degrees for marking coordinates by latitude and longitude, with the prime meridian at Rhodes, and described earthquakes, volcanoes, and various geological activities, such as the upheaval of landmasses from seabeds, proven by images of sea animals in rock quarries high in the mountains. Clearly Strabo’s interest was in descriptive or regional geography and not mathematical geography, although in one chapter he discussed the shape of the earth and methods for measuring distance on the earth. His contributions were primarily with a description of peoples and regions, a discourse made all the better by his extensive travels. Westward he traveled only as far as northern Italy, and he had to rely on others’ accounts for information about Spain, Gaul, and northern Europe. His habitable world was between the Caspian, Arabian, Persian, and Mediterranean Seas and the Atlantic Ocean. Strabo’s extensive geography does not appear to be known or, at least, widely disseminated to western Romans and may have been published only in the Greek-speaking world.

Geography in Latin. Pomponius Mela (flourishing 44 c.e.) wrote the first and only treatise in Latin dedicated to geography. Systematically he described the habitable world with its three continents (Asia, Africa, and Europe). Being a native of southern Spain, he started with Atlantic Europe and moved eastward. Uninterested in mathematical geography, Mela assumed the spherical shape of the world. His descriptions of northern Europe (including Britain, Ireland, and northern Germany) were superior to Strabo’s account. Along with good narrations about peoples and places, occasionally he relied on previous authorities to relate fantastic accounts of customs and habits of curious peoples.

Geography and “Natural History”. Writing in Latin, although he employed many Greek sources, Pliny the Elder (circa 23-79 c.e.) wrote the Natural History, in which books three through six are devoted to geography. Much of their value lies with his relating data from Agrippa’s map with some harsh criticisms and occasional lapses of praise. Pliny’s pen, like his tongue, had a sting that hurt many a writer whom he freely criticized. Like Mela, Pliny was primarily concerned with descriptive geography. Pliny had served in the Roman army along the German frontiers, and he wrote a history, now lost, of the German wars and peoples; his account of Germany is restricted to less than two pages in modern print. Pliny regarded nature as divine and terra (“the earth”) largely as benignitas naturae (“the kindness of nature”) and man’s friend; he excused nature for making some areas poor in rainfall, with inadequate soil, and subject to severe weather. Such conditions, he argued, should instruct people how to live within the rules established by nature. Ironically, Pliny died in one of nature’s less benign moments, the eruption of the volcano Mount Vesuvius, during a daring rescue attempt.

The World According to Ptolemy. Around 150 c.e. Ptolemy wrote a Geographia (Geographical Manual) in eight books that was antiquity’s greatest expression of world geographical knowledge. The text consists of place-names, each with longitudes and latitudes. It was written with maps, including a world map, but likely some alterations were made during the centuries following. The area of the Roman map was fairly accurately depicted, whereas northern European regions, Africa (especially western coastlines), and India have major errors. India, for example, has its triangular shape between its east and west coasts depicted less acute than it is, and the tip of the island of Sri Lanka is drawn far too large. To the north, in Asia he has China and what may be the Malay Peninsula. The easternmost site identified is Cattigara, which some scholars believe may be modern Hanoi. The continuation of a wraparound between China and Africa is cause for speculation that the continent of South America is implied, but there is no direct evidence for the assertion.

Sources

Mary Beagon, Roman Nature: The Thought of Pliny the Elder (Oxford: Clarendon Press, 1992; New York: Oxford University Press, 1992).

O. A. W. Dilke, Greek and Roman Maps (Ithaca, N.Y.: Cornell University Press, 1985).

Roger French and Frank Greenaway, eds., Science in the Early Roman Empire: Pliny the Elder, His Sources and Influence (Totowa, N.J.: Barnes & Noble, 1986).

J. B. Harley and David Woodward, eds., The History of Cartography, volume 1, Cartography in Prehistoric, Ancient, and Medieval Europe and the Mediterranean (Chicago: University of Chicago Press, 1987)

About this article

geography

All Sources -
Updated Aug 13 2018 About encyclopedia.com content Print Topic