Deserts

views updated May 09 2018

DESERTS

DESERTS . In areas of continuous occupation, the presence of the sacred transcends and resolves the stresses produced by the environment. In the desert, humankind, deprived of the support of social solidarity and helplessly confronted by supernatural forces, is beset by anguish and fear.

The Desert and Personal Religious Experience

The first visions of the desert, therefore, are pessimistic. It is the region of the savage beasts and malevolent spirits, of demons of all kinds. In primitive societies it is the place of trials, of initiations. It is the place to which the rejected and the exiled are banished: Cain (Gn. 4:1116), Hagar and Ishmael (Gn. 21:915), and the scapegoat that was burdened with the sins of Israel (Lv. 16:810). Particularly characteristic of the most ancient sedentary societies of the Middle East (Haldar, 1950), this conception was long lived. For the prophets of Israel (e.g., Ez. 20) and in the accounts of the Exodus, the time in the desert is that of infidelity (Ex. 17:7), of the golden calf (Ex. 32), and of punishment before the entrance into the Promised Land.

But another, parallel attitude also developed: the desert as apprenticeship and self-knowledge. As a terrain of struggle, the desert leads to the discovery of one's own being and, thereby, to the affirmation of the individual. At a more evolved stage of religious thought, it is the privileged place of divine revelation, of the betrothal of Israel with Yahveh (Gillet, 1949), of the offer of alliance, and of law that brings liberation. After the infidelities in the land of Canaan, it was by means of a return to the desert, the place of love and intimacy with the divine, that reconciliation with Yahveh was achieved (Hos. 2:1416; Jer. 2:23). The desert thus becomes a refuge from corruption and depravity. Philo Judaeus (d. 4550 ce) adds to this specifically Jewish conception a theme of Hellenistic mysticismthe romantic yearning of the world-weary city dweller for solitude, for retreat to the desert, where he can find peace. The desert, where the air is pure and light (On the Contemplative Life 2223), assumes for Philo an absolute value. It was for this reason that God gave his laws to his people "in the depths of the desert" (On the Decalogue 2). This idea leaves its trace, then, through a whole series of Christian authors, for example, Origen: "John the Baptist, fleeing the tumult of the cities, went into the desert, where the air is purer, the sky more open, and God more intimate" (Homilies on Luke 11). It is the point of departure of the entire Christian monastic movement toward settlement in the desert after the fourth centurya movement that to a large extent regains the primitive pessimistic vision in its land of choice, Egypt.

The desert is where the devil is encountered and where Christ contended with him (Mauser, 1963). Monastic asceticism developed as a struggle in a fearsome place that was the land of demons par excellence. But this struggle was victorious. The presence of the pious anchorites integrated the desert into the realm of faith either by transforming it into a city, desertum civitas (Athanasius, Life of Anthony 14; cf. Chitty, 1966), or by fertilizing it and making it bloom according to the prophecy of Isaiah 35:1: desertum floribus vernans (Jerome, Letters 14). Finally, it was in the desert that the monks would find hēsuchia, the serenity of solitude. Christian tradition would thus, in the course of centuries, base the movement toward the solitary life that was to become as essential component of it on the image of the desert as a place of solitude (ēremia ).

This approbation of solitude, brought to its apogee in Christianity, was to be more or less present in all higher religions in which the ascetic imperative is based on meditation, which is facilitated by life in the desert: Buddhism, particularly in the Tantric forms, Daoism, and Islam. However, in the case of Islam, the acceptance was relatively cautious. For the Muslims, the desert was above all a thème d'illustration (Arnaldez, 1975). It is necessary to "realize" the desert of spiritual solitude before the sole existent being, God. But it is dangerous to actually take abode in the desert. It is true that there man can avoid hypocritical ostentation (riʾāʾ ) and the artificial social role that destroys authentic sincerity (ikhlās ). But he thereby runs the risk of arrogance, of developing the cult of his inner self. This more reserved attitude of Islam in regard to the reality of the desert does not affect the value attributed to the desert as a synonym for solitude and retreat, ending in the solitariness (infirād ) that culminates in mystical ascension. Both by its nature and its symbols, the desert brings man closer to God.

The Desert in the History of Religions

Does this proximity to the divine lead to the development of a particular religious structure? Ernest Renan thought so. In his Histoire générale des langues sémitiques (1885), inspired by the long Christian tradition, he wrote: "The desert is monotheistic. Sublime in its immense uniformity, it first revealed to man the idea of infinity, but not the perception of an unceasingly creative life that a more fertile nature inspired in other races" (p. 6). He later returned to this idea and defined it more precisely in his Histoire du peuple d'Israël (vol. 1, 1887, pp. 45, 59). He found the basis for the development of primitive monotheism, which he attributed to the Semitic peoples, to be "the customs of nomadic life," where there is little room for cultic practice and where "philosophical reflection, exercised intensely within a small circle of observation, leads to extremely simple ideas." More than a century after Renan, his idea of the desert as source and origin of Jewish mono-theistic thought would again inspire the works of a master of biblical archaeology, William Albright (1964, pp. 154156).

Actually, this idea is now largely outdated and has been vigorously disputed. All the studies on pre-Islamic Arab religion in particular (Wellhausen, 1897; Ryckmans, 1951; Henninger, 1959) have drawn a picture of it that has little to do with monotheism and that associates with the supreme deity, Allah, a numerous and varied cortege of deities. The desert is the domain of polymorphous and diffuse ritual. It is, to repeat the expression of the Qurʾān, "associationistic," and the desert bedouin are "the most obdurate in their impiety and hypocrisy" (sura 9:97).

However, the situation remains ambiguous. Beside and above the other gods was Allāh. He was incontestably a god of the nomads, a provider of rain (Brockelmann, 1922), and one can easily imagine the extreme importance that he assumed for a nomad whose survival depends entirely on the condition of the grazing land. Not only among the Arabs was the god of rain the unique god. It has been possible to reconstruct the special characteristics of the religion of pastoral peoples in general in conjunction with the peculiarities of their social structure and their way of life; such a reconstruction has been made for the first time in an environment very different from the desert, namely, that of the high grassy savannas of East Africa (Meinhof, 1926). The creation of powerful personages, of heroic saviors who are then frequently enrobed in historic myths connected with the origins of the tribe, is an expression of the instability of the pastoral tribe, which assembles or disperses in accordance with the importance of the individuals that direct and guide it within a context of aggressive relationships between groups. This orientation often accompanies that which consists of making the god of rain the unique god, dispenser of all benefices. The herdsman soon breaks free of polytheism. Ancestor worship is unknown to him, and the dead are forgotten. The herdsman is knowledgeable and intrepid, little inclined toward fear and superstition.

One is here indeed on the way to the monotheistic god, but under the impetus of a somewhat different logic. It is not that the desert is monotheistic but rather that the pastoral nomad has the tendency, at least, to become monotheistic. In contrast with the profusion of rituals in the Australian desert traversed by primitive hunters or gatherers is the evolution toward the monotheistic god of the warlike pastoral tribe, herders of large beasts of the African savannas or the bedouin of the deserts of the Old World. This trend toward monotheism is a late development in the cultural history of humanity. Contrary to Wilhelm Schmidt's opinion, this "great god of the herdsmen" is not the legacy of a primitive monotheism. Today we know that the pastoral nomads were, in the main, descendants of the first agrarian civilizations of the Old World, or were at least posterior to them. They constitute, in the world of the deserts and the steppes, a relatively recent cultural development, first in the form of the pre-bedouin herders of bovines, and then, after the domestication of the horse and the dromedary, in the form of the widespread, aggressive, warlike nomadism of the bedouin type, which in the Middle East does not go back further than the second millennium bce.

But the appearance of monotheistic tendencies in the tribes of pastoral nomads can be rapid, as is shown by the analysis of neopastoral civilizations of the New World (Planhol, 1975). The Navajo of the Colorado plateaus of North America, whose pastoral mode of life did not emerge until the second half of the eighteenth century, still do not recognize a supreme deity (Reichard, 1950). But among the Goajira of Colombia and Venezuela, whose aggressive, cavalieristic, pastoral life goes back at least two or three centuries, there is a process of elaboration that seems much more advanced. Involved here is the predominance of a demiurgic creator (Maleiwa) and a rain giver (Juya), who are, however, not yet confused with each other, although the first signs of such a confusion are evident (Perrin, 1976). Among the herdsmen of East Africa, such as the Maasai, whose formation of a pastoral system goes back at least a thousand years, pastoral monotheism is well defined.

But, rather than being a true monotheism, it is in fact a monolatry in that it is willing to recognize the existence of other gods, who fulfil the same functions for the benefit of neighboring tribes or peoples; it is a protomonotheism in Baly's sense of the term (1970, pp. 258259). The moral monotheism (Baly's "absolute" or "transcendent" monotheism) is a much more complex and revolutionary structure (Pettazzoni, 1950), the birth of which implies a break, not a simple evolution. Its occurrence exclusively in the Middle Eastern and Old World cultural environment reflects conditions of conflict. Here the presence of groups of pastoral nomads with still rather primitive monotheistic tendencies certainly played an essential role within the orbit of sedentary, sacerdotal civilizations that were polytheistic but much more highly developed (as Weindl, 1935, demonstrated). These conflicts could not be resolved except by a universalistic aspiration such as that of which the birth of Islam, following that of Judeo-Christian monotheism and Zoroastrian dualism, constitutes the final manifestation (Watt, 1953, 1956; Rodinson, 1961). Although Yahvism, as Nyström (1946) has shown, surpasses and in many particulars contradicts and transcends the bedouin ideal, this ideal is nonetheless necessary to it.

There is, therefore, no "religion of the desert." But, in the historical evolution of humanity in the Old World, the deserts have indeed been the privileged place for the development of the pastoral nomadic cultures that evolved precociously toward monotheism and that constituted an essential component in the genesis of the great monotheistic religions.

See Also

Arabian Religions; Eremitism; Monotheism; Retreat.

Bibliography

Albright, William F. History, Archaeology, and Christian Humanism. New York, 1964.

Arnaldez, Roger. "Le thème du désert dans la mystique musulmane, thème d'inspiration ou thème d'illustration." In Les mystiques du désert, pp. 8996. Gap, France, 1975.

Baly, Denis. "The Geography of Monotheism." In Translating and Understanding the Old Testament, edited by Harry Thomas Frank and William L. Reed, pp. 258278. New York, 1970.

Bartelink, G. J. M. "Les oxymores desertum civitas et desertum floribus vernans. " Studia Monastica 15 (1973): 715.

Brockelmann, Carl. "Allah und die Götzen, der Ursprung des islamischen Monotheismus." Archiv für Religionswissenschaft 21 (1922): 99121.

Chitty, Derwas J. The Desert a City. Oxford, 1966.

Gillet, Jacques. "Thème de la marche au Désert dans l'Ancien et le Nouveau Testament." Recherches de science religieuse 36 (AprilJune 1949): 161181.

Guillaumont, Antoine. "La conception du désert chez les moines d'Égypte." In Les mystiques du désert, pp. 2538. Gap, France, 1975.

Haldar, Alfred O. The Notion of the Desert in Sumero-Accadian and West-Semitic Religions. Uppsala, 1950.

Henninger, Joseph. "La religion bédouine pré-islamique." In L'antica societa beduina, edited by Francesco Gabrieli, pp. 115140. Rome, 1959.

Mauser, Ulrich W. Christ in the Wilderness: The Wilderness Theme in the Second Gospel and Its Basis in the Biblical Tradition. London, 1963.

Meinhof, Carl. "Religionen der Hirtenvölker." In his Die Religionen der Afrikaner in ihrem Zusammenhang mit dem Wirtschaftsleben, pp. 7184. Oslo and Cambridge, Mass., 1926.

Les mystiques du désert dans l'Islam, le judaïsme, et le christianisme. Gap, France, 1975. Papers delivered at a conference of the Association des Amis de Sénanque, 28 July3 August 1974.

Nyström, Samuel. Beduinentum und Jahwismus: Eine soziologisch-religionsgeschichtliche Untersuchung zum Alten Testament. Lund, 1946.

Perrin, Michel. Le chemin des Indiens morts: Mythes et symboles goajiro. Paris, 1976.

Pettazzoni, Raffaele. "La formation du monothéisme." Revue de l'Université de Bruxelles 2 (1950): 209219.

Planhol, Xavier de. "Le désert, cadre géographique de l'expérience religieuse." In Les mystiques du désert, pp. 516. Gap, France, 1975.

Reichard, Gladys A. Navajo Religion. 2 vols. New York, 1950.

Rodinson, Maxime. Mahomet. Paris, 1961. Translated by Anne Carter as Mohammed (New York, 1971).

Ryckmans, Gonzague. Les religions arabes préislamiques. 2d ed. Louvain, 1951.

Schmidt, Wilhelm. Der Ursprung der Gottesidee, vols. 712, Die Religionen der Hirtenvölker. Münster, 19401955.

Watt, W. Montgomery. Muammad at Mecca. London, 1953.

Watt, W. Montgomery. Muammad at Medina. London, 1956.

Weindl, Theodor. Monotheismus und Dualismus in Indien, Iran und Palästina als Religion junger, kriegerisch nomadistischer Völker im Gravitations bereich von Völkern alter Kultur. Vienna, 1935.

Wellhausen, Julius. Reste arabischen Heidentums. 2d ed. Berlin, 1897.

Xavier de Planhol (1987)

Translated from French by Roger Norton

Deserts

views updated May 23 2018

Deserts

Desert ecosystems are characterized by an extremely arid, arid, or semi-arid climate, low relative humidity, high air and soil temperatures, strong winds, high solar radiation, low precipitation levels, extended drought periods, soils low in organic matter, low net primary productivity, and a spatially patchy distribution of vegetation and soil resources. In them, water is the predominant controlling factor for most biological processes; precipitation is highly variable and occurs as infrequent and discrete events throughout the year; and precipitation events are highly unpredictable in both space and time. Desert ecosystems may be classified into three groups based on annual precipitation: extremely arid (less than 60 millimeters), arid (60 to 250 millimeters), and semiarid (250 to 500 millimeters). The plant communities of arid lands expand and contract in accordance with alternating wet and dry periods as well as with anthropogenic activities that contribute to desertification (also known as land degradation). While arid ecosystems occur on all continents in both hot and cold environments, this article will not focus on polar deserts.

Distribution of Deserts Worldwide

Earth's major deserts lie within the tropics of Cancer and Capricorn where stable, high atmospheric pressure creates an arid climate at or near latitudes 30°N and 30°S. Deserts are generally located in the interior of large continents. Continental deserts are separated from ocean moisture by large distances or topographic barriers, such as large mountain ranges, which create a rainshadow. Deserts may also be situated on the west coast of large continents adjacent to cold ocean currents, which draw moisture away from the land. Subtropical deserts, such as the Mojave Desert of California, lie within the latitudes of 30°N and 30°S. Cool coastal deserts, including the Peruvian Atacama Desert, occur where cold offshore currents generate high atmospheric pressure and large masses of dry air, which create arid conditions upon their descent. Rainshadow deserts, including the Great Basin Desert in the United States or the Gobi Desert in Mongolia, occur where a topographical barrier such as a mountain range interrupts the flow of moist oceanic air. As moisture-laden air masses travel inland, they are deflected upward on the windward side of a mountain range, lose their moisture, and descend as dry air masses on the leeward side of the mountains. Continental interior deserts, such as the Great Sandy Desert in Australia, occur far from marine moisture.

Plants in the Desert Environment

In order to understand the ways in which plants have adapted to arid lands, it is essential to consider the physical environment. Of all the abiotic constraints imposed on plant activityhigh air temperatures, extremely high soil temperatures, high winds, intense solar radiation, and limited moisturehigh temperatures and limited water are the two factors that severely limit plant growth. Summer air temperatures in the Sonoran Desert in Arizona may reach 40°C during the day but drop to 15°C at night. Soil temperatures may reach 80°C or higher. High temperatures generally are accompanied by strong winds in coastal deserts, such as the Atacama in South America and the Namib in Africa, as well as in continental deserts, including the Chihuahuan and Sonoran in the United States. As well as producing spectacular dust storms and dust devils (small whirlwinds containing sand or dust), wind also abrades and desiccates desert plants.

Water is the single-most limiting factor to the growth and productivity of desert vegetation. The highly sporadic nature of desert rainfall creates a pulse-reserve system of water and nutrient availability that influences many biological processes, especially plant productivity. In the Chihuahuan Desert of New Mexico, gentle winter rainfall penetrates deep into the soil profile and provides most of the moisture for the growth of perennial shrubs, such as creosote bush and mesquite. In contrast, the high-intensity, brief summer thunderstorms provide minimal water for plant growth because most of the water runs off of the soil surface. Many plant species take advantage of rainfall immediately and grow rapidly following precipitation events, then slow their growth when soils dry and moisture once again becomes limiting.

Second only to moisture, the availability of soil nutrients, primarily nitrogen and phosphorus, limits plant productivity in deserts. Nitrogen is the key limiting nutrient in North American deserts, phosphorus is most limiting in Australian deserts, while nitrogen, phosphorus, and potassium are limiting in sand dune communities in Africa's Namib Desert. Soil nutrients and organic matter tend to be concentrated in the upper 2 to 5 centimeters of soil with the greatest amounts underneath the canopies of individual desert shrubs in "islands of fertility." These resource islands harbor greater concentrations of water, soil nutrients, and microorganisms than adjacent soils.

Certain plant species, such as creosote bush, are often referred to as nurse plants. Nurse plants effectively reduce high-incident solar radiation and high temperatures under their canopies and create ideal sites for seed germination and seedling growth. The concentration of limiting resources in islands of fertility or under nurse plants generates a spatially patchy distribution of vegetation across the desert. Competition for water maintains this spacing of plants. While this phenomenon has been most studied in U.S. deserts, it occurs in arid lands worldwide.

Desert Soils

Hot deserts exhibit generally similar soil types. Immature and alkaline with weakly developed soil horizons , desert soils are dry most of the year, and poor in soil organic matter, nitrogen, and phosphorus, but are rich in inorganic ions, carbonate, and gypsum. The main soil orders of hot deserts are Entisols, soils without well-defined layers that are formed from recently exposed rock, and Aridisols. Aridisols, exclusive to arid regions, contain two dominant suborders: Orthids and Argids. Orthids are young calcareous and gypsipherous soils with a caliche (or calcium carbonate hardpan) within 1 meter of the soil surface. The thickness of the caliche layer has been correlated with the size of creosote bush shrubs in Arizona's Sonoran Desert: the thicker the layer, the smaller the shrubs. Argids are older soils and lack the carbonate hardpan layer, but are clay-rich and may be good agricultural soils when water is available.

Plant Adaptations to the Desert Environment

Desert plant species show various physical, physiological, and pheno-logical (timing of growth and reproduction) characters that enable them to survive and grow in arid, nutrient-limited environments. Some plants, such as summer and winter desert ephemerals , restrict all growth and flowering to periods when water is available. They are able to withstand droughts and high water stress because their underground rhizomes or bulbs remain dormant during the dry season. In extreme droughts, desert ephemerals may remain completely dormant, eliminate reproduction, or limit growth to the vegetative phase. Other species, such as the California poppy and other desert annuals, complete their entire life cycle during the rainy season. Their long-lived seeds germinate only under suitable environmental conditions. As a result, they respond to the pulse-reserve system of resource availability, showing high rates of primary production in favorable years and minimal, or no production, in drought years. Ephemerals and annuals, while showy, produce minimal biomass .

In deserts worldwide, perennial shrubs and subshrubs, such as the creosote bush and jojoba, produce most of the desert plant biomass. These species limit water loss and reduce heat loads at the leaf surface by limiting the surface area to many small single, dissected, or compound leaves covered with a waxy cuticle or leaf hairs. Most shrubs have canopies with a compact globe or inverted cone shape. This morphology allows water to funnel directly to the plant roots and reduces the amount of surface area that is exposed to sunlight. Perennials have a large root-to-shoot ratio, and most roots are distributed in the soil in one of two ways. The roots may be confined to the upper meter of the soil profile and fan out horizontally from the base of the shrub, enabling shrubs access to even the slightest rainfall. Alternatively, the roots may extend deep into the soil profileup to 12 meters with mesquiteand allow plants to obtain water that is stored at these depths. As with other desert plants, perennials may also limit or suppress flowering and fruiting in years of extreme drought.

Perennials are able to remain metabolically active at very low soil- and plant-water potentials, high internal water deficits, and high temperatures. They have sensitive regulation of leaf stomata as a function of internal and external conditions, including water stress, temperature, atmospheric humidity, and light intensity. Most shrub species acquire carbon throughout the C3 photosynthetic pathway, despite the fact that the alternative C4 pathway is thought to increase the amount of carbon gain per unit of water used (water-use efficiency [WUE]). The only desert perennials that have the C4 pathway are the halophytic (salt-tolerant) species, such as tamarisk, short-lived summer active perennials, and most grasses.

Cacti, common to deserts, show unique adaptations to the desert environment. They have shallow, horizontally extended root systems, an upright, ribbed trunk that reduces the midday heat and solar radiation load and water storage within their trunks. Saguaro cacti, located near Tucson, Arizona, expand and contract like an accordion depending on the moisture conditions. In wet years the cacti are plump and green, but in dry years they are slim and yellow-green in color. Because cacti lack typical broad leaves, the overall green coloring derives from the photosynthetic trunk. Over evolutionary time, cactus "leaves" have been reduced to hairlike spines that reflect solar radiation or spikelike spines that protect the plant from herbivores . Other noncactus species, such as ocotillo and the boojum trees native to Baja California, produce photosynthetically active leaves only in wet years and limit photosynthesis to the stems when drought prevails. Cacti and other succulent species obtain carbon through the crassulacean acid metabolism (CAM) photosynthetic pathway. CAM photosynthesis allows the cacti to open their stomata only at night in order to reduce water loss.

see also Cacti; Desertification; Photosysthesis, Carbon Fixation and.

Anne Fernald Cross

Bibliography

Caldwell, M. M., J. H. Manwaring, and S. L. Durham. "The Microscale Distribution of Neighboring Plant Roots in Fertile Soil Microsites." Functional Ecology 5 (1991): 765-72.

Cross, A. F., and W. H. Schlesinger. "Plant Regulation of Soil Nutrient Distribution in the Northern Chihuahuan Desert." Plant Ecology 145 (1999): 11-25.

Fox, G. A. "Drought and the Evolution of Flowering Time in Desert Annuals." American Journal of Botany 77 (1990): 1508-18.

Le Houérou, H. N. "Climate, Flora and Fauna Changes in the Sahara Over the Past 500 Million Years." Journal of Arid Environments 37 (1997): 619-47.

Mahall, B. E., and R. M. Callaway. "Root Communication Mechanisms and Intra-community Distributions of Two Mojave Desert Shrubs." Ecology 73 (1992): 2145-51.

Martinez-Meza, E., and W. G. Whitford. "Stemflow, Throughfall, and Channelization of Stemflow by Three Chihuahuan Desert Shrubs." Journal of Arid Environments 32 (1996): 271-87.

McAuliffe, J. R. "Markovian Dynamics of Simple and Complex Desert Plant Communities." American Naturalist 131 (1988): 459-90.

Nobel, P. S. Environmental Biology of Agaves and Cacti. Cambridge, UK: Cambridge University Press, 1988.

Schlesinger, W. H., J. F. Reynolds, G. L. Cunningham, L. F. Huenneke, W. M. Jarrell, R. A. Virginia, and W. G. Whitford. "Biological Feedbacks in Global Desertification." Science 247 (1990): 1043-48.

, J. Raikes, A. E. Hartley, and A. F. Cross. "On the Spatial Pattern of Soil Nutrients in Desert Ecosystems." Ecology 77 (1996): 364-74.

MAJOR DESERTS OF THE WORLD

North America:

Great Basin, Sonoran, Mojave, Baja California, Chihuahuan

South America:

Patagonian, Puna, Monte, Chaco, Espinal, Peruvian-Chilean, Atacama

Asia:

Gobi, Takla Makan, Iranian, Thar, Syrian, Arabian, Sinai, Negev

Africa:

Sahara, Sahel, Somalian, Namib, Karoo, Kalahari, Madagascar

Australia:

Great Sandy, Gibson, Great Victoria, Arunta, Stuart

Deserts

views updated May 21 2018

DESERTS

Predominant landscape of the Middle East and North Africa.

Stretching from the Atlantic coast in the west to Pakistan in the east, a band of arid land (15° and 30° north latitude) dominates this region. The North African expanse is generally known as the Sahara, although subdivisions within it have individual names indicating the nature of the surface. The terms erg (as in the Great Eastern Erg of Algeria) and serir (as the Serir of Kalanshu in Libya) indicate a region of sand dunes. Where the surface is rocky underfoot the terms used are reg or hamada (for example, the Hamada of Dra south of the Anti-Atlas mountains). Individual areas may also be given the name desert, the Western Desert and the Eastern Desert in Egypt, although they are smaller parts of the whole. On the peninsula of the same name, the Arabian Desert is an extension of the Sahara and is divided into the Rub al-Khali (Empty Quarter, a region of vast sand dunes) and the Nafud and Najd. To the north is the Syrian Desert, and to the east the two deserts of the Iranian plateau are known as the Dasht-e Kavir and the Dasht-e Lut.

The term desert is one in common usage and therefore difficult to define. Most experts prefer to speak of "drylands" or "arid lands" and to define such places through various measures of the availability of water for plant growth (implying that not all deserts are hot.) A common definition of desert, however, is those regions of Earth's surface having fewer than 10 inches (250 mm) of precipitation annually and extreme high temperatures. This classical approach relates such measures to areas with types of vegetation adapted to hot, arid conditions. In areas with much sunshine and small amounts of precipitation and/or natural moisture from the soil, only plants called xerophytes survivethose adapted to such conditions. In certain hyperarid locations, precipitation may be even less and no vegetation of any kind is found.

Desert rainfall is not only sparse but is also extremely variable in time and space as well as in quantity. Such variance means that human occupancy of the desert must depend for survival on reliable springs and rivers for irrigation rather than on precipitation. Traditional pastoral nomadism, located on the desert margins, was adapted to this environment by moving its productive units (i.e., herds and flocks) to where grass and water seasonally occurred. But even nomads ventured into the true desert only for travel as transporters and raiders. The few permanent inhabitants of the deserts were those oasis dwellers dependent upon perennial springs for intensive agriculture and the growing of date palms.

Desert soils are usually of poor quality except for those in the valleys of rivers where alluvial deposits have accumulated. True desert soilscalled aridisols have low biomass, very sparse or no organic acids and gases, few or no bacteria, and are essentially mineral in character. Any rain or sheet flooding and runoff that percolate beneath the surface rapidly evaporate. As a result, soluble salts are precipitated and redeposited, forming a crusty layer on the surface or just beneath it. Repeated leaching and deposition can result in concentrations of sodium chloride (NaCl), white alkali (salt), or similar deposits of sodium carbonate (Na2CO3), black alkali, which poison the soil and make agriculture impossible. Under desert conditions agriculture is extremely difficult, and even the use of irrigation water can cause salinity, through evaporation and the precipitation of the dissolved salts it may carry, which leads to the abandonment of such farmland.

The natural xerophytic vegetation found in deserts has adapted to conditions of high temperatures and scant and irregular amounts of precipitation. Xerophytes often occur as drought-resisting plants with heavy cuticles, which reduce transpiration, or with stomata, which can be closed for the same purpose. Other xerophytes reduce water use by shedding their leaves and remaining leafless during the dry season. Among these plants are the euphorbia and the cacti, the latter originally found only in the Western Hemisphere.

Phreatphytes constitute another class of desert vegetation, which includes palms. These plants have developed long taproots, which reach the water table, allowing them to survive the driest of surface conditions. Other plants evade drought by flowering and seeding only during brief rainy periods.

During the intervening months and years of drought, the seeds remain dormant.

Desert vegetation under such conditions is sparse, and soil-forming conditions (including the creation of humus) are poor. Rainstorms can be intense, although of short duration, and often soil particles are carried away from desert surfaces by sheet flooding. The result of these conditions is erosionwhich results in hills lacking deep layers of soil. Their profiles are characteristically steep sided with thick strata forming cliff faces rising vertically from the surrounding plains. Flat-topped mesas and steep buttes dominate the landscape, while valleys are flat bottomed with vertical side slopes. Wind erosion and deposition are also significant factors in desert landscape formation. Crescent-shaped barchan dunes are found where sands are insufficient to completely mantle the underlying surface. Copious sands form "seas," with longitudinal sief dunes and star-shaped rhourd dunes. Such seas, however, are the exceptions and rocky desert surfaces are common.

In desert areas, underground supplies of water assume great importance. Porous and permeable strata deep beneath the surface sometimes contain large quantities of water. Such aquifers may have impervious layers (aquicludes) above and below them that confine the water and keep it from escaping except in limited amounts at oases. Other aquifers occur in unconsolidated alluvial materials in river valleys (Arabic, wadis ). This water is recharged from river seepage and/or rainfall. In the Middle East, most of the major aquifers are non-renewable and contain fossil water, which once usedextracted or minedwill not be replaced. Desert countries, such as Libya and Saudi Arabia, with few or no surface streams have in the last two decades turned to the exploitation of such aquifers as part of their economic development plans. An ambitious agricultural program in Saudi Arabia has used tube wells and central pivot irrigation to produce bumper wheat crops in an otherwise hostile desert environment. Libya is engaged in constructing a "Great Manmade River"actually a gigantic system of pumps and pipelineswith which to bring water from aquifers beneath the central Sahara to coastal locations, for municipal and agricultural use. In both these cases and others, the critical element is the quantity of water available and whether it will last long enough to justify such expensive projects. Many experts counsel caution in undertaking such attempts to remake, or "green," the desert.

See also Climate; Desalinization; Eastern Desert; Geography; Nafud Desert; Syrian Desert; Water.


Bibliography

Beaumont, Peter. Environmental Management and Development in Drylands. London and New York: Routledge, 1989.

Goudie, Andrew, and Wilkinson, John. The Warm Desert Environment. Cambridge, U.K., and New York: Cambridge University Press, 1977.

Whitehead, Emily E.; Hutchinson, Charles F.; Timmerman, Barbara N.; et al., eds. Arid Lands: Today and Tomorrow: Proceedings of an International Research and Development Conference. Boulder, CO: Westview Press, 1988.

john f. kolars

Deserts

views updated May 09 2018

DESERTS

DESERTS. Definition has been the central problem in the history of the deserts of the United States. The need to ascertain the limits of arability and the difficulty of establishing such boundaries where precipitation fluctuates unpredictably constitute a basic developmental theme for more than half the nation. Archaeological evidences of prehistoric Native American communities indicate that droughts occasioned recurrent disaster to agricultural


societies long ago, as now, in border areas. In 1803 President Thomas Jefferson, seeking congressional support for exploration of the upper Missouri River, summarized existing knowledge of newly purchased Louisiana in describing it as a region of "immense and trackless deserts" but also, at its eastern perimeter, as "one immense prairie"—a land "too rich for the growth of forest trees." The subsequent expedition of Meriwether Lewis and William Clark (1804–1806) marked the official beginning of American efforts to elaborate the description.

Until the 1860s a conception prevailed that the vast province west from the meridian of Council Bluffs, on the Missouri River, to the Rocky Mountains, between thirty-five and forty-nine degrees north latitude, was a "Great American Desert." The explorations of Lewis and Clark, Zebulon Pike, and Stephen Harriman Long, followed by the experiences of traders to Santa Fe, Rocky Mountain fur trappers, immigrants to Oregon and California, soldiers along the Gila Trail, surveyors for transcontinental railroads, and prospectors throughout the West confirmed the appellation.

While commentators agreed that agriculture could have no significant role in the region, they did occasionally recognize that the Great Plains, the mountain parks, and the interior valleys of California and the Northwest afforded excellent pasturage. As livestock industry developed in these areas during the period from 1866 to 1886, redefinition of the limits of aridity evolved. Maj. John Wesley Powell's surveys and, notably, his Report on the Lands of the Arid Region (1878) expressed the new point of view; agriculture, Powell asserted, could be profitably conducted in many parts of the West, but only as an irrigated enterprise and generally as a supplement to stock growing. The collapse of open-range ranching in the mid-1880s emphasized the need for expanded hay and forage production and gave impetus to development of irrigation programs. But Powell's efforts to classify the public lands and the passage of the Carey Desert Land Grant Act of 1894 raised controversy. States east of the 104th meridian were excluded, at the request of their representatives, from the application of the Carey legislation. Farmers during the 1880s had expanded cultivation without irrigation nearly to that meridian in the Dakotas and even beyond it in the central plains. Many were convinced that "rainfall follows the plow." They saw no need to assume the costs and the managerial innovations of supple-mental watering. A new conception of the boundaries of aridity was emerging.

Drought in the mid-1870s had driven a vanguard of settlers eastward from the James River Valley, a prairie zone normally receiving more than twenty inches of annual rainfall. Drought in the period 1889–1894 forced thousands back from the plains farther west, where average precipitation ranges between fifteen and twenty inches annually. As normal conditions returned, however, farmers in the first two decades of the twentieth century expanded cultivation across the plains to the foothills of the Rockies—in Montana, Colorado, and New Mexico—and in many areas beyond—Utah, Idaho, the interior valleys of California, and eastern Oregon and Washington. Irrigation supplied water to only a small portion of these lands. Dry farming—a specialized program that, ideally, combines use of crop varieties adapted to drought resistance, cultivation techniques designed to conserve moisture, and management systems that emphasize large-scale operations—provided a new approach to the problem of aridity. The deserts, promoters claimed, could be made to "blossom like the rose."

When severe droughts again returned from 1919 to 1922, and from 1929 to 1936, assessment of the effectiveness of dry farming raised new concern for defining the limits of aridity—an outlook most strongly expressed in the reports of the National Resources Board of the mid-1930s but one that still permeates the writings of agricultural scientists. Long-term precipitation records, with adjustment for seasonality and rate of variability in rainfall, humidity, temperature, and soil conditions, now afford some guidance to the mapping of cultivable areas.

By established criteria a zone of outright desert (less than five inches average annual precipitation) ranges from southeastern California, northward through the western half of Nevada, nearly to the Oregon border. Because cropping without irrigation is impracticable when rainfall averages less than ten inches annually, climatic pockets found in all states west of the 104th meridian—most prevalently in Arizona, central New Mexico, eastern Nevada, Utah, and the lee side of the Cascades in Oregon and Washington—may also be defined as arid. Semiaridity—an average precipitation of from ten to fifteen inches annually—characterizes the western Dakotas, much of Montana, and large sections of eastern New Mexico, Colorado, Wyoming, Idaho, Oregon, and Washington. There dry farming may be successful but only when management programs include allowances for recurrent drought. Throughout much of the semiarid region livestock production predominates, with cropping to afford feed and forage supplementary to native short-grass pasturage. In many areas, however, the possibility of raising wheat of superior milling quality, which commands premium prices, encourages alternative land utilization. The costs of marginal productivity must be carefully weighed.

Eastward, roughly from the Missouri River to the ninety-eighth meridian and curving to the west through the central and southern plains, is a subhumid zone, in which rainfall averages from fifteen to twenty inches annually, an amount sufficient, if well distributed, to permit cultivation without recourse to specialized programs but so closely correlated to the margin of general farming requirements that a deficiency occasions failure. Almost every spring, alarms are raised that some areas of the vast wheat fields extending from the central Dakotas, through western Kansas and Nebraska and eastern Colorado and New Mexico, into the panhandles of Oklahoma and Texas have suffered serious losses. There the problem of defining limits of arability is yet unresolved; the boundaries of America's deserts and arid regions remain uncertain.

BIBLIOGRAPHY

Fite, Gilbert C. The Farmers' Frontier, 1865–1900. New York: Holt, Rinehart and Winston, 1966.

Goetzmann, William H. Exploration and Empire: The Explorer and the Scientist in the Winning of the American West. New York: Knopf, 1966.

Hargreaves, Mary W. M. Dry Farming in the Northern Great Plains, 1900–1925. Lawrence: University Press of Kansas, 1993.

Limerick, Patricia Nelson. Desert Passages: Encounters with the American Deserts. Albuquerque: University of New Mexico, 1985.

Teague, David W. The Southwest in American Literature and Art: The Rise of a Desert Aesthetic. Tucson: University of Arizona Press, 1997.

Mary W. M.Hargreaves/a. r.

See alsoAgriculture ; Death Valley ; Great Plains .

Deserts

views updated Jun 11 2018

Deserts

Introduction

A desert is a biome where precipitation is rare and unpredictable. Many deserts lie in the continental interior, such as the Sahara of northern Africa, and the Mojave, which covers parts of California, Nevada, Arizona, and Utah. Few deserts are completely barren. Most have at least some vegetation, although it is adapted to dry conditions. For instance, desert cacti have spines rather than leaves to reduce water loss. Desert animals include invertebrates and small mammals, like kangaroo mice.

Deserts are always dry but they are not necessarily hot. For instance, Antarctica is a cold desert, while coastal deserts have cool winters and long, warm summers. A desert is a vulnerable environment whose vegetation is readily damaged, causing the soil to become drier still. Many formerly productive lands are turning into deserts, especially in Africa and China.

Historical Background and Scientific Foundations

Deserts cover one fifth of Earth’s surface and are characterized by their dryness. Precipitation, more usually in the form of rain, is rare and unpredictable. In the desert biome, the overall average amount of water falling is fewer than 12 in (30 cm) in a year. More rain than this may fall, but some of it is balanced out by the amount of water lost from the soil through evaporation. In some years, there may be no rain at all. On the other hand, there may sometimes be sudden, heavy rainfall which can even cause flooding.

Deserts are favored by warm, dry climates with high atmospheric pressure, which is why many are found in continental interiors at latitudes of around 30 degrees, North and South. However, coastal deserts also exist, such as the Atacama Desert in Chile, which is one of the driest places on Earth. Although the presence of the ocean normally provides plenty of precipitation through evaporation of surface waters, the Atacama Desert lies within the rain shadow of the Andes mountains.

There are four types of desert, classified according to climate. Hot, dry deserts are warm all year round, with very hot summers. Examples include the Mojave, Sahara, and Australian deserts. The atmosphere of a hot, dry desert contains very little moisture to filter the sun’s rays, which means the ground surface absorbs a lot of heat during the day and loses a lot of heat at night. This leads to daily temperature extremes with temperatures that can be as high as 115°F (46°C) during the day, while plunging to around 17°F (-8°C) at night. Rain tends to occur in short, intense, bursts and sometimes evaporating before it even reaches the ground. Soil is coarse, shallow, or rocky with good drainage.

The semi-arid desert, such as those of Utah and Montana, has moderately long, dry summers. Daily temperatures are less extreme than in hot, dry deserts. Soil texture varies from coarse to fine and sandy. Coastal deserts are found in moderately cool to warm areas, and are characterized by their cool winters and long warm summers. The coastal desert enjoys more rainfall than the hot, dry, and semi-arid deserts, amounting to a few inches or more a year. The soil of a coastal desert is fine-textured with a significant salt content. It is fairly porous and drains well. The cold desert, occurring in the Antarctic, Arctic, and Greenland regions, has snowfall and high rainfall in the winter. Summers are short and moderately warm. The soil of a cold desert is heavy, salty, and full of silt. Mean winter temperature will be between 35–39°F (2–4°C) and mean summer temperature is 77°F (25°C).

Deserts are home to a diverse range of plant and animal life. Species tend to be specially adapted to the dry environment with special features to conserve water. For

WORDS TO KNOW

BIOME: A well-defined terrestrial environment (e.g., desert, tundra, or tropical forest) and the complex of living organisms found in that region.

PRECIPITATION: Moisture that falls from clouds as a result of condensation in the atmosphere.

SEMIARID: Receiving very little annual rainfall, roughly 10-20 in (25-50 cm), and characterized by short grasses and shrubs.

instance, desert plants, like cacti, have tough leathery skin and spines, rather than leaves. Many are salt-tolerant. Some desert plants are also capable of blooming and setting seed very quickly, to make the most of sudden, intense rainfall.

Desert animals include reptiles and small mammals, such as kangaroo rats and pocket mice. They tend to pursue a nocturnal lifestyle, sheltering from the desert heat during the day. Desert rodents can get most of their moisture from their food and convert its fat con-

tent into water in their bodies. They also have highly concentrated urine and dry feces, which also helps reduce water loss. Large mammals are rare in the desert for they cannot store or conserve enough water to survive such dry conditions. The camel, whose hump stores fat, is an exception.

Impacts and Issues

The desert environment is easily damaged. For instance, tracks left by army tanks during World War II (1939–1945) are still visible in the California deserts. Vegetation has an important protective role in a desert, retaining moisture in the soil. Overgrazing causes the soil to drift and become even drier. Another important issue is desertification, which is the creation of new desert from previously productive land, with the accompanying risk of drought, crop failure, and hunger. The United Nations says that 80% of the world’s grassland is suffering from overgrazing and soil degradation. Three quarters of this area is now showing some signs of desertification.

See Also Desertification; Drought

BIBLIOGRAPHY

Books

Cunningham, W.P., and A. Cunningham. Environmental Science: A Global Concern. New York: McGraw-Hill International Edition, 2008.

Kaufmann R., and C. Cleveland. Environmental Science. New York: McGraw-Hill International Edition, 2008.

Web Sites

University of California Museum of Paleontology. “The Desert Biome.” http://www.ucmp.berkeley.edu/exhibits/biomes/deserts.php (accessed February 14, 2008).

Deserts

views updated Jun 08 2018

116. Deserts

See also 142. ENVIRONMENT .

eremite
a religious hermit living alone, often in the desert. eremitic , adj.
eremology
the systematic study of desert features and phenomena.
xerophobia
an abnormal fear of dryness and dry places, as deserts.