State (polity)

From Citizendium
Jump to navigation Jump to search
This article is developed but not approved.
Main Article
Discussion
Related Articles  [?]
Bibliography  [?]
External Links  [?]
Citable Version  [?]
 
This editable, developed Main Article is subject to a disclaimer.
This article is about the type of political entity. For other uses of the term State, please see State (disambiguation).

A state is a set of institutions that possess the authority to make the rules that govern the people in one or more societies. Although the term often refers broadly to all institutions of government or rule—ancient and modern—the term in its current meaning only came into broad circulation after the 16th century, and in a strict sense the word is only applied to the political systems that began to develop in western Europe in the 15th century.

Within a federal system, the term state also refers to political units, not entirely sovereign themselves, but subject to some extent to the authority of the larger state, or federal union, such as the "states" in the United States of America.

In casual usage, the terms "country," "nation," and "state" are often used as if they were synonymous. More strictly, however, we can distinguish them as follows:

  • country is the geographical area
  • nation designates a people, however national and international both confusingly refer as well to matters pertaining to what are strictly states, as in national capital, international law
  • state refers to set of governing institutions with sovereignty over a definite territory; it may be a subset of a nation. "State of X" is also used to distinguish a government from a geographical area

Etymology

The word "state" and its cognates in other European languages (it:stato, fr:État, de:Staat) ultimately derive from the Latin status, meaning "condition" or "status." With the revival of the Roman law in the 14th century in Europe, this Latin term was used to refer to the legal standing of persons, and in particular the special status of the king. The word was also associated with Roman ideas (dating back to Cicero) about the "status rei publicae", the "condition of the republic."[1]

Empirical and juridical aspects of the state

The word 'state' has both an empirical and a juridical sense. Entities can be states either de facto or de jure or both.[2]

Empirically (or /de facto/), an entity is a state if, as in Max Weber's influential definition, it is that organization that has a 'monopoly on legitimate violence' over a specific territory.[3] From this point of view, the state includes such institutions as the armed forces, civil service or state bureaucracy, courts, and the police.

Juridically (or /de jure/), an entity is a state in international law if it is recognized as such by other states, even if it does not actually have a monopoly on the legitimate use of force over a territory. Only an entity juridically recognized as a state can enter into international agreements. Conversely, an organization may have all the empirical attributes of a state, monopolizing the means of legitimate violence over a specified territory, and yet not be juridically recognized as a state (e.g., the Somali region of Somaliland).

States, government types, and political systems

It is important to distinguish between states and forms of government or regime types, such as democracy or dictatorship. The form of government identifies only one aspect of the state, namely, the way in which the highest political offices are filled and their relationship to each other and to society. It does not include other aspects of the state that may be very important in its everyday functioning, such as the quality of its bureaucracy. For example, two democratic states may be quite different if one has a capable, well-trained bureaucracy while the other does not. Thus generally speaking the term "state" refers to the instruments of political power, while the term regime or form of government refers more to the way in which such instruments can be accessed and employed.[4]

Some scholars have suggested that the term "state" is too imprecise and loaded to be used productively in sociology and political science, and ought to be substituted by the more comprehensive term "political system." The "political system" refers to the ensemble of all social structures that function to produce collectively binding decisions in a society. In modern times, these would include the political regime, political parties, and various sorts of organizations. The term "political system" thus denotes a broader concept than the state.[5]

Types of states

States can be classified according to a variety of criteria. Modern sociology and political science, following Max Weber, usually distinguish between traditional and modern states, depending on the kind of legitimation that the rulers have and their associated organizational forms.[6]

Traditional States

A traditional state, in Weber's view, is a state whose main principle of legitimation is found in tradition. In a traditional state, the ruler thus holds power by virtue of tradition. This means that political power is both restricted and personalized: the ruler's actions are constrained by tradition, but subjects owe obedience to his or her person, rather than to an impersonal order such as the law. Traditional states tend to have undeveloped bureaucracies (which require impersonal rules to function properly) and their staffs tend to appropriate, to greater or lesser degrees, the means of administration.

There is more than one kind of traditional state, and little consensus on the appropriate terminology to designate them. Agrarian empires, some city states and feudal states are some historical types. Weber himself distinguished between patriarchal, patrimonial, and feudal forms of the traditional state, and some scholars have used the idea of a neopatrimonial state to analyze contemporary states where power is personalized and there is a great deal of corruption.

Modern States

By contrast, a modern state is a state whose main principle of legitimation is legal-rational. In such states, loyalty is to an impersonal order (e.g., a constitution) and power is depersonalized. Organizationally, modern states are characterized by bureaucracies whose staffs do not own the means of administration. Actual modern states are not always sharply differentiated from traditional states, and they may have more than one principle of legitimation.

Other Classifications

Other classifications of states are possible. Michael Mann, for example, has proposed a classification of states based on two dimensions of political power: "infrastructural power" and "despotic power." These measure, respectively, the capacity of a society for centralized administration (manifested in its technical and organizational structures) and the ability of rulers to impose their will on society without prior negotiation with powerful groups. Variations in these four dimensions yield four basic types: "feudal" states (low in both counts), "imperial" states (low in infrastructural power, high in despotic power), "bureaucratic" states (high in infrastructural power, low in despotic power), and "authoritarian" states (high in both counts).[7]

Historical states

The earliest forms of the state emerged whenever it became possible to centralize power in a durable way. Agriculture and writing are almost everywhere associated with this process: agriculture because it allowed for the emergence of a class of people who did not have to spend most of their time providing for their own subsistence, and writing (or the equivalent of writing, like Inca quipus) because it made possible the centralization of vital information.[8]

The clear advantages of centralizing some activities (for example, the distribution of food in the absence of a money economy, as in the Inca empire, or the defense of a cultural area) probably made it possible for early states to take root. Even today, states can accumulate infrastructural power when it appears advantageous to much of society that certain activities be managed in a centralized way (e.g., coinage, monetary policy, defense). Centralization, however, creates resources that state elites can then use for their own purposes. From this point of view, the history of the state appears as a constant oscillation between functionally useful increases in centralization and the misappropriation by rulers of the resources such centralization generates.[7]

The state in classical antiquity

In classical antiquity, the state took a variety of forms, none of them very much like the modern state. There were monarchies whose power (like that of the Egyptian Pharaoh) was based on the religious function of the king and his control of a centralized army. There were also large, quasi-bureaucratized empires, like the Roman empire, which depended less on the religious function of the ruler and more on effective military and legal organizations and the cohesiveness of an Aristocracy.

Perhaps the most important political innovations of classical antiquity came from the Greek city-states and the Roman Republic. The Greek city-states before the 4th century granted citizenship rights to their free population, and in Athens these rights were combined with a directly democratic form of government that was to have a long afterlife in political thought and history.

In contrast, Rome did not introduce direct democracy but developed from a monarchy into a republic, governed by a senate dominated by the Roman aristocracy. The Roman political system contributed to the development of law, constitutionalism and to the distinction between the private and the public spheres.

From feudalism to the absolutist state

The story of the development of the modern state in the West typically begins with the dissolution of the western Roman empire. The fragmentation of the imperial state resulted in the devolution of power into the hands of private lords whose political, judicial, and military roles corresponded to the organization of economic production. In these conditions, according to Marxists, their estate—the basic economic unit of society—was the state.

The state-system of feudal Europe was an unstable configuration of suzerains and anointed kings. A monarch, formally at the head of a hierarchy of sovereigns, was not an absolute power who could rule at will; instead, relations between lords and monarchs were mediated by varying degrees of mutual dependence, which was ensured by the absence of a centralized system of taxation. This reality ensured that each ruler needed to obtain the 'consent' of each estate in the realm. Given the legal underpinnings of the feudal organization of society, and the Roman Catholic Church's claim to act as a law-making power coequal to rather than subordinate to secular authorities, the conception of the 'modern state' is not a basis for understanding politics in medieval feudalism. The formalization of these struggles over taxation thus gave rise to the Standestaat, or the state of Estates.[9]

In continental Europe the struggles between the king and the nobility eventually tilted in favor of the centralizing activities of the kings. As Europe's dynastic states—England under the Tudors, Spain under the Habsburgs, and France under the Bourbons—embarked on a variety of programs designed to increase centralized political and economic control, they increasingly exhibited many of the institutional features that characterize the "modern state." This centralization of power involved the delineation of political boundaries, as European monarchs gradually defeated or co-opted other sources of power, such as the Church and lesser nobility. In place of the fragmented system of feudal rule, with its often-indistinct territorial claims, large, unitary states with extensive control over definite territories emerged. This process gave rise to the highly centralized and increasingly bureaucratic forms of absolute monarchical rule of the 17th and 18th centuries, when the principal features of the contemporary state system took form, including the introduction of a standing army, a central taxation system, diplomatic relations with permanent embassies, and the development of state economic policy—mercantilism. These were the classical absolutist states.

The modern state

The rise of the "modern state" as a public power constituting the supreme political authority within a defined territory is associated with western Europe's gradual institutional development beginning in earnest in the late 15th century, culminating in the rise of absolutism and capitalism.

Cultural and national homogenization figured prominently in the rise of the modern state system. Since the absolutist period, states have largely been organized on a national basis. The concept of a national state, however, is not synonymous with the concept of a nation-state. Even in the most ethnically homogenous societies there is not always a complete correspondence between state and nation, hence the active role often taken by the state to promote nationalism through emphasis on shared symbols and national identity.[10]

It is in this period that the term "the state" is first introduced into political discourse in more or less its current meaning. Although Niccolò Machiavelli is often credited with first using the term to refer to a territorial sovereign government in the modern sense in The Prince, published in 1532, it is not until the time of the British thinkers Thomas Hobbes and John Locke that the concept in its current meaning is fully developed.[1]

State and civil society

Given the increasing institutional access to the state and role in the development of public policy by many parts of civil society, it is increasingly difficult to identify the boundaries of the state. The boundaries of the state are continually changing, for example, through privatization, nationalization, and the creation of new regulatory bodies. Often the nature of quasi-autonomy organizations is unclear, generating debate among political scientists on whether they are part of the state or civil society.

The distinction between the state to the public sphere or civil society has been the subject of considerable attention in analyses of state development has underwritten a diverse array of inquiries. Jürgen Habermas, in his conception of the public sphere, has defined civil society in terms of its role as a site of extra-institutional engagement with matters of public interest autonomous from the state. Earlier Western philosophers, in contrast, emphasized the supremacy of state over society, such as Thomas Hobbes and G.W.F. Hegel.

Some Marxist theorists, such as Antonio Gramsci, question the distinction between the state and civil society altogether, arguing that the former is integrated into many parts of the latter. Others, such as Louis Althusser, maintain that civil organizations such as church, schools, and even trade unions are part of an 'ideological state apparatus.' In this sense, the state can fund a number of groups within society that, while autonomous in principle, are dependent on state support.

The state in the international system

The state and supranationalism

International relations theorists have traditionally posited the existence of an international system, where states take into account the behavior of other states when making their own calculations. From this point of view, states embedded in an international system face internal and external security and legitimation dilemmas. Recently the notion of an 'international community' has been developed to refer to a group of states who have established rules, procedures, and institutions for the conduct of their relations. In this way the foundation has been laid for international law, diplomacy, formal regimes, and organizations.

In the late 20th century, globalization generated a debate as to whether the state can retain any of the freedom of action formerly associated with sovereignty. The rising mobility of labor and capital and the emergence of many international institutions appear to constrain the ability of states to set policy within their own borders. These constraints on the state's freedom of action are accompanied in some areas, notably Western Europe, with projects for interstate integration such as the European Union. However, the state remains the basic political unit of the world, as it has been since the 16th century.

The state and international law

The legal criteria for statehood are not obvious. Often, the laws are surpassed by political circumstances. However, one of the documents often quoted on the matter is the Montevideo Convention from 1933, the first article of which states:

The state as a person of international law should possess the following qualifications: (a) a permanent population; (b) a defined territory; (c) government; and (d) capacity to enter into relations with the other states.

States today

Since the late 19th century, the entirety of the world's inhabitable land has been parceled up into states; earlier, quite large land areas had been either unclaimed or uninhabited, or inhabited by nomadic peoples who were not organized as states. Currently more than 200 states comprise the international community, with the vast majority of them represented in the United Nations.

Contemporary theories of the state

Since Weber, an extensive literature on the processes by which the "modern state" emerged has been generated. Marxist scholars assert that the formation of states can be explained primarily in terms of the interests and struggles of social classes, while non-Marxist scholars—often in the tradition of Weber or Emile Durkheim—place greater emphasis on non-class actors. Similarly, some scholars have tended to understand the development of the modern state in terms of internal dynamics and conflicts, while others - drawing on a tradition going back to Weber and Otto Hintze- emphasize instead international processes such as war and imperialism.

Marxists generally argue that there is a discernible historical pattern in the emergence of capitalist states, in which the formation of national states in the West is related to the emergence of capitalism.[11] By contrast, scholars in the Weberian tradition have tended to emphasize the development of military power, though without necessarily dismissing the relationship of the modern state with the development of capitalism.[12][8]

There are three main traditions within political science that shape 'theories of the state': the Marxist, the pluralist, and the institutionalist. Each of these theories has been employed to gain understanding on the state, while recognizing its complexity. Several issues underlie this complexity. First, the boundaries of the state are not closely defined, but constantly changing. Second, the state is not only the site of conflict between different organizations, but also internal conflict and conflict within organizations. Some scholars speak of the 'state's interest,' but there are often various interests within different parts of the state that are neither solely state-centered nor solely society-centered, but develop between different groups in civil society and different state actors.

Marxism

For Marxist theorists, the role of modern states is determined or related to their position in capitalist societies. Many contemporary Marxists offer a liberal interpretation of Marx's comment in The Communist Manifesto that the state is but the executive committee for managing the common affairs of the whole bourgeoisie. Ralph Miliband argued that the ruling class uses the state as its instrument to dominate society by virtue of the interpersonal ties between state officials and economic elites. For Miliband, the state is dominated by an elite that comes from the same background as the capitalist class. State officials therefore share the same interests as owners of capital and are linked to them through a wide array of interpersonal and political ties.

By contrast, other Marxist theorists argue that the question of who controls the state is irrelevant. Heavily influenced by Gramsci, Nicos Poulantzas, a Greek neo-Marxist theorist argued that capitalist states do not always act on behalf of the ruling class, and when they do, it is not necessarily the case because state officials consciously strive to do so, but because the 'structural' position of the state is configured in such a way to ensure that the long-term interests of capital are always dominant. Poulantzas' main contribution to the Marxist literature on the state was the concept of 'relative autonomy' of the state. While Poulantzas' work on 'state autonomy' has served to sharpen and specify a great deal of Marxist literature on the state, his own framework came under criticism for its 'structural functionalism.'

Pluralism

While neo-Marxist theories of the state were relatively influential in continental Europe in the 1960s and 1970s, pluralism, a contending approach, gained greater adherence in the United States. Within the pluralist tradition, Robert Dahl sees the state as either a neutral arena for contending interests or its agencies as simply another set of interest groups. With power competitively arranged in society, state policy is a product of recurrent bargaining. Although pluralism recognizes the existence of inequality, it asserts that all groups have an opportunity to pressure the state. The pluralist approach suggests that the state's actions are the result of pressures applied for both polyarchy and organized interests.

Institutionalism

Both the Marxist and pluralist approaches view the state as reacting to the activities of groups within society, such as classes or interest groups. In this sense, they have both come under criticism for their 'society-centered' understanding of the state by scholars who emphasize the autonomy of the state with respect to social forces.

In particular, the "new institutionalism," an approach to politics that holds that behavior is fundamentally molded by the institutions in which it is embedded, asserts that the state is not an 'instrument' or an 'arena' and does not 'function' in the interests of a single class. Scholars working within this approach stress the importance of interposing civil society between the economy and the state to explain variation in state forms.

"New institutionalist" writings on the state, such as the works of Theda Skocpol, suggest that state actors are to an important degree autonomous. In other words, state personnel have interests of their own, which they can and do pursue independently (at times in conflict with) actors in society. Since the state controls the means of coercion, and given the dependence of many groups in civil society on the state for achieving any goals they may espouse, state personnel can to some extent impose their own preferences on civil society.[13]

'New institutionalist' writers, claiming allegiance to Weber, often utilize the distinction between 'strong states' and 'weak states,' claiming that the degree of 'relative autonomy' of the state from pressures in society determines the power of the state—a position that has found favor in the field of international political economy.

The state in modern political thought

Beginning with Thomas Hobbes, the question of the justification of the state becomes central to modern political thought. Indeed, the concept of the modern state emerged from the attempt to find a basis for the legitimate use of political power that was not connected with revealed religion. Thus the great absolutist thinkers such as Hobbes and Bodin were viewed with suspicion by other more orthodox figures such as Sir Robert Filmer insofar as they defended the absolute sovereignty of the king on "popular" grounds.

The development of a concept of the state that emphasized the use of coercion by a centralized authority but was nevertheless dissociated from any religious justification was theoretically problematic. Modern political thought thus often had recourse to state of nature reasoning. The state of nature is an imagined depiction of a condition in which there is no centralized, coercive power to enforce agreements. Depending on the assumptions one makes about such a state, one can then justify a more or less extensive "state of society" in which there is a centralized, coercive authority that can enforce agreements.

Thus, for example, Thomas Hobbes argued that conditions in the state of nature would be so horrifying (life would be "solitary, nasty, brutish, and short") that they would justify an all-powerful sovereign. John Locke, by contrast, argued that the state of nature imposed some moral constraints on the kind of state that is justified, and Jean Jacques Rousseau suggested that a true state of nature would be much better in some important ways than most existing states. Though state of nature reasoning was never the only form of thinking about the justification of the state in modern political thought (neither David Hume nor John Stuart Mill thought it necessary to engage in such an exercise) it still has some proponents (e.g., Robert Nozick).

State of nature reasoning is usually, but not always, connected to some "contractual" theory of the justification of the modern state. In these theories, states are justified when they are the result of a social contract, variously specified, that determines the rights and duties of individuals vis a vis political authority, and in a sense constitutes the state. Social contract theories usually emphasize the need for consent to legitimate state power, and in some cases (e.g., Rousseau) also stress the sovereignty of the people as the fundamental subject of the contract.

See also

References

  1. 1.0 1.1 Skinner, Quentin. 1989. The State. In Political Innovation and Conceptual Change, edited by T. Ball, J. Farr and R. L. Hanson. Cambridge: Cambridge University Press. ISBN 0521359783
  2. Jackson, Robert H., and Carl G. Rosberg. 1982. Why Africa's Weak States Persist: The Empirical and The Juridical in Statehood. World Politics 35 (1):1-24.[1]
  3. Weber, Max. 1994. The Profession and Vocation of Politics. In Political Writings. Cambridge: Cambridge University Press. ISBN 0521397197.
  4. Bobbio, Norberto. 1989. Democracy and Dictatorship: The Nature and Limits of State Power. Minneapolis: University of Minnesota Press. ISBN 0816618135.
  5. Easton, David. 1990. The Analysis of Political Structure. New York: Routledge.
  6. Weber, Max. 1978. Economy and Society. Translated by G. Roth and K. Wittich. 2 vols. Berkeley: University of California Press. ISBN 0520035003. See vol. 1, part I, chapter 3.
  7. 7.0 7.1 Mann, Michael. 1984. The Autonomous Power of the State. Archives europeénnes de sociologie 25:185-213. PDF
  8. 8.0 8.1 Giddens, Anthony. 1987. Contemporary Critique of Historical Materialism. 3 vols. Vol. II: The Nation-State and Violence. Cambridge: Polity Press. ISBN 0520060393. See chapter 2.
  9. Poggi, G. 1978. The Development of the Modern State: A Sociological Introduction. Stanford: Stanford University Press.
  10. Breuilly, John. 1982. Nationalism and the State. New York: St. Martin's Press.
  11. Miliband, Ralph. 1983. Class power and state power. London: Verso.
  12. Tilly, Charles. 1992. Coercion, Capital, and European States, AD 990-1992. Cambridge, Massachusetts: B. Blackwell.
  13. Rueschemeyer, Dietrich, Theda Skocpol, and Peter B. Evans, eds. 1985. Bringing the State Back In. Cambridge: Cambridge University Press.