geochronology

geochronology
geochronologic /jee'oh kron'l oj"ik/, geochronological, adj.geochronologist, n.
/jee'oh kreuh nol"euh jee/, n.
the chronology of the earth, as based on both absolute and relative methods of age determination.
[1890-95; GEO- + CHRONOLOGY]

* * *

Dating and interpretation of geologic events in the history of the Earth.

The classical technique of geochronology was stratigraphy, including faunal succession. Since the mid 20th century, radiometric dating has provided absolute age data to supplement the relative dates obtained from the fossil record. Radiometric dating is based on the principle that radioactive isotopes in geologic material decay at constant, known rates to daughter isotopes. See also carbon-14 dating.

* * *

Introduction

      field of scientific investigation concerned with determining the age and history (stratigraphy) of the Earth's rocks and rock assemblages. Such time determinations are made and the record of past geologic events is deciphered by studying the distribution and succession of rock strata, as well as the character of the fossil organisms preserved within the strata.

      The Earth's surface is a complex mosaic of exposures of different rock types that are assembled in an astonishing array of geometries and sequences. Individual rocks in the myriad of rock outcroppings (or in some instances shallow subsurface occurrences) contain certain materials or mineralogic information that can provide insight as to their “age.”

      For years investigators determined the relative ages of sedimentary rock strata (stratification) on the basis of their positions in an outcrop and their fossil content. According to a long-standing principle of the geosciences, that of superposition, the oldest layer within a sequence of strata is at the base and the layers are progressively younger with ascending order. The relative ages of the rock strata deduced in this manner can be corroborated and at times refined by the examination of the fossil forms present. The tracing and matching of the fossil content of separate rock outcrops (i.e., correlation) eventually enabled investigators to integrate rock sequences in many areas of the world and construct a relative geologic time scale.

      Scientific knowledge of the Earth's geologic history has advanced significantly since the development of radiometric dating, a method of age determination based on the principle that radioactive atoms in geologic materials decay at constant, known rates to daughter atoms. Radiometric dating has provided not only a means of numerically quantifying geologic time but also a tool for determining the age of various rocks that predate the appearance of life-forms.

Early views and discoveries
      Some estimates suggest that as much as 70 percent of all rocks outcropping from the Earth's surface are sedimentary. Preserved in these rocks is the complex record of the many transgressions and regressions of the sea, as well as the fossil remains or other indications of now extinct organisms and the petrified sands and gravels of ancient beaches, sand dunes, and rivers.

      Modern scientific understanding of the complicated story told by the rock record is rooted in the long history of observations and interpretations of natural phenomena extending back to the early Greek (ancient Greek civilization) scholars. Xenophanes of Colophon (560?–478? BC), for one, saw no difficulty in describing the various seashells and images of life-forms embedded in rocks as the remains of long-deceased organisms. In the correct spirit but for the wrong reasons, Herodotus (5th century BC) felt that the small discoidal nummulitic petrifactions (actually the fossils of ancient lime-secreting marine protozoans) found in limestones outcropping at al-Jīzah (Jīzah, Al-), Egypt, were the preserved remains of discarded lentils left behind by the builders of the pyramids.

      These early observations and interpretations represent the unstated origins of what was later to become a basic principle of uniformitarianism, the root of any attempt at linking the past (as preserved in the rock record) to the present. Loosely stated, the principle says that the various natural phenomena observed today must also have existed in the past (see below The emergence of modern geologic thought: Lyell's promulgation of uniformitarianism (geochronology)).

      Although quite varied opinions about the history and origins of life and of the Earth itself existed in the pre-Christian era, a divergence between Western and Eastern thought on the subject of natural history became more pronounced as a result of the extension of Christian (Christianity) dogma to the explanation of natural phenomena. Increasing constraints were placed upon the interpretation of nature in view of the teachings of the Bible. This required that the Earth be conceived of as a static, unchanging body, with a history that began in the not too distant past, perhaps as little as 6,000 years earlier, and an end, according to the scriptures, that was in the not too distant future. This biblical history of the Earth left little room for interpreting the Earth as a dynamic, changing system. Past catastrophes, particularly those that may have been responsible for altering the Earth's surface such as the great flood of Noah, were considered an artifact of the earliest formative history of the Earth. As such, they were considered unlikely to recur on what was thought to be an unchanging world.

      With the exception of a few prescient individuals such as Roger Bacon (c. 1220–92) and Leonardo da Vinci (1452–1519), no one stepped forward to champion an enlightened view of the natural history of the Earth until the mid-17th century. Leonardo seems to have been among the first of the Renaissance scholars to “rediscover” the uniformitarian dogma through his observations of fossil marine organisms and sediments exposed in the hills of northern Italy. He recognized that the marine organisms now found as fossils in rocks exposed in the Tuscan Hills (Tuscany) were simply ancient animals that lived in the region when it had been covered by the sea and were eventually buried by muds along the seafloor. He also recognized that the rivers of northern Italy, flowing south from the Alps and emptying into the sea, had done so for a very long time.

      In spite of this deductive approach to interpreting natural events and the possibility that they might be preserved and later observed as part of a rock outcropping, little or no attention was given to the history—namely, the sequence of events in their natural progression—that might be preserved in these same rocks.

The principle of superposition of rock strata
      In 1669 the Danish-born natural scientist Nicolaus Steno (Steno, Nicolaus) (née Niels Steensen) published his noted treatise De solido intra solidum naturaliter contento dissertationis prodromus (Eng. trans. The Prodromus of Nicolaus Steno's Dissertation Concerning a Solid Body Enclosed by Process of Nature Within a Solid). This seminal work laid the essential framework for the science of geology by showing in very simple fashion that the layered rocks of Tuscany exhibit sequential change—that they contain a record of past events. Following from this observation, Steno concluded that the Tuscan rocks demonstrated superpositional relationships: rocks deposited first lie at the bottom of a sequence, while those deposited later are at the top. This is the crux of what is now known as the principle of superposition. Steno put forth still another idea—that layered rocks were likely to be deposited horizontally. Therefore, even though the strata of Tuscany were (and still are) displayed in anything but simple geometries, Steno's elucidation of these fundamental principles relating to the formation of stratified rock made it possible to work out not only superpositional relationships within rock sequences but also the relative age of each layer.

      With the publication of the Prodromus and the ensuing widespread dissemination of Steno's ideas, other natural scientists of the latter part of the 17th and early 18th centuries applied them to their own work. The early English geologist John Strachey (Strachey, John), for example, produced in 1725 what may well have been the first modern geologic maps of rock strata. He also described the succession of strata associated with coal-bearing sedimentary rocks in Somersetshire, the same region of England where he had mapped the rock exposures.

Classification of stratified rocks
      In 1756 Johann Gottlob Lehmann (Lehmann, Johann Gottlob) of Germany reported on the succession of rocks in the southern part of his country and the Alps, measuring and describing their compositional and spatial variation. While making use of Steno's principle of superposition, Lehmann recognized the existence of three distinct rock assemblages: (1) a successionally lowest category, the Primary (Urgebirge), composed mainly of crystalline rocks, (2) an intermediate category, or the Secondary (Flötzgebirge), composed of layered or stratified rocks containing fossils, and (3) a final or successionally youngest sequence of alluvial and related unconsolidated sediments (Angeschwemmtgebirge) thought to represent the most recent record of the Earth's history.

      This threefold classification scheme was successfully applied with minor alterations to studies in other areas of Europe by three of Lehmann's contemporaries. In Italy, again in the Tuscan Hills in the vicinity of Florence, Giovanni Arduino (Arduino, Giovanni), regarded by many as the father of Italian geology, proposed a four-component rock succession. His Primary and Secondary divisions are roughly similar to Lehmann's Primary and Secondary categories. In addition, Arduino proposed another category, the Tertiary (Tertiary Period) division, to account for poorly consolidated though stratified fossil-bearing rocks that were superpositionally older than the (overlying) alluvium but distinct and separate from the hard (underlying) stratified rocks of the Secondary.

      In two separate publications, one that appeared in 1762 and the second in 1773, Georg Christian Füchsel (Füchsel, Georg Christian) also applied Lehmann's earlier concepts of superposition to another sequence of stratified rocks in southern Germany. While using upwards of nine separate categories of sedimentary rocks, Füchsel essentially identified discrete rock bodies of unique composition, lateral extent, and position within a rock succession. (These rock bodies would constitute formations in modern terminology.)

      Nearly 1,000 kilometres (620 miles) to the east, the German naturalist Peter Simon Pallas (Pallas, Peter Simon) was studying rock sequences exposed in the southern Urals of eastern Russia. His report of 1777 differentiated a threefold division of rock, essentially reiterating Lehmann's work by extension.

      Thus, by the latter part of the 18th century, the superpositional concept of rock strata had been firmly established through a number of independent investigations throughout Europe. Although Steno's principles were being widely applied, there remained to be answered a number of fundamental questions relating to the temporal and lateral relationships that seemed to exist among these disparate European sites. Were these various German, Italian, and Russian sites at which Lehmann's threefold rock succession was recognized contemporary? Did they record the same series of geologic events in the Earth's past? Were the various layers at each site similar to those of other sites? In short, was correlation among these various sites now possible?

The emergence of modern geologic thought
      Inherent in many of the assumptions underlying the early attempts at interpreting natural phenomena in the latter part of the 18th century was the ongoing controversy between the biblical view of Earth processes and history and a more direct approach based on what could be observed and understood from various physical relationships demonstrable in nature. A substantial amount of information about the compositional character of many rock sequences was beginning to accumulate at this time. Abraham Gottlob Werner (Werner, Abraham Gottlob), a scholar of wide repute and following from the School of Mining in Freiberg, Ger., was very successful in reaching a compromise between what could be said to be scientific “observation” and biblical “fact.” Werner's theory was that all rocks (including the sequences being identified in various parts of Europe at that time) and the Earth's topography were the direct result of either of two processes: (1) deposition in the primeval ocean, represented by the Noachian flood (his two “Universal,” or Primary, rock series), or (2) sculpturing and deposition during the retreat of this ocean from the land (his two “Partial,” or disintegrated, rock series). Werner's interpretation, which came to represent the so-called Neptunist conception of the Earth's beginnings, found widespread and nearly universal acceptance owing in large part to its theological appeal and to Werner's own personal charisma.

      One result of Werner's approach to rock classification was that each unique lithology in a succession implied its own unique time of formation during the Noachian flood and a universal distribution. As more and more comparisons were made of diverse rock outcroppings, it began to become apparent that Werner's interpretation did not “universally” apply. Thus arose an increasingly vocal challenge to the Neptunist theory.

James Hutton's (Hutton, James) recognition of the geologic cycle
      In the late 1780s the Scottish scientist James Hutton launched an attack on much of the geologic dogma that had its basis in either Werner's Neptunist approach or its corollary that the prevailing configuration of the Earth's surface is largely the result of past catastrophic (catastrophism) events which have no modern counterparts. Perhaps the quintessential spokesman for the application of the scientific method in solving problems presented in the complex world of natural history, Hutton took issue with the catastrophist and Neptunist approach to interpreting rock histories and instead used deductive reasoning to explain what he saw. By Hutton's account, the Earth could not be viewed as a simple, static world not currently undergoing change. Ample evidence from Hutton's Scotland provided the key to unraveling the often thought but still rarely stated premise that events occurring today at the Earth's surface—namely erosion, transportation and deposition of sediments, and volcanism—seem to have their counterparts preserved in the rocks. The rocks of the Scottish coast and the area around Edinburgh proved the catalyst for his argument that the Earth is indeed a dynamic, ever-changing system, subject to a sequence of recurrent cycles of erosion and deposition and of subsidence and uplift. Hutton's formulation of the principle of uniformitarianism, which holds that Earth processes occurring today had their counterparts in the ancient past, while not the first time that this general concept was articulated, was probably the most important geologic concept developed out of rational scientific thought of the 18th century. The publication of Hutton's two-volume Theory of the Earth in 1795 firmly established him as one of the founders of modern geologic thought.

      It was not easy for Hutton to popularize his ideas, however. The Theory of the Earth certainly did set the fundamental principles of geology on a firm basis, and several of Hutton's colleagues, notably John Playfair (Playfair, John) with his Illustrations of the Huttonian Theory of the Earth (1802), attempted to counter the entrenched Wernerian influence of the time. Nonetheless, another 30 years were to pass before Neptunist and catastrophist views of Earth history were finally replaced by those grounded in a uniformitarian approach.

      This gradual unseating of the Neptunist theory resulted from the accumulated evidence that increasingly called into question the applicability of Werner's Universal and Partial formations in describing various rock successions. Clearly, not all assignable rock types would fit into Werner's categories, either superpositionally in some local succession or as a unique occurrence at a given site. Also, it was becoming increasingly difficult to accept certain assertions of Werner that some rock types (e.g., basalt) are chemical precipitates from the primordial ocean. It was this latter observation that finally rendered the Neptunist theory unsustainable. Hutton observed that basaltic rocks exposed in the Salisbury Craigs, just on the outskirts of Edinburgh, seemed to have baked adjacent enclosing sediments lying both below and above the basalt. This simple observation indicated that the basalt was emplaced within the sedimentary succession while it was still sufficiently hot to have altered the sedimentary material. Clearly, basalt could not form in this way as a precipitate from the primordial ocean as Werner had claimed. Furthermore, the observations at Edinburgh indicated that the basalt intruded the sediments from below—in short, it came from the Earth's interior, a process in clear conflict with Neptunist theory.

      While explaining that basalt may be intrusive, the Salisbury Craigs observations did not fully satisfy the argument that some basalts are not intrusive. Perhaps the Neptunist approach had some validity? The resolution of this latter problem occurred at an area of recent volcanism in the Auvergne area of central France. Here, numerous cinder cones and fresh lava flows composed of basalt provided ample evidence that this rock type is the solidified remnant of material ejected from the Earth's interior, not a precipitate from the primordial ocean.

Lyell's (Lyell, Sir Charles, Baronet) promulgation of uniformitarianism
      Hutton's words were not lost on the entire scientific community. Charles Lyell, another Scottish geologist, was a principal proponent of Hutton's approach, emphasizing gradual change by means of known geologic processes. In his own observations on rock and faunal successions, Lyell was able to demonstrate the validity of Hutton's doctrine of uniformitarianism and its importance as one of the fundamental philosophies of the geologic sciences. Lyell, however, imposed some conditions on uniformitarianism that perhaps had not been intended by Hutton: he took a literal approach to interpreting the principle of uniformity in nature by assuming that all past events must have conformed to controls exerted by processes that behaved in the same manner as those processes behave today. No accommodation was made for past conditions that do not have modern counterparts. In short, volcanic eruptions, earthquakes, and other violent geologic events may indeed have occurred earlier in Earth history but no more frequently nor with greater intensity than today; accordingly, the surface features of the Earth are altered very gradually by a series of small changes rather than by occasional cataclysmic phenomena.

      Lyell's contribution enabled the doctrine of uniformitarianism to finally hold sway, even though it did impose for the time being a somewhat limiting condition on the uniformity principle. This, along with the increased recognition of the utility of fossils in interpreting rock successions, made it possible to begin addressing the question of the meaning of time in Earth history.

Determining the relationships of fossils with rock strata
The hypothesis of fossil succession in the work of Georges Cuvier
      During this period of confrontation between the proponents of Neptunism and uniformitarianism, there emerged evidence resulting from a lengthy and detailed study of the fossiliferous strata of the Paris Basin that rock successions were not necessarily complete records of past geologic events. In fact, significant breaks frequently occur in the superpositional record. These breaks affect not only the lithologic character of the succession but also the character of the fossils found in the various strata.

      An 1812 study by the French zoologist Georges Cuvier (Cuvier, Georges, Baron) was prescient in its recognition that fossils do in fact record events in Earth history and serve as more than just “follies” of nature. Cuvier's thesis, based on his analysis of the marine invertebrate and terrestrial vertebrate fauna of the Paris Basin, showed conclusively that many fossils, particularly those of terrestrial vertebrates, had no living counterparts. Indeed, they seemed to represent extinct forms, which, when viewed in the context of the succession of strata with which they were associated, constituted part of a record of biological succession punctuated by numerous extinctions. These, in turn, were followed by a seeming renewal of more advanced but related forms and were separated from each other by breaks in the associated rock record. Many of these breaks were characterized by coarser, even conglomeratic strata following a break, suggesting “catastrophic” events that may have contributed to the extinction of the biota. Whatever the actual cause, Cuvier felt that the evidence provided by the record of faunal succession in the Paris Basin could be interpreted by invoking recurring catastrophic geologic events, which in turn contributed to recurring massive faunal extinction, followed at a later time by biological renewal.

William Smith's (Smith, William) work with faunal sequence
      As Cuvier's theory of faunal succession was being considered, William Smith, a civil engineer from the south of England, was also coming to realize that certain fossils can be found consistently associated with certain strata. In the course of evaluating various natural rock outcroppings, quarries, canals, and mines during the early 1790s, Smith increasingly utilized the fossil content as well as the lithologic character of various rock strata to identify the successional position of different rocks, and he made use of this information to effect a correlation among various localities he had studied. The consistency of the relationships that Smith observed eventually led him to conclude that there is indeed faunal succession (faunal succession, law of) and that there appears to be a consistent progression of forms from more primitive to more advanced. As a result of this observation, Smith was able to begin what was to amount to a monumental effort at synthesizing all that was then known of the rock successions outcropping throughout parts of Great Britain. This effort culminated in the publication of his “Geologic Map of England, Wales and Part of Scotland” (1815), a rigorous treatment of diverse geologic information resulting from a thorough understanding of geologic principles, including those of original horizontality, superposition (lithologic, or rock, succession), and faunal succession. With this, it now became possible to assume within a reasonable degree of certainty that correlation could be made between and among widely separated areas. It also became apparent that many sites that had previously been classified according to the then-traditional views of Arduino, Füchsel, and Lehmann did not conform to the new successional concepts of Smith.

Early attempts at mapping and correlation
      The seminal work of Smith at clarifying various relationships in the interpretation of rock successions and their correlations elsewhere resulted in an intensive look at what the rock record and, in particular, what the fossil record had to say about past events in the long history of the Earth. A testimony to Smith's efforts in producing one of the first large-scale geologic maps of a region is its essential accuracy in portraying what is now known to be the geologic succession for the particular area of Britain covered.

      The application of the ideas of Lyell, Smith, Hutton, and others led to the recognition of lithologic and paleontologic successions of similar character from widely scattered areas. It also gave rise to the realization that many of these similar sequences could be correlated.

      The French biologist Jean-Baptiste de Monet, Chevalier de Lamarck (Lamarck, Jean-Baptiste), in particular, was able to demonstrate the similarity of fauna from a number of Cuvier's and Alexandre Brongniart's (Brongniart, Alexandre) collections of fossils from the Paris Basin with fossil fauna from the sub-Apennines of Italy and the London Basin. While based mainly on the collections of Cuvier and Brongniart, Lamarck's observations provided much more insight into the real significance of using fossils strictly for correlation purposes. Lamarck disagreed with Cuvier's interpretation of the meaning of faunal extinction and regeneration in stratigraphic successions. Not convinced that catastrophes caused massive and widespread disruption of the biota, Lamarck preferred to think of organisms and their distribution in time and space as responding to the distribution of favourable habitats. If confronted with the need to adapt to abrupt changes in local habitat—Cuvier's catastrophes—faunas must be able to change in order to survive. If not, they became extinct. Lamarck's approach, much like that of Hutton, stressed the continuity of processes and the continuum of the stratigraphic record. Moreover, his view that organisms respond to the conditions of their environment had important implications for the uniformitarian approach to interpreting Earth history.

      Once it was recognized that many of the rocks of the Paris Basin, London Basin, and parts of the Apennines apparently belonged to the same sequence by virtue of the similarity of their fossil content, Arduino's term Tertiary (proposed as part of his fourfold division of rock succession in the Tuscan Hills of Italy) began to be applied to all of these diverse locations. Further work by Lyell and Gérard-Paul Deshayes resulted in the term Tertiary being accepted as one of the fundamental divisions of geologic time.

The concepts of facies, stages, and zones
      During the latter half of the 18th and early 19th centuries, most of the research on the distribution of rock strata and their fossil content treated lithologic boundaries as events in time representing limits to strata that contain unique lithology and perhaps a unique fossil fauna, all of which are the result of unique geologic processes acting over a relatively brief period of time. Hutton recognized early on, however, that some variations occur in the sediments and fossils of a given stratigraphic unit and that such variations might be related to differences in depositional environments. He noted that processes such as erosion in the mountains of Scotland, transportation of sand and gravels in streams flowing from these mountains, and the deposition of these sediments could all be observed to be occurring concurrently. At a given time then, these diverse processes were all taking place at separate locations. As a consequence, different environments produce different sedimentary products and may harbour different organisms. This aspect of differing lithologic type or environmental or biological condition came to be known as facies. (It was Steno who had, in 1669, first used the term facies in reference to the condition or character of the Earth's surface at a particular time.)

      The significance of the facies concept for the analysis of geologic history became fully apparent with the findings of the Swiss geologist Amanz Gressly (Gressly, Amanz). While conducting survey work in the Jura Mountains in 1838, Gressly observed that rocks from a given position in a local stratigraphic succession frequently changed character as he traced them laterally. He attributed this lateral variation to lateral changes in the depositional environments responsible for producing the strata in question. Having no term to apply to the observed changes, he adopted the word facies. While Gressly employed the term specifically in the context of lithologic character, it is applied more broadly today. As now used, the facies concept has come to encompass other types of variation that may be encountered as one moves laterally (e.g., along outcroppings of rock strata exposed in stream valleys or mountain ridges) in a given rock succession. Lithologic facies, biological facies, and even environmental facies can be used to describe sequences of rocks of the same or different age having a particularly unique character.

Stages and zones
      The extensive review of the marine invertebrate fauna of the Paris Basin by Deshayes and Lyell not only made possible the formalization of the term Tertiary but also had a more far-reaching effect. The thousands of marine invertebrate fossils studied by Deshayes enabled Lyell to develop a number of subdivisions of the Tertiary of the Paris Basin based on the quantification of molluskan (mollusk) species count and duration. Lyell noted that of the various assemblages of marine mollusks found, those from rocks at the top of the succession contained a large number of species that were still extant in modern environments. Progressively older strata yielded fewer and fewer forms that had living counterparts, until at the base of the succession, a very small number of the total species present could be recognized as having modern counterparts. This fact allowed Lyell to consider subdividing the Tertiary of the Paris Basin into smaller increments, each of which could be defined according to some relative percentage of living species present in the strata. The subdivision resulted in the delineation of the Eocene (Eocene Epoch), Miocene (Miocene Epoch), and Pliocene epochs (Pliocene Epoch) in 1833. Later this scheme was refined to further divide the Pliocene into an Early and a Late Pliocene.

      Lyell's biostratigraphically defined concept of sequence, firmly rooted in concepts of faunal succession and superposition, was developed on mixed but stratigraphically controlled collections of fossils. It worked, but it did not address the faunal composition of the various Paris Basin strata other than in gross intervals—intervals that were as much lithologically as paleontologically (paleontology) defined. Alcide d'Orbigny (Orbigny, Alcide Dessalines d'), a French geologist, demonstrated correlational and superpositional uniqueness by utilizing paleontologically distinct intervals of strata defined solely on the basis of their fossil assemblages in his study of the French Jurassic (Jurassic Period) Terrains Jurassiques (1842). This departure from a lithologically based concept of paleontologic succession enabled d'Orbigny to define paleontologically unique stages. Each stage represented a unique period in time and formed the basis of later work that resulted in the further subdivision of d'Orbigny's original stages into 10 distinct stage assemblages. In spite of the work of Smith and to a lesser extent Lyell and others, d'Orbigny's approach was essentially that of a catastrophist. Stage boundaries were construed to represent unusual extrinsic geologic events, with significant implications for faunal continuity. The applicability of d'Orbigny's stages to areas outside of France had only limited success. At this point in the development of paleontology as a science, little was understood about the geologic time range of various fauna. Even less was known about the habitats—the environmental limits—of ancient fauna. Could certain groups of organisms have sufficiently widespread distribution in the rock record to enable correlations to be made with certainty? The Jurassic of western Europe consisted mostly of shallow marine sediments widely deposited throughout the area. It is now known that some of the mollusks with which d'Orbigny worked were undergoing very rapid evolutionary change; they were thus relatively short-lived as distinct forms in the geologic record and had a wide-ranging environmental tolerance. The result was that some forms, notably of the group of mollusks called ammonite (ammonoid) cephalopods (cephalopod), were distributed extensively within a variety of sedimentary facies. The correlating of strata based on the faunal stage approach was widely accepted. Interestingly, most of d'Orbigny's Jurassic stages, with refinements, are still in use today.

      Only a short time after d'Orbigny's original analysis of Jurassic strata, the German mineralogist and paleontologist Friedrich A. Quenstedt (Quenstedt, Friedrich August) challenged (in 1856–58) the validity of using stages to effect correlations in cases where the actual geologic ranges and bed-by-bed distribution of individual component fossils of an assemblage were unknown. In retrospect, this seems blatantly obvious, but at the time the systematic stratigraphic documentation of fossil occurrence was not always carried out. Much critical biostratigraphic data necessary for the proper characterization of faunal assemblages was simply not collected. As argued, individual fossil ranges and their distributions could have profound influence on the concept of faunal succession and evolutionary dynamics.

      Several of Quenstedt's students at the University of Tübingen followed up on this latter concern. One in particular, Carl Albert Oppel (Oppel, Albert), essentially refined his mentor's concepts by paying particular attention to the character of the range of individual species in a succession of fauna. These intervals of unique biological character, which he called zones (biozone), were essentially subdivisions of the stages proposed by Quenstedt. Oppel's recognition of the earliest occurrence of a fossil species (or its first appearance), its range through a succession of strata, and its eventual loss from the local record (or its last appearance) led him to compare such biostratigraphic data from many species. By making use of such data on species that overlap in some or all of their stratigraphic ranges and from widely separated areas, Oppel was able to erect a biochronology based on a diverse record of first appearances, last appearances, and individual and overlapping range zones. This fine-scale refinement of a biologically defined sense of succession found wide applicability and enabled not only biochronological (or temporal) but also biofacies (spatial) understanding of the succession in question.

      Six of the Jurassic zones and some of the fossil species on which they are based are illustrated in the figure. The zone of Arietites bucklandi is based on the joint occurrences of two species with differing ranges in time, here called “species J” and “G,” of the Arnioceras semicostatum zone by overlap in the range of “species G” and “Q,” and so on. The zonal fossils are characteristic but are not necessarily confined in their occurrence to the zone.

V. Standish Mallory Gary Dean Johnson

Completion of the Phanerozoic (Phanerozoic Eon) time scale
      With the development of the basic principles of faunal succession and correlation and the recognition of facies variability, it was a relatively short step before large areas of Europe began to be placed in the context of a global geologic succession. This was not, however, accomplished in a systematic manner. Whereas the historical ideas of Lehmann and Arduino were generally accepted, it became increasingly clear that many diverse locally defined rock successions existed, each with its own unique fauna and apparent position within some sort of “universal” succession.

      As discussed above, Arduino's Tertiary was recognized in certain areas and was in fairly common use after 1760, but only rudimentary knowledge of other rock successions existed by the later part of the 18th century. The German naturalist Alexander von Humboldt (Humboldt, Alexander von) had recognized the widespread occurrence of fossil-bearing limestones (limestone) throughout Europe. Particular to these limestones, which formed large tracts of the Jura Mountains of Switzerland, were certain fossils that closely resembled those known from the Lias and Oolite formations of England, which were then being described by William Smith. Subsequently, Humboldt's “Jura Kalkstein” succession, as he described it in 1795, came to be recognized throughout Europe and England. By 1839, when the geologist Leopold Buch (Buch, Leopold, Freiherr von) recognized this rock sequence in southern Germany, the conceptual development of the Jurassic System was complete.

      The coal-bearing strata of England, known as the Coal Measures, had been exploited for centuries, and their distribution and vertical and lateral variability were the subject of numerous local studies throughout the 17th and early 18th centuries, including those of Smith. In 1808 the geologist Jean-Baptiste-Julien d'Omalius d'Halloy (Omalius d'Halloy, Jean-Baptiste-Julien d') described a coal-bearing sequence in Belgium as belonging to the Terrain Bituminifére. Although the name did not remain in common usage for long, the Terrain Bituminifére found analogous application in the work of two English geologists, William D. Conybeare (Conybeare, William Daniel) and William Phillips, in their synthesis of the geology of England and Wales in 1822. Conybeare and Phillips coined the term Carboniferous (Carboniferous Period) (or coal-bearing) to apply to the succession of rocks from north-central England that contained the Coal Measures. The unit also included several underlying rock formations extending down into what investigators now consider part of the underlying Devonian System. At the time, however, the approach by Conybeare and Phillips was to encompass in their definition of the Carboniferous all of the associated strata that could be reasonably included in the Coal Measures succession.

      D'Omalius mapped and described a local succession in western France. While doing so, he began to recognize a common sequence of soft limestones, greensands (glauconite-bearing sandstones), and related marls (marl) in what is today known to be a widespread distribution along coastal regions bordering the North Sea and certain regions of the Baltic. The dominant lithology of this sequence is frequently the soft limestones or chalk beds so well known from the Dover region of southeast England and Calais in nearby France. D'Omalius called this marl, greensand, and chalk-bearing interval the Terrain Crétacé. Along with their adoption of the term Carboniferous in 1822, Conybeare and Phillips referred to the French Terrain Crétacé as the Cretaceous System (Cretaceous Period).

      Clearly, surficial deposits and related unconsolidated material, variously relegated to the categories of classification proposed by Arduino, Lehmann, Werner, and others as “ alluvium” or related formations, deserved a place in any formalized system of rock succession. In 1829 Jules Desnoyers of France, studying sediments in the Seine valley, proposed using the term Quaternary to encompass all of these various post-Tertiary formations. At nearly the same time, the important work of Lyell on the faunal succession of the Paris Basin permitted finer-scaled discrimination of this classic Tertiary sequence. In 1833 Lyell, using various biostratigraphic evidence, proposed several divisions of the Tertiary System that included the Eocene, Miocene, and Pliocene epochs. By 1839 he proposed using the term Pleistocene (Pleistocene Epoch) instead of dividing his Pliocene Epoch into older and newer phases. The temporal subdivision of the Tertiary was completed by two German scientists, Heinrich Ernst Beyrich and Wilhelm Philipp Schimper. Beyrich introduced the Oligocene (Oligocene Epoch) in 1854 after having investigated outcrops in Belgium and Germany, while Schimper proposed adding the Paleocene (Paleocene Epoch) in 1874 based on his studies of Paris Basin flora.

      Werner's quadripartite division of rocks in southern Germany was applied well into the second decade of the 19th century. During this time, rock sequences from the lower part of his third temporal subdivision, the Flötzgebirge, were subsequently subdivided into three formations, each having fairly widespread exposure and distribution. Based on his earlier work, Friedrich August von Alberti identified in 1834 these three distinct lithostratigraphic units, the Bunter Sandstone, the Muschelkalk Limestone, and the Keuper Marls and Clays, as constituting the Trias or Triassic System (Triassic Period).

      Perhaps one of the most intriguing episodes in the development of the geologic time scale concerns the efforts of two British geologists and in large measure their attempts at unraveling the complex geologic history of Wales. Adam Sedgwick (Sedgwick, Adam) and Roderick Impey Murchison (Murchison, Sir Roderick Impey) began working, in 1831, on the sequence of rocks lying beneath the Old Red Sandstone (which had been included in the basal sequence of the Carboniferous, as defined by Conybeare and Phillips, earlier in 1822). What started as an earnest collaborative attempt at deciphering the structurally and stratigraphically complicated rock succession in Wales ended in 1835 with a presentation outlining two distinct subdivisions of the pre-Carboniferous succession. Working up from the base of the post-Primary rock succession of poorly fossiliferous clastic rocks in northern Wales, Sedgwick identified a sequence of rock units defined primarily by their various lithologies. He designated this succession the Cambrian (Cambrian Period), after Cambria, the Roman name for Wales. Murchison worked downward in the considerably more fossiliferous pre-Old Red Sandstone rock sequence in southern Wales and was able to identify a succession of strata containing a well-preserved fossil fauna. These sequences defined from southern Wales were eventually brought into the context of Sedgwick's Cambrian. Murchison named his rock succession the Silurian (Silurian Period), after the Roman name for an early Welsh tribe. In a relatively short time, Murchison's Silurian was expanding both laterally and temporally as more and more localities containing the characteristic Silurian fauna were recognized throughout Europe. The major problem created by this conceptual “expansion” of the Silurian was that it came to be recognized in northern Wales as coincident with much of the strata in the upper portion of Sedgwick's Cambrian. With Sedgwick's Cambrian based mainly on lithologic criteria, the presence of Silurian fauna created correlational difficulties. As it turned out, Sedgwick's Cambrian was of little value outside of its area of original definition. With it being superseded by the paleontologically based concept of the Silurian, some sort of compromise had to be worked out.

      This compromise came about primarily as a result of the work of Charles Lapworth (Lapworth, Charles), the English geologist who in 1879 proposed the designation Ordovician System for that sequence of rocks representing the upper part of Sedgwick's Cambrian succession and the lower (and generally overlapping) portion of Murchison's Silurian succession. The term Ordovician (Ordovician Period) is derived from yet another Roman-named tribe of ancient Wales, the Ordovices. A large part of Lapworth's rationale for this division was based on the earlier work of the French-born geologist Joachim Barrande (Barrande, Joachim), who investigated the apparent Silurian fauna of central Bohemia. Barrande's 1851 treatise on this area of Czechoslovakia demonstrated a distinct succession from a “second” Silurian fauna to a “third” Silurian fauna. This divisible Silurian, as well as separate lines of evidence gathered by Lapworth in Scotland and Wales, finally enabled the individual character of the Cambrian, Ordovician, and Silurian systems to be resolved.

      While involved in their work on Welsh stratigraphic successions, Sedgwick and Murchison had the opportunity to compare some rock outcroppings in Devonshire, in southwest England, with similar rocks in Wales. The Devon rocks were originally thought to belong to part of Sedgwick's Cambrian System, but they contained plant fossils very similar to basal Carboniferous (Old Red Sandstone) plant fossils found elsewhere. Eventually recognizing that these fossil-bearing sequences represented lateral equivalents in time and perhaps temporally unique strata as well, Sedgwick and Murchison in 1839 proposed the Devonian System (Devonian Period).

      During the early 1840s, Murchison traveled with the French paleontologist Edouard de Verneuil and the Latvian-born geologist Alexandr Keyserling to study the rock succession of the eastern Russian (Russia) platform, the area of Russia west of the Ural Mountains. Near the town of Perm, Murchison and Verneuil identified fossiliferous strata containing both Carboniferous and a younger fauna at that time not recognized elsewhere in Europe or in the British Isles. Whereas the Carboniferous fossils were similar to those they had seen elsewhere (mainly from the Coal Measures), the stratigraphically higher fauna appeared somewhat transitional to the Triassic succession of Germany as then understood. Murchison coined the term Permian (Permian Period) (after the town of Perm) to represent this intermediate succession.

      With continued refinement of the definition of the Carboniferous in Europe, particularly in England, what at one time comprised the Old Red Sandstone, Lower Coal Measures (Mountain Limestone and Millstone Grit), and Upper Coal Measures now stood as just the Lower and Upper Coal Measures. It was beginning to be recognized that certain rock sequences in the Catskill Mountains of eastern New York state in North America resembled the Old Red Sandstone of western England. Furthermore, coal-bearing strata exposed in Pennsylvania greatly resembled the similar coal-bearing strata of the Upper Coal Measures. Lying beneath these coal-bearing rocks of Pennsylvania was a sequence of limestones that could be traced over thousands of square kilometres and that occurred in numerous outcrops along various tributary streams to the Ohio and Mississippi rivers in Indiana, Kentucky, Missouri, Illinois, and Iowa. This “subcarboniferous” strata, identified by the American geologist David Dale Owen in 1839, was subsequently termed Mississippian (Mississippian Subperiod) in 1870 as a result of work conducted by another American geologist, Alexander Winchell, in the upper Mississippi valley area. Eventually the overlying strata, the coal-bearing rocks originally described from Pennsylvania, were formalized as Pennsylvanian (Pennsylvanian Subperiod) in 1891 by the paleontologist and stratigrapher Henry Shaler Williams.

      The North American-defined Mississippian and Pennsylvanian systems were later correlated with presumed European and British successions. Although approximately similar in successional relationship, the Mississippian–Pennsylvanian boundary in North America is now considered slightly younger than the Lower–Upper Carboniferous boundary in Europe.

      By the 1850s, with the development of the geologic time scale nearly complete, investigators were beginning to recognize that a number of major paleontologically defined boundaries were common and recurrent regardless of where a succession was studied. By this time rock successions were being defined according to fauna they contained, and the relative time scale, which was being erected, was based on the principle of faunal succession; consequently, any major hiatus or change in faunal character was bound to be interpreted as important. In 1838 Sedgwick proposed that all pre-Old Red Sandstone sediments be included in the rock succession designated the Paleozoic (Paleozoic Era) Series (or Era) that contained generally primitive fossil fauna. John Phillips, another English geologist, went on to describe the Mesozoic Era to accommodate what then was the Cretaceous, Jurassic, Triassic, and partially Permian strata, and the Kainozoic (Cainozoic, or Cenozoic (Cenozoic Era)) era to include Lyell's Eocene, Miocene, and Pliocene. This subdivision of the generally fossiliferous strata that lay superpositionally above the so-called Primary rocks of many of the early workers resulted in the recognition of three distinct eras. Subsequent subdivision of these eras into specific geologic periods finally provided the hierarchy for describing the relative dating of geologic events.

Development of radioactive dating methods and their application
      As has been seen, the geologic time scale is based on stratified rock assemblages that contain a fossil record. For the most part, these fossils allow various forms of information from the rock succession to be viewed in terms of their relative position in the sequence. Approximately the first 87 percent of Earth history occurred before the evolutionary development of shell-bearing organisms. The result of this mineralogic control on the preservability of organic remains in the rock record is that the geologic time scale—essentially a measure of biologic changes through time—takes in only the last 13 percent of Earth history. Although the span of time preceding the Cambrian period—the Precambrian (Precambrian time)—is nearly devoid of characteristic fossil remains and coincides with some of the primary rocks of certain early workers, it must, nevertheless, be evaluated in its temporal context.

Early attempts at calculating the age of the Earth
      Historically, the subdivision of Precambrian rock sequences (and, therefore, Precambrian time) had been accomplished on the basis of structural or lithologic grounds. With only minor indications of fossil occurrence (mainly in the form of algal stromatolites), no effective method of quantifying this loosely constructed chronology existed until the discovery of radioactivity enabled dating procedures to be applied directly to the rocks in question.

      The quantification of geologic time remained an elusive matter for most human enquiry into the age of the Earth and its complex physical and biological history. Although Hindu teachings accept a very ancient origin for the Earth, medieval Western concepts of Earth history were based for the most part on a literal interpretation of Old Testament references. Biblical scholars of Renaissance Europe and later considered paternity as a viable method by which the age of the Earth since its creation could be determined. A number of attempts at using the “begat” method of determining the antiquity of an event—essentially counting backward in time through each documented human generation—led to the age of the Earth being calculated at several thousand years. One such attempt was made by Archbishop James Ussher (Ussher, James) of Ireland, who in 1650 determined that the Creation had occurred during the evening of Oct. 22, 4004 BC. By his analysis of biblical genealogies, the Earth was not even 6,000 years old!

      From the time of Hutton's refinement of uniformitarianism, the principle found wide application in various attempts to calculate the age of the Earth. As previously noted, fundamental to the principle was the premise that various Earth processes of the past operated in much the same way as those processes operate today. The corollary to this was that the rates of the various ancient processes could be considered the same as those of the present day. Therefore, it should be possible to calculate the age of the Earth on the basis of the accumulated record of some process that has occurred at this determinable rate since the Creation.

      Many independent estimates of the age of the Earth have been proposed, each made using a different method of analysis. Some such estimates were based on assumptions concerning the rate at which dissolved salts or sediments are carried by rivers, supplied to the world's oceans, and allowed to accumulate over time. These chemical and physical arguments (or a combination of both) were all flawed to varying degrees because of an incomplete understanding of the processes involved. The notion that all of the salts dissolved in the oceans (ocean) were the products of leaching from the land was first proposed by the English astronomer and mathematician Edmond Halley (Halley, Edmond) in 1691 and restated by the Irish geologist John Joly (Joly, John) in 1899. It was assumed that the ocean was a closed system and that the salinity of the oceans was an ever-changing and ever-increasing condition. Based on these calculations, Joly proposed that the Earth had consolidated and that the oceans had been created between 80 and 90 million years ago. The subsequent recognition that the ocean is not closed and that a continual loss of salts occurs due to sedimentation in certain environments severely limited this novel approach.

      Equally novel but similarly flawed was the assumption that, if a cumulative measure of all rock successions were compiled and known rates of sediment accumulation were considered, the amount of time elapsed could be calculated. While representing a reasonable approach to the problem, this procedure did not or could not take into account different accumulation rates associated with different environments or the fact that there are many breaks in the stratigraphic record. Even observations made on faunal succession proved that gaps in the record do occur. How long were these gaps? Do they represent periods of nondeposition or periods of deposition followed by periods of erosion? Clearly sufficient variability in a given stratigraphic record exists such that it may be virtually impossible to even come to an approximate estimate of the Earth's age based on this technique. Nevertheless, many attempts using this approach were made.

      William Thomson (later Lord Kelvin (Kelvin, William Thomson, Baron)) applied his thermodynamic principles to the problems of heat flow, and this had implications for predicting the age of a cooling Sun and of a cooling Earth. From an initial estimate of 100 million years for the development of a solid crust around a molten core proposed in 1862, Thomson subsequently revised his estimate of the age of the Earth downward. Using the same criteria, he concluded in 1899 that the Earth was between 20 and 40 million years old.

      Thomson's calculation was based on the assumption that the substance of the Earth is inert and thus incapable of producing new heat. His estimate came into question after the discovery of naturally occurring radioactivity by the French physicist Henri Becquerel in 1896 and the subsequent recognition by his colleagues, Marie and Pierre Curie, that compounds of radium (which occur in uranium minerals) produce heat. As a result of this and other findings, notably that of Ernest Rutherford (Rutherford, Ernest, Baron Rutherford of Nelson, of Cambridge) (see below), it became apparent that naturally occurring radioactive elements in minerals common in the Earth's crust are sufficient to account for all observed heat flow. Within a short time another leading British physicist, John William Strutt (Rayleigh, John William Strutt, 3rd Baron), concluded that the production of heat in the Earth's interior was a dynamic process, one in which heat was continuously provided by such materials as uranium. The Earth was, in effect, not cooling.

An absolute age framework for the stratigraphic time scale
      In his book Radio-activity (1904), Rutherford explained that radioactivity results from the spontaneous disintegration of an unstable element into a lighter element, which may decay further until a stable element is finally created. This process of radioactive decay involves the emission of positively charged particles (later to be recognized as helium nuclei) and negatively charged ones (electrons) and in most cases gamma rays (a form of electromagnetic radiation) as well. This interpretation, the so-called disintegration theory, came to provide the basis for the numerical quantification of geologic time.

      In 1905 Strutt succeeded in analyzing the helium content of a radium-containing rock and determined its age to be 2 billion years. This was the first successful application of a radiometric technique to the study of Earth materials, and it set the stage for a more complete analysis of geologic time. Although faced with problems of helium loss and therefore not quite accurate results, a major scientific breakthrough had been accomplished. Also in 1905 the American chemist Bertram B. Boltwood (Boltwood, Bertram Borden), working with the more stable uranium–lead system, calculated the numerical ages of 43 minerals. His results, with a range of 400 million to 2.2 billion years, were an order of magnitude greater than those of the other “quantitative” techniques of the day that made use of heat flow or sedimentation rates to estimate time.

      Acceptance of these new ages was slow in coming. Perhaps much to their relief, paleontologists now had sufficient time in which to accommodate faunal change. Researchers in other fields, however, were still conservatively sticking with ages on the order of several hundred million, but were revising their assumed sedimentation rates downward in order to make room for expanded time concepts.

      In a brilliant contribution to resolving the controversy over the age of the Earth, Arthur Holmes, a student of Strutt, compared the relative (paleontologically determined) stratigraphic (stratigraphy) ages of certain specimens with their numerical ages as determined in the laboratory. This 1911 analysis provided for the first time the numerical ages for rocks from several Paleozoic geologic periods as well as from the Precambrian. Carboniferous-aged material was determined to be 340 million years, Devonian-aged material 370 million years, Ordovician (or Silurian) material 430 million years, and Precambrian specimens from 1.025 to 1.64 billion years. As a result of this work, the relative geologic time scale, which had taken nearly 200 years to evolve, could be numerically quantified. No longer did it have merely superpositional significance, it now had absolute temporal significance as well.

Gary Dean Johnson

Nonradiometric dating
      In addition to radioactive decay, many other processes have been investigated for their potential usefulness in absolute dating. Unfortunately, they all occur at rates that lack the universal consistency of radioactive decay. Sometimes human observation can be maintained long enough to measure present rates of change, but it is not at all certain on a priori grounds whether such rates are representative of the past. This is where radioactive methods frequently supply information that may serve to calibrate nonradioactive processes so that they become useful chronometers. Nonradioactive absolute chronometers may conveniently be classified in terms of the broad areas in which changes occur—namely, geologic and biological processes, which will be treated here.

Geologic processes as absolute chronometers
weathering processes
      During the first third of the 20th century, several presently obsolete weathering chronometers were explored. Most famous was the attempt to estimate the duration of Pleistocene interglacial intervals through depths of soil development. In the American Midwest, thicknesses of gumbotil and carbonate-leached zones were measured in the glacial deposits (tills (till)) laid down during each of the four glacial stages. Based on a direct proportion between thickness and time, the three interglacial intervals were determined to be longer than postglacial time by factors of 3, 6, and 8. To convert these relative factors into absolute ages required an estimate in years of the length of postglacial time. When certain evidence suggested 25,000 years to be an appropriate figure, factors became years—namely, 75,000, 150,000, and 200,000 years. And, if glacial time and nonglacial time are assumed approximately equal, the Pleistocene Epoch lasted about 1,000,000 years.

      Only one weathering chronometer is employed widely at the present time. Its record of time is the thin hydration layer at the surface of obsidian artifacts. Although no hydration layer appears on artifacts of the more common flint and chalcedony, obsidian is sufficiently widespread that the method has broad application.

      In a specific environment the process of obsidian hydration is theoretically described by the equation D = Kt1/2, in which D is thickness of the hydration rim, K is a constant characteristic of the environment, and t is the time since the surface examined was freshly exposed. This relationship is confirmed both by laboratory experiments at 100° C (212° F) and by rim measurements on obsidian artifacts found in carbon-14 dated sequences. Practical experience indicates that the constant K is almost totally dependent on temperature and that humidity is apparently of no significance. Whether in a dry Egyptian tomb or buried in wet tropical soil, a piece of obsidian seemingly has a surface that is saturated with a molecular film of water. Consequently, the key to absolute dating of obsidian is to evaluate K for different temperatures. Ages follow from the above equation provided there is accurate knowledge of a sample's temperature history. Even without such knowledge, hydration rims are useful for relative dating within a region of uniform climate.

      Like most absolute chronometers, obsidian dating has its problems and limitations. Specimens that have been exposed to fire or to severe abrasion must be avoided. Furthermore, artifacts reused repeatedly do not give ages corresponding to the culture layer in which they were found but instead to an earlier time, when they were fashioned. Finally, there is the problem that layers may flake off beyond 40 micrometres (0.004 centimetre, or 0.002 inch) of thickness—i.e., more than 50,000 years in age. Measuring several slices from the same specimen is wise in this regard, and such a procedure is recommended regardless of age.

Accumulational processes
      Sediment in former or present water bodies, salt dissolved in the ocean, and fluorine in bones are three kinds of natural accumulations and possible time indicators. To serve as geochronometers, the records must be complete and the accumulation rates known.

      The fossiliferous part of the geologic column includes perhaps 122,000 metres of sedimentary rock if maximum thicknesses are selected from throughout the world. During the late 1800s, attempts were made to estimate the time over which it formed by assuming an average rate of sedimentation. Because there was great diversity among the rates assumed, the range of estimates was also large—from a high of 2.4 billion years to a low of 3 million years. In spite of this tremendous spread, most geologists felt that time in the hundreds of millions of years was necessary to explain the sedimentary record.

      If the geologic column (see below) were made up entirely of annual layers, its duration would be easy to determine. Limited sedimentary deposits did accumulate in this way, and they are said to be varved; one year's worth of sediment is called a varve, and, in general, it includes two laminae per year.

      Varves arise in response to seasonal changes. New Mexico's Castile Formation, for example, consists of alternating layers of gypsum and calcite that may reflect an annual temperature cycle in the hypersaline water from which the minerals precipitated. In moist, temperate climates, lake sediments collecting in the summer are richer in organic matter than those that settle during winter. This feature is beautifully seen in the seasonal progression of plant microfossils found in shales at Oensingen, Switz. In the thick oil shales of Wyoming and Colorado in the United States, the flora is not so well defined, but layers alternating in organic richness seem to communicate the same seasonal cycle. These so-called Green River Shales also contain abundant freshwater-fish fossils that confirm deposition in a lake. At their thickest, they span 792 vertical metres. Because the average thickness of a varve is about 0.015 centimetre (0.006 inch), the lake is thought to have existed for more than 5 million years.

      Each of the examples cited above is of a floating chronology—i.e., a decipherable record of time that was terminated long ago. In Sweden, by contrast, it has been possible to tie a glacial varve chronology to present time, and so create a truly absolute dating technique. Where comparisons with radiocarbon dating are possible, there is general agreement.

      As early as 1844, an English chemist named Middleton claimed that fossil bones contain fluorine in proportion to their antiquity. This idea is sound in principle, provided that all the other natural variables remain constant. Soil permeability, rainfall, temperature, and the concentration of fluorine in groundwater all vary with time and location, however. Fluorine dating is therefore not the simple procedure that Middleton envisioned.

      Still, the idea that hydroxyapatite (hydroxylapatite) in buried bone undergoes gradual change to fluorapatite is a correct one. In a restricted locality where there is uniformity of climate and soil, the extent of fluorine addition is at least a measure of relative age and has been so used with notable success in dating certain hominid remains. Both the Piltdown hoax, for example, and the intrusive burial of the Galley Hill skeleton were exposed in part by fluorine measurements. Supplementing them were analyses of uranium, which resembles fluorine in its increase with time, and nitrogen, which decreases as bone protein decays away.

      Fluorine changes could conceivably be calibrated if bone samples were found in a radiometrically dated sequence. Conditions governing fluorine uptake, however, are so variable even over short distances that it is risky to use fluorine content as an absolute chronometer much beyond the calibration site itself. In short, fluorine dating is not now and probably never will be an absolute chronometer. Even when used in relative dating, many fluorine analyses on diverse samples are needed, and these must be supplemented by uranium and nitrogen measurements to establish confidence in the chronological conclusions.

      Based on three centuries of direct measurement, the Earth's magnetic field is known to be varying slowly in both its intensity and direction. In fact, change seems to have been the rule throughout all of the Earth's past. Magnetic minerals in rocks (and in articles of fired clay) provide the record of ancient change, for they took on the magnetic field existing at the time of their creation or emplacement.

      Polar (polar wandering) reversals were originally discovered in lava rocks and since have been noted in deep-sea cores. In both cases the time dimension is added through radiometric methods applied to the same materials that show the reversals. Potassium–argon is the commonest chronometer used. A magnetic-polarity (or paleomagnetic) time scale has been proposed along the line of the geologic time scale; time divisions are called intervals, or epochs.

Biological processes as absolute chronometers
Tree-ring growth
      In the early 1900s an American astronomer named Andrew E. Douglass (Douglass, Andrew Ellicott) went looking for terrestrial records of past sunspot cycles and not only found what he sought but also discovered a useful dating method in the process. The focus of his attention was the growth rings (growth ring) in trees (tree)—living trees, dead trees, beams in ancient structures, and even large lumps of charcoal.

      The key documents for tree-ring dating, or dendrochronology, are those trees that grow or grew where roots receive water in direct proportion to precipitation. Under such a situation, the annual tree rings vary in width as a direct reflection of the moisture supplied. What is important in tree-ring dating is the sequence in which rings vary. Suppose, for example, that a 100-year-old tree is cut down and its ring widths are measured. The results can be expressed graphically, and, if a similar graph were made from a small stump found near the 100-year-old tree, the two graphs could be compared until a match of the curves was obtained. The time when the small stump was made would thereby be determined from the position of its outer ring alongside the 100-year record.

      Not every tree species nor even every specimen of a suitable species can be used. In the American Southwest, success has been achieved with yellow pine, Douglas fir, and even sagebrush. Unfortunately, the giant sequoia of California does not live in a sufficiently sensitive environment to provide a useful record. The even older bristlecone pine in California's White Mountains does have a climate-sensitive record, but its area of growth is so limited and so inaccessible that no bristlecone specimens have so far appeared in archaeological sites. This shortcoming notwithstanding, dead bristlecone pine trees are presently providing rings as old as 8,200 years for dating by carbon-14. The purpose is to check the carbon-14 method.

coral growth
      Certain fossil corals have long been used to date rocks relatively, but only recently has it been shown that corals may also serve as absolute geochronometers. They may do so by preserving a record of how many days there were in a year at the time they were growing. The number of days per year has decreased through time because the rate of rotation of the Earth has decreased; geophysical evidence suggests that days are currently lengthening at the rate of 20 seconds per million years. If this were typical of the slowdown during the past, a year consisted of 423 days about 600 million years ago.

      It is thought that horn corals (horn coral) indicate the number of days per year by means of their exceedingly fine external ridges of calcium carbonate, each of which is believed to represent a day's growth. Several hundred of the fine ridges also seem to cluster as a unit that presumably corresponds to one year. In certain modern West Indian corals the number of fine ridges in a presumed annual increment is approximately 360, suggesting that coral patterns are being properly interpreted.

      Not many fossil corals are in a state of preservation that permits the counting of ridges, but those that are seem to lend themselves well to this procedure. Several Middle Devonian corals indicate between 385 and 410 ridges, with an average of about 400. It remains to be seen whether this method of dating, so elegant in concept and so simple in application, will blossom or wither away in the years to come.

Edwin A. Olson Ed.

Additional Reading
Overviews are presented in Claude C. Albritton, Jr., The Abyss of Time, Changing Conceptions of the Earth's Antiquity After the Sixteenth Century (1980); Don L. Eicher, Geologic Time, 2nd ed. (1976); William B.N. Berry, Growth of a Prehistoric Time Scale: Based on Organic Evolution, rev. ed. (1987); Henry Faul and Carol Faul, It Began with a Stone: A History of Geology from the Stone Age to the Age of Plate Tectonics (1983); Robert H. Dott, Jr., and Roger Lyman Batten, Evolution of the Earth, 4th ed. (1988); and Reed Wicander and James S. Monroe, Historical Geology: Evolution of the Earth and Life Through Time (1989). A. Hallam, Great Geological Controversies, 2nd ed. (1989), traces the development of the history of geology and of various, often contradictory, concepts. For the early recognition of the geologic cycle and the promulgation of uniformitarianism, see the classics themselves: James Hutton, Theory of the Earth, With Proofs and Illustrations, 2 vol. (1795, reissued 1972); John Playfair, Illustrations of the Huttonian Theory of the Earth (1802, reprinted 1964); and Charles Lyell, Principles of Geology, 3 vol. (1830–33), available also in many later editions. Relevant developments in modern geologic sciences are discussed in Donald R. Prothero, Interpreting the Stratigraphic Record (1989); Ruth E. Moore, Man, Time, and Fossils: The Story of Evolution, 2nd rev. ed. (1961); Martin J.S. Rudwick, The Meaning of Fossils: Episodes in the History of Palaeontology, 2nd rev. ed. (1976, reprinted 1985); W. Lee Stokes, Essentials of Earth History: An Introduction to Historical Geology, 4th ed. (1982); Don L. Eicher and A. Lee McAlester, History of the Earth (1980); and Don L. Eicher, A. Lee McAlester, and Marcia L. Rottman, The History of the Earth's Crust (1984).Gary Dean Johnson

* * *


Universalium. 2010.

Игры ⚽ Нужен реферат?

Look at other dictionaries:

  • geochronology — 1893, from GEO (Cf. geo ) + CHRONOLOGY (Cf. chronology) …   Etymology dictionary

  • geochronology — [jē΄ōkrə näl′ə jē] n. [ GEO + CHRONOLOGY] the branch of geology dealing with the age of the earth and its materials, the dating of evolutionary stages in plant and animal development, etc. geochronological [jē΄ōkrän΄ə läj′i kəl] adj …   English World dictionary

  • Geochronology — A schematic depiction of the major events in the history of our planet. Geochronology is the science of determining the age of rocks, fossils, and sediments, within a certain degree of uncertainty inherent to the method used. A variety of dating… …   Wikipedia

  • geochronology — noun a) The science of dating samples of rock or sediment. Geochronology allows us to see more precisely how the world has evolved. b) Any one of the methods by which the age of different samples of rock can be determined. The history of the… …   Wiktionary

  • geochronology — noun Date: 1893 1. the chronology of the past as indicated by geologic data 2. the study of geochronology • geochronological also geochronologic adjective • geochronologically adverb • geochronologist noun …   New Collegiate Dictionary

  • geochronology — geochronologija statusas T sritis ekologija ir aplinkotyra apibrėžtis Mokslas, tiriantis Žemės plutos uolienų susidarymo chronologiją ir amžių. Skirstoma į reliatyviąją ir absoliučiąją. atitikmenys: angl. geochronology vok. Geochronologie, f rus …   Ekologijos terminų aiškinamasis žodynas

  • geochronology — n. [Gr. ge, earth; chronos, time; logos, discourse] The measurement of time in relation to the evolution of the earth …   Dictionary of invertebrate zoology

  • geochronology — study of measuring geological time Sciences and Studies …   Phrontistery dictionary

  • geochronology — noun the branch of geology concerned with the dating of rock formations and geological events. Derivatives geochronological adjective geochronologist noun …   English new terms dictionary

  • geochronology — geo·chronology …   English syllables

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”