Volume 5

Daniel R. Muhs , in Encyclopedia of Geology (Second Edition), 2021

Pre-Quaternary Paleosols

Within the fields of sedimentary petrology and paleoclimatology, there has been an increasing interest in pre-Quaternary paleosols in the longer geologic record. Some of this work was pioneered by Jay Quade and his students at the University of Arizona. Much of their effort involves using the stable isotope record in pedogenic carbonates (calcic or petrocalcic horizons) of pre-Quaternary paleosols to infer paleoclimate and past vegetation (e.g., Quade et al., 1995). This is a very specific use (isotopic composition) of particular soil components (carbonate accumulations), but a prerequisite for such work is accurate identification of such features as truly pedogenic calcic and petrocalcic horizons. Although such horizons can often be identified confidently in many pre-Quaternary paleosols, other types of horizons pose greater challenges for recognition in the rock record.

Mack et al. (1993), Mack and James (1994), Birkeland (1999), and Kraus (1999) provide reviews of paleosols in the longer, pre-Quaternary geologic record. Two examples of Tertiary paleosols formed in very different geologic settings are offered here. On the Canary Islands of Spain, situated off the northwestern coast of Africa, paleosols intercalated with carbonate-rich eolian or marine sand can be identified by their distinctive red colors, which differ from the host rock or sediment (Fig. 12A ). These paleosols likely formed in an arid environment, as the host calcareous sands are not leached. However, the paleosols contain fine-grained quartz and mica, minerals that are rare in Canary Islands rocks. Muhs et al. (2019) infer that part of the genesis of these paleosols is due to accretion of African dust after the host eolian sands were stabilized. During each period of paleosol formation, the dust likely accreted on a vegetated, geomorphically stable surface, and the soil gradually "grew upward" until the next phase of eolian sand accumulation buried it.

Fig. 12

Fig. 12. Examples of pre-Quaternary paleosols: (A) reddish-brown Pliocene paleosols (red arrows) on the Canary Islands, Spain, intercalated with carbonate-rich marine and eolian sands, Fuerteventura (see Meco et al., 2011; Muhs et al., 2019); (B) red, Oxisol-like paleosol (locally referred to as a laterite) intercalated between Paleocene basalt flows, Giant's Causeway, County Antrim, Northern Ireland. Stratigraphy is from Wilson and Manning (1978); interpretation of Oxisol/laterite is from Hill et al. (2000); ages of basalt flows are from Ganerød et al. (2010, 2011).

Photographs by D.R. Muhs.

The other pre-Quaternary example presented here is the spectacular Oxisol-like paleosol (commonly referred to as a "laterite;" see Wilson and Manning, 1978) found between Paleocene basalt flows at Giant's Causeway in Northern Ireland (Fig. 12B). Ages of the underlying basalt are 63–62   Ma, and while the overlying basalt is not as well dated, it is very likely less than ~   61   Ma (Ganerød et al., 2010, 2011). Hill et al. (2000) report that this paleosol, which reaches a maximum thickness of ~   30   m, is rich in kaolinite, gibbsite, goethite, and hematite, all minerals that typify highly weathered Oxisols. Extreme chemical weathering in the Giant's Causeway paleosol is indicated by SiO2 contents of 13–21%, compared to SiO2 contents in the basalt of ~   44%. In contrast, the paleosol has Al2O3 and Fe2O3 contents of 17–34% and 25–45%, respectively, whereas the basalt has Al2O3 and Fe2O3 contents of 15% and ~   14%, respectively. These differences indicate significant loss of primary silicate minerals by chemical weathering and formation of Al-rich and Fe-rich secondary minerals. Hill et al. (2000) interpret the Giant's Causeway paleosol to have formed in a time period of perhaps a million years under a humid, subtropical (or tropical?) climate. The distinctive red color of this paleosol (Fig. 12B), due to its high hematite content, certainly resembles the red Oxisols of subtropical and tropical climates.

It is important to note, however, that red colors also can be produced by diagenetic alteration, as demonstrated decades ago by Walker (1967). Indeed, one of the biggest problems with interpreting or even identifying paleosols in the rock record is that it is often very difficult to distinguish features or components produced by pedogenesis from those produced by diagenesis. Both Dahms and Holliday (1998) and Birkeland (1999) discuss many of the problems in identifying pre-Quaternary paleosols. Birkeland (1999) points out that many hypothesized soil features in suspected paleosols must be carefully distinguished through what he calls the "veil of diagenesis." Nevertheless, the examples given here serve to demonstrate that despite the problems of differentiating diagenetic processes and pedogenic processes, pre-Quaternary paleosols can be identified in the longer geologic record.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780124095489120020

HISTORY OF GEOLOGY FROM 1835 TO 1900

D.R. Oldroyd , in Encyclopedia of Geology, 2005

Rocks and their Formation

There were few major contributions to sedimentary petrology in the period here under review. The Irishman Patrick Ganly's 1830s discovery of the 'way-upness' criterion offered by current bedding was not utilized until the twentieth century. In 1839 Christian Ehrenberg published his microscopic studies of chalks and limestones. Formal distinction of sedimentary, igneous and metamorphic rocks was made by Henri Coquand in 1857.

Various igneous rocks such as basalts, granites, gabbros, syenites, or porphyries had been recognised since antiquity and many classifications were proposed in the nineteenth century, according to chemical and/or mineral composition, texture, or supposed mode of formation, but the field was confused. Examination of rocks in thin section by Henry Sorby assisted in a sense, but the proliferation of information also added to the confusion. While Huttonian theory was triumphant as regards Werner's original theory, there was continued interest in the role of water in the formation of igneous and metamorphic rocks. Notably, in 1857 and 1859 Gabriel Auguste Daubrée of the French Mines Department subjected materials to high temperatures and pressures, with or without water, and concluded that new minerals could crystallize without wholesale melting. He thought that past conditions could have been radically different from those at present and that there might have been an ocean primitif. Granite was thought to be produced by 'aqueous plasticity', not igneous melting. At Freiberg, Bernhard von Cotta held analogous views. But in 1860 Daubrée thought that foliations were due to pressure during 'regional metamorphism'.

The variety of igneous rocks raised the question whether there were different kinds of subterranean magma, or whether some processes of differentiation occurred from essentially the same starting material. In 1844 Darwin had the idea of differentiation of magma by gravity settling of first-formed crystals; and in 1846 he distinguished cleavage, foliation, and stratification (while regarding gneisses as stratified rocks). Dana thought that differentiation of magma might precede crystallization. By contrast, on the basis of observations in Iceland, in 1851 and 1853 the chemist Robert Bunsen proposed that there were two separate magma chambers under the island, producing 'trachytic' and 'pyroxenic' rocks or intermediate mixtures. In 1853 Wolfgang Sartorius von Waltershausen hypothesized the existence of different subterranean zones; and the eruption of more siliceous types preceded the more basic, thus relating igneous compositions to age. In 1857 Joseph Durocher asserted the liquation model and priority for the idea over Bunsen. Metallic lodes were ascribed to 'émanations'. Following work in Hungary and California, Ferdinand von Richthofen envisaged a succession of magmas, the earliest being more siliceous, the great outpourings of Tertiary basalts being due to earlier depletion of siliceous magma. In 1878 in America, Clarence King thought pressure release could facilitate fusion. In 1880 Dutton suggested that fusion could follow pressure release, local temperature elevation, or water absorption.

The master petrologists of the period were Harry Rosenbusch in Strasbourg, Ferdinand Zirkel in Leipzig and Ferdinand Fouqué and Auguste Michel-Lévy in Paris, who specialized in the study of feldspars. All were adept with the use of the petrographic microscope. In 1873 Rosenbusch published a catalogue of all then known magmatic and metamorphic rock types. Rosenbusch's 1877 study of metamorphism around the Barr–Andlau granite in the Vosges was important for his recognition of zones of contact metamorphism (schists, knotted schists, hornfelses), seemingly without feldspars. But Michel-Lévy found feldspars in the contact aureole of the Flamandville granite in Normandy and he and Fouqué thought there was no fundamental distinction between contact and regional metamorphism. Throughout this period Continental petrologists continued, in the Wernerian tradition, to try to find relationships between age and 'hard-rock' composition, whereas their British counterparts chiefly concerned themselves with biostratigraphy.

However, in 1893 the British surveyor George Barrow, working on metamorphic rocks in the southern Scottish Highlands, found characteristic metamorphic minerals (sillimanite, kyanite, and staurolite) around a granitic mass, and this gave him mappable subdivisions of the region. These were developed in the twentieth century as 'Barrovian zones', but the useful idea was not initially followed up.

Experimental petrology was undertaken in the nineteenth century, but until high-pressure and pressure techniques were developed in the twentieth century for simulating rock formations, work on phase diagrams was developed, and ideas about the structure of the Earth's interior could be pursued through seismology, petrological understanding remained speculative and somewhat at the level of natural history.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B0123693969003695

Calcium Carbonate Features

Nicolas Durand , ... Eric P. Verrecchia , in Interpretation of Micromorphological Features of Soils and Regoliths (Second Edition), 2018

1.3 Terminology

Geologists have long shown an interest in calcitic features in thin sections. A rich literature exists on diagenetic carbonate features in sedimentary petrology (e.g., Dunham, 1969a, 1969b; Walls et al., 1973; Harrison & Steinen, 1978; Adams, 1980; Freytet & Plaziat, 1982; Esteban & Klappa, 1983; James & Choquette, 1984; Harwood, 1988). A comprehensive overview of continental calcitic features was presented by Reeves (1976). In addition, a number of soil scientists who have studied microscopically discrete calcitic features in soils have proposed a classification of those features (Blokhuis et al., 1969; Sehgal & Stoops, 1972; Wieder & Yaalon, 1974; Bal, 1975a, 1975b; Ruellan et al., 1979; Verrecchia, 1987; Wright, 1990; Becze-Deák et al., 1997; Zamanian et al., 2016), including a classification developed for the discussion of carbon sequestration by pedogenic carbonates (Monger et al., 2015). Other accounts of the genesis and properties of petrocalcic horizon and calcrete profiles are given by Gile et al. (1966), Elloy and Thomas (1981), Braithwaite (1983), Goudie (1983), Esteban and Klappa (1983), Verrecchia and Freytet (1987), Wright and Tucker (1991) and Alonso-Zarza and Tanner (2010).

For general descriptive terms of pedofeatures, the reader is referred to Bullock et al. (1985) and Stoops (2003). Some of the terms and concepts used for pedogenic calcitic features are taken from sedimentary petrography. These terms include micrite (crystal size <4   μm in diameter, seen but not readily measured under the optical microscope), microsparite (5-20   μm) and sparite (>20   μm) (Folk, 1962; Chilingar et al., 1965). Isolated pedogenic crystals are considered as crystalline pedofeatures only when larger than 20   μm, while smaller crystals are part of the micromass.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780444635228000097

Redoximorphic Features

David L. Lindbo , ... Mike J. Vepraskas , in Interpretation of Micromorphological Features of Soils and Regoliths, 2010

Publisher Summary

Redoximorphic features have been referred to by various terms over the past decades. The terminology used to describe redoximorphic features is derived from sedimentary petrology and soil micromorphology. Redoximorphic features are common in most soils where water saturation occurs and their presence is used extensively in making land use decisions. Although redoximorphic features are visible in the field with both the naked eye and a hand lens, micromorphological analysis can further enhance our understanding of how these features form and how to interpret them correctly. These interpretations are best made when supportive information such as water table and climatic data are available, but they can to a large extent be based on extensive micromorphological data correlating redoximorphic pedofeatures to environmental conditions.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780444531568000088

Preliminary Scientific Results of SK-1

Wang Chengshan , ... Stephan A. Grahamet al., in Initial Report of Continental Scientific Drilling Project of the Cretaceous Songliao Basin (SK-1) in China, 2019

4.3.3 Preliminary Results

1.

Meter-scale cycle identification was conducted on the Quantou Formation in SK-1s. Sedimentary cycles identified were then compared with cycles determined by deep laterolog (RD) (Fig. 4.5).

Figure 4.5. Fischer analysis of the Quantou Formation in SK-1s.

2.

Relations between sedimentary petrology and geophysical response were analyzed. Natural gamma spectrum logging was considered to better respond to the determination of climate change, and was chosen as an important parameter for paleoclimate studies.

3.

According to coherent analysis, logging parameters, such as natural gamma spectrum U, Th/U, acoustic travel time AC, and deep resistivity R D, are all related to source rock organic content (Fig. 4.6). Through multiple stepwise regression, a logging interpretation model was used to determine source rock organic content, which may more effectively predict source rock distribution.

Figure 4.6. Correlation of organic content between calculated data derived from logging and actual measurement.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780128129289000049

Calcium Carbonate Features

Nicolas Durand , ... Matthew G. Canti , in Interpretation of Micromorphological Features of Soils and Regoliths, 2010

1.3 Terminology

Geologists have long shown an interest in calcitic features in thin sections and more recently in studying them with submicroscopic techniques (e.g. James, 1972; Freytet, 1973; Freytet & Plaziat, 1982; Tucker & Wright, 1990 ). A rich literature exists on diagenetic carbonate features in sedimentary petrology (e.g. Dunham, 1969a,b; Walls et al., 1973; Harrison & Steinen, 1978; Adams, 1980; Esteban & Klappa, 1983; James & Choquette, 1984; Harwood, 1988). A comprehensive list of continental calcitic features is given by Reeves (1976).

A number of soil scientists who have studied discrete calcitic features in soils microscopically have proposed a classification of those features (Blokhuis et al., 1969; Sehgal & Stoops, 1972; Wieder & Yaalon, 1974; Bal, 1975a,b ; Ruellan et al., 1979; Wright, 1990).

Previous accounts of the genesis and properties of petrocalcic horizon and calcrete profiles are given by Gile et al. (1966), Elloy and Thomas (1981), Braithwaite (1983), Goudie (1983), Esteban and Klappa (1983), Verrecchia and Freytet (1987), Wright and Tucker (1991).

For general descriptive terms of pedofeatures the reader is referred to Bullock et al. (1985) and Stoops (2003). Some of the terms and concepts used for pedogenic calcitic features are taken from sedimentary petrography. These terms include micrite (crystal size <4   µm diameter, seen but not readily measured under the optical microscope), microsparite (5–20   μm) and sparite (>20   μm) (Folk, 1962; Chilingar et al., 1965). Isolated pedogenic crystals are considered as crystalline pedofeatures only when larger than 20   μm, while smaller crystals are part of the micromass.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B978044453156800009X

Surficial Geochemical Exploration Methods

P. Sarala , in Mineral Deposits of Finland, 2015

Lithogeochemistry

Lithogeochemistry is a geochemical method that involves the sampling of bedrock. It can be considered for the recognition and definition of mineralized and anomalous areas, and to distinguish dispersion halos at different mapping scales (Hawkes, 1957). In a broad sense, lithogeochemistry covers the research of igneous, sedimentary, and metamorphic petrology; hydrothermal alteration; weathering; and diagenesis. Also, it has a close connection to the mining-related fields of applied geochemistry (including exploration and environmental research and monitoring), genesis of mineral deposits, metallurgy, and deep hydrogeochemistry.

Lithogeochemical surveys can be carried out on a grid or on traverses of an area, with samples taken of all available rock outcrops or at some specific interval. After lithological mapping, one or several rock types may be selected for sampling and analyzed for various elements. Using symbols and/or contours, results are compiled in geochemical maps to allow interpretation. Under favorable conditions, mineralized zones or belts may be outlined in which more detailed work can be concentrated. In the case of large territories, geochemical provinces may be outlined.

Isotopic analyses can help build stratigraphical and geochronological models. Applicable isotopic systems include stable isotopes such as 16O/18O, 12C/13C, 32S/34S, and common Pb and radiogenic isotopes such as samarium-neodymium and rhenium-osmium (Faure and Mensing, 2004). For example, the isotopic composition of sulfur can help identify the origin of hydrothermal fluids and the conditions within the depositional environment. Isotopic ratios are also used for age determination of rock types (geochronology) relying on radiogenic isotopes (e.g., U series, U-Pb, K-Ar, Sm-Nd) (Allègre, 2008). Although isotopic geochemistry is an important part of the geochemical research, it is beyond the scope of this chapter and not handled in detail here.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780124104389000273

Sedimentologist's Guide for Recognition, Description, and Classification of Paleosols

N.J. Tabor , ... L.A. Michel , in Terrestrial Depositional Systems, 2017

Conclusion

Although paleosols are a relatively minor component of the rock record in terms of volume, the paleoclimatic and paleoenvironmental information they contain makes them especially valuable to geoscientists. The rapid increase in the number of paleosol studies in peer-reviewed literature over the past 30 years ensures that paleopedology will remain a significant discipline for the foreseeable future. Given the close relationship of paleopedology with sedimentology and sedimentary petrology, it is imperative that sedimentologists and stratigraphers have a working knowledge of paleopedology and be capable of recognizing, describing, and communicating information about paleosols.

Paleosol research requires careful attention to observation, measurement, and recording of details in the field. Paleopedology studies should also incorporate laboratory-based petrographic, mineralogic, elemental, and isotopic measurements in order to provide additional data related to paleoecology, paleoenvironments, and diagenesis. The information and photos presented here offer a broad overview of many types of paleosols and paleosol features as they appear in the field. We have attempted to differentiate the rare and exotic examples of features from those that are frequently encountered. We have tried to indicate which soil features are resistant to diagenesis and are likely to be preserved in paleosols without significant alteration, and tried to contrast these with mutable characteristics that should not be relied on for paleoenvironmental reconstruction. We also hope that we have convinced readers that modern soils and paleosols are fundamentally different entities that require different methods of analysis and classification. While no single classification scheme has been developed to date that appears to adequately address the needs of field-based sedimentologists, stratigraphers, and paleopedologists, continued research and communication among these workers may arrive at a reasonable system to bridge the importance of modern soils and their fossil records, and help to inform our understanding of past environments. Most of all, we hope that this guide conveys the utility and versatility of paleosols as complex data storage media and will inspire readers to continue to explore the expanding discipline of paleopedology.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780128032435000042

Volume 2

John D. Winter , in Encyclopedia of Geology (Second Edition), 2021

Specific Metamorphic Rock Types

As mentioned above, some rock types are sufficiently common that they have been given special names, typically based on a common and specific protolith, but many also imply a specific range of metamorphic grade. It may also be proper to name a metamorphic rock by adding the prefix meta- to a term that indicates the protolith, such as meta-pelite, meta-ironstone, etc.

The commonly used names that specify a particular metamorphic rock type are listed below. As a rule, these names take precedence over the textural names described above.

Marble. A metamorphic rock composed predominantly of calcite or dolomite. The protolith is typically limestone or dolostone.

Quartzite . A metamorphic rock composed predominantly of quartz. The protolith is typically sandstone. Some confusion may result from the use of this term in sedimentary petrology for a pure quartz sandstone.

Greenschist/Greenstone. A low-grade metamorphic rock that typically contains chlorite, actinolite, epidote, and albite. Note that the first three minerals are green, which imparts the color to the rock. Such a rock is called greenschist if foliated, and greenstone if not. The protolith is either a mafic igneous rock or graywacke.

Amphibolite. A metamorphic rock dominated by hornblende + plagioclase. Amphibolites may be foliated (schistose or gneissose) or non-foliated. The protolith is either a mafic igneous rock or graywacke.

Serpentinite. An ultramafic rock metamorphosed at low grade, so that it contains mostly serpentine.

Blueschist. A blue-amphibole-bearing metamorphosed mafic igneous rock or mafic graywacke. This term is so commonly applied to such rocks that it is even applied to non-schistose samples. Glaucophane is the most common blue amphibole, and glaucophane schist has been applied to rocks known to contain it.

Eclogite. A green and red metamorphic rock that contains clinopyroxene and garnet (omphacite + pyrope). The protolith is typically basaltic. Eclogites contain no plagioclase.

Calc-silicate rock (granofels or schist). A rock composed of various Ca-Mg-Fe-Al silicate minerals such as grossular, epidote, tremolite, vesuvianite, etc. The protolith is typically a limestone or dolostone with silica either originally present as clastic grains or introduced metasomatically.

Skarn. A calc-silicate rock (see immediately above) formed by contact metamorphism and silica metasomatism from a pluton into an adjacent carbonate rock. Tactite is a synonym.

Granulite. A high grade rock of pelitic, mafic, or quartzo-feldspathic parentage that is predominantly composed of OH-free minerals. Muscovite is absent, and plagioclase and orthopyroxene are common.

Migmatite. A composite silicate rock that is heterogeneous on the 1- to 10-cm scale, typically having a dark gneissic matrix (melanosome) and lighter felsic portions (leucosome). Migmatites may appear layered, or the leucosomes may occur as pods or form a network of cross-cutting veins.

Any of these terms, or a term indicating the protolith, may be combined with the textural classification terms if it helps describe a rock more fully. One may choose to use either term as a modifier. One may thus call a rock a schistose amphibolite, an amphibolitic schist, a pelitic schist, or a phyllitic meta-tuff, etc.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780124095489125424

Sediment generation and provenance: processes and pathways

Eduardo Garzanti , in Sedimentary Geology, 2016

1 Introduction

The origin of sedimentary petrology dates back to the late 19th century, with the invention of thin-section petrography by H.C. Sorby. Provenance studies flourished in the first half of the 20th century, when P.D. Krynine – inspired in Moscow by the ideas of his teacher M.S. Shvetsov – became the first strong advocate of tectonic control on sandstone composition. A feverish activity in sedimentary petrology followed the first classification schemes proposed in the late 1940s ( Krynine, 1948; Pettijohn, 1954), and new classifications continued to proliferate through the 1950s and 1960s (McBride, 1963; Dott, 1964; Folk, 1980). Virtually all of these were projected onto triangular diagrams, which force us to consider three parameters only at a time (quartz, feldspars, rock fragments, or in other cases micas or "clay matrix"). The underlying conceptual schemes were not based on a thorough investigation of modern settings (Suttner, 1974) and did not benefit yet from the breakthrough of the plate-tectonic revolution. Major progress came with the work of Dickinson (1970), who a few years after Gazzi (1966) established operational rules to improve the reproducibility of detrital modes and showed how to connect them with paleogeodynamic scenarios in a seemingly straightforward univocal way (Dickinson and Suczek, 1979). The new paradigm hinged on the axiom that "detrital modes of sandstone suites primarily reflect the different tectonic settings of provenance terranes" (Dickinson, 1985, p. 333). A parallel, somewhat milder statement applied to geochemical composition – echoing Crook (1974) – is Bhatia's (1983, p. 611) "close correlation exists between the geochemical composition of sandstones and tectonic settings of sedimentary basins." Such claims reflect the enthusiasm of those years, when plate-tectonic theory finally made geologists aware of the fundamental controls of tectonic processes and seemed able to provide a fresh new explanation for each geological phenomenon. The power of suggestion was strong, and the enchantment persisted through time. Even though other schools refused to adopt the same radical attitude and focused more on compositional modifications during erosion, transport, and deposition (e.g., Suttner et al., 1981), sedimentary petrology has remained nailed to that vision for three decades, confirming that "knowledge is in the end based on acknowledgment" (Wittgenstein, 1974, #378).

In provenance analysis, as in other fields, we may either trust a simplified model based on what is believed to be essential or surrender to nature's baffling complexities and interpret each case as the unique result of multiple competing causes, giving up the quest for a general theory (Paola and Leeder, 2011). To overcome such trade-off and to preserve the benefits that theory offers as a basis for interpretation (Weltje, 2012), we need to acknowledge that the lithological characteristics of parent rocks as inferred from detrital modes of a daughter sandstone cannot be used blindly as a proxy for plate-tectonic setting. The path leading from a handful of sand to a geodynamic scenario has long been known to be winding and fraught with difficulties (Basu, 1985; Johnsson, 1993). And yet, if we make the most of their great potential, then traditional petrographic methods can provide us with a simple and very effective means to identify the nature and tectonostratigraphy of source terranes, and thus with an unexcelled key to track their erosional evolution through space and time. Whenever basic tools such as the optical microscope are left aside in favor of advanced efficient machines that produce a great deal of numbers on a very minor and possibly non-representative fraction of the sediment, the ultimate risk is to find ourselves facing a sea of data without a suitable vessel to sail. The rapid development of technological devices should not make us feel that culture is superfluous, and thus induce us to throw away the traditional keys to understanding.

Read full article

URL:

https://www.sciencedirect.com/science/article/pii/S0037073815001608