Theology and Science

Dirk Evers

This article begins with an introduction briefly sketching the concepts of Western Christian theology and science in a systematic perspective informed by the historical shaping of the terms. It then offers an inquiry into the biblical traditions and their relations to the development of science in the modern age. This is followed by a general account of the development of the interactions between theology and science from late modernity until today, and relates this to methodological questions and positions. Finally, the article presents selected issues and themes discussed in such discourses which today are labelled ‘theology and science’.

1 Introduction

‘Science’ and ‘theology’ as concepts reflect time-dependent cultural ways of structuring reality, rather than referring to given phenomena in the natural, empirical world. Both are high-level general terms, and it takes careful consideration not to misunderstand them as references to quasi-objective entities, or even quasi-agents acting in history and culture. When we speak of science we refer to certain practices, communities, institutions, and a whole variety of methods and bodies of knowledge, and at the same time we refer to discourses in which the label science is used as a value-laden expression for certain ways to understand reality. Discourses on theology and science regularly suffer from distortions when they compete for public attention and belief formation, and mix facts with values.

Taken literally, the term ‘theology’ refers to discourse on God or the divine. Consequently, Augustine of Hippo defined the Latin equivalent for the original Greek term, theologia, as ‘thought or speech explaining the divine essence’ (‘de divinitate ratio sive sermo’, Augustine 1968: 4–5). Theology as understood in this entry is the public and communal reflection on the divine and human-divine relations from within a particular religious tradition and practice. It critically and constructively reflects the respective religion’s truth claims from within its tradition and develops concepts, bodies of doctrine, and common forms of speech. Since God or the divine is not a given object of reference, the notions of God or the divine can vary immensely, and they vary not only among different faith traditions but also within a particular tradition. Theology today must also reflect on this plurality and develop means and open hermeneutic perspectives for constructive dialogue across different traditions. For any contemporary academic theology this includes interdisciplinary dialogue with other disciplines, including the sciences.

Thus ‘theology’ and ‘science’ cannot interact as such – be it in conflict, dialogue, or integration. Only people can communicate and participate in public and academic discourses on scientific and religious worldviews, and they do so in different situations, under different premises, and with different goals. Therefore, if we want to reflect on the interactions and challenges between science and theology (with theology understood as the self-reflective form of religion), we have to consider carefully the differences between the fields, the different levels of possible discourse, and the premises involved. In the case of this article, the approach is from an emic (insider) perspective of Western Christian theology, and it begins with reflections on biblical traditions in relation to what are today identified as issues between theology and science.

2 Scripture

It would be an inappropriate anachronism to look for ‘theology and science’ in the biblical traditions. Even the New Testament does not use the term ‘theology’, and, obviously, ‘science’ in any modern sense of the word is not present in the biblical traditions. However, although the concepts are missing, we can identify central themes and issues of the modern debates in many biblical texts. Furthermore, biblical texts, concepts, and stances have had an important impact on the development of modern science, and they provided vehemently debated reference points in discourses surrounding it. Even side-remarks like Josh 10:12 (that – because of God’s miraculous intervention – the sun stood still at Gibeon) could be interpreted as evidence that the Bible promotes a geocentric model of the universe. Thus the modern theological debates around the character of scripture as a document of divine revelation and its authority were significantly fuelled by debates about the difference between biblical views of creation and modern scientific accounts. This first chapter will begin by addressing certain traditions within scripture, in which we find themes that can be understood as belonging to the field of discourse that today we label ‘theology and science’. It will then take a short glance at some ways in which scripture is related to early modern science and how that shaped scriptural hermeneutics.

2.1 Old Testament

Many important texts in scripture deal with the natural world, its creation, and its order with reference to God the creator. They usually do so by employing notions and categories which are derived from cosmological concepts contemporary to their times. The Hebrew Bible begins with two narratives about creation (Gen 1:1–2:4a; 2:4b–25), which deal with the origins of the structure of the world as a whole, as it was known in the Iron Age. Historical criticism has pointed to the fact that both narratives draw from different cosmological backgrounds and originated in different centuries – with the second narrative being the significantly older, more archaic, and more mythical of the two. The setting of the second narrative is provided by the garden of Eden, which in its primordial state is described as a desert without water or rain. Creation begins when God waters this desert and converts it into a fertile mythical garden. There are many other details that resemble not only Babylon’s creation myth, Enuma Elish, but also the Akkadian epic Atra-Hasis. For this article it is important to note that the biblical traditions developed the notion of creation and the creator God – who was identified with Yahweh, the God of Israel – in close contact with different forms of common knowledge and cosmological models as part of ancient oriental ‘science’. And the early editors of the now-canonical texts saw the necessity of compiling different cosmological models and notions from different sources without seeking to integrate them into one coherent cosmology.

This is demonstrated by the fact that the younger narrative (Gen 1:1–2:4a) is put first in the canonical text and serves as general account of the primeval creation of the cosmos and all living beings, including humanity. Genesis 1 is usually ascribed to a source with priestly background, often located in Judah in the fifth century BCE when it was under Persian imperial rule. Here a more abstract cosmology is developed, again along the lines of contemporary (fifth-century) models and notions. The stage is set, not in the garden as the habitat for humans, but in a structured cosmos that is created out of a void or chaos, with heavenly bodies such as the sun, moon, and stars, and with an Earth divided into different habitats and populated by different classes of plants and animals. The different acts of creation are conceptualized as acts of speech and/or as creational works. God’s creational activity is summarized with the word bara, an activity solely ascribed to Creator God, while no human nor any other creature can engage in bara.

All in all it seems obvious that the biblical texts borrow themes, models, and notions from ancient mythology and cosmology but adapt them according to their monotheistic view of the Creator God, and demythologize all other divine and demonic forces within creation. Some have described this as part of a transformation of religions around the world, which German philosopher Karl Jaspers called the Axial Age (Jaspers 2017). Others have followed him (see Eisenstadt 1986 and Joas and Bellah 2012). According to this thesis, from about 900 to 200 BCE, in four different cultural spheres on the globe, conceptions were formed which to this day prove to be pivotal to humanity as we know it. Although this notion is highly contested on both philosophical and historical grounds, it points to the fact that in the biblical traditions as well as in other contexts a fundamental change of cosmology took place, which overcame mythological narratives by establishing views of the world as a whole, and distinguishing this from history as the flow of events that happen within the created world. That led to a demythologizing as well as de-demonizing of the world. This can be substantiated in the biblical narrative. Sun and moon – divine beings in the ancient world – are transformed into a greater and a lesser light, with the purpose of giving light during day and night. Also, the Creator God does not have to bring creation into being by combatting chaotic and antagonistic forces as occurs in other ancient creation stories, but does so by his word, which brings light and matter into being and lets the earth bring forth plants etc. Such a view of creation overcomes magical thinking, and allows for a certain distance between human beings and the reality of the cosmos, which is necessary in order to understand creation as a structured and organized whole. Therefore, ‘creation’ in the Bible not only designates the act of creation but also its result: the cosmos as creation.

Accordingly, the main focus of the creation narratives in the Bible is an analysis of the conditions and structures of the world, not a detailed reconstruction of its origin, and different aspects of such a religious view of reality can be expressed by different narratives and stories. This not only applies to the creation narratives. We find concepts of creation in the Psalms and in the prophets. Second Isaiah (Isa 40–55), for example, uses the term bara in order to express God’s power to restore Israel after exile. The plurality of cosmological notions thus reflects different issues of concern, like the question of the stability of the world, its goodness, the power of the God of Israel in comparison to other forces and divine figures, the purpose of human existence, and the congruence of cosmological, anthropological, social, and political order.

In this context, a very early form of science emerged in the ancient Near East in which the biblical traditions also participate: the traditions of wisdom as found in the wisdom literature of the Hebrew Bible. Again, this is not a tradition unique to the Bible, but an international type of knowledge in the ancient Near East. Close parallels are found in Mesopotamia, and even more in Egypt. Wisdom reflects the order of creation, where everything has its place and time, and it is intellectual and pragmatic at the same time. It is a guide to dealing appropriately with nature as well as other human beings. It contributes to the successful conduct of life and helps to avoid crises or to deal with them. Foundational for this attitude towards nature was the conviction that there are regular basic structures of reality which form the background of individual experiences and allow for generalizations. The Bible depicts King Solomon as the exemplary character of a wise man, who is not only a wise ruler but also a learned man with profound knowledge of natural things:

And God gave Solomon wisdom and understanding beyond measure, and breadth of mind like the sand on the seashore […] He also spoke 3,000 proverbs, and his songs were 1,005. He spoke of trees, from the cedar that is in Lebanon to the hyssop that grows out of the wall. He spoke also of beasts, and of birds, and of reptiles, and of fish. And people of all nations came to hear the wisdom of Solomon, and from all the kings of the Earth, who had heard of his wisdom. (1 Kgs 4:29, 32–34)

The readers of wisdom texts are addressed as potential competent participants in such knowledge of the natural world, and they are challenged to get wisdom by the fear of God leading to honest desire for knowledge and by studying God’s words and works. The notion of personified wisdom became important in later wisdom texts (cf. Prov 1–9) which inspired the notion of an all-pervading rationality of nature. For early modern science referring to mathematical laws of nature, Wis 11:20 provided inspiration, claiming that God has ‘arranged all things by measure and number and weight’.

2.2 New Testament

The New Testament presupposes the Old Testament notions of creation and wisdom, as well as the main features of its cosmology. That creation has a beginning in time set by the Creator is self-evident for all biblical traditions. Paul refers to the creational acts of God through words according to Genesis 1, by which God ‘calls into existence the things that do not exist’ (Rom 4:17; cf. 2 Cor 4:6 and Col 1). In certain texts central motives and notions of the personified wisdom are transferred to Jesus Christ (cf. Matt 11–13) and, especially in hymnal texts, Jesus Christ is understood as the eternal, pre-existent Logos, through whom creation was established and ordered (John 1:1–18; Col 1:15–20; Heb 1:1–4). However, although the world is often referred to as ‘cosmos’ (kosmos), because the world is also the place of sin and suffering, it is not seen as a coherent whole (as in some Hellenistic traditions, where ‘cosmos’ resonated with categories such as ‘order’ or ‘ornament’), but as the epitome of everything that actually and factually is. Although the New Testament texts refer to cosmological notions of their time, and to Jewish as well as Hellenistic concepts of wisdom and understanding of created reality, the relations between religious notions and philosophical thinking are rarely discussed. Paul, however, refers to the cross of Jesus Christ as in opposition to the wisdom of the world (1 Cor 1:18–31).

2.3 Scripture and science: biblical hermeneutics

Early Christianity did not consider the biblical texts as particularly relevant sources for an understanding of nature or cosmology. The predominant understanding of the scriptures was as spiritual texts with different layers of meaning, with the literal, historical understanding as a shallow surface that conceals rather than reveals the decisive spiritual meaning. According to the early Alexandrian theologian Origen, many erroneous notions and inappropriate, impious understandings of the divine arise when scripture is not understood according to its spiritual meaning but rather according to the literal (Origen of Alexandria 1966: 271, IV/2,2). According to his view, the biblical texts present types rather than historical tokens, and prophecies rather than reports, so that their real meaning has to be found by going beyond their literal surface. He thus distinguished between body, soul, and spirit (soma, psyche, pneuma, see Origen of Alexandria 1966: 276, IV/2,4) of a text. The body is the literal or historical sense directed towards the beginners of faith, the soul is edifying those progressing in faith, while the deep spiritual meaning leads the wise believer towards full knowledge of the divine. Origen even claimed that there are biblical texts which have no historical, literal sense at all, while all have a deeper, psychological, or spiritual sense. Accordingly, Origen’s notion of creation refers to spiritual rather than cosmological affairs. Many among the early Christian theologians combined the biblical creation accounts with philosophical thinking, mainly Platonic and Neo-Platonic notions. Augustine, for example, in his exegesis of the book of Genesis in its literal meaning (De genesi ad litteram), tried to identify what from the biblical texts can and must be defended as factual. The six days of creation, for example, he understood as a logical framework, not as a distinct period of time. He saw it as futile and a waste of time to reflect too much on cosmology, the celestial spheres, the form of the Earth, or the particulars of the creation of living beings.

In medieval times, scripture was interpreted along the lines of tradition as prefigured by the fathers of the early church. After the establishment of universities and theological faculties, and with a fresh and full reception of the work of Aristotle as mediated by the Islamic world, scholastic theology merged biblical notions with Aristotelian cosmology and used rational philosophy as a preamble to the articles of faith. Scripture was understood as having different layers of meaning: the historical or literal sense was understood as fundamental but superseded by spiritual senses (allegorical, moral, and anagogic). However, the Reformation movement established a mainly literal approach to scripture. Martin Luther and the other reformers, at least in their claims, disposed of the spiritual sense of scripture and allowed only the literal sense. They saw any spiritual sense, which was usually linked to tradition and authority, as a means to control faith, thus spoiling the gospel of grace. That focus, on a literal understanding of the biblical texts in their original language as accessible to scholars, was accompanied by a move towards a practical rather than contemplative view of nature, with nature seen as the primary realm of human vocation, while religious ceremonies like worship and prayer changed their meaning from religious works of merit to tools linking the personal life to God. That promoted an experimental and more technical approach towards nature. Important parts of the Bible, like the creation narratives in the first chapters of Genesis or the accounts of Noah and the flood, were now read not as mysterious texts with a spiritual meaning, but as reporting historical events (Rudwick 2014). Thus the

Bible – its contents, the controversies it generated, its varying fortunes as an authority, and most importantly, the new way in which it was read by the Protestants – played a central role in the emergence of natural science in the seventeenth century. (Harrison 1998: 4–5)

However, when more and more tensions and outright contradictions between biblical notions and new scientific insights (as well as within the biblical traditions) became obvious, biblical texts lost their informative credibility. Historical criticism since the mid-eighteenth century began to understand biblical texts as religious products of their time, undermining faith in the historical content and the accuracy of scripture, and revealing the non-factual character of many of its genres. It was against this background that the strong Conflict Thesis in the late nineteenth century emerged, claiming a necessary and inherent conflict between science and religion.

3 Theology and science since the nineteenth century

3.1 Intrinsic conflict?

Present-day debates on the relations and tensions between theology and science often presuppose a deep and inherent conflict between religion and science. This has become known as the Warfare or Conflict Thesis of science and religion (see Lindberg and Numbers 1987; Brooke 1998: 45–48; Ungureanu 2019). In its strong form this thesis claims that there exists an inherent and necessary conflict between scientific and religious views, which are contradictory in their truth claims and antagonistic in their cultural significance. Historically, the Conflict Thesis in its strong form was a result of debates in the aftermath of the late eighteenth century (the Enlightenment and the French revolution), reaching the broader public – including the newly emancipated and educated working class – by the end of the nineteenth century. Political upheavals, new media such as newspapers, journals, and popular books, as well as a concept of education as self-cultivation, shaped a new understanding of science and religion as antagonists:

Between 1750 and 1870 – from the publication of the Encyclopédie to the early work of Nietzsche and of Darwin’s Descent of man – the relationship of science and religion in the western world passed from fruitful co-operation and modest tensions to harsh public conflict, a situation that many observers have since come incorrectly to assume to be a permanent fact of modern cultural life. (Turner 2010: 87)

In the nineteenth century, university philosophical faculties were broken up and the natural sciences were organized in independent institutions in and outside universities. Thus the profession of a scientist was established – a term first coined in 1834 by William Whewell, in place of the older terms ‘natural philosopher’ or ‘man of science’. Scientific research gained enormous momentum, was subsidized with public money, and built up social prestige by unravelling hitherto unsolved mysteries of nature, discovering numerous new phenomena from molecules to galaxies to evolution to electro-magnetism. Many scientists became public figures and defended the claim that science was the modern means to explain the totality of reality. In the English-speaking world, the Conflict Thesis became popular and widespread when John William Draper (1811–1882), an English-born North American scientist, wrote a History of the Conflict Between Religion and Science (1874). He described the history of science as shaped by a fundamental conflict between intellect and religion, so that it must be depicted as ‘a narrative of the conflict of two contending powers, the expansive force of the human intellect on one side, and the compression arising from traditionary faith and human interests on the other’ (Draper 1874: vi). Draper’s argument was mainly directed against the Catholic Church and its anti-liberal encyclicals, which brought the book onto the index of forbidden books of the Catholic Church, while he himself followed a liberal Protestant agenda. The other important protagonist of the Conflict Thesis was the North American historian and first president of Cornell University Andrew Dickson White (1832–1918), with his popular History of the Warfare of Science with Theology in Christendom (1896). He also employed the metaphor of war and envisioned a profound antagonism between science and dogmatic theology. Again, White was not an atheist, but favoured a liberal, non-doctrinal version of the Christian religion and put science over religious tradition, claiming that ‘all untrammelled scientific investigation, no matter how dangerous to religion some of its stages may have seemed for the time to be, has invariably resulted in the highest good both of religion and science’ (White 1896: viii).

These popular accounts merged with a general political liberalism, which in the aftermath of the American and French revolutions all over Europe challenged the political establishment with demands of emancipation, religious toleration, democratic reforms, and socialist revolutionary propaganda. Early modern materialism was endemic in England since the times of Hobbes and was critical towards Christianity, especially with the French materialists of the eighteenth century such as La Mettrie and d’Holbach who believed that reality was made of a single material substance. It took an anthropological twist with German idealist philosopher Ludwig Feuerbach, who interpreted religion as the outward projection of the alienated inner human nature. He deeply influenced Marxist political thinking, and – together with Nietzsche and others – shaped a view of religion as an expression of human longing, a longing exploited by religious institutions and priestly castes in cooperation with political powers in order to suppress the masses and detain them from education and liberation. Science was identified as a driving force of intellectual progress at the cost of religion, at least in its traditional doctrinal form. This merged with an interest in great heroic figures of the past who could serve as role models, fighting against unyielding institutions and dogmatic prejudices. Christopher Columbus was presented as fighting for the spherical form of the Earth against a stubborn council of Salamanca – a literary fiction invented by American author Washington Irving (see Irving 1828) and included in Draper’s History – while Galileo became a heroic fighter against the Inquisition. With Darwin’s theory of evolution, which integrated human beings into a developmental history of life on this planet driven by purely natural causes, the second half of the nineteenth century saw the rise of naturalistic worldviews in which science seemed to be a natural associate of secularity.

The formation of the alleged ‘eternal’ antagonism between religion and science in the nineteenth century shows the constructive character of the notions involved:

The fundamental weakness of the conflict thesis is its tendency to portray science and religion as hypostatized forces, as entities in themselves. They should rather be seen as complex social activities involving different expressions of human concern, the same individuals often participating in both. (Brooke 1998: 56)

This does not imply that there is a fundamental consonance between scientific views and religious convictions which will emerge if one only finds the appropriate bridging concepts. Since the beginnings of early modern philosophy, conflicts, tensions, and dissonances have inevitably shaped the relations between religious views, theological reflection, ratio-empirical knowledge and philosophical thinking. But not all conflicts are destructive. They can also be an engine of change and of progress, both for religious perspectives (including theological reflection) and science-based views of reality.

What deserves more attention, however, is the fact that the thesis of a fundamental conflict between science and religion in the nineteenth century emerged together with antagonistic movements of an amalgamation of religion and science in the modern form of esotericism. The latter were inspired by the late-colonial global exchange of ideas and had profound repercussions for countries and cultures of the Global South (Bergunder 2016). Concepts of religion, science, (natural) theology, and esotericism (including a notion of ‘magic’ to re-enchant the modern world, see Milbank 2022) have a common history emerging out of the latter half of the nineteenth century, interlinked with colonialism, the politics of knowledge, and the formation of cultural self-awareness of non-Western vs. Western cultures.

3.2 Peaceful separation?

The response of many theologians in the nineteenth century was a methodological separation of science and religion, with religion held responsible for religious affections and ethical judgements guided by values and the sciences responsible for causal analysis in accordance with the laws of nature. Many were convinced that materialistic and deterministic worldviews were essentially incomplete and unable to account for the phenomena of life and human existence because they abandoned teleological thinking and left no room for aims and goals. In its liberal strand, modern Protestant theology sketched religion as inwardness and identified it with ‘the divine movement [of the mind] which works in a mysterious way in the unconscious depths of the universal human mind’ (Troeltsch 1913: 340; current author’s translation). Against the claims of naturalistic, monist worldviews it was argued that

a belief, which assumes nature and matter to be everything and which tries to deduce everything from it, is impossible, as is shown by the self-reliance of the spiritual world. The only question that we have to pose in relation to natural science is if the spiritual world with its oughts and cultural values is [recognized as] self-reliant and self-efficient with respect to nature; while in all other issues we can peacefully let science continue its ways. (Troeltsch 1913: 332–333, current author’s translation)

Higher criticism of the biblical texts and traditional doctrine was seen as an important deconstruction leading to an undogmatic, non-authoritarian, and individual Christianity.

From a very different viewpoint, the movement of dialectical theology in the beginning of the twentieth century came to a similar result. It claimed that what the Christian faith, guided by revelation, understands as creation has nothing to do with scientific investigations, as long as science refrains, as it should, from developing scientific theories into religion-like worldviews (Weltanschauungen): ‘theology can and must move freely where science which really is science, and not secretly a pagan Gnosis or religion, has its appointed limit’ (Barth 1958: 3).

The official Catholic reaction was different. The Roman magisterium refused to understand religion as an inward feeling or stance untouched by scientific investigation, and insisted that ‘faith is not a blind religious feeling’ (in the so-called Oath Against the Errors of Modernism, Denzinger 1955: 550). It is supported by science, as long as science is undertaken along the lines of a Neo-Thomist understanding of ontology and in accordance with rational and ethical laws of nature along the teaching of the church. Then, as the first Vatican council from 1871 stated, the one true creator God can ‘be known with certainty from those things which have been made, by the natural light of human reason’ (Denzinger 1955: 449). In any case, reason must not contradict religious faith, but, rightly understood, ‘demonstrates the basis of faith […] while faith frees and protects reason from errors and provides it with manifold knowledge’ (Denzinger 1955: 448). It was only with the encyclical ‘Humani Generis’ from 1950 and the Second Vatican Council that Catholic teaching came to a more open and self-critical relation to the natural sciences, including a rehabilitation of Galileo Galilei.

As another party in this discussion one can identify conservative confessional or evangelical theologies, which tried to limit scientific explanations to certain domains and subordinated them to their reading of the biblical texts. They disapproved of biblical criticism, defended traditional doctrines such as the historicity of a supernatural creation, the Fall and the Flood from Genesis, the virgin birth, and the deity of Christ, and argued against the decadence of Darwinism and socialism. A set of ninety essays in twelve volumes published between 1910 and 1915 called The Fundamentals: A Testimony to The Truth documents that frontline, in which science was enclosed within the boundaries of traditional Christian doctrine. These essays were distributed free of charge, and proved influential in the US fundamentalist movement to which they gave their name. Since then, debates about creationism, the authority of scripture and its significance for public life and morality, as well as the interplay between modern science, atheism, and left-wing politics have been part of public discourse in the US and beyond. In any case, what the different groups and issues illustrate is the fact that debates between religion, theology, and science from the 1850s until today do not represent a conflict between two clearly distinguishable domains named ‘science’ and ‘religion’ but are heated debates about different worldviews linked to cultural-political agendas, which usually involve faith-based positions on all sides.

3.3 Theology and science: developing the field since the 1960s

Apart from individual figures like French archaeologist and Jesuit Pierre Teilhard de Chardin (1881–1955) or German Protestant theologian Karl Heim (1874–1958), the first half of the twentieth century saw little constructive exchange between academic theology and the natural sciences. That slowly changed after the Second World War, when physicists and biologists started to approach theologians and initiated discourses on ethical questions (nuclear weapons and the responsibility of scientists, for example), as well as on epistemological and ontological issues linked to science. In the English-speaking world since the 1960s, ‘science and religion’ took on a more distinct nature as an academic discourse. A major starting-point was the book Issues in Science and Religion from 1966 written by Ian Barbour (1923–2013), an American physicist, philosopher, and process theologian. Barbour presented a history of science in relation to religion, reflected on methodological issues between science and religion, and developed his own integrative view of the matter inspired by process-philosophical thought. Also in 1966 the first issue of ZYGON was published in Chicago, to this day the most widely read specialist journal in the field. Its name is derived from the Greek word zygon meaning ‘yoke’, depicting the intention of the journal to harness science and religion together in order to pull more effectively, i.e. to have a significant positive impact on the modern world. ZYGON is sponsored by the Institute on Religion in an Age of Science (IRAS).

Barbour’s work was followed by, among others, British biochemist and theologian Arthur Peacocke’s (1924–2006) Creation and the World of Science (2004a). Peacocke identified himself as a panentheist (see below) and envisioned the cosmos as a creative process. He was an advocate for Theistic Evolution, the view that the Christian concept of creation is compatible and in fact must be synthesized with the modern view of the theory of evolution (Peacocke 2004b). Accordingly, Peacocke understood God as the first cause and immanent sustainer and guider of the process of evolution.

A third important voice was British physicist and theologian John Polkinghorne (1930–2021). One of his first books on the matter was One World from 1986 (Polkinghorne 1993). Polkinghorne tried to develop a comprehensive and intellectually satisfying view of reality including science and Christian religious perspectives, sometimes described by him as a ‘Theory of Everything’ (Polkinghorne 2000: 25). He, among others, engaged with issues of divine agency, the features of the cosmos, quantum theory, and the problem of evil. There was a fundamental accordance among these scientists as theologians (Polkinghorne 1996): (1) they presupposed an epistemological view with regard to both science and religion, called ‘critical realism’ (Losch 2009); (2) they understood reality in a non-reductive view as multi-layered with different levels of organization out of which new properties emerge; (3) they saw human nature as called towards a meaningful existence and therefore striving towards knowledge and orientation, both through science and religion; (4) they questioned traditional concepts of God as omniscient and omnipotent with regard to chance in nature and human free will. However, there were significant differences in their understanding of divine agency, a personal God, or the nature and meaning of Jesus Christ (Barbour 2010).

Since the 1980s a number of initiatives and journals evolved to foster constructive dialogue, including academic societies like the European Society for the Study of Science and Theology (ESSSAT, founded in 1988), or the International Society for Science and Religion (ISSR, founded in 2001), and journals like Theology and Science or Philosophy, Theology, and the Sciences (PTSc). Academic institutions dedicated to theology and science are the Ian Ramsey Centre in Oxford and the Center for Theology and the Natural Sciences (CTNS) in Berkeley, California. Since 1987, many projects in this field have been sponsored by the John Templeton Foundation, a philanthropic organization located in the US and dedicated to promoting harmony between scientific and spiritual knowledge.

In other countries and different academic and religious settings, however, the formation of discourses on theology and science took different paths. Celebrating its fiftieth year of publication, the 2015 volume of ZYGON contained essays presenting the developments of religion and science around the world, from different European settings to the US, China, Japan, and South Africa, to the Islamic world and Indian traditions, thus presenting a rich prospect of concepts, approaches, and debates. In Germany, for example, epistemology was shaped by Neo-Kantian, phenomenological, or post-structuralist epistemologies rendering critical realism implausible, as well as any natural theology built on scientific knowledge. Academic theology in Germany is mainly related to cultural studies, and, with notable exceptions like Wolfhart Pannenberg (1928–2014), Jürgen Moltmann (1926–) and Michael Welker (1947–), it has not been engaged intensively in theology and science (Evers 2015).

4 Systematic perspectives on the field

4.1 Barbour’s fourfold typology

Philosophers of science like Thomas S. Kuhn (1922–1996) and Michael Polanyi (1891–1976) have argued since the early 1960s that science does not straightforwardly discover and present objective truths about reality by accumulating more and more accepted facts and theories. Rather, science is in itself a dynamic process working within historical parameters, operating with specific practices and assumptions according to internal standards. These insights opened new opportunities for the science-theology dialogue. Ian Barbour’s seminal work Issues in Science and Religion (1966b) is praised as having established the science and religion field (see above). It discussed methods of scientific discovery and developed the concept of critical realism, which rests on a theory of truth affirming correspondence between theories and factual reality. It thus repudiates instrumentalism while at the same time acknowledging the limitations of scientific descriptions, avoiding a naïve scientific realism suggesting an immediate access to reality (Barbour 1966a: 29). Limits, for example, lie in the symbolic character of the models used:

One has to use models, but one has to recognize their limitations; one has to realize that they are partial and limited, that each one selects certain aspects and emphasizes those, that none of them corresponds exactly in any simple way to reality. (Barbour 1966a: 30)

Later, in his 1990 Religion in an Age of Science, Barbour presented his widely-debated fourfold typology of relating science and religion (Berg 2004): conflict, independence, dialogue, and integration. Barbour described scientific materialism and biblicism as the two opposing extremes under the conflict model. Members of both camps promote absolute truth claims, derived either from science or from religion (biblical texts), and do not accept truth claims based on arguments from the other field. For him this explained the ideological and political nature of the conflicts between those groups, as experienced today in heated debates about creationist views, or anti-religious naturalism like that of Richard Dawkins who sees religions as being trapped in a ‘God delusion’ (Dawkins 2006). Barbour saw no value in the conflict model and intended to overcome sterile, ideological confrontation.

In Barbour’s view the second model, independence, is aware of the fruitlessness of the conflict model and therefore develops an understanding of religion and science as two categorically different endeavours working with very different methods and using independent language games in order to make sense of different aspects of reality. Barbour subsumed a variety of approaches under this label, like dialectical theology, theological existentialism such as that of Rudolf Bultmann, and philosophers and theologians like George Lindbeck, in the succession of Wittgenstein, who see science and religion as different language games. Although not listed by Barbour, because he developed his model shortly after Barbour’s book was published, Stephen Jay Gould’s idea of ‘non-overlapping magisteria’ (NOMA) falls into this category as well (see Gould 1997; 1999). Gould understood science as investigating the natural world of objects in space and time, while religion shapes a moral world:

Science tries to document the factual character of the natural world, and to develop theories that coordinate and explain these facts. Religion, on the other hand, operates in the equally important, but utterly different, realm of human purposes, meanings, and values – subjects that the factual domain of science might illuminate, but can never resolve. (Gould 1999: 4)

Insofar as these realms of competence do not overlap, any serious conflict is excluded, and no unification or synthesis ‘under any common scheme of explanation or analysis’ (Gould 1999: 4) is possible.

Barbour valued the independence model as a first step into the right direction, because it respected the specific character of each discipline. But for him this model did not allow for constructive dialogue and mutual enrichment. Reality is not divided into different compartments of human activity and understanding, but presents itself as an organic whole, and God as creator and Lord of life must not be confined to a religious sphere. That would also prevent religion from addressing the pressing issues of ecology. Therefore, Barbour wanted to go beyond the independence model and advocated a dialogue model combined with a certain version of integration. Barbour identified common historical roots of both science and Christianity, and he saw common ground in how both disciplines debate boundary questions, methodology, and the spiritual dimension of nature. The integration model, then, goes beyond identifying certain issues of common interest and concern, and tries to work towards an ultimate integration of science and religion, resulting in a comprehensive worldview. For Barbour, Teilhard de Chardin was an impressive representative of this model, providing a synthesis of scientific and religious concepts within the framework of a theology of nature. However, in order to do more justice to specific religious concerns, Barbour himself argued for a prudent form of process philosophy in the tradition of Alfred North Whitehead. With all this he wanted to do justice to valid concerns of the independence model, while practicing dialogue with regard to methodological issues and aiming at integration with respect to notions of creation and anthropology.

Barbour seemed to imply that three of his four models – namely independence, dialogue, and integration – are compatible with each other. Others have understood them as mutually exclusive (Stenmark 2013: 2310). However, the fact that two disciplines are methodologically independent does not necessarily imply that – on another level – they cannot relate to each other, be it in dialogue or as parts of an integrative view of reality. It seems obvious that Barbour understood his four models to be ordered according to their strength in bringing religion and science together – from conflict through independence to dialogue and integration. That is underpinned by a historical trajectory that implies progress moving from conflict, as illustrated by the case of Galileo towards integration in contemporary debates.

However, since religion, theology, and science involve complex interplays of different perspectives, interpretations, traditions, and concerns, conflict seems inevitable. Even within science and within theology conflicts can have a productive function, if one thinks of the conflict between Leibniz and Clarke/Newton (Leibniz, Clarke, and Newton 1998) or Einstein and Bohr (Marage and Wallenborn 1999). The main concern might be to make conflict productive, not to eliminate it. The issue is to appropriately identify the real sources and levels of conflict, to avoid idle dispute rooted in misunderstanding, to employ suitable ways of potentially solving superficial issues, and to address hermeneutical shortcomings in order to start a fruitful exchange over the subject matter. Since science has not left our views of reality untouched – and since it provides powerful technological means for practical purposes – traditional forms and convictions of religion and its practices, which were deeply embedded in pre-modern cosmologies and anthropologies, are also affected. Still, there will often remain a plurality of defensible options of how to deal with certain conflictual constellations.

4.2 Natural theology or theology of nature

Since the beginning of philosophical reflective thinking, religious concepts and convictions as expressed in narratives and myths and derived from sacred books were related to other forms of knowledge of the divine independent of religious sources. The Stoic tradition distinguished between three species of theology: mythical, natural, and political. Mythical theology referred to fabulous narrations about deities, which were rejected as popular superstition. Political or civil theology was related to the set of rules regulating worship in the official public cult. Natural theology, however, was the theology of the philosophers who reflected on the true essence of the divine, criticizing both mythical and civil theology. This notion of natural philosophical theology was taken up in Christianity and passed on since Augustine down to the debates in early modernity. It was distinguished from revealed theology as found in scripture. Since Augustine, the relationship between natural and revealed theology was expressed in the metaphor of the two books, the book of nature and the book of scripture, the two sources of knowledge about God, but the term natural theology was only occasionally used. Thomas Aquinas preferred to refer to the natural light of the intellect, which is capable of developing fundamental notions of an entity that everybody named God. The reformers of the sixteenth century, however, considered any natural knowledge of God as fundamentally blurred and soteriologically ambivalent; it is scripture which provides the means to regain access to the original meaning of the book of nature.

Starting with Spinoza, thinkers in early modernity began to argue for a consequential separation of philosophy and theology. More and more, natural theology became a discipline within philosophy, independent of Christian notions and the context of salvation. Francis Bacon defined natural theology cognitively as ‘knowledge or rudiment of knowledge concerning God which may be obtained by the contemplation of his creatures’ (Bacon 1862: 211–212). The significance and plausibility of such knowledge, however, remained disputed. In the seventeenth century, two pathways of natural theology can be distinguished. One comprised different varieties of physico-theologies, most popular in England (John Ray, William Derham), the Netherlands, Germany, and Switzerland, drawing on the empirical study of nature and developing arguments of design. The other comprised all forms of rational theology. Both strands were fundamentally criticized by David Hume on epistemological grounds, and by Immanuel Kant with transcendental arguments; there is no certainty of rational knowledge beyond the realm of experience. German Protestant theology since Friedrich Schleiermacher (1768–1834) followed the Kantian criticism and insisted on the independence of religion from theoretical reason or empirical knowledge. Religion rests neither on sense data nor pure reasoning, but is deeply rooted in human nature. According to Schleiermacher, religion is born out of the feeling of absolute dependence, in which our self-consciousness in general represents the finitude of our being’, and it ‘is therefore […] a universal element of life; and the recognition of this fact entirely takes the place […] of all the so-called proofs of the existence of God’ (Schleiermacher 2016: 133–134).

All in all, ‘in nineteenth century German Protestantism there existed no natural theology anymore; the term appeared only in historical references’ (Sparn 1994: 90, current author’s translation). However, Karl Barth’s ‘Nein!’ to natural theology in the beginning of the twentieth century identified the attempts of nineteenth century theology as nothing but an actualized version of natural theology, trying to build Christian thinking on arbitrary anthropological concepts and thus preventing the living word of the living God from being heard. An alternative to Barth’s theology of revelation was Paul Tillich’s method of correlation, which confined natural theology to philosophical analysis of existence as ‘the question of reason about its own ground and abyss’, which is ‘asked by reason, but reason cannot answer it’ (Tillich 1951: 120), and to which theology and revelation related with appropriate accounts of religious responses based on revelation. Debates in the 1970s, however, led to the consensus that ‘natural theology’ addressed a profound theological issue, although the ways of traditional natural theology are usually considered obsolete. Wolfhart Pannenberg had argued for a renewal of natural theology not as a foundation for theology but as an explanation of the factual human potential for experiencing and conceiving the divine, and for apologetic means (Pannenberg 2001a: 73–118). Others understood natural theology as a reminder that Christian theology is called to translate and interpret its doctrine as universally significant and in relation to human existence in general.

In England, however, natural theology enjoyed a better reputation. This is reflected in the importance that is given to public discourse on these matters, as in the Boyle Lectures, established in 1692 and revived in 2004 after abeyance from 1903, or the Gifford Lectures, established in 1887 with the intention of ‘“Promoting, Advancing, Teaching, and Diffusing the study of Natural Theology,” in the widest sense of that term […] without reference to or reliance upon any supposed special exceptional or so-called miraculous revelation’ (Gifford 2022). While special doctrine based on revelation is seen as a constant source of division and dispute, the natural knowledge of God is understood as ‘the true foundations of all ethics and morals, […] the means of man’s highest well-being, and the security of his upward progress’ (Gifford 2022).

The standard textbook for this kind of natural theology was the 1802 work of English clergyman William Paley, Natural Theology or Evidences of the Existence and Attributes of the Deity, which was republished many times throughout the nineteenth century. The young Charles Darwin studied two other texts by Paley during his time as student of theology, and later carefully read Paley’s Natural Theology, but in his autobiography concluded that the ‘old argument of design in nature, as given by Paley, which formerly seemed to me so conclusive, fails, now that the law of natural selection has been discovered’ (Darwin 1958: 87). However, some historians of science assume that the biggest blow to natural theology ‘was not necessarily theological liberalism or scientific positivism, rather, it was the moral devastation of the First World War’ (Eddy 2013: 113–114), which marked the end of an optimistic worldview built on general and natural knowledge of divine purpose and benevolent guidance of reality, and finally led to the present situation of increased plurality and choice in a postmodern world. Contemporary versions of natural theology, like that of Alister McGrath’s Scientific Theology (2017), therefore argue rather for a theology of nature, which must not and cannot develop ‘persuasive grounds of faith outside the bounds and scope of revelation, but is rather a demonstration that, when the natural world is “seen” through the lens of the Christian revelation, the outcome is imaginatively compelling’ (McGrath 2017: 131).

4.3 Frameworks: theism; panentheism; naturalism/immanentism/pantheism; cultural studies and hermeneutics

These considerations suggest that, instead of relating two allegedly independent bodies of knowledge named science and theology, it might be more helpful to distinguish between different approaches towards reality and its potential relation to the divine. We can then distinguish between different frameworks in which God and reality are envisioned. In a theistic framework, God is understood as a transcendent supreme being, fundamentally distinct from reality in space and time. In such a framework, questions of how to relate the divine to the process of reality become crucial. In early modernity, beginning with Baruch Spinoza, pantheistic frameworks were developed in which nature or reality was identified with the divine. A religious view, then, is nothing but an understanding of nature as having a deeper dimension out of which it develops (natura naturans as the original creative substance of nature brings forth natura naturata as the phenomenal world of our experiences). In Christian theology, Schleiermacher tended towards such a model, as when he understood ‘divine causality as equivalent in compass to the sum-total of the natural order’ (Schleiermacher 2016: 201). Today, proponents of what is called religious naturalism interpret nature as sacred and deny any supernatural entity or agent and thus fall into the pantheistic camp as well (Wildman 2014).

Many Christian theologians today, however, as well as philosophers of religion, promote a middle way called panentheism, according to which the world is in God, but God is more than the world (Hartshorne 1978: 90). This is the view of process theology, which understands God as having absolute and eternal aspects that distinguish the divine from created reality, as well as relative, historical, and responsive aspects that intimately entangle the divine with the process of creation. For Whitehead, God has no significance apart from the process of the cosmos, and God’s very nature requires a world to relate to in activity and response. ‘God is the aboriginal instance of […] creativity’, thus expressing the creativity of reality itself, and there is ‘no meaning to “God” apart from the “creativity” and the “temporal creatures,” and no meaning to the “temporal creatures” apart from “creativity” and “God”’ (Whitehead 1978: 225).

However, all of these are variations of an ontological approach considering the divine in connection with or as an aspect of the physical world as such. If one does not share the presupposition that such an approach to reality is possible, even within critical limits, then theology and science become human enterprises of understanding and interpretation that can only be explored from within, in a perspective of participation rather than of observation. Hermeneutics and cultural studies focus not on scientific knowledge or religious truth claims but on method, intentions, language, allegations, interest, power structures, hierarchy, model-building, and the like, thus questioning the claim that universal, essentialist, and objective features of reality are accessible at all. However, a reduction of science to cultural discourse does not do justice to the rigour of scientific methods and the extraordinary success of many scientific theories, not only in the general laws of physics but also, for example, in synthesizing a vaccine against a virus on the basis of its genome. Only such an account of science ‘that does not make the success of science a miracle’ (Putnam 1975: 73) can be appropriate. Any interpretation that ascribes scientific knowledge a marginal, coincidental function seems to be paternalistic in the sense that it reduces science to something which is at odds with its self-understanding, standards, history, and effects.

Thus, in the present author’s understanding, any relevant discourse about what we today identify as theology and science must take into account the interplay of three independent but interrelated dimensions: objective knowledge claims in a third-person perspective and their justification, in relation to which Putnam’s No-Miracle Argument seems valid; subjective, individual, and existential intentions and concerns of the dialogue partners involved in a first-person perspective; and anthropological considerations of human existence and its moral nature, human cultural traditions (including historical religions), political constellations, and hermeneutical frameworks, which one can address in second-person contexts. Any meaningful discourse will take into account all three dimensions as related yet irreducible, so that their triangular interplay cannot be reduced to binary relations – even if, for a certain purpose, one focuses on facts, on existential meaning, or on historical dynamics in particular.

5 Relevant issues between theology and science

The following subsections will address four significant questions which are debated in current theology-and-science discourses. They refer to central issues at stake: in 5.1, the relation between notions of creation and scientific cosmologies; in 5.2, the possibility of divine action in nature, which seems to be in tension with scientific concepts of causality; in 5.3, the question whether or not the theory of evolution rules out a purposeful design of creatures by a benevolent creator; and in 5.4, the question how an explanation of religious concepts and behaviour from the perspective of cognitive science within an evolutionary framework might affect the epistemic status of religious beliefs. Any deeper inquiry into these issues, however, would require an extensive discussion of the literature in the respective field, which cannot be achieved is this overview.

5.1 Cosmology and creation

A central and recurrent theme in discourses between science, religion, and theology is the question of whether or not our universe, as it is presented in a scientific perspective, must or can be explained as intentionally created by a Creator God. This is linked to a number of issues. One concerns the question of whether our universe is eternal or has a contingent origin. An eternal world not only contradicts biblical creation accounts but also raises the question of how a creative act could be seen as its origin, if there is no beginning in time. On the other hand, ‘creation’ as a theological notion does not necessarily imply an initial causal act but denotes a principal origin, the transition from non-being into being. Augustine suggested that this implies the origin of time itself. Immanuel Kant, on the other hand, argued that reasoning cannot answer the question of a beginning on its own grounds. The idea of an eternal universe is incomprehensible because our notion of causality is trapped in an infinite regress, while an original beginning cannot answer the question what was before such a beginning.

Since the middle of the twentieth century, these issues are debated in connection with cosmological models based on Einstein’s general theory of relativity and strongly corroborated by empirical evidence such as the cosmic background radiation, which suggests that there was indeed a primordial state of our cosmos as we observe it. This state, which even in a scientific description might imply an absolute beginning of physical time, is usually referred to as the Big Bang. In addition, more or less right from the beginning the expanding cosmos is characterized by a set of fundamental natural constants such as the strength of the electric, magnetic, and gravitational forces which have to fall into very small margins in order to bring about a cosmos in which life (as we know it) is possible. This is often referred to as the fine-tuning of the universe. By some, this fact of a contingent and well-tempered beginning has been regarded as supportive of the idea of an intelligent creator. Others point to the fact that such a designer of the primordial state of the cosmos is not an explanation but a cover-up of a missing explanation. Alternative models of explanation seem possible, which assume a multitude of universes and an understanding of chance within the limits of boundary conditions, which allows for the unintentional formation of universes (Hawking and Mlodinow 2012). However, suffering and evil within a largely void and transitory cosmos have been regarded as an indication that our universe does not point towards an intelligent and purposeful creator God.

This whole discourse challenges any naïve theological notion of creation as just the initial designing and manufacturing of the physical world. Lately, theological concepts of creation have employed notions of process philosophy, a panentheistic notion of God, and of an ongoing creation (traditionally called creatio continua), which allow for a responsive relationship between a creator God and creation (see for example Moltmann 2005; Pannenberg 2001b; or the anthology Clayton and Peacocke 2004). Such views can take up the traditional concept of divine providence by which the creator accompanies, guides, influences, and directs creation. That concept seems to distinguish deism from theism, because in a deist’s view God created physical reality in the beginning, got it started, and then left it all to natural causes. But any notion of continuous divine agency raises the question how and on which levels such a Creator God can interfere with physical creation. For some time, this was an extensively debated issue within the science and religion discourse, as documented by the so-called Divine Action Project, a long-term collaboration between the Center for Theology and the Natural Sciences and the Vatican Observatory (see the assessment of this project in Shults, Murphy, and Russell 2009).

5.2 Does God act in nature?

Wonders, miracles, and sacramental rituals have a central place in religious traditions. On a cognitive level, they are often referred to as support for or even proof of God’s existence. There are manifold narratives about God’s supernatural interventions in scripture, and they seem to be necessary in order to testify to the power and existence of a supernatural God and to make petitionary prayer (as well as sacraments) meaningful. However, in the early Enlightenment period the concept of miracles was put under scrutiny and related to the apparent immutability of natural laws. Divine action seemed to imply a (supernatural?) ‘violation of the laws of nature’ (Hume 1902: 114). But this raised the question if, and under which conditions, such a violation could be corroborated, since according to Hume there are only rare and dubious accounts of miracles, while scientific investigation pictures a reality fully determined by natural law – a view which today is called the causal closure of nature. It also implies that special divine agency is at work only occasionally, leaving the world take its own course in most cases. However, others have pointed to the fact that, according to standard statistical accounts, the probability of any singular event is close to zero in relation to the totality of all other events, and Hume’s definition of a miracle might have ruled out any nonzero probability of miracles right from the start (see Earman 2000; for a more extensive discussion see Saunders 2002).

Philosophers and theologians have therefore questioned the idea that God intervenes in nature, also because God would then act against the course of things which he as creator has set in motion. Some refer to scholastic concepts, such as the distinction between God as the primary cause of all reality and secondary causes as the causal power of creatures granted by God’s primary causality: ‘Therefore God is the cause of every action, inasmuch as every agent is an instrument of the divine power operating’ (ST q. 3, a. 7; Aquinas 1932). However, Aquinas’ notion of cause is shaped by the Aristotelian concept of four causes (comprising material, formal, and efficient, as well as final causes), and it is hardly compatible with the modern notion of natural law. With reference to scientific laws it is by no means clear how exactly one would coordinate the causal agency of God and creatures in such a concept of double agency (Ritchie 2017).

Others have pointed to the fact that modern physics has abandoned a causally closed concept of nature, but is employing non-deterministic laws of nature that allow for a non-interventional divine agency in, with, and under natural processes. Prominent candidates for such domains of non-interventional influences for divine providence are the fundamental quantum level of physical reality, or notions of complex systems, including so-called chaos theory. The argument rests on the idea that divine providence might use this supposed elasticity of natural processes to act without disruptive intervention in the regularities of nature. From this general providential agency some still want to distinguish what tradition called special providence, which comprises interventional miracles and divine actions which actually abrogate natural causality. However, given the evil that we encounter in this world, this does not answer the question of why God would act on some occasions – for example as narrated in the biblical miracle stories – rather than on others. In any case, some notion of providence, albeit a modest one, seems to be an indispensable aspect of Christian faith (Fergusson 2010).

5.3 Evolution and creation

The question of teleology, that is the idea that there is a purpose for creation, is linked to another central issue of discourse between science and religion, namely the interpretation of evolution by means of natural selection (De Smedt and De Cruz 2020). When Darwin presented his theory in 1859, the main issue was not its discrepancy with the biblical narratives. Biblical exegesis had long pointed to the fact that a literal understanding of the biblical narratives is incompatible with geological and historical timescales, and the mythological nature of the biblical narratives had been thoroughly established – at least in academic theology. The main challenges were the means of natural selection, which according to Darwin are the only and sufficient causes that account for the diversity of life-forms on this planet, including mankind. This was seen as incoherent with the Christian (as well as any other religious) view of God’s providential ordering of nature. Or, as the German theologian and religionist Rudolf Otto stated, Darwin’s theory ‘only becomes definitely anti-theological because it is anti-teleological’ (Otto 1913: 140).

In public debates of Darwin’s time, Darwinism was attractive for all who wanted to establish a naturalistic understanding of creation independent from ecclesiastical theological doctrine, whether in a liberal Christian, general religious, or purely naturalistic understanding. An example is David Friedrich Strauss’ last work from 1872, translated into English a year later as The Old Faith and the New (1873), which raised nearly as great a scandal as his book on the life of Jesus. Strauss finally abandoned Christianity altogether and gave a critique of theism, employing Darwin’s theory as a replacement for the Christian creation account, including an evolutionary view on hominization as the modern scientific understanding of the creation of human beings. This naturalistic worldview was linked to progressive political democratic claims. Thus in public discourse Darwinism was seen as supportive of a new view on humanity, integrating it into the ongoing process of the natural world, while theologians and representatives of the church protested against the affront to human dignity and the alleged dissolution of morality and social order as well as all higher purposes of human life. Debates about evolution and its religious challenges, from Darwin’s time until today’s debates about creationism, have always pivoted around conflicts between competing moral and political worldviews.

By challenging teleology, divine purpose, and design, evolution was also seen as challenging human origins and traditional views of humanity as created in the image of God, as well as the doctrine of original sin and the fall as ascribed to the first human couple Adam and Eve. The historicity of Adam and Eve, as well as the disruption of creation by a historical fall, became questioned. Ever since, concepts of original sin as a stage in a process of moral awakening or emancipation from the immediate embeddedness in the natural world rather than a historical event gained prominence in theology. Examples are Schleiermacher’s concept of original sin as a social category, or Tillich’s description of the fall. Other theologians try to adapt traditional notions without fully abandoning them (van den Brink 2020). However, an integration of humanity into the whole of creation in an evolutionary perspective is appreciated by others as transcending a sharp distinction between humans and the rest of creation. This allows for a broader view of God’s creative purpose, including an understanding of redemption comprising all living creatures (see Niels Gregersen’s concept of Deep Incarnation, Gregersen 2015). This view contributes to a theological notion of creation as different from purposeful manufacturing: the creation of all forms of living beings is made to develop itself. This allows for new arguments for an evolutionary theodicy when, for example, Christopher Southgate in the face of natural evil arising in evolution develops an only-way argument: evolution is ‘the only, or at least the best, process by which creaturely values of beauty, diversity, and sophistication could arise’ (Southgate 2008: 48).

5.4 Cognitive science of religion

In recent decades a new scientific approach to religion emerged, which focused on the origins of religiosity from the perspective of human cognitive capacities. This has become known as the Cognitive Science of Religion (CSR), and has been extensively discussed in the science-theology field (White 2021). However, one must also understand that only certain branches and figures of such research participate in this dialogue, while others promote it as a shift in the fields of religious studies away from the dominance of methods of cultural, sociological, and political hermeneutics. Theological debate of issues related to such research concentrates on certain key questions. One is the question of whether, and in which sense, humans are naturally religious. Sometimes this is linked to classical theological concepts such as natural religion, John Calvin’s sensus divinitatis (sense of deity), or Schleiermacher’s feeling of ultimate dependency, and some have argued that it might be possible to develop a natural theology based on a natural human receptivity for the divine (Watts 2013: 483). However, CSR tends to explain religious beliefs and practices as products of normal, everyday cognitive processes and not as founded in some special faculties for religion – whatever that could mean. That leaves room for critical interpretations which understand religious beliefs and practices as unsupported by-products of cognitive processes otherwise functional in mundane situations. A much-discussed concept is that of a cognitive agency detection device, which becomes hyperactive and then produces personal concepts of spiritual and divine beings (Barrett 2000: 31). This can be understood as an idle and illusionary malfunction or as an amplification of human psychological propensities, even as an outward projection of the collective human mind, as Ludwig Feuerbach had concluded (Feuerbach 1854). Others see it as consistent with a creator who arranged for the development of receptivity to make authentic religious experience possible. In any case, there is some agreement even between those who argue for and against religion that religion might be natural, whatever its value may be (see the discussion in van Eyghen 2020). On the other hand, there are processes of an erosion of religion and religious experiences in certain Western societies that are very profound and presumed to be irreversible. Charles Taylor, for example, has argued that religious indifference or a fully secular option is one of the major historical achievements of modernity, when for the first time in history it is culturally possible, socially acceptable, and intellectually satisfying for significant parts of Western societies to live without reference to any transcendent power and within an impersonal order, to which the individual has nothing but an instrumental relation (Taylor 2007).

That leads to a second key question in these debates, concerning the relation of a natural propensity towards religious experiences, including beliefs and rituals, to institutional forms of religion and theological doctrine, as well as scientific research itself. In recent years, the naturalness of religion has been contrasted with the unnaturalness of both theology and science (see Barrett 2012 and McCauley 2013). From its beginnings in the New Testament, Christian theology has always expressed the need to critically and constructively reflect on spiritual concepts of reality, which we today identify as religious. Faith in God in a Christian sense strives towards understanding (cf. Anselm’s famous formula fides quaerens intellectum, faith seeking understanding) and has always been not the straightforward affirmation or cultivation of natural religious experiences but a reflective life-form shaped by God’s revelation, critical of what seems self-evident in a religious perspective. Thus Karl Barth had insisted ‘that every genuine proclamation of the Christian faith is a force disturbing to, even destructive of, the advance of religion, its life and richness and peace’ (Barth 1957: 444). Reference to the alleged benefits of a religious mindset or religious practices might miss the point of religious faith, turning it into a spiritual policy or coping strategy. From an emic perspective, from within many religions, this approach would not be believing in God, but believing in the best thing for oneself.

6 Conclusion

As has been pointed out repeatedly, there is no perennial fundamental conflict between theology and science, rather a conflictual history of human believing and reflecting, rich with more or less productive tensions. Issues of conflict, congruence, or affirmation change according to the settings of the relevant discourses. With modern scientific theories and their consolidation, especially in the nineteenth century, certain issues have come to the front that shape the debate to this very day. However, theology should realize that religion, faith, and revelation must not be separated from facts and empirical evidence but must provide interpretative perspectives of reality. At the same time, science naturally tends towards shaping belief systems and provides deep insights into the structure of the world we live in. While it is important not to blur methodological and hermeneutical differences between theology and science, it is equally important to find levels of exchange for the benefit of theology, science, and belief systems.

This is relevant not only for any discourse with regards to understanding reality, but even more so for discourse on existential and ethical questions which involve science and religion. The question of theodicy with special reference to what the tradition has named natural evil, for example, incorporates factual and theoretical scientific knowledge, existential concerns, and communal contexts of social strategies of coping with evil. Present debates on climate change analogously involve scientific theories and models about human induced climate change, human concerns of what a good life means (including the appreciation of the dignity of non-human creatures), and political decision-making and social dynamics. These three dimensions are all involved, but at the same time the theological reflections of religious traditions, against the background of a modern understanding of reality as shaped and informed by science, bring about recurrent issues and challenges for religious views of reality. In the case of Christian theology, it is faith seeking understanding that drives theological reflection to develop an understanding of reality as created, guided, redeemed, and finally consummated and perfected by God, as disclosed by the teaching, life, and fate of Jesus Christ.

Attributions

Copyright Dirk Evers ORCID logo (CC BY-NC)

Bibliography

  • Further reading

    • Deane-Drummond, Celia E., and Agustìn Fuentes (eds). 2020. Theology and Evolutionary Anthropology: Dialogues in Wisdom, Humility, and Grace. Abingdon/New York: Routledge.
    • Dixon, Thomas, and Adam Shapiro. 2022. Science and Religion: A Very Short Introduction. Oxford: Oxford University Press. 2nd edition.
    • Harrison, Peter. 2015. The Territories of Science and Religion. Chicago: University of Chicago Press.
    • McGrath, Alister. 2020. Science and Religion: A New Introduction. Oxford: Wiley-Blackwell. 3rd edition.
    • Perry, John, and Joanna Leidenhag (eds). 2021. ‘Themed Issue: Science-Engaged Theology’, Modern Theology 37: 243–563.
    • Southgate, Christopher (ed.). 2011. God, Humanity and the Cosmos: A Textbook in Science and Religion. London: T&T Clark. 3rd edition.
  • Works cited

    • Aquinas, Thomas. 1932. On the Power of God (Quaestiones Disputatae de Potentia Dei). London: Burns Oates & Washbourne.
    • Augustine. 1968. City of God, Volume III: Books 8-11. Loeb Classical Library 413. Translated by David S. Wiesen. Cambridge, MA: Harvard University Press.
    • Bacon, Francis. 1862. ‘Advancement of Learning’, in The Works of Francis Bacon. Volume 6. Edited by James Spedding, Douglas D. Heath, and Robert Leslie Ellis. Boston, MA: Houghton Mifflin, 77–412.
    • Barbour, Ian G. 1966a. ‘Commentary on Theological Resources from the Physical Sciences’, Zygon 1, no. 1: 27–30.
    • Barbour, Ian G. 1966b. Issues in Science and Religion. London: SCM Press.
    • Barbour, Ian G. 2010. ‘John Polkinghorne on Three Scientist-Theologians’, Theology and Science 8, no. 3: 247–264.
    • Barrett, Justin L. 2000. ‘Exploring the Natural Foundations of Religion’, Trends in Cognitive Sciences 4, no. 1: 29–34.
    • Barrett, Justin L. 2012. ‘The Naturalness of Religion and the Unnaturalness of Theology’, in Is Religion Natural? Issues in Science and Theology 7. Edited by Dirk Evers, Antje Jackelén, and Michael Fuller. London: Continuum International Publishing, 1–23.
    • Barth, Karl. 1957. Church Dogmatics: Volume II/1. Edited by Geoffrey W. Bromiley and Thomas F. Torrance. Edinburgh: T&T Clark.
    • Barth, Karl. 1958. Church Dogmatics: Volume III/1. Edited by Geoffrey W. Bromiley and Thomas F. Torrance. Edinburgh: T&T Clark.
    • Berg, Christian. 2004. ‘Barbour’s Way(s) of Relating Science and Theology’, in Fifty Years in Science and Religion: Ian G. Barbour and His Legacy. Edited by Robert J. Russell. Ashgate: Aldershot, 61–75.
    • Bergunder, Michael. 2016. ‘“Religion” and “Science” Within a Global Religious History’, Aries – Journal for the Study of Western Esotericism 16: 86–141.
    • Brooke, John Hedley. 1998. Science and Religion: Some Historical Perspectives. Cambridge: Cambridge University Press.
    • Clayton, Philip, and Arthur Robert Peacocke (eds). 2004. In Whom We Live and Move and Have Our Being: Panentheistic Reflections on God’s Presence in a Scientific World. Grand Rapids: William B. Eerdmans.
    • Darwin, Charles. 1958. The Autobiography of Charles Darwin 1809-1882. With the Original Omissions Restored. Edited by Nora Barlow. London: Collins.
    • Dawkins, Richard. 2006. The God Delusion. Boston: Houghton Mifflin.
    • De Smedt, Johan, and Helen De Cruz. 2020. The Challenge of Evolution to Religion. Cambridge Elements Cambridge/New York: Cambridge University Press.
    • Denzinger, Henry. 1955. The Sources of Catholic Dogma. Fitzwilliam, NH: Loreto Publications.
    • Draper, John William. 1874. History of the Conflict Between Religion and Science. New York: D. Appleton.
    • Earman, John. 2000. Hume’s Abject Failure: The Argument Against Miracles. Oxford: Oxford University Press.
    • Eddy, Matthew D. 2013. ‘Nineteenth-Century Natural Theology’, in The Oxford Handbook of Natural Theology. Edited by Russell Re Manning. Oxford: Oxford University Press, 100–117.
    • Eisenstadt, Shmuel N. (ed.). 1986. The Origins and Diversity of Axial Age Civilization. Albany, NY: State University of New York Press.
    • Evers, Dirk. 2015. ‘Religion and Science in Germany’, Zygon 50, no. 2: 503–533.
    • Fergusson, David. 2010. ‘The Theology of Providence’, Theology Today 67, no. 3: 261–278.
    • Feuerbach, Ludwig. 1854. The Essence of Christianity. London: John Chapman.
    • Gifford, Adam. 2022. ‘Lord Adam Gifford’s Will’, https://www.giffordlectures.org/lord-gifford/will
    • Gould, Stephen Jay. 1997. ‘Nonoverlapping Magisteria’, Natural History 106: 16–22; 60–62.
    • Gould, Stephen Jay. 1999. Rocks of Ages: Science and Religion in the Fullness of Life. New York: Ballantine Books.
    • Gregersen, Niels Henrik. 2015. Incarnation: On the Scope and Depth of Christology. Minneapolis, MN: Fortress Press.
    • Harrison, Peter. 1998. The Bible, Protestantism, and the Rise of Natural Science. Cambridge/New York: Cambridge University Press.
    • Hartshorne, Charles. 1978. The Divine Relativity: A Social Conception of God. New Haven, CT: Yale University Press. 7th edition.
    • Hawking, Stephen, and Leonard Mlodinow. 2012. The Grand Design. New York: Bantam Books.
    • Hume, David. 1902. ‘An Enquiry Concerning Human Understanding’, in Enquiries Concerning Human Understanding and Concerning the Principles of Morals. Edited by Lewis A. Selby-Bigge. Oxford: Clarendon Press, 5–165. 2nd edition.
    • Irving, Washington. 1828. A History of the Life and Voyages of Christopher Columbus. 4 vols. London: John Murray.
    • Jaspers, Karl. 2017. Vom Ursprung und Ziel der Geschichte. Karl Jaspers Gesamtausgabe 1. Edited by Kurt Salamun. Basel: Schwabe.
    • Joas, Hans, and Robert N. Bellah (eds). 2012. The Axial Age and Its Consequences. Cambridge, MA: Belknap Press of Harvard University Press.
    • Leibniz, Gottfried Wilhelm, Samuel Clarke, and Isaac Newton. 1998. The Leibniz-Clarke Correspondence: With Extracts from Newton’s ‘Principia’ and ‘Optiks’. Edited by Hubert G. Alexander. Manchester: Manchester University Press.
    • Lindberg, David C., and Ronald L. Numbers. 1987. ‘Beyond War and Peace: A Reappraisal of the Encounter Between Christianity and Science’, Perspectives on Science and Christian Faith 39, no. 3: 140–149.
    • Losch, Andreas. 2009. ‘On the Origins of Critical Realism’, Theology and Science 7, no. 1: 85–106.
    • Marage, Pierre, and Grégoire Wallenborn. 1999. ‘The Debate Between Einstein and Bohr, or How to Interpret Quantum Mechanics’, in The Solvay Councils and the Birth of Modern Physics. Edited by Pierre Marage and Grégoire Wallenborn. Basel: Birkhäuser, 161–174.
    • McCauley, Robert N. 2013. Why Religion Is Natural and Science Is Not. New York: Oxford University Press.
    • McGrath, Alister E. 2017. Re-Imagining Nature: The Promise of a Christian Natural Theology. Chichester: Wiley Blackwell.
    • Milbank, John. 2022. ‘Religion, Science and Magic: Rewriting the Agenda’, in After Science and Religion: Fresh Perspectives from Philosophy and Theology. Edited by Peter Harrison, John Milbank, and Paul Tyson. Cambridge/New York: Cambridge University Press, 75–143.
    • Moltmann, Jürgen. 2005. God in Creation: An Ecological Doctrine of Creation. Translated by Margaret Kohl. London: SCM Press. 5th edition.
    • Origen of Alexandria. 1966. On First Principles: Being Koetschau’s Text of the de Principiis Translated into English. Edited by George W. Butterworth. New York: Harper & Row.
    • Otto, Rudolf. 1913. Naturalism and Religion. London: Williams and Norgate. 2nd edition.
    • Pannenberg, Wolfhart. 2001a. Systematic Theology. Volume 1. Grand Rapids: Eerdmans. 3rd edition.
    • Pannenberg, Wolfhart. 2001b. Systematic Theology. Volume 2. Grand Rapids: Eerdmans. 3rd edition.
    • Peacocke, Arthur Robert. 2004a. Creation and the World of Science: The Re-Shaping of Belief. Oxford/New York: Oxford University Press. 2nd edition.
    • Peacocke, Arthur Robert. 2004b. Evolution: The Disguised Friend of Faith? Selected Essays. Philadelphia: Templeton Foundation Press.
    • Polkinghorne, John C. 1993. One World: The Interaction of Science and Theology. London: SPCK. 8th edition.
    • Polkinghorne, John C. 1996. Scientists as Theologians: A Comparison of the Writings of Ian Barbour, Arthur Peacock and John Polkinghorne. London: SPCK.
    • Polkinghorne, John C. 2000. Faith, Science and Understanding. London: SPCK.
    • Putnam, Hilary. 1975. Mathematics, Matter and Method. Volume 1. Cambridge: Cambridge University Press.
    • Re Manning, Russell (ed.). 2013. The Oxford Handbook of Natural Theology. With the Assistance of J. H. Brooke and F. Watts. Oxford: Oxford University Press.
    • Ritchie, Sarah Lane. 2017. ‘Dancing Around the Causal Joint: Challenging the Theological Turn in Divine Action Theories’, Zygon 52, no. 2: 361–379.
    • Rudwick, Martin J. S. 2014. Earth’s Deep History: How It Was Discovered and Why It Matters. Chicago/London: The University of Chicago Press.
    • Saunders, Nicholas. 2002. Divine Action and Modern Science: A Bradford Book. Cambridge: Cambridge University Press.
    • Schleiermacher, Friedrich Daniel Ernst. 2016. The Christian Faith. Edited by Paul T. Nimmo. London: Bloomsbury Academic. 3rd edition.
    • Shults, F. LeRon, Nancey C. Murphy, and Robert J. Russell (eds). 2009. Philosophy, Science and Divine Action. Philosophical Studies in Science and Religion 1. Leiden/Boston: Brill.
    • Southgate, Christopher. 2008. The Groaning of Creation: God, Evolution, and the Problem of Evil. Louisville, KY: Westminster John Knox Press.
    • Sparn, Walter. 1994. ‘Natürliche Theologie’, in Theologische Realenzyklopädie. Volume 24. Edited by 85–98. Berlin: de Gruyter.
    • Stenmark, Mikael. 2013. ‘Typologies in Science and Religion’, in Encyclopedia of Sciences and Religions. Edited by Anne L. C. Runehov and Lluis Oviedo. Dorrecht: Springer, 2309–2315.
    • Strauss, David Friedrich. 1873. The Old Faith and the New: A Confession. Edited by Mathilde Blind. New York: Henry Holt and Company.
    • Taylor, Charles. 2007. A Secular Age. Cambridge, MA: Belknap Press of Harvard University Press.
    • Tillich, Paul. 1951. Systematic Theology. Volume 1. Chicago: University of Chicago Press.
    • Troeltsch, Ernst. 1913. ‘Christentum und Religionsgeschichte [1897]’, in Gesammelte Schriften Bd. 2: Zur religiösen Lage, Religionsphilosophie und Ethik. Tübingen: Mohr Siebeck, 328–363.
    • Turner, Frank M. 2010. ‘The Late Victorian Conflict of Science and Religion as an Event in Nineteenth-Century Intellectual and Cultural History’, in Science and Religion: New Historical Perspectives. Edited by Thomas M. Dixon, Geoffrey N. Cantor, and Stephen Pumfrey. Cambridge/New York: Cambridge University Press.
    • Ungureanu, James C. 2019. Science, Religion, and the Protestant Tradition: Retracing the Origins of Conflict. Pittsburgh: University of Pittsburgh Press.
    • van den Brink, Gijsbert. 2020. Reformed Theology and Evolutionary Theory. Grand Rapids: William B. Eerdmans.
    • van Eyghen, Hans. 2020. Arguing from Cognitive Science of Religion: Is Religious Belief Debunked? London: Bloomsbury Academic.
    • Watts, Fraser. 2013. ‘Natural Theology and the Mind Sciences’, in The Oxford Handbook of Natural Theology. Edited by Russell Re Manning. Oxford: Oxford University Press, 475–487.
    • White, Andrew Dickson. 1896. A History of the Warfare of Science with Theology in Christendom. 2. vols. Translated by J. Cottingham, R. Stoothoff, and D. Murdoch. New York: D. Appleton.
    • White, Claire. 2021. An Introduction to the Cognitive Science of Religion: Connecting Evolution, Brain, Cognition and Culture. Abingdon/New York: Routledge.
    • Whitehead, Alfred North. 1978. Process and Reality. New York: Free Press.
    • Wildman, Wesley J. 2014. ‘Religious Naturalism: What It Can Be, and What It Need Not Be’, Philosphy, Theology and the Sciences 1, no. 1: 36–58.

Academic tools