Creation in the Old Testament

William P. Brown

Creation in the Old Testament (or Hebrew Bible) is not limited to the first two creation accounts in Genesis (1:1–2:3; 2:4b–3:24). It is a widespread theme featured not only in Genesis but also in the Psalms, the Prophets, and the Wisdom literature. This article surveys representative creation texts in the Old Testament, noting their particularities and differences as well as exploring their theological and ecological relevance for contemporary communities.

1 Introduction

‘In the beginning God created the heaven and the earth.’ Such are the familiar words from the King James Version that open the Bible. A more accurate rendering of the Hebrew would be: ‘When God began to create the heavens and the earth’ (CEB, similarly NJPS). In any case, the Bible begins with creation. Not coincidentally, the Christian Bible also concludes with creation, specifically new creation (Rev 21–22). In between its bookends, the Bible has so much more to say about creation.

1.1 The pervasive theme of creation in the Bible

Beyond Genesis 1, the story continues with creation’s destruction by the Flood (7:10–12, 17–24), followed by an account of re-creation (8:13–21) that concludes with a covenantal promise from God to never destroy the earth again (9:1–17). But even as the story turns more narrowly toward Israel’s ancestors, beginning with Abraham and Sarah (Gen 11:26–12:3), creation never quite fades into the background. It remains a continuing focus: Abraham is beckoned to count the ‘stars of heaven’ as a sign of God’s promise of progeny (15:5), so also the ‘sand’ of the seashore (22:17). The ‘God of Abraham, Isaac, and Jacob’ is identified unequivocally as the ‘maker of heaven and earth’ (14:19, 22). Beyond Genesis, as we shall see, creation finds its home in the wisdom literature, the Psalms, and even the Prophets. Indeed, the two accounts of creation in Genesis are only the first of several accounts interspersed throughout the Old Testament, including Psalm 104 ; Job 38–41; Prov 8:22–31; Eccl 1:3–11; and portions of Isaiah 40–55. Moreover, the poets and prophets of the Old Testament drew heavily from the natural realm for imagery to press their various messages. From Genesis to Revelation, the theme of creation is manifold (cf. Ps 104:24).

1.2 The emerging interest in biblical creation

Theologically, the word ‘creation’ regards the physical world as the work of God, from the universe, or perhaps the multiverse, as a whole to the subatomic realm of quantum physics. That theology and science share this overlapping subject matter has resulted in occasional bouts of conflict as well as moments of shared understanding and mutual respect. Given this shared domain, interest in the theme of creation in the Old Testament has ebbed and waned ever since the emergence of science as an independent discipline of study in the sixteenth century. Most recently, however, such interest has gained positive momentum as theologians and scientists enter into dialogue to explore common ground and shared concerns, including concern for the wellbeing of the planet (Brown 2010; Rasmussen 2013).

In modern biblical studies, interest in the biblical creation accounts burst onto the scene with the 1849 discovery of Enuma elish, often called the ‘Babylonian Epic of Creation’. Named after its first two words (‘When above’), this collection of seven clay tablets recounts the creation of the gods and the world, as well as the battle of the gods and the ascendancy of Marduk, the patron deity of Babylon. This account, in other words, enfolds together theogony, theomachy, and cosmogony. Biblical scholars took note of this discovery, particularly Hermann Gunkel. In his 1895 book Schöpfung und Chaos in Urzeit und Endzeit (‘Creation and Chaos in the Primeval Era and the Eschaton’), Gunkel argued that this Mesopotamian myth underlay the story of creation in Genesis 1, specifically with respect to the motif of cosmic combat or chaos conflict (Chaoskampf), which depicts the hero god vanquishing the deified force of chaos (i.e. Tiamat in Enuma elish) before establishing creation (Gunkel 2006, first published in 1895), a motif evident also in other biblical texts such as Psalms 74 and 89 (see Levenson 1994).

Since Gunkel’s study, the 1929 discovery of the royal archives of Ugarit, an ancient city in northern Syria off the Mediterranean coast (modern Ras Shamra), provided more evidence of this motif. The so-called Ba‘al Cycle features a dramatic struggle between the storm god of fertility, Ba‘al, and the chaos monster Yamm (‘Sea’), indicating a West-Semitic origin of this combat motif. In addition, the serpent monster Tunnan is featured, highlighting the role of deified chaos in these myths in various forms, including the god of death (Mot). Although the Ba‘al Cycle lacks a bona fide creation account, parallels of the Chaoskampf motif are evident between certain biblical creation texts and the Ugaritic literature in consonance with the Mesopotamian Enuma elish.

While the comparative literature of the ancient Near East directed scholarly attention toward creation in the Bible, the broader field of Old Testament theology throughout much of the twentieth century tended to regard creation as peripheral to, if not a foil of, ancient Israelite faith, given its alleged associations with ‘pagan’ religion, from Canaanite to Babylonian. Israelite religion, by contrast, was regarded as based primarily on the ‘mighty acts of God’ in history (Wright 1950 and 1952), ‘salvation history’ (von Rad 1966: 131–142), or ‘covenant’ (Eichrodt 1961). The theological marginalization of creation in biblical scholarship during this time was due in part to the unfounded association of natural religion with Nazi ideology, on the one hand, and the theological vilification of Canaanite fertility cults, on the other (see Brueggemann 1996: 177–190).

Since then, however, ‘creation’ has become an object of intense, if not urgent, interest among theologians and biblical scholars for at least three reasons: (1) a greater appreciation of ancient Israel’s cultural ties to other ancient Near Eastern cultures (Mesopotamian, Canaanite, and Egyptian), whose own creation accounts and epic myths influenced, in varying degrees, Israel’s own creation traditions; (2) the mounting ecological crises facing the planet; and (3) growing interest in fostering dialogue between religion and science. This article surveys the theme of creation in the Old Testament with a particular focus on various creation texts, exploring their significance theologically and ecologically (see Ecotheology for more on this approach).

2 Creation in the Old Testament

While the theme of creation is widespread throughout the Bible, particularly in the Old Testament, there is no specific word for ‘creation’ in biblical Hebrew as there is in New Testament Greek (i.e. ktisis). Lacking in biblical Hebrew is a nominal cognate to the verb bārā’ (‘create’), which has or implies only God as its subject. Its abstract nominal form, bry’h, is first attested in the Dead Sea Scrolls by the mid-second century BCE and is subsequently well attested in later rabbinic literature (see Martínez 2007: 219–240). Closest is the poetic pair (or merismus) ‘heaven and earth’ or ‘the heavens and the earth’, signifying the totality of creation (e.g. Gen 2:1; Exod 20:11; Deut 4:26; Ps 113:6; Jer 32:7; Joel 3:16; Hag 2:6, 21), with God referenced as ‘maker of heaven and earth’ (Gen 14:19, 22; Ps 134:3). Creation in the Old Testament, in other words, reflects the dialectic of transcendence and immanence, a dialectic that is far from dualistic. While ‘heaven(s)’ is related to the divine realm in varying degrees, sometimes depicted as the place of God’s residence, the ‘earth’ covers all life and domains ‘below’, what scientists call the ‘biosphere’, including the land, the oceans/seas, and the sky (also referred to as ‘the heavens’; Hebrew haššāmayim). In certain traditions, the temple is considered the portal or conduit between heaven and earth, given that God is said to dwell in both (e.g. 1 Kings 8). Scholars often refer to the overall picture of creation as the ‘astrodome’ model or the three-tiered universe:

Diagram of the three-tiered universe model in the Bible, with the firmament as a dome separating the heavens or waters above from the sky, Earth, Sheol, and the Great Deep below image/svg+xml Raqia or Firmament Sheol Tehom or Great Deep Earth Mountains Chambers in heavens Sun Moon Stars Windows Portals Pillars of the Earth Fountains of the Deep
Figure 1. Diagram of the ‘astrodome’ model of biblical cosmology (source: Wikimedia Commons; copyright Tom Lemmens, used under licence CC BY 4.0)

Terminology aside, the prominence of creation in the Old Testament is reflected particularly in the variety of creation-oriented texts interspersed throughout the ancient scriptures, beginning with Genesis but including most notably Job 38–41, Psalm 104, and Prov 8:22–31, not to mention numerous passages that treat some aspect of creation in the prophetic literature and elsewhere. Like snowflakes, no two creation accounts are alike, each one sharing a distinctive perspective or emphasis. We begin with the creation account that serves as the gateway to the Jewish and Christian scriptures: Gen 1:1–2:3 (hereafter referred to as Genesis 1).

2.1 Genesis 1: the cosmic account

The first creation account in the Bible, typically called the ‘Priestly’ account by scholars, is arguably the most densely structured text in the entire Hebrew Bible. This is no accident: Genesis 1 depicts a methodically ordered creation patterned after the architecture of the temple (see section 2.1.2). Compared to other ancient Near Eastern accounts, Genesis 1 reads like a dispassionate treatise, resembling more an itemized list than a flowing, let alone gripping, narrative (McBride 2000: 6).

In Genesis 1, God creates the world in six days, each day building on the previous, through eight acts and ten utterances, with the seventh day marking creation’s completion by God’s sanctified ‘rest’. The curtain rises to reveal a pre-creative state of benign ‘chaos’, a condition of watery emptiness and formlessness signalled by the assonantal expression tōhû wābōhû (Gen 1:2). In this state of primordial soup, God commences to create, methodically and purposefully, beginning with the creation, or better, unleashing of light (v. 3; see Smith 2010: 71–78). The narrative’s step-by-step character is governed by the alternating rhythm of divine commandment (‘Let . . .’) and fulfillment, by word and deed. God acts creatively in various ways: calling into existence, dividing, making, naming, seeing, and blessing. Repeated seven times is God’s approbation (‘God saw that it was good’), which concludes nearly every single act (1:4, 10, 12, 18, 21, 25), culminating in the pronouncement ‘very good’ to cover creation as a whole (v. 31). God’s specific creations are as follows:

  1. light and its separation from the darkness, resulting in ‘day’ and ‘night’;
  2. the ‘dome’ or ‘firmament’ (Hebrew rāqîa‘), named ‘sky’ (Hebrew šāmayim), separating the waters ‘above’ from the waters ‘below’;
  3. the collections of the waters ‘below’, called ‘seas’ (Hebrew yammîm) allowing for the ‘land/earth’ (Hebrew ’ereṣ) to emerge;
  4. plants bearing seeds;
  5. lights in the sky;
  6. aquatic and avian creatures;
  7. land animals;
  8. human beings.

The process of creation in Genesis 1 proceeds as a series of separations: separations of light from darkness, day from night, the ‘waters above’ from the ‘waters below’, and the land from the waters. As the final step, the last day is separated from all previous days, given its uniquely holy status. As the crown of creation, the seventh day signals creation’s completion. In sum, creation in Genesis 1 is a graduated process of differentiation leading to a diversity of domains, on the one hand, and the diversity of life, on the other.

The valuation ‘good’ (Hebrew ṭōv), repeated seven times in Genesis 1, spans a range of meaning. Completed as ‘good’, creation is a bona fide cosmos, an ordered, harmonious, and beautiful whole, the opposite of chaos (cf. 1:2). ‘Goodness’ also acknowledges the plenitude and diversity of life, both animal and botanical, as well as creation’s own capacities for sustaining life. A ‘good’ creation is one that is stable, habitable, sustainable, and flourishing. Indeed, part of creation’s goodness is that creation itself is creative. In fact, God in Genesis 1 does not always create alone: God enlists the ‘land/earth’ and the ‘waters’ to bring forth life (1:11, 20, 24). The earth and the waters exercise their own creative agency at the behest of God’s command. Far from being inert elements, the ‘earth’ and the ‘waters’ prove to be ‘empowering environments’ (Welker 1999: 40, 42). Creation in Genesis turns out to be a collaborative affair, including the creation of humanity (v. 26). In other words, God creates a creating creation.

2.1.1 Cosmic symmetry

But there is more to this well-ordered cosmos than its ‘goodness’. The Priestly account of creation reflects an unmistakable literary symmetry that unfolds for the first six days of creation:

Diagram showing the literary symmetry in the biblical description of the first six days of creation 'Day 0' (1:1–2) (formless and empty) Day 1 (1:3–5) Day 4 (1:14–19) Day 2 (1:6–8) Day 5 (1:20–23) Day 3 (1:9–13) Day 6 (1:24–31) (form-full and filled) Day 7 (2:1–3) Lights Light SkyWaters (below) Avian lifeMarine life LandVegetation Land lifeHuman lifeFood Sabbath
Figure 2. Cosmic symmetry in the six days of creation (Genesis 1:1–2:3)

According to their thematic correspondences, the first six days line up to form two parallel columns. Days 1–3 delineate the cosmic domains (i.e. sky, water, and land), which are then populated by various inhabitants of these various domains (Days 4–6). Vertically, the two columns address the two abject conditions of lack referenced in 1:2, formlessness and emptiness (Hebrew tōhû wābōhû). The left column (Days 1–3) recounts the cosmos being formed, while the right column (Days 4–6) describes the cosmos being filled. Day 3, moreover, serves as the link by depicting the land as fully ‘vegetated’, equipped to provide the means for sustaining animal life. Days 4–6 report the filling of these domains with their respective inhabitants, from the celestial spheres (1:14–18) to human beings (1:26–28). With the stars, sun, and moon set in the heavens and the various forms of life, each according to its ‘kind’, filling the sky, land, and sea, creation proceeds from emptiness to fullness in the right column, just as it had progressed from formlessness to form-fullness in the left.

The symmetrical structure of creation reveals a rather complex view of life. While Genesis 1 seems to define life as anything that can ‘be fruitful and multiply’, that is, reproduce (1:22, 28), such a definition proves to be too limited in its larger schema. Vegetation, for example, is created on the third day and thus seems not to be deemed living. (Nowhere are plants commanded to ‘be fruitful and multiply’.) Otherwise, the author would have placed the creation of vegetation on the fifth or sixth day accompanied by God’s command to ‘fill’ the earth. Instead, plants are deemed an indelible and edible feature of the land, an integral part of the earth’s domain. Elsewhere in biblical tradition, however, plants are considered alive (e.g. Job 14:7–10; Ezek 17:24; Zech 11:2). The emphasis, moreover, on plants bearing ‘seed’ in Genesis 1 points to the power and means of botanical succession. Indeed, it is no coincidence that the term for ‘seed’ in Hebrew (zera‘) often designates progeny, that is, procreated life (e.g. Gen 4:25; 15:3; Jer 31:27; cf. 1 Cor 15:36). In Genesis, vegetation covers the ‘dry land’ (Hebrew yabbāšâ), called ‘earth’ (Hebrew ’ereṣ), to provide for life, making it inhabitable for animals, both human and nonhuman (Gen 1:11–12, 29–30). As the land hosts plant ‘life’, so plants host animal life, turning the land into a landscape fit for the proliferation of animal life. With plants as an essential part of its domain, the once ‘dry land’ is transformed into a living landscape.

Also suggestive is the creation of ‘lights’ on the fourth day of creation: sun (‘the greater light’), moon (‘the lesser light’), and ‘the stars’ (1:14–18). These astral bodies are considered just as much inhabitants of their own domain (‘light’) as the birds and the fish are of theirs (‘sky/heavens’ and the ‘waters’). Moreover, the celestial spheres and human beings share common function: the former are to ‘rule’ both day and night (1:14–18), while the latter are to exercise ‘dominion’ over all other creatures (1:26–28). Stars and human beings both share in the task of ruling.

Although life seems to be defined categorically by reproduction in Genesis 1, structural considerations noted above suggest a more complex and expansive picture. The mobility of celestial bodies and their designation as members of their primordial domain, coupled with their assigned tasks to rule both day and night, suggest a functional correspondence with human life on earth. As often observed, Genesis 1 de-divinizes, or de-mythologizes, the sun and the moon, which are given only functional designations in the text rather than their common names (Hebrew šemeš and yārēaḥ). Nevertheless, Genesis does not ‘de-animate’ them. As bona fide creations of God, the sun and the moon exhibit some sense of ‘life’ in their prescribed agency vis-à-vis life on earth. Thus, we have in Genesis 1 alternative forms of life that distinguish themselves from procreating life yet exercise active agency.

2.1.2 Cosmic temple

The structure of Genesis 1 commends itself not only by its unfolding symmetry but also by its architecture, with the final day serving as its capstone. Without the seventh day, the sabbath day, the creation pattern would lose a distinction that remains hidden to modern readers not acquainted with the ancient architecture of sacred space. Many temples of the ancient Near East, including the Solomonic temple, exhibited a threefold or tripartite structure that can also be found in the literary symmetry of the Genesis text (see Jenson 1992; Hundley 2013: 749–767).

The Solomonic temple described in 1 Kings 6, for example, consists of an outer vestibule or portico, the nave or main room, and an inner sanctum or holy of holies, as diagramed below.

Diagram of the Temple structure showing portico (top), nave (middle), and Holy of Holies (bottom) Portico Nave Holy of Holies
Figure 3. Diagram of the Solomonic temple

This threefold arrangement of sacred space corresponds to the way in which Genesis 1 recounts the various days of creation as they are distributed both chronologically and thematically. The first six days, by virtue of their correspondence, establish the architectural boundaries of sacred space, while the last day inhabits, as it were, the most holy space.

Diagram of the Temple structure showing portico (top), nave (middle), and Holy of Holies (bottom), with six days of creation superimposed around the portico and day seven within the Holy of Holies Portico Nave Holy of Holies Day 2 Day 1 Day 3 Day 6 Day 7 Day 5 Day 1 Day 4
Figure 4. Temple space as it corresponds to the six days of creation in Gen 1:1–2:3

As creation unfolds ‘daily’, it becomes constructed in the imago templi, in the model of a temple (cf. Isa 66:1–2; Job 38:4–7). The correspondence is unmistakable: in the temple’s holiest recess, God dwells; on the holiest day of the week, God rests (see 1 Kgs 8:10–13; cf. Ps 132:8, 13–14). This temple-oriented structure highlights the importance of the sabbath day, the crown of creation, serving also as a model for human conduct: as God rested on the primordial seventh day after completing creation, so Israel is to rest by observing the sabbath after six days of work (Exod 20:8–11; 31:13–17). With the sabbath serving as creation’s capstone, a ‘temple in time’ (Heschel 1951), the theological message of Genesis 1 is complete: the cosmos, indeed all creation, is God’s cosmic temple (see also Walton 2009: 72–92).

2.1.3 Imago Dei

The cosmic temple of Genesis 1 also reveals something significant about humanity’s role and place in creation. As a rule, temples throughout the ancient Near East contained an image of its resident deity. In Jerusalem, however, the physical representation of God was expressly forbidden (Exod 20:4; Deut 5:8). Because God was considered to be without form, the deity could not be represented (see Deut 4:12–18).

Genesis 1, however, does not reject the language of divine ‘image’ (Hebrew ṣelem). Rather, it recasts it by identifying the imago Dei with human beings, created on the sixth day (Gen 1:26a, 27). Human beings alone, according to the text, bear an iconic relation to the divine. Used elsewhere to refer to idols such as the cult statues of other deities (see Middleton 2005: 45–46), the Hebrew term for ‘image’ is applied to human beings to acknowledge their near-divine status. While God lacks a blatantly anthropomorphic profile in Genesis 1, human beings are unequivocally theomorphic by design. Made in God’s ‘image’, women and men reflect and refract God’s presence in the world.

Also distinctive of humanity’s creation in Genesis 1 is the use of first-person plural language (‘let us make’) in divine discourse. While, as noted above, God had earlier enlisted the creative agencies of earth and water to create non-human life, here God enlists the members of the divine council, minor divine beings that serve the supreme, enthroned deity, variously referenced as ‘the sons of God’ (Hebrew bĕnê hā’ĕlōhîm; Job 1:6), ‘sons of the Most High’ (Hebrew bĕnê ‘elyôn; Ps 82:6), or simply ‘gods’ (Hebrew ’ĕlōhîm; Ps 82:1; cf. 8:5 [Heb. 6]), including heavenly ‘messengers’ or angels (Hebrew malā’kîm; e.g. Gen 19:1, 15; 32:2; Ps 103:20; Job 4:1). The Priestly account of creation, however, refers to the plurality of the divine realm only here, most likely to reflect or match the plural nature of humanity, created ‘male and female’ (Gen 1:27b). In other words, the Priestly account acknowledges humanity’s plural, social nature as a dimension of the imago Dei, reflecting something of the plural, social nature of the divine realm within this ancient cosmogony.

While so much more could be said about the significance of humanity made in the ‘image of God’ (see Brown 2010: 41–44; Schüle 2005: 5–7), it is clear that such language points foremost to humanity’s elevated status, specifically a royal, function-based status that is tied to exercising ‘dominion’ over creation. In ancient Near Eastern royal traditions, it is the king who assumes the status as the deity’s ‘image’. But in Genesis 1, such status extends to all humanity, a ‘democratizing’ move on the part of the Priestly author. But what does such ‘dominion’ entail? The language of Genesis 1 is admittedly harsh, particularly with the call to ‘subdue’ (Hebrew kābaš) the earth (v. 28). Outside of Genesis, the verb can connote military conquest (Zech 9:15 and Num 32:29) and rape (Esth 7:8), as well as enslavement (Jer 34:11; Neh 5:5). While it is difficult to soften its tone in Genesis, it would be inaccurate to claim that Gen 1:28 within its context envisions creation as something to be conquered or exploited, as if the earth were humanity’s enemy. Creation is not chaos, warranting conquest either by God or by humanity. Indeed, God pronounced creation ‘good’ six times before humanity arrived on the scene (vv. 4, 10, 12, 18, 21, 25). Nevertheless, reference to ‘subdue’ gives the motif of ‘dominion’ an aggressive, imperial edge to it, affirming a hierarchy in which human beings assume supremacy over the rest of creation.

What this passage meant to an ancient agrarian society does not imply war with creation or mandate creation’s exploitation (contra White 1967: 1203–1207). This passage provided a theological mandate to the farmer who had to toil painfully to turn rocky, hilly, and heavily forested land into arable fields. Backbreaking effort to make the land productive for human subsistence was no ‘peaceful’ matter. The land’s rocky soil was resistant to plowing and thus had to be forced into productivity (cf. Gen 3:17–19). The hills had to be terraced, and domesticated animals had to be raised and bred. The passage depicts the farmer as a royal warrior in the cultivation of the soil and in the care of flocks and herds, all within a prescribed hierarchy of life. The ancient blessing of humanity’s ‘dominion’ over creation would affirm for twenty-first century readers the new geological epoch of the Anthropocene, which acknowledges humanity’s power as a species to determine the very course of creation, no less, for woe or for weal (see, e.g., Grinspoon 2016).

2.2 Genesis 2–3: the garden account

In Gen 2:4b–3:24, creation becomes more narrowly focused. The God of the cosmos exchanges the royal sceptre for a garden spade, as it were. Known as the Yahwist account for its prominent use of the divine name cast in composite form, ‘the Lord God’ (Hebrew yhwh ’ĕlōhîm), this second creation story is altogether different in content and tone from Genesis 1. Compared to the measured cadences of its canonical predecessor, this account reads more like a folktale. While the Priestly account spans the breadth of the cosmos with systematic rigour, the Yahwist story takes place in a garden with its focus on humanity’s first family.

The opening scene in Genesis 2 is strikingly different: whereas the Priestly account begins with dark, watery ‘chaos’ (1:2), the Yahwist account opens with dry, barren terrain – a land of lack (2:5). Nevertheless, both scenes emphasize an initial setting of emptiness. Here the ‘ground’ (Hebrew ’ădāmâ) takes centre stage, cast as a major character in the narrative. The grand cosmogony of Genesis 1 is followed by a soil-bound anthropogony in Genesis 2–3.

The Yahwist account unfolds in four scenes, each containing parallel elements (Gen 2:4b–17; 2:18–25; 3:1–7; 3:8–24). Every scene identifies a deficiency or lack followed by God’s response. The plot is driven by the fits and starts of ongoing creative activity, suggesting more an improvisation on God’s part than a meticulously executed program; or if put to music, more like a jazz performance than a Bach fugue. The Yahwist tale is the story of God’s responsiveness to certain free variables in the plot, particularly the serpent and the human couple. In contrast to Genesis 1, creation according to the Yahwist is a series of ‘not-goods’ made good, but with a tragic turn at the end.

The first recounted act of creation in Genesis 2 is not the creation of light out of darkness (1:3) but the fashioning of a human being out of the ground (2:7). The ’ādām (‘human one’) is made not in the imago Dei but in the imago terrae, in the image of the earth, as it were. Humanity’s identity is bound to the ground by a remarkable Hebrew wordplay: the ’ādām comes from the ‘dust’ or topsoil of the ’ădāmâ (‘ground’), which makes humanity a ‘groundling’. But humanity’s tie to the land is more than an etymological accident; the ’ādām’s origin from the ‘ground’ also reflects the ’ādām’s vocational identity to work or ‘serve’ the ground (2:5, 15). Contrary to Genesis 1, the ’ādām is the first rather than the last of God’s creatures to be formed. Yet this hominid of the humus remains a work in progress. For now, Yhwh places the ’ādām in a garden, a horticultural feast for the eyes and the appetite (2:9). As ‘the divine farmer’, Yhwh cultivates the soil and plants trees, as well as shaping clay (Hiebert 1996: 67). Yhwh works unbegrudgingly in the soil to raise a garden and to fashion a human being to maintain and harvest it.

Providing water for much of the earth, Eden is no ordinary garden. The name ‘Eden’ has more to do with condition than with location; it designates a setting of delight and plenty (Hebrew ‘dn), the sheer opposite of lack. The Garden of ‘Plenty’ is the Garden of ‘Eden’ (cf. Gen 18:12; Ps 36:8; Neh 9:25; Jer 31:12; 51:34; Ezek 36:35; Joel 2:3). Moreover, this garden is where God strolls in the cool ‘breeze of the day’ (Gen 3:8), evoking a sense of sanctity about the garden, perhaps even the setting of a temple garden. In any case, Eden’s garden is no pristine paradise of leisure: it must be tended. This ‘groundling’ is given the task to ‘serve’ (Hebrew ‘bd) and ‘preserve’ (Hebrew šmr) the garden and its soil (2:15). The ’ădām is tied to the ’ădāmâ in service, as the ’ădāmâ yields to the ’ādām its productivity. Both are reciprocally bound together.

Unlike certain ancient Near Eastern creation accounts, humans in Genesis 2 are not created to perform menial labour for the gods, who themselves loathe such toil. The God of Genesis 2 creates the ’ādām and bestows upon him the commission to preserve the garden for his own benefit. The garden exists for the ‘groundling’ and the ‘groundling’ for the garden. The ’ādām is given wide-ranging freedom with only one specific restriction: he is granted access to ‘every tree of the garden’ except one, the ‘tree of the knowledge of good and evil’ (2:16–17). Judgment, moral and otherwise, which comes either with maturity or as a gift of God (2 Sam 14:17; 1 Kgs 3:9), is not meant to be part of the groundling’s cognitive toolkit, at least not yet. The tree of knowledge represents improper and premature access to such capacity.

God then resorts to an alternative plan by fashioning another creature, this time from the ’ādām’s own body. This new creation requires, however, a new lack: a part of the ’ādām must be removed. This new being is no one-way derivation, however: from the creation of the woman (Hebrew ’iššâ), the ’ādām gains ‘his’ gender as a ‘man’ (Hebrew ’îš). The narrative acknowledges that the man and the woman are mutually engendered and, at the same time, of common substance and form (2:23). With the creation of the woman, humanity is now separated by gender. In Genesis 1, separation plays a key role in cosmic creation. In Genesis 2, the separation of flesh makes possible sexual/gender differentiation and, in turn, sexual union (v. 25).

With the creation of the woman, the man now takes on a dual identity: he remains kin to the ground in his humanity as he has become kin to the woman in his gendered identity. No subordination pertains in the garden. Life in the garden is one of fruitful work, abundance, and mutual companionship. In the garden, there is no fear or shame, even before God. The absence of fear and shame, that is, the condition of innocence, was meant to endure. But, alas, it does not. The creation story of the garden in Genesis 2 leads to a crime and punishment story in Genesis 3. Suffice it to say, things do not turn out well. The primal couple’s act of disobedience in partaking from the forbidden fruit leads to an awakening of consciousness: their eyes ‘were opened’ to recognize their own naked vulnerability (3:7), while at the same time becoming ‘like God/gods’ (3:5), which compels Yhwh to expel them (3:11) to face the hardships of mortal life outside the garden, including the pain of labour, whether in bearing children or in cultivating crops (3:16–19).

Death, moreover, becomes a painful reality (v. 19). The issue of whether the primal couple was created immortal or not in the garden remains open to interpretation. Whereas the weight of Christian tradition favours the former, the story can be just as easily, if not more plausibly, read as an account of how the first humans lost the opportunity to become immortal. Yhwh expels the couple out of concern that they will become immortal (Gen 3:22), implying that the couple had not partaken from the tree of life prior to their disobedience. Adam’s ‘curse’ in v. 19, thus, is not itself a death penalty but an expulsion penalty, anticipating what is to come in v. 22 (see Wright 1996: 305–329). In any case, through disobedience and expulsion, humanity becomes more fully human: complex and self-conflicted, prone to violence and capable of doing good. Nevertheless, Yhwh does not abandon them, and the story continues generation after generation.

Such are the first two creation accounts in the Bible, both very different in orientation, form, and content. Set side by side, these two stories illustrate the complex, multiple nature of human identity, made both in the ‘image of God’ and from ‘the dust of the ground’. On the one hand, humanity is given the power to exercise dominion over all creation; on the other, humanity is commissioned to ‘serve’ creation. Humanity is both ‘groundling’ and ‘godling’. In conjoining these two stories together (see 2:4a), the ancient editors acknowledge the complexity of human identity, place, and vocation in the world. God, too, is a study in complexity: cosmically royal and rurally immanent. As for creation, the Priestly author constructs a cosmic temple while the Yahwist cultivates a sanctuary garden. Thus creation ‘in the beginning’ in the Bible is no uniform story told from a single perspective. And there are more stories to come, as one finds in the Psalms.

2.3 Creation in the Psalms

The Psalms offer various perspectives of creation and humanity’s place in it, including Psalms 8, 33, 104, and 148. In addition, the central portion of Psalm 74 features the West-Semitic Chaoskampf motif (see section 1.2 above), a feature widely attested in ancient Near Eastern literature concerning the gods and creation. (All translations are the author’s unless otherwise noted.)

But O God, my king from of old,
who works salvation within the earth,
you split the sea with your might;
you shattered the heads of the sea-dragons upon the waters.
You crushed the heads of Leviathan;
you gave it as food to a pack of desert dwellers.
You broke open springs and wadis;
you dried up perennial streams.
Yours is the day; yours also is the night;
you established moonlight and sun.
You set all the boundaries of the earth;
summer and winter—you made them.

(Ps 74:12–17)

As ‘king from of old’, God’s work of creation is deemed salvific. The psalm describes God ‘shattering’ the sea-dragons and the multi-headed Leviathan. According to Isa 27:1, Leviathan is the ‘swift serpent’ (Hebrew nāḥāš bārīaḥ) that is consigned to destruction ‘on that day’ (cf. Job 26:3). This fearsome creature of chaos is fully depicted in Job 41, a sea monster without rival. God’s victory over chaos has its creational counterpart in God’s opening of ‘springs and wadis’, reminiscent of Marduk’s piercing of Tiamat’s eyes to unleash the Tigris and Euphrates in Enuma elish. God is depicted as the divine warrior-king who begins creation by conquest and concludes with placing boundaries and ordering time, a cosmic drama that proceeds from chaos to order. Such is the movement of salvation in a creational context. Similarly in Psalm 89, God is said to have defeated the sea monster Rahab:

Yhwh, God of hosts, who is like you?
Mighty Yh, your faithfulness surrounds you!
You rule over the sea’s swelling;
you still the surging of its waves.
It is you who crushed Rahab like a corpse;
with your strong arm you scattered your enemies.

(Ps 89:8–10 [Heb. 9–11])

‘Rahab’, like Leviathan, is a mythological monster associated with the sea’s billowing waves, consigned to destruction by God, ‘crushed . . . like a corpse’ (see Isa 51:9; Job 9:13; 26:12). With the conquering of chaos comes the ‘scattering’ of enemies. The successful subduing of chaos serves as the litmus test of God’s sovereignty in and over creation – an ancient association found, for example, in Enuma elish: creation effected through conquest. In near parallel fashion to the Babylonian Epic of Creation, reference to the chaos monster’s ‘corpse’ in Psalm 89 is immediately followed by acclamation of Yhwh’s ‘founding’ (y-s-d) of all creation (v. 12). Creation of the cosmos presupposes the conquest of chaos, as part of the poetic, epic sequence of conquest, creation, and cosmic sovereignty.

The Chaoskampf motif, however, does not find such a ready home in other creation accounts in the Old Testament. As noted above, it is absent in Genesis 1: the waters of ‘chaos’ in 1:2 are benign in so far that they present no threat to God. Instead, the primordial ‘waters’ are used by God in creation to create aquatic life. Similarly, in Psalms 8 and 104 watery chaos plays little or no role (see 104:6–9). On the other hand, these two psalms differ widely regarding humanity’s place in creation.

2.3.1 Psalm 8: from divine glory to human glory

As the first praise psalm in the Psalter, Psalm 8 is ‘bookended’ by praise of Yhwh’s ‘majestic’ and sovereign name, within which the psalm proceeds from praise of God’s glory to a claim about humanity’s glory. Structurally, the psalm sets divine glory and human glory in mirror-like apposition, but what distinguishes them are the contrasting scales by which glory is measured: cosmic and earthly. The middle section (vv. 3–4 [Heb. 4–5]) pivots from God’s cosmic glory to humanity’s earthly glory. It is amid God’s cosmic glory reflected in the night sky that the speaker inquires of humanity’s identity and place in the form of a question that cannot be asked except theologically: ‘What is humanity that you call them to mind, humankind that you care for them?’ (v. 4 [Heb. 5]). The question does not begin with ‘who?’ () but with ‘what?’ (), conferring a sense of insignificance. More importantly, the interrogative echoes the refrain regarding God’s majesty (‘How [] majestic is your name’), infusing the question with a sense of praise-filled wonder. The psalmist’s central question, thus, carries a double significance: (1) a sense of humanity’s insignificance before creation’s cosmic splendour; and (2) a sense of awe and mystery that humanity remains the object of Yhwh’s attention. The former emphasizes the latter.

As for the latter, the psalm’s final section spells out the reason behind Yhwh’s mindfulness of humanity (vv. 5–8 [Heb. 6–9]). Yhwh has created humanity for a certain status and function. Compared to the divine realm, the human realm is only slightly ‘less’ (Hebrew ḥ-s-r) in terms of God’s ‘glory’ and ‘grandeur’, a difference in degree rather than category. When compared to the divine, a slight deficiency turns out to be a great surfeit, specifically a surplus of splendour. As God shares in ‘splendour’ (Hebrew hôd), so humanity shares in ‘grandeur’ (Hebrew hādār). ‘How great thou art’ is complemented, in effect, by ‘how great we are’. But humanity’s greatness, the psalm makes clear, is a derived glory: it rests entirely on God’s glory. Whereas God’s glory is thoroughly cosmic, humanity’s glory is localized and limited on the earth. Moreover, humanity’s ‘glory’ is distinctly royal, the glory of dominion, whose exercise is found in ruling over the biological order, just as in Gen 1:26–28. All nonhuman animals are set ‘under [humanity’s] feet’, a vivid image of royal subjugation (cf. Ps 18:38 [Heb. 39]; Lam 3:34). Humanity constitutes the holy human empire on earth; Homo imperiosus is the functional sign of humanity’s ‘glory’.

Psalm 8 echoes Gen 1:26–28 in its casting of humanity as dominant over all other creatures. From the psalmist’s perspective, human beings are made ‘slightly less than the divine’ (v. 5a [Heb. 6a]). In Genesis, they are made ‘in the image of God’ (1:27). In Genesis, they are to have ‘dominion over the fish of the sea, and over the birds of the air, and over the cattle, and over all the wild animals of the earth, and over every creeping thing that creeps upon the earth’ (v. 26b). In Psalm 8, domestic and wild animals, birds, and all sea life are placed ‘under’ humanity’s ‘feet’ ( vv. 6–8 [Heb. 7–9]). Whether prosaically or poetically, both accounts apply the language of kingship to identify humanity’s place in God’s creation.

2.3.2 Psalm 104: Homo sapiens in a Terra sapiens

A very different picture of creation and humanity’s place is found in Psalm 104, which celebrates creation’s biodiversity and commends it all for Yhwh’s enjoyment. As the most extensive psalm of creation in the Psalter, Psalm 104 is a poetic revelry of praise to Yhwh for having created the world in all of its diversity, a product of divine wisdom (v. 24). This cosmic liturgy proceeds from the meteorological to the zoological to the anthropological, all expressed in doxological fashion. Scholars often observe several parallels with the ‘Great Hymn to the Aten’, the sun-disk deity whose worship was mandated under the Eighteenth Egyptian Dynasty of Pharaoh Akhenaten (14th century BCE). Psalm 104 and this Egyptian hymn share such motifs as the ‘call of admiration’ to the deity, the diversity of creation, aquatic life, ships plying the sea, and divine provision, from food to springs of water (see Schipper 2014: 61–69). While the parallels are strong, direct literary dependence seems unlikely.

Also well noted is the parallel movement shared in Genesis 1 and Psalm 104, so evident that one can easily delineate the psalm’s structure in terms of the ‘days’ of creation enumerated in Genesis: Day 1 = Ps 104:1–2; Day 2 = vv. 3–4; Day 3 = vv. 5–9; Day 4 = vv. 19–20; Day 5 = vv. 12, 17, 25–26; Day 6 = v. 23. But such synoptic comparison works well only for the first three days, those days that establish the creational domains of light, heaven, and land (vs. sea). Thereafter, the psalmist veers away from Genesis to revel in the wonder of creation’s diversity, particularly among the animals sustained by Yhwh’s providential care. Creation in Psalm 104 is not simply a matter of the primordial past, as one might infer from Genesis 1 with creation completed on the seventh day. Creation, moreover, is ongoing in God’s sustained provision for creation (vv. 27–30). Creation is as much a beginning point as it is the continuation of provision and renewal in Psalm 104.

As Genesis 1 testifies to how ‘goodness’ is built into creation at every step in the process, so Psalm 104 acknowledges God’s ‘wisdom’ in and throughout creation:

How many are your works, Yhwh!
With wisdom you have made them all.
The earth is stock full of your creatures!

(Ps 104:24)

The psalmist exclaims God’s wisdom reflected in creation (cf. Prov 3:19–20). The result is a Terra sapiens, a ‘wise world’ that hosts a staggering variety of creatures, including Homo sapiens (see Grinspoon 2016). To illustrate just how manifold creation is, the psalmist describes a host of animals: onagers, birds, cattle, cedars and other trees, storks, mountain goats, coneys, lions, and, yes, Leviathan – all attesting to God’s wise handiwork. As the species of life are wonderfully varied and numerous, so also are their habitats and niches, from towering trees and flowing wadis to mountainous crags and the deep dark sea. The psalm acknowledges that each species has its rightful habitat: the trees are for the birds (vv. 12, 17), the mountains ‘belong to’ the wild goats, the crags provide ‘refuge for’ the coneys (or rock hyrax, specifically Procavia capensis, v. 18), the lions have their dens (v. 22), and people have their homes. In God’s cosmic mansion there are many dwelling places, each one ‘fit’ for each species.

Remarkably, this panoramic overview of creation does not directly mention human beings until v. 23. As in Genesis 1, humanity is a literary latecomer, but not because it is the most dominant of creatures. Humanity is more of an afterthought.

You bring on the darkness and it is night;
when every animal of the forest prowls.
The young lions roar for their prey;
seeking their food from God.
When the sun rises, they withdraw,
and to their dens they retire.
Humans go forth to their work,
to their labour until evening.

(Ps 104: 20–23)

If one did not know any better, one might think that the only difference between lions and humans is that lions take the ‘night shift’, while humans take the ‘day shift’ to earn their living. All in all, humanity in Psalm 104 is considered merely one species among many, all coexisting together, each with its niche in creation and its rhythm for flourishing. But not one of them is placed ‘under’ humanity’s ‘feet’ in Psalm 104. Instead of assuming a dominant place in the ‘great chain of being’, humanity in Psalm 104 is given a non-anthropocentric entry into what might be called the ‘great encyclopaedia of being’. Creation is not simply habitat for humanity; it is habitat for diversity.

As for the matter of chaos, Leviathan raises its head in Psalm 104, but it is not an ugly head.

There is the sea, both vast and wide.
Therein are the sea creatures teeming beyond count,
living things both small and large.
There go the ships,
and Leviathan, with which you fashioned to play.

(Ps 104: 25–26)

Providing evidence of the creator’s wisdom (v. 24), the psalmist points to the vast teeming array of sea creatures, but Leviathan is singled out for special attention. Elsewhere in certain biblical traditions, as noted above, Leviathan is a creature clearly not for play but for destruction.

Not so in Psalm 104. No Chaoskampf rages here. The psalmist has taken a symbol of monstrous chaos, a figure of abject terror, and turned it into an object of playful wonder. In the poet’s hands, Leviathan is stripped of all the trappings of terror and becomes Yhwh’s partner in play. So if Leviathan is not the purveyor of chaos, who or what is? The psalm refrains from revealing the true monster until the end.

May Yhwh’s glory endure forever!
May Yhwh rejoice in his works!
Who looks at the earth, and it trembles;
he touches the mountains, and they erupt in smoke.
I will sing to Yhwh as long as I have life;
I will sing praise to my God while I still live.
May my poetic meditation be pleasing to him;
I will rejoice in Yhwh.
Let sinners vanish from the earth,
and the wicked be no more.
Bless Yhwh, my whole being!
Hallelujah!

(Ps 104: 31–35)

Psalm 104 concludes with acclamations of Yhwh’s ‘glory’ and ‘joy’, cast as indirect requests (‘may…’). Indeed, the entire psalm builds up to this culminating moment, revealing its primary agenda: to provide credible cause for Yhwh’s joy. By so vividly describing creation’s diversity and abundant flourishing, the psalm gives abundant reason for why Yhwh’s joy should endure.

But creation is not all perfectly harmonious: Yhwh can also cause the earth to ‘tremble’ (Hebrew r-‘-d) and the mountains to ‘erupt in smoke’ (Hebrew ’-š-n). Such is the language of theophany. Yhwh need only appear, and all creation convulses. It is to such power that the speaker appeals with an imprecation against ‘sinners’ and the ‘wicked’ at the end (v. 35a). The psalm’s cosmic scope, which includes even the monstrous Leviathan within the orbit of Yhwh’s providential (and playful) care, has no room for the wicked. By exhorting Yhwh to destroy the wicked, the psalmist transfers the chaos traditionally associated with mythically monstrous figures like Leviathan and places it squarely on human shoulders. The purveyors of chaos in Psalm 104 are not mythically theriomorphic, that is, monsters made in the image of a wild animal (thērion). Rather, they are monstrously anthropomorphic.

2.4 Creation in the wisdom literature

The so-called ‘wisdom’ literature of the Hebrew Bible (Proverbs, Job, and Ecclesiastes) is eclectically diverse in content and form, making the notion of wisdom difficult to define. Nevertheless, certain salient features can be identified in this corpus, including the pervasive theme of creation. In the place of national history and worship, as one finds elsewhere in the Old Testament, including the Psalms, creation and common human experience constitute the sages’ primary foci. Three texts in particular profile creation in remarkably divergent ways, one from each wisdom book. We begin with Proverbs.

2.4.1 Proverbs 8:22–31: Wisdom at play in creation

One of the most exquisitely crafted creation poems in all of scripture is found in Prov 8:22–31, a creation account unlike any other. The poem lingers over Yhwh at work in the methodical construction of the cosmos. But this account is punctuated repeatedly by an ‘I’-witness who turns this cosmic litany into an intensely personal testimony. That ‘I’-witness is Wisdom herself. As for its genre, scholars liken Wisdom’s poem to an ‘aretalogy’, a form of self-praise often associated with the goddess Isis (Schipper 2019: 292–294). But it is through her self-praise that Wisdom bears witness to Yhwh’s praiseworthy creation.

The poem opens with Wisdom placing herself at the beginning of Yhwh’s creative acts (vv. 22–23). Wisdom is ‘created’ or ‘acquired’ (the Hebrew verb q-n-h is ambiguous – cf. Exod 15:16 and Gen 14:19–22) by Yhwh prior to the creation of the world, she testifies, thus asserting her preeminent status vis-à-vis creation. Wisdom describes her own genesis and development: she is conceived in v. 22, gestated in v. 23, birthed in vv. 24–25, present during creation in v. 27, and ‘playing’ in vv. 30–31. Her self-identity in v. 30a remains enigmatic in the history of interpretation, all dependent on the translation of one word: ’āmôn, whose various meanings proposed by the ancient witnesses and scholarly proposals alike range from ‘master craftsman’ to ‘child’ to the adverbial translation ‘continually’ (for discussion, see Schipper 2019: 312–314). Regardless of translation, the verbal language of vv. 30–31 describes Wisdom as a playing child. While her origin is sharply distinguished from the origins of the cosmos, Wisdom nevertheless shares an intimate bond with the ‘inhabited world’ (v. 31).

Wisdom recounts Yhwh at work in carving, anchoring, stabilizing, establishing, circumscribing, and securing creation. Unlike in Genesis 1, no divine ‘word’ or command is spoken. Instead, Yhwh is all action. Yhwh sets boundaries to limit the sea from overwhelming the land. Yhwh draws a ‘circle’ upon ‘the face of the deep’ and lays the ‘foundations of the earth’ with exacting measurements. Yhwh ‘sinks’ the mountains to serve as weight-bearing pillars to hold up the heavens so as to prevent cosmic collapse. Yhwh constructs a universe that is safe and secure, setting the cosmic infrastructures and boundaries firmly in place, all to maintain the world’s order and stability.

Except for a glancing reference in the final verse, absent is any specific reference to the creation of life, human or otherwise, in Wisdom’s grand soliloquy. In relation to Wisdom, Yhwh is not just the architect of the cosmos; the Deity of design is also a doting, playful (not to mention single) parent. To put it pointedly, Wisdom presents herself as God’s only begotten daughter. No wonder this passage, particularly v. 22, was fought over in the Arian christological controversy of the early church (see Ticciati 2019: 179–190). Conceiving Wisdom as a child, in any case, highlights a profoundly sapiential paradox: Wisdom grows in wisdom, and she requires a world to do so. All the world was made for her delight and edification. As a complement to God’s cosmic temple in Genesis 1, creation serves as Wisdom’s playhouse in Proverbs 8.

2.4.2 Job 38–41: God’s wild kingdom

Job features one of the most evocative and detailed portrayals of creation in all of scripture, surprisingly so for a book that focuses almost exclusively on a single person’s suffering. In two fell swoops, a man of unassailable moral rectitude and great wealth, the ‘greatest of all the people of the east’ (1:1, 3), is stripped of all security, prosperity, and health; all the while his character is attacked with increasing vehemence by his friends in the guise of ‘comfort’ (2:11). With Job’s own world turned upside down, socially, economically, and existentially, Yhwh responds by presenting to Job a world, indeed a cosmos, that extends far beyond his own imagination (chs. 38–41).

Yhwh’s answer consists of two speeches (Job 38:1–40:2 and 40:6–41:34), each of which is introduced with the challenge for Job to ‘gird’ himself. The first challenge addresses Yhwh’s cosmic ‘design’ (Hebrew ‘ēṣâ, 38:1); the second deals with Yhwh’s ‘justice’ or governance (Hebrew mišpāṭ, 40:8). The overall movement of Yhwh’s twofold answer is telling: it begins with detailing the cosmic expanses and moves toward recounting various phenomena, meteorological and biological, concluding with a detailed ‘study’ of one creature, Leviathan. As creation’s purview zooms from the cosmic to the particular, Yhwh’s poetic account runs counter to the narrative logic of the ancient mythos of creation, which typically begins with chaos, proceeds to conquest, and concludes with creation. In Enuma elish, Tiamat’s defeat sets the stage for creation, which concludes with the gods’ adulation of the conquering creator Marduk. In Genesis 1, God begins creation amid the benign ‘chaos’ of dark waters (Gen 1:2) to fashion a cosmic temple.

Yhwh’s speeches in Job, however, move in the opposite direction. Yhwh appears to Job in the whirlwind, as a storm god. And yet, like Ba‘al in Ugaritic myth, such a terror-provoking manifestation of the divine has its salutary side: a whirlwind in the wilderness delivers rain to the parched desert (see Job 38:25–27). Divine terror proves to be a fertile terror, and it sets the tone for all that follows. From Job’s perspective, Yhwh’s discourse teaches even as it terrorizes, and it begins with the construction of a well-established earth, whose founding prompts celestial rejoicing.

Where were you when I founded the earth?
Say so, if you have understanding!
Who determined its measurements? Surely you know!
Or who extended a measuring line upon it?
On what were its footings sunk?
Or who laid its cornerstone,
when the morning stars rejoiced together,
and all the divine beings shouted for joy?

(Job 38:4–7)

The first act of creation begins with an admiring look at the earth’s foundations: measurements are determined, lines are drawn, bases are sunk, and a capstone is laid. The celestial joy that erupts from the cosmic choir suggests that the earth is the foundation of no ordinary edifice but of a cosmic temple (cf. Ezra 3:10–13). Yhwh begins in Job where God in Genesis 1 ends, proceeding from the astronomical to the meteorological and finally to the zoological, concluding with the monstrous Leviathan of the watery abyss. The Joban account of creation, in other words, proceeds in the opposite direction of most creation counts: from creation to chaos! Lacking, moreover, is any human created in the ‘image of God’ to rule the earth.

With the temple’s foundation laid, next are the sea’s confinement (Job 38:8–11) and the unleashing of light (vv. 12–15). The sea is depicted as a tempestuous newborn whose birth is facilitated by Yhwh in the role of midwife (cf. Prov 30:4). Yhwh swaddles the sea and sets firm boundaries against its ‘proud waves’. Light, however, does what it is told: it shakes out the wicked from the earth’s ‘skirts’ (v. 13). Next come creation’s depths and breadth, including the places of light and darkness (vv. 16–21). More detail is disclosed about the meteorological realm. Snow and hail are stored for ‘the day of battle’ (v. 23). Such precipitous violence has its salutary side: torrents are channelled ‘to satisfy the waste and desolate land’ (v. 25–27). Creation even at its most seemingly desolate bears life and beauty, Yhwh points out to Job. Yhwh challenges Job to bind and guide the constellations (vv. 30–33) and command the weather (vv. 34–37). Job remains silent.

Continuing the cosmic tour, Yhwh showcases five pairs of animals: lion and raven, mountain goat and deer, onager and auroch, ostrich and warhorse, hawk and vulture. With the exception of the raven and the warhorse, all of the animals listed constituted wild game for Egyptian and Mesopotamian kings. The royal hunts were not conducted for entertainment purposes, thrilling as they may have been. They were staging grounds for the king’s prowess on the battlefield, a symbolic exertion of royal power over the world. By slaying wild animals, the king was ‘fulfilling his coronation requirement to extend the kingdom beyond the city to include the wilderness’ (Dick 2006: 255). In the lion hunt specifically, the king identified himself as both the hunter and the lion; hence, the leonine carcass was never mutilated (Dick 2006: 244–245).

In Job, Yhwh presents these denizens of the wild not for Job to kill and thereby prove his physical prowess, as if he himself were a king (cf. 29:25). To the contrary, these animals are deemed free and untamable. In a remarkably ironic turn, Yhwh begins with the lion, the quintessential predator of the wild and the most prized wild game of kings, itself a symbol of royalty, and asks, ‘Can you hunt prey for the lion?’ (v. 38:39). Job is not to gird up his loins to kill lions; he does so to provide for them.

Nor are these animals to be named or defined in any way by Job, as in the ādām’s case in the garden (Gen 2:19–20). Far from it: Job is driven through the power of divine poetry into the wilderness to encounter the beasts on their own turf, in situ. Yet he discovers the wild to be full of alien life filled with inalienable value, denizens endowed with strength, dignity, and freedom. The mountain goat kids ‘go forth and do not return’ (39:4); the onager roams freely beyond human reach (v. 5); the auroch resists domestication (vv. 9–12); the ostrich fearlessly flaps its pinions before the hunter (vv. 16–18); the warhorse exults in its thunderous strength (v. 22); and the raptors soar, spying out their prey and cleaning up the battlefield (vv. 26–30).

All these animals live and move and have their being as Yhwh intended, who serves as their provider, hunting the lion’s prey (38:39), responding to the raven’s cry (v. 41), and directing the raptor’s flight (39:26). Yhwh admires each in loving detail, and with such detail Job is afforded a perspective that lies far beyond himself, a perspective that is Yhwh’s, but one that the animals also share. Job is invited to see the looming battle through the eyes of the warhorse, to spy out corpses through the eyes of the vulture, to roar for prey as the lion, to cry for food like the raven’s brood, to roam free on the vast plains, to laugh at fear, and to play in the mountains. Job’s journey so far is no descent but an ascent to Nature (Wilson 2006: 13, 163).

In Yhwh’s second speech, Job’s journey does make a deep descent. Yhwh profiles two magnificent animals that loom mythically large: Behemoth and Leviathan, perhaps drawn in part from the water buffalo (or hippopotamus) and the crocodile, both formidable creatures in their own right. (For an artistic rendering, see William Blake’s famous watercolour depiction of these two creatures from Illustrations to the Book of Job at The William Blake Archive). Whatever they are, these larger-than-life beasts are the quintessential embodiments of chaos, yet they are highly esteemed by Yhwh. Nothing is said of Yhwh’s intent to subjugate either Behemoth or Leviathan; freedom reigns for both these fearsome creatures. Behemoth is claimed as the ‘first (or chief) of God’s works’ (40:19), taking on the preeminent status of Wisdom in Proverbs 8 and cosmic light in Genesis 1.

Behold Behemoth, which I made with you!
It eats grass like an ox.
Behold its potency in its loins,
and its power in the muscles of its belly.
It stiffens its tail like a cedar;
the sinews of its thighs are intertwined.
Its bones are tubes of bronze;
its limbs are like a rod of iron.
It is the first of God’s works;
[Only] the one who made it can approach it with sword.

(Job 40:15–19)

Nowhere in Yhwh’s answer is mention made of humanity, let alone humanity’s dominion. This is no anthropocentric world that is profiled by Yhwh. The world, rather, is a glorious hodgepodge of life in all its awe-filled and fierce variety, a wondrous plurality devoid of a centre or summit. But here in Yhwh’s presentation of Behemoth, Job receives a clue to his place in Yhwh’s wild creation. ‘Behold Behemoth, which I made with you (‘immāk)’. Job shares a connection ‘with’ the monstrous Behemoth, likely connoting a fraternal connection such as the one that Job complains about regarding the jackals and ostriches in 30:29. In any case, Behemoth and Job are deemed fellow creatures, and by extension all the creatures of the wild. For all the alien otherness of creation, Job finds his place in the company of such creatures, a stranger among strangers. This single preposition invites reflection on what Job shares with these creatures of the wild, beginning with Behemoth: alien identity, resistance to control, fierceness. In Yhwh’s creation, Job not only discovers himself sharing common creaturehood with the wild; he also sees something of himself in each of these creatures, all sharing in the irrepressible exercise of life. In his bewilderment Job is ‘be-wilded’.

Yhwh’s answer to Job concludes with Job 41, the only chapter in the Bible devoted entirely to a single (albeit mythic) animal. With Leviathan, Job takes the plunge into the depths of chaos. This monstrous figure marks the culmination of creation in Job with these final words:

On the earth there is nothing like it,
a creature made without fear.
It surveys all who are lofty;
it is king over all the sons of pride.

(Job 41:33–34 [Heb. 25–26])

In Yhwh’s world, this monster of the deep not only thrives but also assumes unrivalled royal status (41:26; cf. 40:11–12). It is Leviathan, not Job, who bears such status. So much for Job’s self-fancy as king (29:25).

What kind of world does Yhwh present to Job? A world that is terrifyingly and wondrously vast and alien, teeming with life characterized by fierce strength, inalienable freedom, and wild beauty (see O’Connor 2004: 48–56). Limits, to be sure, are set in place: the earth rests on stable foundations, the tempestuous sea is contained like a swaddled infant, and the dawn renews the earth with some semblance of order (38:4–15). Nevertheless, the world is not an object of divine micromanagement or control (Fretheim 2005: 235, 239). Land, sea, and sky are host to myriad life-forms, all alien to the human eye and untamable to the human hand, but all affirmed and sustained by Yhwh. Yhwh’s world is filled with scavengers and predators, even monsters (cf. Gen 1:21), all co-existing and thriving. This world is God’s wild kingdom. It pulses with ‘pizzazz’ (Davis 2001: 139).

Job and Genesis offer contrasting portrayals of creation. Whereas creation in Genesis 1 is conceived as a cosmic temple, creation in Job is more of a cosmic tempest, bursting with vitality and freedom. In Genesis 1, creation proceeds methodically from benign ‘chaos’ to cosmic order. In Job, the order of presentation moves from foundational order to royal chaos (Leviathan). In Genesis 2, the animals are presented to the ’ādām in the garden for their naming. Job, on the other hand, experiences something of the native habitats and habits of these animals, from which Job learns their names. In Genesis 1, creation is deemed ‘good’. While some sense of goodness is presupposed in the Joban account of creation, added to this goodness is the value of alterity: creation is filled with strange and alien creatures endowed with inalienable worth by God. In Genesis 1, as well as in Psalm 8, creation is hierarchically defined with humanity receiving the ‘blessing’ of dominion. In Job, humanity assumes no such role; indeed, humanity as a class or species is absent in Yhwh’s presentation. If one wants to find a royal figure in creation, Leviathan, the quintessential creature of chaos, is the only candidate that qualifies in Job. Likewise, ‘image of God’ language applied to humanity is nowhere evident in Job, perhaps because the Joban poet considers all creation made in God’s image in so far that creation reflects in varying degrees God’s wisdom and might. Often noted is the theophanic imagery associated particularly in the figures of Leviathan and the warhorse (see Newsom 2003: 243, 251, 261; Habel 1985: 547). In any case, Job offers a radical revision of Irenaeus’s often quoted line, ‘The glory of God is a living human being’ (Gloria Dei est vivens homo [Adversus Haereses, 4.20]). In Job the glory of God is a fully living creation.

2.4.3 Ecclesiastes 1: creation without cause, pause, and effect

Featuring the musings of the sage Qoheleth (‘Teacher’ in most translations), the book of Ecclesiastes contains the most unconventional perspective on creation in the Bible. In fact, the entire book, except for its epilogue, is so unorthodox that it has been often called the ‘strangest book in the Bible’ (Scott 1965: 191). Ecclesiastes opens not with a creation account per se but with what could be called a dynamic ‘snapshot’ of the cosmos (1:4–11). Moreover, the book concludes with further cosmological reflections in 12:2–7. Creation, thus, frames Qoheleth’s reflections on the purpose and meaning of life, so also the sage’s assessment that everything is ‘vanity’ (hevel), or better translated ‘futility’ (1:2; 12:8).

By questioning the purpose of human toil in 1:3, Qoheleth also questions the purpose of creation in vv. 4–11. Generations come and go, the sun rises and sets, the wind blows hither and yon, and the streams flow perpetually, all the while both earth and sea remain unchanged. Nothing is gained; neither progress nor purpose is discernible. The sage observes the weary sun ‘panting’ to the place of its rising, ever repeating its ‘revolutions’. Sun, wind, and streams are all set in constant motion, all returning to where they began without pause. Activity abounds, but nothing is achieved. Ever in motion but never changing, the cosmos is uniformly indifferent to human living, from birth to death. The perpetual cycles exhibit neither beginning nor ending, much less a ‘new’ beginning. This is a world without pause and effect, a world without discernible history.

The same can also be said of human agency, according to Qoheleth. As the sea is never filled, so human yearning (‘eye’ and ‘ear’) is never satisfied (v. 8b). Despite their efforts, the people of past generations will be forgotten by those who come after (v. 11). Indeed, the same fate applies to every generation. Nothing of significance is left for posterity. Establishing a legacy is a futile venture. Both natural and ‘man-made’ history are doomed to repetition, much like the sun and the wind. The past is the future; ‘there is nothing new under the sun’ (1:9). Any ‘new’ thing is simply a replay or variation of the past. Genuine change is a mirage.

For Qoheleth, the cosmos moves on its own frenetic inertia, with human history mirroring its futile movements (see Eccl 3:1–8). Accompanying the lack of newness or change in cosmic history is the lack of memory in human history (1:11). With the passing of each generation, memory is by and large wiped clean (cf. 9:5). A life that strives to ensure its legacy by pursuing gain pursues only the wind, for the past cannot be remembered any more than the future can be controlled. Likewise, creation itself is pointless: for all the energy expended in creation, nothing is gained and everything is to lose (see 2:22; 3:9; 6:11). Like cosmos, like humanity: death casts its long shadow over both (12:1–7).

Creation according to Qoheleth is emptied of telos and filled with toil, a cosmos without purpose and devoid of its own genesis. Elsewhere in the Bible, genesis and purpose are inseparably wedded. But here there is nothing, properly speaking, creative or purposeful about Qoheleth’s cosmos. Indeed, God appears not even to be involved, at least not directly. Whereas the creation traditions of Genesis, Psalms, Proverbs, and Job all claim the world as created by a beneficent deity, Qoheleth’s cosmology, for all intents and purposes, excludes cosmogony and deems God as inscrutable (3:11). As there is no beginning, there also is no discernible point to creation. Qoheleth’s world is a creation void of creation, and hevel is its name (1:2; 12:8).

The most often used word in Ecclesiastes, hevel is the book’s single-word thesis. Throughout his reflections, Qoheleth presents one example after another of hevel, from the cosmic to the personal. The word itself conjures the image of ‘vapour’, something entirely insubstantial, perhaps even noxious (see Miller 2002). And yet hevel bears a rich and varied function in Qoheleth’s discourse. There are many nuances of hevel conveyed in Ecclesiastes, for the term can be translated in a number of related ways, depending on the context: ‘futility’, ‘transience’, ‘worthlessness’, ‘absurdity’, ‘farce’, ‘mirage’, and even ‘shit’ have all been proposed (see Crüsemann 1984: 57; Fox 1986: 409–427; Brown 2000: 21–22; Sneed 2017: 879–894). But regardless of its specific nuance, it is indubitable for Qoheleth that hevel happens all the time. The incessant cycling of the elements is for him the stellar example of ‘vanity’, a cosmic exercise of futility that eventually proves ephemeral. But as science has demonstrated and documented, the (re)cycling of the elements, from carbon to water, is precisely what sustains creation (Dell 2009: 181–189).

2.5 Creation in the prophets

The prophets of ancient Israel revelled in images drawn from the world of nature. Wolves, jackals, lions, acacia trees, towering cedars, cascading waters, and blossoming deserts all populate the prophets’ messages, for woe or for weal. Judgment from their lips invoke harsh scenes of devastated landscapes, while prophetic hope and consolation conjure scenes of nature’s renewal. Creation, moreover, according to certain prophets, demonstrates God’s unrivalled sovereignty over the nations.

2.5.1 Creation in travail and restoration

Perhaps the most devastating judgment found among the prophets that bears a creational perspective comes from Jeremiah:

I looked on the earth, and it was formless and void;
and to the heavens, and their light was null.
I looked on the mountains,
and they were quaking, and all the hills were rocking back and forth.
I looked, and there was no one left,
and all the birds of the air had taken flight.
I looked, and the fruitful land was a desert,
and all its cities were laid in ruins before Yhwh, before his fierce anger.
For thus says Yhwh:
The whole land shall be a desolation; yet I will not make a full end.

(Jer 4:23–27)

Because of Yhwh’s judgment, Jeremiah sees all the earth ‘formless and void’ (tōhû wābōhû), just as it was in the beginning in Gen 1:2. What would provoke such ecological catastrophe in the prophet’s eyes? In Yhwh’s words, articulated by Jeremiah, the people ‘have no understanding’ and have become ‘skilled in doing evil’ (v. 22). The consequences of evil conduct are ecologically devastating.

But the judgment of disaster is not the final word among the prophets. On the other side of judgment is restoration and renewal, about which the prophets speak with equal boldness. Hosea, for example, speaks of the renewal of God’s marital covenant with Israel, a covenant that extends not just to Israel but, like Noah’s covenant in Genesis 9, also to the ‘wild animals, the birds of the air, and the creeping things of the ground’ (Hos 2:18). God’s new covenant is a cosmic covenant of peace that also heralds the renewal of the earth’s fertility:

On that day I will answer, says Yhwh,
I will answer the heavens
and they will answer the earth;
and the earth will answer the grain, the wine, and the oil,
and they will answer Jezreel;
and I will sow him for myself in the land.

(Jer 2:21–23)

Jezreel, the valley whose name means ‘God sows’, had suffered the blood of massacre (Jer 1:4; see 2 Kgs 9–10) but is now destined to recapture its namesake. Hosea describes the earth’s fructification in the language of discourse: ecological renewal unfolds as a chain of communication from the top down. Yhwh ‘answers the heavens’ with life-giving sustenance, the heavens do the same for the earth, and so forth. Creation is an interdependent whole, with each domain contributing to another to include a people sown in the land. This ‘food chain’ is one of provision, not predation.

2.5.2 Creation in Amos

Three hymnic passages in the book of Amos acclaim God as creator, each one embedded within a series of prophetic judgments or indictments (Amos 4:13; 5:8–9; 9:5–6). They profile the deity as one who both ‘forms the mountains’ and ‘treads on the earth’s heights’ (4:13), who ‘turns deep darkness into morning’ and ‘darkens the day into night’ (5:8), who ‘touches the earth, melting it’, and ‘founds his vault upon the earth’ (9:6). In Amos, the God who creates is also the God who destroys, cosmically mirroring salvation and judgment.

2.5.3 Creation in Second Isaiah

Similarly in Isaiah, God is said to ‘form light and create darkness, [...] make prosperity and create disaster’ (45:7), indicating a more comprehensive span of creational capacity on the part of God than found in Genesis (cf. Gen 1:2). The creation passages in Isaiah are found primarily in chapters 40–55, the portion of the book that reflects historically the time when much of Israel was exiled to Babylon after Jerusalem’s destruction (586–539 BCE) and theologically the rise of monotheism in ancient Israel’s understanding of God (e.g. Isa 45:5–6, 14, 18, 21, 46:9).

In a series of divine speeches, cast as trial speeches, Israel’s God, Yhwh, deploys the language of creation to assert not only Yhwh’s supreme sovereignty but also Israel’s international commission, as in Isa 42:5–6:

Thus says God, Yhwh,
who has created the heavens and stretched them out,
who has hammered out the earth and what emerges from it,
who has given breath to the people upon it,
and spirit to those who walk in it.
I am Yhwh,
I have called you in righteousness,
I have taken you by the hand and kept you;
I have given you as a covenant to the people,
a light to the nations.

(Isa 42:5–6)

The God who created the universe is the same God who sustains and commissions a people. The heavens are unfurled like fabric (cf. Ps 104:2); the earth is hammered out like sheet metal; Israel is called to be a ‘light to the nations’. For all its hymnic grandeur, the creation language of v. 5 serves to introduce the bold announcement in v. 6, which apart from v. 5 has ostensibly nothing to do with creation and all to do with Israel’s history. But joined to v. 5, Israel’s mission to the nations becomes an extension of God’s cosmimc, creative activity.

The prophet’s view of creation reflects the powerful historical forces that gripped his people in exile. Case in point: the prophetic poet views celestial creation as unfurled fabric:

Thus says Yhwh, who redeemed you,
and who formed you in the womb:
I am Yhwh, who makes all things,
who alone stretched out the heavens,
who by myself hammered out the earth.

(Isa 44:24)


I made the earth,
and created humankind upon it.
It was my hands that stretched out the heavens,
and I commanded all their host.

(Isa 45:12)


Indeed, my hand laid the foundation of the earth,
and my right hand spread out the heavens;
When I call them,
they stand together at attention.

(Isa 48:13)


You have forgotten Yhwh, your Maker,
who stretched out the heavens,
and laid the foundations of the earth.

(Isa 51:13a; cf. 16)

The heavens and the earth are God’s works, but in very different ways. Whereas the earth is ‘hammered out’ and its foundations laid, the heavens are ‘spread’ or ‘stretched out’. The contrast between the celestial and the terrestrial is unmistakable; so also the divergence between Genesis and Isaiah. In Genesis 1, the heavens are associated with ‘hammering out’ to form a solid ‘firmament’ (rāqîa‘), firmly demarcating the realm of the transcendent (see Gen 1:6–7). In Isaiah, however, the heavens are likened to stretchable, unfurled fabric (see also Ps 104:2b), with the earth itself cast as a kind of firmament, associated with a foundation. The ancient poet’s cosmic perspective, in fact, corresponds to his grand vision of a restored people:

Sing, O barren one who did not bear!
Burst into song and shout,
you who have not been in labour!
For the children of the desolate one will become more numerous
than the children of the married one, says Yhwh.
Enlarge the site of your tent,
and let the curtains of your habitations be stretched out.
Do not hold back;
lengthen your cords and strengthen your stakes.
For you will spread out to the right and to the left,
and your progeny will possess the nations,
and the desolate towns you will settle.

(Isa 54:1–3)

This passage clarifies the purpose of the celestial metaphor: the heavens are like the canvas of a tent. Such imagery suggests that all creation was fashioned for habitation, including Israel’s. The ‘desolate’ daughter Zion is commanded to sing as a mother of many children and to stretch out ‘the curtains of [her] habitations’, her ‘tent’ (v. 2). The promise of ‘spreading out’ in all directions follows the charge. Zion’s actions, in other words, are to replicate God’s heavenly act of creation. With the heavens cast as a celestial tent, Zion is deemed an extension of heaven on earth.

As the heavens and the earth were made for habitation, so Zion beckons the exiles to return and reinhabit the land from which they were taken. The reference to God laying the earth’s foundation also resonates with God’s promise that the temple’s ‘foundation’ be laid, to be fulfilled by the Persian king Cyrus II (Isa 44:28). The temple’s rebuilding, in fact, mirrors the earth’s founding. With the temple restored and the tent spread out – as the earth is founded and the heavens are stretched – daughter Zion becomes mother Zion, the bearer of a people’s rebirth.

As the passages above make clear, creation in Isaiah has as much to do with Israel as with the cosmos. The same God who created the heavens also ‘created’ Israel:

Now listen up, O Jacob my servant,
Israel whom I have chosen!
Thus says Yhwh, who has made you,
who has formed you in the womb and will help you:
Do not fear, my servant Jacob,
Jeshurun, whom I have chosen.

(Isa 44:1–2; see also vv. 21, 24; 54:5)

Israel’s creation is bound up with Israel’s election. To choose a people is to create a community. Israel was formed from the womb of God’s decision to constitute and sustain a people. In these passages and elsewhere in Isaiah, cosmos and community are wedded together.

Finally, in one densely-packed passage, God reveals the essential purpose of creation.

For thus says Yhwh,
who creates the heavens (he is God!),
who forms the earth and fashions it;
He established it;
he did not create it a waste (tōhû);
he formed it to be inhabited!
‘I am Yhwh,
and there is no other.
I did not speak in secret,
in a land of darkness.
I did not say to Jacob’s offspring,
‘Seek me in waste (tōhû)’.
I Yhwh speak the truth,
I declare what is right’.

(Isa 45:18–19)

This compact poem divides itself evenly into two parts. The creation part (v. 18) introduces the divine declaration that follows (v. 19). God’s intention for creation is habitation, not ‘waste’ or ‘darkness’; thus, seeking God can only take place in a reconstituted land, not one emptied and left dark (cf. Jer 4:23–26). For the poet of the exile, a land ‘emptied’ by exile and darkened by the lengthening shadows of its ruins constitutes the very antithesis of creation from the beginning. ‘Seek me’ in an ordered, life-sustaining creation is God’s invitation to a displaced and despairing people. On the other hand, seeking God in a wasteland is simply a ‘waste’.

3 Science and creation

To read these biblical creation texts scientifically – a temptation among some interpreters – is a hermeneutical dead end, given the simple fact that ancient texts about the physical world are not scientifically informed. Biblical texts are theological texts, not scientific ones. Nevertheless, there is rising interest among some scholars in holding the separate disciplines of theology and science in constructive dialogue regarding the natural world, and to do so would include interpreting the biblical text in conversation with science (see Brown 2010).

For example, one can acknowledge that Genesis 1 describes how the world was created in quite ‘naturalistic’ ways. As essayist and novelist Marilynne Robinson observes, ‘If ancient people had consciously set out to articulate a worldview congenial to science, it is hard to imagine how, in terms available to them, they could have done better’ (Robinson 1998: 39). Readily evident, for example, is the narrative movement from simplicity to complexity regarding life, culminating in the creation of humanity, a movement that runs somewhat parallel to biological evolution. Moreover, the creation of light in Gen 1:3 marks the beginning of time, making possible the alternation between day and night, beginning with the first ‘day’ (v. 5). For modern readers, it is hard not to think of the Big Bang of modern cosmology, that explosive moment of cosmic energy (‘inflation’) that made possible both time and space, resulting eventually in all physical matter as we know it. The association, albeit anachronistic, does underline the momentous import of light as the beginning point of creation in Genesis. One can also recognize the creation of the celestial spheres on the fourth ‘day’ as distinct from the creation of light on the first ‘day’ bearing a measure of astrophysical resonance, given the fact that the first generation of stars did not emerge until roughly half a billion years after the Big Bang (for more discussion, see Brown 2010: 36–41).

On the other hand, the specific order of creation in Genesis 1 differs significantly from cosmic evolution as reconstructed by modern cosmologists: most obvious is the creation of land or ‘earth’ before that of the sun. Planets are formed by a process of ‘accretion’, which presupposes the existence of a host star. Moreover, Genesis claims creation to be completed in seven ‘days’, whereas modern cosmology discerns the universe to be circa 13.8 billion years old. Because the ‘days’ of Genesis 1 are defined by the alternation of ‘evening and morning’, they cannot be considered anything other than twenty-four-hour days in the ancient context of the text. Attempts to harmonize Genesis 1 and cosmic evolution by allegorizing the ‘days’ of Genesis (cf. Ps 90:4; 2 Pet 3:8) falter on having to assign wildly different chronological lengths for each day, not to mention failing to reconcile the order of creation presented in Genesis with that of modern cosmology. Like all other biblical accounts of creation, Genesis 1 is no scientific text. Biblical creation stories are theologically driven, culturally based descriptions of the world as understood by the ancient authors of scripture. They are, in other words, myths; and because they are myths, they share profound truths about the world and humanity’s place in it that reach beyond science.

There are, however, ways to appreciate the biblical text from a scientific perspective that highlight its wisdom. Take, for example, the Garden of Eden account in Genesis 2–3. This ancient and deceptively simple story is no account of human evolution. It is an etiology, a story situated in the primordial past that is meant to explain or justify the author’s physical and culture-based realities, such as the culture of patriarchy (3:16b), as well as why snakes slither on the ground and women experience excruciating pain during childbirth (vv. 14, 16a). Nevertheless, there are certain aspects of the story that are broadly resonant with anthropological observations about Homo sapiens and its evolution. First, the ’ādām in Genesis is cast developmentally, a creature who initially is only task-oriented but ultimately gains consciousness and moral awareness. Nakedness and the need for clothing, moreover, are anthropological distinctions that set humans apart from other primates. The pain of childbirth is also considered a uniquely human distinction within the ancient narrative, due to the comparatively large head of the human infant traveling through the birth canal. Finally, the Yahwist account acknowledges the common origin and substance of all animal life, the ‘ground’. Call it the ‘ground of being’ (cf. Tillich 1951: 235–236). Biologists identify DNA as the common genetic thread of all life on earth.

Reading Psalm 104 through the lens of biology compels the reader to appreciate the way in which the psalmist celebrates the breadth of life in all its diversity. Each species has its home and sustenance in creation. As life’s species are varied and numerous, so also are their habitats and niches. While about 1.8 million species have been officially catalogued, most experts believe that around 10 million species actually exist (Pimm and Bernstein 2008: 14). Considering all the microbes and other tiny organisms awaiting discovery, some suggest 100 million or more (Baker 2006: 132). Organisms not only respond to their respective environments by their ‘fit’; they also transform them, thereby becoming agents of their own evolution. Not only do species evolve, so do their environments. Organisms are not passive agents as they achieve fitness within their respective environments. They are niche constructors (Odling-Smee 1988: 73–132). Termites, for example, create a new environment by constructing mounds for shelter and climate control. Mounds, burrows, and dams are all examples of such niche construction. Equally illustrative is the example given by the psalmist: birds building nests (Ps 104:17). The animals referenced in Psalm 104 are active agents in turning their environments into habitats, their places into niches.

Although biblical texts are by no means scientific and in many cases cannot be reconciled with scientific understandings of the natural world, familiarity with science can highlight certain emphases and nuances in these ancient texts that might otherwise be overlooked, as noted above. Because these texts are theological, affirming in different ways God as creator and sustainer of creation, there is warrant for exploring the biblical accounts of creation in conversation with science. As a discipline, theology strives to relate everything to God, including the natural world. If theology is ‘faith seeking understanding’, to quote Anselm, and science is a form of understanding seeking further understanding, then theology has nothing to fear from science and all to gain from it. Reading the ancient creation texts of the Bible in dialogue with science constitutes an integral part of the theological task.

4 Conclusion and relevance

It is well-nigh impossible to summarize the Bible’s perspective on creation, given its irreducible diversity of accounts and passages, from Genesis to Ecclesiastes, including various psalms and portions from the Prophets. By depicting creation established in seven days, Genesis 1 offers a view of creatio completa, whereas Psalm 104, with its emphasis on God’s providential care, provides a vivid depiction of creatio continua. While Genesis 1 does not begin absolutely ‘in the beginning’ with creation conceived out of nothing, Isaiah comes closer to the so-called doctrine of creatio ex nihilo with the statement: ‘I fashion light and create darkness, I make prosperity and create disaster; I am Yhwh who does all these things’ (Isa 45:7; cf. Gen 1:2). Even closer is the statement in 2 Macc 7:28: ‘I beg you, my child, to look at the heaven and the earth and see everything that is in them, and recognize that God did not make them out of things that existed’ (NRSV). By contrast, the perspective offered in Ecclesiastes leaves little room for creation as created. Everything that is has always been, according to Qoheleth; ‘there is nothing new under the sun’ (Eccl 1:9b).

The place of humanity in creation, moreover, varies widely, from human dominion over the rest of creation (Gen 1; Ps 8) to equal co-existence with other creatures (Ps 104). With its emphasis on the freedom and dignity of wild creatures, Yhwh’s depiction of the creation in Job deliberately omits humanity in its panoramic sweep, countering provincial, anthropocentric models of creation. By commanding all creatures to give praise to God, both human and nonhuman, animate and ‘inanimate’, Psalm 148 gives all creatures an equal playing field (or better ‘praising’ field). Or in the final verse of the final psalm of the Hebrew Psalter: ‘Let everything that breathes praise Yhwh!’ (Ps 150:6). Call it an ‘ecology of praise’.

The creation traditions of the Old Testament offer guidance and inspiration for communities of faith to see the world not as a warehouse of resources for human consumption but as a planetary community of life whose diversity and flourishing are valued by God. The Old Testament affirms that human beings, constituting the most powerful species on the planet, have a pivotal role to play in order to ‘serve’ and ‘preserve’ creation (cf. Gen 2:15). For too long the ‘dominion’ model exemplified in Genesis 1 and Psalm 8 has been interpreted as free licence to exploit the earth in complete disregard of the damaging effects on the planet. The Bible challenges its readers to consider new ways of exercising human power, tremendous as it is, to ensure the flourishing of creation rather than its diminishing. While the ‘image of God’ language of Genesis 1 acknowledges the God-given dignity of every human being (Gen 1:26–28), each ‘fearfully and wonderfully made’ (Ps 139:14), creation according to the book of Job acknowledges that all creatures reflect something of God’s glory. Moreover, Genesis 2 acknowledges the interdependence of all life as emergent from the ‘ground’.

As for the ‘ground’, certain biblical traditions treat not only biologically-based creatures but also domain-based creations as living: ‘mountains and all hills’ give praise (Ps 148:9a); the ‘earth’ and the ‘waters’ are agents of creation (Genesis 1); the land can ‘vomit’ out its inhabitants (Lev 18:25, 28). They are ‘living landscapes’ (Jørstad 2019). While we tend to view such entities from our rationalistic mindset as inert or ‘inanimate’, the Bible underlines their vitality, given their vital importance in the creation and sustenance of life. As ‘groundling’ (’ādām), humanity is inseparably tied to the ‘ground’.

So also human community and creation: their link is solidified in justice, or lack thereof, according to Hosea:

Hear the word of Yhwh, people of Israel:
Yhwh has an indictment against the inhabitants of the land.
There is no faithfulness or covenant loyalty (ḥesed),
and no knowledge of God in the land.
Swearing, lying, and murder, and stealing and adultery burst forth;
bloodshed follows bloodshed.
Consequently, the land mourns,
and all who live in it languish;
together with the wild animals and the birds of the air,
even the fish of the sea are perishing.

(Hos 4:1–3)

The prophet correlates the violation of justice within the community, cast in part as violations of the Decalogue, to the languishing of the land and its inhabitants. Justice within the community, in turn, results in justice for creation, and vice versa. Such texts today call for the ‘beloved community’ to act on behalf of the biotic community, to act toward healing, repairing, restoring, and renewing creation. Black liberation theologian James Cone said it best: ‘The fight for justice cannot be segregated but must be integrated with the fight for life in all its forms’ (Cone 2000: 36).

Attributions

Copyright William P. Brown ORCID logo (CC BY-NC)

Bibliography

  • Further reading

    • Brown, William P. 2000. Ecclesiastes. Interpretation: A Bible Commentary for Teaching and Preaching. Louisville: John Knox Press.
    • Brown, William P. 2010. The Seven Pillars of Creation: The Bible, Science, and the Ecology of Wonder. New York: Oxford University Press.
    • Brueggemann, Walter. 1996. ‘The Loss and Recovery of Creation in Old Testament Theology’, Theology Today 53, no. 2: 177–190.
    • Dell, Katharine J. 2009. ‘The Cycle of Life in Ecclesiastes’, Vetus Testamentum 59: 181–189.
    • Fretheim, Terrence E. 2005. God and World in the Old Testament: A Relational Theology of Creation. Nashville, TN: Abingdon.
    • Grinspoon, David. 2016. Earth in Human Hands: Shaping Our Planet’s Future. New York: Grand Central.
    • Hiebert, Theodore. 1996. The Yahwist’s Landscape: Nature and Religion in Early Israel. New York: Oxford University Press.
    • Middleton, Richard J. 2005. The Liberating Image: The Imago Dei in Genesis 1. Grand Rapids: Brazos.
    • Rasmussen, Larry. 2013. Earth-Honoring Faith: Religious Ethics in a New Key. Oxford: Oxford University Press.
    • Smith, Mark S. 2010. The Priestly Vision of Genesis 1. Minneapolis: Fortress.
    • Walton, John H. 2009. The Lost World of Genesis One: Ancient Cosmology and the Origins Debate. Downers Grove: IVP.
    • Welker, Michael. 1999. Creation and Reality. Edited by John F. Hoffmeyer. Minneapolis: Fortress.
    • White, Lynn Townsend Jr. 1967. ‘The Historical Roots of Our Ecological Crisis’, Science 155: 1203–1207.
  • Works cited

    • Baker, Catherine. 2006. The Evolution Dialogues: Science, Christianity, and the Quest for Understanding. Edited by James B. Miller. Washington, DC: AAAS.
    • Brown, William P. 2010. The Seven Pillars of Creation: The Bible, Science, and the Ecology of Wonder. New York: Oxford University Press.
    • Brueggemann, Walter. 1996. ‘The Loss and Recovery of Creation in Old Testament Theology’, Theology Today 53, no. 2: 177–190.
    • Cone, James. 2000. ‘Whose Earth Is It Anyway?’, Cross Currents 50: 36–46.
    • Crüsemann, Frank. 1984. ‘The Unchangeable World: The “Crisis of Wisdom” in Koheleth’, in The God of the Lowly: Socio-Historical Interpretations of the Bible. Edited by W. Schttroff and W. Stegemann. Translated by M. J. O’Connell. Maryknoll, NY: Orbis Books, 57–77.
    • Davis, Ellen F. 2001. Getting Involved with God: Rediscovering the Old Testament. Cambridge, MA: Cowley.
    • Dick, Michael B. 2006. ‘The Neo-Assyrian Royal Lion Hunt and Yahweh’s Answer to Job’, Journal of Biblical Literature 125, no. 2: 243–270.
    • Eichrodt, Walther. 1961. Theology of the Old Testament. Old Testament Library. 2 vols. Translated by J. A. Baker. Philadelphia, PA: Westminster.
    • Fox, Michael V. 1986. ‘The Meaning of Hebel for Qoheleth’, Journal of Biblical Literature 105, no. 34: 409–427.
    • Fretheim, Terrence E. 2005. God and World in the Old Testament: A Relational Theology of Creation. Nashville, TN: Abingdon.
    • Grinspoon, David. 2016. Earth in Human Hands: Shaping Our Planet’s Future. New York: Grand Central.
    • Gunkel, Hermann. 2006. Creation and Chaos in the Primeval Era and the Eschaton: A Religio-Historical Study of Genesis 1 and Revelation 12. Translated by K. William Whitney, Jr. Grand Rapids: Eerdmans.
    • Habel, Norman C. 1985. The Book of Job: A Commentary. Old Testament Library. Philadelphia, PA: Westminster.
    • Heschel, Abraham Joshua. 1951. Sabbath: Its Meaning for Modern Man. New York: Farrar, Straus, and Young.
    • Hiebert, Theodore. 1996. The Yahwist’s Landscape: Nature and Religion in Early Israel. New York: Oxford University Press.
    • Hillel, Daniel. 2006. The Natural History of the Bible: An Environmental Exploration of the Hebrew Scriptures. New York: Columbia University Press.
    • Hundley, Michael B. 2013. ‘Sacred Spaces, Objects, Offerings, and People in the Priestly Texts: A Reappraisal’, Journal of Biblical Literature 132, no. 4: 749–767.
    • Jenson, Philip P. 1992. Graded Holiness: A Key to the Priestly Conception of the World. JSOT Supplement 106. Sheffield: JSOT Press.
    • Jørstad, Mari. 2019. The Hebrew Bible and Environmental Ethics: Humans, Non-Humans, and the Living Landscape. Cambridge: Cambridge University Press.
    • Levenson, Jon D. 1994. Creation and the Persistence of Evil: The Jewish Drama of Divine Omnipotence. Princeton: Princeton University Press. 2nd edition. First published 1988.
    • Martínez, Florentino García. 2007. Qumranica Minora II. Studies on the Texts of the Desert of Judah 64. Edited by E. J. C. Tigchelaar. Leiden: Brill.
    • McBride, S. Dean Jr. 2000. ‘Divine Protocol: Genesis 1:1–2:3 as Prologue to the Pentateuch’, in God Who Creates: Essays in Honor of W. Sibley Towner. Edited by William P. Brown and S. Dean McBride Jr. Grand Rapids: Eerdmans, 3–41.
    • Middleton, Richard J. 2005. The Liberating Image: The Imago Dei in Genesis 1. Grand Rapids: Brazos.
    • Miller, Douglas B. 2002. Symbol and Rhetoric in Ecclesiastes: The Place of Hebel in Qohelet’s Work. Academia Biblica 2. Atlanta: Society of Biblical Literature.
    • Newsom, Carol A. 2003. The Book of Job: A Context of Moral Imaginations. Oxford: Oxford University Press.
    • O’Connor, Kathleen. 2004. ‘Wild, Raging Creativity: Job in the Whirlwind’, in Earth, Wind, & Fire: Biblical and Theological Perspectives on Creation. Edited by Carol J. Dempsey and Mary Margaret Pazdan. Collegeville, MN: Liturgical Press, 48–56.
    • Odling-Smee, F. J. 1988. ‘Niche-Constructing Phenotypes’, in The Role of Behavior in Evolution. Edited by J. C. Plotkin. Cambridge, MA: MIT Press, 73–132.
    • Pimm, Eric Chivian, Stuart L., Maria Alice S. Alves, and Aaron Bernstein. 2008. ‘What Is Biodiversity?’, in Sustaining Life: How Human Health Depends on Biodiversity. Edited by Erich Chivian and Aaron Bernstein. New York: Oxford University Press, 3–28.
    • von Rad, Gerhard. 1966. ‘The Theological Problem of the Old Testament Doctrine of Creation’, in The Problem of the Hexateuch and Other Essays. New York: McGraw-Hill, 131–142. First published 1936
    • Robinson, Marilynne. 1998. The Death of Adam: Essays on Modern Thought. Boston: Houghton Mifflin.
    • Schipper, Bernd U. 2014. ‘Egyptian Background to the Psalms’, in The Oxford Handbook of the Psalms. Edited by William P. Brown. New York/Oxford: Oxford University Press, 57–75.
    • Schipper, Bernd U. 2019. Proverbs 1-15. Hermeneia. Edited by Thomas Krüger. Translated by Stephen Germany. Minneapolis: Fortress.
    • Schüle, Andreas. 2005. ‘Made in the “Image of God”: The Concepts of Divine Images in Gen 1-3’, Zeitschrift Für Die Alttestamentliche Wissenschaft 117: 1–20.
    • Scott, R. B. Y. 1965. Proverbs, Ecclesiastes. Anchor Bible 18. Garden City, NY: Doubleday.
    • Smith, Mark S. 2010. The Priestly Vision of Genesis 1. Minneapolis: Fortress.
    • Sneed, Mark. 2017. ‘Hebel as “Worthless” in Qoheleth: A Critique of Michael V. Fox’s “Absurd” Thesis’, Journal of Biblical Literature 136, no. 4: 879–894.
    • Ticciati, Susannah. 2019. ‘Proverbs 8:22 and the Arian Controversy’, in Reading Proverbs Intertextually. Library of Hebrew Bible/Old Testament Studies 629. Edited by Katharine J. Dell and Will Kynes. London/New York: T&T Clark, 179–190.
    • Tillich, Paul. 1951. Systematic Theology. Volume 1. Chicago: University of Chicago Press.
    • Walton, John H. 2009. The Lost World of Genesis One: Ancient Cosmology and the Origins Debate. Downers Grove: IVP.
    • Welker, Michael. 1999. Creation and Reality. Edited by John F. Hoffmeyer. Minneapolis: Fortress.
    • White, Lynn Townsend Jr. 1967. ‘The Historical Roots of Our Ecological Crisis’, Science 155: 1203–1207.
    • Wilson, Edward O. 2006. The Creation: An Appeal to Save Life on Earth. New York: W. W. Norton.
    • Wright, David P. 1996. ‘Holiness, Sex, and Death in the Garden of Eden’, Biblica 77: 305–329.
    • Wright, G. Ernest. 1950. The Old Testament Against Its Environment. London: SCM.
    • Wright, G. Ernest. 1952. God Who Acts: Biblical Theology as Recital. London: SCM.

Academic tools