How the Periodic Table of Available Elements Shaped Natural History

The billion-year scale of natural history alters the common relationship between ... as active, then a natural history can be told in which elements w...
1 downloads 0 Views 1MB Size
Chapter 3

How the Periodic Table of Available Elements Shaped Natural History Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

Benjamin J. McFarland* Department of Chemistry and Biochemistry, Seattle Pacific University, Seattle, Washington 98119, United States *E-mail: [email protected].

The narrative of natural history, normally told in terms of geology and biology, can be better understood by accounting for environmental and chemical factors represented in the periodic table. The inorganic chemist R.J.P. Williams wrote extensively about how the elements were constrained in form, function, and availability throughout deep time. Geological and biological observations from natural history can be explained with chemistry. For example, the distinction between “old” elements used by simpler, earlier microbes and “new” elements used by more complex organisms can be explained using solubilities and redox potentials.

Introduction The title “Elements Old and New” provokes two questions: “How old?” and “How new?” Ancient humans recognized several “old” elements thousands of years ago that were isolated and purified by geological processes themselves billions of years old. Those same processes actively sorted elements and made them available in different environments throughout natural history. Some elements predominated the ecosphere four billion years ago and can be considered truly old, while others came to predominate a mere billion years ago and can be considered new (or at least newer). The billion-year scale of natural history alters the common relationship between chemistry and biology. In broad terms, biology studies living organisms on the meter-to-micrometer scale, biochemistry studies macromolecules on the micrometer-to-nanometer scale, chemistry studies molecules and atoms on the © 2017 American Chemical Society

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

nanometer-to-picometer scale, and particle physics studies subatomic particles at even smaller scales. However, large scales in time and space include geochemical changes of large amounts of atoms and molecules, which can be explained by appealing to chemical priciples of reactivity such as periodic trends. This is chemistry on the megameter scale, describing reactions that influenced the entire planet. At this scale, chemical changes defined the geological arena in which biological competition, mutation, and evolution could take place. Life filled the available space and flowed through diverse environments on the planet — yet it was all built from the same periodic table, on which, for example, oxygen holds a unique place in its distinctive abundance and reactivity. Constant chemical laws, expressed through elemental availability, shaped the contingencies and uncertainties of individual organisms and ecosystems. In this essay, the title “Elements Old and New” is interpreted ecocentrically rather than anthropocentrically. Elemental isolation, discovery, and refinement are usually considered to be human activities. However, if natural processes may be described as active, then a natural history can be told in which elements were isolated, discovered, and before humans evolved:

1. 2. 3.

Isolation. The oceans isolated different elements in the liquid phase according to their solubility and abundance. Discovery. Living cells discovered the chemical utility of particular elements for novel reactions through natural selection. Refinement. Through pumps and proteins, microbes, animals, and plants refined certain elements into particular extracellular and subcellular compartments.

As a result, the earth over time underwent a redox-dependent sequence of elemental availability, dominated by certain elements long ago and different elements more recently, as the periodic table of elements available to life expanded and contracted. I have described this natural history at book length for a general audience in A World from Dust: How the Periodic Table Shaped Life (1). This essay provides new data supporting this chemical narrative by emphasizing how the Earth’s changing environment shaped chemical availability, which in turn shaped the chemical capabilities and functions available to life. In this way, the periodic table continues to explain the grand arc of the narrative of natural history.

Chemical Logic Explains Elemental Abundance in Different Environments The periodic table is by definition incomplete, because larger elements can always be formed, but it has become more complete over time as more elements were made. The first periodic table would contain two elements representing the mix of hydrogen and helium immediately after the Big Bang. Stellar nucleosynthesis fused lighter elements into heavier elements (2), and the periodic 68 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

table grew over time into the diverse set of 90 naturally occurring elements that now constitute the environment. Even 13 billion years of this process has not been sufficient to alter the overall dominance of the two lightest elements; for example, in a diffusion cartogram of the periodic table weighted by current abundance in the universe, only hydrogen and helium can be clearly seen (3). A graph of universal abundance organized by atomic number is also dominated by these two elements, even when using a logarithmic scale (Figure 1). The universe differentiated into diverse environments, with different proportions of elements sorted into each by physical processes. Within stars, gravity pulled heavier elements toward the core. Gravity also pulled heavier elements together outside of stars, so that rocky planets like the Earth are enriched in heavier elements like magnesium, calcium, iron, aluminum, silicon, and oxygen. Another diffusion cartogram of the periodic table, this one weighted by elemental abundance in the Earth’s crust, lists these six elements most prominently (3). The same six elements comprise Robert Hazen’s “Big Six” classification of abundant geological elements (4). Gravity collected these rocky elements together in the proto-Earth as the solar wind blew hydrogen and helium into interplanetary space. From the earliest days of the Earth, the periodic table of elements abundant on the planet was arguably more complex than that of most of the universe. As the Earth cooled, a solid crust formed on its liquid mantle, on which liquid water condensed to form oceans beneath a nascent, gaseous atmosphere. Three of these four environments were fluid, allowing elements to mix, react, and exchange. Even the solid environment of the crust was mixed by subduction into the mantle, moved through the processes of plate tectonics. Because of this contact and mixing, each of these environments gathered a different combination of available elements through a combination of chemical properties such as melting points and solubility products. Because these dynamic processes form stable, persistent cycles such as the water cycle, together they resemble a chemical reactor, so that environmental chemists are fond of describing the earth as a “giant chemical reactor.” This exact phrase can be found in diverse places in the atmospheric chemistry and geochemistry literature (5–7), because it describes well how chemical laws mixed and sequestered different forms of the elements in different environments. The limited availability of elements and distinct patterns of chemical reactivity combine to imply that all possible combinations of elements can not be realized in all environments. The partitioning of elements into different environments can be understood with chemical logic, by comparing chemical properties and thermodynamic parameters. One example of this explains why free ozygen was absent from the ancient atmosphere. Both oxygen and sulfur would have been abundant on the early earth, and each could have combined with many elements because of the physical processes mixing the planet’s environments. If for each available element, we compare the heats of formation of oxides to heats of formation of sulfides, a thermodynamic advantage is found for oxygen reacting with aluminum, silicon, carbon, and magnesium because of their larger standard enthalpies of formation relative to their sulfides (Figure 2). 69

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

Figure 1. Relative abundance of all elements in the universe, arranged by atomic number. Reproduced with permission from ref. (1). Copyright 2016 Oxford University Press.

These four elements would have been so abundant on the surface of the early earth that most of the oxygen would have reacted with them, leaving little left to react with the other elements and still less to reside in the atmosphere as O2 gas. Therefore, the low levels of oxygen on the early Earth despite oxygen’s elemental abundance, as observed by multiple lines of geological evidence (4), can be assigned a chemical explanation on the basis of thermodynamic reference tables. When the earth was first formed, oxygen was incorporated into the Earth’s crust as oxides and into the oceans as water, and its free, diatomic form was essentially absent from the atmosphere. After that, over billions of years, oxygen levels increased to today’s levels, as shown by several lines of geological evidence, including the presence of iron oxide in the global geological oxidized precipitate known as the Banded Iron Formations (Figure 3). 70 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

Figure 2. Elements sorted by preference for oxygen over sulfur, calculated from standard heats of formation. Reproduced with permission from ref. (1). Copyright 2016 Oxford University Press.

Life Requires Elements That Are Both Available and Functional In the anoxic Archaean environment, microbial life harnessed and dissipated energy gradients in order to persist and do work, which required a persistent chemical environment inside the cell relative to the fluctuations in the external environment. Biochemical processes used the liquid phase in order to respond to fluctuations in the gradients, and so the material used for life’s reactions had to be soluble in water. Many have hypothesized that thermal and geochemical gradients at underwater vents would have provided the matter and energy needed for life (8), although others suggest that geothermal pools on the surface could have served a similar function (9). Wherever it happened, life emerged and evolved multiple chemical mechanisms to persist, drawing material from the environment in order to build and maintain dynamic, complex structures of proteins, sugars, lipids, and DNA. Biological possibilities were constrained and restricted by the chemical form, reactivity, and availability of elements, as illustrated by the “arsenic life” controversy. In late 2010, a team of scientists presented evidence from Mono Lake in California, a high-arsenic, low-phosphorus environment, that bacteria not only tolerated high concentrations of arsenic but used arsenic instead of phosphorus (10), possibly incorporated into the DNA backbone as arsenate. This was contested because arsenate esters have much faster hydrolysis half-lives in water 71

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

than phosphate esters. Arsenate is therefore much less suitable for maintaining DNA structure in solution, despite its chemical similarity to phosphate (11). Later experiments supported the conclusion that the bacteria from Mono Lake do not incorporate arsenate and remain dependent on phosphate (12, 13), like all known life forms. Phosphate-binding proteins from this particular bacterials strain are able to discriminate subtle structural differences between arsenate and phosphate by precise placement of hydrogen bonds (14), allowing arsenate to be rejected even in high-arsenate environments. The availability of arsenate in Mono Lake did not overcome its instability and therefore its inability to build structures persistent enough for aqueous life.

Figure 3. Oxygen levels over time, expressed as percentage of modern atmospheric levels and aligned with significant geological and biological events. Reproduced with permission from ref. (1). Copyright 2016 Oxford University Press.

The interior of a living cell is separated from the environment by a semipermeable membrane, perforated with channels and pumps that maintain the internal environment. Bacteria persist in a high-arsenate environment by evolving these proteins to maintain internal consistency despite external challenges. The prominent bioinorganic chemist R.J.P. Williams (1926-2015) proposed that the elements kept inside each cell are selected based on availability and function, and he outlined three broad functional categories for the different biochemical uses of the elements (15), which can be re-stated as the following: 72 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

1.

2.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

3.

Balance. Highly soluble elements must be pumped across the membrane to equalize osmotic pressure. Large amounts of these elements are required. Building. Elements that can form covalent bonds are used to make stable structures inside and immediately outside the cell. Moderate amounts of these elements are required. (Calcium, sulfur, and phosphate can be used for both balance and building and therefore are considered members of both categories.) Biochemistry. Elements with useful chemical properties, such as transition metals, can be used to catalyze particular reactions. Trace amounts of these elements are required.

Williams argued that cells selected phosphate rather than sulfate or arsenate for the backbone of DNA because of phosphate’s availability and chemistry (e.g., its ability to form stable phosphate esters, its flexibility, its negative charge, etc.), which are well suited for the information storage and transfer functions of DNA (15). The selection of phosphate led to the selection of magnesium to balance phosphate’s negative charge, because magnesium binds well to phosphate (e.g., the ionic radius of magnesium fits well into the spacing of oxygens in a diphosphate ester, as seen in the association of magnesium with ATP as Mg-ATP) (16). Williams theorized that the presence of charged phosphate and magnesium inside the cell required the rejection of sodium and the retention of potassium for osmotic balance in the chemical environment of the ocean (17). With this style of reasoning, Williams gave a chemical rationale for the elemental concentrations inside living cells. Broad chemical features of the environment derived from the periodic table are predictable in a way that specific biological features like gene sequences are not. The first discovery that Williams published (from research accomplished as an undergraduate) was the Irving-Williams series describing the stability of transition-metal complexes with various ligands (18). Williams later described how this series predicted the free transition-metal ion concentrations inside cells, because the inorganic chemistry of transition metal ion binding is universal despite extensive biological variability (17). Cellular pumps and channels must maintain levels of metal ions at the concentration where there is enough of the ion to bind and use, but not so much that it cross-links and congeals the fluid and dynamic structures necessary for life. Williams’s insight that large-scale chemical predictability co-exists with and shapes biological unpredictability can be seen in cases in which the environment shapes life, such as in recent large-scale “-omics” studies. Multiomic sequence information of microorganisms in the Saanich Inlet were explained with a biogeochemical model in which a chemical feature of the environment (elemental fluxes across a redox gradient) determines microbial community structure and gene expression patterns (19). In a global study of the ocean microbiome, the geochemical category of environmental conditions, including the elemental forms and concentrations that shape biological niches, was found to be more strongly determinative and predictive of the patterns of microbial life than the biological category of taxonomy (20). On a different scale, a study of microbial communities 73

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

sampled along an intercontinental cycling trip concluded that local environmental factors select from a pool of globally distributed microbial species, so that “most of the [biological] variation can be explained by environmental and not spatial patterns (21).” This may apply to how cancer responds to its environment as well. When metastatic cancer cells encounter a new environment in a different part of the human body, tumors originating from different primary sites will predictably converge to similar phenotypes, being shaped by the adaptive landscape of the local environment (22). Biological species and genes, and geographical spatial patterns, are contigent upon various degrees of unoredictability, but the chemistry of the environment plays a major role in shaping the characteristics of life in a manner consistent with the chemistry of the elements. This implies that when the chemistry of the planetary environment changed fundamentally over billions of years and as the array of elements available from the periodic table changed, that these environmental-chemical changes could have fundamentally reshaped life.

Figure 4. Concentrations of elements in the modern ocean, with elements categorized by biological functions of balance (dark gray), building (medium gray), and biochemistry (light gray). Reproduced with permission from ref. (1). Copyright 2016 Oxford University Press. 74 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Environmental Availability Restricted the Biochemical Periodic Table Differently Billions of Years Ago

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

In both ancient and modern oceans, the more soluble forms of the elements would have been more available for biological functions. This solubility depends on the presence of counterions that may form less soluble precipitates and is dependent on the concentrations of different elemental forms. Williams noted that his three broad categories of elemental function correlate with aqueous solubility in the ocean today (Figure 4) (15): 1. 2. 3. 4.

Elements available at millimolar solubilities are used for balance. Elements available at micromolar concentrations are used for building. Elements available at nanomolar concentrations are used for biochemistry and catalysis. Elements available at concentrations lower than nanomolar are not used.

Concentrations in the ocean are directly related to elemental abundance in the universe, which is why most of the elements used for biochemical processes are those with smaller atomic weights, which have been formed in more abundance from hydrogen and helium by stellar nucleosynthesis (Figure 1). The three elements present in two categories in Figure 4 each have chemical rationalizations. Calcium’s availability in the low-millimolar/high-micromolar range of concentrations and its high charge density explain its ability to provide both balancing and building functions. Sulfur is used for balance as sulfate, while it is used for building as sulfide (e.g., in proteins as disulfide bonds). Phosphate’s unique structural and energetic utility for building DNA means that it is enriched in the cell and must be accounted for in electrostatic balance as well as in biomolecular building. Several elements are redox-sensitive, and as the global redox state changed, their availability would have changed, and so would their use in life. The vertical arrows on Figure 4 show how Williams predicted that redox-sensitive elements would have responded to increasing levels of free oxygen in the atmosphere and ocean. Over billions of years of oxidation, iron, nickel, and cobalt decreased, while molybdenum and copper increased. Also, reduced forms of carbon, sulfur, and nitrogen would have been more prominent in the ancient ocean compared to the more oxidized forms present today. If these changes in concentrations were significant enough to change the availability of these elements in the ocean, their patterns of usage in life would have changed as well. Abundance and solubility influence availability and function, restricting the biochemical periodic table today to about two dozen elements. The anoxic environment three billion years ago would have resulted in a different emphasis on the table, especially in the transition metal block, with the elements on the left of the block used more and the elements on the right used less. As predicted by Williams based on solubility constants and redox potentials, older elements included iron, cobalt, nickel, and reduced forms of carbon, sulfur, and nitrogen (especially methane, sulfide, and ammonia) while newer elements included copper, molybdenum, and oxidized forms of carbon, sulfur, and nitrogen 75

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

(especially sulfate and nitrate) (23). This sequence is illustrated in Figure 5 by an arrow that places the elements in approximate order of oxidation from a reduced form to an oxidized form. The x-axis in this figure is both redox potential and time.

Figure 5. Redox potential (bottom) correlates with old and new forms of the elements used by life (top). Reproduced with permission from ref. (1). Copyright 2016 Oxford University Press. These predictions were made by Williams based on chemical properties and biochemical observations of how the metals used by archaea and ancient bacteria differ from those used by animals today. In the first decade of the twenty-first century, enough genomic data had accumulated that this chemical sequence could be more rigorously tested by multiple authors (24–26). These analyses upheld the general sequence shown in Figure 5, with more reduced forms of the elements used early in natural history and more oxidized forms used later, and with the overall patterns of metalloprotein usage in line with the previous publications by Williams.

Recent Evidence for How Oxygen Affected the Ancient Environment The chemical sequence of elements old and new throughout natural history is a continuing area of research. For example, recent studies using selenium isotopes support the general sequence of oxygen levels shown in Figure 3, including transient “whiffs” of oxygen before the Great Oxidation Event accompanied by low ocean oxygen levels even during the “overshoot” in atmospheric oxygen levels that followed that event (27–29). A more contested point is more whether increased atmospheric oxygen was a cause or effect of the widespread diversification of complex animal fossils termed the Cambrian explosion, as 76 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

described in Chapter 9 of A World from Dust (1). An interesting new technique measuring oxygen levels in halite inclusions supports high oxygen levels before the Great Oxidation Event (30), providing additional evidence that oxygen preceded biological complexity and could have been its cause. Most investigations to date have focused on oxygen levels, but recent investigations have started to determine the levels and forms of other elements. A pattern in nitrogen isotopes coincident with the Great Oxidation Event was analyzed and interpreted as showing that the rise in oxygen led to an increase in nitrate, allowing evolution of new aerobic metabolic cycles based on this element (31). This geochemical evidence supports Williams’s chemical sequence, and provides initial evidence for how oxygen shaped both the geochemistry and biochemistry of another element. One story recounted in A World from Dust that was challenged by recent evidence is from Chapter 9, in which molecular clock genetic evidence was used to suggest that newly-evolved lignin-degrading peroxidases caused a decrease of carbon burial and the end of the Carboniferous era (32). However, a later analysis of the genetic evidence concluded that, as the title of the study put it, “Delayed fungal evolution did not cause the Paleozoic peak in coal production (33).” Another study highlighted the complexity of lignin degradation genes, emphasizing that there are many genes with poorly understood functions and implying that our understanding is not complete enough to encapsulate in a simple cause-and-effect story (34). Our understanding of the interplay between chemistry and biology continues to evolve in cases like this.

The Unique Perspective of Chemistry on Natural History Throughout the narrative of A World from Dust, Williams’s ideas are contrasted with those of the naturalist Stephen Jay Gould. Gould explained the narrative of evolution in his book Wonderful Life with the metaphor of rewinding and replaying the “tape of life” to repeat natural history (35). Gould’s ultimate conclusion emphasized biological contingency: “But I suspect, from the rarity of Pikaia in the Burgess and the absence of chordates in other Lower Paleozoic Lagerstiitten, that our phylum did not rank among the great Cambrian success stories, and that chordates faced a tenuous future in Burgess times. … And so, if you wish to ask the question of the ages -- why do humans exist? -- a major part of the answer, touching those aspects of the issue that science can treat at all, must be: because Pikaia survived the Burgess decimation. … The survival of Pikaia was a contingency of ‘just history.’ I do not think that any ‘higher’ answer can be given, and I cannot imagine that any resolution could be more fascinating.” (p. 322-323) Although my perspective as a chemist causes me to de-emphasize the role of biological contingency, I agree that Gould’s “tape of life” metaphor expresses deep truths about our interconnected and complex universe. This is not a debate about 77

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

the facts but about the prioritization and interpretation of the facts. Yet Gould’s fundamental conclusion in this case is built upon one type of observation: the relative scarcity of the protochordate Pikaia in the fossil record at certain times. Simon Conway Morris, cited by Gould as the scientist who first “reached a firm conclusion” about the nature of Pikaia (35), interprets the same data differently and emphasizes the predictable importance of evolutionary convergence. Conway Morris argues that some Ediacaran body plans were lost during the Cambrian era because they were less efficient and more poorly suited to the increased competition after the Cambrian explosion (36). This is a biochemical judgment built partly on chemistry, including the ability of biological structures to utilize carbon and oxygen. If chemical inefficiency provides a reason why these body plans were removed from the array of possible biological shapes, then the same chemical inefficiency would remove those same shapes upon repetition of the tape of life. There may be similar chemical and/or environmental factors that led to the selection of Pikaia despite its unassuming profile in the fossil record. If so, then the proliferation of chordates would be more predictable than Gould concluded because of the ‘higher’ reason of chemical necessity. Evidence for chemical constraints shaping biology can be found in the records of natural history, as predicted by chemists like Williams. In the case of the Cambrian Explosion, current evidence suggests that a major geochemical event involving atmospheric oxygen increase immediately preceded a period of rapid biological innovation. Biological complexity can be connected to both the energy and chemical structures provided by oxygen, and to the chemistry of oxidized elements and elements released by oxidative weathering. If geological evidence continues to place oxygen increase before biological complexity, then the unique chemical characteristics of oxygen can connect the two in a causal chain, turning what is now correlation into something that may be considered chemical causation. Such a chain of reasoning would apply across the universe and would predict a Cambrian explosion event wherever photosynthetic water-based life can persist for billions of years, producing eventual oxygen increase. One concern with an emphasis on the predictability of natural history is that it may smuggle unwarranted teleology into scientific explanations, supervenient on the non-teleological processes of evolution. However, some scientists have proposed ways that evolutionary direction could emerge from large-scale physical constraints and laws. R.J.P. Williams is one example; another is Terrence Deacon, who theorized in Incomplete Nature about how goal-oriented behavior can develop from stochastic processes, in the context of functional and adaptive organisms (37): “… natural selection is indeed a thoroughly non-teleological process. Yet the specific organic processes which this account ignores, and on which it depends, are inextricably bound up with teleological concepts, such as adaptation, function, information, and so forth.” (p. 137) “Thus, although the evolutionary process is itself non-normative (i.e., is not intrinsically directed toward a goal), it produces organisms which are capable of making normative assessments of the information they receive.” (p. 413) 78

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

Deacon’s conclusions apply to how the random proliferation of particular genes and body plans may be constained by the chemistry of the environment to give functions, goals, and an overall direction to evolution. Chemical laws impose constraints on the available elements and the types of reactions those elements can undertake, eliminating possibilities and restricting evolution. Deacon calls these processes “absential” because they are primarily shaped by what is absent from the system rather than what is present. In living organisms, predictability can emerge from randomness through constraint in a pattern consistent with Deacon’s arguments. For example, mammalian cells transcribe genes in a messy and unpredictable process, but transport through the nuclear membrane dampens the random fluctuations and results in transcript levels in the cytoplasm that are tightly adapted to the microenvironment (38), and therefore predictable from detailed knowledge of the chemical and biological features of the microenvironment. Here, nuclear transport constrains the randomness of transcription, keeping the random fluctuations in the nucleus absent from the cytoplasm, which produces predictable results in the cytoplasm. (This is similar to how the pumps and metal-binding proteins in a cell maintain free metal ion concentrations consistent with the Irving-Williams series.) Deacon applies these concepts to explain how biological function and goal-directed behavior can emerge from evolution’s random processes, such as those involved in gene mutation and processing. Deacon’s concepts of functional emergence complicate simple identification of cause and effect. Instead of causality, he discusses goal-oriented behavior in terms of work, which is physically defined as directed movement: “Work is a more complex concept than mere cause, because it is a function of relational features. And relations, both actual and potential, are precisely where the action is. More important, work is only possible because of limitation. To abuse an old metaphor: the fabric of mind is not merely the thread that composes it.” (p.141). If directed movement and work can emerge from random processes that are limited by external constraints, then direction in evolution may emerge as well from global chemical constraints. In Deacon’s ideas, the keys to explaining how life shaped the planet and the planet shaped life can be found in relational. absential features such as feedback loops and the need for life to operate within the constraints imposed by its environment. Some have suggested that life has survived for so many billions of years on this planet because global feedback loops have allowed the planet to traverse a “Gaian bottleneck” as the sun’s luminosity changed over time (39). This interpretation sees the Earth’s surface oceans and atmosphere as an effect rather than a cause of life. Chemistry constrains all of these processes, including the Gaian bottleneck, because they take place within the chemical constraints of phase diagrams: the temperatures and pressures needed to retain surface oceans and a thick atmosphere given the energy output of the sun. To use Deacon’s terminology, this kind of planetary feedback loop is absential and directed by the requirements of dynamic persistence within a set of physical-chemical constraints. A chemical perspective that takes into account patterns of availability, solubility, and reactivity connects to both geological and biological perspectives on natural history, providing explanatory insights in a conversation with geology, 79

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

biology, and other fields of knowledge. The rise of oxygen and subsequent global chemical transformation distinguishes old elements from new over the course of natural history. The geological effects of this change are well documented, while the biological causes and effects are still being disentangled. The importance of chemistry in understanding also implies that this perspective is useful in teaching, which is why I have used these stories in physical chemistry, biochemistry, and science writing courses to help student learning. When natural history is discussed, the discussion should include not only rocks and animals, but the periodic table of elements as well.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

References 1. 2. 3. 4. 5.

6. 7. 8. 9.

10.

11. 12.

13.

McFarland, B. A World from Dust: How the Periodic Table Shaped Life; Oxford University Press: New York, 2016. Viola, V. E. Formation of the Chemical Elements and the Evolution of Our Universe. J. Chem. Educ. 1990, 67, 723. Winter, M. J. Diffusion Cartograms for the Display of Periodic Table Data. J. Chem. Educ. 2011, 88, 1507–1510. Hazen, R. M. The Story of Earth: The First 4.5 Billion Years, from Stardust to Living Planet; Penguin: New York, 2012. Ehhalt, D.; Zellner, R.; Georgii, H.-W.; Warneck, P.; Zellner, R.; Mentel, T.; Sausen, R. Global Aspects of Atmospheric Chemistry; Springer: New York, 1999. Avogadro, A.; Ragaini, R. C. Technologies for Environmental Cleanup: Soil and Groundwater; Springer: New York, 1993. Council, N. R. Frontiers in Chemical Engineering: Research Needs and Opportunities; National Academies Press: Washington, DC, 1988. Lane, N.; Martin, W. F. The Origin of Membrane Bioenergetics. Cell 2012, 151, 1406–1416. Mulkidjanian, A. Y.; Bychkov, A. Y.; Dibrova, D. V.; Galperin, M. Y.; Koonin, E. V. Origin of First Cells at Terrestrial, Anoxic Geothermal Fields. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, E821–E830. Wolfe-Simon, F.; Blum, J. S.; Kulp, T. R.; Gordon, G. W.; Hoeft, S. E.; PettRidge, J.; Stolz, J. F.; Webb, S. M.; Weber, P. K.; Davies, P. C. W.; Anbar, A. D.; Oremland, R. S. A Bacterium That Can Grow by Using Arsenic Instead of Phosphorus. Science 2011, 332, 1163–1166. Benner, S. A. Comment on “a Bacterium That Can Grow by Using Arsenic Instead of Phosphorus”. Science 2011, 332, 1149–1149. Reaves, M. L.; Sinha, S.; Rabinowitz, J. D.; Kruglyak, L.; Redfield, R. J. Absence of Detectable Arsenate in DNA from Arsenate-Grown Gfaj-1 Cells. Science 2012, 337, 470–473. Erb, T. J.; Kiefer, P.; Hattendorf, B.; Günther, D.; Vorholt, J. A. Gfaj-1 Is an Arsenate-Resistant, Phosphate-Dependent Organism. Science 2012, 337, 467–470.

80 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

14. Elias, M.; Wellner, A.; Goldin-Azulay, K.; Chabriere, E.; Vorholt, J. A.; Erb, T. J.; Tawfik, D. S. The Molecular Basis of Phosphate Discrimination in Arsenate-Rich Environments. Nature 2012, 491, 134–137. 15. Williams, R. J. P.; Da Silva, J. F. Bringing Chemistry to Life: From Matter to Man; Oxford University Press: Oxford, England, 1999. 16. Bojovschi, A.; Liu, M. S.; Sadus, R. J. Mg2+ Coordinating Dynamics in Mg:Atp Fueled Motor Proteins. J. Chem. Phys. 2014, 140, 115102. 17. Da Silva, J. F.; Williams, R. J. P. The Biological Chemistry of the Elements: The Inorganic Chemistry of Life; Oxford University Press: Oxford, England, 2001. 18. Irving, H.; Williams, R. J. P. 637. The Stability of Transition-Metal Complexes. J. Chem. Soc. 1953, 0, 3192–3210. 19. Louca, S.; Hawley, A. K.; Katsev, S.; Torres-Beltran, M.; Bhatia, M. P.; Kheirandish, S.; Michiels, C. C.; Capelle, D.; Lavik, G.; Doebeli, M.; Crowe, S. A.; Hallam, S. J. Integrating Biogeochemistry with Multiomic Sequence Information in a Model Oxygen Minimum Zone. Proc. Natl. Acad. Sci. U.S.A. 2016, 113, E5925–E5933. 20. Louca, S.; Parfrey, L. W.; Doebeli, M. Decoupling Function and Taxonomy in the Global Ocean Microbiome. Science. 2016, 353, 1272–1277. 21. Nagler, M.; Ascher, J.; Gómez-Brandón, M.; Insam, H. Soil Microbial Communities Along the Route of a Venturous Cycling Trip. Appl. Soil Ecol. 2016, 99, 13–18. 22. Cunningham, J. J.; Brown, J. S.; Vincent, T. L.; Gatenby, R. A. Divergent and Convergent Evolution in Metastases Suggest Treatment Strategies Based on Specific Metastatic Sites. Evol. Med. Pub. Health. 2015, 2015, 76–87. 23. Williams, R. J. P.; Da Silva, J. F. The Chemistry of Evolution: The Development of Our Ecosystem; Elsevier: Amsterdam, The Netherlands, 2005. 24. Dupont, C. L.; Yang, S.; Palenik, B.; Bourne, P. E. Modern Proteomes Contain Putative Imprints of Ancient Shifts in Trace Metal Geochemistry. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 17822–17827. 25. Dupont, C. L.; Butcher, A.; Valas, R. E.; Bourne, P. E.; Caetano-Anollés, G. History of Biological Metal Utilization Inferred through Phylogenomic Analysis of Protein Structures. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 10567–10572. 26. David, L. A.; Alm, E. J. Rapid Evolutionary Innovation During an Archaean Genetic Expansion. Nature 2011, 469, 93–96. 27. Pogge von Strandmann, P. A. E.; Stüeken, E. E.; Elliott, T.; Poulton, S. W.; Dehler, C. M.; Canfield, D. E.; Catling, D. C. Selenium Isotope Evidence for Progressive Oxidation of the Neoproterozoic Biosphere. Nature Comm. 2015, 6, 10157. 28. Stüeken, E. E.; Buick, R.; Anbar, A. D. Selenium Isotopes Support Free O2 in the Latest Archean. Geology 2015, 43, 259–262. 29. Kipp, M. A.; Stüeken, E. E.; Bekker, A.; Buick, R. Selenium Isotopes Record Extensive Marine Suboxia During the Great Oxidation Event. Proc. Natl. Acad. Sci. U.S.A. 2017, 114, 875–880. 81

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on November 9, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch003

30. Blamey, N. J. F.; Brand, U.; Parnell, J.; Spear, N.; Lécuyer, C.; Benison, K.; Meng, F.; Ni, P. Paradigm Shift in Determining Neoproterozoic Atmospheric Oxygen. Geology 2016, 44, 651–654. 31. Zerkle, A. L.; Poulton, S. W.; Newton, R. J.; Mettam, C.; Claire, M. W.; Bekker, A.; Junium, C. K. Onset of the Aerobic Nitrogen Cycle During the Great Oxidation Event. Nature 2017, 542, 465–467. 32. Floudas, D.; Binder, M.; Riley, R.; Barry, K.; Blanchette, R. A.; Henrissat, B.; Martínez, A. T.; Otillar, R.; Spatafora, J. W.; Yadav, J. S.; Aerts, A.; Benoit, I.; Boyd, A.; Carlson, A.; Copeland, A.; Coutinho, P. M.; de Vries, R. P.; Ferreira, P.; Findley, K.; Foster, B.; Gaskell, J.; Glotzer, D.; Górecki, P.; Heitman, J.; Hesse, C.; Hori, C.; Igarashi, K.; Jurgens, J. A.; Kallen, N.; Kersten, P.; Kohler, A.; Kües, U.; Kumar, T. K. A.; Kuo, A.; LaButti, K.; Larrondo, L. F.; Lindquist, E.; Ling, A.; Lombard, V.; Lucas, S.; Lundell, T.; Martin, R.; McLaughlin, D. J.; Morgenstern, I.; Morin, E.; Murat, C.; Nagy, L. G.; Nolan, M.; Ohm, R. A.; Patyshakuliyeva, A.; Rokas, A.; Ruiz-Dueñas, F. J.; Sabat, G.; Salamov, A.; Samejima, M.; Schmutz, J.; Slot, J. C.; St. John, F.; Stenlid, J.; Sun, H.; Sun, S.; Syed, K.; Tsang, A.; Wiebenga, A.; Young, D.; Pisabarro, A.; Eastwood, D. C.; Martin, F.; Cullen, D.; Grigoriev, I. V.; Hibbett, D. S. The Paleozoic Origin of Enzymatic Lignin Decomposition Reconstructed from 31 Fungal Genomes. Science 2012, 336, 1715–1719. 33. Nelsen, M. P.; DiMichele, W. A.; Peters, S. E.; Boyce, C. K. Delayed Fungal Evolution Did Not Cause the Paleozoic Peak in Coal Production. Proc. Natl. Acad. Sci. U.S.A. 2016, 113, 2442–2447. 34. Nagy, L. G.; Riley, R.; Bergmann, P. J.; Krizsán, K.; Martin, F. M.; Grigoriev, I. V.; Cullen, D.; Hibbett, D. S. Genetic Bases of Fungal White Rot Wood Decay Predicted by Phylogenomic Analysis of Correlated Gene-Phenotype Evolution. Mol. Biol. Evol. 2017, 34, 35–44. 35. Gould, S. J. Wonderful Life: The Burgess Shale and the Nature of History; Norton: New York, 1989. 36. Hoyal Cuthill, J. F.; Conway Morris, S. Fractal Branching Organizations of Ediacaran Rangeomorph Fronds Reveal a Lost Proterozoic Body Plan. Proc. Natl. Acad. Sci. U.S.A. 2014, 111, 13122–13126. 37. Deacon, T. W. Incomplete Nature: How Mind Emerged from Matter; Norton: New York, 2011. 38. Battich, N.; Stoeger, T.; Pelkmans, L. Control of Transcript Variability in Single Mammalian Cells. Cell 2015, 163, 1596–1610. 39. Chopra, A.; Lineweaver, C. H. The Case for a Gaian Bottleneck: The Biology of Habitability. Astrobiology 2016, 16, 7–22.

82 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.