Download Goldilocks and the three inorganic equilibria: how Earth`s chemistry

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts
no text concepts found
Transcript
Downloaded from http://rsta.royalsocietypublishing.org/ on April 29, 2017
rsta.royalsocietypublishing.org
Goldilocks and the three
inorganic equilibria: how
Earth’s chemistry and life
coevolve to be nearly in tune
Discussion
R. E. M. Rickaby
Cite this article: Rickaby REM. 2015
Goldilocks and the three inorganic equilibria:
how Earth’s chemistry and life coevolve to be
nearly in tune. Phil. Trans. R. Soc. A 373:
20140188.
http://dx.doi.org/10.1098/rsta.2014.0188
Department of Earth Sciences, University of Oxford,
South Parks Road, Oxford OX1 3AN, UK
One contribution of 18 to a discussion meeting
issue ‘The new chemistry of the elements’.
Subject Areas:
geochemistry, inorganic chemistry
Keywords:
coevolution, essential nutrients, trace metals,
inorganic equilibria, periodic table, toxicity
Author for correspondence:
R. E. M. Rickaby
e-mail: [email protected]
Life and the chemical environment are united in
an inescapable feedback cycle. The periodic table
of the elements essential for life has transformed
over Earth’s history, but, as today, evolved in tune
with the elements available in abundance in the
environment. The most revolutionary time in life’s
history was the advent and proliferation of oxygenic
photosynthesis which forced the environment
towards a greater degree of oxidation. Consideration
of three inorganic chemical equilibria throughout this
gradual oxygenation prescribes a phased release of
trace metals to the environment, which appear to have
coevolved with employment of these new chemicals
by life. Evolution towards complexity was chemically
constrained, and changes in availability of notably
Fe, Zn and Cu paced the systematic development of
complex organisms. Evolving life repeatedly catalysed
its own chemical challenges via the unwitting release
of new and initially toxic chemicals. Ultimately, the
harnessing of these allowed life to advance to greater
complexity, though the mechanism responsible for
translating novel chemistry to heritable use remains
elusive. Whether a chemical acts as a poison or a
nutrient lies both in the dose and in its environmental
history.
1. The periodic table of life
Goldilocks happens upon the house of the three bears in
the woods, and discovers among the too hot or too cold
porridge, a bowl of porridge that is just right; among the
three chairs that are too big or too small, one that is just
right; and among the beds that are too hard or too soft,
one that is just right. Like this tale, after 4.5 Byr of Earth’s
2015 The Author(s) Published by the Royal Society. All rights reserved.
Downloaded from http://rsta.royalsocietypublishing.org/ on April 29, 2017
2
14.7
C
Mg
B
13.7
Li
10.4K 13
Ca
14.7
12.4
11.6
O
N
Si
13
Sr
Al
10.9
V
4.81
7.3
Mn
CrMo 8.0
6.7
6.6
Cs
6.34
Ba
8
Sc Y Ti 4.18
Zr
4
5.4
5.1
Hf
Lu
Ni
Fe
6
6.9 Cu 6.6
Zn
10.5
Co
Cd
4.7
Nb
4.6
W
4.7
Ta
4.1
Ag
5.7
Ru
Re
Tc 4.3
3.6
Rh
Ar
Pt
2.9
Os
bulk biological
elements
1.5
10.2
Ne
7.5
Ga
5.4
5.8
Ge
4.8
Br
7.3
Sb
4.4
3
Se
3.7
8.6Kr
6.5
Xe
6.0
Bi
6.8
I7.9
6.2
Te
Sn
3.6
InTl 4.7 Pb
Au
3
4.4 Hg
2.6 Pd 3.3
2.4
10.8
9.3
As
6.7
13.4
11.3 F
11
P
Rb
Be9.15
S
5.7
4.7
5
4
Ir 1.7
trace elements believed
to be essential for bacteria,
plants or animals
possibly essential trace
elements for some
species
Figure 1. A periodic table constructed to demonstrate the similarity between the abundant elements within seawater and
the elements essential for life. The periodic table of the chemistry of seawater with concentrations indicated by the height of
the bar of each element in log10 [M] fmol kg−1 with the concentrations also marked under the element symbol; concentrations
were derived from the compilation of literature courtesy of http://www.marscigrp.org/ocpertbl.html. Columns are coloured
according to the bulk biological elements considered essential to life (yellow), trace elements considered essential for life (green)
and trace elements thought of as possibly being essential for some species (red).
history, to a first order, the elements that are required for life appear to have evolved to be in
tune with the elements that are available in abundance to life within the ocean (figure 1). But how
has this general matching of abundance and need arisen? Why, of the smorgasbord of elements
within the periodic table, have only 25 of the 117 been demonstrated to be essential to all life with
a further seven being useful to only some species (figure 1)? At least two factors are at play. First
is the unique chemical nature of different elements that perform key parts of biological reactions,
which has enabled the evolution of complex and sentient multicellular life. Second, and apparent
from the broad correlation between concentration in the environment and biological requirement
(figure 1), is the role of evolving availability, dictated by three inorganic equilibria: solubility,
complexation and redox chemistry [1–3]. Metal availability for life is denoted as the concentration
or activity of the free ion, acknowledging that complexation by key ligands can enhance metal
availability in some forms for life today. With this perspective and delving into deep time, the key
conclusion that can be drawn is that the elements required for life must have evolved over time
in a movable feast. A periodic table of elements essential for life 2 billion years ago (Ba) would
have looked very different to that of the modern day.
2. Geological controls of the chemistry of the ocean where life evolved
The abundance of elements within bulk Earth was determined by the Big Bang and subsequent
supernovae processes. These elements were the starting ingredients, but Earth’s surface is
chemically distinct from bulk Earth and the interior, owing to melting and differentiation. This
.........................................................
He
Cl 12.3
Na
rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140188
log10[M] fmol kg–1
H 16.7
Downloaded from http://rsta.royalsocietypublishing.org/ on April 29, 2017
It is the remarkable similarity between the abundance of the major elements in life and those in
the sea that points to a coevolution between these two chemistries (figure 1). Na and Cl are by
far the most abundant, followed by Mg, Ca, K and HCO3 − . With such long residence times in the
ocean, and concentrations being dominantly controlled by solubility and buffering, these major
ions have maintained a near-constant relatively high concentration on million year time scales
and for much of Earth’s history. Although there is an intimate relationship between the degree of
employment for life and complexity of life, whereby the number of proteins using for example
Ca increases exponentially as life advances from single bacterial cells to compartmentalized
eukaryotic cells, and ultimately to multicellular forms [4], this is unlikely driven by changes
in the environment. In fact, as we shall explore below, the increased use of Ca occurs
contemporaneously with the increased use of both Cu and Zn and, we shall argue, is enabled as
a result of the advent of those trace metals into the environment triggering enhanced complexity
of life.
By contrast to the major ions that, today and always, have the highest concentration in
the oceans and life, the transition (and trace) metals have the potential for an extraordinarily
flexible redox state, owing to their partially filled d-orbitals in the ground state. This flexibility
has two important implications for life. First, the transition metals are unique in providing life
with unparalleled opportunities for performing electron transfer, redox and acid–base chemistry.
Second, their availability is highly sensitive to the average redox state of Earth’s surface.
The major oxidation/reduction chemical changes of the surface of Earth arose from the uptake
of carbon dioxide and nitrogen in combination with hydrogen either as H2 from hydrogen sulfide,
or water, into organic compounds, i.e. reduction (in cells), and release of sulfur or oxygen. O2
release (as waste) led to the oxidation of surface minerals and ions and molecules in solution.
All this activity was and is driven by light aided by essential catalysts largely based on bound
inorganic ions.
hv
H2 O + CO2 (+elements) −−−−−→ reduced organic compounds + O2 .
catalysts
These changes, small at first, gained systematically as the rate of absorption of light by chemicals,
the pool of reduced forms of carbon as a product of life and the inherent release of oxygen all
increased.
.........................................................
3. Changing major and trace element concentrations
3
rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140188
process segregated Earth into the continental and oceanic crust, mantle and core, and partitioned
the more labile elements according to their siderophile or lithophile nature.
Water is the only chemical reagent that redistributes elements between the rock reservoir
and Earth surface water via solubility. At low temperatures, rivers leaching rocks and soils
(accelerated by plants since their Carboniferous land invasion) form the primary conduit for
transport of elements from the continental surface to the ocean after chemical weathering.
At high temperatures, seawater circulates through newly forming and young ocean crust
exchanging elements via leaching and clay formation. Additional loss of elements from the
ocean occurs through chemically and biologically mediated precipitates to form ocean sediments
that ultimately become subducted or lifted back onto continents. The relative balance of the
inputs and outputs of elements to the ocean through time sets the evolving chemical nature of
seawater.
The major ions in seawater (Na, Mg, Ca, K, Cl, SO4 and HCO3 − ) are those that are most soluble
from rocks but, importantly, do not have a major sink or output. For example, Na, Cl and K are
only removed via the infrequent occurrence of intense evaporite mineral formation that lends
these elements a long ‘residence time’ of tens of millions of years and makes them relatively
invariant for the majority of Earth’s history.
Downloaded from http://rsta.royalsocietypublishing.org/ on April 29, 2017
4. Three chemical equilibria controlling metal availability in the ocean
The solubility at equilibrium is described by equilibrium solubility products KS = [Mn+ ][An− ],
where A is an anion. The order of solubility products of salts of the abundant metal ions, M2+ ,
is that Ca2+ > Mg2+ for carbonates and phosphates but Mg2+ > Ca2+ for silicates while the
order is Ba2+ > Sr2+ > Ca2+ > Mg2+ for sulfate insolubility. These metals do not form sulfides.
Such orders and the abundances of the elements have meant that free Mg2+ ions were more
concentrated in the sea than Ca. The general order of insolubility of salts in the series of divalent
d-block ions, M2+ , is
Cu2+ > Ni2+ , Zn2+ > Co2+ > Fe2+ > Mn2+ .
This order holds for sulfides and oxides. A major mineral in Earth’s mantle is olivine,
Fe2+ Mg2+ SiO4 , which formed through the abundance of Mg and Fe. It is thought that the early
sea had high Mg2+ and Fe2+ concentrations, as olivine is relatively soluble. The high Fe2+
reduced the amount of sulfide somewhat as the iron precipitated it as pyrite though it is also
not extremely insoluble. Even so the residual sulfide greatly restricted the free ion concentrations
of Cu (probably Cu+ which has a very insoluble sulfide), Zn2+ and some heavier metal ions on
early Earth.
(b) Equilibrium 2: complexation
The hydrated free ion concentrations, availability in solution, are related to their combinations
with ligands in the sea according to the equilibrium K = [Mn+ ][An− ]/[MA]. In the earliest times,
the restriction of [M] in the sea would have been primarily due to the presence of hydroxide,
carbonate, silicate and sulfide but later by oxyanions of stronger acids, for example, sulfate. The
affinity of complex formation for metals closely parallels the above trends in solubility of metal
salts. Hydroxide and oxide greatly reduce the free ion concentrations of cations with a charge
of more than two. Thus at pH = 7, M3+ such as Al3+ could be held by OH− in complexes or
precipitates. In the presence of silica Al3+ also forms large soluble aluminosilicates. Decreasing
pH increases free Al3+ concentration. The only divalent ion that may have been restricted by
complex formation with aluminosilicate is nickel. The later metal ions of the series Mn2+ , Fe2+ ,
Co2+ , Ni2+ , Cu2+ , Zn2+ formed sulfide complexes of increasing strength, 1/K, in this order but
with Cu2+ > Zn2+ , the Irving–Williams series [6]. The ions Mg2+ and Ca2+ form few complexes
in the sea, while Na+ and K+ form hardly any complexes. Anions can also bind to one another,
but rarely, or to organic surfaces.
(c) Equilibrium 3: redox
The metal ion oxidation changes are relatively fast and systematic due to the equilibrium
thermodynamics of oxidation/reduction even if full equilibrium is not quite achieved. Redox
potential, E, for the equilibrium between two redox states say M+ and M2+ is written as
E = Eo +
RT [M2+ ]
ln
,
nF
[M+ ]
where T is the temperature, F is the Faraday constant and [ ] indicates concentration. Eo , the
standard redox potential for [H+ ]/[H2 ] at pH = 0 is 0.0 V and is −0.42 V at pH = 7.0. The range of
.........................................................
(a) Equilibrium 1: solubility
rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140188
The increasing redox potential of the environment affects elemental availability via the simple
chemistry of three equilibria [3,5]. The general restrictions on the availability of elements as free
cations, M+ and M2+ in the sea are insolubility and complex ion formation with redox chemistry
controlling the redox state of the ions.
4
Downloaded from http://rsta.royalsocietypublishing.org/ on April 29, 2017
Although the redox history of the ocean and atmosphere may be more dynamic than originally
thought [7], a consensus is emerging of the timing of the major oxidative switches (figure 2).
The most powerful evidence for the great oxidation event (GOE), the first major transition from
initially reducing to oxidized conditions at the planet’s surface, is the cessation of the massindependent fractionation of δ 33 S, δ 34 S and δ 36 S between sulfides and sulfates, dated to between
2090 and 2450 Ma [8] (figure 2). Gas-phase chemical reactions prior to 2450 Ma are likely to lend
the mass-independence to the sulfur isotopic fractionation. This mass-independent signature
was diluted or eliminated subsequently in the presence of more abundant oxygen. Possible
mechanisms to homogenize the mass-independent isotope signatures of the reduced and oxidized
forms include oxidative weathering, or suppression of the atmospheric reactions of sulfur under
oxidative conditions.
Molybdenum has become an element of increased interest for providing a window into the
past redox conditions of the oceans. Under oxic conditions, this relatively abundant element
exists dominantly as the molybdate ion (MoO4 2− ) [10] but under euxinic conditions with free
hydrogen sulfide, Mo is almost entirely scavenged from the water column as the particlereactive oxythiomolybdate ion (MoOx S4−x 2− ) [11] which then captures whole ocean Mo isotopic
compositions, sensitive to the proportion of oxic and reducing sediments. The Mo content,
reactive iron content and the Mo isotopic signatures of shales all point towards at least one
or two further steps in oxygenation significantly at the Precambrian/Cambrian boundary
(543 Ma), coincident with the rise of animal life [9,12] (figure 2) and potentially entering into the
Carboniferous period some 300 Ma [13]. Crucially despite the GOE approximately 2.2 Ba, only
the surface ocean is thought to have become oxic at that time, and the subsequent approximately
2 Byr until 543 Ma are thought to have maintained a euxinic and sulfidic deep ocean, limiting the
trace metal solubility [14].
From this broad oxidation history of the Earth’s surface (figure 2), it becomes apparent that the
oxidation state and complexity of life have advanced hand in hand. The major jump in oxygen
approximately 543 Ma accompanied the ‘Cambrian explosion’ of life. This explosion in diversity
in the skeletal fossil record represents the transition from single celled eukaryotes to differentiated
cells, complex multicellularity and to biomineralization.
6. Evidence for evolving transition metal availability
Within the framework of the advancing oxidation potential of the Earth’s surface, the three
chemical equilibria prescribe the evolving transition metal availability: declining Fe, Mn and Co
and increasing Cu, Zn and Cd across each step of oxygen rise. The geological record provides one
approach to testing this chemical framework. Chemical sediments from Proterozoic time may
capture the chemistry of palaeoseawater. Potential archives include the banded iron formations
(e.g. [15]), pyrite (in framboidal form) [16] and shales [9,17]. Banded iron formations consist
of alternating magnetite (haematite) and chert. Framboidal pyrite (FeS2 ), an early diagenetic
precipitate, is thought to form quickly after deposition and is therefore best poised to capture
seawater as opposed to porewater chemistry [16]. Shales can be thought of as fine-grained
detritus, which may adsorb bulk seawater chemical properties as the particles fall through the
.........................................................
5. Probing the past oxidation state of the planet
5
rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140188
possible oxidation/reduction potentials is constrained throughout the evolving slow switch from
H2 to O2 limiting conditions, where the O2 /H2 O standard potential at pH = 7.0 is +0.8 V.
These three equilibria then prescribe the inevitable direction of change of element availability
that occurs in a given order in the environment, according to the reconstructed degree of
oxidation. It is proposed that this provides a control of cell chemistry too in evolution. Some
elements were directly affected by change of oxidation conditions of the sea as described while
other elements were not affected by the presence of hydrogen or oxygen, for example, Na+ ,
K+ , F− , Cl− .
Downloaded from http://rsta.royalsocietypublishing.org/ on April 29, 2017
1st step in O2
2nd step in O2
2.5
6
0.5
0
300
–0.5
200
–1.0
100
–1.5
0
1000
2000
3000
Mo (ppm)
D33S
1.0
0
4000
age (Ma)
Figure 2. Adapted figure to compile datasets used to support the two major steps to the evolving oxidation of Earth including
(i) the disappearance of the signature of mass-independent fractionation approximately 2 Ba of sulfur (33 S) (open squares on
y1-axis) [8] and (ii) the increasing Mo content of euxinic shales (black circles on the y2-axis) [9].
water column. For each archive however, the chemical contents of the deposit itself cannot be
translated directly to seawater concentrations. For the deposit to act as a faithful recorder of
evolving metal availability, the metal must partition into the sediment independently of the
oxidation state of the ocean, and remain unchanged by post-depositional diagenesis. To date,
the picture of evolving Zn, for example, availability from various archives is anything but clear
[16–18] although there are hints of an increase in seawater Zn concentration towards the modern
day [16]. Emerging work suggests that a more promising archive may be within the chemistry of
greenalite (an abundant authigenic Fe2+ -silicate in Precambrian iron formation, ironstones and
shales). Long recognized as an abundant component of Precambrian lithologies but regarded
as enigmatic in origin, greenalite occurs throughout the depositional spectrum and is one of
the few phases preserved in Precambrian sediments that directly archives multiple trace metal
inventories [19].
Intriguingly, the chemistry and genes of biota seem to provide better recorders than the
geological record for the chemical framework of the altering trace metal availability as oxygen
evolved and for the order of availability. The chemical make-up of extant phytoplankton which
combine the trace metal requirement of the host phagoautotroph with the endosymbiont hints
at the later increased availability of Cu, Zn and Cd [20,21]. The more recently evolved red algal
superfamily has lower requirements for Fe and Mn, with higher requirements for Cd, Co and Mo
than the green algal superfamily. But more than the metallome, an intimate look at comparative
genomics and protein structures demonstrates that the prevalence of fold superfamilies specific
metals within protein motifs [4] reiterates a phased employment of transition metals within
organisms. Fe, Mn and Mo are preferentially selected by early life before Cu and Zn (and Ca,
see above) increased their utilization only after the GOE. Furthermore, bioinformatic approaches
to the characterization of the zinc and copper proteomes from all organisms with a full genetic
sequence show distinct correlations between the number of zinc- and copper-containing proteins
encoded within the genomes, and the complexity of the organism. Of note is the step change
in use of zinc-finger containing proteins and zinc hydrolytic enzymes, the copper chaperones,
homeostatic proteins and redox enzymes within eukaryotic and multicellular organisms [22,23].
There appears a clear trend of novel and multiple copies of the genetic code for these metal
proteins as the metal availability rises in the environment due to oxidation. It is thought that the
availability of these key d-block metals acted as a handbrake on the evolution of life (e.g. [1,5,14]),
.........................................................
1.5
rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140188
2.0
Downloaded from http://rsta.royalsocietypublishing.org/ on April 29, 2017
The early establishment and evolution of life on the Earth, then, revolves around a cycle of (1)
invention, (2) unwitting generation of initially toxic waste (e.g. O2 , Cu and Zn), but which via
the (3) evolution of management systems of those toxins towards (4) biological use, ultimately
forces or yields invention of novel life chemistry and so on. Even Cd, a chemical latecomer to the
ocean in the redox sequence, and commonly thought of as toxic to life due to its striking chemical
similarity to Ca and ability to bind irreversibly to Ca enzymes, may have become of use to life.
Vertical profiles of dissolved Cd show a near linear relationship with dissolved phosphate (e.g.
[24]), an essential and limiting nutrient in the ocean, with the implication that requirement for life
must also extend to Cd. The first enzymatic use of Cd has recently been proposed within a novel
form of carbonic anhydrase (CDCA). Despite the inference that the earliest enzymes were likely
cambialistic and evolved towards specificity [25], the relative scarcity of CDCA, so far found
only with the diatom class of algae [26], speaks to the geologically recent novelty in the use of
trace metals for life. Whether the uptake of Cd into CDCA accounts quantitatively for the global
correlation between Cd and phosphate remains contentious [27,28]. A question for future research
will be to disentangle from the widely explored biological quota of metals for phytoplankton [29],
and inference of nutrient-like metals from seawater profiles, the entwined absolute requirement
of metals for life from the biological sequestration of mistakenly imported metals.
8. The evolving toxic threshold and dose–response curves
From consideration of the evolution of the new chemistry of the elements, it transpires that
it is not only the concentration of an element which dictates its degree of toxicity (Paracelsus
‘[a]ll things are poison and nothing is without poison, only the dose permits something not to
be poisonous’) but also its availability within evolutionary history. The evolutionary cycle of
invention and unwitting poisoning of life predicts a systematic temporal evolution to the dose–
response curve of increasingly complex life as the concentration of a new chemical increased in
the environment, e.g. O2 , Cu and Zn (figure 3). Across each step in elevated concentration of
newly introduced elements in the environment, increasingly complex life would have gained an
increase in the toxic threshold, as the initial metal toxicity management systems (recognition,
binding and accumulation and/or expulsion) transitioned towards biological use. Furthermore,
once biological use of a novel metal is established then any response curve must incur some
degree of limitation if that metal confers an inducible physiological advantage, or limitation if
that metal becomes employed constitutively for a core physiological use. In this case, invention of
a high-affinity uptake system may be additionally triggered to fulfil this biological requirement
at low metal concentrations (figure 3).
Since the bioinformatic approaches, detailed above, yield only patterns and phasing of trace
metal evolution through time, interrogation of both ends of the dose–response curves of extant
organisms with diverse emergences may yield additional insight into absolute concentrations of
metals at different times. If these response curves of organisms are reflective of the environment
experienced since emergence, as suggested for coccolithophores in response to carbon [30], then
probing a variety of extant organisms of diverse complexity and times of emergence for their
toxic thresholds to, for example, Cu or Zn chemistries may reflect the highest concentrations of
metals attained in the environment. Similarly, the specificity of the limiting arm of the response
curve for different organisms might also approximate the environmental concentrations at the
time the metal transitioned from being managed to becoming essential to life, since cambialism
characterizes earliest metabolic processes [25]. Both low-affinity and inducible high-affinity Cu
and Zn transporters have been demonstrated (with the switch to high affinity for Zn2+ occurring
.........................................................
7. The new chemistry of the elements for life
7
rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140188
only permitting the functions necessary for eukaryotic cells and multicellularity to arise as and
when the metals such as Zn and Cu, best suited to the metallo-centres of enzymes responsible for,
e.g. cellular communication, splitting and cohesion between cells, were sufficiently available.
emerging
tolerance 2
toxic
1
concentration
Figure 3. A hypothetical schematic to show the evolution of the dose–response curve for evolving organisms as a novel
chemical/toxin (e.g. O2 , Cu, Zn) is introduced stepwise into the environment through steps 1–5. In the first stage, for the
earliest organisms approximately 3.5 Ba, that metal is always toxic and any increase in concentration poisons growth (1).
But upon prolonged exposure to increasing metal availability, and evolution of metal homeostasis, growth can plateau with
increasing concentration as the metals can be internally managed before reaching the toxic threshold (2). Further advancement
of the metal management system evolving towards inducible biological use boosts growth to higher levels and pushes the
toxic threshold towards higher concentrations while starting to show decreased success at very low concentrations (3). Upon
employment of the new metal towards core biological use, low availability of that metal must then yield limitation of the growth
and the full dose–response curve (4). There may be a final step of evolving higher affinity uptake mechanisms to fulfil the core
physiological requirements (5).
between 10−10.5 and 10−9.5 M, and for Cu+ between 500 and 2000 nM) in marine phytoplankton
[31,32], with induction linked to lower availability but essential need of those metals.
Lastly, such dose–response curves which display some degree of limitation may allow
exploration of the farther reaches of the periodic table for additional essential elements for life
and provide the first indicator that life could be finding novel use for ‘pollutants’ introduced by
human modification of environmental chemistry during this anthropogenic era.
9. Concluding remarks
Subsequent to the introduction of oxygen as a waste product into the environment as an inevitable
consequence of the emergence of photosynthesis, the periodic table of the elements essential to life
evolved systematically according to element availability in the environment, directly controlled
by the three inorganic equilibria. These equilibria enabled the advancement of life towards
complexity via phased release of key elements into the environment, initially toxic to life but
ultimately employed for use. The mechanism by which the systematic inorganic equilibria
chemical changes of the environment interact with the chance alterations of the genomes of
organisms to result in heritable novel life chemistry and the Goldilocks-like tuning of life’s
chemistry to that of the environment remains elusive.
Acknowledgements. I would like to extend enormous thanks to Bob Williams for his endless inspiration, mindexpanding discussions and undinted support and guidance; and Dave Sansom for invaluable help in
preparing the figures.
Funding statement. I would like to thank the Philip Leverhulme Award and the ERC Starting Grant ERC SP2GA-2008-200915 for financial support.
8
.........................................................
5
homeostasis
and first use
(early organisms)
3
homeostasis
and core physiological
use yields limitation
(complex organisms)
4
rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140188
growth/success
Downloaded from http://rsta.royalsocietypublishing.org/ on April 29, 2017
Downloaded from http://rsta.royalsocietypublishing.org/ on April 29, 2017
References
.........................................................
rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140188
1. de Silva FJJR, Williams RJP. 1991 Biological chemistry of the elements. Oxford, UK: Oxford
University Press.
2. Williams RJP, de Silva FJJR. 2006 The chemistry of evolution: the development of our ecosystem.
Amsterdam, The Netherlands: Elsevier.
3. Williams RJP, Rickaby REM. 2012 Evolution’s destiny. Cambridge, UK: RSC Publishing.
(doi:10.1039/9781849735599)
4. Dupont CL, Butcher A, Valas RE, Bourne PE, Caetano-Anolles G. 2010 History of biological
metal utilization inferred through phylogenomic analysis of protein structures. Proc. Natl
Acad. Sci. USA 107, 10 567–10 572. (doi:10.1073/pnas.0912491107)
5. Saito MA, Sigman DM, Morel FMM. 2003 The bioinorganic chemistry of the ancient ocean:
the evolution of cyanobacterial metal requirements and the biochemical cycles of the archaen
proterozoic boundary. Inorg. Chim. Acta 356, 308–318. (doi:10.1016/S0020-1693(03)00442-0)
6. Irving H, Williams RJP. 1948 Order of stability of metal complexes. Nature 162, 746–747.
(doi:10.1038/162746a0)
7. Lyons TJ, Reinhard CT, Planavsky NJ. 2014 The rise of oxygen in Earth’s early ocean and
atmosphere. Nature 506, 307–315. (doi:10.1038/nature13068)
8. Farquhar J, Bao H, Thiemens M. 2000 Atmospheric influence of Earth’s earliest sulfur cycle.
Science 289, 756–758. (doi:10.1126/science.289.5480.756)
9. Scott C, Lyons TW, Bekker A, Shen Y, Poulton SW, Chu X, Anbar AD. 2008 Tracing the stepwise
oxygenation of the proterozoic biosphere. Nature 452, 456–459. (doi:10.1038/nature06811)
10. Morford JL, Emerson S. 1999 The geochemistry of redox sensitive trace metals in sediments.
Geochim. Cosmochim. Acta 63, 1735–1750. (doi:10.1016/S0016-7037(99)00126-X)
11. Erickson BE, Helz GR. 2000 Molybdenum(VI) speciation in sulfidic waters: stability
and lability of thiomolybdates. Geochim. Cosmochim. Acta 64, 1149–1158. (doi:10.1016/
S0016-7037(99)00423-8)
12. Canfield DE, Poulton SW, Narbonne GM. 2007 Late-Neoproterozoic deep-ocean oxygenation
and the rise of animal life. Science 315, 92–95. (doi:10.1126/science.1135013)
13. Dahl TW et al. 2010 Devonian rise in atmospheric oxygen correlated to the radiations of
terrestrial plants and large predatory fish. Proc. Natl Acad. Sci. USA 107, 17 911–17 915.
(doi:10.1073/pnas.1011287107)
14. Anbar AD, Knoll AH. 2002 Proterozoic ocean chemistry and evolution: a bioinorganic bridge?
Science 297, 1137–1142. (doi:10.1126/science.1069651)
15. Klein C, Beukes NJ. 1993 The sedimentology and geochemistry of the glaciogenic
late protozoic rapstan iron formation in Canada. Econ. Geol. 88, 542–565. (doi:10.2113/
gsecongeo.88.3.542)
16. Large RR et al. 2014 Trace element content of sedimentary pyrite as a new proxy for
deep-time ocean–atmosphere evolution. Earth Planet. Sci. Lett. 389, 209–220. (doi:10.1016/
j.epsl.2013.12.020)
17. Scott C et al. 2013 Bioavailability of zinc in marine systems through time. Nat. Geosci. 6,
125–128. (doi:10.1038/ngeo1679)
18. Robbins LJ et al. 2013 Authigenic iron oxide proxies for marine zinc over geological
time and implications for eukaryotic metallome evolution. Geobiology 11, 295–306.
(doi:10.1111/gbi.12036)
19. Tosca NJ, Guggenheim S, Pufahl PK. In press. The origin of greenalite in Precambrian
sediments: implications for Archaean and Palaeoproterozoic seawater chemistry. GSA
Bulletin.
20. Quigg A, Finkel ZV, Irwin AJ, Rosenthal Y, Ho T-Y, Reinfelder JR, Schofield O, Morel
FMM, Falkowski PG. 2003 The evolutionary inheritance of elemental stoichiometry in marine
phytoplankton. Nature 425, 291–294. (doi:10.1038/nature01953)
21. Quigg A, Irwin AJ, Finkel ZV. 2011 Evolutionary inheritance of elemental stoichiometry in
phytoplankton. Proc. R. Soc. B 278, 526–534. (doi:10.1098/rspb.2010.1356)
22. DeCaria L, Bertini I, Williams RJP. 2010 Zinc proteomes, phylogenetics and evolution.
Metallomics 2, 706–709. (doi:10.1039/C0MT00024H)
23. DeCaria L, Bertini I, Williams RJP. 2011 Copper proteomes, phylogenetics and evolution.
Metallomics 3, 56–60. (doi:10.1039/c0mt00045k)
24. Elderfield H, Rickaby REM. 2000 Oceanic Cd/P ratio and nutrient utilisation in the glacial
Southern Ocean. Nature 405, 305–310. (doi:10.1038/35012507)
9
Downloaded from http://rsta.royalsocietypublishing.org/ on April 29, 2017
10
.........................................................
rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140188
25. Caetano-Anollés G, Yafremava LS, Gee H, Caetano-Anollés D, Kim HS, Mittenthal JE. 2009
The origin and evolution of modern metabolism. Int. J. Biochem. Cell Biol. 41, 285–297.
(doi:10.1016/j.biocel.2008.08.022)
26. Xu Y, Feng L, Jeffrey PD, Shi Y, Morel FMM. 2008 Structure and metal exchange in
the cadmium carbonic anhydrase of marine diatoms. Nature 452, 56–61. (doi:10.1038/
nature06636)
27. Horner TJ, Lee RBY, Henderson GM, Rickaby REM. 2013 Nonspecific uptake and homeostasis
drive the oceanic cadmium cycle. Proc. Natl Acad. Sci. USA 110, 2500–2505. (doi:10.1073/
pnas.1213857110)
28. Morel FMM. 2013 The oceanic cadmium cycle: biological mistake or utilization? Proc. Natl
Acad. Sci. USA 110, E1877. (doi:10.1073/pnas.1304746110)
29. Twining BS, Baines SB. 2013 The trace metal composition of marine phytoplankton. Annu. Rev.
Mar. Sci. 5, 191–215. (doi:10.1146/annurev-marine-121211-172322)
30. Henderiks J, Rickaby REM. 2007 A coccolithophore concept for constraining the Cenozoic
carbon cycle. Biogeosciences 4, 323–329. (doi:10.5194/bg-4-323-2007)
31. Sunda WG, Huntsman SA. 1992 Feedback interactions between zinc and phytoplankton in
seawater. Limnol. Oceanogr. 37, 25–40. (doi:10.4319/lo.1992.37.1.0025)
32. Guo J, Annett AL, Taylor RL, Lapi S, Ruth TJ, Maldonado MT. 2010 Copper-uptake kinetics of
coastal and oceanic diatoms. J. Phycol. 46, 1218–1228. (doi:10.1111/j.1529-8817.2010.00911.x)