Nothing Special   »   [go: up one dir, main page]

10 1016@j Earscirev 2020 103150

Download as pdf or txt
Download as pdf or txt
You are on page 1of 96

Journal Pre-proof

Love is in the Earth: A review of tellurium (bio)geochemistry in


Earth surface environments

O.P. Missen, R. Ram, S.J. Mills, B. Etschmann, F. Reith, J.


Shuster, D.J. Smith, J. Brugger

PII: S0012-8252(19)30699-3
DOI: https://doi.org/10.1016/j.earscirev.2020.103150
Reference: EARTH 103150

To appear in: Earth-Science Reviews

Received date: 24 October 2019


Revised date: 21 February 2020
Accepted date: 7 March 2020

Please cite this article as: O.P. Missen, R. Ram, S.J. Mills, et al., Love is in the Earth:
A review of tellurium (bio)geochemistry in Earth surface environments, Earth-Science
Reviews(2019), https://doi.org/10.1016/j.earscirev.2020.103150

This is a PDF file of an article that has undergone enhancements after acceptance, such
as the addition of a cover page and metadata, and formatting for readability, but it is
not yet the definitive version of record. This version will undergo additional copyediting,
typesetting and review before it is published in its final form, but we are providing this
version to give early visibility of the article. Please note that, during the production
process, errors may be discovered which could affect the content, and all legal disclaimers
that apply to the journal pertain.

© 2019 Published by Elsevier.


Journal Pre-proof
REVISION 2.1

Love is in the Earth: a review of tellurium (bio)geochemistry in Earth

surface environments

O.P. Missen a,b,* , R. Ram a, S.J. Mills b, B. Etschmann a, F. Reith c,d† , J. Shuster c,d, D.J.

Smith e, J. Brugger a,*

f
oo
a
School of Earth, Atmosphere and Environment, 9 Rainforest Walk, Monash University,

Clayton 3800, Victoria, Australia


b pr
Geosciences, Museums Victoria, GPO Box 666, Melbourne 3001, Victoria, Australia
e-
c
School of Biological Sciences, The University of Adelaide, Adelaide, SA 5005, Australia
d
Pr

CSIRO Land and Water, Contaminant Chemistry and Ecotoxicology, PMB2, Glen Osmond,

SA 5064, Australia
al

e
School of Geography, Geology & the Environment, University of Leicester, Leicester, UK
u rn
Jo

*Corresponding authors’ e-mails: omissen@museum.vic.gov.au, joel.brugger@monash.edu



Passed away prior to completion of the manuscript

Keywords: Tellurium cycling, environments, (bio)geochemistry, mobility, mineralogy


Journal Pre-proof
ABSTRACT

Tellurium (Te) is a rare metalloid in the chalcogen group of the Periodic Table. Tellurium is

regularly listed as a critical raw material both due to its increased use in the solar industry,

and to the dependence on other commodities in its supply chain. A thorough understanding of

the geo(bio)chemistry of Te in surface environments is fundamental for supporting the search

for future sources of Te (geochemical exploration); developing innovative processing

techniques for extracting Te; and quantifying the environmental risks associated with rapidly

increasing anthropogenic uses. The present work links existing research in inorganic Te

f
oo
geochemistry and mineralogy with the bio(geo)chemical and biological literature towards

developing an integrated Te cycling model.


pr
Although average crustal rocks contain only a few µg/kg of Te, hydrothermal fluids and
e-
vapours are able to enrich Te to levels in excess of mg/kg. Tellurium is currently recovered as
Pr

a by-product of base- metal mining; in these deposits, it occurs mainly in common sulphides

substituting for sulphur. Extreme Te enrichment (up to wt.%) is found in association with the
al

precious metals Au and Ag in the form of telluride and sulphosalt minerals. Tellurium also
rn

forms a large variety of oxygen-containing secondary minerals as a result of weathering of


u

Te-containing ores in (near-)surface environments. Anthropogenic activities introduce


Jo

significant amounts of Te into surficial environments, both through processing materials that

contain minor Te, and through breakdown of used Te-containing materials. Additionally,
132
radioactive Te is produced in nuclear reactors, and can contaminate surrounding and distal

environments.

Environmental contamination of Te poses concern to organisms due to the acute toxicity of

some Te compounds, especially the soluble tellurite and tellurate anions. A small percentage

of microorganisms, however, are able to tolerate elevated levels of Te by detoxifying it

through precipitation or volatilisation. Bioaccumulation of Te compounds can occur in some

plants of the garlic family. A variety of interlinked organic and inorganic processes governs
Journal Pre-proof
Te environmental chemistry. The Te cycle in surface environments incorporates (oxidative)

dissolution of Te from primary ore minerals, inorganic precipitation and redissolution

processes in which secondary minerals are formed, and bioreductive reprecipitation and

volatilisation processes governed mainly by microbes. Our integrated Te cycling model

highlights the interplay between anthropogenic, geochemical and biogeochemical processes

on the distribution and mobility of Te in surface environments.

f
oo
HIGHLIGHTS

 Tellurium has a complex environmental geochemistry




pr
The many modes of Te bonding govern its surface behaviour
Outcropping Te deposits are an analogue for anthropogenic contamination
 We propose a cycling model for tellurium in surface environments
e-
 We highlight future areas for tellurium biogeochemical research
Pr
al
u rn
Jo
Journal Pre-proof
TABLE OF CONTENTS

ABSTRACT 2

HIGHLIGHTS 3

TABLE OF CONTENTS 4

1. INTRODUCTION 5

2. PHYSICAL A ND CHEMICAL PROPERTIES OF TELLURIUM 9

3. TELLURIUM MINERA LOGY 11

4. TELLURIUM DISTRIBUTION AND ORE DEPOSITS 13


4.1 Overview of tellurium distribution in the Earth’s crust 13
4.2 Tellurium ore deposits 15

5. TELLURIUM IN THE ENVIRONMENT 22

f
5.1 Tellurium in the oxidation zone of primary Te-rich ores 22

oo
5.2 Tellurium deportment in soils and deeper regolith environments 25
5.3 Anthropogenic tellurium 25
5.4 Tellurium toxicity 28

6. BIOGEOCHEMISTRY OF TELLURIUM
6.1 Biooxidation (bioleaching)
pr 30
31
e-
6.2 Biosorption 33
6.3 Bioreduction 34
6.4 Bioaccumulation 41
Pr

6.5 Applications 42

7. A TELLURIUM BIOGEOCHEMICAL CYCLING MODEL 44


7.1 An integrated Te cycling model – comparison with Se 44
al

7.2 Tellurium dispersion around Au-Te deposits 46

8. CONCLUSIONS 48
rn

ACKNOWLEDGEMENTS 49
u

FOOTNOTES 49
Jo

REFERENCES 51

TABLES 70

FIGURES A ND CAPTIO NS 79
Journal Pre-proof
1. INTRODUCTION

Tellurium (Te) was discovered in 1783 by Franz Joseph Müller von Reichenstein, but fully

publicised only over a decade later by Martin Heinrich Klaproth, who named the new

element after the Latin word for "earth", tellus (Emsley, 2011; Klaproth, 1798). With an

estimated crustal abundance of ~5 µg/kg (reported range 1 to 27 µg/kg; Emsley, 2011;

Vaigankar et al., 2018; Wedepohl, 1995), it is one of the least abundant elements in the

lithosphere and comparable to crustal abundances of precious metals gold (Au) and platinum

(Pt) (Emsley, 2011). Recently, Te has come into prominence due to new industrial

f
oo
applications, including in cadmium telluride (CdTe) solar panels (Diso et al., 2016; Goldfarb,

2014; Reese et al., 2018), thermoelectric devices (Bae et al., 2016; Knockaert, 2011; Lin et al.,
pr
2016), batteries (Ding et al., 2015; He et al., 2016; He et al., 2017), and nanomaterials like
e-
CdTe quantum dots (Mahdavi et al., 2018). Furthermore, the recent nuclear incident at
132
Fukushima led to severe contamination by the radioisotope Te, renewing interest in the
Pr

biogeochemical mechanisms involved in Te transport in the environment (e.g. Gil-Díaz, 2019;


al

Tagami et al., 2013).


rn

Tellurium is distributed unevenly through the Earth’s crust. Hydrothermal and magmatic
u

processes are the key mechanisms leading to high Te concentrations and the formation of
Jo

primary Te minerals (Brugger et al., 2016). Tellurium is an essential element in over 180

minerals, making it the most anomalously diverse element in mineralogy, i.e. it forms the

greatest number of minerals relative to its crustal abundance (Christy, 2015; Pasero, 2020).

Tellurides are primary minerals containing reduced Te (formal oxidation state -II to 0; e.g.,

calaverite, krennerite, sylvanite; Figure 1a-c) and elemental tellurium (Figure 1d); secondary

minerals comprise tellurites (oxidation state +IV) and tellurates (+VI) (e.g., teineite,

zemannite, and jensenite, Figure 1d- f). The designations ‘primary’ and ‘secondary’ minerals

relate to formation conditions. Primary minerals form deeper in the crust under anoxic

conditions from hydrothermal fluids or silicate melts (Ciobanu et al., 2006; Zhang and Spry,
Journal Pre-proof
1994); whereas secondary minerals form via weathering of primary minerals under the

oxidising conditions of near-surface environments (Christy et al., 2016a). Some tellurites and

tellurates possess non- linear optical properties (Norman, 2017; Weil, 2018; Yu et al., 2016)

with potential applications in the electronics industry. Nonetheless, most industrial

applications for Te utilise tellurides (Amatya and Ram, 2012; Woodhouse et al., 2013; Yeh et

al., 2008).

In 2010, the US Department of Energy classified Te as a critical metal with an anticipated

global supply shortfall by 2025 (Bauer et al., 2010), and Te remains on the list of critical

f
oo
metals published by the US Department of the Interior in 2018 (USDOI, 2018). The global

Te industry is still in its infancy with a global production of 440 metric tonnes and estimated
pr
reserves of 31,000 metric tonnes from Te contained in copper ores (Anderson, 2019).
e-
Currently, >90% of Te (along with Se) is recovered from copper anode slimes as a by-
Pr

product of the electrolytic refining of copper (Kyle et al., 2011; Makuei and Senanayake,

2018), and Te supply is thus intrinsically linked to the Cu mining industry. Recent advances
al

in industrial uses of Te focus on CdTe solar panels, which currently supply five percent of the
rn

global solar panel market (USDOE, 2019). Due to a growing world population and
u

concurrent attempts to limit man- made climate change, renewable energy industries including
Jo

CdTe solar panels are growing in prominence (see Figure 2; Frishberg, 2017; Nuss, 2019;

Wang, 2011).

The increased demand for Te will inevitably result in increasing Te contamination around

mining and industrial sites (e.g. Kagami et al., 2012), and the decommissioning of CdTe solar

panels also has the potential to be a source of contamination, in particular to groundwater

(Fthenakis and Wang, 2006; Marwede and Reller, 2012; Ramos-Ruiz et al., 2017b; Xu et al.,

2018). Recent improvements to the detection of low levels of Te via cathodic stripping

voltammetry (Biver et al., 2015) and inductively-coupled plasma mass spectrometry (Filella

and Rodushkin, 2018) have been made, although further improvements in the detection limits
Journal Pre-proof
of environmental Te are ideally required (Filella, 2019). Another short-term anthropogenic
132 129m
source of Te contamination are the radiogenic isotopes Te and Te released to the

environment in nuclear spills or explosions. This comprises both nuclear weapons testing

(particularly from the 1940s to the 1970s) and accidental spillage from power plant failure

such as the Chernobyl and Fukushima Daiichi nuclear disasters (Dickson and Glowa, 2019;
132
Yoschenko et al., 2018). The radioactive and biologically active decay product of Te, 132 I,

is of most concern, and a greater understanding of Te biogeochemical cycling could have

provided better and longer- lasting solutions for cleaning up radioactive materials following

f
oo
the Fukushima spill (Gil-Díaz, 2019).

Research in Te biogeochemistry remains overall in its infancy. The cycling of this element in
pr
near-surface environments is dynamic, as the transformation of Te oxidative states can form
e-
both inorganic and organic compounds (Belzile and Chen, 2015; Bonificio and Clarke, 2014;
Pr

Chasteen et al., 2009). In terms of biogeochemical processes at the cellular level, mechanisms

for detoxifying often involve reduction of soluble Te oxyanions (i.e., tellurite and tellurate
al

anions) (Piacenza et al., 2017; Taylor, 1996; 1999). These soluble Te oxyanions are toxic to
rn

most microorganisms at low concentrations, i.e., 1 mg/L or an equivalent 4 µM (Presentato


u

et al., 2019), which are orders of magnitude less than cytotoxic concentrations of mercury or
Jo

cadmium (Chasteen et al., 2009; Presentato et al., 2016). As such, Te solubility is a key factor

contributing to its toxicity; reduced Te compounds (e.g., meta llic Te) have low solubility and

are therefore considered less toxic as they are not bioavailable. Reduction of Te oxyanions by

microorganisms follows two major pathways, which may be either active or passive:

bioprecipitation and, to a lesser extent, biovolatilisation (Chasteen and Bentley, 2003). Active

bioprecipitation (also known as biomineralisation) results in the formation of nanoparticles of

elemental Te, and biovolatilisation in the formation of volatile, organic forms of Te such as

dimethyl telluride. In terms of passive bioprecipitation/biomineralisation, microorganisms as

well as other organic material, e.g., extracellular polymeric substances (EPS), can act as a
Journal Pre-proof
sorbent material as well as a reductant for Te oxyanions. As such, the accumulation of

reduced Te can occur over time as long as Te oxyanions are supplied to the system and a

consistent presence of biomass is available to serve as a sorbent material and reductant

(Tanaka et al., 2010). Biooxidation of Te leading to its solubility (and intuitively, its mobility

in the environment) is the less-studied process of biogeochemical cycling relative to Te

oxyanion reduction. Metal-tolerant microorganisms are capable of indirectly oxidising

tellurides by producing an oxidant (e.g. Fe-oxidisers producing Fe3+) as a by-product of their

metabolism (Climo et al., 2000b). Collectively, the summation of reduction and oxidative

f
processes contribute to the biogeochemical cycle of Te under near-surface environmental

oo
conditions and provide both organic and inorganic pathways for precipitation, sorption and

oxidation (Filella et al., 2019). pr


e-
Here we focus on the environmental (bio)geochemistry of Te, by first tying together existing
Pr

research in mineralogy, geochemistry and microbiology – each discipline extensively studied

in their own right, but not often explicitly linked. Over the past decade, various aspects of the
al

biosphere and lithosphere have been shown to influence the dissolution, re-precipitation and
rn

mobility of Se, the chalcogen element located above Te in the Periodic Table (Bailey, 2017;
u

Nancharaiah and Lens, 2015; Sharma et al., 2015; Tan et al., 2016; Ullah et al., 2018). In
Jo

addition, precious metals such as Au (Reith et al., 2013; Sanyal et al., 2019; Southam et al.,

2009) and Pt (Reith et al., 2014; 2016; 2019), which have had been generally considered to

be inert in the biosphere, have been shown to display complex biogeochemical cycling. To

our knowledge, an interdisciplinary research perspective of Te cycling is still lacking in the

literature; therefore, here we develop a model for a global biogeochemical cycling of Te in

near-surface environments. This Te biogeochemical cycling model helps to identify gaps in

our current understanding of Te biogeochemistry, and allows critical evaluation of var ious

environmental factors contributing to Te mobility, which is becoming increasingly important


Journal Pre-proof
as anthropogenic activity associated with Te applications is a greater factor in environmental

contamination.

2. PHYSICAL AND CHEMICAL PROPERTIES OF TELLURIUM

Tellurium has the atomic number 52, belongs to the chalcogen group, and is located below

oxygen (O), sulphur (S; Kagoshima et al., 2015), and selenium (Se; Ullah et al., 2018), and

above radioactive polonium (Po; Ram et al., 2019) in the Periodic Table (see Supplementary

Table 1). Tellurium has 39 known isotopes, eight of which are naturally occurring. Half of

f
those isotopes, 122 Te (natural abundance 2.5%), 124 Te (4.6%), 125 Te (6.9%), and 126 Te (18.7%),

oo
120 123
are stable. The other four naturally occurring isotopes are unstable, i.e., Te (0.09%), Te

(0.87%), 128
Te (31.8%) and 130
pr
Te (34.5%), but with long to extremely long half- lives of

~1016 , 9.2 × 1016 , 2.2 × 1024 (the longest known half- life of any isotope) and 7.9 × 1020 years,
e-
respectively (Emsley, 2011). The long- lived radioactive isotopes of Te are more abundant
Pr

than the stable isotopes by a ratio of ca. 2:1. A ninth isotope, 132 Te (half- life 3.204 days), is of

considerable concern as a nuclear waste product and was the third most abundant element
al

released by the Fukushima Daiichi nuclear disaster (Endo et al., 2018). Aside from trace
rn

132
levels found in U ores from natural nuclear fission events (Emsley, 2011), Te is of
u

129m
anthropogenic origin. Additionally, Te (‘m’ indicating a metastable isotope with the
Jo

nucleus in a long- lived excited state; 129m Te decays via gamma radiation to 129
Te with a half-

life 33.6 days) is produced in nuclear fission waste and spills (Watanabe et al., 2012; Endo et
127m 131m
al., 2018). The metastable isotopes Te and Te are also generated in smaller amounts
119m
by nuclear fission (Takahashi et al., 2019), and some other radioactive isotopes like Te

(half- life 4.70 days) are generated synthetically as precursors for medical uses (Bennett et al.,

2019).

Of the chalcogens, Te has the highest melting and boiling points, at 449.5 and 988 °C,

respectively. However, in polymetallic systems, Te is one of the “low-melting point


Journal Pre-proof
chalcophile metals” together with Ag, As, Au, Bi, Hg, Sb, Se, Sn, and Tl, forming melts at

metamorphic and hydrothermal temperatures between 500 and 600 °C, depending on

pressure (Frost et al., 2002; Tooth et al., 2011). Elemental Te is a semiconductor that is also

photoconductive (increased conductance when exposed to light; Liu et al., 2010), although Te

is more commonly combined with other elements in semiconducting applications (Marwede

and Reller, 2012).

In terms of its aqueous chemistry, Te is similar to Se, sharing fewer parallels with S and Po,

but previous studies have shown the danger of assuming that Te behaves analogous to other

f
oo
chalcogens (Chivers and Laitinen, 2015). Like S and Se, Te occurs in four main formal

oxidation states in aqueous solution: -II, +II (least stable and not known naturally), +IV and
pr
+VI. The +IV and +VI may coexist in the same solution due to slow reaction kinetics of
e-
aqueous Te oxidation and reduction (Filella and May, 2019). Reduced Te compounds (formal
Pr

oxidation state -II and elemental Te) are typically poorly soluble; for example, the solubility

product (log Ksp ; reaction 2.1) for silver telluride (hessite) is -71.7 (Chen et al., 2002):
al

Ag 2 Te(𝑠) ↔ 2Ag + (𝑎𝑞) + Te2− (𝑎𝑞), 𝐾sp = [Ag + ]2 [Te2− ] (2.1)


rn

Nonetheless, cadmium telluride, bismuth telluride and elemental Te were shown to be


u
Jo

sparingly soluble in ambient conditions by Bonificio and Clarke (2014), diffusing across an

agar plate by a poorly understood mechanism.

The simplest Te anion is telluride (Te2-). Unsurprisingly, the Te2- anion (2.21 Å) has a larger

radius than the selenide (Se2-, 1.98 Å) and sulphide (S2-, 1.84 Å) anions. The difference in

radii is 11% for Se2- but 20% for S2-, exceeding the 15% limit of the Goldschmidt rule, and

meaning that Te2- substitution for S2- involves significant structural strain (Blundy and Wood,

1994; Shannon and Prewitt, 1969); yet sulphide minerals commonly can incorporate mg/kg

levels of Te (see Table 2). Additionally, as is the case for other p block elements, the hydride

(hydrogen telluride, H2 Te(g)) is volatile (Cooke and McPhail, 2001; Grundler et al., 2013),
Journal Pre-proof
although, unlike Se, Te does not readily form a volatile chloride (Te 2+Cl2 ) when heated to

200°C in the presence of HCl with pH below 0 (Chen et al., 2016). Tellurium atoms often

polymerise to form more complex polytelluride anions with formal oxidation states between -

II and 0; the simplest aqueous polytelluride ion is the linear Te 2 2- ditelluride ion (Brugger et

al., 2012; Ruck and Locherer, 2015; Singh and Sharma, 2000). In minerals, Te participates in

varied forms of intermetallic bonding in tellurides (Bindi and Biagioni, 2018; Helmy et al.,

2007; Moëlo et al., 2008), for instance forming Te–Ag bonds in hessite; and Te–Au and Te–

Te bonds in calaverite (Figure 3). Te–Te bonds have metallic character, and metal–Te bonds

f
tend to have strong covalent character due to the small difference in Pauling electronegativity

oo
between Te (2.1) and many metals such as Cu (1.90), Ag (1.93) and Pt (2.28). The non- metal

pr
O, by comparison, has a higher electronegativity of 3.44, allowing Te 4+ and Te6+ to form
e-
ionic bonds to O (Figure 3). Geometric arrangements of Te–O bonds around a Tem+ centre

form [Tem+On ]m-2n oxyanions (Christy et al., 2016a). Tetravalent Te has a stereochemically
Pr

active lone electron pair, and as a result forms a variety of distorted (hemidirected)

coordination polyhedra (Christy and Mills, 2013), e.g., [Te 4+O 3 ]2-, [Te4+O4 ]4- and [Te4+O5 ]6-;
al

sometimes there is more than one coordination in the same structure (Missen et al., 2019). In
rn

contrast, Te6+ lacks a lone electron pair, and occurs exclusively in minerals as Te6+O6
u

octahedra (Christy et al., 2016a; Mills and Christy, 2013) (Figure 3). The structural diversity
Jo

of Te–O compounds is further discussed in the following section.

3. TELLURIUM MINERALOGY

As a chalcophile element (Christy, 2018), Te typically occurs in sulphide ore deposits with

other chalcophile elements such as Ag, Cu and Pb (see Table 2). Rather than forming

separate telluride minerals, which tend to be minor phases in most deposits, the bulk of the

primary Te is typically found associated with sulphides, i.e., Te substitutes for S in common

sulphide minerals such as chalcopyrite, covellite, galena, pyrite, pyrrhotite and

sphalerite (Brugger et al., 2016; Dill, 2010; Hattori et al., 2002; Simon and Essene, 1996;
Journal Pre-proof
Vikentyev, 2006). Pyrite is the most studied host sulphide for Te (Cook et al., 2009a;

Deditius et al., 2010, 2014; Dmitrijeva et al., 2020; Keith et al., 2018; King et al., 2014). The

mechanisms of Te incorporation in pyrite are complex and dynamic, involving for example

the preferential incorporation of Te on the (1 1 0) pyrite surface (Chouinard et al., 2005).

Tellurium minerals form in a variety of geological environments (see Section 4.2). As

previously discussed, Te mineralogy is the most ‘anomalously diverse’ of all elements

despite its low crustal abundance (Christy, 2015). As of 2020, the International Mineralogical

Association (IMA) recognises more than 180 Te minerals (Pasero, 2020), comprising almost

f
oo
equal numbers of primary (oxidation states almost entirely#1 -II to 0) and secondary

(oxidation states +IV and +VI) minerals. Important Te minerals are listed in Table 1, and a
pr
full current list is provided in Supplementary Table 2.
e-
Primary minerals form the bulk of Te minerals by weight/natural abundance. Gold and/or Ag
Pr

tellurides are the most important primary Te minerals (See also Table 1). There are fourteen

Au–Te minerals, the most common being calaverite (AuTe 2 , Figure 1a), krennerite
al

(Au3 AgTe8 , Figure 1b), sylvanite [(Au,Ag)2 Te4 , Figure 1c] and petzite (Ag3 AuTe2 ); and
rn

twenty-two primary Ag–Te minerals, the most common of which is hessite, Ag2 Te.
u

Elemental Te is also a primary mineral in a few deposits (Figure 1d). There are 36 Bi–Te
Jo

minerals (29 primary and 7 secondary minerals)#2 . Primary Bi–Te minerals are abundant in

some deposits (Ciobanu et al., 2009), for example tetradymite (Bi2 Te2 S) is the main Te ore

mineral in the Dashuigou deposit, China (Mao et al., 2002). Reliable thermodynamic data are

available for only a handful of telluride minerals, including calaverite (Mills, 1974) and two

Pt telluride minerals (Olivotos and Economou-Eliopoulos, 2016).

Secondary Te minerals develop in the topmost reaches of deposits, where primary minerals

come into contact with oxygenated groundwaters (Dill, 2010; Williams, 1990). Although

secondary minerals of some elements are prevalent enough to form economic deposits (e.g.,
Journal Pre-proof
non-sulphide Zn deposits; Boni et al., 2003; Hitzman et al., 2003), secondary Te minerals are

rare and occur only in sub-economic amounts. Additionally, obtaining thermodynamic data

for these minerals is near impossible because of the difficulty in obtaining sufficient amounts

of pure material; synthesis usually leads to fine-grained mixtures.

Figure 4 shows the structural diversity of Te4+–O minerals, including isolated Te4+On

polyhedra, and Te4+On polyhedra linked in non-cyclic finite units, infinite chains, infinite

layers and infinite frameworks. This diversity explains the large amount of tellurite minerals,

and is due to the flexible and asymmetric coordination environment of the Te 4+ ion, which

f
oo
results from its lone electron pair (Christy and Mills, 2013). Similar structural diversity, from

units to frameworks, is possible for Te6+–O minerals, although most tellurate minerals
pr
contain Te6+O6 octahedra isolated from each other, with much of the structural diversity
e-
driven by the varied crystal chemistry of associated cations like Cu2+ (Christy et al., 2016b).
Pr

Cyclic finite units of TeO n polyhedra are not yet known in any Te4+ or Te6+ mineral (Christy

et al., 2016a).
al
rn

4. TELLURIUM DISTRIBUTION AND ORE DEPOSITS

4.1 Overview of tellurium distribution in the Earth’s crust


u
Jo

Tellurium is much more abundant in the solar system – 2.28-2.32 mg/kg Te based on C1

chondrite (Anders and Grevesse, 1989; Lodders, 2010) – than in the Earth’s crust (5 µg/kg).

The low crustal abundance of Te results from three processes active during the Earth’s

formative years. First, the Earth formed from Te- and other volatile-depleted chondritic

meteorites (Braukmüller et al., 2019). Second, volatile Te compounds such as hydrogen

telluride (H2 Te) formed and were subsequently lost to outer space during accretion (O'Neill

and Palme, 2008). Finally, Te acts as a siderophile under highly reducing conditions,

resulting in its sequestration with metallic Fe and Ni in the Earth’s core (Rose-Weston et al.,

2009; Wang and Becker, 2013). Hence, most Te in the crust is thought to have been added in
Journal Pre-proof
a ‘late veneer’ (Wang and Becker, 2013), which over geologic timescales has differentiated

into the characteristic uneven distribution of Te known today.

The mobility of Te from the mantle to the atmosphere can be grouped according to prevalent

physical, geochemical and biochemical processes (Figure 5). Major Te sinks include mineral

deposits, weathered rocks, soils, and sediments, with the processes governing the interplay

between these sinks further discussed in Section 7.1. Tellurium is generally below the level of

ng/m3 in airborne particulates (Belzile and Chen, 2015), locally higher near Te-rich areas due

to release of volatiles. Tellurium is also highly depleted in seawater (<2 ng/L) and surface

f
oo
freshwater (<20 ng/L) compared to its crustal abundance (Table 3) (Belzile and Chen, 2015;

Emsley, 2011; Gil-Díaz et al., 2019a; Wedepohl, 1995). This is due to the strong sorption of
pr
Te(IV/VI) oxyanions onto mineral surfaces (Section 5.2; Qin et al., 2017), and as a result, Te
e-
tends to be enriched in soils (average around 35 µg/kg) (Figure 5); Te is particularly enriched
Pr

locally near Te deposits and in ferromanganese nodules on the ocean floor (Baturin, 2012).

Ferromanganese nodules contain on average ~1 mg/kg Te, i.e. three orders of magnitude
al

higher than the average crustal concentration, with some nodules containing over 200 mg/kg
rn

(Figure 5; Baturin, 2012; Hein et al., 2003). Elevated levels of Te are also found in red beds
u

and black shales. Reduction spheroids in red beds (British Isles Triassic, Parnell et al., 2016;
Jo

worldwide Mesoproterozoic, Parnell et al., 2018) commonly display average enrichment to 1-

10 mg/kg, and often contain micron-scale grains of telluride minerals. Neoproterozoic black

shales from the Gwna Group in the British Isles contain up to 30 mg/kg Te (Armstrong et al.,

2018).

Volcanoes are another source of environmental Te, where volatile forms of Te (e.g., H2 Te(g))

are released through eruptions, quiescent degassing (e.g., fumaroles), or hydrothermal

activity. For example, Te occurs at 10-1000 mg/kg levels in volcanogenic sulphur (Figure 5;

Yu et al., 2019). The high temperature (~600 ˚C) fumaroles at the Avacha Volcano,

Kamchatka, Russia, contain up to 15.9 mg/kg Te (Okrugin et al., 2017) (Table 3), and native
Journal Pre-proof
Te has been identified in fumarolitic deposits from Vulcano, Italy (Fulignati and Sbrana,

1998). Deep-sea black-smoker massive-sulphide and volcanogenic-sulphur deposits typically

contain Te enriched by four orders of magnitude compared to average crustal levels (Butler

and Nesbitt, 1999; De Ronde et al., 2015; Greenland and Aruscavage, 1986; Yu et al., 2019).

Tellurium is also enriched in organic-rich rocks and coals (Belzile and Chen, 2015), with

concentrations up to 2 mg/kg in the pyritic coals of Brora, Scotland (Bullock et al., 2017).

Tellurium stored in coal is subsequently released into the atmosphere through the burning of

coal to produce energy (see Figure 5). There are relatively few studies on the Te content of

f
oo
crude oil and natural gas. Green (2011) calculated that 4 to 220 tonnes of Te were processed

during crude oil refinement in 2010, and <100 tonnes of Te processed during natural gas
pr
refinement, though the average amounts in each source are likely to be <1 mg/kg. Parnell et
e-
al. (2015) used pyrite Te levels as a proxy for Te content of natural oil reservoirs. The
Pr

average Te content of pyrite associated with biodegraded palaeo-oil reservoirs is in the 100s

of µg/kg, as opposed to less than 30 µg/kg for other pyrites in central England (Parnell et al.,
al

2015).
rn

4.2 Tellurium ore deposits


u

4.2.1 Overview
Jo

Tellurium is produced from only a handful of localities, and no deposit is currently mined

solely for Te. For instance, Te is recovered as a by-product of Au mining from the Kankberg

mine in Sweden (Goldfarb, 2014; Goldfarb et al., 2017), which has local epithermal- like

mineralisation in a volcanogenic massive sulphide district; and together with Au from the

epithermal vein deposits of Dashuigou and Majiagou in the Sichuan and Shaanxi provinces of

China (Mao et al., 2002). Additionally, the company Deer Horn Capital expects to begin

production of Te together with Au and Ag from the Deer Horn area of British Columbia,

Canada. The average Te grade at Deer Horn is 118 mg/kg Te, with an expected total amount
Journal Pre-proof
of 67 tonnes recoverable (Meintjes et al., 2018). Other major producers of Te are the United

States, Canada, Peru and Japan (Anderson, 2019; Emsley, 2011). Although it is rarely

recovered, Te is enriched (100’s of µg/g to wt.%) in many base- and precious- metal deposits.

The largest reserves based on the amounts of Te produced by Cu mining are in China

(6,600 metric tons), Peru (3,600 metric tons) and the United States (3,500 metric tons)

(Anderson, 2019). Nonetheless, the disconnect between relatively large Te reserves and the

relatively small amounts actually produced means that Te is classified as a critical mineral

commodity (USDOI, 2018).

f
oo
The most important deposit types are of magmatic and hydrothermal origin. Note that in this

and the following sections, we list only the most Te-rich deposits, or those where Te is an
pr
important constituent of the economic mineral assemblage. Tellurium enrichment occurs in
e-
association with Au and/or Ag in many (usually hydrothermal) ore deposits, and with
Pr

Platinum Group Minerals (PGMs) in a smaller number of deposits (usually magmatic, e.g.

Ciobanu et al., 2006; Keith et al., 2018; Pals et al., 2003). Several magmatic Copper–Nickel–
al

Platinum- Group-Metal (Cu–Ni–PGM) sulphide deposits (e.g. Noril’sk (Genkin and


rn

Evstigneeva, 1986) and Kola Peninsula (Subbotin et al., 2019) in Russia; Bushveld in South
u

Africa (Kingston, 1966); and Sudbury in Canada (Dare et al., 2014)) are enriched in Te (e.g.,
Jo

up to 150 mg/kg in chalcopyrite-rich massive sulphide ores from Sudbury; Dare et al., 2014).

The major sulphide minerals in Cu–Ni–PGM deposits are pyrrhotite, pentlandite and

chalcopyrite, which may host Te themselves (Table 2), although Te-rich Cu–Ni–PGM

deposits tend to contain small amounts of tellurides (Holwell et al., 2017). Even though Cu–

Ni–PGM sulphide deposits are the only non- hydrothermal Te deposit type, they host some of

the larger Te reserves.

Outside of magmatic deposits, the majority of Te deposits are hydrothermal in nature,

encompassing the following deposit types (Goldfarb et al., 2017); numbers in square brackets

refer to locality number in the map shown in Figure 6:


Journal Pre-proof
 Epithermal deposits, e.g. Cripple Creek [49], Colorado, USA (Kelley et al., 1998;

Thompson et al., 1985), Emperor [4], Fiji (Pals and Spry, 2003; Pals et al., 2003),

Moctezuma mines [56], Mexico (Deen and Atkinson Jr, 1988) and Săcărâmb [37],

Romania (Ciobanu et al., 2004).

 Orogenic gold deposits, e.g. Sunrise Dam [78], Australia (Sung et al., 2007, 2009) and

Ashanti [13], Ghana (Bowell, 1992).

 Volcanogenic Massive Sulphide (VMS) deposits; Ural Mountains [28], Russia is by

far the best described (Vikentyev, 2006).

f

oo
Iron Oxide-Copper-Gold (IOCG) deposits, e.g. Olympic Dam [1], Australia (Rollog et

al., 2019).

 pr
Porphyry deposits, e.g. Almalyk [19], Uzbekistan (Cheng et al., 2018) and Mount
e-
Milligan [74], British Columbia, Canada (LeFort et al., 2011).


Pr

Intrusion-related Au deposits, e.g. Dongping Au-Te field [79] in Hebei, China

(Sillitoe and Thompson, 1998) and Bjorkdal [81], Sweden (Roberts et al., 2006).
al

 Skarn deposits, e.g. Ortosa [49], Spain (Cepedal et al., 2006) and Geodo, [16A] South
rn

Korea (Kim et al., 2012b).

 Carlin-type gold deposits, e.g. Deep Star [69], Nevada, USA (Fleet and Mumin, 1997;
u
Jo

Heitt et al., 2003) and Zarshuran [31], Iran (Asadi et al., 2000; Mehrabi et al., 1999).

As the majority of Te deposits are hydrothermal, and indeed hydrothermal processes are key

to determining the modern surface distribution of Te, the hydrothermal geochemistry of Te is

discussed in the following section.

4.2.2 Hydrothermal geochemistry of tellurium

In contrast to surface waters, hydrothermal fluids can be enriched in Te, although few data

are currently available. Waters from modern geothermal systems routinely carry Te at the

µg/L level (seawater <2 ng/L), and fluid-inclusions from Te-rich epithermal and porphyry

systems may carry as much as 100’s of mg/L Te (Table 3). Tellurium transport in
Journal Pre-proof
hydrothermal fluids remains “incompletely understood” (Goldfarb et al., 2017) due to the

paucity of experimental data at elevated temperatures. The seminal work of McPhail (1995)

still provides an extensive review of Te thermodynamic data at room temperature, as well as

isocoulombic extrapolations, allowing prediction of Te transport and deposition in waters and

vapours to 350˚C. Experimental studies by Brugger et al. (2012), Grundler et al. (2013) and

Etschmann et al. (2016) provide the only direct evidence of the nature and geometry of Te

complexes at elevated temperatures, as well as updated thermodynamic properties for a

number of key complexes. Filella and May (2019) provide a modern reassessment of the

f
available thermodynamic properties for the Te–O–H system at 25 ˚C; we chose to retain the

oo
properties of McPhail (2005) and Grundler et al. (2013), because (i) the choice of properties

pr
changes little in the speciation diagram shown in Figure 7A; (ii) these properties allow
e-
calculations at elevated temperatures; (iii) they are freely available for use (supplementary

material in Ram et al. 2019); and (iv) they more accurately reproduce the few available
Pr

experimental data (Figure 8). A list of recommended thermodynamic properties is listed in


al

Table 4.
rn

Tellurium is generally extremely poorly soluble under reducing conditions, in the form of

tellurides (e.g., Te2- and HTe-) and polytelluride (e.g., Te2 2-) complexes (Figure 7). Brugger et
u
Jo

al. (2012) showed that polytelluride species are stable to high temperatures (599 °C; 800 bar),

and are thus expected to be an important form of Te in basic fluids at high temperatures (e.g.,

CO2-rich fluids) (Cook et al., 2009c; Cooke and McPhail, 2001; Gao et al., 2017; Grundler et

al., 2013; Keith et al., 2018). Brugger et al. (2012) also showed that branched polytelluride

species are stable at low temperature, but only the Te2 2- dimer remains at high temperature.

Under more oxidising conditions, the tellurite complexes H2 TeO 3 (aq), HTeO 3 -, and TeO 3 2-

become increasingly stable as temperature increases (Figure 7). At 300˚C, they are stable

close to the hematite/magnetite buffer (Figure 7), leading Grundler et al. (2013) to propose

that these complexes account for Te transport in some Te-rich systems (e.g., porphyry-
Journal Pre-proof
epithermal systems). In situ XAS measurements by Grundler et al. (2013) showed that these

tellurite complexes share a trigonal pyramidal [TeO 3 ] geometry (with Te at the apex and the

three O atoms at the vertices), and the electron lone pair above the pyramid. The one-sided

geometry of Te4+ complexes plays a key role in their interaction with other species. In general,

Au and Te tend to have higher solubility in basic fluids (Brugger et al., 2012; Figures 7,8),

meaning that alkaline and silica-undersaturated host rocks support more efficient transport

and enrichment of Te and Au (Smith et al., 2017). In contrast to S and Se, Te can form

complexes with halides (metal- like behaviour). Etschmann et al. (2016) showed that the

f
square pyramidal [TeCl4 (aq)] complex occurs in very acidic brines, but this complex is likely

oo
to play a minor role in natural environments. Figure 7 also illustrates that tellurate complexes

pr
[typically octahedral, e.g. H6 TeO 6 (aq)] exist only at low temperature in the presence of
e-
significant amounts of molecular oxygen, and hence are important only in (near)-surface

environments.
Pr

An important aspect of the hydrothermal geochemistry of Te is the fact that Te partitions


al

strongly into the vapour phase upon separation under reducing conditions in the form of
rn

H2 Te(g), leading to relative enrichment of Te in the vapour phase. These volatile phases are

believed to play an important role in the formation of “bonanza” Au–Te ores in epithermal
u
Jo

systems (Cook et al., 2009c; Cooke and McPhail, 2001; Gao et al., 2017; Grundler et al.,

2013; Keith et al., 2018).

Finally, as a “low-melting point chalcophile metal”, Te will also be enriched together with Ag,

As, Au, Bi, Hg, Sb, Se, Sn, and Tl in melts at temperatures between 500 and 600 °C (Frost et

al., 2002; Tooth et al., 2011). These melts can play a key role in Te enrichment and Au-Ag

mineralisation, either via physical migration (Tomkins et al., 2006), via interaction with

hydrothermal fluids (McFall et al., 2018; Tooth et al., 2008; 2011), or with magmatic vapours

(Okrugin et al., 2017; Yu et al., 2019).


Journal Pre-proof
4.2.3 Distribution of Te deposits

Figure 6 shows the worldwide distribution Te deposits with respect to the underlying crustal

provinces. Localities are itemised in Supplementary Table 3. The majority of Te deposits

occur in Phanerozoic Orogenic belts, which reflects the predominance of epithermal and

porphyry-type deposits, the role of subduction in the formation of these deposits, and the

destruction of these shallow deposits by erosion in older terrains (Kesler and Wilkinson,

2008). This relation is particularly noticeable on the western side of North America and

through the central Orogen provinces of Eurasia. Links between magmatic Cu–Ni–PGM

f
oo
deposits and porphyry or epithermal Cu–Au(–Te) deposits have recently been determined,

with Te a useful tracer of the metallogenic continuum between these deposit types (Holwell

et al., 2019). pr
e-
Epithermal Au and Te deposits are some of the largest and highest grade Te resources.
Pr

Epithermal deposits are typically associated with alkaline volcanic rocks, and although

limited in their spatial distribution, they are economically important (du Bray, 2017).
al

Tellurium- rich epithermal deposits typically contain abundant tellurides that can host a large
rn

proportion of the precious metals Au and Ag, as well as heavy metals such as Pb (e.g., altaite),
u

Bi (e.g., tetradymite) and Hg (e.g., coloradoite) (Dill, 2010). Telluride minerals typically
Jo

form late in the paragenesis, after the bulk of associated S has been removed from solution to

form sulphide minerals (Watterson et al., 1977).

Porphyry Cu–Au–Te deposits are mainly found in western North America in the Orogenic or

Extended Crust provinces, and include famous localities like Bingham Canyon, Utah and the

world-class but not yet producing Pebble deposit in remote western Alaska.

Although many orogenic gold deposits contain minor amounts of telluride minerals, a few

have significant amounts of Au bound to Te; these include California’s gold-rush deposits in

the Mother Lode region, and Sunrise Dam in Western Australia (Sung et al., 2009; 2007).
Journal Pre-proof
Note that the genesis of the Te-rich mineralisation in some orogenic Au deposits remains

controversial; the most famous example is the giant Golden Mile deposit, Western Australia

(Bateman and Hagemann, 2004; Shackleton et al., 2003), in which the Te-rich ores have been

attributed to an epithermal- like genesis (Clout et al., 1990), but more recently have been

linked to higher temperature (400˚C) deep magmatic-derived fluids (Mueller et al., 2020).

Tellurium- rich skarns are rare and feature only a handful of examples, spread across several

continents, from Geodo in South Korea (Kim et al., 2012b) to Hedley in British Columbia,

Canada (Ray et al., 1987). Gold tends to be the main commodity in Te-bearing skarns, with

f
oo
lower levels of Te associated with skarn Au deposits than in other deposit types.

pr
Volcanogenic massive sulphide (VMS) deposits only comprise a handful of Te-rich examples,

including the Central Asian Au mines of Almalyk, Uzbekistan (Cheng et al., 2018) and Zod,
e-
Armenia (Konstantinov and Grushin, 1970), with the Te associated with the sulphide
Pr

minerals. Kankberg, Sweden produces around 10% of the world’s Te as a by-product of Au

processing and occurs in a VMS region, but the local mineralisation at Kankberg itself is
al

epithermal (Goldfarb et al., 2017).


rn

A handful of Carlin- type and intrusion-type Au deposits also contain small amounts of Te.
u

Aside from a cluster of three Carlin-type deposits in Nevada (Getchell, Meikle and Deep
Jo

Star), which contain up to a maximum of 200 mg/kg Te in the ores, the only other Carlin

deposit with significant Te is the Zarshuran deposit in Iran (Asadi et al., 2000).

Perhaps the most unusual Te deposit is the IOCG deposit of Olympic Dam, Australia – the

world’s largest IOCG deposits. The ore contains ~2.5 mg/kg Te, which translates into

~24,000 tons of Te out of 9,576 Mt ore (Ehrig et al., 2012; Rollog et al., 2019).

Magmatic deposits commonly occur in Shield and Platform geologic provinces, which

encompass areas with relatively flatter terrain than the Orogenic provinces. Montana’s

Stillwater complex is the only exception, being located in the Rocky Mountains on the
Journal Pre-proof
boundary of the Orogenic and Large Igneous provinces. Te whole rock grades at the

Stillwater complex are relatively low, not exceeding 20 mg/kg (Zientek et al., 1990), but still

exceeding baseline Te levels by four orders of magnitude.

4.2.4 Tellurium as a tool in mineral exploration

The enrichment of Te in precious and base metal deposits, coupled with its occurrence in

distinctive zones around ore deposits, means that it may be used as a geochemical tracer in

mineral exploration. Watterson et al. (1977) showed that proximal haloes containing at least

100 µg/kg Te extend up to 10 km away from central orebodies in the United States, with

f
oo
proximal concentrations in excess of 10 mg/kg. For instance, the Ely Au-Cu-Ag porphyry

deposit in Nevada also has a Te-containing halo, with an average of 100 mg/kg Te recorded
pr
in the jasperoids, gossans, and highly altered sedimentary carbonates in the mineralised zone
e-
of the district. The levels of Te actually increase toward the boundaries of the Ely deposit,
Pr

which in total encompasses around 40 km2 surface area and approximately 60 km3 of Te-

enriched rock (Gott and McCarthy, 1966; Watterson et al., 1977). However, the use of Te in
al

geochemical exploration for base and precious metals remains limited, possibly because its
rn

usefulness is limited by our poor understanding of Te dispersion processes.


u
Jo

5. TELLURIUM IN THE ENVIRONMENT

5.1 Tellurium in the oxidation zone of primary Te-rich ores

The preservation of secondary minerals in the oxidation zone provides a window into the

dissolution process, allowing us to see secondary minerals that record processes and

conditions when the original hypogene ore is at complete equilibrium with atmospheric

oxygen (Williams, 1990). Overall, an improved understanding of the deportment of Te from

exposed Te deposits will also provide a useful analogy for anthropogenic Te-contamination

(Filella et al., 2019). The first step in generating secondary Te minerals is the release of Te

oxyanions into solution via dealloying of tellurides and/or the oxidation of other primary Te-
Journal Pre-proof
bearing minerals, predominantly sulphides (Table 2). Solubilised Te oxyanions may

precipitate by reaction with other solubilised cations of metals like Fe 3+, Zn2+, Cu2+, UO 22+

and/or Pb2+ (Christy et al., 2016b; Figure 9A), and dealloying often leaves behind the

precious (and less soluble) metals in native form (Okrugin et al., 2014; Figure 9B-D).

Secondary minerals often surround primary minerals as coatings or halos (Figure 1d). The

exact formation conditions and crystallisation processes of individual secondary minerals are

not fully understood, although contributing factors include local chemical environment, pH

and redox potential (Christy et al., 2016b). A lack of thermodynamic data (due in part to the

f
oo
lack of synthetic analogues) on Te minerals contributes to the difficulty in understanding

their geochemical formation conditions (Christy et al., 2016b; Filella and May, 2019). Upon
pr
deeper weathering, as observed above the high- grade Bambolla vein at Moctezuma, Mexico,
e-
Te occurs predominantly in association with Fe-(Mn)-oxy-hydroxides (Figure 10),
Pr

presumably in sorbed forms (Hayes and Ramos, 2019).

Several studies have addressed the weathering of tellurides under mild (mostly ≤ 220˚C)
al

hydrothermal conditions; under these conditions the reaction occurs over a time-scale of
rn

hours, whereas at ambient conditions months to years are required (Tenailleau e t al., 2006).
u

Studied minerals include calaverite, AuTe2 (Zhao et al., 2009; Equation 5.1); krennerite,
Jo

Au3 AgTe8 (Xu et al., 2013; Equation 5.2); and sylvanite, (Au,Ag)2 Te4 (Zhao et al., 2013;

Equations 5.3, showing dissolution of sylvanite, and 5.4 showing the coupled precipitation of

calaverite-I). The oxidation of these tellurides leads to the formation of “mustard gold”

(Figure 9C,D) through dissolution of the parent telluride, subsequent reprecipitation of gold

(or gold-silver alloy) and diffusion of the aqueous Te away from the reaction front in the

solution (Altree-Williams et al., 2015; Zhao et al., 2009; Figure 9B). Although to date these

reactions have been studied in abiotic conditions, in the weathering environment they may be

microbially mediated.
Journal Pre-proof
AuTe2 (𝑠) + 2O2 (𝑎𝑞) + 2H2 O ↔ 2H2 TeO3 (𝑎𝑞) + Au(𝑠) (5.1)

Au3.28 Ag 0.72 Te8 (𝑠) + 8.05O2 (𝑎𝑞) + 7.89H2 O + 0.23HCl(𝑎𝑞)

↔ 3.77Au0.87 Ag 0.13 (𝑠) + 0.23AgCl(𝑠) + 8H2 TeO3 (𝑎𝑞) (5.2)

Aux Ag 2−𝑥 Te4 (𝑠) + 4.5O2 (𝑎𝑞) + 2HCl(𝑎𝑞) + 3H2 O

↔ 𝑥AuCl(𝑎𝑞) + (2 − 𝑥 )AgCl(𝑎𝑞) + 4H2 TeO3 (𝑎𝑞) (5.3)

(1 − 𝑥 )AuCl(𝑎𝑞) + 𝑥AgCl(𝑎𝑞) + (2 − 𝑥 )H2 TeO3 (𝑎𝑞)

↔ (Au1−𝑥 Ag 𝑥 )Te2−𝑥 (𝑠) + HCl(𝑎𝑞) + (1.5 − 𝑥 )H2 O

f
oo
+ (2.25 − 𝑥 )O2 (𝑎𝑞) (5.4)

pr
Some of the best-known secondary Te mineral localities are relatively small epithermal Au

and/or Ag deposits or prospects that are also rich in Te, most notably the Moctezuma mines,
e-
Sonora, Mexico (21 new Te minerals; Jacobson et al., 2018; see Figure 6, locality 56) and the
Pr

Otto Mountain mines (16 new exclusively Te–O minerals; Christy et al., 2016b), California,

United States; locality 59). The prevalence of these deposits and description of many rare
al

secondary minerals on the western side of North America is partly due to avid mineral
rn

collector interest in the area, meaning that most potential Te mineral localities have been
u

extensively explored (Localities 49-71 in Figure 6). Many secondary Te minerals are known
Jo

from just one locality, indicating that highly specific conditions are required for the fo rmation

of some of these phases (Christy et al., 2016a; Kampf et al., 2016).

Despite the prevalence and diversity of secondary Te minerals in some localities, others

display little or no secondary Te mineralisation. For example, secondary Te minerals are

virtually unknown in Australia, despite a number of Te-rich deposits and occurrences

(Figure 6). These deposits occur in old cratons, and have been subjected to weathering since

at least the Mesozoic. This suggests that climate, intensity of weathering, and the metallic

elements in the parent tellurides play an important role in controlling the paragenesis of
Journal Pre-proof
secondary Te minerals. Extreme, prolonged weathering results in the dispersion of Te: deep

weathering generates large amounts of Fe oxides, and these scavenge Te via sorption,

limiting Te solubility and precipitation of secondary Te minerals (Hayes and Ramos, 2019;

Figure 10). In young (<~10 My) deposits, weathering is limited, and although some

secondary Te minerals do form, for example at the Late Miocene (6.9–7.1 My) Aginskoe

deposit, Kamchatka, Russia (Andreeva et al., 2013; Takahashi et al., 2013; Figure 9), the

extent of the secondary Te mineral occurrence is highly limited (Okrugin et al., 2014). Mills

and Christy (2019) dated the secondary U-tellurite minerals schmitterite and moctezumite

f
from the Tertiary Moctezuma deposit, Mexico, to 436.5 ±27.1 and 502.8±46.0 (schmitterite)

oo
and 31.9±0.2 and 274.8±9.1 (moctezumite) ky old. These data suggest that erosion rate and

pr
climate play a key role in the development and preservation of secondary Te minerals.
e-
5.2 Tellurium deportment in soils and deeper regolith environments
Pr

In the regolith, immediately following the dissolution of tellurides, Te is generally bound by

sorption onto clay-sized soil particles rather than in minerals (Fairbrother et al., 2012;
al

Goldfarb et al., 2017; Hayes and Ramos, 2019). In particular, a strong association has been
rn

observed between Fe3+ oxide minerals and Te (Qin et al., 2017). Te6+ can be incorporated
u

into the structures of Fe3+ oxides, whereas Te4+ tends to be bound more weakly to Fe3+ oxides
Jo

by surface interactions only (Qin et al., 2017). Both Qin et al. (2017) and Hayes and Ramos

(2019) studied Te speciation in tailings piles, meaning that subsequent studies are required to

analyse Te behaviour in different contexts, namely sites which are less contaminated by

anthropogenic activities.

5.3 Anthropogenic tellurium

To date, the major anthropogenic activities involving Te are mining and ore processing (Te

being a by-product of Cu mining in particular) and fossil fuel burning, with emission

concentrations of Te on the order of mg/kg (Figure 5). Given the ~five-fold increase in Te
Journal Pre-proof
usage since 1940 (Nuss, 2019), environmental release of Te from Te-producing processes,

manufacturing Te-bearing industrial products, and subsequent decommissioning of these

products is expected to escalate. Highly toxic Te was also released during the Chernobyl and

Fukushima nuclear disasters; this is covered in the following section.

Currently, the production of CdTe solar cells consumes 40% of global Te output. Other

applications include thermoelectric production (30%), metallurgy/alloys (15%), rubber (5%),

and as a paint or pigment for glasses, enamels, and plastics (Anderson, 2019). Further uses of

Te include manufacture of CdTe quantum dots, readily synthesised by pyrolysis (Murray et

f
oo
al., 1993). The semiconducting properties of these quantum dots make them useful for a wide

variety of medical (William et al., 2006) and electronics (Kumar and Kumar, 2015)
pr
applications. CdTe solar panels have their CdTe firmly encapsulated in protective casing,
e-
meaning that the operational risk of toxic exposure to humans is virtually zero (Biver and
Pr

Filella, 2016).

Nearly 90% of Te is produced by extraction from anode slimes using a series of


al

pyrometallurgical and hydrometallurgical operations to remove Se, Te and other precious


rn

metals (Kyle et al., 2011; Makuei and Senanayake, 2018). The mined ore is subjected to
u

milling, flotation, smelting, casting and finally electrolysis in which Te is separated from Cu
Jo

and collects alongside other impurities as tellurides in the anode slime, now enriched 10 5

times compared to the Te concentration in the Cu ore (Makuei and Senanayake, 2018). The

methods for recovering Te from the slimes are determined by chemical and phase

composition and content of associated metals. The tellurides are subjected to

hydrometallurgical treatment either by direct leaching of the raw slimes with sulphuric acid

in the presence of oxygen (pressure) or aeration (atmosphere), or by pressurised leaching in

alkaline solution. Tellurium has also been extracted from lead(–zinc) smelting processes

(Makuei and Senanayake, 2018). Other potential sources for Te extraction are flue dusts and

gases generated during the smelting of bismuth, copper, and gold ores (Kyle et al., 2011).
Journal Pre-proof
Tellurium contamination is commonly a by-product of the processing of Te-bearing raw

materials via Te-rich wastewaters (e.g. Cu-processing, Kagami et al., 2012; Shibasaki et al.,

1992; sulphuric acid production, Zonaro et al., 2017) or emissions from smelters (Chien and

Han, 2009; Perkins, 2011). Mining activities and in particular tailings piles and dams may

also leach Te, especially from Au mining, or even the mining of o ther sulphide deposits, for

instance at the Kawazu mine, Shizuoka, Japan (Qin et al., 2017; see Figure 6, locality 18) and

at Rodalquilar, Spain (Wray, 1998; see Figure 6, locality 34). Tellurium contamination also

results from electronics waste processing and recycling (Shuva et al., 2016). Sorption of Te is

f
one method which may be used to remediate a Te-contaminated site or waterway, through

oo
carbon-based sorbents (Dimpe and Nomngongo, 2017) or biosorption (Piacenza et al., 2017),

pr
potentially recycling the sorbed Te for future use. In this manner current sources of Te
e-
contamination may be transformed into future sources of Te (Shuva et al., 2016).
Pr

CdTe has a band gap of 1.45 eV at 300 K, making it perfectly matched to the peak of the

solar spectrum (Wu, 2004), leading to ever- increasing use in the solar panel industry.
al

Discarded cadmium telluride (CdTe) solar panels and to a lesser extent, other thermoelectric
rn

materials like Bi2 Te3 (with specific uses in the electronics industry) are likely to increase the
u

anthropogenic output of Te into the environment (Cyrs et al., 2014; Marwede and Reller,
Jo

2012; Zeng et al., 2015). The dissolution of commonly used tellurides must be understood so

that the Te (and associated elements) do not become environmental contaminants (Ramos-

Ruiz et al., 2017b). Dealloying of CdTe occurs most efficiently when the decomposing solar

panels are exposed to oxidising, acidic conditions (Ramos-Ruiz et al., 2017b), with the

reduction rate decreasing monotonically with increasing pH (Biver and Filella, 2016). Unlike

for some chalcogenides, the CdTe dissolution rate does not stop completely under anoxic

conditions, and increases near-linearly with increasing dissolved oxygen. Bi2 Te3 behaves

markedly differently, with no oxygen dependence, and a minimum rate of reduction at pH 5.3

(Biver and Filella, 2016). These two examples show the importance of understanding the
Journal Pre-proof
dissolution regimes of individual tellurides, which may vary markedly from telluride to

telluride.

5.4 Tellurium toxicity

132
Soluble Te oxyanions, the short- lived radioactive isotope Te, and nanoparticles of both

elemental Te and cadmium telluride (CdTe) are the most common toxic forms of Te. Soluble

forms of Te are more toxic than their insoluble counterparts based on the increased

bioavailability of Te from soluble compounds, with toxicity beginning at extremely low

levels for microbes (1 mg/L for tellurite; Presentato et al., 2019). In general, the tellurate

f
oo
anion is 2-10 times less toxic than tellurite (Cunha et al., 2009). Both soluble oxyanions are

most likely to be encountered in contaminated wastewaters or around the weathering zone s of

Te deposits.
pr
e-
For humans, ca. 90% of ingested Te accumulates in the bones (Gerhardsson, 2015), while
Pr

accidental ingestion of microgram amounts o f TeO 2 or Na2 TeO 3 results in body odour and

garlic breath due to the metabolism of the Te 4+ oxysalts to gaseous dimethyl telluride,
al

(CH3 )2 Te (Cunha et al., 2009; Gerhardsson, 2015). Te is expected to be processed in a similar


rn

manner to Se, although unlike Se, Te is not an essential micronutrient (Ogra, 2009;
u

Presentato et al., 2019). No cases of severe Te poisoning have been recorded, though workers
Jo

at a Canadian silver refinery exposed to Te during their work tended to report experiencing

garlic odour when their Te content surpassed 1 µmol Te per mol creatinine (Berriault and

Lightfoot, 2011). The late Alan Criddle (mineralogist at the Natural History Museum,

London) contracted one of the few documented recent cases of tellurium breath following a

visit to the Au–Te deposit at Moctezuma, Mexico. “[He] breathed in dust rich in the

secondary tellurate ochres that occur there. He said his breath stank so bad that dogs would

run howling from him. He found it difficult to live with himself for a few days ” (oral

communication, Ben Grguric).


Journal Pre-proof
One of the most toxic forms of Te is the short- lived, anthropogenically produced, radioactive

isotope, 132 Te. 132 Te is a component of nuclear disaster contamination, and was the third-most

released radionuclide following the Fukushima Daiichi accident (Figure 11), with the total

released activity of 180 petabecquerel (PBq) only less than that from 131 I and 133 Xe (Gil-Díaz,

2019). 132 Te is a β-particle emitter with a half- life of 3.2 days, decaying to 132 I, which itself is

radioactive with a half- life of <3 hours and releases a β-particle to form the stable 132 Xe. The

main toxicity mechanism of 132 Te is the in corpore formation of the thyroid-critical element I,
132 132
as radioactive I. The I is subsequently transported to the thyroid, where it decays,

f
132
causing radiation damage (Drozdovitch et al., 2019). Exposure to Te is believed to have

oo
contributed to the high incidence of thyroid cancers in people directly exposed to radiation

pr
after the Fukushima Daiichi nuclear disaster (Drozdovitch et al., 2019). Additionally, 129m
Te
e-
is also produced in nuclear waste, and was detected in aerosols around Fukushima following

the meltdown (Foreman, 2015; Kanai, 2015) and may also be transported aqueously (Gil-
Pr

129m 129
Díaz et al., 2019b). Te decays via gamma radiation to Te, then follows a similar decay
132 129 129 129
Te, i.e. via two β-particles to form stable
al

path to Xe via radioactive I. However, I

has a half- life of 1.58 × 107 years, meaning 129


I and thus 129m
Te are of considerably less
rn

132 132 129m 132


concern biologically than Te and I, with Te instead finding application as a Te
u

129 129m
signaller (Tagami et al., 2013). The long half- life of I means most Te produced in
Jo

anthropogenic nuclear reactions remains either in storage or in the environment today.

Although less toxic than their soluble and bioavailable oxyanion counterparts, Se and Te

nanoparticles are themselves toxic to many microorganisms. Many Te-resistant

microorganisms actively produce Te nanoparticles from Te oxya nions as a detoxifying

mechanism (see Section 6.3). The toxicity mechanism for non-resistant microorganisms is

(1) believed to be due to reaction of the nanoparticles with intracellular thiols, producing

reactive oxygen species (ROS) causing oxidative stress, much like the oxyanions of Se and

Te (e.g. Zonaro et al., 2015), and (2) may contribute to functional damage of cell membranes
Journal Pre-proof
by changing their composition (Pi et al., 2013). An advantage of nanoparticles over

traditional antimicrobial agents is that their high surface-to-volume ratios provide a larger

area of interaction with biological systems (Zonaro et al., 2015). Selenium and Te

nanoparticles are particularly noted for their antibiofilm ability, as many conventional

antibiotics are more effective against planktonic than biofilm bacteria, a problem which Se

and Te nanoparticles readily overcome (e.g. Zonaro et al., 2015, Vaigankar et al., 2018).

Tellurium nanoparticles produced from Bacillus sp. BZ have been documented for their

antifungal properties against the fungus Candida albicans (Zare et al., 2014). Tellurium and

f
Se nanoparticles could thus have future antimicrobial and/or antifungal roles, with different

oo
morphologies to target different classes of organism (Abo Elsoud et al., 2018; Estevam et al.,

pr
2017; Tran and Webster, 2013; Vaigankar et al., 2018; Zonaro et al., 2015).
e-
A different type of toxic Te-containing nanoparticle comes in the form of CdTe quantum dots,
Pr

which are now known to be cytotoxic (Chen et al., 2012; Lovrić et al., 2005). The formation

of ROS when CdTe nanoparticles interact directly with the plasma membrane, mitochondria
al

and nucleus of cells is a key factor in their toxicity (Lovrić et al., 2005). Other toxicity factors
rn

are the release of soluble Cd2+ ions and the overall intracellular distribution of the quantum
u

dots (Chen et al., 2012). Consequently, CdTe quantum dots must either be firmly
Jo

encapsulated in protective coatings for safe usage (e.g. glutathione ; Zheng et al., 2007), and

CdTe quantum dots have sometimes been replaced by quantum dots formed from other

materials (Li et al., 2009; Pons et al., 2010).

6. BIOGEOCHEMISTRY OF TELLURIUM

Many essential processes in the biosphere are carried out by microbes, which generally

operate on quicker timescales than purely inorganic processes (Ehrlich and Newman, 2009;

Figure 12). Tellurium, unlike the other non-transient chalcogens, is not known to be an

essential biological nutrient (Ogra, 2017), but Te is occasionally found substituting for S in
Journal Pre-proof
the two sulphur-containing amino acids cysteine and methionine (Anan et al., 2013; Ramadan

et al., 1989). Additionally, some microorganisms can respire Te oxyanions (e.g. Yurkov et al.

1996; further detail provided below); this suggests that Te should possibly be included in the

list of biological trace elements. In the environment, microorganisms typically interact with

part per trillion levels of Te, given its extreme rarity – especially in surface waters (see

Section 4.1). Tellurium has more similarity to S than other common biologically active

elements – especially when present in high concentrations, where it can interact with

sulphate-specific enzymes such as sulphate reductases (Ottosson et al., 2010) and transporters

f
(Goff and Yee, 2017). More recently, Te toxicity has been explored at a more detailed

oo
biochemical level, examining the speciation of Te in different biological environments (e.g.

pr
Ogra, 2017; Turner et al., 2012). While there is still conjecture on Te microbiology as applied
e-
to natural (including mineral transformation) contexts, specific aspects of Te microbiology

are well understood, particularly its bioreduction. This has been extensively investigated by
Pr

several researchers including Taylor (1999), Zannoni et al. (2007), Turner et al. (2012) and
al

Presentato et al. (2019).


rn

Broadly, Te biogeochemistry may be divided into biooxidation, biosorption/bioaccumulation,


u

and bioreduction (Figure 12). Biooxidation typically results in species which are more
Jo

soluble and toxic than their (usually solid) precursors. Coping mechanisms for dealing with

high concentrations of Te oxyanions include biosorption, as well as bioprecipitation and

biovolatilisation, which are both forms of bioreduction.

6.1 Biooxidation (bioleaching)

Although there is no current evidence for direct environmental biooxidation of Te, recent

characterisation of ‘Se-oxidising bacteria’ and their role in biogeochemical Se cycling

(Nancharaiah and Lens, 2015) suggests that there may be ‘Te-oxidising bacteria’ in certain

Te-rich (micro)-environments. Se-oxidising bacteria are mostly chemautotrophs, often

thiobacilli. Biooxidation processes for Se are typically 3-4 times slower than reduction,
Journal Pre-proof
meaning that biologically mediated processes alone result in locally greater amounts of

reduced Se species (Nancharaiah and Lens, 2015). Se oxidisers may themselves be capable of

oxidising Te and tellurides to release Te oxyanions back to the environment, but to our

knowledge this potential process has not been studied.

Environmental oxidative processes initiated by microbes are commonly caused by ‘indirect’

mechanisms. One common way metals are released from insoluble minerals in local (micro-

)environments is that the bacteria use an ‘inert’ substrate to grow, and EPS or other biogenic

molecules produced as the colony grows slowly dissolve the surface of the

f
oo
substrate (Fairbrother et al., 2009; Reith et al., 2009; 2019). Another common pathway occurs

by the production of an oxidant in situ. Only a handful of acidophilic chemolithotrophic


pr
bacteria – all S- and/or Fe-oxidisers that produce inorganic oxidants as a by-product of their
e-
metabolism – have been studied for their oxidising behaviour applied to Te compounds, with
Pr

almost all of these studies aiming to enhance bioleaching (Choi et al., 2018; Climo et al.,

2000a; 2000b; Guo et al., 2012a; 2012b; Kim et al., 2015). Bioleaching is an attractive
al

method for extracting metals from sulphide ores as it uses gentler chemicals than traditional
rn

leaching (Khaing et al., 2019). The leaching bacteria do not directly oxidise Te, but instead

Fe-oxidising bacteria oxidise Fe2+ to Fe3+ ions. In the presence of S-oxidising bacteria, which
u
Jo

tend to produce pH- lowering H+, the Fe3+ ions are then able to oxidise tellurides, releasing the

constituent metals into solution (Climo et al., 2000b). Equation 6.1 shows the

thermodynamically favourable (log K = 25.7 at 25°C, increasing to 29.0 at 60°C) oxidation of

calaverite by (biologically produced) ferric ions:

AuTe2 (𝑠) + 8Fe3+ (𝑎𝑞) + 6H2 O ↔ Au(𝑠) + 2H3 TeO3+ (𝑎𝑞) + 6H + (𝑎𝑞) + 8Fe2+ (𝑎𝑞) (6.1)

The most common S- or Fe-oxidising bacteria are capable of Te ore bioleaching, as discussed

in the following papers:

 Acidothiobacillus ferrooxidans (Choi et al., 2018; Kim et al., 2015)


Journal Pre-proof
 Leptospirillum ferrooxidans (Climo et al., 2000b)

 Thiobacillus ferrooxidans (Climo et al., 2000b; Guo et al., 2012b), T. thiooxidans, T.

caldus (Climo et al., 2000b)

In conclusion, in natural environments, Te and tellurides may be oxidised either inorganically

(e.g. in the presence of an oxidant like Fe 3+) or organically (by as yet-uncharacterised Te

oxidisers). Currently, inorganic oxidation of tellurides is the best-defined way of releasing Te

to the environment, but the oxidants may be produced by biological means; Fe and S

oxidisers are both capable of leaching Te from tellurides in anthropogenic contexts. To date,

f
oo
no microorganism has been described as purposefully dissolving tellurides or metallic Te via

a direct biooxidative process (Bonificio and Clarke, 2014).

6.2 Biosorption
pr
e-
Biosorption of Te is typically a precursor to bioreductive or bioaccumulative processes, as
Pr

microorganisms remove (soluble) Te from the environment in a process which begins with

the interaction of EPS with Te oxyanions (Figure 12). The biosorption mechanisms of soluble
al

Te oxyanions are not fully understood. One pathway for E. coli involves the tellurate anion
rn

entering E. coli cells via the SulT-type sulphate transporter CysPUWA (Goff and Yee, 2017).
u

Removing the CysPUWA transporter in mutant strains of E. coli resulted in the accumulation
Jo

of less cellular Te, and higher resistance to Te compared to the wild-type strain (Goff and

Yee, 2017). The tellurite anion appears to enter cells via different transporters, e.g. phosphate

transporters move tellurite into E. coli (Borghese et al., 2016a). Tellurite enters Rhodobacter

capsulatus cells via acetate permease RcActP2, and expressing RcActP2 of R. capsulatus

results in a fourfold increase in the rate of tellurite uptake in E. coli (Borghese et al., 2016a).

In case of fatal Te poisoning, Te may be bound in the dead microorganism’s biomass,

preventing it from interacting with other microorganisms. There is little prior literature

examining the biosorption of Te as a biochemical process, other than acknowledging that it is


Journal Pre-proof
a precursor to bioreduction (Section 6.3) and bioaccumulation (Section 6.4). More commonly,

biosorption (followed by bioaccumulation) is simply discussed as a method of bioremediation

of areas Te-contaminated areas (Piacenza et al., 2017; see Section 6.5).

6.3 Bioreduction

Tellurium bioreduction comes in two main forms, namely bioprecipitation and

biovolatilisation, with the former being the more common Te detoxification mechanism.

Bioprecipitation of Te oxyanions to metallic Te was first noticed over a century ago ; indeed,

the black precipitate produced remains in use to test for the presence of certain Te-resistant

f
oo
strains (Corper, 1915; King and Davis, 1914). Biovolatilisation of Te oxyanions was also first

noticed in the 19th century due to the potent smell of volatile Te compounds. The conversion
pr
of Te oxyanions to alkylated, gaseous Te compounds is mediated by S-adenosyl- methionine
e-
(SAM), as described by the Challenger mechanism (Basnayake et al., 2001). Early
Pr

experiments suggesting that S-adenosyl methionine (SAM) would bind to a telluro- methylase

enzyme (TehB) as part of the volatilisation process (Liu et al., 2000; Presentato et al., 2019)
al

were verified by the crystal-structure determination of TehB isolated from E. coli


rn

(Choudhury et al., 2011). Simplified versions of the two main bioreduction types are given
u

here using the soluble tellurite anion as the starting material:


Jo

𝐁𝐢𝐨𝐩𝐫𝐞𝐜𝐢𝐩𝐢𝐭𝐚𝐭𝐢𝐨𝐧 𝐨𝐯𝐞𝐫𝐚𝐥𝐥 (𝐮𝐧𝐛𝐚𝐥𝐚𝐧𝐜𝐞𝐝): 𝐓𝐞𝐎𝟑𝟐− (𝒂𝒒) ↔ 𝐓𝐞𝟎 (𝒔) (𝟔. 𝟐)

𝐁𝐢𝐨𝐯𝐨𝐥𝐚𝐭𝐢𝐥𝐢𝐬𝐚𝐭𝐢𝐨𝐧 𝐨𝐯𝐞𝐫𝐚𝐥𝐥 (𝐮𝐧𝐛𝐚𝐥𝐚𝐧𝐜𝐞𝐝): 𝐓𝐞𝐎𝟑𝟐−(𝒂𝒒) ↔ (𝐂𝐇𝟑 )𝟐𝐓𝐞(𝒈) (𝟔. 𝟑)

The importance of determining the mechanism of tellurite toxicity is growing as

anthropogenic uses of Te increase (Ottosson et al., 2010); as a result, Te bioreduction has

entered a scientific ‘Renaissance period’ over the past decade (Goff and Yee, 2017;

Presentato et al., 2019; Turner et al., 2012). Figure 12 shows our summary of the currently

known and postulated biogeochemical processes for Te.


Journal Pre-proof
6.3.1 Mechanism of Te bioprecipitation

Bioprecipitation provides a method for removing toxic soluble Te oxyanions from the

environment, depositing Te as (relatively) inert nanoparticles (Figure 12). The environmental

fate of these nanoparticles is not well understood, with a paucity of studies on Te

nanoparticles in the environment. Tellurium-resistant microorganisms are in general not

adversely affected by the nanoparticles they produce, even though the nanoparticles

themselves are toxic for many microorganisms (Presentato et al., 2019). Bioprecipitation can

occur as a result of microorganisms respiring Te oxyanions in anaerobic respiration processes:

f
oo
some microorganisms are able to use either the tellurite or tellurate anions as terminal

electron acceptors. Under anoxic conditions, both Bacillus selenitireducens and


pr
Sulphurospirillum barnesii are capable of respiring the Te oxyanions anaerobically, turning a
e-
fatal concentration of Te oxyanions for many microorganisms into an evolutionary advantage
Pr

(Baesman et al., 2007).

Microorganisms capable of reducing the oxyanions of Te occur in a variety of extreme


al

environments, and a selection of the most important are listed in Table 5. A community of
rn

123 separate Te-resistant bacteria was isolated from the Antarctic peninsula (Arenas et al.,
u

2014), where the resistance to Te oxyanions is thought to be endowed by cross-resistance to


Jo

other ROS generators like UV light. Other Te-resistant microorganisms were isolated from

the extreme high pressure and temperature environment of the Juan de Fuca Ridge black

smoker field, in the depths of the Pacific Ocean (Maltman et al., 2016; 2017). Tellurium and

selenium reducing bacteria are thought to have a symbiotic relationship with the tube worms

which they associate with near the Juan de Fuca Ridge black smoker field (Maltman et al.,

2016; 2017). Vent chimneys are a rich natural Te source (see Table 3), therefore

microorganisms from these environments are more likely to be Te resistant (Bonificio and

Clarke, 2014). Similar communities of Te oxyanion-reducing bacteria occur in a variety of

anthropogenically contaminated sites (Chien and Han, 2009; Kagami et al., 2012; Qin et al.,
Journal Pre-proof
2017; Zonaro et al., 2017; Table 5). Highly saline lakes or springs like California’s Mono

Lake (Figure 12; Baesman et al., 2006; 2007; 2009) or Iran’s Neidasht spring (Etezad et al.,

2009; Soudi et al., 2009) also harbour Te-resistant organisms. The diversity of these localities

shows that bacterial communities with varying degrees of Te resistance (for instance,

different rates of reduction and minimum inhibitory concentrations) exist in a large variety of

terrestrial environments.

The majority of microorganisms capable of reducing Te oxyanions are Gram- negative

bacteria, along with some alpha-Proteobacteria, Gram-positive bacteria (Presentato et al.,

f
oo
2016) and fungi (Abo Elsoud et al., 2018; Gharieb et al., 1999). Interestingly, as well as

containing more representative species which can reduce Te oxyanions, Gram- negative
pr
bacteria are in general more susceptible to Te oxyanion (especially tellurite) toxicity (Castro
e-
et al., 2008), as the peptidoglycan-based cell wall of Gram-positive bacteria is better able to
Pr

prevent Te oxyanions from entering the cells.

The Gram- negative Te-sensitive E. coli was used in studies of how non-Te resistant
al

microorganisms cope with high levels of soluble Te (e.g. Wang et al., 2011). Some strains of
rn

E. coli have higher levels of resistance to Te oxyanions than others, and genes encoding
u

resistance in other microorganisms have been expressed successfully in E. coli – giving the
Jo

modified E. coli greater resistance to Te oxyanions (e.g. Castro et al., 2009). The effects of

elevated tellurite in E. coli include decreased dNADH/NADH (Nicotinamide Adenine

Dinucleotide) dehydrogenase activity, alteration of oxidases, augmented lipid peroxidation,

and Fe–S centre dismantling (Díaz-Vásquez et al., 2015). Exposure to elevated levels of Te

oxyanions for non-resistant organisms usually proves fatal due to the generation of ROS and

to an unbalancing of the thiol:redox cell buffering system, (e.g. Arenas et al., 2014; Díaz-

Vásquez et al., 2015). One method of reducing tellurite toxicity is by exposure to the less

toxic selenite, which protects E. coli from tellurite damage in certain situations (Vrionis et al.,

2015). Microorganisms which have been conditioned (i.e. pre-exposed to low level
Journal Pre-proof
concentrations of Te oxyanions) also tend to survive better when exposed to larger levels of

tellurite (Presentato et al., 2016).

Like E. coli, other microorganisms also use a NADH-dependent pathway to reduce tellurite

e.g. Salinicoccus iranensis (Alavi et al., 2014). Some Te-resistant microorganisms may have

specific tellurite reductases such as the E3 (dihydrolipoamide dehydrogenase) component of

a pyruvate dehydrogenase of Aeromonas Caviae (Castro et al., 2008; 2009), the zinc- (Zn)

and molybdenum- (Mo) containing enzyme of B. sp. STG-83 (Etezad et al., 2009) and the

two distinct enzymes responsible for Te and Se oxyanion reduction in the ER-Te-48 strain of

f
oo
bacteria isolated from the Juan de Fuca Ridge black smokers (Maltman et al., 2017). The

exact structure of specific Te-reducing enzymes is not fully understood. One tellurate
pr
reductase enzyme in E. coli is likely to contain molybdopterin, as mutants with genes deleted
e-
in the molybdopterin synthesis pathway were unable to reduce tellurate (Theisen et al., 2013)
Pr

– potentially this molybdopterin is the Mo-containing unit postulated by Etezad et al. (2009)

in B. sp. STG-83. Resistance to Te and other metals may also be associated with antibiotic
al

resistance due to co-resistance (i.e. genetic linkage) between the resistance genes (Argudín et
rn

al., 2018).
u

Unsurprisingly, Te oxyanion-reducing bacteria operate best when in the presence of a reliable


Jo

source of organic carbon, typically also an electron donor which facilitates bioreductive

processes (e.g. acetate, lactate and pyruvate). The presence of redox mediator molecules

(generally quinone-based) typically results in greater efficiency of Te oxyanion reduction as

they can quickly transport electrons from cells to oxidised compounds (Ramos-Ruiz et al.,

2016; Van der Zee and Cervantes, 2009). The presence of other anions or molecules may also

alter the rates of Te oxyanion reduction. The pathways for reduction of selenite and tellurite

are typically rather different. Microorganisms that were studied for Te or Se oxyanion

reduction activity alone are expected to also reduce the other, but with different efficiencies

and maximum allowable concentrations of the oxyanions – as shown by comparative studies


Journal Pre-proof
(e.g. Etezad et al., 2009). The reduction rates of tellurite and tellurate anions are generally

also different for this reason. Tellurite is typically reduced faster, as the transition to

elemental Te requires two less electrons per Te atom than the reduction of tellurate. For

instance, the tellurite anion reduction rate by the methanogenic microbial consortium studied

by Ramos-Ruiz et al., (2016) is seven times faster than the rate for tellurate: the reduction of

tellurate to tellurite is the rate-limiting step.

The presence of selenite increases tellurite reduction 13- fold in some bacterial cultures (Bajaj

and Winter, 2014) due to stimulation of tellurite reduction by parallel reduction with selenite.

f
oo
Conversely, a mixture of selenite and tellurite has a more prohibitive effect on growth of the

white-rot fungus Phanerochaete chrysosporiumhan than either anion alone, indicating that
pr
reduction mechanisms are organism-specific (Espinosa-Ortiz et al., 2017). In Shewanella
e-
oneidensis MR-1, elemental chalcogen precipitates are produced extracellularly for Se, and
Pr

intracellularly for Te (Klonowska et al., 2005).

Some fungi and yeasts are also capable of bioreducing Te oxyanions. Gharieb et al. (1999)
al

found that a species of Fusarium reduced tellurite to elemental Te, and in the same study
rn

noted that Penicillium citrinum reduced tellurite to both elemental Te nanoparticles and also
u

to a volatile Te species (see Section 6.3.3) The previously mentioned white-rot fungus, P.
Jo

chrysosporium, is capable of reducing both Se and Te oxyanions (Espinosa-Ortiz et al., 2017).

Abo Elsoud et al. (2018) analysed six different fungal isolates, and found that all were

capable of producing elemental Te, and new Te-reducing fungi are regularly identified (Liang

et al., 2019). It is likely that many more fungi will reduce Te oxyanions, particularly in

telluriferous areas. Some fungi are particularly noted for their biovolatilisation (see Section

6.3.3).

6.3.2 Morphology of biogenic Te nanoparticles


Journal Pre-proof
The morphology of precipitates formed via microbial reduction of Te oxyanions varies

considerably (Figure 14), and includes both intracellular and extracellular Te nanoparticles,

typically as rods or spheres (e.g. Baesman et al., 2009; Sepahei and Rashetnia, 2009;

Figure 14a). The mediating factors controlling whether the Te is produced intracellularly

and/or extracellularly seem to be organism-dependent, although in general the particles are

produced near the cytoplasm (White et al., 1995). Kagami et al. (2012) found that elemental

Te produced extracellularly can migrate away from the parent cell once produced. The size of

the nanoparticles often exceeds 100 nm (Jain et al., 2014)#3 . Redox mediators tend to promote

f
the production of extracellular Te nanostructures (Borghese et al., 2016b; Ramos-Ruiz et al.,

oo
2016; Wang et al., 2011). Most studies describe Te oxyanion reduction under anoxic

pr
conditions, where the reduction of an alternative oxidant to oxygen is preferable; however,
e-
Presentato et al. (2016) describes the reliable synthesis of Te nanorods by an aerobic

mechanism (Figure 14b). Bioprecipitated compounds aside from elemental Te have only
Pr

rarely been reported, for instance the extracellular Se–Te composite nanoparticles formed by
al

reduction of a mixture of selenite and tellurite (Bajaj and Winter, 2014; Figure 13c).
rn

6.3.3 Mechanism of Te biovolatilisation


u

Alkylation of p block elements (in particular methylation) is a common method by which


Jo

cells metabolise metals (Fatoki, 1997; Frankenberger Jr, 1993), and is a key method of

releasing Te to the atmosphere (Figure 5). Biovolatilisation typically accounts for less than 5%

of total Te oxyanion conversion, with the remainder of the Te bioprecipitated (Figure 12;

Ollivier et al., 2008), i.e. there are no specific Te biovolatilisers. Biovolatilisation produces

volatile, smelly gases containing covalent Te–C or Te–S bonds, the simplest of which is

dimethyl telluride [DMTe; (CH3 )2 Te]. Dimethyl ditelluride (DMDTe; CH3 TeTeCH3 ) is also

common, while dimethyltellurenyl sulphide is rarer (DMTeS; CH3 TeSCH3 ) (see Figure 12;

Ollivier et al., 2008; Zonaro et al., 2015). Organosulphur compounds like dimethyl sulphide

are also produced during the biovolatilisation processes producing organotellurium


Journal Pre-proof
compounds (Ollivier et al., 2008). Biovolatilisation thus results in the release of Te to the

atmosphere. This process may occur in areas naturally rich in Te, or at contaminated

anthropogenic sites.

Specific organisms which display natural volatilising behaviour, listed by alphabetical order

of genus then by species, are listed in Table 2 (DMTe, DMDTe and DMTeS in second

column from right). Penicillium molds are some of the only fungi studied for their

biovolatilising behaviour (White et al., 1995), with Penicillium citrinum biovolatilising as

well as bioprecipitating Te (Gharieb et al., 1999). One unexpected source of Se and Te

f
oo
volatiles (along with volatile compounds of five other heavy metals) in certain local areas is

duck manure compost, the microbes in which proved capable of methylating and/or
pr
ethylating Se and Te (Pinel-Raffaitin et al., 2008).
e-
It is unclear whether biovolatilisation is truly a detoxification measure for microorganisms
Pr

which convert soluble Te to volatile forms, as unlike elemental Te, volatile organotellurium

compounds are not necessarily less toxic than their soluble Te oxyanion counterparts (White
al

et al., 1995). Nonetheless, even if the toxicity is similar, volatile compounds do have more
rn

chance of migrating away from the microbes, suggesting that this mechanism is overall a
u

detoxification measure (Basnayake et al., 2001). In highly Te-rich areas, the smell of onions
Jo

or garlic may be detected in the air from the action of bioreducing microbes.

6.3.4 Other mechanisms of Te detoxification

Bioreduction is the most common method of Te detoxification, but it is not the only method

by which microorganisms process high levels of soluble Te. Some bacteria (e.g.

Erythromicrobium ezovicum and Roseococcus thiosulphatophilus) have been found to

process Te oxyanions without using a reduction pathway, depending on the source of organic

carbon available (Yurkov et al., 1996). These alternative methods of Te detoxification

include tellurite efflux or complexing (Sepahei and Rashetnia, 2009; Yurkov et al., 1996),
Journal Pre-proof
with complexing of Te oxyanions being a possible precursor to their bioaccumulation. The

existence of microorganisms that resist high levels of Te oxyanions without using

bioreductive mechanisms adds a further layer of complexity to the biogeochemistry of Te

(Turner, 2001).

6.4 Bioaccumulation

Bioaccumulation begins for Te at the smallest biological level as Te-resistant microbes

absorb soluble Te oxyanions and convert them to elemental Te by bioreduction. Elemental Te

may accumulate in Te-resistant cells as Te nanoparticles or as Te bound in proteins.

f
oo
Tellurium nanoparticles are typically produced near cell boundaries (Turner et al., 2012), and

pr
some are released extracellularly (e.g. Bajaj and Winter, 2014; Borghese et al., 2016b) and

may act as toxins for other microorganisms – in a different fashion to the toxicity mechanism
e-
for soluble Te oxyanions (e.g. Abo Elsoud et al., 2018). In general, heavy metals and other
Pr

toxins are concentrated up the food chain; plants and animals have higher levels of Te than

the surrounding environment.


al
rn

Some onion and garlic family (Allium) plants owe their distinctive smell to organosulphur

compounds – and to lesser extents to organoselenium and organotellurium compounds.


u
Jo

Although Te conversion in these species is not an essential feature of their metabolism, plants

like Allium sativum are capable of incorporating Te into two usually S-containing amino

acids, cysteine and methionine (Anan et al., 2013), with some plants in the Allium family

containing extreme levels of Te enrichment up to 300 mg/kg (Dunn, 2011). However in

general, Te levels in plants do not exceed 1 mg/kg, even in telluriferous soils (Cowgill, 1988),

and commonly plants contain less than 0.02 mg/kg Te (Dunn, 2011). The flowers of plants

generally contain the most Te, and tree leaves contain more Te than branches (Cowgill, 1988;

Dunn, 2011).
Journal Pre-proof
Bioaccumulation of Te provides one potential sink for Te in the natural environment. The

rarity of Te means that natural bioaccumulation only occurs to a significant extent in areas

with naturally or anthropogenically raised levels of Te. As a relatively reactive element, Te is

unlikely to remain bioaccumulated for long periods of time, especially as biogenically

produced nanoparticles (see below).

6.5 Applications

All of the common biotransformations of Te have some practical uses, which are briefly

discussed in this section. Biooxidation of gold tellurides has become a relatively common

f
oo
pre-treatment step in gold mining. Gold tellurides host a significant amount of the gold grade

in many deposits, but traditional leaching using cyanide, thiourea, or ammoniacal thiosulfate
pr
solutions are not efficient for gold tellurides; thus pre-treatment stages have been introduced
e-
prior to leaching, the most commonly used being energy intensive high temperature roasting
Pr

(Zhao et al., 2010). Adding a culture of oxidising microbes to Au-telluride ore piles may

allow release of higher amounts of metals than would otherwise be possible (Climo et al.,
al

2000a; 2000b), while at the opposite end of the process, adding a culture of reducing
rn

microbes to waste streams could allow recovery of toxic elements, preventing them from
u

becoming pollutants (Ramos-Ruiz et al., 2017a).


Jo

Biosorption of a variety of heavy metals has been discussed as an environmentally friendly

method for their remediation (Alluri et al., 2007; Banerjee et al., 2018; Dhankhar and Hooda,

2011; Gavrilescu, 2004; Piacenza et al., 2017; White et al., 1995). Using either living or dead

biomass to entrap metals from polluted streams is particularly viable for large-scale low-level

contamination, when sorption using more toxic chemicals is less effective (Gavrilescu, 2004).

Currently, such techniques have not been implemented for Te sorption on an industrial level

However, the most likely area of future application involves the biosorption of Te (and Cd)

released from decommissioned CdTe solar panels (Alluri et al., 2007; Dhankhar and Hooda,

2011; Rajwade and Paknikar, 2003). Biosorption may one day be used for both remediation
Journal Pre-proof
of a contaminated area and recycling, by collecting the sorbed nanoparticles and repurposing

them for other uses (Dhankhar and Hooda, 2011; Piacenza et al., 2017). Two studies

published near-concurrently in 2017 both outlined a preliminary design for an Upflow

Anaerobic (granular) Sludge Bed (UASB) for Te biosorption. Both papers theorised that their

model reactor beds could be scaled up to recycle Te-containing wastewaters as nanoparticles

of elemental Te (Mal et al., 2017; Ramos-Ruiz et al., 2017a). The UASBs provide a double

benefit of cleaning the wastewater and also potentially result in a recycled supply of Te. Mal

et al. (2017) found that their model reactor was able to remove 90% of the tellurite from

f
20 mg/L tellurite solutions, with most of this Te associating with the granular sludge as

oo
elemental Te. 78% of this Te was reclaimed from the sludge by an EPS method. Ramos-Ruiz

pr
et al. (2017) obtained comparable percentages (83-96%) of continuous removal of Te from
e-
20 mg/L tellurite solutions without a redox mediator to assist in the electron transfer, and

greater than 99.5% efficiency with the redox mediator (riboflavin, better known as vitamin
Pr

B2). Similar UASBs have also been designed and described for Se alone (e.g. Jain et al.,
al

2016). More recently, Wadgaonkar et al. (2018a) described a UASB capable of bioreducing

selenite and tellurite simultaneously. Biosorption – and often subsequently, bioprecipitation –


rn

thus forms an intermediate step in the removal of soluble Te from the environment, and on a
u

larger scale may gain prominence in future as a method of removing and recycling Te from
Jo

waste materials (Nancharaiah et al., 2016).

Tellurium and Se nanoparticles can be produced via bioreduction in morphologies that can be

difficult to obtain via inorganic means (Bansal et al., 2012; Zonaro et al., 2017). Biosynthesis

also has environmental advantages, typically requiring lower amounts of toxic chemicals than

alternative methods (Presentato et al., 2016). As well as antimicrobial applications (Section

5.4), the nanoparticles may potentially be used in a variety of highly accurate detecting

applications, such as for detecting hydrogen peroxide (Manikandan et al., 2017; Wang et al.,

2010), chlorine gas (Sen et al., 2009) and Hg2+ cations (Wei et al., 2011). Further applications
Journal Pre-proof
of Se and Te nanoparticles make use of their ability to photocatalytically degrade organic

pollutants such as methylene blue in the presence of sunlight (Vaigankar et al., 2018). Overall,

these applications already show the versatility of Te biotransformations, and we predict that

future study of Te biogeochemistry will lead to many more.

7. A TELLURIUM BIOGEOCHEMICAL CYCLING MODEL

7.1 An integrated Te cycling model – comparison with Se

An integrated Te cycling model is presented in Figure 5, linking inorganic and organic

f
processes. Our Te cycle is based on those known for other elements, especially Se, and on

oo
our current understanding of the geochemistry and microbial ecology of Te. Elemental

pr
cycling has long been known for biologically essential elements, most notably the carbon (C),
e-
nitrogen (N) and sulphur (S) cycles. Even elements which were long assumed to be

geochemically inert are involved in complex biogeochemical cycling (Stolz, 2017), including
Pr

Au and Pt (Reith et al., 2014; 2007). Te is the last non-transient chalcogen (any attempts to
al

define such a cycle for polonium (Po) would be essentially meaningless) without a detailed

biogeochemical cycling description, following the description of the Se cycle by Nancharaiah


rn

and Lens (2015).


u
Jo

The Te cycle begins with the transformation of Te-rich crustal rock through either

magmatism or hydrothermal processes (Figure 5), leading to the formation of Te-rich rocks

and deposits (Section 4). The ore from these deposits may be processed by mining and

smelting, most likely for the Au or Cu content rather than for recovery of Te itself, or the Te

may be released to the environment by weathering processes, especially if anthropogenic

activities have exposed more of the ore (Figure 5). Weathering and emission of Te to the

environment during mining, usage and recycling/breakdown may lead to solid, aqueous or

gaseous forms of Te, depending on the processes involved. Solid forms include Te sorbed

onto clays and Fe oxides in soils, oxidative dissolution leads to aqueous Te oxyanions, and
Journal Pre-proof
volatilisation (mostly biological, except at high temperatures during earlier smelting or

mining processing steps) leads to gaseous forms of Te (Figure 5). Deposition into sediments

and dissolution into groundwater result in higher Te concentrations than occur in freshwater

and ocean water (Figure 5). Te also sparingly enters biomass through microorganisms,

though as Te is nonessential for life, its intake is incidental compared to the minute but

essential amounts of Se that all organisms must absorb. The breakdown of these organisms

over millions of years leads to the presence of Se and Te in coal at levels greater than their

crustal abundances, and burning coal is one mode of Te and Se release to the atmosphere

f
oo
(Figure 5).

The bacterium in Figure 12 shows the biogeochemical processes present for Te in full.
pr
Nancharaiah and Lens (2015) postulate that Se is cycled between oxic and anoxic
e-
environments in natural environments, and influences carbon and nitrogen mineralisation
Pr

processes through bacterial anaerobic respiration. As is the case for Te, Se bioreduction

converts Se from soluble Se oxyanions to elemental Se and volatile organo-selenium


al

compounds. Selenium oxidisers provide one method of release of Se from elemental Se and
rn

selenides, allowing solubilised Se to again travel in waterways until a reduction source


u

(selenium resistant microbes and soluble inorganic cations which precipitate Se oxyanions) is
Jo

encountered, whereas Te oxidisers are currently only postulated to exist. The relation

between the Te and Se cycles may be even closer, with some microorganisms capable of

acting in similar ways on both Se and Te species – for instance Se and Te reducers both

produce either elemental chalcogen nanoparticles or volatile organochalcogens (Wallschläger

and Feldmann, 2010). In summary, the processes governing the Te biogeochemical cycle are

clear, but the extent to which microbially-catalysed pathways for reduction, oxidation or

other processes predominate in surficial environments is not yet known, and will form the

core of future work on the environmental geochemistry of Te.


Journal Pre-proof
7.2 Tellurium dispersion around Au-Te deposits

In Figure 15, we apply our cycling model to the example of elemental dispersion of Te

around a Au–Te deposit. We have chosen to discuss this example of Te cycling, as (1) Te and

Au are typically found together, in association with metal sulphides; (2) the Au cycle is

already well understood (Rea et al., 2018; Sanyal et al., 2019; Shuster and Reith, 2018),

(3) Te may be used as a pathfinder for Au through application of their linked biogeochemical

cycles, and (4) this is an important analogue for predicting the fate of anthropogenic Te

contamination.

f
oo
This model describes the movement of Te and Au through the environment around a buried

Au–Te hydrothermal vein. Gold and Te occur as gold tellurides, along with small Au nuggets,
pr
‘invisible Au’ associated with other primary minerals, and elemental Te (Sung et al., 2009).
e-
Once these minerals reach the surface and interact with groundwater within the saprolith,
Pr

they may begin to undergo dissolution reactions, in particular inorganic acidic weathering

(see Section 5.1). Passive oxidation by different classes of microorganisms may also have an
al

effect on the release of Au and Te to the soil. As soluble species, Au and Te in surface
rn

aqueous solutions are toxic to macro- and microorganisms in sufficiently high concentrations.
u

Once solubilised, reprecipitation and dissolution reactions occur in a cyclical fashion, and
Jo

other chemical species such as Cu2+, Pb2+, Fe3+, Cl- interact with soluble forms of Te.

Tellurium released into natural waters as tellurite may undergo further oxidation to tellurate.

As mobile Te oxyanions, Te may be transported away from the weather ing zone of the Te

deposit by groundwater. Eventually, Te may be removed from environmental aqueous

solution by three major methods:

(1) Purely inorganic methods (precipitation by a cation, possibly leading to secondary

mineralisation). This scenario is especially prominent if pH levels increase towards neutral

further from weathering sites, lowering the solubility of soluble Te oxyanions.


Journal Pre-proof
(2) Microbial bioreduction (formation of Te nanoparticles or gaseous organotellurium

compounds, especially in areas where the microorganisms display Te resistance). Some

microorganisms have been shown to successfully detoxify both Se and Te oxyanions,

suggesting that this is one area where the Se and Te cycles may overlap (Bajaj and Winter,

2014; Espinosa-Ortiz et al., 2017). The plethora of studies on how microorganisms interact

with soluble Te in controlled laboratory settings gives us an insight into how such

microorganisms might interact with soluble Te in natural environments. Bioprecipitated Te

nanoparticles are relatively reactive, although if they aggregate together the rate of reaction is

f
slowed. The elemental Te nanoparticles produced by bioprecipitation may remain insoluble,

oo
or depending on conditions could be oxidised again to soluble Te species. Alkylated forms of

pr
Te produced by biovolatilisation, although produced as minority products in bioreductive
e-
processes, can travel as airborne material (although organotellurium compounds are denser

than air, meaning these compounds are unlikely to travel long distances), spreading Te away
Pr

from natural or anthropogenic Te rich areas. Other methods of Te release to the atmosphere
al

include anthropogenic methods (e.g. burning Te-containing coal) and the release of Te-

containing volcanic gases.


u rn

(3) Removal of Te and Au from aqueous solutions by macrobiota, although in sufficient


Jo

amounts this may result in health problems for the new host organisms. This effect is most

prominent in metal-rich haloes around ore deposits.

Eventually Te returns to the regolith, either as atmospheric organotellurium compounds

interacting with soils to form aqueous or solid forms of Te, or by release of Te and Au from

their hosts. Microbes and fungi are likely to be involved in these conversion steps, from

where Te may continue its biogeochemical cycle. Thus, through a variety of inorganic and

biological processes, Te may be cycled through the environment, with these cycles especially

active near locally Te-rich areas (e.g. Au mining sites) or anthropogenically contaminated

areas (e.g. Cu refining plants or sulphuric acid factories). We expect that future advances in
Journal Pre-proof
this area will lead to a more complete description of the biogeochemical Te cycle, including

the percentage split between inorganic and biological pathways for the same processes, thus

providing valuable insight into the environmental mobility of a rare, yet important element.

8. CONCLUSIONS

Tellurium biogeochemistry is a fascinating and under-studied area of research.

Tellurium is regularly listed as a critical raw material both due to its increased use in the solar

industry, and to the dependence on other commodities in its supply chain. As Te usage

f
increases and our exposure to Te correspondingly rises, a thorough understanding of the

oo
geo(bio)chemistry of Te in surface environments is fundamental for supporting the search for

pr
future sources of Te (geochemical exploration); developing inno vative processing techniques

for extracting Te; and quantifying the environmental risks associated with new anthropogenic
e-
uses. The present work links existing research in inorganic Te geochemistry and mineralogy
Pr

with the bio(geo)chemical and biological literature towards developing an integrated Te

cycling model.
al
rn

A key challenge to understanding the dynamics of Te mobility in the environments in

the quantification of the role of microbes. Although microbes capable of bioreduction of Te


u
Jo

compounds have been well-studied in laboratory contexts and can be used in bioremediation,

their role in environmental systems currently remains poorly quantified. In particular, it is

noticeable that metallic Te nanoparticles have not been reported from the environment to date.

Similarly, biooxidation of Te compounds has only been observed occurring indirectly, yet it

is likely that direct Te oxidisers exist but have not yet been effectively characterised. Since

Te in present in reduced, insoluble forms in ores, its oxidatio n is the first step in its liberation.

Coupled to both of these processes is the need to better understand the biosorption and

bioaccumulation of Te, in particular, the fate of Te nanoparticles (presumably produced by

Te bioreduction) in the environment. Quantification of these fundamental processes is


Journal Pre-proof
required to decipher the interplay between anthropogenic, geochemical and biogeochemical

processes and its role on the distribution and mobility of Te in surface environments.

ACKNOWLEDGEMENTS

The authors acknowledge support funding provided to OPM by an Australian

Government Research Training Program (RTP) Scholarship, a Monash Graduate Excellence

Scholarship (MGES) and a Monash-Museums Victoria Scholarship (Robert Blackwood). The

authors further acknowledge The Ian Potter Foundation grant ‘tracking tellurium’ to SJM.

f
Further funding support was provided Natural Environment Research Council (UK) grant

oo
NE/M010848/1 in the Security of Supply of Minerals programme to DJS. The authors would

pr
like to acknowledge the ARC Research Hub on Australian Copper-Uranium (project number:

IH130200033), funded by the Australian Research Council, BHP Olympic Dam and the
e-
South Australian Department of State Development; for their support and assistance. We are
Pr

grateful to Brent Thorne (Figure 1), Georges Favreau (Figure 2b), Victor Okrugin (Figure 9a)

and Stefan Ansermet (Figure 10c) for providing images.


al
rn

FOOTNOTES
u

#1
The only exception to this rule is goldfieldite, Cu10 Te4 S13 , which has Tellurium (IV)
Jo

surrounded by three S atoms.

#2
Note that all numbers of total minerals in this section include overlaps as some Te minerals

have complex chemical compositions. Montbrayite, for instance, has the recently revised

formula [(Au,Ag,Sb,Pb,Bi)23 (Te,Sb,Pb,Bi)38 ; Bindi et al., 2018].

#3
When studying applications for biogenic Te nanoparticles, dimensions less than 30 nm can

be desirable due to the large increases in surface area achievable with a given mass of

nanoparticles, compared to, for example, the same mass of nanoparticles which exceed 100

nm in at least one dimension.


Journal Pre-proof

f
oo
pr
e-
Pr
al
rn
u
Jo
Journal Pre-proof
REFERENCES

Abo Elsoud, M.M., Al-Hagar, O.E.A., Abdelkhalek, E.S. and Sidkey, N.M., 2018. Synthesis
and investigations on tellurium myconanoparticles. Biotechnology Reports, 18:
e00247.
Afifi, A.M., Kelly, W.C. and Essene, E.J., 1988. Phase-relations among tellurides, sulfides,
and oxides .1. Thermochemical data and calculated equilibria. Economic Geology, 83:
377-394.
Alavi, S., Amoozegar, M.A. and Khajeh, K., 2014. Enzyme(s) responsible for tellurite
reducing activity in a moderately halophilic bacterium, Salinicoccus iranensis strain
QW6. Extremophiles, 18(6): 953-961.
Alluri, H.K., Ronda, S.R., Settalluri, V.S., Bondili, J.S., Suryanarayana, V. and
Venkateshwar, P., 2007. Biosorption: An eco- friendly alternative for heavy metal
removal. African Journal of Biotechnology, 6(25): 2924-2931.
Altree-Williams, A., Pring, A., Ngothai, Y. and Brugger, J., 2015. Textural and
compositional complexities resulting from coupled dissolution–reprecipitation

f
oo
reactions in geomaterials. Earth-Science Reviews, 150: 628-651.
Amatya, R. and Ram, R.J., 2012. Trend for thermoelectric materials and their earth
abundance. Journal of Electronic Materials, 41(6): 1011-1019.
pr
Amoozegar, M.A., Ashengroph, M., Malekzadeh, F., Razavi, M.R., Naddaf, S. and Kabiri,
M., 2008. Isolation and initial characterization of the tellurite reducing moderately
halophilic bacterium, Salinicoccus sp. strain QW6. Microbiological Research, 163(4):
e-
456-465.
Anan, Y., Yoshida, M., Hasegawa, S., Katai, R., Tokumoto, M., Ouerdane, L., Łobiński, R.
and Ogra, Y., 2013. Speciation and identification of tellurium-containing metabolites
Pr

in garlic, Allium sativum. Metallomics, 5(9): 1215-1224.


Anders, E. and Grevesse, N., 1989. Abundances of the elements: Meteoritic and solar.
Geochimica et Cosmochimica Acta, 53(1): 197-214.
al

Anderson, C.S., 2019. Tellurium. USGS Mineral Commodity Summaries, https://prd-wret.s3-


us-west-2.amazonaws.com/assets/palladium/production/atoms/files/mcs-2019-
rn

tellu.pdf.
Andreeva, E.D., Matsueda, H., Okrugin, V.M., Takahashi, R. and Ono, S., 2013. Au–Ag–Te
Mineralization of the Low ‐ Sulfidation Epithermal Aginskoe Deposit, Central
u

Kamchatka, Russia. Resource Geology, 63(4): 337-349.


Jo

Arenas, F.A., Pugin, B., Henríquez, N.A., Arenas-Salinas, M.A., Díaz-Vásquez, W.A., Pozo,
M.F., Muñoz, C.M., Chasteen, T.G., Pérez-Donoso, J.M. and Vásquez, C.C., 2014.
Isolation, identification and characterization of highly tellurite-resistant, tellurite-
reducing bacteria from Antarctica. Polar Science, 8(1): 40-52.
Argudín, M., Hoefer, A. and Butaye, P., 2018. Heavy metal resistance in bacteria from
animals. Research in Veterinary Science, 122: 132-147.
Armstrong, J.G., Parnell, J., Bullock, L.A., Perez, M., Boyce, A.J. and Feldmann, J., 2018.
Tellurium, selenium and cobalt enrichment in Neoproterozoic black shales, Gwna
Group, UK: Deep marine trace element enrichment during the Second Great
Oxygenation Event. Terra Nova, 30(3): 244-253.
Asadi, H., Voncken, J., Kühnel, R. and Hale, M., 2000. Petrography, mineralogy and
geochemistry of the Zarshuran Carlin- like gold deposit, northwest Iran. Mineralium
Deposita, 35(7): 656-671.
Ba, L.A., Döring, M., Jamier, V. and Jacob, C., 2010. Tellurium: an element with great
biological potency and potential. Organic & Biomolecular Chemistry, 8(19): 4203-
4216.
Journal Pre-proof
Bae, E.J., Kang, Y.H., Jang, K.-S. and Cho, S.Y., 2016. Enhancement of thermoelectric
properties of PEDOT: PSS and tellurium-PEDOT: PSS hybrid composites by simple
chemical treatment. Scientific Reports, 6: 18805.
Baesman, S.M., Bullen, T., Dewald, J., Zhang, D., Curran, S., Islam, F., Beveridge, T. and
Oremland, R., 2006. Formation of Unique Tellurium Nanocrystals with Anaerobic
Growth of Bacillus selenitireducens and Sulfurospirillum barnesii using Te-oxyanions
as Electron Acceptors, AGU Fall Meeting Abstracts.
Baesman, S.M., Bullen, T.D., Dewald, J., Zhang, D., Curran, S., Islam, F.S., Beveridge, T.J.
and Oremland, R.S., 2007. Formation of tellurium nanocrystals during anaerobic
growth of bacteria that use Te oxyanions as respiratory electron acceptors. Applied
and Environmental Microbiology, 73(7): 2135-2143.
Baesman, S.M., Stolz, J., Kulp, T. and Oremland, R.S., 2009. Enrichment and isolation of
Bacillus beveridgei sp. nov., a facultative anaerobic haloalkaliphile from Mono Lake,
California, that respires oxyanions of tellurium, selenium, and arsenic. Extremophiles,
13(4): 695-705.
Bailey, R.T., 2017. Selenium contamination, fate, and reactive transport in groundwater in

f
relation to human health. Hydrogeology Journal, 25(4): 1191-1217.

oo
Bajaj, M. and Winter, J., 2014. Se (IV) triggers faster Te (IV) reduction by soil isolates of
heterotrophic aerobic bacteria: formation of extracellular SeTe nanospheres.
Microbial Cell Factories, 13(1): 168.
pr
Banerjee, A., Jhariya, M.K., Yadav, D.K. and Raj, A., 2018. Micro-remediation of Metals: A
New Frontier in Bioremediation. In: C.M. Hussain (Editor), Handbook of
e-
Environmental Materials Management. Springer International Publishing, Cham, pp.
479-513.
Bansal, V., Bharde, A., Ramanathan, R. and Bhargava, S.K., 2012. Inorganic materials using
Pr

‘unusual’microorganisms. Advances in Colloid and Interface Science, 179: 150-168.


Barin, I., 1995. Thermochemical Data of Pure Substances, third edition, VCH, Weinheim.
Basnayake, R.S., Bius, J.H., Akpolat, O.M. and Chasteen, T.G., 2001. Production of dimethyl
al

telluride and elemental tellurium by bacteria amended with tellurite or tellurate.


Applied Organometallic Chemistry, 15(6): 499-510.
Bateman, R. and Hagemann, S., 2004. Gold mineralisation throughout about 45 Ma of
rn

Archaean orogenesis: protracted flux of gold in the Golden Mile, Yilgarn craton,
Western Australia. Mineralium Deposita, 39(5-6): 536-559.
u

Baturin, G.N., 2012. The geochemistry of manganese and manganese nodules in the ocean,
Vol 2. Springer Science & Business Media.
Jo

Bauer, D.J., Diamond, D.B., Li, J., 2010. Critical Materials Strategy. U.S. Department of
Energy, https://www.hsdl.org/?view&did=703261.
Belousov, I., Large, R., Meffre, S., Danyushevsky, L., Steadman, J. and Beardsmore, T.,
2016. Pyrite compositions from VHMS and orogenic Au deposits in the Yilgarn
Craton, Western Australia: Implications for gold and copper exploration. Ore Geology
Reviews, 79: 474-499.
Belzile, N. and Chen, Y.-W., 2015. Tellurium in the environment: A critical review focused
on natural waters, soils, sediments and airborne particles. Applied Geochemistry, 63:
83-92.
Bennett, K.T., Bone, S.E., Akin, A.C., Birnbaum, E.R., Blake, A.V., Brugh, M., Daly, S.R.,
Engle, J.W., Fassbender, M.E. and Ferrier, M.G., 2019. Large-Scale Production of
119m
Te and 119 Sb for Radiopharmaceutical Applications. ACS Central Science, 5(3):
494-505.
Berriault, C.J. and Lightfoot, N.E., 2011. Occupational tellurium exposure and garlic odour.
Occupational Medicine, 61(2): 132-135.
Bethke, C.M. (2008) Geochemical and biogeochemical reaction modeling (second edition).
Cambridge University Press, New York. 543 pages.
Journal Pre-proof
Bindi, L. and Biagioni, C., 2018. A crystallographic excursion in the extraordinary world of
minerals: the case of Cu- and Ag-rich sulfosalts. Acta Crystallographica Section B,
74(6): 527-538.
Bindi, L., Paar, W.H. and Lepore, G.O., 2018.
Montbrayite,(Au,Ag,Sb,Pb,Bi)23 (Te,Sb,Pb,Bi)38 , from the Robb- montbray Mine,
Montbray, Quebec: Crystal Structure and Revision of the Chemical Formula. The
Canadian Mineralogist, 56(2): 129-142.
Binns, R., Dotter, L. and Blacklock, K., 2004. Chemistry of Borehole Fluids Collected at
PACMANUS, Papua New Guinea, OPD Leg 193.
Biver, M., Quentel, F. and Filella, M., 2015. Direct determination of tellurium and its redox
speciation at the low nanogram level in natural waters by catalytic cathodic stripping
voltammetry. Talanta, 144: 1007-1013.
Biver, M. and Filella, M., 2016. Bulk dissolution rates of cadmium and bismuth tellurides as
a function of pH, temperature and dissolved oxygen. Environmental Science &
Technology, 50(9): 4675-4681.
Blundy, J. and Wood, B., 1994. Prediction of crystal–melt partition coefficients from elastic

f
moduli. Nature, 372(6505): 452.

oo
Boni, M., Gilg, H.A., Aversa, G. and Balassone, G., 2003. The" calamine" of southwest
Sardinia: Geology, mineralogy, and stable isotope geochemistry of supergene Zn
mineralization. Economic Geology, 98(4): 731-748.
pr
Bonificio, W.D. and Clarke, D.R., 2014. Bacterial recovery and recycling of tellurium from
tellurium‐containing compounds by Pseudoalteromonas sp. EPR 3. Journal of Applied
e-
Microbiology, 117(5): 1293-1304.
Borghese, R., Canducci, L., Musiani, F., Cappelletti, M., Ciurli, S., Turner, R.J. and Zannoni,
Pr

D., 2016a. On the role of a specific insert in acetate permeases (ActP) for tellurite
uptake in bacteria: Functional and structural studies. Journal of Inorganic
Biochemistry, 163: 103-109.
Borghese, R., Brucale, M., Fortunato, G., Lanzi, M., Mezzi, A., Valle, F., Cavallini, M. and
al

Zannoni, D., 2016b. Extracellular production of tellurium nanoparticles by the


photosynthetic bacterium Rhodobacter capsulatus. Journal of Hazardous Materials,
rn

309: 202-209.
Bowell, R., 1992. Supergene gold mineralogy at Ashanti, Ghana: implications for the
supergene behaviour of gold. Mineralogical Magazine, 56(385): 545-560.
u

Braukmüller, N., Wombacher, F., Funk, C. and Münker, C., 2019. Earth’s volatile element
Jo

depletion pattern inherited from a carbonaceous chondrite- like source. Nature


Geoscience, 12(7): 564-568.
Brugger, J., Etschmann, B.E., Grundler, P.V., Liu, W., Testemale, D. and Pring, A., 2012.
XAS evidence for the stability of polytellurides in hydrothermal fluids up to 599°C,
800 bar. American Mineralogist, 97(8-9): 1519-1522.
Brugger, J., Liu, W., Etschmann, B., Mei, Y., Sherman, D.M. and Testemale, D., 2016. A
review of the coordination chemistry of hydrothermal systems, or do coordination
changes make ore deposits? Chemical Geology, 447: 219-253.
Bullock, L., Parnell, J., Perez, M. and Feldmann, J., 2017. Tellurium Enric hment in Jurassic
Coal, Brora, Scotland. Minerals, 7(12): 231.
Butler, I. and Nesbitt, R., 1999. Trace element distributions in the chalcopyrite wall of a
black smoker chimney: insights from laser ablation inductively coupled plasma mass
spectrometry (LA–ICP–MS). Earth and Planetary Science Letters, 167(3-4): 335-345.
Castro, M.E., Molina, R., Díaz, W., Pichuantes, S.E. and Vásquez, C.C., 2008. The
dihydrolipoamide dehydrogenase of Aeromonas caviae ST exhibits NADH-dependent
tellurite reductase activity. Biochemical and Biophysical Research Communications,
375(1): 91-94.
Journal Pre-proof
Castro, M.E., Molina, R.C., Díaz, W.A., Pradenas, G.A. and Vásquez, C.C., 2009. Expression
of Aeromonas caviae ST pyruvate dehydrogenase complex components mediate
tellurite resistance in Escherichia coli. Biochemical and Biophysical Research
Communications, 380(1): 148-152.
Cepedal, A., Fuertes-Fuente, M., Martin-Izard, A., Gonzalez-Nistal, S. and Rodriguez-Pevida,
L., 2006. Tellurides, selenides and Bi- mineral assemblages from the Río Narcea Gold
Belt, Asturias, Spain: genetic implications in Cu–Au and Au skarns. Mineralogy and
Petrology, 87(3-4): 277-304.
Chasteen, T.G. and Bentley, R., 2003. Biomethylation of selenium and tellurium:
microorganisms and plants. Chemical Reviews, 103(1): 1-26.
Chasteen, T.G., Fuentes, D.E., Tantaleán, J.C. and Vásquez, C.C., 2009. Tellurite: history,
oxidative stress, and molecular mechanisms of resistance. FEMS Microbiology
Reviews, 33(4): 820-832.
Chen, N., He, Y., Su, Y., Li, X., Huang, Q., Wang, H., Zhang, X., Tai, R. and Fan, C., 2012.
The cytotoxicity of cadmium-based quantum dots. Biomaterials, 33(5): 1238-1244.
Chen, R., Xu, D., Guo, G. and Gui, L., 2002. Silver telluride nanowires prepared by DC

f
electrodeposition in porous anodic alumina templates. Journal of Materials Chemistry,

oo
12(8): 2435-2438.
Chen, Y.-W., Alzahrani, A., Deng, T.-L. and Belzile, N., 2016. Valence properties of
tellurium in different chemical systems and its determination in refractory
pr
environmental samples using hydride generation–atomic fluorescence spectroscopy.
Analytica Chimica Acta, 905: 42-50.
e-
Cheng, Z., Zhang, Z., Chai, F., Hou, T., Santosh, M., Turesebekov, A. and Nurtaev, B., 2018.
Carboniferous porphyry Cu–Au deposits in the Almalyk orefield, Uzbekistan: the
Sarycheku and Kalmakyr examples. International Geology Review, 60(1): 1-20.
Pr

Chien, C.C. and Han, C.T., 2009. Tellurite Resistance and Reduction by a Paenibacillus Sp.
Isolated from Heavy Metal‐Contaminated Sediment. Environmental Toxicology and
Chemistry, 28(8): 1627-1632.
al

Chivers, T. and Laitinen, R.S., 2015. Tellurium: a maverick among the chalcogens. Chemical
Society Reviews, 44(7): 1725-1739.
rn

Choi, N.-C., Cho, K.H., Kim, B.J., Lee, S. and Park, C.Y., 2018. Enhancement of Au–Ag–Te
contents in tellurium-bearing ore minerals via bioleaching. International Journal of
Minerals, Metallurgy, and Materials, 25(3): 262-270.
u

Choudhury, H.G., Cameron, A.D., Iwata, S. and Beis, K., 2011. Structure and mechanism of
Jo

the chalcogen-detoxifying protein TehB from Escherichia coli. Biochemical Journal,


435(1): 85-91.
Chouinard, A., Paquette, J. and Williams-Jones, A.E., 2005. Crystallographic controls on
trace-element incorporation in auriferous pyrite from the Pascua epithermal high-
sulfidation deposit, Chile–Argentina. The Canadian Mineralogist, 43(3): 951-963.
Christy, A.G., 2015. Causes of anomalous mineralogical diversity in the Periodic Table.
Mineralogical Magazine, 79(1): 33-50.
Christy, A.G., 2018. Quantifying lithophilicity, chalcophilicity and siderophilicity. European
Journal of Mineralogy, 30(2): 193-204.
Christy, A.G. and Mills, S.J., 2013. Effect of lone-pair stereoactivity on polyhedral volume
and structural flexibility: application to Te IVO6 octahedra. Acta Crystallographica
Section B, 69(5): 446-456.
Christy, A.G., Mills, S.J. and Kampf, A.R., 2016a. A review of the structural architecture of
tellurium oxycompounds. Mineralogical Magazine, 80(3): 415-545.
Christy, A.G., Mills, S.J., Kampf, A.R., Housley, R.M., Thorne, B. and Marty, J., 2016b. The
relationship between mineral composition, crystal structure and paragenetic sequence:
the case of secondary Te mineralization at the Bird Nest drift, Otto Mountain,
California, USA. Mineralogical Magazine, 80(2): 291-310.
Journal Pre-proof
Ciobanu, C.L., Cook, N.J., Damian, G.H., Damian, F. and Buia, G., 2004. Telluride and
sulphosalt associations at Sacarimb. Au-Ag-Telluride Deposits of the Golden
Quadrilateral, Apuseni Mts., Romania, 12: 145-186.
Ciobanu, C.L., Cook, N.J. and Spry, P.G., 2006. Preface–Special Issue: Telluride and
selenide minerals in gold deposits–how and why? Mineralogy and Petrology, 87(3-4):
163-169.
Ciobanu, C.L., Cook, N.J., Pring, A., Brugger, J., Danyushevsky, L.V. and Shimizu, M.,
2009. ‘Invisible gold’ in bismuth chalcogenides. Geochimica et Cosmochimica Acta,
73(7): 1970-1999.
Climo, M., Watling, H. and van Bronswijk, W., 2000a. Bio-oxidation of a Telluride-rich
Gold Concentrate. Chemeca 2000: Opportunities and Challenges for the Resource and
Processing Industries: 46.
Climo, M., Watling, H.R. and Van Bronswijk, W., 2000b. Biooxidation as pre-treatment for a
telluride-rich refractory gold concentrate. Minerals Engineering, 13(12): 1219-1229.
Clout, J., Cleghorn, J. and Eaton, P., 1990. Geology of the Kalgoorlie gold field, Geology of
the mineral deposits of Australia and Papua New Guinea. The Australasian Institute

f
of Mining and Metallurgy, Melbourne, pp. 411-431.

oo
Cook, N.J., Ciobanu, C.L. and Mao, J., 2009a. Textural control on gold distribution in As-
free pyrite from the Dongping, Huangtuliang and Hougou gold deposits, North China
Craton (Hebei Province, China). Chemical Geology, 264(1-4): 101-121.
pr
Cook, N.J., Ciobanu, C.L., Pring, A., Skinner, W., Shimizu, M., Danyushevsky, L., Saini-
Eidukat, B. and Melcher, F., 2009b. Trace and minor elements in sphalerite: A LA-
e-
ICPMS study. Geochimica et Cosmochimica Acta, 73(16): 4761-4791.
Cook, N.J., Ciobanu, C.L., Spry, P.G. and Voudouris, P., 2009c. Understanding gold-(silver)-
telluride-(selenide) mineral deposits. Episodes, 32(4): 249-263.
Pr

Cooke, D.R. and McPhail, D., 2001. Epithermal Au-Ag-Te mineralization, Acupan, Baguio
district, Philippines: numerical simulations of mineral deposition. Economic Geology,
96(1): 109-131.
al

Corper, H.J., 1915. Sodium Tellurite as a Rapid Test for the Viability of Tubercle Bacilli
Studies on the Biochemistry and Chemotherapy of Tuberculosis, XIII. The Journal of
Infectious Diseases: 47-53.
rn

Cowgill, U.M., 1988. The tellurium content of vegetation. Biological Trace Element
Research, 17(1): 43-67.
u

Cunha, R.L., Gouvea, I.E. and Juliano, L., 2009. A glimpse on biological activities of
tellurium compounds. Anais da Academia Brasileira de Ciências, 81(3): 393-407.
Jo

Cyrs, W.D., Avens, H.J., Capshaw, Z.A., Kingsbury, R.A., Sahmel, J. and Tvermoes, B.E.,
2014. Landfill waste and recycling: Use of a screening- level risk assessment tool for
end-of- life cadmium telluride (CdTe) thin- film photovoltaic (PV) panels. Energy
Policy, 68: 524-533.
Dare, S.A., Barnes, S.-J., Prichard, H.M. and Fisher, P.C., 2014. Mineralogy and
geochemistry of Cu-rich ores from the McCreedy East Ni-Cu-PGE deposit (Sudbury,
Canada): implications for the behavior of platinum group and chalcophile elements at
the end of crystallization of a sulfide liquid. Economic Geology, 109(2): 343-366.
De Ronde, C., Chadwick, W., Ditchburn, R., Embley, R., Tunnicliffe, V., Baker, E., Walker,
S., Ferrini, V. and Merle, S., 2015. Molten sulfur lakes of intraoceanic arc volcanoes,
Volcanic lakes. Springer, pp. 261-288.
Deditius, A.P., Utsunomiya, S., Reich, M., Kesler, S.E., Ewing, R.C., Hough, R. and Walshe,
J., 2010. Trace-metal nanoparticles in pyrite. Ore Geology Reviews, 42: 32-46.
Deditius, A.P., Reich, M., Kesler, S.E., Utsunomiya, S., Chryssoulis, S.L., Walshe, J. and
Ewing, R.C., 2014. The coupled geochemistry of Au and As in pyrite from
hydrothermal ore deposits. Geochimica et Cosmochimica Acta, 140: 644-670.
Journal Pre-proof
Deen, J.A. and Atkinson Jr, W.W., 1988. Volcanic stratigraphy and ore deposits of the
Moctezuma District, Sonora, Mexico. Economic Geology, 83(8): 1841-1855.
Dhankhar, R. and Hooda, A., 2011. Fungal biosorption–an alternative to meet the challenges
of heavy metal pollution in aqueous solutions. Environmental Technology, 32(5):
467-491.
Díaz-Vásquez, W.A., Abarca-Lagunas, M.J., Cornejo, F.A., Pinto, C.A., Arenas, F.A. and
Vásquez, C.C., 2015. Tellurite- mediated damage to the Escherichia coli NDH-
dehydrogenases and terminal oxidases in aerobic conditions. Archives of
Biochemistry and Biophysics, 566: 67-75.
Dickson, R.S. and Glowa, G.A., 2019. Tellurium behaviour in the Fukushima Dai- ichi
Nuclear Power Plant accident. Journal of Environmental Radioactivity, 204: 49-65.
Dill, H.G., 2010. The “chessboard” classification scheme of mineral deposits: mineralogy and
geology from aluminum to zirconium. Earth-Science Reviews, 100(1-4): 1-420.
Dimpe, K.M. and Nomngongo, P.N., 2017. A review on the efficacy of the application of
myriad carbonaceous materials for the removal of toxic trace elements in the
environment. Trends in Environmental Analytical Chemistry, 16: 24-31.

f
Ding, N., Chen, S.F., Geng, D.S., Chien, S.W., An, T., Hor, T.A., Liu, Z.L., Yu, S.H. and

oo
Zong, Y., 2015. Tellurium@ Ordered Macroporous Carbon Composite and Free‐
Standing Tellurium Nanowire Mat as Cathode Materials for Rechargeable Lithium–
pr
Tellurium Batteries. Advanced Energy Materials, 5(8): 1401999.
Diso, D., Fauzi, F., Echendu, O., Olusola, O. and Dharmadasa, I., 2016. Optimisation of
CdTe electrodeposition voltage for development of CdS/CdTe solar cells. Journal of
e-
Materials Science: Materials in Electronics, 27(12): 12464-12472.
Dmitrijeva, M., Cook, N.J., Ehrig, K., Ciobanu, C.L., Metcalfe, A.V., Kamenetsky, M.,
Pr

Kamenetsky, V.S. and Gilbert, S., 2020. Multivariate Statistical Analysis of Trace
Elements in Pyrite: Prediction, Bias and Artefacts in Defining Mineral Signatures.
Minerals, 10(1): 61.
Drozdovitch, V., Kryuchkov, V., Chumak, V., Kutsen, S., Golovanov, I. and Bouville, A.,
al

2019. Thyroid doses due to Iodine-131 inhalation among Chernobyl cleanup workers.
Radiation and Environmental Biophysics, 58(2): 183-194.
rn

du Bray, E.A., 2017. Geochemical characteristics of igneous rocks associated with epithermal
mineral deposits—a review. Ore Geology Reviews, 80: 767-783.
Dunn, C.E., 2011. Biogeochemistry in mineral exploration. Vol. 9 of Handbook of
u

Exploration and Environmental Chemistry. Elsevier Publishers, Amsterdam, The


Jo

Netherlands.
Echmaeva, E.A. and Osadchii, E.G., 2009. Determination of the thermodynamic properties of
compounds in the Ag–Au–Se and Ag–Au–Te systems by the EMF method. Geology
of Ore Deposits, 51: 247–258.
Ehrig, K., McPhie, J. and Kamenetsky, V., 2012. Geology and mineralogical zonation of the
Olympic Dam iron oxide Cu-U-Au-Ag deposit, South Australia. 2012 Society of
Economic Geologists, Inc. Special Publication 16, pp. 237–267
Ehrlich, H. and Newman, D., 2009. Geomicrobiology CRC Press, Boca Raton, USA (606z
pp.).
Emsley, J., 2011. Nature's Building Blocks: an A-Z Guide to the Elements. Oxford
University Press, Oxford.
Endo, S., Fujii, K., Kajimoto, T., Tanaka, K., Stepanenko, V., Kolyzhenkov, T., Petukhov, A.,
Akhmedova, U. and Bogacheva, V., 2018. Comparison of calculated beta- and
gamma-ray doses after the Fukushima accident with data from single-grain
luminescence retrospective dosimetry of quartz inclusions in a brick sample. Journal
of Radiation Research, 59(3): 286-290.
Espinosa-Ortiz, E.J., Rene, E.R., Guyot, F., van Hullebusch, E.D. a nd Lens, P.N.L., 2017.
Biomineralization of tellurium and selenium-tellurium nanoparticles by the white-rot
Journal Pre-proof
fungus Phanerochaete chrysosporium. International Biodeterioration &
Biodegradation, 124: 258-266.
Estevam, E.C., Griffin, S., Nasim, M.J., Denezhkin, P., Schneider, R., Lilischkis, R.,
Dominguez-Alvarez, E., Witek, K., Latacz, G. and Keck, C., 2017. Natural selenium
particles from Staphylococcus carnosus: Hazards or particles with particular promise?
Journal of Hazardous Materials, 324: 22-30.
Etezad, S.M., Khajeh, K., Soudi, M., Ghazvini, P.T.M. and Dabirmanesh, B., 2009. Evidence
on the presence of two distinct enzymes responsible for the reduction of selenate and
tellurite in Bacillus sp. STG-83. Enzyme and Microbial Technology, 45(1): 1-6.
Etschmann, B.E., Liu, W., Pring, A., Grundler, P.V., Tooth, B., Borg, S., Testemale, D.,
Brewe, D. and Brugger, J., 2016. The role of Te (IV) and Bi(III) chloride complexes in
hydrothermal mass transfer: an X-ray absorption spectroscopic study. Chemical
Geology, 425: 37-51.
Fairbrother, L., Shapter, J., Brugger, J., Southam, G., Pring, A. and Reith, F., 2009. Effect of
the cyanide-producing bacterium Chromobacterium violaceum on ultraflat Au
surfaces. Chemical Geology, 265(3): 313-320.

f
Fairbrother, L., Brugger, J., Shapter, J., Laird, J., Southam, G. and Reith, F., 2012. Supergene

oo
gold transformation: Biogenic secondary and nano-particulate gold from arid
Australia. Chemical Geology, 320: 17-31.
Filella, M. and Rodushkin, I., 2018. A concise guide for the determination of less-studied
pr
technology-critical elements (Nb, Ta, Ga, In, Ge, Te) by inductively coupled plasma
mass spectrometry in environmental samples. Spectrochimica Acta Part B: Atomic
e-
Spectroscopy, 141: 80-84.
Filella, M., 2019. Foreword to the Research Fro nt on ‘Tellurium in Biological and
Environmental Systems: After Fukushima’. Environmental Chemistry, 16(4): 213-214.
Pr

Filella, M. and May, P.M., 2019. The aqueous chemistry of tellurium: critically- selected
equilibrium constants for the low-molecular-weight inorganic species. Environmental
Chemistry, 16(4): 289-295.
al

Filella, M., Reimann, C., Biver, M., Rodushkin, I. and Rodushkina, K., 2019. Tellurium in
the environment: current knowledge and identification of gaps. Environmental
Chemistry, 16(4): 215-228.
rn

Fleet, M.E. and Mumin, A.H., 1997. Gold-bearing arsenian pyrite and marcasite and
arsenopyrite from Carlin Trend gold deposits and laboratory synthesis. American
u

Mineralogist, 82(1-2): 182-193.


Foreman, M.R.S.J., 2015. An introduction to serious nuclear acc ident chemistry. Cogent
Jo

Chemistry, 1(1): 1049111.


Frishberg, M., 2017. Advances in Solar Power Light the Future. Research Technology
Management, 60(2): 7.
Frost, B.R., Mavrogenes, J.A. and Tomkins, A.G., 2002. Partial melting of sulfide ore
deposits during medium-and high-grade metamorphism. The Canadian Mineralogist,
40(1): 1-18.
Fthenakis, V. and Wang, W., 2006. Extraction and separation of Cd and Te from cadmium
telluride photovoltaic manufacturing scrap. Progress in Photovoltaics: Research and
Applications, 14(4): 363-371.
Fulignati, P. and Sbrana, A., 1998. Presence of native gold and tellurium in the active high-
sulfidation hydrothermal system of the La Fossa volcano (Vulcano, Italy). Journal of
Volcanology and Geothermal Research, 86(1-4): 187-198.
Gao, S., Xu, H., Li, S., Santosh, M., Zhang, D., Yang, L. and Quan, S., 2017. Hydrothermal
alteration and ore- forming fluids associated with gold-tellurium mineralization in the
Dongping gold deposit, China. Ore Geology Reviews, 80: 166-184.
Gavrilescu, M., 2004. Removal of heavy metals from the environment by biosorption.
Engineering in Life Sciences, 4(3): 219-232.
Journal Pre-proof
Genkin, A. and Evstigneeva, T., 1986. Associations of platinum- group minerals of the
Noril'sk copper-nickel sulfide ores. Economic Geology, 81(5): 1203-1212.
George, L.L., Cook, N.J., Crowe, B.B. and Ciobanu, C.L., 2018. Trace Elements in
Hydrothermal Chalcopyrite. Mineralogical Magazine, 82(1): 59-88.
Gerhardsson, L., 2015. Tellurium, Handbook on the Toxicology of Metals. Elsevier, pp.
1217-1228.
Gharieb, M.M., Kierans, M. and Gadd, G.M., 1999. Transformation and tolerance of tellurite
by filamentous fungi: accumulation, reduction, and volatilization. Mycological
Research, 103(3): 299-305.
Gil-Díaz, T., 2019. Tellurium radionuclides produced by major accidental events in nuclear
power plants. Environmental Chemistry, 16(4): 296–302.
Gil-Díaz, T., Schäfer, J., Dutruch, L., Bossy, C., Pougnet, F., Abdou, M., Lerat-Hardy, A.,
Pereto, C., Derriennic, H. and Briant, N., 2019a. Tellurium behaviour in a major
European fluvial–estuarine system (Gironde, France): fluxes, solid/liquid partitioning
and bioaccumulation in wild oysters. Environmental Chemistry, 16(4): 229–242.
Gil-Díaz, T., Schäfer, J., Dutruch, L. and Eyrolle-Boyer, F., 2019b. Tellurium radionuclide

f
dispersion scenarios in aquatic systems: coupling of adsorption kinetics, radionuclide

oo
decay and estuarine hydrodynamics. Book of Abstracts of the Advanced Workshop on
Solution Chemistry of TCEs (Technology Critical Elements), Bialystok, pp. 26.
Goff, J. and Yee, N., 2017. Tellurate enters Escherichia coli K-12 cells via the SulT-type
pr
sulfate transporter CysPUWA. FEMS microbiology letters, 364(24): fnx241.
Goldfarb, R.J., 2014. Tellurium: The Bright Future of Solar Energy. US Department of the
e-
Interior, US Geological Survey.
Goldfarb, R.J., Berger, B.R., George, M.W., and Seal, R.R., II, 2017, Tellurium, chap. R of
Schulz, K.J., DeYoung, J.H., Jr., Seal, R.R., II, and Bradley, D.C., eds., Critical
Pr

mineral resources of the United States—Economic and environmental geology and


prospects for future supply: U.S. Geological Survey Professional Paper 1802, p. R1–
R27, DOI: 10.3133/pp1802R.
al

Gott, G.B. and McCarthy, J.H.J., 1966. Distribution of gold, silver, tellurium, and mercury in
the Ely mining district, White Pine County, Nevada. In: U.S. Dept of the Interior
(Editor), pp. 5 p.
rn

Green, M.A., 2011. Is sour crude or sour gas a potential source of Se and Te? Progress in
Photovoltaics: Research and Applications, 19(8): 991-995.
u

Greenland, L.P. and Aruscavage, P., 1986. Volcanic emission of Se, Te, and As from Kilauea
volcano, Hawaii. Journal of Volcanology and Geothermal Research, 27(1): 195-201.
Jo

Grundler, P.V., Brugger, J., Etschmann, B.E., Helm, L., Liu, W., Spry, P.G., Tian, Y.,
Testemale, D. and Pring, A., 2013. Speciation of aqueous tellurium (IV) in
hydrothermal solutions and vapors, and the role of oxidized tellurium species in Te
transport and gold deposition. Geochimica et Cosmochimica Acta, 120: 298-325.
Guo, Y.F., Deng, T.L., Zhang, N. and Liao, M.X., 2012a. Tellurium Extraction from the
Unique Independent Tellurium Ores in China by Bioleaching, Advanced Materials
Research. Trans Tech Publ, pp. 625-630.
Guo, Y.F., Zhang, N., Li, D.C., Tang, F.M. and Deng, T.L., 2012b. Tellurium recovery from
the unique tellurium ores, Advanced Materials Research. Trans Tech Publ, pp. 1060-
1063.
Hardardóttir, V., Brown, K., Fridriksson, T., Hedenquist, J., Hannington, M. and
Thorhallsson, S., 2009. Metals in deep liquid of the Reykjanes geothermal system,
southwest Iceland: Implications for the composition of seafloor black smoker fluids.
Geology, 37(12): 1103-1106.
Hattori, K.H., Arai, S. and Clarke, D.B., 2002. Selenium, tellurium, arsenic and antimony
contents of primary mantle sulfides. The Canadian Mineralogist, 40(2): 637-650.
Journal Pre-proof
Hayes, S.M. and Ramos, N.A., 2019. Surficial geochemistry and bioaccessibility of tellurium
in semiarid mine tailings. Environmental Chemistry, 16(4): 251-265.
He, J., Chen, Y., Lv, W., Wen, K., Wang, Z., Zhang, W., Li, Y., Qin, W. and He, W., 2016.
Three-dimensional hierarchical reduced graphene oxide/tellurium nanowires: a high-
performance freestanding cathode for Li–Te batteries. ACS Nano, 10(9): 8837-8842.
He, J., Lv, W., Chen, Y., Wen, K., Xu, C., Zhang, W., Li, Y., Qin, W. and He, W., 2017.
Tellurium- impregnated porous cobalt-doped carbon polyhedra as superior cathodes
for lithium–tellurium batteries. ACS Nano, 11(8): 8144-8152.
Hein, J.R., Koschinsky, A. and Halliday, A.N., 2003. Global occurrence of tellurium-rich
ferromanganese crusts and a model for the enrichment of tellurium. Geochimica et
Cosmochimica Acta, 67(6): 1117-1127.
Heitt, D.G., Dunbar, W.W., Thompson, T.B. and Jackson, R.G., 2003. Geology and
geochemistry of the Deep Star gold deposit, Carlin trend, Nevada. Economic Geology,
98(6): 1107-1135.
Helmy, H.M., Ballhaus, C., Berndt, J., Bockrath, C. and Wohlgemuth-Ueberwasser, C., 2007.
Formation of Pt, Pd and Ni tellurides: experiments in sulfide–telluride systems.

f
Contributions to Mineralogy and Petrology, 153(5): 577-591.

oo
Hitzman, M.W., Reynolds, N.A., Sangster, D., Allen, C.R. and Carman, C.E., 2003.
Classification, genesis, and exploration guides for nonsulfide zinc deposits. Economic
Geology, 98(4): 685-714.
pr
Holwell, D.A., Adeyemi, Z., Ward, L.A., Smith, D.J., Graham, S.D., McDonald, I. and Smith,
J.W., 2017. Low temperature alteration of magmatic Ni-Cu-PGE sulfides as a source
e-
for hydrothermal Ni and PGE ores: A quantitative approach using automated
mineralogy. Ore Geology Reviews, 91: 718-740.
Holwell, D.A., Fiorentini, M., McDonald, I., Lu, Y., Giuliani, A., Smith, D.J., Keith, M. and
Pr

Locmelis, M., 2019. A metasomatized lithospheric mantle control on the metallogenic


signature of post-subduction magmatism. Nature Communications, 10(1): 3511.
Hu, Z. and Gao, S., 2008. Upper crustal abundances of trace elements: a revision and update.
al

Chemical Geology, 253(3-4): 205-221.


Jacobson, M., W. Keller, J. and W. Atkinson Jr, W., 2018. The Where of Mineral Names:
Moctezumite, Moctezuma Mine (La Bambolla Mine), Moctezuma, Municipality of
rn

Moctezuma, State of Sonora, Mexico. Rocks & Minerals, 93(5): 466-471.


Jain, R., Gonzalez-Gil, G., Singh, V., Van Hullebusch, E.D., Farges, F. and Lens, P.N.L.,
u

2014. Biogenic selenium nanoparticles: production, characterization and challenges.


Nanobiotechnology, Studium Press LLC, USA: 361-390.
Jo

Jain, R., Matassa, S., Singh, S., van Hullebusch, E.D., Esposito, G. and Lens, P.N., 2016.
Reduction of selenite to elemental selenium nanoparticles by activated sludge.
Environmental Science and Pollution Research, 23(2): 1193-1202.
Kagami, T., Fudemoto, A., Fujimoto, N., Notaguchi, E., Kanzaki, M., Kuroda, M., Soda, S.,
Yamashita, M. and Ike, M., 2012. Isolation and characterization of bacteria capable of
reducing tellurium oxyanions to insoluble elemental tellurium for tellurium recovery
from wastewater. Waste and Biomass Valorization, 3(4): 409-418.
Kagoshima, T., Sano, Y., Takahata, N., Maruoka, T., Fischer, T.P. and Hattori, K., 2015.
Sulphur geodynamic cycle. Scientific Reports, 5: 8330.
Kampf, A.R., Cooper, M.A., Mills, S.J., Housley, R.M. and Rossman, G.R., 2016. Lead-
tellurium oxysalts from Otto Mountain near Baker, California, USA: XII.
Andychristyite, PbCu2+Te6+O 5 (H2 O), a new mineral with hcp stair-step layers.
Mineralogical Magazine, 80(6): 1055-1065.
Kanai, Y., 2015. Geochemical behavior and activity ratios of Fukushima-derived
radionuclides in aerosols at the Geological Survey of Japan, Tsukuba, Japan. Journal
of Radioanalytical and Nuclear Chemistry, 303(2): 1405-1408.
Journal Pre-proof
Kavlak, G. and Graedel, T., 2013. Global anthropogenic tellurium cycles for 1940–2010.
Resources, Conservation and Recycling, 76: 21-26.
Keith, M., Smith, D.J., Jenkin, G.R.T., Holwell, D.A. and Dye, M.D., 2018. A review of Te
and Se systematics in hydrothermal pyrite from precious metal deposits: Insights into
ore-forming processes. Ore Geology Reviews, 96: 269-282.
Kelley, K.D., Romberger, S.B., Beaty, D.W., Pontius, J.A., Snee, L.W., Stein, H.J. and
Thompson, T.B., 1998. Geochemical and geochronological constraints on the genesis
of Au-Te deposits at Cripple Creek, Colorado. Economic Geology, 93(7): 981-1012.
Kesler, S.E. and Wilkinson, B.H., 2008. Earth's copper resources estimated from tectonic
diffusion of porphyry copper deposits. Geology, 36(3): 255-258.
Khaing, S.Y., Sugai, Y. and Sasaki, K., 2019. Gold Dissolution from Ore with Iodide-
Oxidising Bacteria. Scientific Reports, 9(1): 4178.
Kim, D.-H., Kanaly, R.A. and Hur, H.-G., 2012a. Biological accumulation of tellurium
nanorod structures via reduction of tellurite by Shewanella oneidensis MR-1.
Bioresource Technology, 125: 127-131.
Kim, E.-J., Park, M.-E. and White, N.C., 2012b. Skarn gold mineralization at the Geodo mine,

f
South Korea. Economic Geology, 107(3): 537-551.

oo
Kim, P., Kim, H., Myung, E., Kim, Y. and Lee, Y., 2015. The enhancing of Au-Ag-Te
content in tellurium-bearing ore mineral by bio-oxidation- leaching, EGU General
Assembly Conference Abstracts.
pr
King, J., Williams-Jones, A., van Hinsberg, V. and Williams-Jones, G., 2014. High-
sulfidation epithermal pyrite- hosted Au (Ag-Cu) ore formation by condensed
e-
magmatic vapors on Sangihe Island, Indonesia. Economic Geology, 109(6): 1705-
1733.
King, W.E. and Davis, L., 1914. Potassium tellurite as an indicator of microbial life.
Pr

American Journal of Public Health, 4(10): 917-932.


Kingston, G.A., 1966. The occurrence of platinoid bismuthotellurides in the Merensky Reef
at Rustenburg platinum mine in the western Bushveld. Mineralogical Magazine,
al

35(274): 815-834.
Klaproth, M.H., 1798. XVIII. Extract from a memoir on a new metal ca lled tellurium. The
Philosophical Magazine, 1(1): 78-82.
rn

Klonowska, A., Heulin, T. and Vermeglio, A., 2005. Selenite and tellurite reduction by
Shewanella oneidensis. Applied and Environmental Microbiology, 71(9): 5607-5609.
u

Knockaert, G., 2011. Tellurium and tellurium compounds. In ‘Ullmann’s encyclopedia of


industrial chemistry’. Wiley‐VCH Verlag GmbH & Co. KGaA: Weinheim.
Jo

Konstantinov, M. and Grushin, V., 1970. Geologic position of the Zod-Agduzdag gold-ore
node in Transcaucasia. International Geology Review, 12(12): 1447-1453.
Kumar, M. and Kumar, S., 2015. Luminescent CdTe quantum dots incarcerated in a
columnar matrix of discotic liquid crystals for optoelectronic applications. RSC
Advances, 5(2): 1262-1267.
Kyle, J., Breuer, P., Bunney, K., Pleysier, R. and May, P., 2011. Review of trace toxic
elements (Pb, Cd, Hg, As, Sb, Bi, Se, Te) and their deportment in gold processing.
Part 1: Mineralogy, aqueous chemistry and toxicity. Hydrometallurgy, 107(3-4): 91-
100.
LeFort, D., Hanley, J. and Guillong, M., 2011. Subepithermal Au-Pd Mineralization
Associated with an Alkalic Porphyry Cu-Au Deposit, Mount Milligan, Quesnel
Terrane, British Columbia, Canada. Economic Geology, 106(5): 781-808.
Li, L., Daou, T.J., Texier, I., Kim Chi, T.T., Liem, N.Q. and Reiss, P., 2009. Highly
luminescent CuInS2 /ZnS core/shell nanocrystals: cadmium- free quantum dots for in
vivo imaging. Chemistry of Materials, 21(12): 2422-2429.
Journal Pre-proof
Liang, X., Perez, M.A.M.-J., Nwoko, K.C., Egbers, P., Feldmann, J., Csetenyi, L. and Gadd,
G.M., 2019. Fungal formation of selenium and tellurium nanoparticles. Applied
Microbiology and Biotechnology, 103(17): 7241-7259.
Lin, S., Li, W., Chen, Z., Shen, J., Ge, B. and Pei, Y., 2016. Tellurium as a high-performance
elemental thermoelectric. Nature Communications, 7: 10287.
Liu, M., Turner, R.J., Winstone, T.L., Saetre, A., Dyllick-Brenzinger, M., Jickling, G., Tari,
L.W., Weiner, J.H. and Taylor, D.E., 2000. Escherichia coli TehB RequiresS-
Adenosylmethionine as a Cofactor To Mediate Tellurite Resistance. Journal of
bacteriology, 182(22): 6509-6513.
Liu, J.-W., Zhu, J.-H., Zhang, C.-L., Liang, H.-W. and Yu, S.-H., 2010. Mesostructured
assemblies of ultrathin superlong tellurium nanowires and their photoconductivity.
Journal of the American Chemical Society, 132(26): 8945-8952.
Lodders, K., 2010. Solar system abundances of the elements, Principles and perspectives in
cosmochemistry. Springer, pp. 379-417.
Lovrić, J., Cho, S.J., Winnik, F.M. and Maysinger, D., 2005. Unmodified cadmium telluride
quantum dots induce reactive oxygen species formation leading to multiple organelle

f
damage and cell death. Chemistry & Biology, 12(11): 1227-1234.

oo
Mahdavi, S., Khanmohammadi, H. and Masteri-Farahani, M., 2018. Surface functionalized
cadmium telluride quantum dots for the optical detection and determination of
herbicides. Journal of Materials Science: Materials in Electronics: 1-6.
pr
Makuei, F.M. and Senanayake, G., 2018. Extraction of tellurium from lead and copper
bearing feed materials and interim metallurgical products–A short review. Minerals
e-
Engineering, 115: 79-87.
Mal, J., Nancharaiah, Y.V., Maheshwari, N., van Hullebusch, E.D. and Lens, P.N.L., 2017.
Continuous removal and recovery of tellurium in an upflow anaerobic granular sludge
Pr

bed reactor. Journal of Hazardous Materials, 327: 79-88.


Maltman, C., Walter, G. and Yurkov, V., 2016. A diverse community of metal (loid) oxide
respiring bacteria is associated with tube worms in the vicinity of the Juan de Fuca
al

Ridge black smoker field. PloS One, 11(2): e0149812.


Maltman, C., Donald, L.J. and Yurkov, V., 2017. Two distinct periplasmic enzymes are
responsible for tellurite/tellurate and selenite reduction by strain ER-Te-48 associated
rn

with the deep sea hydrothermal vent tube worms at the Juan de Fuca Ridge black
smokers. Archives of Microbiology, 199(8): 1113-1120.
u

Manikandan, M., Dhanuskodi, S., Maheswari, N., Muralidharan, G., Revathi, C., Kumar, R.R.
and Rao, G.M., 2017. High performance supercapacitor and non-enzymatic hydrogen
Jo

peroxide sensor based on tellurium nanoparticles. Sensing and Bio-Sensing Research,


13: 40-48.
Mao, J., Wang, Y., Ding, T., CHEN, Y., WEI, J. and YIN, J., 2002. Dashuigou tellurium
deposit in Sichuan Province, China: S, C, O, and H isotope data and their implications
on hydrothermal mineralization. Resource Geology, 52(1): 15-23.
Marwede, M. and Reller, A., 2012. Future recycling flows of tellurium from cadmium
telluride photovoltaic waste. Resources, Conservation and Recycling, 69: 35-49.
Maslennikov, V., Maslennikova, S., Large, R., Danyushevsky, L., Herrington, R. and Stanley,
C., 2013. Tellurium-bearing minerals in zoned sulfide chimneys from Cu-Zn massive
sulfide deposits of the Urals, Russia. Mineralogy and Petrology, 107(1): 67-99.
McDonough, W.F. and Sun, S.-S., 1995. The composition of the Earth. Chemical geology,
120(3-4): 223-253.
McFall, K.A., Naden, J., Roberts, S., Baker, T., Spratt, J. and McDonald, I., 2018. Platinum-
group minerals in the Skouries Cu-Au (Pd, Pt, Te) porphyry deposit. Ore Geology
Reviews, 99: 344-364.
McPhail, D.C., 1995. Thermodynamic properties of aqueous tellurium species between 25
and 350ºC. Geochimica et Cosmochimica Acta, 59(5): 851-866.
Journal Pre-proof
Mehrabi, B., Yardley, B. and Cann, J., 1999. Sediment-hosted disseminated gold
mineralisation at Zarshuran, NW Iran. Mineralium Deposita, 34(7): 673-696.
Meintjes, T., Bob, L., Gary, G. and Marc, S., 2018. NI-43-101 Technical report on the
preliminary economic assessment for the Deer Horn gold-silver-tellurium property.
Mills, K.C., 1974. Thermodynamic data for inorganic sulphides, selenides and tellurides.
Mills, S.J. and Christy, A.G., 2013. Revised values of the bond- valence parameters for TeI V-
O, TeVI-O and TeIV-Cl. Acta Crystallographica Section B, 69(2): 145-149.
Mills, S.J. and Christy, A.G., 2019. Mineral extinction. Mineralogical Magazine, 83(5): 621-
625.
Missen, O.P., Kampf, A.R., Mills, S.J., Housley, R.M., Spratt, J., Welch, M.D., Coolbaugh,
M.F., Marty, J., Chorazewicz, M. and Ferraris, C., 2019. The crystal structures of the
mixed-valence tellurium oxysalts tlapallite,
(Ca,Pb)3 CaCu6 [Te 3 Te O 12 ]2 (Te O3 )2 (SO 4 )2 ·3H2 O, and carlfriesite, CaTe 2 Te6+O8 .
4+ 6+ 4+ 4+

Mineralogical Magazine, 83(4): 539-549.


Moëlo, Y., Makovicky, E., Mozgova, N.N., Jambor, J.L., Cook, N., Pring, A., Paar, W.,
Nickel, E.H., Graeser, S. and Karup-Møller, S., 2008. Sulfosalt systematics: a review.

f
Report of the sulfosalt sub-committee of the IMA Commission on Ore Mineralogy.

oo
European Journal of Mineralogy, 20(1): 7-46.
Mueller, A.G., Hagemann, S.G., Brugger, J., Xing, Y. and Roberts, M.P., 2020. Early
Fimiston pyrite and late Oroya pyrite-telluride gold ore, Paringa, South mine, Golden
pr
Mile, Kalgoorlie: 4. Mineralogy and PTX constraints on superimposed magmatic-
hydrothermal systems. Mineralium Deposita, https://doi.org/10.1007/s00126-019-
e-
00939-8.
Murray, C.B., Norris, D.J. and Bawendi, M.G., 1993. Synthesis and characterization of nearly
monodisperse CdE (E = sulfur, selenium, tellurium) semiconductor nanocrystallites.
Pr

Journal of the American Chemical Society, 115(19): 8706-8715.


Nancharaiah, Y.V. and Lens, P., 2015. Ecology and biotechnology of selenium-respiring
bacteria. Microbiology and Molecular Biology Reviews, 79(1): 61-80.
al

Nancharaiah, Y.V., Mohan, S.V. and Lens, P., 2016. Biological and bioelectrochemical
recovery of critical and scarce metals. Trends in Biotechnology, 34(2): 137-155.
Norman, M., 2017. Copper tellurium oxides–A playground for magnetism. Journal of
rn

Magnetism and Magnetic Materials.


Nuss, P., 2019. Losses and environmental aspects of a byproduct metal: tellurium.
u

Environmental Chemistry, 16(4): 243-250, https://doi.org/10.1071/EN18282.


O'Neill, H.S.C. and Palme, H., 2008. Collisional erosion and the non-chondritic composition
Jo

of the terrestrial planets. Philosophical Transactions of the Royal Society A:


Mathematical, Physical and Engineering Sciences, 366(1883): 4205-4238.
Ogra, Y., 2009. Toxicometallomics for research on the toxicology of exotic metalloids based
on speciation studies. Analytical Sciences, 25(10): 1189-1195.
Ogra, Y., 2017. Biology and toxicology of tellurium explored by speciation analysis.
Metallomics, 9(5): 435-441.
Okrugin, V.M., Andreeva, E., Etschmann, B., Pring, A., Li, K., Zhao, J., Griffiths, G.,
Lumpkin, G.R., Triani, G. and Brugger, J., 2014. Microporous gold : Comparison of
textures from Nature and experiments. American Mineralogist, 99: 1171-1174.
Okrugin, V.M., Favero, M., Liu, A., Etschmann, B., Plutachina, E., Mills, S., Tomkins, A.G.,
Lukasheva, M., Kozlov, V., Moskaleva, S., Chubarov, M. and Brugger, J., 2017.
Smoking gun for thallium geochemistry in volcanic arcs: Nataliyamalikite, TlI, a new
thallium mineral from an active fumarole at Avacha Volcano, Kamchatka Peninsula,
Russia. American Mineralogist, 102(8): 1736-1746.
Olivotos, S. and Economou- Eliopoulos, M., 2016. Gibbs free energy of formation for
selected platinum group minerals (PGM). Geosciences, 6(1): 2.
Journal Pre-proof
Ollivier, P.R., Bahrou, A.S., Marcus, S., Cox, T., Church, T.M. and Hanson, T.E., 2008.
Volatilization and precipitation of tellurium by aerobic, tellurite-resistant marine
microbes. Applied and Environmental Microbiology, 74(23): 7163-7173.
Ottosson, L.-G., Logg, K., Ibstedt, S., Sunnerhagen, P., Käll, M., Blomberg, A. and
Warringer, J., 2010. Sulfate assimilation mediates tellurite reduction and toxicity in
Saccharomyces cerevisiae. Eukaryotic Cell, 9(10): 1635-1647.
Pages, D., Rose, J., Conrod, S., Cuine, S., Carrier, P., Heulin, T. and Achouak, W., 2011.
Heavy Metal Tolerance in Stenotrophomonas maltophilia, Environmental Chemistry.
Apple Academic Press, pp. 128-136.
Pals, D.W. and Spry, P., 2003. Telluride mineralogy of the low-sulfidation epithermal
Emperor gold deposit, Vatukoula, Fiji. Mineralogy and Petrology, 79(3-4): 285-307.
Pals, D.W., Spry, P.G. and Chryssoulis, S., 2003. Invisible gold and tellurium in arsenic-rich
pyrite from the Emperor gold deposit, Fiji: implications for gold distribution and
deposition. Economic Geology, 98(3): 479-493.
Parnell, J., Bellis, D., Feldmann, J. and Bata, T., 2015. Selenium and tellurium enrichment in
palaeo-oil reservoirs. Journal of Geochemical Exploration, 148: 169-173.

f
oo
Parnell, J., Spinks, S. and Bellis, D., 2016. Low‐temperature concentration of tellurium and
gold in continental red bed successions. Terra Nova, 28(3): 221-227.
Parnell, J., Spinks, S. and Brolly, C., 2018. Tellurium and selenium in Mesoproterozoic red
pr
beds. Precambrian Research, 305: 145-150.
Pasero, M., 2020. The New IMA List of Minerals, available via http://cnmnc.main.jp/.
Perkins, W.T., 2011. Extreme selenium and tellurium contamination in soils—An eighty
e-
year-old industrial legacy surrounding a Ni refinery in the Swansea Valley. Science of
the Total Environment, 412: 162-169.
Pr

Pi , J., Yang, F., Jin, H., Huang, X., Liu, R. and Yang, P., 2013. Selenium nanoparticles
induced membrane biomechanical property changes in MCF-7 cells by disturbing
membrane molecules and F-actin. Bioorganic & Medicinal Chemistry Letters, 23:
6296–6303.
al

Piacenza, E., Presentato, A., Zonaro, E., Lampis, S., Vallini, G. and Turner, R.J., 2017.
Microbial-based bioremediation of selenium and tellurium compounds. Biosorption,
rn

published by IntechOpen, pp 117-147.


Pinel- Raffaitin, P., Pécheyran, C. and Amouroux, D., 2008. New volatile selenium and
tellurium species in fermentation gases produced by composting duck manure.
u

Atmospheric Environment, 42(33): 7786-7794.


Jo

Pons, T., Pic, E., Lequeux, N., Cassette, E., Bezdetnaya, L., Guillemin, F., Marchal, F. and
Dubertret, B., 2010. Cadmium- free CuInS2 /ZnS quantum dots for sentinel lymph node
imaging with reduced toxicity. ACS Nano, 4(5): 2531-2538.
Presentato, A., Piacenza, E., Anikovskiy, M., Cappelletti, M., Zannoni, D. and Turner, R.J.,
2016. Rhodococcus aetherivorans BCP1 as cell factory for the production of
intracellular tellurium nanorods under aerobic conditions. Microbial Cell Factories,
15(1): 204.
Presentato, A., Turner, R.J., Vásquez, C.C., Yurkov, V. and Zanno ni, D., 2019. Tellurite-
dependent blackening of bacteria emerges from the dark ages. Environmental
Chemistry, 16(4): 266-288.
Pudack, C., Halter, W., Heinrich, C.A. and Pettke, T., 2009. Evolution of magmatic vapor to
gold-rich epithermal liquid: The porphyry to epithermal transition at Nevados de
Famatina, northwest Argentina. Economic Geology, 104(4): 449-477.
Qin, H.-B., Takeichi, Y., Nitani, H., Terada, Y. and Takahashi, Y., 2017. Tellurium
distribution and speciation in contaminated soils from abandoned mine tailings:
comparison with selenium. Environmental Science & Technology, 51(11): 6027-6035.
Journal Pre-proof
Rajwade, J. and Paknikar, K., 2003. Bioreduction of tellurite to elemental tellurium by
Pseudomonas mendocina MCM B-180 and its practical application. Hydrometallurgy,
71(1-2): 243-248.
Ram, R., Vaughan, J., Etschmann, B. and Brugger, J., 2019. The aqueous chemistry of
polonium (Po) in environmental and anthropogenic processes. Journal of Hazardous
Materials, 380: 120725.
Ramadan, S.E., Razak, A.A., Ragab, A.M. and El-Meleigy, M., 1989. Incorporation of
tellurium into amino acids and proteins in a tellurium-tolerant fungi. Biological Trace
Element Research, 20(3): 225.
Ramos-Ruiz, A., Field, J.A., Wilkening, J.V. and Sierra-Alvarez, R., 2016. Recovery of
elemental tellurium nanoparticles by the reduction of tellurium oxyanions in a
methanogenic microbial consortium. Environmental Science & Technology, 50(3):
1492-1500.
Ramos-Ruiz, A., Sesma-Martin, J., Sierra-Alvarez, R. and Field, J.A., 2017a. Continuous
reduction of tellurite to recoverable tellurium nanoparticles using an upflow anaerobic
sludge bed (UASB) reactor. Water Research, 108: 189-196.

f
Ramos-Ruiz, A., Wilkening, J.V., Field, J.A. and Sierra-Alvarez, R., 2017b. Leaching of

oo
cadmium and tellurium from cadmium telluride (CdTe) thin- film solar panels under
simulated landfill conditions. Journal of Hazardous Materials, 336: 57-64.
Ray, G.E., Dawson, G.L. and Simpson, R., 1987. The Geology and Controls of Skarn
pr
Mineralization in the Hedley Gold Camp Southem British Columbia* (92W8, 82315),
Geological Fieldwork. BC Ministry of Energy, Mines and Petroleum Resources.
e-
Rea, M.A.D., Zammit, C.M. and Reith, F., 2016. Bacterial biofilms on gold grains—
implications for geomicrobial transformations of gold. FEMS Microbiology Ecology,
92(6): fiw082.
Pr

Rea, M.A.D., Wulser, P.-A., Brugger, J., Etschmann, B., Bissett, A., Shuster, J. and Reith, F.,
2018. Effect of physical and biogeochemical factors on placer gold transformation in
mountainous landscapes of Switzerland. Gondwana Research, 66: 77-92.
al

Reese, M.O., Glynn, S., Kempe, M.D., McGott, D.L., Dabney, M.S., Barnes, T.M., Booth, S.,
Feldman, D. and Haegel, N.M., 2018. Increasing markets and decreasing package
weight for high-specific-power photovoltaics. Nature Energy, 3(11): 1002.
rn

Reith, F., Lengke, M.F., Falconer, D., Craw, D. and Southam, G., 2007. The
geomicrobiology of gold. The ISME Journal, 1(7): 567-584.
u

Reith, F., Etschmann, B., Grosse, C., Moors, H., Benotmane, M.A., Monsieurs, P., Grass, G.,
Doonan, C., Vogt, S., Lai, B., Martinez-Criado, G., George, G.N., Nies, D.H.,
Jo

Mergeay, M., Pring, A., Southam, G. and Brugger, J., 2009. Mechanisms of gold
biomineralization in the bacterium Cupriavidus metallidurans. Proceedings of the
National Academy of Sciences, 106(42): 17757-17762.
Reith, F., Brugger, J., Zammit, C.M., Nies, D.H. and Southam, G., 2013. Geobiological
cycling of gold: from fundamental process understanding to exploration solutions.
Minerals, 3(4): 367-394.
Reith, F., Campbell, S., Ball, A., Pring, A. and Southam, G., 2014. Platinum in Earth surface
environments. Earth-Science Reviews, 131: 1-21.
Reith, F., Zammit, C.M., Shar, S.S., Etschmann, B., Bottrill, R., Southam, G., Ta, C., Kilburn,
M., Oberthür, T. and Ball, A.S., 2016. Biological role in the transformation of
platinum-group mineral grains. Nature Geoscience, 9(4): 294-298.
Reith, F., Nolze, G., Saliwan-Neumann, R., Etschmann, B., Kilburn, M.R. and Brugger, J.,
2019. Unravelling the formation histories of placer gold and platinum- group mineral
particles from Corrego Bom Successo, Brazil: A window into noble metal cycling.
Gondwana Research: 76, 246-259
Journal Pre-proof
Roberts, S., Palmer, M.R. and Waller, L., 2006. Sm-Nd and REE characteristics of
tourmaline and scheelite from the Bjorkdal gold deposit, northern Sweden: Evidence
of an intrusion-related gold deposit? Economic Geology, 101(7): 1415-1425.
Robie, R.A. and Hemingway, B.S., 1995. Thermodynamic Properties of Minerals and
Related Substances at 298.15 K and 1 bar (105 Pascals) Pressure and Higher
Temperatures. U.S. Geological Survey Bulletin, p. 2131.
Rollog, M., Cook, N.J., Guagliardo, P., Ehrig, K., Ciobanu, C.L. and Kilburn, M., 2019.
Detection of Trace Elements/Isotopes in Olympic Dam Copper Concentrates by
nanoSIMS. Minerals, 9(6): 336.
Rose-Weston, L., Brenan, J.M., Fei, Y., Secco, R.A. and Frost, D.J., 2009. Effect of pressure,
temperature, and oxygen fugacity on the metal-silicate partitioning of Te, Se, and S:
Implications for earth differentiation. Geochimica et Cosmochimica Acta, 73(15):
4598-4615.
Ruck, M. and Locherer, F., 2015. Coordination chemistry of homoatomic ligands of bismuth,
selenium and tellurium. Coordination Chemistry Reviews, 285: 1-10.
Sanyal, S.K., Shuster, J. and Reith, F., 2019. Cycling of biogenic elements drives

f
biogeochemical gold cycling. Earth-Science Reviews, 190: 131-147.

oo
Sen, S., Sharma, M., Kumar, V., Muthe, K., Satyam, P., Bhatta, U.M., Roy, M., Gaur, N.,
Gupta, S. and Yakhmi, J., 2009. Chlorine gas sensors using one-dimensional tellurium
nanostructures. Talanta, 77(5): 1567-1572.
pr
Sepahei, A.A. and Rashetnia, V., 2009. Tellurite Resistance and Reduction During Aerobic
and Anaerobic Growth of Bacteria Isolated from Sarcheshme Copper Mine. Iranian
e-
Journal of Environmental Health, Science and Engineering, 6(4): 253-260.
Shackleton, J.M., Spry, P.G. and Bateman, R., 2003. Telluride mineralogy of the golden mile
deposit, Kalgoorlie, Western Australia. The Canadian Mineralogist, 41(6): 1503-1524.
Pr

Shannon, R.D. and Prewitt, C.T., 1969. Effective ionic radii in oxides and fluorides. Acta
Crystallographica Section B, 25(5): 925-946.
Sharma, V.K., McDonald, T.J., Sohn, M., Anquandah, G.A., Pettine, M. and Zboril, R., 2015.
al

Biogeochemistry of selenium. A review. Environmental Chemistry Letters, 13(1): 49-


58.
Shibasaki, T., Abe, K. and Takeuchi, H., 1992. Recovery of tellurium from decopperizing
rn

leach solution of copper refinery slimes by a fixed bed reactor. Hydrometallurgy,


29(1-3): 399-412.
u

Shuster, J. and Reith, F., 2018. Reflecting on Gold Geomicrobiology Research: Thoughts and
Considerations for Future Endeavors. Minerals, 8(9): 401.
Jo

Shuva, M., Rhamdhani, M., Brooks, G., Masood, S. and Reuter, M., 2016. Thermodynamics
data of valuable elements relevant to e-waste processing through primary and
secondary copper production: a review. Journal of Cleaner Production, 131: 795-809.
Sillitoe, R.H. and Thompson, J.F.H., 1998. Intrusion–Related Vein Gold Deposits: Types,
Tectono ‐ Magmatic Settings and Difficulties of Distinction from Orogenic Gold
Deposits. Resource Geology, 48(4): 237-250.
Simmons, S.F. and Brown, K.L., 2006. Gold in magmatic hydrothermal solutions and the
rapid formation of a giant ore deposit. Science, 314(5797): 288-291.
Simon, G. and Essene, E.J., 1996. Phase relations among selenides, sulfides, tellurides, and
oxides; I, Thermodynamic properties and calculated equilibria. Economic Geology,
91(7): 1183-1208.
Singh, A.K. and Sharma, S., 2000. Recent developments in the ligand chemistry of tellurium.
Coordination Chemistry Reviews, 209(1): 49-98.
Smith, D.J., Naden, J., Jenkin, G.R.T. and Keith, M., 2017. Hydrothermal alteration and fluid
pH in alkaline-hosted epithermal systems. Ore Geology Reviews, 89: 772-779.
Soudi, M.R., Ghazvini, P.T.M., Khajeh, K. and Gharavi, S., 2009. Bioprocessing of seleno-
oxyanions and tellurite in a novel Bacillus sp. strain STG-83: A solution to removal of
Journal Pre-proof
toxic oxyanions in presence of nitrate. Journal of Hazardous Materials, 165(1-3): 71-
77.
Southam, G., Lengke, M.F., Fairbrother, L. and Reith, F., 2009. The biogeochemistry of gold.
Elements, 5(5): 303-307.
Stolz, J.F., 2017. Gaia and her microbiome. FEMS microbiology ecology, 93(2).
Subbotin, V., Vymazalová, A., Laufek, F., Savchenko, Y., Stanley, C., Gabov, D. and Plášil,
J., 2019. Mitrofanovite, Pt3 Te4 , a new mineral from the East Chuarvy deposit,
Fedorovo-Pana intrusion, Kola Peninsula, Russia. Mineralogical Magazine, 83(4):
523-530.
Sung, Y.-H., Ciobanu, C., Pring, A., Brügger, J., Skinner, W., Cook, N. and Nugus, M., 2007.
Tellurides from Sunrise Dam gold deposit, Yilgarn Craton, Western Australia: a new
occurrence of nagyágite. Mineralogy and Petrology, 91(3-4): 249-270.
Sung, Y.-H., Brugger, J., Ciobanu, C., Pring, A., Skinner, W. and Nugus, M., 2009. Invisible
gold in arsenian pyrite and arsenopyrite from a multistage Archaean gold deposit:
Sunrise Dam, Eastern Goldfields Province, Western Australia. Mineralium Deposita,
44(7): 765.

f
Tagami, K., Uchida, S., Ishii, N. and Zheng, J., 2013. Estimation of Te-132 distribution in

oo
Fukushima Prefecture at the early stage of the Fukushima Daiichi nuclear power plant
reactor failures. Environmental Science & Technology, 47(10): 5007-5012.
Takahashi, R., Matsueda, H., Okrugin, V.M., Shikazo no, N., Ono, S., Imai, A., Andreeva,
pr
E.D. and Watanabe, K., 2013. Ore ‐ forming Ages and Sulfur Isotope Study of
Hydrothermal Deposits in Kamchatka, Russia. Resource Geology, 63(2): 210-223.
e-
Takahashi, S., Kawashima, S., Hidaka, A., Tanaka, S. and Takahashi, T., 2019. Estimation of
the Release Time of Radio-Tellurium During the Fukushima Daiichi Nuclear Power
Pr

Plant Accident and Its Relationship to Individual Plant Events. Nuclear Technology,
205(5): 646-654.
Tan, L.C., Nancharaiah, Y.V., van Hullebusch, E.D. and Lens, P.N., 2016. Selenium:
environmental significance, pollution, and biological treatment technologies.
al

Biotechnology Advances, 34: 886–907.


Tanaka, M., Arakaki, A., Staniland, S.S. and Matsunaga, T., 2010. Simultaneously Discrete
rn

Biomineralization of Magnetite and Tellurium Nanocrystals in Magnetotactic Bacteria.


Applied and Environmental Microbiology, 76(16): 5526-5532.
Taylor, A., 1996. Biochemistry of tellurium. Biological Trace Element Research, 55(3): 231-
u

239.
Jo

Taylor, D.E., 1999. Bacterial tellurite resistance. Trends in microbiology, 7(3): 111-115.
Tenailleau, C., Pring, A., Etschmann, B., Brugger, J., Grguric, B. and Putnis, A., 2006.
Transformation of pentlandite to violarite under mild hydrothermal conditions.
American Mineralogist, 91(4): 706-709.
Theisen, J., Zylstra, G.J. and Yee, N., 2013. Genetic evidence for a molybdopterin-containing
tellurate reductase. Applied and Environmental Microbiology, 79(10): 3171-3175.
Thompson, T.B., Trippel, A.D. and Dwelley, P.C., 1985. Mineralized veins and breccias of
the Cripple Creek district, Colorado. Economic Geology, 80(6): 1669-1688.
Tomkins, A.G., Pattison, D.R. and Frost, B.R., 2006. On the initiation of metamorphic sulfide
anatexis. Journal of Petrology, 48(3): 511-535.
Tooth, B., Brugger, J., Ciobanu, C. and Liu, W., 2008. Modeling of gold scavenging by
bismuth melts coexisting with hydrothermal fluids. Geology, 36(10): 815-818.
Tooth, B., Ciobanu, C.L., Green, L., O’Neill, B. and Brugger, J., 2011. Bi- melt formation and
gold scavenging from hydrothermal fluids: An experimental study. Geochimica et
Cosmochimica Acta, 75(19): 5423-5443.
Tran, P.A. and Webster, T.J., 2013. Antimicrobial selenium nanoparticle coatings on
polymeric medical devices. Nanotechnology, 24(15): 155101.
Journal Pre-proof
Turner, R.J., 2001. Tellurite toxicity and resistance in Gram-negative bacteria. Recent
research developments in microbiology, pp. 69-77.
Turner, R.J., Borghese, R. and Zannoni, D., 2012. Microbial processing of tellurium as a tool
in biotechnology. Biotechnology Advances, 30(5): 954-963.
Ullah, H., Liu, G., Yousaf, B., Ali, M.U., Irshad, S., Abbas, Q. and Ahmad, R., 2018. A
comprehensive review on environmental transformation of selenium: recent advances
and research perspectives. Environmental Geochemistry and Health, 41(2), 1003-
1035.
U.S. Department of Energy, 2019. Cadmium Telluride. Office of Energy Efficiency and
Renewable Energy, Washington, DC, https://www.energy.gov/eere/solar/cadmium-
telluride.
USDOI, 2018. Final List of Critical Minerals 2018. Federal Register Archives, Document
number 2018-10667, pp. 23295-23296.
https://www.federalregister.gov/documents/2018/05/18/2018-10667/final- list-of-
critical- minerals-2018.
Vaigankar, D.C., Dubey, S.K., Mujawar, S.Y., D’Costa, A. and S.K, S., 2018. Tellurite

f
biotransformation and detoxification by Shewanella baltica with simultaneous

oo
synthesis of tellurium nanorods exhibiting photo-catalytic and anti-biofilm activity.
Ecotoxicology and Environmental Safety, 165: 516-526.
Van der Zee, F.P. and Cervantes, F.J., 2009. Impact and application of electron shuttles on
pr
the redox (bio)transformation of contaminants: A review. Biotechnology Advances,
27(3): 256-277.
e-
Vikentyev, I., 2006. Precious metal and telluride mineralogy of large volcanic-hosted
massive sulfide deposits in the Urals. Mineralogy and Petrology, 87(3-4): 305-326.
Vrionis, H.A., Wang, S., Haslam, B. and Turner, R.J., 2015. Selenite protection of tellurite
Pr

toxicity toward Escherichia coli. Frontiers in Molecular Biosciences, 2: 69.


Wadgaonkar, S.L., Mal, J., Nancharaiah, Y.V., Maheshwari, N.O., Esposito, G. and Lens,
P.N., 2018. Formation of Se(0), Te(0), and Se(0)–Te(0) nanostructures during
al

simultaneous bioreduction of selenite and tellurite in a UASB reactor. Applied


Microbiology and Biotechnology, 102(6): 2899-2911.
Wallier, S., Rey, R., Kouzmanov, K., Pettke, T., Heinrich, C.A., Leary, S., O’Connor, G.,
rn

Tamas, C.G., Vennemann, T. and Ullrich, T., 2006. Magmatic fluids in the breccia-
hosted epithermal Au-Ag deposit of Rosia Montana, Romania. Economic Geology,
u

101(5): 923-954.
Wallschläger, D. and Feldmann, J., 2010. 10: Formation, Occurrence, Significance, and
Jo

Analysis of Organoselenium and Organotellurium Compounds in the Environment,


Organometallics in Environment and Toxicology, pp. 319-364.
Wang, S., 2011. Tellurium, its resourcefulness and recovery. Jom, 63(8): 90.
Wang, T., Yang, L., Zhang, B. and Liu, J., 2010. Extracellular biosynthesis and
transformation of selenium nanoparticles and application in H2 O2 biosensor. Colloids
and Surfaces B: Biointerfaces, 80(1): 94-102.
Wang, X., Liu, G., Zhou, J., Wang, J., Jin, R. and Lv, H., 2011. Quinone- mediated reduction
of selenite and tellurite by Escherichia coli. Bioresource Technology, 102(3): 3268-
3271.
Wang, Z. and Becker, H., 2013. Ratios of S, Se and Te in the silicate Earth require a volatile-
rich late veneer. Nature, 499(7458): 328.
Watanabe, T., Tsuchiya, N., Oura, Y., Ebihara, M., Inoue, C., Hirano, N., Yamada, R.,
Yamasaki, S.- i., Okamoto, A. and Nara, F.W., 2012. Distribution of artificial
radionuclides (110mAg, 129m Te, 134 Cs, 137 Cs) in surface soils from Miyagi Prefecture,
northeast Japan, following the 2011 Fukushima Dai- ichi nuclear power plant accident.
Geochemical journal, 46(4): 279-285.
Journal Pre-proof
Watterson, J.R., Gott, G.B., Neuerburg, G.J., Lakin, H.W. and Cathrall, J.B., 1977. Tellurium,
a Guide to Mineral Deposits. In: C.R.M. Butt and I.G.P. Wilding (Editors),
Developments in Economic Geology. Elsevier, pp. 31-48.
Wedepohl, K.H., 1995. The composition of the continental crust. Geochimica et
Cosmochimica Acta, 59(7): 1217-1232.
Wei, T.Y., Chang, H.Y., Lee, Y.F., Hunga, Y.L. and Huang, C.C., 2011. Selective Tellurium
Nanowire‐based Sensors for Mercury (II) in Aqueous Solution. Journal of the Chinese
Chemical Society, 58(6): 732-738.
Weil, M., 2018. Crystal structures of the triple perovskites Ba2 K2 Te2 O9 and Ba2 KNaTe2 O9 ,
and redetermination of the double perovskite Ba 2 CaTeO 6 . Acta Crystallographica
Section E, 74(7): 1006-1009.
White, C., Wilkinson, S.C. and Gadd, G.M., 1995. The role of microorganisms in biosorption
of toxic metals and radionuclides. International Biodeterioration & Biodegradation,
35(1-3): 17-40.
William, W.Y., Chang, E., Drezek, R. and Colvin, V.L., 2006. Water-soluble quantum dots

f
for biomedical applications. Biochemical and biophysical research communications,

oo
348(3): 781-786.
Williams, P.A., 1990. Oxide zone geochemistry. Ellis Horwood Limited, pp. 286.
Woodhouse, M., Goodrich, A., Margolis, R., James, T., Dhere, R., Gessert, T., Barnes, T.,
pr
Eggert, R. and Albin, D., 2013. Perspectives on the pathways for cadmium telluride
photovoltaic module manufacturers to address expected increases in the price for
tellurium. Solar Energy Materials and Solar Cells, 115: 199-212.
e-
Wray, D.S., 1998. The impact of unconfined mine tailings and anthropogenic pollution on a
semi-arid environment–an initial study of the Rodalquilar mining district, south east
Pr

Spain. Environmental Geochemistry and Health, 20(1): 29-38.


Wu, X., 2004. High-efficiency polycrystalline CdTe thin- film solar cells. Solar Energy, 77(6):
803-814.
Xu, W., Zhao, J., Brugger, J., Chen, G. and Pring, A., 2013. Mechanism of mineral
al

transformations in krennerite, Au3 AgTe8 , under hydrothermal conditions. American


Mineralogist, 98(11-12): 2086-2095.
rn

Xu, Y., Li, J., Tan, Q., Peters, A.L. and Yang, C., 2018. Global status of recycling waste solar
panels: A review. Waste Management, 75: 450-458.
Yeh, K.-W., Huang, T.-W., Huang, Y.- l., Chen, T.-K., Hsu, F.-C., Wu, P.M., Lee, Y.-C., Chu,
u

Y.-Y., Chen, C.-L. and Luo, J.-Y., 2008. Tellurium substitution effect on
Jo

superconductivity of the α-phase iron selenide. Europhysics Letters, 84(3): 37002.


Yoschenko, V., Ohkubo, T. and Kashparov, V., 2018. Radioactive contaminated forests in
Fukushima and Chernobyl. Journal of Forest Research, 23(1): 3-14.
Yu, H., Young, J., Wu, H., Zhang, W., Rondinelli, J.M. and Halasyamani, P.S., 2016.
Electronic, crystal chemistry, and nonlinear optical property relationships in the
dugganite A3 B3 CD2 O14 family. Journal of the American Chemical Society, 138(14):
4984-4989.
Yu, M.-Z., Chen, X.-G., Garbe-Schönberg, D., Ye, Y. and Chen, C.-T.A., 2019. Volatile
Chalcophile Elements in Native Sulfur from a Submarine Hydrothermal System at
Kueishantao, Offshore NE Taiwan. Minerals, 9(4): 245.
Yurkov, V., Jappe, J. and Vermeglio, A., 1996. Tellurite resistance and reduction by
obligately aerobic photosynthetic bacteria. Applied and Environmental Microbiology,
62(11): 4195-4198.
Zare, B., Sepehrizadeh, Z., Faramarzi, M.A., Soltany ‐ Rezaee ‐ Rad, M., Rezaie, S. and
Shahverdi, A.R., 2014. Antifungal activity of biogenic tellurium nanoparticles against
Candida albicans and its effects on squalene monooxygenase gene expression.
Biotechnology and Applied Biochemistry, 61(4): 395-400.
Journal Pre-proof
Zannoni, D., Borsetti, F., Harrison, J.J. and Turner, R.J., 2007. The bacterial response to the
chalcogen metalloids Se and Te. Advances in Microbial Physiology, Vol. 53, pp.1-
312.
Zeng, C., Ramos-Ruiz, A., Field, J.A. and Sierra-Alvarez, R., 2015. Cadmium telluride
(CdTe) and cadmium selenide (CdSe) leaching behavior and surface chemistry in
response to pH and O2. Journal of Environmental Management, 154: 78-85.
Zhang, X. and Spry, P.G., 1994. Calculated stability of aqueous tellurium species, calaverite,
and hessite at elevated temperatures. Economic Geology, 89(5): 1152-1166.
Zhao, J., Brugger, J., Grundler, P.V., Xia, F., Chen, G. and Pring, A., 2009. Mechanism and
kinetics of a mineral transformation under hydrothermal conditions: Calaverite to
metallic gold. American Mineralogist, 94(11-12): 1541-1555.
Zhao, J., Xia, F., Pring, A., Brugger, J., Grundler, P.V. and Chen, G., 2010. A novel pre-
treatment of calaverite by hydrothermal mineral replacement reactions. Minerals
Engineering, 23(5): 451-453.
Zhao, J., Brugger, J., Xia, F., Ngothai, Y., Chen, G. and Pring, A., 2013. Dissolution-
reprecipitation vs. solid-state diffusion: Mechanism of mineral transformations in

f
sylvanite,(AuAg)2 Te4 , under hydrothermal conditions. American Mineralogist, 98(1):

oo
19-32.
Zheng, Y., Gao, S. and Ying, J.Y., 2007. Synthesis and cell ‐ imaging applications of
pr
glutathione‐capped CdTe quantum dots. Advanced Materials, 19(3): 376-380.
Zientek, M.L., Fries, T.L. and Vian, R.W., 1990. As, Bi, Hg, S, Sb, Sn and Te geochemistry
e-
of the J-M Reef, Stillwater Complex, Montana: constraints on the origin of PGE-
enriched sulfides in layered intrusions. Journal of Geochemical Exploration, 37(1):
51-73.
Pr

Zonaro, E., Lampis, S., Turner, R.J., Qazi, S.J.S. and Vallini, G., 2015. Biogenic selenium
and tellurium nanoparticles synthesized by environmental microbial isolates
efficaciously inhibit bacterial planktonic cultures and biofilms. Frontiers in
al

Microbiology, 6: 584.
Zonaro, E., Piacenza, E., Presentato, A., Monti, F., Dell’Anna, R., Lampis, S. and Vallini, G.,
2017. Ochrobactrum sp. MPV1 from a dump of roasted pyrites can be exploited as
rn

bacterial catalyst for the biogenesis of selenium and tellurium nanoparticles.


Microbial Cell Factories, 16(1): 215.
u
Jo
Journal Pre-proof
TABLES

Table 1: Selected Te minerals (including wt% to nearest % of Te for tellurides). A full table
of Te minerals is presented in Supplementary Table 1.

Mineral name Ideal chemical formula Crystal Ideal Wt%


symmetry Te*
PRIMARY (TELLURIDES)
Altaite PbTe Isometric 38 %
Calaverite AuTe2 Monoclinic 56 %
Cervelleite Ag4 TeS Monoclinic 22 %
Coloradoite HgTe Isometric 39 %

f
Empressite AgTe Orthorhombic 54 %

oo
Frohbergite FeTe2 Orthorhombic 82 %
Goldfieldite Cu10 Te4 S13 Isometric 33 %
Hessite
Kawazulite
Ag2 Te
Bi2 Te2 Se
pr Monoclinic
Trigonal
37 %
34 %
e-
Kostovite CuAuTe4 Orthorhombic 66 %
Pr

Krennerite Au3 AgTe8 Orthorhombic 59 %


Melonite NiTe2 Trigonal 81 %
Montbrayite (Au,Ag,Sb,Bi,Pb)23 - Triclinic 45-50 %
al

(Te,Sb,Bi,Pb)38
Muthmannite AuAgTe2 Monoclinic 46 %
rn

Nagyágite [Pb3 (Pb,Sb)3 S6 ](Au,Te)3 Monoclinic ~30 %


u

Petzite Ag3 AuTe2 Isometric 33 %


Jo

Rickardite Cu7 Te5 Orthorhombic 59 %


Rucklidgeite PbBi2 Te4 Trigonal 45 %
Stützite Ag5-xTe3 ; Hexagonal
43 %
0.24 ≤ x ≤ 0.36
Sylvanite (Au,Ag)2 Te4 Monoclinic 60-61 %
Tellurobismuthite Bi2 Te3 Trigonal 48 %
Tetradymite Bi2 Te2 S Trigonal 36 %
(ELEMENTAL)
Tellurium Te Trigonal
SECONDARY (TELLURITES)
Emmonsite Fe3+2 (Te4+O 3 )3 ·3H2 O Triclinic
3+
Mackayite Fe (Te4+2 O 5 )(OH) Tetragonal
Journal Pre-proof
Magnolite Hg1+2 Te4+O3 Orthorhombic
Moctezumite Pb(UO 2 )(Te4+O3 )2 Monoclinic
Paratellurite α-Te4+O 2 Tetragonal
Spiroffite Mn2+2 Te4+3 O 8 Monoclinic
Teineite CuTe4+O 3 ·2H2 O Orthorhombic
Tellurite β-Te4+O 2 Orthorhombic
3+ 4+
Zemannite Mg0.5 ZnFe (Te O3 )3 Hexagonal
·(3+n)H2 O; 0 ≤ n ≤ 1.5
(TELLURATES)
Burckhardtite Pb2 (Fe3+Te6+)[AlSi3 O 8 ]O6 Trigonal
6+
Dugganite Pb3 Zn3 (AsO 4 )2 (Te O 6 ) Trigonal

f
Jensenite Cu3 Te6+O6 ·2H2 O Monoclinic

oo
6+
Khinite PbCu3 (Te O 6 )(OH)2 Orthorhombic
6+
Markcooperite Pb2 UO2 (Te O6 ) Monoclinic
Mcalpineite 6+
Cu3 (Te O6 )
6+
pr Isometric
e-
Ottoite Pb2 Te O5 Monoclinic
6+
Quetzalcoatlite Zn6 Cu3 (Te O6 )2- Trigonal
Pr

(OH)6 ·AgxPbyClx+2y
Timroseite Pb2 Cu5 (Te6+O6 )2 (OH)2 Orthorhombic
Xocolatlite Ca2 Mn4+2 (Te6+O6 )2 ·H2 O Monoclinic
al

(MIXED-VALENCE)
Carlfriesite CaTe4+2 Te6+O8 Monoclinic
rn

Tlapallite (Ca,Pb)3 CaCu6 - Trigonal


[Te4+3 Te6+O 12 ]2-
u

(Te4+O3 )2 (SO 4 )2 ·3H2 O


Jo
Journal Pre-proof
Table 2: Quantification of amounts of Te associated with common sulphides. Some
measurements are averages across a range of samples, others particular to a specific locality.

Ideal Te content
Mineral name chemical Detail/locality average Reference
formula (mg/kg)
Covellite CuS Average Up to 430 (Dill, 2010)

Chalcopyrite CuFeS2 Average Up to 70 (Dill, 2010)


(George et al.,
Average <1
2018)
Broken Spur vent, (Butler and
45
Mid-Atlantic Ridge Nesbitt, 1999)
Baita Bihor; (George et al.,
1.9

f
epithermal 2018)

oo
Kanmantoo;
(George et al.,
metamorphosed 1.9
2018)
sulphide ore

Galena PbS Average


pr Up to 200 (Dill, 2010)
e-
Pyrite FeS2 Average Up to 70 (Dill, 2010)
Pr

Volcanogenic
massive sulphide
Up to 860, often (Belousov et al.,
average, Yilgarn
less than 2 2016)
Craton, Western
al

Australia
Orogenic average,
rn

Up to 710, often (Belousov et al.,


Yilgarn Craton,
less than 100 2016)
Western Australia
u

Pyrite,
arsenian pyrite FeS2 ,
Jo

(Keith et al.,
and Fe(S,As)2 Carlin average 573
2018)
arsenopyrite and FeAsS
compilation
Epithermal high
(Keith et al.,
sulphidation 343
2018)
average
Epithermal low
(Keith et al.,
sulphidation 600
2018)
average
Epithermal alkaline (Keith et al.,
546
average 2018)
(Keith et al.,
Orogenic average 306
2018)
(Keith et al.,
Porphyry average 26.3
2018)
Pyrite, As- Dongping, China, 4.0-20823 (i.e. (Cook et al.,
FeS2
deficient alkaline epithermal microinclusions 2009a)
Journal Pre-proof
of tellurides)
Huangtuliang, 3.0-18285 (i.e.
(Cook et al.,
China, alkaline microinclusions
2009a)
epithermal of tellurides)
Hougou, China, (Cook et al.,
2.6-19.6
alkaline epithermal 2009a)

Pyrrhotite Fe1-xS Average Up to 60 (Dill, 2010)

(Cook et al.,
Sphalerite ZnS Average < 0.05
2009b)
Magura epithermal
(Cook et al.,
Au deposit, 16-665
2009b)
Romania

f
oo
pr
e-
Pr
al
u rn
Jo
Journal Pre-proof

Table 3: Concentrations of dissolved Te in geothermal and hydrothermal fluids.

Te
Temperature
Locality concentration Comments Reference
(˚C)
(µg/L)
Three samples
Reykjanes geothermal with chlorinity
(Hardardóttir
system, Iceland, at 284-295 16.5 to 18.5 close to that of
et al., 2009)
1350– 1500 m depth seawater (0.51-
0.53 M).
Lihir geothermal
Lihir water also
system, Papua New (Simmons and
260 Up to 4 contained up to
Guinea, at depth Brown, 2006)
13 µg/L Au.
550 m
Wairakei geothermal
Used as a field (Simmons and
 255

f
system, New Zealand, Up to 0.4
blank sample Brown, 2006)

oo
at depth 950 m
Based on
Pacmanus seafloor hot extrapolations
springs field, Papua from samples (Binns et al.,
New Guinea, at depth
130 m
 55 pr
3 to 18
heavily
contaminated by
2004)
e-
seawater (88%)
233-255 Liquid-vapour
Up to 14,000
(homogenisation T) inclusion
Pr

Au–Te epithermal Unique liquid-


358 340,000 (Wallier et al.,
system of Roşia vapour inclusion
2006)
Montană, Romania Two co-existing
al

5,500 and low-density


ND
180,000 vapour
rn

inclusions
The final stage
Porphyry Cu–Mo–Au of this
stage and transitional
u

mineralizing
quartz–sericite–pyrite (Pudack et al.,
278  23 Up to 670,000 system is a Te-
Jo

stage of the Nevados de 2009)


rich high
Famatina deposit
sulphidation
(northwest Argentina) system
High-
temperature gas
High temperatures condensate
fumaroles at the contained ca. (Okrugin et
Avacha Volcano,  600 Up to 16,000
100 times more al., 2017)
Kamchatka Peninsula, Te than low-
Far East Russia
temperature
condensates
Te in scarcely
hydrothermally
High-sulphidation
altered rocks in
hydrothermal system (Fulignati and
250–520 5000-19,000 correspondence
of the La Fossa volcano Sbrana, 1998)
to the highest-
(Vulcano, Italy) temperature
fumarolic vents
Journal Pre-proof

Table 4: Thermodynamic properties for Te aqueous species, minerals, and vapor species adapted from selected references.

Temperature [ºC]
Reaction Reference
25 50 60 100 150 200 250 300 350
Aqueous Species
H2 Te(aq) + 1.5O2 (g) = H2 TeO3 (aq) 99.06 90.59 87.57 77.09 66.81 58.72 52.21 46.87 42.46 McPhail (1995)
HTe- + H+ + 1.5O2 (g) = H2 TeO3 (aq) 101.70 93.30 90.32 80.10 70.24 62.67 56.75 52.16 48.88 McPhail (1995)
Te2- + 2H+ + 1.5O2 (g) = H2 TeO3 (aq)
Te2 2- + 2H+ + H2 O(l) + 2.5O2 (g) = 2H2 TeO3 (aq)
H3 TeO3 + = H2 TeO3 (aq) + H+ (log K a0 )
113.86
151.77
–2.79
105.04
139.20
–2.43
101.90
139.74
–2.31
91.45
119.47
–1.88
81.36
104.79
–1.46
o f
73.85
93.55
–1.14
68.14
84.83
–0.87
63.89
78.14
–0.62
61.31
73.54
–0.32
McPhail (1995)
McPhail (1995)

o
Grundler et al. (2013)

r
H2 TeO3 (aq) = HTeO3 - + H+ (log Ka1 ) –5.22 –5.57 –5.69 –6.07 –6.39 –6.61 –6.84 –7.17 –7.98 Grundler et al. (2013)
HTeO3 - = TeO3 2- + H+ (log K a2 ) –10.07 –9.89 –9.83 –9.58 –9.29 –9.04 –8.86 –8.84 –9.25

p
Grundler et al. (2013)

-
H6 TeO6 (aq) = H2 TeO3 (aq) + 2H2 O(l) + 0.5O2 (g) -2.77 -1.97 -1.67 -0.58 0.57 1.56 2.42 3.17 3.81 McPhail (1995)
H5 TeO6 - + H+ = H6 TeO6 (aq)

e
7.70 7.40 7.31 7.01 6.81 6.76 6.85 7.14 7.73 McPhail (1995)

r
H4 TeO6 2- + H+= H5 TeO6 - 10.95 10.57 10.46 10.16 10.05 10.12 10.37 10.84 11.72 McPhail (1995)
Minerals
Te(s) + H2 O(l) + O2 (g) = H2 TeO3 (aq)
TeO2 (s) + H2 O(l) = H2 TeO3 (aq) (log K s )
42.6
–4.61

l
38.6
–4.21
P37.2
–4.07
32.3
–3.59
27.4
–3.14
23.5
–2.77
20.5
–2.44
18.0
–2.07
16.2
–1.53
Robie & Hemingway (1995)
Grundler et al. (2013)
AuTe2 (s) (calaverite) + 1.5H2 O(l) + 2.25O2 (g) + H+
= 2H2 TeO3 (aq) + Au +

n a74.4 67.5 65.0 56.4 47.9 41.3 36.0 31.8 28.8 Mills (1974)

ur
+ + Mills (1974), Afifi et al.
Ag 2 Te(s) (hessite) + 2H + 1.5O2 (g) = 2Ag + 2H2 TeO3 (aq) 49.8 45.4 43.7 38.2 32.9 28.8 25.5 22.9 21.1
(1988)

o
Echmaeva & Osadchii,
Ag 3 AuTe2 (s) (petzite) + 4H+ + 3O2 (g)
86.8 78.9 76.1 66.3 56.9 49.4 43.6 – – ( 2009), from Grundler et al.

J
= 3Ag + + Au + + 2H2 TeO3 (aq)
( 2013)*
Gaseous species
H2 Te(g) + 1.5O2 (g) = H2 TeO3 (aq) 98.1 88.9 85.7 74.4 63.3 54.7 47.7 41.9 (37.1) McPhail (1995)
Te2 (g) + 2H2 O(l) + 2O2 (g) = 2H2 TeO3 (aq) 102.9 92.9 89.3 77.0 64.9 55.3 47.6 41.2 (35.9) McPhail (1995)
H2 TeO3 (aq) = TeO2 (g) + H2 O(l) –31.1 –28.0 –26.8 –22.9 –19.0 –16.0 –13.5 –11.6 –10.3 Barin (1995)
H2 TeO3 (aq) = TeO2 (H2 O)(g) –20.5 –17.6 –16.5 –13.1 –9.90 –7.56 –5.83 –4.58 –3.81 Grundler et al. (2013)
H2 TeO3 (aq) + H2 O(l) = TeO2 .2(H2 O)(g) –38.5 –31.4 –29.1 –21.3 –14.7 –10.3 –7.31 –5.31 –4.14 Grundler et al. (2013)
* Grundler et al. (2013) have a typo in their Table 13 – the properties for petzite are listed using O 2 (aq), rather than O 2 (g).
Journal Pre-proof

Table 5: Microbiological identities, locations, and produced Te nanoparticle morphology for selected Te oxyanion-reducing forms of bacteria

Product of reduction,
Genus and species, with strain Taxonomic
Location collected (if applicable) nanostructure size in nm Reference
or enzyme (if applicable) designation
if applicable
Aeromonas caviae ST Gram-negative rod Previously isolated Te0 (Castro et al., 2008)
0
Aspergillus welwitschiae Sources such as rotting garlic, kitchen Te nanospheres and
Fungus (Abo Elsoud et al., 2018)

f
sink slime nanoellipsoids (~60)
Gram-positive rod, Sediment slurries from Mono lake, Te0 nanospheres and

o
Bacillus beveridgei (Baesman et al., 2009)
facultative anaerobe California, USA nanorods
Bacillus filicolonicus, B.
laterosporus
Gram-positive rods,
facultative anaerobes
Sarcheshme copper mine in Kerman
Province, Iran

r o Te0 nanospheres (Sepahei and Rashetnia, 2009)

p
Mud samples from salt marsh bordering
Intracellular Te0 nanorods

-
Bacillus halodenitrificans Gram-positive rod,
the Indian River inlet, in Rehoboth Beach, (Ollivier et al., 2008)
facultative anaerobe (<100)

e
Delaware, USA

Bacillus pumilis
Gram-positive rod,
facultative anaerobe
r
Anzali Lagoon in Gilan province, Iran
and the Neidasht spring in the north of

P
Iran
Te0 ; volatilisation of
unspecified gases
(Soudi et al., 2009)

Bacillus selenitireducens
Gram-positive rod,
facultative anaerobe l
Sediment slurries from Mono lake,

a California, USA
Extracellular Te 0 nanorods
(<200) clustered into larger
rosettes
(Baesman et al., 2006;
Baesman et al., 2007)

n
ur
Bacillus species, Enzyme STG- Gram-positive rod, Neidasht spring in the north of Iran (also
Te0 (Etezad et al., 2009)
83 facultative anaerobe see above)
Gram negative
Extracellular Se0 –Te0
Duganella violacienigra

J o
heterotrophic non-
halophilic and aerobic
proteobacterium
Obligate aerobic
Se-rich soils from agricultural fields in
India
composite nanoparticles
(~100)
(Bajaj and Winter, 2014)

Erythrobacter litoralis photosynthetic Previously isolated Intracellular Te0 (Yurkov et al., 1996)
bacterium
Obligate aerobic Intracellular Te0 or resistant
Erythromicrobium ezovicum photosynthetic Previously isolated but no apparent reduction, (Yurkov et al., 1996)
bacterium depending on carbon source
Erythromicrobium
Obligate aerobic
hydrolyticum, E. ramosum, E. Previously isolated Intracellular Te0 (Yurkov et al., 1996)
photosynthetic bacteria
sibiricum, E. ursincola
Magnetospirillum magneticum Facultative anaerobic Coprecipitated Te 0 and
Previously isolated (Tanaka et al., 2010)
strain AMB-1 magnetotactic spiral magnetite nanocrystals
Journal Pre-proof

Drainage water from a metal refining Extracellular Te 0 (100) and


Ochrobactrum anthropi TI-2
Gram-negative rod(s) plant in Amagasaki City, Hyogo DMTe, DMDTe and (Kagami et al., 2012)
and TI-3
Prefecture, Japan DMTeS
Gram-negative rod,
Dump of roasted (arseno)pyrites from
Ochrobactrum sp. MPV1 Aerobic Intracellular Te0 (Zonaro et al., 2017)
near a factory in Tuscany, Italy
α-proteobacterium
Sediment from the Er-Jen River in Tainan
Paenibacillus Sp. Strain TeW Gram-positive rod County, Taiwan, contaminated by run-off Te0 (Chien and Han, 2009)

f
from electroplating factories and smelters
Pseudoalteromonas sp. strain Gram-negative aerobic
EPR3 rods
Previously isolated from volcanic vents

o
Te0 ; DMTe (Bonificio and Clarke, 2014)

ro
Gram-negative
Pseudomonas Kesterson Reservoir in the San Joaquin
fluorescens K27 facultative anaerobic Te0 ; DMTe and DMDTe (Basnayake et al., 2001)
Valley of California, USA

-p
rod
Pseudomonas mendocina MCM
Gram-negative rod Previously isolated Te0 (Rajwade and Paknikar, 2003)

re
B-180
Photosynthetic Gram-
Intracellular Te0 nanorods
Rhodobacter capsulatus negative anaerobic α-

P
Previously isolated (Borghese et al., 2016)
(200-700)
proteobacterium
Rhodococcus aethivorans BCP1
Aerobic Gram-positive
sphere

a l
Previously isolated
Intracellular Te0 non-
aggregated nanorods (<700)
(Presentato et al., 2016)

n
Mud samples from salt marsh bordering
Obligately aerobic,

ur
Rhodotorula mucilaginosa the Indian River inlet, in Rehoboth Beach, Te0 (Ollivier et al., 2008)
Gram-positive rods
Delaware, USA
Intracellular Te0 or resistant

o
Obligately aerobic
Roseococcus thiosulfatophilus Previously isolated but no apparent reduction, (Yurkov et al., 1996)
photosynthetic sphere

Salinicoccus iranensis
J
Gram-positive
halophilic sphere
Gram-positive
Previously isolated
Salty environments including saline or
depending on carbon source
Te0 (Alavi et al., 2014)

Salinicoccus sp. QW6 hypersaline soils, hypersaline brackish Intracellular Te0 (Amoozegar et al., 2008)
halophilic sphere
water and textile factory effluents of Iran
Facultative Gram- Water from the Zuari Estuary, Goa state, Te0 nanorods (8-75
Shewanella baltica (Vaigankar et al., 2018)
negative anaerobic rod India diameter)
Kim: Intracellular Te 0
Facultative Gram- nanorods (100-200); (Kim et al., 2012a; Klonowska
Shewanella oneidensis MR-1 Previously isolated
negative anaerobic rod Klonowska: also et al., 2005)
intracellular (size
Journal Pre-proof

unspecified)
Pages: Te0 intracellularly;
Pages: Isolated from contaminated
Kagami: Te0 intracellularly
Pseudomonas cultures; Kagami: Drainage
Stenotrophomonas maltophilia and extracellularly (200- (Kagami et al., 2012; Pages et
Gram-negative rod water from a metal refining plant in
300) and DMTe and al., 2011; Zonaro et al., 2015)
Amagasaki City, Hyogo Prefecture,
DMDTe; Zonaro: Te0 (75-
Japan; Zonaro: Environmental isolates
80)
Extracellular Te0

f
Sulphurospirillum barnesii Sediment slurries from Mono lake, nanospheres (<50)
Anaerobic spiral (Baesman et al., 2006)
California, USA coalescing into larger

o ocomposites

r
Abbreviations: Te0 – solid elemental tellurium, DMTe – gaseous dimethyl telluride, DMDTe – gaseous dimethyl ditelluride and DMTeS –

p
-
gaseous dimethyl tellurenyl sulphide. Ordered by scientific name.

r e
l P
n a
u r
J o
Journal Pre-proof

FIGURES AND CAPTIONS

(a) (b)

f
oo
(c) (d)
pr
e-
Pr
al

(e) (f)
u rn
Jo

Figure 1: Visual and chemical diversity of Te minerals. Three primary minerals: (a)
Calaverite (AuTe2 ) from Cripple Creek, Colorado, USA, field of view (FOV) 1.3 mm, (b)
Krennerite (Au3 AgTe8 ) also from Cripple Creek, FOV 3 mm and (c) Sylvanite [(Au,Ag) 2 Te]
from Emperor Mine, Viti Levu, Fiji, FOV 5 mm. (d) One example of elemental Te with the
wsecondary mineral, teineite (CuTe4+O3 ·2H2 O) from Teine Mine, Japan, FOV 3 mm. Finally,
two secondary minerals: (e) Zemannite [Mg0.5 ZnFe3+(Te4+O3 )3 ·(3+n)H2 O)], from Moctezuma
mines, Mexico, FOV 2 mm and (f) Jensenite (Cu2+3 Te6+O6 ·2H2 O) from Centennial Eureka
Mine, Utah, USA, FOV 1 mm. Images credit: Mr. Brent Thorne (Utah, USA, private
collection).
Journal Pre-proof

f
oo
pr
e-
Pr
al
u rn
Jo

Figure 2: a) World consumption of Te, showing the dramatic growth in Te usage over the
past decade in thermoelectric devices and photoreceptors, with cadmium telluride (CdTe)
solar panels one of the main drivers. a) Reproduced from Nuss (2019) with permission from
CSIRO Publishing. b) One of the many new CdTe solar panel arrays: a solar-powered pump
in the village of Angarf, Ouarzazate Province, Morocco. Image credit: Mr. Georges Favreau
(Aix-en-Provence, France, private collection).
Journal Pre-proof

f
oo
pr
e-
Pr
al
u rn
Jo

Figure 3: The wide variety of bonding Te participates in contributes to its rich surface
chemistry. Te participates in intermetallic bonding between elements like Ag, Au, and other
Te atoms in primary minerals. Te also forms a variety of arrangements with oxygen, with
different bonding modes for Te4+ and Te6+. The bonding arrangements between Te and O are
further explored in Figure 4.
Journal Pre-proof

o f
r o
- p
r e
l P
n a
u r
J o
Figure 4: Various Te–O bonding networks as illustrated by Te4+ minerals (see Supp. table 1 for formulae). The bonding networks vary from
simple, finite units to complex, three-dimensional structures. All structures except cyclo are known in minerals, with the simplest neso
compounds forming the largest number of structures across both Te4+ and Te6+ minerals. Although other ligands may bind to Te 4+ and Te6+ in
synthetic contexts, in minerals, all strong bonds are to oxygen atoms.
Journal Pre-proof

f
oo
pr
e-
Pr
al
u rn
Jo

Figure 5: Tellurium environmental cycling schematic, bringing together the various sinks of
Te (highlighted in green) and the processes and transformations which occur between
different Te reservoirs (highlighted in pink). Particularly high concentrations of Te are found
in volcanogenic sulphur, some ore deposits and in Cu conce ntrates and anode slimes during
some Cu processing. The complex biological interactions shown in the top part of the
diagram are further explored in Section 6. a (Anders and Grevesse, 1989); b (McDonough and
Sun, 1995); c (Hattori et al., 2002); d (Hu and Gao, 2008); e (Dill, 2010); f (Maslennikov et al.,
2013); g (Hein et al., 2003); h (Goldfarb et al., 2017); i (Makuei and Senanayake, 2018); j
(Kavlak and Graedel, 2013); k (Nuss, 2019); l (Filella et al., 2019); m (Belzile and Chen,
2015); and n (Ba et al., 2010).
Journal Pre-proof

f
oo
pr
e-
Pr
al
rn
u
Jo
Journal Pre-proof

o f
r o
- p
r e
l P
n a
u r
J o
Figure 6: World map showing locations of selected tellurium-enriched mineral occurrences, by deposit type. The deposit types shown here are
classified as follows: volcanogenic massive sulfide (VMS), iron oxide-copper-gold (IOCG), orogenic gold, porphyry, epithermal, skarn, Carlin-
type gold, magmatic copper–nickel–platinum- group metal (Cu-Ni-PGM), and other deposit types (43, 46 and 50, Te-rich waste from massive
sulphide ores; 79 and 81, Intrusion-related Au deposit). Adapted from mindat.org, Goldfarb et al. (2017) and Keith et al. (2018), using the crust-
type world map template of the USGS (https://earthquake.usgs.gov/data/crust/type.html).
Journal Pre-proof

f
oo
Figure 7: Solubility of various Te species at ~1 (solid red lines) and 100 ppb levels (dashed
pr
red lines) as a function of pH and oxygen fugacity at (a) 25 and (b) 300 °C and water-
saturated pressures, using thermodynamic properties collected by McPhail (1995) and
e-
Grundler et al. (2013). The Fe diagram in a Fe-S-Cl-O-H system is drawn as blue dashed
Pr

lines for reference (conditions listed in the legend inset). The S diagram is shown as purple
dashed lines. Abbreviations in top two diagrams: py - pyrite; hm - hematite; mt - magnetite;
po – pyrrhotite. Horizontal dotted lines at the top and bottom of the two diagrams indicate the
al

stability field of H2 O.
u rn
Jo
Journal Pre-proof

f
oo
pr
e-
Pr

Figure 8: Solubility of tellurite (TeO 2 ) at 25˚C calculated at ionic strength = 1, using the
properties of Grundler et al. (2013) and Filella and May (2019) using Geochemist’s
al

Workbench (Bethke, 2008). The tellurite solubilities measured by Grundler et al. (2013) are
rn

shown using circles; unfilled circles correspond to solutions of buffer-only (ionic strength
<0.5), and filled circles to solutions with ~1 m NaCl and ~0.2 m buffer.
u
Jo
Journal Pre-proof

f
oo
pr
e-
Pr
al
u rn
Jo

Figure 9: Formation of secondary Te minerals at the young (6.9–7.1 My) Aginskoe deposit,
in the active volcanic arc in the Kamchatka Peninsula, Russia. (A) Rich assemblage of
secondary Te minerals are found in a restricted part of the oxidation zone. (B) Weathering of
a chalcopyrite grain containing petzite inclusion. The rim (labelled o-cpy) consists of poorly
crystallised, inhomogeneous and Te-bearing Fe-Cu-oxy-hydroxides, containing inclusions of
fine-grained native gold. (C,D) Microporous gold resulting from the dealloying of primary Te
minerals, most likely calaverite; some of the pores are filled with a Te-rich phase, probably
tellurite. Image credits: (A) Prof. V. Okrugin (Institute Of Volcanology And Seismology,
Russian Academy of Science, Petropavlovsk-Kamchatsky, Russia); (B-D) Copyright
permission received from Okrugin et al. (2014).
Journal Pre-proof

f
oo
pr
Figure 10: Moctezuma, Sonora, Mexico, is famous for the abundance and diversity of
e-
tellurium minerals. (A) On the first few meters below the surface, intense weathering results
Pr

in the formation of abundant Fe(Mn)-oxy- hydroxides; these are shown by portable-XRF to


contain high amounts of Te (>>100 ppm); however, no visible secondary mineral was
observed. Below this level, native tellurium was observed undergoing weathering to tellurite
al

(TeO 2 ) and probably other secondary minerals in smaller amounts. (B) Native tellurium (here
rn

in a calcite matrix) is the most prominent primary (hydrotherma l) Te mineral. (C) Tellurite is
one of the most prominent secondary minerals, here in millimetre-size blades. Image credit:
u

(C) Stefan Ansermet (Musée cantonal de géologie, Lausanne, Switzerland).


Jo
Journal Pre-proof

o f
r o
- p
r e
l P
n a
u r
J o
132
Figure 11: Log scale Te concentration contour maps in the eastern part of Fukushima Prefecture showing the distribution of the
132
anthropogenic and radioactive isotope Te around the Fukushima Daiichi Nuclear Power Plant (FDNPP) (decay-corrected to March 11, 2011)
132
showing (a) Measured Te; (b) estimated from 129m Te, and (c) all data including MEXT. Copyright permission received from Tagami et al.
(2013).
Journal Pre-proof

o f
r o
- p
r e
l P
n a
u r
J o
Figure 12: A schematic of a ‘super- microorganism’ capable of mediating the key processes microbes participate in with Te that control its
immobilisation and mobilisation; these include bioaccumulation, bioprecipitation, bioreduction, biovolatilisation and biooxidation. Biosorption
controls the initial step of interaction between any microorganism and Te, while bioreduction is the most common detoxification step.
Journal Pre-proof

f
oo
pr
e-
Figure 13: Mono Lake, California. Three Te-oxyanion reducing bacteria (Bacillus beveridgei,
Pr

B. selenitireducens and Sulphurospirillum barnesii, see Table 2) have been found from the
saline and alkaline waters of this lake (Baesman et al., 2006; Baesman et al., 2009; Baesman
al

et al., 2007). Note that the salinity is so high that evaporites precipitate onto the rocks sitting
rn

above the water line. Image licensed under a Creative Commons Attribution-Share Alike 3.0
Unported license (https://commons.wikimedia.org/wiki/File:Mono_Lake_Tufa.JPG).
u
Jo
Journal Pre-proof

f
oo
pr
e-
Pr
al
u rn
Jo

Figure 14: Te nanostructures produced by bioprecipitation. (a) SEM micrographs of Te


nanospheres and a cluster of Te nanorods produced by Bacillus beveridgei. Copyright
permission received from Baesman et al. (2009). Licence no.: 4761590947824. (b) TEM
micrograph of Te nanorods produced aerobically by Rhodococcus aetherivorans, visible
protruding outwards from a single cell. Reproduced from Presentato et al. (2016). (c) SEM
micrographs of SeTe nanospheres produced by Duganella violacienigra (C4 label on right
hand side indicates culture number in original paper). Reproduced from Bajaj and Winter
(2014).
Journal Pre-proof

o f
r o
- p
r e
l P
n a
u r
J o
Figure 15: Postulated model for Te and Au cycling in the regolith, modified after the gold dispersion model from Figure 2 of Rea et al. (2016).
Licence no.: 4761590226366. Note the major difference between the two elements is that Te can be made airborne by microbes in the form of
dimethyl telluride or other alkylated forms of Te, whereas Au remains in the soil or groundwater.
Journal Pre-proof

Declaration of interests

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the

work reported in this paper.

o f
r o
- p
r e
l P
n a
u r
J o

You might also like