Next Article in Journal
Inorganic Macro- and Micronutrients in “Superberries” Black Chokeberries (Aronia melanocarpa) and Related Teas
Previous Article in Journal
Chronic Exposure to Uranium from Gestation: Effects on Behavior and Neurogenesis in Adulthood
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cu Isotopic Composition in Surface Environments and in Biological Systems: A Critical Review

1
State Key Laboratory of Environmental Geochemistry, Institute of Geochemistry, Chinese Academy of Sciences, Guiyang 550081, China
2
Institute of Surface-Earth System Science Research, Tianjin University, Tianjin 300072, China
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2017, 14(5), 538; https://doi.org/10.3390/ijerph14050538
Submission received: 31 March 2017 / Revised: 11 May 2017 / Accepted: 11 May 2017 / Published: 18 May 2017
(This article belongs to the Section Environmental Health)

Abstract

:
Copper (Cu) is a transition metal and an essential micronutrient for organisms, but also one of the most widespread toxic inorganic contaminants at very high content. The research on Cu isotopes has grown rapidly in the last decade. Hitherto, a large number of studies have been published on the theoretical fractionation mechanisms, experimental data and natural variations of Cu isotopes in variable environments and ecosystems. These studies reported a large variation of δ65Cu (−16.49 to +20.04‰) in terrestrial samples and showed that Cu isotopes could be fractionated by various biogeochemical processes to different extent. Several papers have previously reviewed the coupling of Cu and Zn isotope systematics, and we give here a tentative review of the recent publications only on Cu isotopesin variable surface repositories, animals and human beings, with a goal to attract much attention to research on Cu (and other metals) behaviors in the environment and biological systems.

1. Introduction

Copper (Cu) is a transition metal and a ubiquitous biolimiting micronutrient that involves in many enzymes and regulatory proteins as a cofactor in almost all living organisms [1]. On the other hand, it is also one of the most widespread toxic inorganic contaminants at very high content [2], thus once released into the environment, it potentially threatens ecosystems and the biota (and even human beings) [3]. Cu has two stable isotopes, 65Cu (69.17%) and 63Cu (30.83%). Cu isotopes are proven to be helpful for understanding sources of Cu and processes that Cu undergoes in surface environments. Cu isotope studies have an apparently longer history, back to the 1940s [4]. Since then, this field had been developed very slowly until the 1990s, when the application of the new generation of multiple-collector magnetic-sector inductively coupled mass spectrometry (MC-ICP-MS) allowed for more precise isotopic analysis [5,6,7]. Though the measurement precision reported in early studies (for example, by thermal ionization mass spectrometry (TIMS)) was relatively low [4], the attempts of this pilot work prompted the following studies on Cu isotopes. Since last decade, researchers have devoted themselves to increasing the accuracy of Cu isotope measurement, including sample inlet system improvement, efficient Cu separation from various materials and mass bias correction [6,7,8,9,10,11,12,13,14]. As a pioneer work during this period, Maréchal et al. [6] analyzed Cu isotope composition by plasma-source mass spectrometry combined with a purification protocol to extract Cu from natural samples. Based on the recent advances in mass spectrometry and separation techniques, Cu stable isotopes could be accurately determined, with the current 2SD precision better than 0.05‰ [10,13,15], and its application has been largely extended to variable surface environments even related human health. Hitherto, a large number of studies regarding theoretical, experimental and application data sets of Cu isotope compositions have been published. These studies have largely improved our knowledge on Cu biogeochemical behaviors in biosphere, hydrosphere and lithosphere, and to different extent on Cu isotopic fractionation mechanisms in various geochemical, biological and metabolism processes [15,16,17,18,19,20,21].
Preceded this work, Albarède has reviewed some earlier papers about the isotope geochemistry of Cu (and Zn), with a focus on the analytical techniques and Cu isotope variations in biological and geological samples [8]. It must be emphasized that, in the last decade, a large number of papers has been published, which considerably enlarged our knowledge on Cu and its isotopes, especially on the mechanisms that fractionate Cu isotopes in variable processes. Most recently, Moynier et al. [22] reviewed the isotope geochemistry of Zn and Cu, such as both isotopic signatures in natural samples (even extra-terrestrial materials), their fractionation factors and experimental controlled mechanisms. Different from the previous contributions, we give here a tentative review of the recent publications on Cu isotopes, with a main focus on Cu isotope systematics in variable surface repositories and in animal and human organisms, and Cu isotopic fractionation mechanisms encountered in different surface biogeochemical and metabolism processes. The main goal of this review is to show the potential of Cu isotopes in studying the behaviors of Cu during variable environmental processes, and the linkage of Cu presence to the relative health effects.

2. Analytical Methodologies

2.1. Cu Separation from Sample Matrix

Before isotope determination by MC-ICP-MS, Cu should be extracted and purified from sample matrix, especially from Zn, Co and Cr that might interfere Cu isotopes during instrumental analysis. Practically, different anion-exchange resins were examined and used to purify Cu from natural samples, including AG1×2 (100–200 mesh, chloride form), AG1×4 (100–200 mesh, chloride form), AG 1×8 (BioRad, 100–200 mesh and 200–400 mesh), AG 1×10 (BioRad, 100–200 mesh), and AG MP-1 (100–200 mesh, chloride form) [6,11,23,24]. In these studies, efforts were made to guarantee: (1) high purity; and (2) high recovery of Cu in the final eluant during extraction and separation of this element from sample matrix, because the unsatisfactory recovery would lead to an unexpected fractionation of Cu isotopes. In a pioneering work, Koster used first AG 1×8 resin to extract Cu from silicate rock samples, but failed to recover all Cu [23]. Over a decade after, Walt et al. [24] found that anionic exchange resin AG MP-1 (100–200 mesh, chloride form) could isolate Cu(II) more efficiently, and then in 1999, Marechal et al. [6] applied this resin to purify Cu from various natural samples including silicates, carbonates and organisms. Larner et al. [11] established a new Cu extraction technique prior to the isotope analysis using AG 1×8 (BioRad, 200–400 mesh). In their method, univalent Cu+ was oxidized to Cu2+ on the resin and then Cu2+ was extracted into the elution, which allowed for a higher recovery (100 ± 2%) and lower procedure blanks. Since the ocean plays an important role in the global biogeochemical cycle of Cu, two separation protocols were developed for purifying trace Cu from seawater [25,26]. Bermin et al. [25] firstly measured Cu isotopes in seawater, and demonstrated that both chelation and Mg(OH)2 coprecipitation followed by anion exchange chromatography can efficiently separate and purify Cu from seawater samples (recovery of about 99 ± 9%). Later, in 2013, Takano et al. [26] developed another protocol composed of two chromatographic columns (Nobias-chelate PA1 resin and AGMP-1 (100–200 mesh, Bio-Rad, Berkeley, CA, USA)), which significantly facilitated the accurate analysis of seawater Cu isotope ratio.

2.2. Cu Isotope Analysis

Nowadays, Cu isotope determinations are generally performed on MC-ICP-MS (Nu instruments, UK or Neptune, Thermo-Fisher, Bremen, Germany). The Cu isotope data for a sample are expressed as δ65Cu (‰) notation, with a comparison to the 65Cu/63Cu ratio of the Cu standard solution, calculated by the formula:
δ65Cu = [(65Cu/63Cu)sample/(65Cu/63Cu)standard − 1] × 1000
where standard represents the international reference material NIST SRM 976 [6]. Recently, researchers have made efforts to improve the measurement precision, in different analytical aspects such as Cu separation from sample matrix [6,11,23,24], highly purified Cu standard development [10,13,27,28], and mass discrimination correction [10,29,30,31].
The NIST SRM 976 Cu solution was the commonly recommended standard for Cu isotopic analysis and widely used to test the instrument stability. Unfortunately, the NIST SRM 976 is no longer available; an alternative reference material is thus needed to replace this standard. This additional standard should be pure and sufficient to meet the global need for long term investigations on Cu isotopes and be (inter-) calibrated against NIST SRM 976. This also helps for testing and verifying accurate Cu isotope analysis [10,13,27]. Researchers are actively engaged in this job. For example, the Romil Cu standard (>99.999%, an industrial single element standard, Romil Ltd., Cambridge, UK) was calibrated relative to NIST SRM 976 and used for isotopic analysis [7,11,32,33]. Other standards such as ERM-AE633 and ERM-AE647 from the Institute for Reference Materials and Measurements (IRMM, Belgium), GSB Cu from Chinese National Reference Material (>99.99%), the National Standard Substances (China), and NWU-Cu-A and NWU-Cu-B (China) were also suggested and displayed δ65Cu values of +0.01 ± 0.05‰, +0.21 ± 0.05‰, +0.44 ± 0.04‰, +0.91 ± 0.03‰ and δ65Cu = −0.05 ± 0.03‰ relative to the NIST SRM 976 [13,27], respectively. Though commercially available, the different batches of one single standard may possess various isotopic ratios [7,8]. Thus, the reference material selected for the future should be calibrated and compared by different laboratories worldwide.

2.3. Mass Bias Correction

Since instrumental fractionation may eventually be produced in MC-ICP-MS measurements, the raw data (instrumental mass bias) must be corrected to obtain reliable isotope ratio [10,28]. Normally, mass bias could be corrected in several ways: (1) sample-standard bracketing (SSB), in which one unknown sample was determined between two standard analyses; (2) external normalization, by measuring an element (dopant) of similar mass (and/or chemical properties) with the analyte added in the sample; (3) a combination of external normalization and sample-standard bracketing, or (4) double-spike [34,35,36]. Because Cu has only two isotopes, double-spike is not applicable. While the other three methods were normally used, Zn was generally chosen as a dopant for mass bias correction in Cu isotopic analysis [14,31,37], and vice versa [6,34]. In addition, Ni and Ga were alternative candidates for such issue [10,38]. To be mentioned, the concentration ratio (and matrix) of dopant/Cu in the samples should be identical to that in the standard in order not to induce additional analysis bias [10].

3. Cu Isotopic Fractionation during Low Temperature Processes Revealed by Experimental Studies

Many experiments have been conducted in laboratory to better understand the Cu isotopic fractionation mechanisms in potential migration and transformation processes involved in the surface biogeochemical cycling of Cu. Previous studies showed that Cu isotope fractionation could be induced by various low temperature processes, such as adsorption, mineral dissolution, redox reaction, complexation with minerals and/or organic matters, and biotic incorporation [38,39,40,41,42,43,44,45,46,47]. In these processes, the isotopic fractionation (expressed as Δ65CuP-I) was determined for describing the enrichment or depletion of Cu isotope (63Cu or 65Cu) in the product compared to the initial materials, and was calculated as following:
Δ65CuP-I = δ65CuP − δ65CuI
where δ65Cup and δ65CuI were the isotope ratio in the product P and initial material I, respectively. Figure 1 summarizes Δ65Cu (−1.46 ~ +4.40‰) in low temperature processes. Studies on leaching of different ores or minerals indicated that, in general, heavy Cu isotope (65Cu) was preferentially released into the acidic solutions (with Δ65Cusolution-solid as large as +3‰), while the lighter remained in the mineral or residual phases [39,42,47,48,49,50,51]. The pH might influence the Cu isotopic fractionation to different degrees depending on the geochemical reactions. For example, the pH variation impact on Cu isotopic fractionation were significant, slight and insignificant respectively during Cu interaction with soil and aquatic microorganisms (and metal oxy(hydro)oxides), adsorption onto kaolinite and complexation with insolubilized humic acid [40,50,52]. In a chalcopyrite oxidation experiment, Rodríguez et al. [45] showed that heavy Cu isotope was easily oxidized and released into solution at the beginning, and as the leaching progressed, the leachate displayed increasingly similar δ65Cu values to the chalcopyrite, close to an equilibrium fractionation in a closed system. Large isotopic fractionation was observed in aqueous Cu(II) reactions with CuS and FeS with heavy isotope (65Cu) enriched in solution (Δ65Cusolution-mineral ~ +3‰) [39,53]. Complextion could produce Cu isotope fractionation with an enrichment of heavy Cu isotope in the organo-Cu complexes (Δ65Cucomplex-free: +0.14 to +0.84‰) [38,40]. During adsorption processes, heavy Cu isotopes could be preferentially adsorbed to Fe and Al oxy(hydr)oxides or bacteria as well thus induced an isotopic offset of Δ65Cusolution-solid [43,50,54,55]. In fact, the distribution amongst different species would also induce large Cu isotope fractionation, as calculated by Fujii et al. [37] that Cu in dissolved carbonates and sulfates was isotopically heavier than free Cu.
Cu isotopic fractionation may also be triggered by biological processes. It is worth mentioning that Mathur et al. [43] observed distinct enrichment of δ65Cu by up to +3.0‰ in Cu sink on bacteria surfaces in contrast to the aqueous phase. Navarrete et al. [44] performed a further surface adsorption research and indicated that the apparent separation factors (Δ65Cusolution–solid) produced by live bacterial cells ranging from +0.20 to +2.60‰, and live-bacteria intracellular incorporation preferred the lighter Cu isotope even produced a significant larger fractionation of Cu isotopes (apparent Δ65Cusolution-solid ranging from ~+1.0 to +4.4‰). They suggested that the cell enrichment of the light Cu isotope was probably a metabolically-driven phenomenon rather than caused by reversible surface adsorption, a result consistent with the previous study [58].

4. Cu Isotope Compositions in Earth Surface Reservoirs

4.1. Solid Reservoirs on the Earth Surface

4.1.1. Soil Weathering Profiles

The transportation and bioavailability of Cu in soils were strongly controlled by the chemical Cu forms and affected by water-induced alterations under variable soil redox conditions. Cu could exist in three forms as elemental Cu, Cu+ and Cu2+ in soils and Cu2+ was the major form [59,60]. Several studies reported Cu isotopic fractionation and the vertical variation in soils, and showed diverse variation tendencies for different types of soil [59,61,62,63]. Bigalke et al. [62] studied the hydromorphic soils and found an interesting correlation of δ65Cu with Fe oxide proportion at different crystallization stage, and suggested an important effect of redox condition (caused by water saturation) on Cu isotope variation. This was further confirmed by the sequential chemical extraction study of the same group on a soil core from a freshwater intertidal mudflat, which showed δ65Cu values increased with depth while the redox conditions changed from oxic to strongly anoxic conditions [61]. Recently, Kusonwiriyawong et al. [64] studied the Cu isotope variations related to the chemical partitioning changes by leaching the soils with different reactants (F1–F5:NH4NO3-extractable, NaOAc-extractable, NH4Ox-extractable, hot H2O2/NH4OAc-extractable and residual fractions, respectively). They observed a strong effect of redox conditions on Cu partitioning and isotope composition in soils and a Cu availability declining under anoxic conditions. In addition, Cu isotope ratios might display seasonal variations with organic matter and redox condition concomitantly changes [61]. Bigalke et al. [59] also found that, unlike the hydromorphic soils, oxic weathering without evident podzolization induced an increasingly light tendency of δ65Cu values with depth.
Other studies reported also large variations of Cu isotope ratio during surface weathering of shales. Black shales, one of the main Cu pools that can release Cu to water systems, displayed large isotope variation (between −6.42 and +19.73‰), with heavy Cu isotope significantly preferred to release into fluids, which was in line with leaching experimental results [65]. For instance, the soils were found depleted in 65Cu (average δ65Cu = −0.5 ± 0.2‰) during shale weathering compared to the parent rocks (average δ65Cu = +0.03 ± 0.15‰), while the pore waters were enriched in 65Cu (average δ65Cu = +1.14 ± 0.44‰) [66].

4.1.2. Lithosphere

Lithosphere is an important Cu reservoir on the earth. About 75% of the world’s copper are supplied by porphyry copper systems [67]. There are three major mineral categories according to Cu isotopic distribution: leached, supergene and hypogenezone [68]. Large variation of δ65Cu (−16.49 to +20.04‰) was observed in ores and minerals [6,59,69,70], averaging +0.11‰, as summarized in Figure 2. Hypogene minerals (e.g., chalcopyrite, chalcocite, covellite, and pyrite enargite) usually had a narrow δ65Cu variation around zero, whereas leached (e.g., hematite, jarosite, and goethite) was isotopically lighter and supergene minerals (e.g., chalcocite, hematite, and Cu oxide) isotopically heavier [57,68,71,72,73,74]. Narrow variations of Cu isotopic composition in silicate reservoirs (e.g., basalts, granites) were reported recently [75,76]. It suggested that high-temperature magmatic processes do not induce large Cu isotope fractionation [30,72,77,78,79]. In addition, Liu et al. [75] found the metasomatized peridotites displayed relatively wider ranges of δ65Cu values (−0.64 to +1.82‰) than non-metasomatized peridotites (from −0.15 to +0.18‰), indicating metasomatism might trigger significant isotopic fractionation.
Cu isotope can be significantly fractionated (~36‰) in low-temperature ore deposits, and this was mainly attributed to the processes such as redox reactions, mineral dissolution and precipitation [14,39,47,48,49,53,66]. Zhu et al. [7] found that Cu formed at low temperature displayed large δ65Cu variations in native Cu, malachite, azurite and chalcopyrite at a single site, suggesting that the δ65Cu variations were mainly caused by low temperature aqueous process fractionation rather than source heterogeneity. In the sedimentary mineralization, the Cu isotopic ratio of Cu sulphide concretions exhibited abimodal distribution (δ65Cu = −0.91 ± 0.51‰ and −3.32 ± 0.29‰), and redox and salinity variations were likely the major regional controlling parameters affecting the Cu isotopic ratio [90]. Interestingly, Wilson et al. [21] measured Cu isotopes with a goal of identifying their origin of mislabeled Cu minerals. Since the δ65Cu (+5.63‰) of their Cu crystal specimen was significantly different from Michigan Cu samples with a tight cluster of Cu isotope ratio (averaging δ65Cu ~ +0.30‰), but in accordance with the Namibian chalcopyrite (δ65Cu: +1.73~ +9.27‰), signifying a possible origin of these minerals from Namibia.

4.2. Biosphere

As a redox-active metal, Cu functions as an enzyme-activator and is a significant prosthetic-group-part of many enzymes [94]. Given the importance of Cu in biological activities, researchers have carried out a large number of studies on Cu isotopic fractionation and relative mechanisms during soil-plant uptake and within-plant Cu translocation [59,95,96]. Recent studies indicated that plants systematically enriched in lighter Cu isotope compared with the soil in which they grew, and reduction was the major ubiquitous uptake mechanism. Free Cu(II) was reduced to Cu(I) at the surface of the plasma membrane by specific reductase and then Cu(I) was transported into plant roots by a transporter protein (COPT1) with an enrichment of isotopically light Cu [18,95,97,98]. Various within-plant Cu isotopic fractionation direction and magnitude were observed among different plant organs (root and stem), different plant tissues (xylem and phloem), and plant senescence caused remobilization [18]. Moreover, different plant components also had different Cu isotope compositions, for example, in the dicot, lentil (Lens culinaris), and two monocots, virginia wild rye (Elymusvirginicus) and hairy-leavedsedge (Carexhirsutella), the shoots were isotopically lighter than the germinated seeds and the Cu isotopic signature of leaves had a correlation with the plant heights [98]. Oat plants exhibited no evident fractionation during translocation, but heavier Cu isotope was preferentially transported to shoots in tomato, and the isotopic variations within different tissues were mainly due to redox cycling [97]. In addition, the magnitude of soil-plant fractionation was found to be impacted by free Cu ion species (modulated by the reduction process) in soil solution and to be correlated with soil pH values [18].
Summary on fractionation factors (Δ65Cu) during plant–soil and within-plant processes are illustrated in Figure 3. As was shown, Δ65Cu variation between plant organs and rhizosphere environment (soil/solution) are −1.43 to −0.11‰, while the in-plant Δ65Cu ranged from −0.90 to +1.35‰. The largest variation of Δ65Cuplant-rhizosphere environment (−1.43‰) was observed between root and solution in tomato, while the maximum in-plant Δ65Cu (+1.35‰) was found between stem and root of tomato.

4.3. Hydrosphere

Studies carried out in last decade have shown that Cu isotope approach could provide new insights into fingerprinting sources and processes (and mechanisms) that Cu involved in various aquatic environments, including rivers, streams and oceans [19,34,83,84,85,87,89,99,100]. According to the literature data, dissolved δ65Cu in rivers ranged from −0.69 to +1.55‰, with the average +0.53‰ ± 2SD; and the suspended particulate matter (SPM) in aquatic environment (wetland, estuaries and rivers) displayed a complementary light pool (−1.02 to +0.09‰, averaging −0.31‰ ± 2SD) [85,88,89,99]. Generally, dissolved Cu in the aqueous repertories was isotopically heavier than the solid Earth such as SPM [89], however, rivers displayed different trend significantly impacted by anthropogenic sources with the lightisotope enriched in waters [85]. For example, in a small Mediterranean vineyard river, dissolved Cu displayed positive δ65Cu (+0.31‰), similar to the SPM (δ65Cu = +0.26‰), and higher than the local bedrock (+0.07‰), due to the significant contribution from long-term application of coppery fungicides [88]. The δ65Cu values of the surface water were found increasingly lighter along the distance from the Pebble Porphyry Cu-Au-Mo deposit, indicating a gradually varied contribution from high-temperature-formed mineral weathering [87].
Up to now, the fractionation of Cu isotopes triggered by in-river processes (adsorption, desorption, biological processes, etc.) has rarely been reported. Ilina et al. [99] showed that the dissolved and the colloidal Cu in Palojoki River displayed similar δ65Cu values, and attributed this to the high isotopic exchange rate between the colloidal and ionic Cu. Another interesting study on wetland showed that the dissolved Cu was depleted in 65Cu when traversing the wetland (Δ65Cuinlet−outlet between +0.03 and +0.77‰), probably caused by Cu adsorption onto organic matters and aluminum minerals [84].
Cu isotopes have also been used to trace the impact of mining or smelting activities on aquatic systems. Sulfide-bearing tailing profiles showed that light Cu isotope was enriched in the oxidized zone than in the original materials [84]. Song et al. [73] observed a large δ65Cu variation (−0.44‰ to +24.4‰) in stream water (impacted by mining activities at the Dexing Mine, the largest Cu mine in Asia), which could be stemmed from pyrite tailings and chalcopyrite weathering. Both results were consistent with the leaching experiments that heavy Cu was preferentially enriched in solutions [42,44,45,48]. Thapalia et al. [20] performed a research on Cu isotope application to trace the sources and history of the Cu deposition in a sediment core from Lake Ballinger adjacent to Seattle, Washington, USA, indicating mean δ65Cu values possibly influenced by the atmospheric deposition of smelter exhaust is about +0.94 ± 0.1‰ (2σ, n = 6) between 1900 and 1979.
Except for continental aquatic environments, researchers also conducted studies on Cu isotope compositions in sea water sampled from the English Channel, Atlantic, Pacific, Indian oceans, andarctic gyre and determined a large δ65Cu variation of +0.38 to +1.44‰, averaging +0.70‰ [19,25,83,89]. In general, the seawater exhibited heavier Cu isotopic ratio than the riverine input (−0.69 to +1.55‰, averaging +0.53‰). This indicated that the heavy isotopic signature of seawater was likely produced by intra-oceanic processes (with light isotope preferential scavenging into sediments), rather than continental input [19,85,89]. In addition to the consistent scavenging effect that increase/decrease 65Cu in deep seawater, Takano et al. [19] also demonstrated that Cu isotopic composition of the surface seawater (from the North and SouthAtlantic, North Pacific and South Indian oceans) was deviated from the mixing of deep seawater, precipitation and river, and thus could be induced bybiological processes. Little et al. [31] compared Cu isotopic composition in Fe-Mn crusts to the dissolved load from the ocean basins, and showed relatively lower δ65Cu (+0.44 ± 0.23‰) than the bulk isotopic ratios (δ65Cu ~ +0.90‰) of dissolved Cu in seawater, comparable to other studies [19,89].

4.4. Atmosphere

Up to now, only few studies reported Cu isotopes in the atmosphere, which could however be an important tracer to trace atmospheric Cu contamination and transportation. The reported δ65Cu variation of the atmospheric precipitation (including snow, rime and rainwater) was −1.44 ~ +0.40‰, averaging −0.12‰ [19,82]. More studies are expected to map the regional and global atmospheric Cu isotope signatures and to investigate the mechanisms fractionating Cu isotopes in atmospheric processes.

5. Animals and Human Beings

Recently, isotope geochemists have made efforts to investigate metal (metalloid) isotopes in organisms of animals and human beings and to link isotopic fractionations to the metabolism processes and thus relative health effects. Significantly, Cu is a micronutrient, a catalytic and structural cofactor for various important enzymes referred to tumor development, and also behave as a modulator of inflammation. It is worth mentioning that Cu isotopic signatures have been reported in organs and body fluids of mammals including human beings, though data from human material were very limited for ethical reasons. Oxidized cations such as Cu(II) and low coordination numbers were found to favor heavy isotopes contrast to the reduced product (Cu(I)) and high coordination numbers [16]. For the aquatic animal, such as fish, Komjarova et al. [101] determined isotope ratios of Cd and Cu in zebrafish gills and found that Cd uptake was evidently restrained by Cu coexistence, suggesting an interaction between Cu and Cd, yet Cu uptake was not interfered by Cd. For terrestrial animals like sheep and mice, the δ65Cu values (inliver, kidney and brain) varied significantly from −1.5‰ to +1.5‰, and it was implied that Cu isotope signature could potentially be modified by the prion protein expression [80,102]. Moreover, in the mammal trophic chains, 65Cu was enriched by about 1.0‰ in bones of carnivores than those of herbivores [103].
Cu isotopic ratio has been investigated in materials from human bodies both for healthy people and cancer patients (Table 1). Cu was translocated by serum albumin, and excess Cu could be stored in metallothione as well [16]. Commonly, Cu was more concentrated in erythrocytes compared with serum. Van Heghe et al. [104] investigated Cu isotope compositions in whole blood samples from healthy volunteers, and found that Cu isotope ratios seemed to be affected by neither gender nordiet, which could be attributed to their failure in Cu extraction from samples. Albarède et al. [105] analyzed Cu isotope ratios of whole blood, serum, and red blood cells of about 50 young blood donors, and for healthy individuals, the Cu isotopic composition variationin human blood was about +1.6‰, with the lowest δ65Cu value in serum (−0.7‰) and the highest in erythrocytes (+0.9‰); Blood of men and women displayed visibly different Cu isotope compositions and with 0.2‰ lighter for women. More interestingly, B-type men erythrocytes could fractionate Cu to a larger extent than all the other blood types, and light Cu was mainly enriched in blood and lymphatic circulation rather than in erythrocytes [105]. Jaouen et al. [81] found that δ65Cu values in human blood decreased with age, and Cu isotope ratios in the Yakut’s blood were significantly lighter than those in blood of European and Japanese populations. In an archeological human bone study, the male’s bones exhibited heavier Cu isotopes than the females’ bones, implying Cu isotopes could be used as a valid tool to distinguish sexes of incomplete human remains [103].
In addition to that of healthy individuals, Cu isotopic characteristics were also studied for cancer patients. Significantly lower δ65Cu values were observed in the serum of patients with breast, liver, and colon cancer in contrast to those in samples from healthy individuals [16]. In the blood of hepatocellular carcinoma (HCC) patients, Cu was 63Cu-enriched by about 0.4‰ compared with that of the control group, and more concentrated Cu was observed in red blood cells and serum of HCC patients [106]. For colorectal and breast cancer patients, the δ65Cu decreased in the serum while increased in cancer cells, which was likely attributed to the extensive oxidative chelation of Cu by cytosolic lactate, and most patients in their study with serum δ65Cu below the threshold level (−0.35‰) did not survive [107]. Particularly, as a marker, the δ65Cu decline preceded biomarkers for several months, thus suggested a strong potential of Cu isotope composition as a new diagnostic tool at least for breast and colorectal cancers.

6. Conclusions

Cu isotopes are increasingly studied and applied in surface biogeochemistry in recent years. As a main Cu reservoir on the earth, the lithosphere displays a large variation of δ65Cu (−16.49 to +20.04‰), which might result from many processes including especially the redox reaction. The surface weathering process (e.g., water–rock interaction) would modify the initial Cu isotopic signature of bedrocks, by, for example, enriching heavy Cu isotope (65Cu) in water, and thus light Cu isotope (63Cu) preferentially remains in the parent materials. The input into the ocean through continental rivers is thus generally characterized by higher Cu isotope composition. However, the anthropogenic contribution may neutralize this Cu isotopic signature by adding Cu with lower δ65Cu value. Though the riverine input (also atmospheric precipitation) could impact the Cu isotopic compositionin offshore seawater, the biological processes may play an important role on the budget of Cu and its isotopes in pelagic water of the open sea. Interestingly, large variation of δ65Cu (−1.45~+1.24‰) has been observed in animal and human materials, implying that the diet intake and metabolism processes would also induce Cu isotope variation. Importantly, several studies reported that healthy individuals and cancer patients possess different Cu isotopic ratios in the same organism, demonstrating a great potential of Cu isotopes in human health studies especially as a new diagnostic toolfor cancers. Though significant progress has been made in the last decade, more studies should be carried out on Cu isotopic fractionation in surface environmental, biogeochemical, and metabolism processes, in order to better understand the transformation of Cu (and other metals) from the environment to ecosystem and even human bodies and the relative toxicity effect of Cu (and metals).

Acknowledgments

This work was financially supported by the Natural Science Foundation of China (U1301231, 41561134017, 41625012, 41603020, and U1612442). The authors acknowledge Qixin Wu from Guizhou University, who improved the quality of the manuscript.

Author Contributions

Jiubin Chen and Zhuhong Wang conceived and designed the systematic reviewand analyzed the data; Zhuhong Wang, Jiubin Chen and Ting Zhang wrote the paper; and Zhuhong Wang, Jiubin Chen and Ting Zhang reviewed and refined the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Semeniuk, D.M.; Cullen, J.T.; Johnson, W.K.; Gagnon, K.; Ruth, T.J.; Maldonado, M.T. Plankton copper requirements and uptake in the subarctic northeast pacific ocean. Deep Sea Res. Part I 2009, 56, 1130–1142. [Google Scholar] [CrossRef]
  2. De Oliveira Filho, E.C.; Lopes, R.M.; Paumgartten, F.J. Comparative study on the susceptibility of freshwater species to copper-based pesticides. Chemosphere 2004, 56, 369–374. [Google Scholar] [CrossRef] [PubMed]
  3. Serra, A.; Guasch, H. Effects of chronic copper exposure on fluvial systems: Linking structural and physiological changes of fluvial biofilms with the in-stream copper retention. Sci. Total Environ. 2009, 407, 5274–5282. [Google Scholar] [CrossRef] [PubMed]
  4. Brown, H.; Inghram, M. The isotopic composition of meteoritic copper. Phys. Rev. 1947, 72, 347. [Google Scholar] [CrossRef]
  5. Chi Fru, E.; Rodríguez, N.P.; Partin, C.A.; Lalonde, S.V.; Andersson, P.; Weiss, D.J.; El Albani, A.; Rodushkin, I.; Konhauser, K.O. Cu isotopes in marine black shales record the great oxidation event. Proc. Natl. Acad. Sci. USA 2016, 113, 4941–4946. [Google Scholar] [CrossRef] [PubMed]
  6. Maréchal, C.N.; Télouk, P.; Albarède, F. Precise analysis of copper and zinc isotopic compositions by plasma-source mass spectrometry. Chem. Geol. 1999, 156, 251–273. [Google Scholar] [CrossRef]
  7. Zhu, X.K.; O’Nions, R.K.; Guo, Y.; Belshaw, N.S.; Rickard, D. Determination of natural Cu-isotope variation by plasma-source mass spectrometry: Implications for use as geochemical tracers. Chem. Geol. 2000, 163, 139–149. [Google Scholar] [CrossRef]
  8. Albarède, F. The stable isotope geochemistry of copper and zinc. Rev. Mineral. Geochem. 2004, 55, 409–427. [Google Scholar] [CrossRef]
  9. Balliana, E.; Aramendia, M.; Resano, M.; Barbante, C.; Vanhaecke, F. Copper and tin isotopic analysis of ancient bronzes for archaeological investigation: Development and validation of a suitable analytical methodology. Anal. Bioanal. Chem. 2013, 405, 2973–2986. [Google Scholar] [CrossRef] [PubMed]
  10. Hou, Q.; Zhou, L.; Gao, S.; Zhang, T.; Feng, L.; Yang, L. Use of Ga for mass bias correction for the accurate determination of copper isotope ratio in the NISTSRM3114 Cu standard and geological samples by MC-ICP-MS. J. Anal. At. Spectrom. 2015, 31, 280–287. [Google Scholar] [CrossRef]
  11. Larner, F.; Rehkämper, M.; Coles, B.J.; Kreissig, K.; Weiss, D.J.; Sampson, B.; Unsworth, C.; Strekopytov, S. A new separation procedure for Cu prior to stable isotope analysis by MC-ICP-MS. J. Anal. At. Spectrom. 2011, 26, 1627. [Google Scholar] [CrossRef]
  12. Petit, J.C.; Taillez, A.; Mattielli, N. A case study of spectral and non-spectral interferences on copper isotope measurements by multi-collector ICP-MS (wet plasma). Geostand. Geoanal. Res. 2013, 37, 319–335. [Google Scholar] [CrossRef]
  13. Yuan, H.; Yuan, W.; Bao, Z.; Chen, K.; Huang, F.; Liu, S. Development of two new copper isotope standard solutions and their copper isotopic compositions. Geostand. Geoanal. Res. 2017, 41, 77–84. [Google Scholar] [CrossRef]
  14. Zhu, X.K.; Guo, Y.; Williams, R.J.P.; O’Nions, R.K.; Matthews, A.; Belshaw, N.S.; Canters, G.W.; Waal, E.C.D.; Weser, U.; Burgess, B.K. Mass fractionation processes of transition metal isotopes. Earth Planet. Sci. Lett. 2002, 200, 47–62. [Google Scholar] [CrossRef]
  15. Vance, D.; Matthews, A.; Keech, A.; Archer, C.; Hudson, G.; Pett-Ridge, J.; Chadwick, O.A. The behaviour of Cu and Zn isotopes during soil development: Controls on the dissolved load of rivers. Chem. Geol. 2016, 445, 36–53. [Google Scholar] [CrossRef]
  16. Albarede, F.; Telouk, P.; Balter, V.; Bondanese, V.P.; Albalat, E.; Oger, P.; Bonaventura, P.; Miossec, P.; Fujii, T. Medical applications of the Cu, Zn, and S isotope effects. Metallomics 2016, 8, 1056–1070. [Google Scholar] [CrossRef] [PubMed]
  17. Chen, J.; Gaillrdet, J.; Louvat, P. Multi-isotopic (Zn, Cu) approach for anthropogenic contamination of suspended sediments of the Seine River, France. Geochim. Cosmochim. Acta Suppl. 2008, 72, 155. [Google Scholar] [CrossRef]
  18. Li, S.-Z.; Zhu, X.-K.; Wu, L.-H.; Luo, Y.-M. Cu isotopic compositions in elsholtzia splendens: Influence of soil condition and growth period on Cu isotopic fractionation in plant tissue. Chem. Geol. 2016, 444, 49–58. [Google Scholar] [CrossRef]
  19. Takano, S.; Tanimizu, M.; Hirata, T.; Sohrin, Y. Isotopic constraints on biogeochemical cycling of copper in the ocean. Nat. Commun. 2014, 5, 5663. [Google Scholar] [CrossRef] [PubMed]
  20. Thapalia, A.; Borrok, D.M.; Van Metre, P.C.; Musgrove, M.; Landa, E.R. Zn and Cu isotopes as tracers of anthropogenic contamination in a sediment core from an urban lake. Environ. Sci. Technol. 2010, 44, 1544–1550. [Google Scholar] [CrossRef] [PubMed]
  21. Wilson, M.; Wilson, D.L.; Mathur, R. Tracing the source of native copper mineral specimens with copper isotope values. Rocks Miner. 2016, 352–356. [Google Scholar] [CrossRef]
  22. Moynier, F.; Vance, D.; Fujii, T.; Savage, P. The isotope geochemistry of zinc and copper. Rev. Mineral. Geochem. 2017, 82, 543–600. [Google Scholar] [CrossRef]
  23. Koster, H.M. Die bestimmung von kupfer(II), eisen(III) und zink(II) mittels atomabsorption bei der chemischen gesteinsanalyse nach abtrennung der losungspartner am cl-be-ladenen anionenaustauscher dowex 1 × 8 oder amberlite cg 400-1. Neues Jahrb. Miner. Abh. 1973, 119, 145–154. [Google Scholar]
  24. Van der Walt, T.N.; Strelow, F.W.E.; Verheij, R. The influence of crosslinkage on the distribution coefficients and anion exchange behaviour of some elements in hydrochloric acid. Solvent Extr. Ion Exch. 1985, 3, 723–740. [Google Scholar] [CrossRef]
  25. Bermin, J.; Vance, D.; Archer, C.; Statham, P.J. The determination of the isotopic composition of Cu and Zn in seawater. Chem. Geol. 2006, 226, 280–297. [Google Scholar] [CrossRef]
  26. Takano, S.; Tanimizu, M.; Hirata, T.; Sohrin, Y. Determination of isotopic composition of dissolved copper in seawater by multi-collector inductively coupled plasma mass spectrometry after pre-concentration using an ethylenediaminetriacetic acid chelating resin. Anal. Chim. Acta 2013, 784, 33–41. [Google Scholar] [CrossRef] [PubMed]
  27. Moeller, K.; Schoenberg, R.; Pedersen, R.B.; Weiss, D.; Dong, S. Calibration of the new certified reference materials ERM-AE633 and ERM-AE647 for copper and IRMM-3702 for zinc isotope amount ratio determinations. Geostand. Geoanal. Res. 2012, 36, 177–199. [Google Scholar] [CrossRef]
  28. Liu, S.A.; Li, D.; Li, S.; Teng, F.Z.; Ke, S.; He, Y.; Lu, Y. High-precision copper and iron isotope analysis of igneous rock standards by MC-ICP-MS. J. Anal. At. Spectrom. 2013, 29, 122–133. [Google Scholar] [CrossRef]
  29. Mason, T.F.D.; Weiss, D.J.; Horstwood, M.; Parrish, R.R.; Russell, S.S.; Mullane, E.; Coles, B.J. High-precision Cu and Zn isotope analysis by plasma source mass spectrometry. J. Anal. At. Spectrom. 2004, 19, 218–226. [Google Scholar] [CrossRef]
  30. Archer, C.; Vance, D. Mass discrimination correction in multiple-collector plasma source mass spectrometry: An example using Cu and Zn isotopes. J. Anal. At. Spectrom. 2004, 19, 656–665. [Google Scholar] [CrossRef]
  31. Little, S.; Sherman, D.; Vance, D.; Hein, J. Molecular controls on Cu and Zn isotopic fractionation in Fe-Mn crusts. Earth Planet. Sci. Lett. 2014, 396, 213–222. [Google Scholar] [CrossRef]
  32. Mason, T.F.D.; Weiss, D.J.; Chapman, J.B.; Wilkinson, J.J.; Tessalina, S.G.; Spiro, B.; Horstwood, M.S.A.; Spratt, J.; Coles, B.J. Zn and Cu isotopic variability in the alexandrinka volcanic-hosted massive sulphide (VHMS) ore deposit, Urals, Russia. Chem. Geol. 2005, 221, 170–187. [Google Scholar] [CrossRef]
  33. Peel, K.; Weiss, D.; Chapman, J.; Arnold, T.; Coles, B. A simple combined sample-standard bracketing and inter-element correction procedure for accurate and precise Zn and Cu isotope ratio measurements. J. Anal. At. Spectrom. 2008, 23, 103–110. [Google Scholar] [CrossRef]
  34. Chen, J.B.; Louvat, P.; Gaillardet, J.; Birck, J.L. Direct separation of Zn from dilute aqueous solutions for isotope composition determination using multi-collector ICP-MS. Chem. Geol. 2009, 259, 120–130. [Google Scholar] [CrossRef]
  35. Chen, J.; Hintelmann, H.; Dimock, B. Chromatographic pre-concentration of Hg from dilute aqueous solutions for isotopic measurement by MC-ICP-MS. J. Anal. At. Spectrom. 2010, 74, A169. [Google Scholar] [CrossRef]
  36. Yang, L. Accurate and precise determination of isotopic ratios by MC-ICP-MS: A review. Mass Spectrom. Rev. 2009, 28, 990–1011. [Google Scholar] [CrossRef] [PubMed]
  37. Fujii, T.; Moynier, F.; Abe, M.; Nemoto, K.; Albarède, F. Copper isotope fractionation between aqueous compounds relevant to low temperature geochemistry and biology. Geochim. Cosmochim. Acta 2013, 110, 29–44. [Google Scholar] [CrossRef]
  38. Ryan, B.M.; Kirby, J.K.; Degryse, F.; Scheiderich, K.; Mclaughlin, M.J. Copper isotope fractionation during equilibration with natural and synthetic ligands. Environ. Sci. Technol. 2014, 48, 8620–8626. [Google Scholar] [CrossRef] [PubMed]
  39. Ehrlich, S.; Butler, I.; Halicz, L.; Rickard, D.; Oldroyd, A.; Matthews, A. Experimental study of the copper isotope fractionation between aqueous Cu(ii) and covellite, Cus. Chem. Geol. 2004, 209, 259–269. [Google Scholar] [CrossRef]
  40. Bigalke, M.; Weyer, S.; Wilcke, W. Copper isotope fractionation during complexation with insolubilized humic acid. Environ. Sci. Technol. 2010, 44, 5496–5502. [Google Scholar] [CrossRef] [PubMed]
  41. Fulda, B.; Voegelin, A.; Maurer, F.; Christl, I.; Kretzschmar, R. Copper redox transformation and complexation by reduced and oxidized soil humic acid. 1. X-ray absorption spectroscopy study. Environ. Sci. Technol. 2013, 47, 10903–10911. [Google Scholar] [CrossRef] [PubMed]
  42. Mathur, R.; Munk, L.A.; Townley, B.; Gou, K.Y.; Gómez Miguélez, N.; Titley, S.; Chen, G.G.; Song, S.; Reich, M.; Tornos, F.; et al. Tracing low-temperature aqueous metal migration in mineralized watersheds with Cuisotope fractionation. Appl. Geochem. 2014, 51, 109–115. [Google Scholar] [CrossRef]
  43. Mathur, R.; Ruiz, J.; Titley, S.; Liermann, L.; Buss, H.; Brantley, S. Cu isotopic fractionation in the supergene environment with and without bacteria. Geochim. Cosmochim. Acta 2005, 69, 5233–5246. [Google Scholar] [CrossRef]
  44. Navarrete, J.U.; Borrok, D.M.; Viveros, M.; Ellzey, J.T. Copper isotope fractionation during surface adsorption and intracellular incorporation by bacteria. Geochim. Cosmochim. Acta 2011, 75, 784–799. [Google Scholar] [CrossRef] [PubMed]
  45. Rodríguez, N.P.; Khoshkhoo, M.; Sandström, Å.; Rodushkin, I.; Alakangas, L.; Öhlander, B. Isotopic signature of Cu and Fe during bioleaching and electrochemical leaching of a chalcopyrite concentrate. Int. J. Miner. Process. 2015, 134, 58–65. [Google Scholar] [CrossRef]
  46. Sherman, D.M. Equilibrium isotopic fractionation of copper during oxidation/reduction, aqueous complexation and ore-forming processes: Predictions from hybrid density functional theory. Geochim. Cosmochim. Acta 2013, 118, 85–97. [Google Scholar] [CrossRef]
  47. Wall, A.J.; Mathur, R.; Post, J.E.; Heaney, P.J. Cu isotope fractionation during bornite dissolution: An in situ x-ray diffraction analysis. Ore Geol. Rev. 2011, 42, 62–70. [Google Scholar] [CrossRef]
  48. Fernandez, A.; Borrok, D.M. Fractionation of Cu, Fe, and Zn isotopes during the oxidative weathering of sulfide-rich rocks. Chem. Geol. 2009, 264, 1–12. [Google Scholar] [CrossRef]
  49. Kimball, B.E.; Mathur, R.; Dohnalkova, A.C.; Wall, A.J.; Runkel, R.L.; Brantley, S.L. Copper isotope fractionation in acid mine drainage. Geochim. Cosmochim. Acta 2009, 73, 1247–1263. [Google Scholar] [CrossRef]
  50. Pokrovsky, O.S.; Viers, J.; Emnova, E.E.; Kompantseva, E.I.; Freydier, R. Copper isotope fractionation during its interaction with soil and aquatic microorganisms and metal oxy(hydr)oxides: Possible structural control. Geochim. Cosmochim. Acta 2008, 72, 1742–1757. [Google Scholar] [CrossRef]
  51. Seo, J.H.; Lee, S.K.; Lee, I. Quantum chemical calculations of equilibrium copper(I) isotope fractionations in ore-forming fluids. Chem. Geol. 2007, 243, 225–237. [Google Scholar] [CrossRef]
  52. Li, D.; Liu, S.A.; Li, S.; Li, D.; Liu, S.A. Copper isotope fractionation during adsorption onto kaolinite: Experimental approach and applications. Chem. Geol. 2015, 396, 74–82. [Google Scholar] [CrossRef]
  53. Pękala, M.; Asael, D.; Butler, I.B.; Matthews, A.; Rickard, D. Experimental study of Cu isotope fractionation during the reaction of aqueous Cu(II) with Fe(II) sulphides at temperatures between 40 and 200 °C. Chem. Geol. 2011, 289, 31–38. [Google Scholar] [CrossRef]
  54. Balistrieri, L.S.; Borrok, D.M.; Wanty, R.B.; Ridley, W.I. Fractionation of Cu and Zn isotopes during adsorption onto amorphous Fe(III) oxyhydroxide: Experimental mixing of acid rock drainage and ambient river water. Geochim. Cosmochim. Acta 2008, 72, 311–328. [Google Scholar] [CrossRef]
  55. Clayton, R.E.; Hudsonedwards, K.A.; Houghton, S.L. Isotopic effects during Cu sorption onto goethite. Geochim. Cosmochim. Acta Suppl. 2005, 69, A216. [Google Scholar] [CrossRef]
  56. Maher, K.C.; Jackson, S.; Mountain, B. Experimental evaluation of the fluid-mineral fractionation of Cu isotopes at 250 °C and 300 °C. Chem. Geol. 2011, 286, 229–239. [Google Scholar] [CrossRef]
  57. Mathur, R.; Schlitt, W. Identification of the dominant Cu ore minerals providing soluble copper at cañariaco, peru through Cu isotope analyses of batch leach experiments. Hydrometallurgy 2010, 101, 15–19. [Google Scholar] [CrossRef]
  58. Borrok, D.M.; Ridley, W.I.; Wanty, R.B. Isotopic fractionation of Cu and Zn during adsorption onto bacterial surfaces. Geochim. Cosmochim. Acta Suppl. 2008, 72, A99. [Google Scholar] [CrossRef]
  59. Bigalke, M.; Weyer, S.; Wilcke, W. Stable Cu isotope fractionation in soils during oxic weathering and podzolization. Geochim. Cosmochim. Acta 2011, 75, 3119–3134. [Google Scholar] [CrossRef]
  60. Weber, F.A.; Voegelin, A.; Kaegi, R.; Kretzschmar, R. Contaminant mobilization by metallic copper and metal sulphide colloids in flooded soil. Nat. Geosci. 2009, 2, 267–271. [Google Scholar] [CrossRef]
  61. Bigalke, M.; Kersten, M.; Weyer, S.; Wilcke, W. Isotopes trace biogeochemistry and sources of Cu and Zn in an intertidal soil. Soil Sci. Soc. Am. J. 2013, 77, 680–691. [Google Scholar] [CrossRef]
  62. Bigalke, M.; Weyer, S.; Wilcke, W. Stable copper isotopes: A novel tool to trace copper behavior in hydromorphic soils. Soil Sci. Soc. Am. J. 2010, 74, 60–73. [Google Scholar] [CrossRef]
  63. Liu, S.-A.; Teng, F.-Z.; Li, S.; Wei, G.-J.; Ma, J.-L.; Li, D. Copper and iron isotope fractionation during weathering and pedogenesis: Insights from saprolite profiles. Geochim. Cosmochim. Acta 2014, 146, 59–75. [Google Scholar] [CrossRef]
  64. Kusonwiriyawong, C.; Bigalke, M.; Abgottspon, F.; Lazarov, M.; Wilcke, W. Response of cu partitioning to flooding: A δ65Cu approach in a carbonatic alluvial soil. Chem. Geol. 2016, 420, 69–76. [Google Scholar] [CrossRef]
  65. Lv, Y.; Liu, S.-A.; Zhu, J.-M.; Li, S. Copper and zinc isotope fractionation during deposition and weathering of highly metalliferous black shales in central China. Chem. Geol. 2016, 445, 24–35. [Google Scholar] [CrossRef]
  66. Mathur, R.; Jin, L.; Prush, V.; Paul, J.; Ebersole, C.; Fornadel, A.; Williams, J.Z.; Brantley, S. Cu isotopes and concentrations during weathering of black shale of the marcellus formation, Huntingdoncounty, Pennsylvania (USA). Chem. Geol. 2012, 304–305, 175–184. [Google Scholar] [CrossRef]
  67. Chiaradia, M. Copper enrichment in arc magmas controlled by overriding plate thickness. Nat. Geosci. 2014, 7, 43–46. [Google Scholar] [CrossRef]
  68. Mirnejad, H.; Mathur, R.; Einali, M.; Dendas, M.; Alirezaei, S. A comparative copper isotope study of porphyry copper deposits in Iran. Geochem. Exp. Environ. Anal. 2010, 10, 413–418. [Google Scholar] [CrossRef]
  69. Gale, N.H.; Woodhead, A.P.; Stos-Gale, Z.A.; Walder, A.; Bowen, I. Natural variations detected in the isotopic composition of copper: Possible applications to archaeology and geochemistry. Int. J. Mass Spectrom. 1999, 184, 1–9. [Google Scholar] [CrossRef]
  70. Dan, A.; Matthews, A.; Bar-Matthews, M.; Halicz, L. Copper isotope fractionation in sedimentary copper mineralization (Timna Valley, Israel). Chem. Geol. 2007, 243, 238–254. [Google Scholar] [CrossRef]
  71. Graham, S.; Pearson, N.; Jackson, S.; Griffin, W.; O’Reilly, S.Y. Tracing Cu and Fe from source to porphyry: In situ determination of Cu and Fe isotope ratios in sulfides from the grasberg Cu–Au deposit. Chem. Geol. 2004, 207, 147–169. [Google Scholar] [CrossRef]
  72. Li, W.; Jackson, S.E.; Pearson, N.J.; Alard, O.; Chappell, B.W. The Cu isotopic signature of granites from the lachlan Fold Belt, SE Australia. Chem. Geol. 2009, 258, 38–49. [Google Scholar] [CrossRef]
  73. Song, S.; Mathur, R.; Ruiz, J.; Chen, D.; Allin, N.; Guo, K.; Kang, W. Fingerprinting two metal contaminants in streams with Cu isotopes near the dexing mine, China. Sci. Total Environ. 2015, 544, 677–685. [Google Scholar] [CrossRef] [PubMed]
  74. Huang, J.; Liu, S.A.; Wörner, G.; Yu, H.; Xiao, Y. Copper isotope behavior during extreme magma differentiation and degassing: a case study on Laacher See phonolite tephra (East Eifel, Germany). Contrib. Mineral. Petrol. 2016, 171, 8–9. [Google Scholar] [CrossRef]
  75. Savage, P.S.; Moynier, F.; Chen, H.; Shofner, G.; Siebert, J.; Badro, J.; Puchtel, I. Copper isotope evidence for large-scale sulphide fractionation during earth’s differentiation. Geochem. Perspect. Lett. 2015, 1, 53–64. [Google Scholar] [CrossRef]
  76. Liu, S.A.; Huang, J.; Liu, J.; Wörner, G.; Yang, W.; Tang, Y.J.; Chen, Y.; Tang, L.; Zheng, J.; Li, S. Copper isotopic composition of the silicate earth. Earth Planet. Sci. Lett. 2015, 427, 95–103. [Google Scholar] [CrossRef]
  77. Bishop, M.C.; Moynier, F.; Weinstein, C.; Fraboulet, J.G.; Wang, K.; Foriel, J. The Cu isotopic composition of iron meteorites. Meteorit. Planet. Sci. 2012, 47, 268–276. [Google Scholar] [CrossRef]
  78. Herzog, G.; Moynier, F.; Albarède, F.; Berezhnoy, A. Isotopic and elemental abundances of copper and zinc in lunar samples, Zagami, Pele’s hairs, and a terrestrial basalt. Geochim. Cosmochim. Acta 2009, 73, 5884–5904. [Google Scholar] [CrossRef]
  79. Moynier, F.; Koeberl, C.; Beck, P.; Jourdan, F.; Telouk, P. Isotopic fractionation of Cu in tektites. Geochim. Cosmochim. Acta 2010, 74, 799–807. [Google Scholar] [CrossRef]
  80. Büchl, A.; Hawkesworth, C.J.; Ragnarsdottir, K.V.; Brown, D.R. Re-partitioning of Cu and Zn isotopes by modified protein expression. Geochem. Trans. 2008, 9, 127–136. [Google Scholar] [CrossRef] [PubMed]
  81. Jaouen, K.; Balter, V.; Herrscher, E.; Lamboux, A.; Telouk, P.; Albarede, F. Fe and Cu stable isotopes in archeological human bones and their relationship to sex. Am. J. Phys. Anthropol. 2012, 148, 334–340. [Google Scholar] [CrossRef] [PubMed]
  82. Novak, M.; Sipkova, A.; Chrastny, V.; Stepanova, M.; Voldrichova, P.; Veselovsky, F.; Prechova, E.; Blaha, V.; Curik, J.; Farkas, J. Cu-Zn isotope constraints on the provenance of air pollution in central Europe: Using soluble and insoluble particles in snow and rime. Environ. Pollut. 2016, 218, 1135–1146. [Google Scholar] [CrossRef] [PubMed]
  83. Little, S.H.; Vance, D.; Walker-Brown, C.; Landing, W.M. The oceanic mass balance of copper and zinc isotopes, investigated by analysis of their inputs, and outputs to ferromanganese oxide sediments. Geochim. Cosmochim. Acta 2014, 125, 673–693. [Google Scholar] [CrossRef]
  84. Babcsányi, I.; Imfeld, G.; Granet, M.; Chabaux, F. Copper stable isotopes to trace copper behavior in wetland systems. Environ. Sci. Technol. 2014, 48, 5520–5529. [Google Scholar] [CrossRef] [PubMed]
  85. Petit, J.C.J.; Schäfer, J.; Coynel, A.; Blanc, G.; Deycard, V.N.; Derriennic, H.; Lanceleur, L.; Dutruch, L.; Bossy, C.; Mattielli, N. Anthropogenic sources and biogeochemical reactivity of particulate and dissolved Cu isotopes in the turbidity gradient of the garonne river (France). Chem. Geol. 2013, 359, 125–135. [Google Scholar] [CrossRef]
  86. Mathur, R.; Titley, S.; Barra, F.; Brantley, S.; Wilson, M.; Phillips, A.; Munizaga, F.; Maksaev, V.; Vervoort, J.; Hart, G. Exploration potential of Cu isotope fractionation in porphyry copper deposits. J. Geochem. Explor. 2009, 102, 1–6. [Google Scholar] [CrossRef]
  87. Mathur, R.; Munk, L.; Nguyen, M.; Gregory, M.; Annell, H.; Lang, J. Modern and paleofluid pathways revealed by Cu isotope compositions in surface waters and ores of the pebble porphyry cu-au-mo deposit, alaska. Econ. Geol. 2013, 108, 529–541. [Google Scholar] [CrossRef]
  88. El Azzi, D.; Viers, J.; Guiresse, M.; Probst, A.; Aubert, D.; Caparros, J.; Charles, F.; Guizien, K.; Probst, J.-L. Origin and fate of copper in a small mediterranean vineyard catchment: New insights from combined chemical extraction and δ65Cu isotopic composition. Sci. Total Environ. 2013, 463, 91–101. [Google Scholar] [CrossRef] [PubMed]
  89. Vance, D.; Archer, C.; Bermin, J.; Perkins, J.; Statham, P.; Lohan, M.; Ellwood, M.; Mills, R. The copper isotope geochemistry of rivers and the oceans. Earth Planet. Sci. Lett. 2008, 274, 204–213. [Google Scholar] [CrossRef]
  90. Asael, D.; Matthews, A.; Bar-Matthews, M.; Harlavan, Y.; Segal, I. Tracking redox controls and sources of sedimentary mineralization using copper and lead isotopes. Chem. Geol. 2012, 310–311, 23–35. [Google Scholar] [CrossRef]
  91. Molnár, F.; Mänttäri, I.; O’Brien, H.; Lahaye, Y.; Pakkanen, L.; Johanson, B.; Käpyaho, A.; Sorjonen-Ward, P.; Whitehouse, M.; Sakellaris, G. Boron, sulphur and copper isotope systematics in the orogenic gold deposits of the archaean hattu schist belt, eastern finland. Ore Geol. Rev. 2016, 77, 133–162. [Google Scholar] [CrossRef]
  92. Huang, J.; Liu, S.A.; Gao, Y.; Xiao, Y.; Chen, S. Copper and Zinc isotope systematics of altered oceanic crust at IODP site 1256 in the eastern equatorial pacific. J. Geophys. Res. Solid Earth 2016, 121, 7086–7100. [Google Scholar] [CrossRef]
  93. Dekov, V.M.; Rouxel, O.; Asael, D.; Hålenius, U.; Munnik, F. Native cu from the oceanic crust: Isotopic insights into native metal origin. Chem. Geol. 2013, 359, 136–149. [Google Scholar] [CrossRef]
  94. Yang, M.J.; Yang, X.E.; Römheld, V. Growth and nutrient composition of elsholtziasplendensnakai under copper toxicity. J. Plant Nutr. 2002, 25, 1359–1375. [Google Scholar] [CrossRef]
  95. Jouvin, D.; Weiss, D.J.; Mason, T.F.; Bravin, M.N.; Louvat, P.; Zhao, F.; Ferec, F.; Hinsinger, P.; Benedetti, M.F. Stable isotopes of Cu and Zn in higher plants: Evidence for Cu reduction at the root surface and two conceptual models for isotopic fractionation processes. Environ. Sci. Technol. 2012, 46, 2652–2660. [Google Scholar] [CrossRef] [PubMed]
  96. Zhu, X.; Li, S.; Luo, Y.; Wu, L. Copper isotope fractionation by higher plants. Geochim. Cosmochim. Acta 2010, 74, A1234. [Google Scholar] [CrossRef]
  97. Ryan, B.M.; Kirby, J.K.; Degryse, F.; Harris, H.; McLaughlin, M.J.; Scheiderich, K. Copper speciation and isotopic fractionation in plants: Uptake and translocation mechanisms. New Phytol. 2013, 199, 367–378. [Google Scholar] [CrossRef] [PubMed]
  98. Weinstein, C.; Moynier, F.; Wang, K.; Paniello, R.; Foriel, J.; Catalano, J.; Pichat, S. Isotopic fractionation of Cu in plants. Chem. Geol. 2011, 286, 266–271. [Google Scholar] [CrossRef]
  99. Ilina, S.M.; Viers, J.; Lapitsky, S.A.; Mialle, S.; Mavromatis, V.; Chmeleff, J.; Brunet, P.; Alekhin, Y.V.; Isnard, H.; Pokrovsky, O.S. Stable (Cu, Mg) and radiogenic (Sr, Nd) isotope fractionation in colloids of boreal organic-rich waters. Chem. Geol. 2013, 342, 63–75. [Google Scholar] [CrossRef]
  100. Rodríguez, N.P.; Engström, E.; Rodushkin, I.; Nason, P.; Alakangas, L.; Öhlander, B. Copper and iron isotope fractionation in mine tailings at the laver and kristineberg mines, northern Sweden. Appl. Geochem. 2013, 32, 204–215. [Google Scholar] [CrossRef]
  101. Komjarova, I.; Bury, N.R. Evidence of common cadmium and copper uptake routes in ZebrafishDaniorerio. Environ. Sci. Technol. 2014, 48, 12946–12951. [Google Scholar] [CrossRef] [PubMed]
  102. Balter, V.; Lamboux, A.; Zazzo, A.; Telouk, P.; Leverrier, Y.; Marvel, J.; Moloney, A.P.; Monahan, F.J.; Schmidt, O.; Albarede, F. Contrasting Cu, Fe, and Zn isotopic patterns in organs and body fluids of mice and sheep, with emphasis on cellular fractionation. Metallomics 2013, 5, 1470–1482. [Google Scholar] [CrossRef] [PubMed]
  103. Jaouen, K.; Gibert, M.; Lamboux, A.; Telouk, P.; Fourel, F.; Albarede, F.; Alekseev, A.N.; Crubezy, E.; Balter, V. Is aging recorded in blood Cu and Zn isotope compositions? Metallomics 2013, 5, 1016–1024. [Google Scholar] [CrossRef] [PubMed]
  104. Van Heghe, L.; Engström, E.; Rodushkin, I.; Cloquet, C.; Vanhaecke, F. Isotopic analysis of the metabolically relevant transition metals Cu, Fe and Zn in human blood from vegetarians and omnivores using multi-collector icp-mass spectrometry. J. Anal. At. Spectrom. 2012, 27, 1327. [Google Scholar] [CrossRef]
  105. Albarede, F.; Telouk, P.; Lamboux, A.; Jaouen, K.; Balter, V. Isotopic evidence of unaccounted for Fe and Cu erythropoietic pathways. Metallomics 2011, 3, 926–933. [Google Scholar] [CrossRef] [PubMed]
  106. Balter, V.; da Costa, A.N.; Bondanese, V.P.; Jaouen, K.; Lamboux, A.; Sangrajrang, S.; Vincent, N.; Fourel, F.; Télouk, P.; Gigou, M. Natural variations of copper and sulfur stable isotopes in blood of hepatocellular carcinoma patients. Proc. Natl. Acad. Sci. USA 2015, 112, 982–985. [Google Scholar] [CrossRef] [PubMed]
  107. Télouk, P.; Puisieux, A.; Fujii, T.; Balter, V.; Bondanese, V.P.; Morel, A.-P.; Clapisson, G.; Lamboux, A.; Albarede, F. Copper isotope effect in serum of cancer patients. A pilot study. Metallomics 2015, 7, 299–308. [Google Scholar] [CrossRef] [PubMed]
  108. Jaouen, K.; Balter, V. Menopause effect on blood Fe and Cu isotope compositions. Am. J. Phys. Anthropol. 2014, 153, 280–285. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Fractionation(Δ65Cu) of Cu isotopesin low temperature processes [37,38,39,40,42,43,44,45,47,48,50,52,53,54,56,57].
Figure 1. Fractionation(Δ65Cu) of Cu isotopesin low temperature processes [37,38,39,40,42,43,44,45,47,48,50,52,53,54,56,57].
Ijerph 14 00538 g001
Figure 2. The δ65Cu (‰) values of various natural materials [6,7,15,19,20,21,26,32,40,59,61,63,64,65,66,68,70,72,73,76,80,81,82,83,84,85,86,87,88,89,90,91,92,93].
Figure 2. The δ65Cu (‰) values of various natural materials [6,7,15,19,20,21,26,32,40,59,61,63,64,65,66,68,70,72,73,76,80,81,82,83,84,85,86,87,88,89,90,91,92,93].
Ijerph 14 00538 g002
Figure 3. Cu isotopic fractionation factors (Δ65Cu) during soil-plant uptake and within-plant processes [18,84,95,97].
Figure 3. Cu isotopic fractionation factors (Δ65Cu) during soil-plant uptake and within-plant processes [18,84,95,97].
Ijerph 14 00538 g003
Table 1. Cu isotope composition in blood and bones of Human beings.
Table 1. Cu isotope composition in blood and bones of Human beings.
Materialδ65Cu Variation (‰)Average δ65Cu (‰, 2SD)References
Healthy women serum−0.80~0.04−0.28 ± 0.41[105]
Healthy men serum−0.64~0.06−0.28 ± 0.40
Healthy womenerythrocytes−0.04~0.800.46 ± 0.47
Healthy menerythrocytes0.23~0.910.67 ± 0.36
Healthy womentotal blood−0.52~0.320.00 ± 0.41
Healthy mentotal blood−0.21~0.430.16 ± 0.33
Cancer patients (HCC) serum−0.66~0.47−0.02 ± 0.54[106]
Control group serum−0.39~0.380.10 ± 0.45
Cancer patients (HCC) red blood cell−0.07~0.920.51 ± 0.56
Control group red blood cell0.57~1.240.88 ± 0.44
Breast cancer patients serum−1.45~0.12−0.51 ± 0.52[107]
Colorectal cancer patients serum−0.65~0.04−0.29 ± 0.30
Aging men blood 0.68 ± 0.49[108]
Control group (Young men) 0.67 ± 0.36
Postmenopausal women 0.71 ± 0.54
Control group (Premenopausal women) 0.43 ± 0.48
Liver −0.26 ± 0.22
Vegetarian female blood−0.75~−0.29−0.51 ± 0.46[104]
Vegetarian male blood−0.22~0.23−0.07 ± 0.52
Omnivorous female blood−0.14~0.17−0.02 ± 0.34
Omnivorous male blood−0.28~0.09−0.05 ± 0.41
Russian and Yakut blood−1.37~−0.22−0.68 ± 0.62[81]
Archeological women bones −0.20 ± 0.25[103]
Archeological men bones −0.11 ± 0.16

Share and Cite

MDPI and ACS Style

Wang, Z.; Chen, J.; Zhang, T. Cu Isotopic Composition in Surface Environments and in Biological Systems: A Critical Review. Int. J. Environ. Res. Public Health 2017, 14, 538. https://doi.org/10.3390/ijerph14050538

AMA Style

Wang Z, Chen J, Zhang T. Cu Isotopic Composition in Surface Environments and in Biological Systems: A Critical Review. International Journal of Environmental Research and Public Health. 2017; 14(5):538. https://doi.org/10.3390/ijerph14050538

Chicago/Turabian Style

Wang, Zhuhong, Jiubin Chen, and Ting Zhang. 2017. "Cu Isotopic Composition in Surface Environments and in Biological Systems: A Critical Review" International Journal of Environmental Research and Public Health 14, no. 5: 538. https://doi.org/10.3390/ijerph14050538

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop