Content uploaded by Kirsti Loukola-Ruskeeniemi
Author content
All content in this area was uploaded by Kirsti Loukola-Ruskeeniemi on Mar 14, 2019
Content may be subject to copyright.
Content uploaded by Annika Parviainen
Author content
All content in this area was uploaded by Annika Parviainen on Mar 13, 2019
Content may be subject to copyright.
Contents lists available at ScienceDirect
Earth-Science Reviews
journal homepage: www.elsevier.com/locate/earscirev
Environmental impact of mineralised black shales
Annika Parviainen
a,b
, Kirsti Loukola-Ruskeeniemi
a,⁎
a
Geological Survey of Finland, Vuorimiehentie 5, FI-02150 Espoo, Finland
b
Universidad de Granada, Instituto Andaluz de Ciencias de la Tierra (IACT, UGR-CSIC), Avda. de las Palmeras 4, Armilla, Granada E-18100, Spain
ARTICLE INFO
Keywords:
Black shale
Sulphides
Organic carbon
Acid mine drainage
Surface water
Regional planning
Mining
Talvivaara
ABSTRACT
Black shales are sedimentary rocks containing > 0.5% of organic carbon. They host polymetallic deposits which
have been mined for Cu, Ni, Zn, Mn, P, Mo, V, U, Au and PGE (platinum group elements). Even sub-economic
occurrences provide potential risk of acid rock drainage when exposed to oxic surface environment. The natural
acid neutralisation potential varies depending on the adjacent rock units, especially on the presence of calcar-
eous rocks. The chemical and mineralogical composition of black shale is reflected in the quality of the surface
waters and groundwater. Cu, Ni, Pb, U and Zn are recognised as major pollutants though the environmental
impact is characteristically polymetallic just like the black shale occurrences. Hence, the environmental impacts
have to be evaluated in each occurrence.
The Proterozoic Ni–Zn–Cu–Co deposit at Talvivaara, Finland, is reviewed in more detail as an example of a
large, low-grade deposit that is currently exploited with open pit mining and a bioleaching process, together
with the Proterozoic Central African Copperbelt, the Cambrian U-Mo deposits in Sweden, the Cambrian
Ni–Mo–PGE deposits in China and the Cambrian-Ordovician U deposits in South-Korea, the Devonian
Ni–Zn–PGE occurrences in Yukon, Canada, and Kentucky, USA, and the Permian Cu-Ag deposits in Poland and
Germany. The mineralised horizons may be merely few centimetres thick like in Yukon or hundreds of metres
thick like at Talvivaara. Both natural and anthropogenic environmental impacts of black shales are reviewed
world-wide, and based on the overview of the state-of-the-art an integrated research approach is suggested for
the comprehensive assessment of the risk.
Black shales are natural sources of soil and water contamination. At Talvivaara, the geochemical background
includes higher than average concentrations of Ni, Cu, Zn and Mn in glacial till, peat, surface waters and
groundwater as well as in stream and lake sediments. Bioaccumulation by plants has been reported in China and
Korea. Even endemic diseases have been proposed to be linked with the contamination derived from the
weathering and leaching of harmful elements from black shale. Anthropogenic actions exposing the black shale
bedrock and associated soils to oxic conditions further intensifies acid rock drainage.
Regional or nation-wide mapping of the black shales is recommended to detect potential risk areas. Finland
has recently completed the country-wide mapping program of black shales with airborne geophysics integrated
with geological, petrophysical and geochemical studies. The black shale database is actively used in regional
planning and by environmental authorities, research institutes and consulting companies. In the case of the
historical black shale mining areas like in the Kupferschiefer in Germany, restoration measures have been ap-
plied to prevent further acid mine drainage. In active and future mining projects, a comprehensive environ-
mental impact assessment with effective monitoring programmes and closure plans play a crucial role in the
prevention of acid mine drainage from the black shale -associated deposits.
1. Introduction
Sulphide oxidation at oxygenated environment and subsequent
migration of trace elements and sulphate from the sulphide-rich bed-
rock and soil to surface waters, known as acid rock drainage, has gained
attention in the past decades (e.g.,Blowes et al., 2003;Nordstrom,
2011). The public awareness of geogenic contamination has grown as a
consequence of the potential human exposure. Chronic exposure even
to low concentrations of As and Cr derived from soil has been suggested
to elevate the risk of developing certain types of cancer (Zhao et al.,
2014;Núñez et al., 2016). The results from a black shale area with
elevated Ni and low Ca concentrations in bedrock and soil indicate that
https://doi.org/10.1016/j.earscirev.2019.01.017
Received 14 September 2018; Received in revised form 19 January 2019; Accepted 20 January 2019
⁎
Corresponding author.
E-mail addresses: aparviainen@iact.ugr-csic.es (A. Parviainen), kirsti.loukola-ruskeeniemi@gtk.fi (K. Loukola-Ruskeeniemi).
Earth-Science Reviews 192 (2019) 65–90
Available online 25 January 2019
0012-8252/ © 2019 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/BY-NC-ND/4.0/).
T
the use of dug well water, as well as the consumption of local mush-
rooms, slightly affects the Ca concentration of the blood serum of local
residents. Low Ca values in the black shale bedrock and soil are re-
flected as lower Ca concentrations in blood serum than in the reference
area in the same municipality with a lower sulphide and higher Ca
content in the bedrock and the overlying glacial till (Loukola-
Ruskeeniemi et al., 2003;Kantola et al., 2000, 2008;Kousa et al.,
2011).
Black shales commonly host sub-economic, low-grade sulphide oc-
currences and may provide a source of base and precious metals, U, Mo,
Ni, Mn, V, Hg, Sb, and W (Fan, 1983;Coveney and Glascock, 1989;
Oszczepalski, 1989;Grauch and Huyck, 1990;Coveney and Chen,
1991;Pašava, 1993;Loukola-Ruskeeniemi et al., 1991;Leventhal,
1993;Wignall, 1994;Pašava et al., 1996;Loukola-Ruskeeniemi, 1999;
Coveney, 2000;Fan et al., 2004;Coveney and Pašava, 2004;Shpirt
et al., 2007;Polgári et al., 2012). As black shales are enriched in a
variety of harmful elements, including radionuclides, they represent
potential sources of soil, groundwater and surface water contamination.
The abundance of organic matter and sulphide minerals makes
black shale susceptible to chemical weathering (Blowes et al., 2003;
Nordstrom, 2011). When black shale outcrops, it is exposed to oxic
conditions. Hence, sulphides are prone to oxidation and dissolution,
liberating protons and trace metals (Jambor, 1994). As a consequence
of low-pH drainage, even the more resistant silicate mineral phases are
susceptible to alteration and dissolution, contributing to the metal load
of the drainage waters. Fractures in the rock promote their weathering
(Jin et al., 2013). The natural weathering processes are enhanced by
anthropogenic actions that bring the black shale to the surface, such as
mining, excavation, road construction and forestry. Therefore, as with
any anthropogenic actions on sulphide-rich rocks, the disturbance of
black shale enhances the release and mobility of the associated metals
(Nordstrom, 2011).
Even though not within the scope of this work, it is worth to men-
tion that, in the geological time scale, black shale plays a role in the
carbon cycle and the oxygenation and de‑oxygenation of the environ-
ment.
Characterisation of the sulphide mineralogy and rock assemblages
of each black shale deposit and occurrence is needed to evaluate their
pollution risk. In this review, we compare eight well-known black shale
deposits where environmental studies have been carried out and where
the concentrations of both S and C
org
in the black shale exceeds 1%,
namely
•the Proterozoic Talvivaara Ni–Zn–Co–Cu deposit in Finland,
•the Proterozoic Cu-Co and Ni deposits in Zambia, in the Central
African Copperbelt,
•the Cambrian U-Mo occurrences in the alum shale units in Sweden,
•the Cambrian Niutitang Ni–Mo–platinum group element (PGE) de-
posits in China,
•the Cambrian-Ordovician Okchon U deposit in South Korea,
•the Devonian Ni–Zn–PGE occurrences in Yukon, Canada,
•the Devonian U-bearing black shale occurrences in Kentucky, USA,
and
•the Permian Kupferschiefer Cu-Ag deposits in Poland and Germany
(please see the references for each deposit in the figure caption of
Fig. 1 and in more detail in sections 5–7 below).
2. Materials and methods
We have compiled data for the selected black shale units in Tables 1
and 2that work as the core of this review. The selected mineralised
black shale occurrences contain > 1% of both organic carbon and sul-
phur and either the black shale layer or organic matter in general has
played an important role in the ore formation processes. A brief de-
scription of the different black shale occurrences is necessary to
understand the geochemical processes and to evaluate their environ-
mental impacts as the thickness, ore mineralogy and the composition of
wall rocks control the potential acid rock drainage. The numeric data is
presented in tables and figures. Sampling and analytical methods for
the data presented in Tables 1–3 are given in detail in the original
publications cited.
In Section 5 below, we describe the environmental impacts of the
Talvivaara black shale deposit in Finland in both pre-mining and active
mining phases. Large geochemical databases serve as a basis for geo-
chemical baseline studies. They were acquired with nation-wide sur-
veys and are maintained by the Geological Survey of Finland (Salminen
and Tarvainen, 1995). Before the nation-wide mapping program
started, pilot studies were carried out in the Talvivaara region because
the chemical characteristics of the bedrock and overlying glacial till
exhibit contrasting geochemical characteristics for black shale and the
adjacent quartzite and granitoid areas (Gustavsson et al., 2012). These
data have subsequently been applied to depict the geochemical back-
ground of Ni and Cu in glacial till, stream sediments and stream water.
The exploitation of the Talvivaara Ni–Zn–Cu–Co deposit began in 2008,
and sampling performed before that is thus considered to represent pre-
mining conditions. Thereafter, the evaluation of the environmental
impacts of mining activities is based on scientific articles and public
monitoring data provided by the mining company in their web-pages.
The present paper is the result of team work where the main re-
sponsibilities were divided as follows: Annika Parviainen was re-
sponsible for the literature study of the environmental impacts and the
compilation of the tables and Fig. 7 while Kirsti Loukola-Ruskeeniemi
was responsible for the compilation of Figs. 1–6. Kirsti Loukola-Rus-
keeniemi was responsible for the black shale research, both geochem-
ical and environmental, in the Talvivaara black shale deposit and other
sites in Finland. The nation-wide black shale mapping program of
Finland was carried out by geophysicists Eija Hyvönen, Hilkka Ar-
kimaa, Jouni Lerssi and Meri-Liisa Airo, geologist Kirsti Loukola-Rus-
keeniemi and research assistant Satu Vuoriainen from the Geological
Survey of Finland. The international review is based on Kirsti Loukola-
Ruskeeniemi's experience from international black shale projects during
1987–2008: the UNESCO International Geological Correlation Pro-
gramme Projects (IGCP) 254, 357 and 429 related to mineralised black
shales and their environmental impacts (e.g.,Pašava and Gabriel,
1988), project “Ore deposits in black shale basins” between Finland and
the Soviet Union /Russia and the European Union research project
BIOSHALE which compared ore geology and environmental impacts of
black shales in the Talvivaara deposit in Finland with the Lubin mine in
Poland (http://www.brgm.eu/project/bioshale-search-sustainable-
way-of-exploiting-black-shale-ores-using-biotechnologies).
3. Black shales and the associated sulphide deposits
Black shales are dark-coloured mudrocks containing organic matter
and silt- and clay-size mineral grains that accumulated together
(Swanson, 1961). In the IGCP 254 of UNESCO titled “Metalliferous
Black Shales”, conducted during 1987–1992, a definition was agreed
that a black shale should contain > 0.5% organic carbon (Huyck,
1990). Many researchers were of the opinion that a black shale should
contain over 1% of both organic carbon and sulphur to represent spe-
cific depositional conditions but since in a black shale unit the content
of organic carbon can vary, 0.5% was agreed to represent the minimum
concentration within the unit. Black shales can be classified due to their
Ca, Al and Si content (Loukola-Ruskeeniemi, 1992) and they comprise a
heterogeneous group of polymetallic rocks (e.g.,Tourtelot, 1979).
Based on the U. S. Geological Survey black shale standard SDO-1 (De-
vonian Ohio Shale from Kentucky), the definition of a mineralised, so
called ‘metalliferous’ black shale is (Huyck, 1990): “A metalliferous
black shale is a black shale that is enriched in any given metal by a
factor 2x (except Be, Co, Mo, and U, for which 1x is sufficient) relative
to the U.S. Geological Survey Standard SDO-1.”
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
66
The unifying feature of black shale is deposition as fine-grained,
laminated sediment on the sea bottom with a high supply of organic
matter. A key issue in the formation of black shale is the preservation of
organic matter in the sediment, which usually requires deposition in
low-oxygen or euxinic bottom waters. However, formation in oxic
conditions has been recently described as well (Reynaud et al., 2018).
At present, the deposition of organic- and sulphur-rich mud occurs, for
example, in the Black Sea (Kaiser et al., 2017). Piper and Calvert (2009)
studied two modern marine basins with anoxic bottom conditions. The
hydrography and trace metal and phytoplankton nutrient budgets ex-
hibit that the rate of burial of labile organic matter promotes anoxic
conditions in the sediment pore waters that enhances the retention of
trace metals deposited from the water column. High supply of clay
minerals (up to 30%) with large reactive surface area contribute to the
sorption and preservation of organic matter and metals (Kennedy et al.,
2002). Macquaker and Bohacs (2007) show that the deposition and
burial of mud is a more dynamic and complex process than traditionally
thought, and findings from hydrocarbon exploration in deep sea
(> 200 m depth) indicate that mud rich in organic carbon is formed by
a combination of processes such as pelagic settling, hemipelagic de-
position and turbidity currents.
The sources of metals include seawater-derived material (e.g.,Mao
et al., 2002;Lehmann et al., 2003, 2007 and 2016;Wallis, 2007;Xu
et al., 2012, 2013;Pagès et al., 2018;Morel and Price, 2003;Piper and
Calvert, 2009;Slack et al., 2015) and hydrothermal and/or multiple
metal sources (Loukola-Ruskeeniemi, 1991;Hulbert et al., 1992;
Coveney et al., 1994;Lott et al., 1999;Steiner et al., 2001;Jiang et al.,
2007;Pašava et al., 2008;Och et al., 2013;Young et al., 2013). The
post-depositional diagenetic and metamorphic processes have an im-
pact on the mineralogy of black shale and the distribution of trace
elements (e.g.,Piper and Calvert, 2009;Rimstidt et al., 2017). In ad-
dition, organic compounds and biomineralization (mineral production
by organisms) may have a role in the genesis of the ore and fluxes at the
sediment-water interface during early diagenesis may have an impact
on the trace element concentrations (Piper and Calvert, 2009;Rimstidt
et al., 2017;Pérez-Huerta et al., 2018). Fine-grained organic-rich mud
with low permeability and a locally reducing environment may capture
the ore-forming fluids underneath and the hydrothermal fluids may
then precipitate in and below this trap. As a potential zone of weakness
during tectonic deformation, a black shale unit may act as a thrust
surface and locally allow the flow of fluids (Loukola-Ruskeeniemi,
1992). Moreover, microbial activity, specifically sulphate-reducing
bacteria, may trigger the precipitation of sulphide minerals, and the
accumulation of metals may be promoted by porphyrins that bind
metals to form complexes (Greenwood et al., 2013).
Black shales host sulphide deposits but are also associated with
many other types of deposits because both black shale and massive
sulphide deposit in most cases require isolation from oxic conditions to
be preserved. In addition to sulphide deposits, black shales are a source
rock for hydrocarbons (e.g.,Sethi and Schieber, 1998) and shale gas
deposits (Zou, 2013).
For the review, we selected eight black shale deposits and occur-
rences containing > 1% organic carbon and over 1% sulphur. The de-
posits represent different metamorphic grades and different geological
terrains. In the Sections 5, 6 and 7 below, we have divided the deposits
into groups according to the main metal (Ni, Cu, U), though the de-
posits are actually polymetallic (Table 1).
4. Weathering of black shale under natural conditions and due to
anthropogenic actions
Black shale occurrences should be considered as sources of acid rock
drainage when they are exposed to oxic conditions. Natural weathering
profiles and excavations have been studied in black shale areas
worldwide, for example in China, Korea, Sweden and the United States
(Falk et al., 2006;Fang et al., 2002;Lavergren et al., 2009b,b;Ling
et al., 2015;Liu et al., 2017;Park et al., 2010;Peng et al., 2004, 2014;
Petsch et al., 2005;Tang et al., 2018;Xu et al., 2013;Yu et al., 2012).
These studies show that black shales are sources of harmful elements in
soils and adjacent water systems, and translocation occurs to plants and
agricultural products cultivated on black shale -associated soils.
Recent studies demonstrate that besides chemical and mechanical
Proterozoic
Finland
Ni–Zn–Co–Cu
50 m -
300 m
Proterozoic
Zambia
Cu–Co
0 m -
35 m
Devonian
0.4 m
Canada
Ni–Zn–PGE
1.2 m
Permian
Poland
Cu–Ag
China
Ni–Mo–PGE
Cambrian
0.6 m
Cambrian
Sweden
U–Mo
20 m -
100 m
25 m
Scale
Mineralised
Black shale
Black shale
Legend
Limestone/
Dolomite
Arenite and
Shales
Arenite/
Sandstone
Shale
Chert
Fig. 1. The relative maximum thickness of mineralised black shale in typical profiles of the Palaeoproterozoic Talvivaara Ni–Zn–Cu–Co deposit in Finland (modified
from Loukola-Ruskeeniemi and Lahtinen, 2013), Proterozoic Cu-Co black shale in Zambia (modified from McGowan et al., 2006), Cambrian U-Mo alum shales in
Sweden (modified from Andersson et al., 1985), Cambrian Ni–Mo–PGE black shale in China (modified from, e.g., Han et al., 2015), Devonian Ni–Zn–PGE black shale
in Yukon (modified from Hulbert et al., 1992), and Permian Cu-Ag black shale in Poland (modified from, e.g.,Pieczonka et al., 2015).
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
67
Table 1
Age, depositional environment and thickness of selected black shale deposits around the world. Additionally, mean (unless mentioned otherwise), min. and max. values of organic C, S and metals of main economic
interest and main ore forming minerals are given.
Location and name of
the black shale deposit
Age
Depositional
environment
Intercalating rocks Thickness of the
sequence
Mean organic C and S content
(Min–Max)
Main sulphides and U
minerals
Estimated ore
reserves
Element concentrations in black
shales (Min–Max)
Finland
Talvivaara
Palaeoproterozoic
Stratified oxic/anoxic,
restricted marine basin
Black meta-carbonate
rocks (calc-silicate
rocks rich in graphite
and tremolite)
50–800 m Organic C:
Median 8.2% (2.4–13.2%)
1
S: Median 9.0% (2.4–26.0%)
1
Pyrite, pyrrhotite,
chalcopyrite, sphalerite,
galena, alabandite,
molybdenite, pentlandite,
ullmanite, stannite, uraninite,
thucholite
Ni–Zn–Co –Cu
2053 Mt (0.22% Ni,
0.13% Cu, 0.50% Zn,
and 0.02% Co)
2
Talvivaara ore (556
samples):median As 86 mg/kg
(4–891), Cd 16 (< 1–41), Co 189
(15–653), Cu 1340 (163–5190),
Mn 2750 (694–59,300), Mo 58
(6–96), Ni 2415 (102–6690), Pb
47 (5–437), U 17 (2–30), V 660
(120–1190), Zn 4955
(941–11,000)
1
China
Niutitang Formation
Cambrian
From a shallow shelf
environment to a deep
basinal facies of black shale
Phosphorite Mineralised layer
5–20 cm, rarely up to 2 m
Organic C:
•10% (0.7–15%)
3
•7.7% (2.6–11.5%)
4
S: -
Bravoite, vaesite, Pyrite,
Millerite, gersdorffite, MoSC
5
Ni–PGE–Mo–U
No estimates
available
Non-mineralised black shale: Ni
(28–1280 mg/kg), Mo (4.3–177)
3
Mineralised black shale: Ni up to
53400 mg/kg, Mo up to 87300
3
Mineralised black shale: As
8969 mg/kg (3340–18054), Cd
481(87–879), Co 281 (109–437),
Cu 2304 (740–3847), Mo 38508
(7346–84922), Ni 48345
(14688–70292), Pb 333
(137–599), Sb 284 (72–564), U
295 (73–520), V 189 (62–552), Zn
48345 (5532–116712)
4
Canada
Yukon
Devonian
Epicratonic marine basin,
proximal to the transition
zones between shallow and
deep water sedimentary
environments
− Ni–Zn–PGE-rich layer
from few cm to 40 cm
Organic C: 1.9% (1.3–2.5%)
6
S: 28% (20–33%)
6
Vaesite, pyrite, marcasite,
sphalerite, wurtzite,
melnicovite
Ni–Zn–PGE
No mining activities,
Other anthropogenic
actions (e.g., road
construction,
excavation)
Nick Property mineralization:
As (0.19–0.42%), Ba (0.2–0.5), Ni
5.3 (2.3–2.8), Mo (0.1–0.4), Re
(1.0–6.1), Se (0.06–0.24), V
(0.037–0.088), Zn 0.73 (0.4–1.2),
U (40–108 mg/kg)
6
Taiga Property: Up to Ni 5.21%,
Au 100 μg/kg, Pt 560, 244 Pd
7
Ni-rich black shale occurrences:
Ni 3.6 % (1.2–7.0), As 6037 mg/kg
(646–10690), Cd 21.6 (3.2–100),
Co 154 (35–330), Cu 298
(89–660), Mo 1995 (390–3300),
Pb 23.4 (7–58), Tl 217 (27–390),
U 150 (3.8–650), Zn 6515
(230–23000)
8
Poland and Germany
(Saxony-Anhalt)
Intracontinental or cratonic
Central European Basin
− Average 0.3 m with up to
1.2 m, footwall and
Organic C:
•0.3-30%
9
Bornite, chalcopyrite, pyrite,
chalcocite, covellite,
neodigenite, native Ag,
Germany
Cu–Ag
(Pb, Zn, Au, V, Mo,
Pitchy shale: Cu 7.1 wt.%
(1.8–39), Pb 0.1 (0.04–0.5), Zn
0.06 (0.8–0.1), Ag 286 mg/kg
(continued on next page)
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
68
Table 1 (continued)
Location and name of
the black shale deposit
Age
Depositional
environment
Intercalating rocks Thickness of the
sequence
Mean organic C and S content
(Min–Max)
Main sulphides and U
minerals
Estimated ore
reserves
Element concentrations in black
shales (Min–Max)
Kupferschiefer
Permian
hanging wall ore up to
50 m
Poland (7.5–8.0%)
Lubin 7.3%
S: -
galena, sphalerite, loellingite,
tedrahedrite, tennantite,
idaite etc.
Ni, Co, Se, Re, Cd, Tl,
Ge, and Te)
(242 Mt)
10
Poland
Cu–Ag–Au–Pb
(Ni–Co–Mo)
Lubin 379 Mt (Cu
1.32%, Ag 54.5 g/
Mg)
17
Rudna 432 Mt (Cu
1.88%, Ag 62.52 g/
Mg)
11
Polkowice-
Sieroszowice 409 Mt
(Cu 2.3% and As
62 g/Mg)
11
(120–3500), Au 1.5 (< 0.6–10),
Hg 15 (1–53), Pt 0.1 (0.01–0.9)
9
Zambia
Central African
Copperbelt
Proterozoic
Deep sea − •0.5–20 m (0.5–35 m)
12
•from a few m to 80 m,
lenses of dark
carbonaceous quartz-
rich rock vary from 1
to 10 m
13
Organic C:
•9–17%
13
S: -
•Pyrite, bornite,
chalcopyrite, chalcocite,
malachite, chrysocolla,
cuprite
12
•Pyrite, bravoite, vaesite,
millerite, chalcopyrite,
molybdenite, pyrrhotite,
carrollite
13
Cu
12
Ni (Cu−PGE)
13
40
Mt (1.07% Ni)
13
•Cu 2–3%, (up to 20%)
12
•Ni up to 17 %, Cu 2.5, Co 0.2, Fe
7, Mo 5000 mg/kg, U 20, Ir
24 μg/kg, Os 15, Ru 66, Rh 66
13
Sweden
Alum shales
Late Cambrian to
early Ordovician
Deep and shallow-marine
environment of the
Baltoscandian Platform
over an extensive period of
tectonic stability
Bituminous
limestones;
occasional chert
bands and
phosphorite nodules
Degerhamn 17 m,
Ranstad (Billingen)
22–24 m
Hällekis: 22 m
Organic C:
•10% (up to 20%)
14
•12% (5–20%)
15
•(2.7–20.6%)
16
Ranstad:
15%
17
Hällekis: 15%
(8–28%)
18
S:
•(4.6–8.4%)
14
•5.5% (4.6–10.4%)
15
•(3.4–8.26%)
16
Ranstad: 15%
17
Hällekis: 5.9%
(4.9–8.4%)
18
Pyrite, chalcopyrite,
sphalerite, galena, brannerite,
uraninite
Degerhamn: alum
salt
Ranstad: U
No estimates
available
•As (59–212 mg/kg), Co
(19–46), Cr (59–93), Cu
(103–224), Mo (104–396), Ni
(96–357), Pb (25–139), Th
(10–14), U (31–8000), V
(303–1464), Zn (36–1190)
15,19
•Co (22–39 mg/kg), Cr (55–210),
Cu (69–210), Mo (47–350), Ni
(110–510), Pb (28–140), U
(22–410), V (380–3100), Zn
(28–600)
16
Degerhamn: As 121 mg/kg
(91–138), Cd 5.5 (1.3–13), Cu 117
(97–158), Ni 107 (88–147), Zn 277
(77–623)
20
Ranstad: As 106 mg/kg, Cd 2.5, Cu
110, Mn 250, Mo 340, Ni 200, Pb 14,
Sb 5, U 300, Zn 130
17
South Korea
Okchon black shales
Cambrian-Ordovician
Okchon sea basin Coaly slates, lenses of
phosphatic nodules
20–40 m Organic C: 21%
S: 2.3%
Pyrite, pyrrhotite,
chalcopyrite, uranothorite,
U
No estimates
available
Okchon Metamorphic Belt: U
(~0.03%), V (~0.3%), Ba
(~1.4%) and Mo (~0.04%)
21
(continued on next page)
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
69
Table 1 (continued)
Location and name of
the black shale deposit
Age
Depositional
environment
Intercalating rocks Thickness of the
sequence
Mean organic C and S content
(Min–Max)
Main sulphides and U
minerals
Estimated ore
reserves
Element concentrations in black
shales (Min–Max)
brannerite, uraninite, ekanite,
and thorutite
Deog-Pyoung: Cd 6.3 mg/kg, Mo
136, Se 8.6
22
Chu-Bu: As 3.0 mg/kg (0.5–6.2),
Ba 1900 (240–4100), Cd 1.1
(0.5–3.9), Cu 38 (16–57), Mo 8
(1–23), Pb 17 (5–38), U 18
(5.4–47), V 199 (64–435), Zn 126
(27–417), Se 17.7 (1.5–33.9), S
451 (48–854)
23-24
Duk-Pyung: As 42 mg/kg
(5.0–110), Ba 2110 (110–6000),
Cd 10.9 (1.0–36), Cu 240
(120–430), Mo 213 (18–650), Pb
124 (19–440), U 83 (3–616), V
938 (70–1900), Zn 394 (42–1100),
Se 1.9 (1.6–2.3), S 215 (15–753)
23-
24
USA
Appalachian basin
black shales
Devonian
Euxinic sea basin Siltstone, sandstone
and limestone
25
0–30 m
25
Organic C: 3–6%
26
S: 2–5%
26
Pyrite, marcasite, U
No estimates
available
As 20–40 mg/kg, Co 20–40, Cu
40–70, Ni 80–150, U 10–40, V
150–300
26
1
Loukola-Ruskeeniemi and Lahtinen, 2013.
2
Talvivaara Mining Company, 2012.
3
Shi et al., 2014.
4
Han et al., 2015.
5
Kao et al., 2001.
6
Hulbert et al., 1992.
7
Butterworth and Caulfied, 1997.
8
Goodfellow et al., 2010.
9
Kucha, 1990.
10
Borg et al., 2012.
11
KGHM Polska Miedz (http://kghm.com/).
12
McGowan et al., 2016.
13
Capistrant et al., 2015.
14
Andersson et al., 1985.
15
Lecomte et al., 2017.
16
Lenventhal, 1991.
17
Allard et al., 1991.
18
Sanei et al., 2014.
19
Schovsbo, 2002.
20
Falk et al., 2006.
21
Kim et al., 2015.
22
Kim et al., 1993.
23
Lee et al., 1998.
24
Park et al., 2010.
25
Roen, 1984.
26
Leventhal and Hosterman, 1982.
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
70
Table 2
Environmental impacts of mining of black shale deposits. Information on mining activities as well as mean, min. and max. values of selected trace elements in mine
waste, soil, stream sediment, lake sediment, groundwater and surface water samples subjected to anthropogenic pollution in black shale areas are given. *Natural
contamination in Korea, as available data on anthropogenic pollution are scarce.
Location and name of the
deposit
Mining activity
Ore enrichment processes
Mean concentrations in mine waste or by-
products
(Min–Max)
Mean concentrations in anthropogenic
(*natural) soil/sediment pollution
(Min–Max)
Mean concentrations in (*natural) surface
water bodies and in groundwater
(Min–Max)
Finland
Talvivaara
2008 – present
Crushing, grinding, heap
bioleaching
Production waste: Cd 1.0 mg/kg, Co 10.0,
Cr 14, Cu 5.0, Fe 30,500, Mn 4840, Ni 420,
Pb 5.1, Th 1.3, U 58, Zn 160
1
Lake sediments: Fe 68,300 mg/kg
(47,000–88,000), Mn 3800 (2400–5100), Ni
760 (460–980), Zn 850 (370–1400), S 68,000
(43,000–110,000)
2
Groundwater: pH (4.4–8.8), SO
4
(< 0.3–9500 mg/L), As (< 0.4–3.3 μg/L), Cd
(< 0.1–2.2), Co (0.1–65), Cu (0.5–1200), Ni
(0.5–6400), Pb (0.25–1.3), Zn (0.25–20,000), Al
(1.5–6500), Fe (< 10–790,000), (U < 0.5–8.7)
3
Lake water (Salminen, Kalliojärvi): As
(< 1.0 μg/L), Cd (< 0.03-1.7), Cu (< 0.1-7.3),
Mn (86-37000), Ni (1.4-4700), Pb (< 0.5-1.4),
Zn (< 5.0-780), U (< 1.0-240)
3
Lake water (Kivijärvi, Lumijärvi, Ylä-
lumijärvi): As (< 1.0-42 μg/L), Cd (0.03-170),
Cu (< 0.1-170), Mn (600-85000), Ni (28-
1700000), Pb (< 0.5-45000)
3
Lake water (Kuusijärvi): pH 9.1 (6.5–11.2),
SO
4
3500 mg/L (1800–6100), Na 690 mg/L
(435–956), Co 4 μg/L (2–30), Cu 9 (0.6–80), Fe
370 (110–6910), Mn 263 (80–5850), Ni 36
(10–510), Zn 64 (10–890)
4
Lake water (Kalliojärvi and Ylä-lumijärvi):
pH 6.1 (4.4–7.5), SO
4
136 mg/L (42–330), Na
3.6 (1.39–457), Co 4 μg/L (<0.2–10), Cu 23
(< 1.2–30), Fe 2230 (80–18,600), Mn 3600
(110–30,500), Ni 19 (10–80), Zn 35
(< 20–130)
4
Stream water (Kalliojoki, Kivijoki,
Tuhkajoki): pH 6.2 (5.1–7.2), SO
4
328 mg/L
(44–473), Na 63 (4.9–282), Cu 20 μg/L
(< 0.12–40), Fe 921 (240–2010), Mn 449
(100–1350), Ni 11 (10–20), Zn 24 (10–70)
4
Drainage pipe: pH 7.9 (6.5–10.4), SO
4
2434 mg/L (1700–3620), Na 452 mg/L
(343–622), Cu 4.3 μg/L (2–40), Fe 4403
(180–837,000), Mn 1704 (170–303,000), Ni 27
(10–220), U, 0.02 (< 0.02–0.04), Zn 27
(10–60)
4
China
Niutitang Formation
Hunan: U mining
Zunyi: Ni–Mo mining,
1985 –
No reported data available
on the ore enrichment
processes
No reported data available Soil: As (21–110 mg/kg), Cd (0.4–1.4), Cu
(62–99), Hg (0.16–0.60), Mo (51–130), Ni
(64–199), Pb (19–25), Zn (139–270), V
(572–1221)
5
Surface water: pH 5.6 (3.0-7.8), As 1.0 μg/L
(0.3–4.2), Cd 41.7 (2.1–297), Co 51.7
(0.05–298), Cu 3.9 (0.2–9.3), Ni 261
(22.8–1531), Pb 0.4 (0.02-2.6), Zn 873
(12.6–4087), U 147 (0.01–1372)
6
Canada
Yukon, Nick Property
No reported mining
activity on black shale
deposits
No reported data available No reported data available Eagle Plain excavation, stream water: pH 3.1
(2.8–4.4), Al 66.3 (44–84 mg/L), Fe 133
(3.0–459), Mg 86 (6.0–105), Mn 2.6 (1.5–3.3),
Zn 4.8 (2.2–7.0)
7
Keno Hill stream water: pH 5.4–7.3, As
(80–220 μg/L), Cd (35–2720), Cr (2–10), Co
(15–149), Cu (< 1–94), Pb (5–131), Fe
(121–25,500), Mo (12–28), Ni (40–621), Zn
(3970–96,200)
8
Poland
Kupferschiefer
Lubin 1968 - (estimated
closure in 2059)
Rudna 1974 - (2045)
Polkowice-Sieroszowice
1968 - (2054)
Cu smelting
Zelazny Most tailings: As 25 mg/kg
(12–37), Cd (< 0.5–0.7), Cu 1309
(1300–1318), Co 16 (5–27), Cr 34 (31–37),
Fe 5550 (5200–5900), Ni 282 (275–288), Pb
245 (206–283), Zn 53 (30–75)
9
Zbiornik Gilow tailings: As 24 mg/kg
(19–29), Cd (< 0.5–0.5), Co 13 (12–14), Cu
1126 (1084–1165), Cr 60 (55–65), Ni 580
(567–592), Pb 221 (219–2222), Zn 47
(32–61)
10
Stream sediments: As (< 5–49 mg/kg), Cd
(< 0.5–2.9), Cu (22–4409), Co (2–62), Cr
(3–114), Fe (4000–200,500), Ni (88–826), Pb
(8–570), Zn (8–700)
9,10
Soil: Cd 4 mg/kg, Cu 7400, Cr 54, Mn 587, Ni
33, Pb 1960, Zn 675
11
No reported data available
(continued on next page)
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
71
Table 2 (continued)
Location and name of the
deposit
Mining activity
Ore enrichment processes
Mean concentrations in mine waste or by-
products
(Min–Max)
Mean concentrations in anthropogenic
(*natural) soil/sediment pollution
(Min–Max)
Mean concentrations in (*natural) surface
water bodies and in groundwater
(Min–Max)
Germany Kupferschiefer
1200–1969 (1990), mining
peak in mid-20
th
century
Cu smelting during the
mining peak
Theisen sludge: Cu 16 g/kg, Cd 0.54, Fe 24,
Zn 209, Pb 131
12-13
Theisen sludge: Zn 207 g/kg, Pb 35, Cu 11,
Cd 0.34, As 5.1, Ag 0.44, U 0.022, S 143
14
Mansfeld river sediments: As
(13–3200 mg/kg), Cu (92–11,400), Cd
(5–23), Ni (10–176), Pb (57–7510), Sb
(2–695), U (4–60), Zn (580–39,000)
14
East-Thuringia soil: As 28 mg/kg (9–112),
Co 49 (13–60), Cu 49 (24–149), Mo 5.3
(BDL–23), Ni 84 (48–170), Pb 19(9–23), S
4740 (936–14,012), U 13 (2–40), Zn 101
(80–164)
15
Sangerhausen surface water: pH (7.6–7.9), As
75 μg/L (< 50–90), Cu 180 (80–360), Pb 70
(60–80), Mo 53 (30–70), Ni 30 (< 30–30), Sb 8
(< 5–10), Zn 713 (410–1050)
16
Mansfeld surface water: pH (7.3–8.0), As
(< 0.5–3.4 μg/L), Cu (3.2–510), Cd (2.8–44), Ni
(3.8–110), Pb (< 0.5–940), Zn (< 10–17,860)
17
East-Thuringia groundwater: pH 4.1
(3.5–5.7), S 2208 mg/L (505–5601), As 4.8 μg/L
(< 0.05–33), Cd 115 (1.5–513), Co 5070
(48–20,550), Cu 669 (2–3433), Fe 3281
(< 20–29,370), Mn 183,067 (3767–808,000),
Ni 21,206 (540–69,580), Pb 2.1(< 0.1–13), U
338 (BDL–3263), Zn 4808 (48–20,360)
15
Zambia
Central African Copperbelt
•Cu-mining from the 20
th
century to present
•Mine development phase in
the Enterprise Ni deposit
Chambishi tailings: Co (790–6175 mg/kg),
Cu 795–9979, Fe (14067–321614)
18
Kitwe soil: pH 4.8 (4.1–8.0), S 400 mg/kg
(40–4500), As 3.1 (0.04–255), Co 57
(2.0–606), Cu 1501 (34–27,410), Pb 15
(4.0–480), Zn 34 (4.0–450)
19
Chambishi stream water: Ni up to 730 μg/L,
Zn 180
20
Chambishi groundwater: pH 9.5, SO
4
1820 mg/L, Co 12 μg/L, Cu 6, Fe 70, Mg 670,
Mn 13
18
Sweden
Alum shales
Skåne (S Sweden): Alum
mining from the 17
th
to the
20
th
century
Degerhamn: Alum mining
from the 18
th
to the 20
th
century
Kvarntorp: Hydrocarbons
1942–1966 (from
associated limestones)
Ranstad: U and V mining
1965-1976
Degerhamn: burning, water
leaching of alum salts,
Ranstad: crushing,
H
2
SO
4
leaching followed by
ion-exchange and solvent
extraction
21
Degerhamn: As 93 mg/kg (33–201), Cd 0.8
(0.2–2.2), Cu 77 (39–157), Ni 47 (21–115),
Zn 41 (7.0–88)
22
Ranstad: As 102 mg/kg, Cd 0.6, Cu 110, Mn
110, Mo 330, Ni 130, Pb 13, Sb 5, U 64, Zn
100
21
Kavarntorp: As 79 mg/kg, Mo 163, Ni 70, U
235, V 650
23
No reported data available Degerhamn groundwater: pH (4.1–7.3), As
(< 1–10 μg/L) Cd (< 0.05–18.7), Cu
(< 1–1050), Ni (0.5–862), Zn (4.3–1300)
22
Degerhamn groundwater: pH 6.5 (4.1–7.7), S
257 mg/L (8.9–662), As 2.5 μg/L (< 0.4–9.5),
Cd 3.1 (0.1–20), Co 98 (0.01–892), Cu 83
(0.2–552), Fe 487 (0.6–5650), Mn 500
(1.1–1860), Mo 37 (0.5–228), Ni 207
(0.1–1640), Pb 0.4 (< 0.01–2.5), Sr 425
(30–1830), U 54 (2.7–198), Zn 76 (1.7–1580)
24
Degerhamn groundwater: U 39 μg/L (0.2-196)
25
Degerhamn stream water: U 33 μg/L (13-82)
25
Ranstad leachates: pH (3–8.4), As
(< 2–43,000 μg/L), Cd (20–60), Cu (1–91,000),
Mn (480–87,000), Ni, (< 2–75,000), Mo
(2900–130,000), Zn (< 20–38,000), Pb
(< 1–15,000)
21
Kvarntorp groundwater: pH (3.2–12.2), Fe
(0.007–1130 mg/L), Mo (< 1–935), Ni
(< 0.5–2420), U (< 0.01–1760 μg/L)
23
Kvarntorp pit lake: pH 6.0 (3.2–7.8), As
0.70 μg/L (< 0.6–1.4), Cd 0.3 (0.02–1.1), Co 7.8
(0.4–27), Cr 0.5 (0.08–1.0), Cu 2.0 (0.6–4.0), Li
310 (220–480), Mo 5.3 (0.2–9.1), Ni 42.6
(1.3–130), Pb 0.7 (0.05–1.9), U 30.2 (4.9–83), V
4.4 (< 0.05–16), Zn 38.3 (2.2–140),
26
South Korea
Okchon black shales
Sporadic U mining
No reported data available
of the ore enrichment
processes
No reported data available No reported data available
*Soil: Cd 1.2 mg/kg, Mo 20, Se 1.5
27
*Soil, Chu-Bu: As 40 mg/kg (5–340), Ba
1570 (490–15,000), Cd 1.6 (0.2–20.1), Cu 64
(23–217), Mo 15 (1–240), Pb 48 (18–182), U
29 (2–450), Zn 172 (50–1100), S 4000
(2000–7300)
28-29
*Soil, Duk-Pyung: As 30 mg/kg (9–61), Ba
996 (93–6380), Cd 0.7 (0.2–7.2), Cu 99
(36–403), Mo 34 (1–134), Pb 57 (12–370), U
78 (5–780), Zn 186 (60–841), S 2200
(900–4800)
28-29
Groundwater: pH 6.6 (4.9–7.9), Cr
(0.10–1.27 μg/L), Cu (0.36–18.9), Fe
(13.1–698), Mn (0.58–1160), Pb (0.12–2.52), Zn
(1.82–1430), V (BDL–2.25), U (BDL–21.3), SO
4
(1400–160,800)
30
*Groundwater: pH 7.2 (6.0–8.2), Cr (ND–5 μg/
L) Cu (ND–312 μg/L), Fe (ND–130), Mn
(ND–84), Pb (ND–17), Zn (ND–996), V (ND–2),
SO
4
(3600–130,900)
31
*Surface water: pH 7.4 (4.1–9.9), Cu (Not
Detected–18 μg/L), Fe (ND–67), Mn (ND–282),
Pb (ND–12), Zn (ND–138), V (ND–3)
31
No reported data available No reported data available
(continued on next page)
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
72
weathering, bioweathering plays an important role in the degradation
of black shale exposed to oxic conditions. Microorganisms may utilise
the organic matter during black shale weathering as a carbon source
controlling the oxidation of sedimentary organic matter (Petsch et al.,
2005). Matlakowska et al. (2012) and Włodarczyk et al. (2016) studied
the bioweathering of the Kupferschiefer black shales in Poland and
found that the oxidation of the fossil organic matter by indigenous li-
thobiotic, heterotrophic and neutrophilic bacteria accounts for the re-
lease of various organic compounds, which in turn can promote the
weathering of sulphides. Subsequently, trace elements and S associated
with sulphides are mobilised, affecting the leachate waters and pre-
cipitation of secondary minerals (Matlakowska et al., 2012;Włodarczyk
et al., 2016). However, other studies show that chemical weathering
processes have higher influence on the oxidation of the rock than the
organic matter degradation by microbial activity (Chi Fru et al., 2016).
Laboratory experiments prove that an indigenous consortium of bac-
teria is involved in the oxidative bioweathering of fossil organic matter
(bitumens and kerogen) and may promote the mobilisation of fossil
organic carbon from the rock (Seifert et al., 2011;Stasiuk et al., 2017).
The mobilisation of organic carbon is documented in the form of oxi-
dised organic compounds, such as monohydroxy and dihydroxy alco-
hols, aldehydes, monocarboxylic and dicarboxylic acids and esters due
to microbial activity (Stasiuk et al., 2017).
The potential for acid rock drainage and soil pollution should be
evaluated case-specifically due to the heterogeneity of black shale oc-
currences (Tables 1 and 2). Table 2 summarises the environmental
impacts of the selected black shale deposits with reported metal con-
centrations in the mining waste, in by-products or in environmental
recipients like in soil, river sediments and water bodies. Fig. 1 high-
lights that the environmental impacts of black shale may vary due to
the volume of sulphide-rich units. Considering the vast extent of many
of the black shale deposits, extending over hundreds of kilometres, they
have the potential to produce acidity in large areas if they outcrop at
the surface. Besides the thickness of the metal-rich sequence, the che-
mical characteristics of under- and overlying rocks and their hydraulic
properties contribute to the issue. Typically, black shales occur in cyclic
Table 2 (continued)
Location and name of the
deposit
Mining activity
Ore enrichment processes
Mean concentrations in mine waste or by-
products
(Min–Max)
Mean concentrations in anthropogenic
(*natural) soil/sediment pollution
(Min–Max)
Mean concentrations in (*natural) surface
water bodies and in groundwater
(Min–Max)
USA
Appalachian basin black
shales
No reported mining
activity on black shale
deposits
Stream sediments, road cut: median As
56 mg/kg (38–98), Co 36 (11–105), Cr 75
(27–92), Cu 30 (12–75), Mn 970 (135–1750),
Mo 78 (3.0–215), Ni 61 (16–340), Pb 42
(25–68), V 335 (87–490), Zn 210 (81–800)
32
Soil, road cut: median As 53 mg/kg (43–57),
Co 5.7 (5.1–5.9), Cr 125 (105–140), Cu 56
(43–63), Mn 52 (44–54), Mo 96 (86–110), Ni
15 (12–18), Pb 23 (21–28), Sb 6.6 (5.6–7.4),
U 23 (20–24), V 465 (395–525), Zn 38
(27–39)
32
BDL: below detection limit; ND: not detected.
1
Tuovinen et al., 2018.
2
Leppänen et al., 2017.
3
Terrafame report, 2017; Terrafame report, 2018.
4
Publically available data by Terrafame Ltd.
5
Pašava et al., 2003.
6
Peng et al., 2009b.
7
Lacelle et al., 2007.
8
Kwong et al., 1997.
9
Sun, 1999.
10
Sun et al., 2000.
11
Helios Rybicka et al., 1994.
12
Daus and Weiß, 2001.
13
Schubert et al., 2003.
14
Wennrich et al., 2004.
15
Carlsson and Büchel, 2005.
16
Bozau et al., 2017.
17
Bobarowski and Bozau, 2006.
18
Sracek et al., 2010.
19
Ettler et al., 2011.
20
Ntengwe et al., 2006.
21
Allard et al., 1991.
22
Falk et al., 2006.
23
Bäckström and Sartz, 2015.
24
Lavergren et al., 2009a.
25
Åström et al., 2009.
26
Allard et al., 2011.
27
Kim and Thornton, 1993a.
28
Lee et al., 1998.
29
Park et al., 2010.
30
Woo et al., 2002.
31
Chon and Oh, 2000.
32
Tuttle et al., 2009.
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
73
alternation with phyllites, sandstones and limestones / dolomites,
making up between 10% and 20% of the succession, rarely over 50%.
Hence, the surrounding rock units reflect the quantity of the natural
neutralisation capacity.
5. Nickel deposits and occurrences
5.1. Talvivaara deposit, Finland
The Talvivaara black shales deposited during 2.1–1.9 Ga and are
characterised by a large amount of organic matter and sulphur
(Table 1). In general, the Palaeoproterozoic of northern and eastern
parts of Finland and the related formations in Karelia and the Kola
peninsula in NW Russia contain the highest total amounts of accumu-
lation and preservation of organic matter in sedimentary rocks reported
so far (Loukola-Ruskeeniemi, 1991;Loukola-Ruskeeniemi, 1999;
Melezhik et al., 1999). The Talvivaara deposit consists of two poly-
metallic ore bodies: Kuusilampi and Kolmisoppi (Ervamaa and Heino,
1983;Loukola-Ruskeeniemi, 1995;Loukola-Ruskeeniemi and Heino,
1996;Loukola-Ruskeeniemi and Lahtinen, 2013;Kontinen, 2012;
Kontinen and Hanski, 2015;Makkonen et al., 2017). The medium-grade
regional metamorphism caused recrystallization of the sulphide phases
and reduction of organic matter to graphite.
The tectonic deformation and isoclinal folding shaped and multi-
plied the black shale layers so that the two ore bodies are now up to
330 m thick even though the original thickness may have been < 50 m
(Figs. 1 and 2). The whole black shale formation reaches a maximum
thickness of over 800 m, including ore-grade (> 0.07% Ni) and barren
black shales (< 0.07% Ni). The Talvivaara mine that exploits the
Kuusilampi orebody has been operating since 2008, and the total esti-
mated low-grade mineral resources are over 2000 Mt. (0.22% Ni, 0.13%
Cu, 0.50% Zn and 0.02% Co, Talvivaara Mining Company, 2012).
However, the Kolmisoppi ore body is in part located under Lake Kol-
misoppi which may limit the possibility to exploit it.
The black shale containing abundant graphite and sulphides is
principally composed of quartz and phlogopitic biotite, with rutile,
apatite, zircon, feldspar, tremolite, diopside, sphene, muscovite and
spessartine garnet as common accessory minerals (Loukola-
Ruskeeniemi and Lahtinen, 2013;Tuovinen et al., 2016). Pyrite [FeS
2
]
and pyrrhotite [Fe
(1-x)
S] are the dominant sulphide minerals, whereas
chalcopyrite [FeCuS
2
], sphalerite [ZnS], galena [PbS], alabandite
[MnS], molybdenite [MoS
2
], pentlandite [(Fe,Ni)
9
S
8
], ullmanite
[NiSbS] and stannite [Cu
2
FeSnS
4
], together with uraninite [UO
2
] and
thucholite [a mixture of hydrocarbons, uraninite and sulphides], occur
in minor amounts (Table 1;Loukola-Ruskeeniemi and Heino, 1996;
Loukola-Ruskeeniemi and Lahtinen, 2013).
5.1.1. Pre-mining conditions and the impact of current mining at Talvivaara
The Talvivaara-type mineralised black shales commonly lie in to-
pographic depressions in Finland due to their susceptibility to erosion
during the last glaciations. They are usually covered by glacial till, peat
and lakes, but exposed outcrops also occur in the Talvivaara area be-
cause the formation was originally that thick. Substantial amounts of
Ni-rich and Ca-poor black shale material were abraded and transported
Table 3
Environmental impact of black shales in Talvivaara under pre-mining conditions. Median, min. and max. values of selected trace elements in glacial till, stream
sediment, lake sediment, surface water, peat and dug well water samples in the black shale area are given.
Sample material Cu Fe Mn Ni Pb Zn Cd Al As Co U
Till mg/kg (N= 586)
1
Min 8.0 6.0 14
Max 2900 3100 17,900
Median 70 56 150
Well-decomposed organic stream sediments mg/kg (N= 88)
2
Min 4.0 9390 73 7.0 7.0 17
Max 660 398,000 4600 2600 195 9800
Median 57 51,300 375 93 38 350
Poorly decomposed organic sediments “peat” mg/kg (N= 100)
2
Min 5.0 6100 91 15 11 68
Max 7200 364,000 22,000 4550 564 10,100
Median 63 83,750 615 106 81 348
Surface lake sediments mg/kg (N= 56)
3
Min 23 8280 93 48 2.61 180 0.39 5720 0.57 6.5 0.54
Max 217 292,000 79,500 1080 845 2500 16 25,500 50 113 4.7
Median 61 45,950 666 173 24 907 4.8 14,450 3.8 29 1.5
Stream water μg/L (N= 51)
2
Min 1.0 0.26 0.02 2.0 1.0 0.02
Max 130 7.9 0.77 920 5.3 3.6
Median 1.2 1.9 0.07 7.8 < 1.00 0.03
Stream water μg/L (N= 26)
3
Min 0.4 170 14 0.5 0.07 1.1 < 0.02 40 0.13
Max 12 4650 973 1.6 1.6 530 0.54 1370 1.1
Median 1.6 1260 82 0.17 0.17 49 0.18 260 0.29
Lake water μg/L (N= 9) Min 0.83 320 5.6 1.69 0.05 0.01 BDL BDL BDL
Max 4.51 910 156 65.5 2.3 177 0.46 3.5 0.09
Median 2.61 410 25 14.8 0.27 94 0.24 1.6 0.06
Peat mg/kg (N= 25)
1
Min 29 2620 6.80 18 0.43 36 1.4 0.61
Max 316 109,000 77 612 24 4360 36 11
Median 180 53,300 33 244 1.3 1490 20 3.7
Dug well μg/L (N= 4)
3
Min 2.00 <30 1.00 10 < 0.03 26 0.22 7 < 0.05
Max 80 290 143 91 2.1 628 3.68 416 0.09
Median 17 75 84 51 0.93 385 2.2 289 0.05
⁎Dug well μg/L (N = 9)
4
Min
Max 220 140 91 82 1.5 130 0.04 990 0.31 7.4 0.23
Median 34 84 8.0 1.9 0.49 24 0.68 40 0.09 0.25 0.05
⁎Well drilled into bedrock μg/L (N= 39)
4
Min
Max 360 17,000 1400 26 8.7 500 0.19 130 0.61 0.73 4.5
Median 31 150 8.0 0.80 0.45 16 0.01 <20 0.07 0.07 0.03
1
Parviainen et al., 2014 (originally from the database of the Geological Survey of Finland).
2
Gustavsson et al., 2012.
3
Loukola-Ruskeeniemi et al., 1998.
4
Karppinen et al., 2012.
⁎
Samples collected in 2011–2012. The mining activities commenced in 2008.
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
74
during the glaciations. This is evidenced by geochemical maps of glacial
till, which show the distribution of ground black shale material far from
the original bedrock source (Loukola-Ruskeeniemi et al., 2003). Simi-
larly, the soil geochemistry is well-correlated with the underlying
bedrock concentrations (Table 3;Gustavsson et al., 2012), and many
studies show the influence of mineralised black shale on the overlying
till and peat cover (e.g.,Loukola-Ruskeeniemi et al., 1998;Mäkinen
et al., 2010;Parviainen et al., 2014). The till deposited over black shale
Fig. 2. Geological map of the Talvivaara area, generalised and modified from the digital geological map database of Finland (DigiKP, Geological Survey of Finland,
2017). At present, open pit mining continues in the Kuusilampi ore body. Cross-section 11400 showing the depth of the deposit is modified from Makkonen et al.
(2017).
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
75
bedrock contains approx. 2-, 3- and 5-fold higher values of Cu, Ni and
Zn, respectively, than the till in the surrounding granite gneiss, quart-
zite and mica schist areas (e.g.,Parviainen et al., 2014). The peat de-
posits in the study site are affected by both the underlying mineralised
rocks and by the acid rock drainage from black shale outcrops. Elevated
pre-mining concentrations of, for instance, Fe, Ni, Zn and S were found
towards the bottom layers under acidic conditions (Mäkilä et al., 2012;
Parviainen et al., 2014).
Weathering of black shale and associated soils have an impact also
on the local water geochemistry. Fig. 3 displays Ni and Cu concentra-
tions in surface waters prior to the mining activities.
Consequently, especially Ni and Zn concentrations in stream and
lake sediments in the mineralised areas reflect the chemical composi-
tion of the bedrock (Table 3): the median concentrations of Cu, Fe, Ni
Fig. 3. Nickel and copper concentrations in surface waters in the Talvivaara area before mining activities. However, it is already obvious that three pilot quarries
have impacted the water quality in streams, lakes and rivers (Gustavsson et al., 2012). The distribution of metamorphosed black shale is modified from the digital
map database of the Geological Survey of Finland (DigiKP, 2017).
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
76
and Zn in these stream sediment samples exhibit 2.7- to 3.5-fold higher
values than in the nearby non-mineralised areas (Gustavsson et al.,
2012). Lake sediment profiles of the adjacent lakes record past events,
showing that acid rock drainage has been active already during the
deglaciation 9000 years ago. The intense metal accumulation took place
when the ice lake in front of the retreating ice sheet found a new
channel to the sea and the water level suddenly dropped, exposing
black shale bedrock to oxic conditions (Loukola-Ruskeeniemi et al.,
1998). Lake sediments were affected, which can be seen at present at
roughly 5.8 m depth, and show 7- to 11-fold higher concentrations of
Cu, Ni and Zn than the recent lake sediments representing pre-mining
conditions. Additionally, U, known to be very mobile in oxic conditions
show 29-fold higher concentrations in comparison to the recent sedi-
ments.
Before mining activities began in the Talvivaara area, the drainage
of peatlands was the main anthropogenic activity that contributed to
the mobilisation of metals from the soils. Shallow lake sediments ex-
hibit a moderate increase in metal concentrations starting from the
early 20th century and a more pronounced increase in the concentra-
tions of Ca, Cr, Cu, K, Mg, Na and Ni in the layers corresponding to the
1970s and 1980s, when the drainage of peatlands was a common
practice in forestry in Finland (Mäkinen et al., 2010). For instance, Ni
exhibited twice as high concentrations in the sediments deposited in the
1980s in comparison to the ones from early 20th century. Furthermore,
Fig. 4. The Talvivaara Ni–Zn–Co–Cu deposit has been mined since 2008 which has modified the landscape. The open pit and primary and secondary bioleaching
heaps are indicated on the map. The distribution of metamorphosed black shale (black schist) is modified from the digital geological map database of Finland
(DigiKP, Geological Survey of Finland, 2017).
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
77
Fig. 5. Distribution of mineralised black shales in four selected regions around the world: the Cambrian alum shales in Sweden (modified from Andersson et al., 1985;
Lecomte et al., 2014), the Cambrian Niutitang black shales in China (modified from Han et al., 2017), the Devonian black shales in Yukon in Canada (modified from
Goodfellow et al., 2010) and the Permian Kupferschiefer in Poland and Germany (modified from Borg et al., 2012).
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
78
northern pike in lakes with large catchment areas and bedrock rich in
black shales may show elevated Hg concentrations (Loukola-
Ruskeeniemi et al., 2003).
The Talvivaara mine project commenced in 2004 and the full-scale
exploitation began in 2008. To recover Ni and other metals of economic
value, the enrichment process takes advantage of bio-leaching to pro-
mote sulphide dissolution (Riekkola-Vanhanen, 2010). The recovery
process consists of heap bioleaching of ground black shale rock cata-
lysed by bacteria like Acidithiobacillus ferrooxidans. The mining opera-
tions produce three types of solid wastes, including waste rock, leached
Fig. 6. Distribution of metamorphosed black shales and graphite-sulphide schists in the Palaeoproterozoic of Finland (Arkimaa et al., 1999, 2000;Loukola-
Ruskeeniemi et al., 2011;Airo et al., 2009;Hyvönen et al., 2013). The map and database were compiled by correlating the aeromagnetic and aeroelectromagnetic
data of the Geological Survey of Finland. The inferred black shale units were verified with geochemical and petrophysical studies of drill core samples. (Colour
version of the figure is available in the web version of this article.)
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
79
ore and production waste consisting of gypsum, calcium carbonate and
Fe hydroxides. After the two-stage (primary and secondary) leaching,
the mine wastes are permanently stored on the secondary leaching site.
The mining activities are leaving a significant footprint on the land-
scape (Fig. 4).
Since the mining activities began, the adjacent water bodies have
been regularly monitored by the mining company under the supervision
of the local environmental authority, the Centre for Economic
Development, Transport and the Environment (the Kainuu ELY Centre).
The water quality in the adjacent rivers and lakes has experienced
changes, and peaks in sulphate and metal concentrations have been
observed. In 2012, a major leakage from the gypsum waste pool was
reported, and the consequences of the spill can be seen in the mon-
itoring data in the surrounding streams and lakes. Here we do not
present the monitoring data directly after the spill, but data from 2015
to 2017 (Table 2). The water quality has recovered lately, although the
monitoring data still reveal elevated concentrations of SO
4
, Fe, Mn and
Na in some of the nearby lakes and streams, as well as elevated Ni and
Zn in some sites. The impact of mining activities has also been recorded
in the lake sediments of Lake Kivijärvi, about 1 km southwest from the
Talvivaara mine. The S, Fe, Mn, Ni and Zn concentrations have in-
creased since the exploitation started (Table 2;Leppänen et al., 2017).
The lake sediments deposited at the time of the major spill from the
gypsum waste pool record for instance 10- and 5.6 -fold higher con-
centrations of Ni and Zn in comparison to pre-mining values. The
concentrations have decreased in the most recent sediments suggesting
that the leakage was successfully blocked.
The Regional Council of Kainuu ordered a study on the water
quality in privately owned dug wells and wells drilled into bedrock
(Karppinen et al., 2012). The objective was to evaluate the natural
impact of metal-enriched black shale on local groundwater. The water
from the studied wells located away from the mine generally exhibited
good quality, although the mean of Mn and Fe exceeded the threshold
values for private household water: 400 μg/L and 100 μg/L, respec-
tively (Table 2;Karppinen et al., 2012). Nickel concentration exceeded
the threshold value (20 μg/L) in only a few of the studied wells
(Table 2;Karppinen et al., 2012). However, the groundwater in-
vestigations in the vicinity of the mine performed by Terrafame Ltd.
demonstrate deterioration in the quality of groundwater (Terrafame
report, 2017;Terrafame report, 2018). Low pH values (3.5–6.1) and
elevated metal concentrations were recorded in the groundwater
(Table 2;Terrafame report, 2017), especially in the shallow ground-
water. Additionally, the water samples from two shallow groundwater
monitoring wells installed in the primary leaching stacks exhibited low
pH values (3.4–4.3) and exceptionally high sulphate and metal con-
centrations (up to 79,000 mg/L SO
4−2
, 0.100 mg/L As, 1.3 mg/L Cd,
10 mg/L Cu, 1000 mg/L Mn, 310mg/L Ni, 660 mg/L Zn, 1700 mg/L Fe,
2.6 mg/L U) (Terrafame report 2017;Terrafame report 2018).
The air quality of the Kainuu Region was evaluated in 2015 by
analysing epiphytic lichen (Hypogymnia physodes). Concentrations of Cr,
Cu, Fe and Ni were higher than average around the Talvivaara mine due
to the emissions from the mine site (Laatikainen and Seppänen, 2017).
5.2. Niutitang black shales, China
The Lower Cambrian black shales hosting the Mo–Ni–PGE ores of
the Niutitang Formation in southwestern China form an over 1600-km-
long belt that outcrops discontinuously (Fig. 5). The mineralised layer
corresponds to a thin accumulation (5–20 cm) of Ni, Mo, Au, Ag, Se, Cr,
V, Zn, U, rare earth elements (REE) and platinum group elements (PGE)
that can be traced along the same stratigraphic horizon (Table 1). Here
Niutitang black shales are classified as Ni deposits as they occur in the
Zunyi area, though U deposits are reported in the Hunan, Jiangxi and
Guangxi Zhuang regions. The exceptional enrichment of metals from
seawater is ascribed to high biological productivity in the surface oxic
layer above sediment-starved euxinic basins (Mao et al., 2002;
Lehmann et al., 2007), and organic matter played an important role in
mineralisation though it was not the sole control (Shi et al., 2014). The
origin of the highly metal-rich layer is also related to hydrothermal
processes, whereas a seawater and terrigenous origin is suggested for
the thicker black shale sequence (Fan, 1983;Lott et al., 1999;Shi et al.,
2014; Han et al., 2015). The chondrite-normalised patterns of PGEs
indicate that PGE enrichment of the polymetallic ores is most likely
originated from mafic rocks (Han et al., 2015). The black shales de-
posited intercalated with phosphorite, barite and sapropelic beds (Mao
et al., 2002). The Cambrian facies shifts from a shallow shelf environ-
ment with phosphatic carbonate rocks in the northwest to deep basinal
facies of black shale in the southeast (Zhu et al., 2003). The Niutitang
Formation unconformably overlies the Ediacaran Dengying Formation
(551–541 Ma) and is mainly composed of organic-rich black shales
containing phosphatic beds, varying between 0.3 and 1.35 m, and tuff
layers. The polymetallic Ni–Mo–PGE–Au sulphide ore is locally present
in the lowermost part of the formation.
5.2.1. Natural and post-mining conditions around the Niutitang black
shales
Peng et al. (2004) reported V, Cr, Co, Cu, Zn, Pb, Cd and Tl en-
richment in soils and Sc, Cr, Fe, Mn, Co, Ni, Cu, Zn, Pb and Cd en-
richment in surface waters near weathering profiles of Lower Cambrian
black shales in Hunan, China, which are associated with a locally high
incidence of endemic diseases, such as cancer. Additionally, notable
enrichment of Mo, Cd, Sb, Sn, U, V, Cu and Ba was observed in soils
associated with another weathering profile of black shale in Hunan (Yu
et al., 2012). The mobilisation of Mn, Sr, Ba, Pb, U, V, Cr, Co, Ni, Cu, Zn
and REE was observed in natural weathering profiles in northeast
Chongqing, China, and their redistribution was dependent on the pH,
weathering intensity and the secondary minerals formed (Ling et al.,
2015). The comparison of an outcropping weathering profile of Ordo-
vician-Silurian black shales in Guizhou Province, China, with the re-
spective unweathered shallow borehole samples reveals mineralogical
changes and the depletion of organic matter, V, Cr, Th, U, Ni and Co in
the weathered samples (Tang et al., 2018).
The most significant U deposits are reportedly in Hunan, Jiangxi
and Guangxi Zhuang regions, whereas Ni-Mo deposits occur in the
Zunyi area. There is limited information available about the black shale
mines in China. However, couple of publications describe soil and water
pollution associated with black shale mining. According to Peng et al.
(2009b), soils in the vicinity of the black shale U mine in the Hunan
region are enriched in Mo, Cd, Sb and U. Besides the pollution of soils
and local water sources, bioaccumulation has occurred also to crop. The
soil pollution and subsequently observed bioaccumulation of poten-
tially harmful metals into maize are related to the chemical con-
centration of the bedrock. Acid mine drainage from another U mine in
Hunan is affected by Cd, Co, Ni, Zn and U (Table 2;Peng et al., 2009a).
Additionally, in the surroundings of the Ni-Mo mines in the Zunyi area
agricultural soils (Table 2) are polluted by drainage waters derived
from the mine waste dumps and consequently bioaccumulation of po-
tentially harmful metals occurs in local tobacco, rice, corn and turnip
(Pašava et al., 2003).
5.3. Yukon black shales, Canada
The Devonian strata in the Yukon Territories includes an extensive
shale sequence that hosts Ni–Zn–PGE occurrences (Fig. 5). The black
shale sequence was deposited in the Selwyn epicratonic marine basin,
and overlies Cambrian and Ordovician carbonate rocks. The sequence
commences with Road River Group calcareous black shales and is fol-
lowed by the Canol Formation in north Yukon and the Earn Group si-
liceous black shales in the central parts of Yukon (Fraser, 2014;Fraser
and Hutchison, 2017). The Selwyn basin black shales have undergone
lower greenschist facies metamorphism.
Here, we focus on the sulphide mineralisation enriched in
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
80
Ni–Zn–PGE and other metals, which occurs in the interface of the Road
River Group and Canol Formation or Earn Group in between the layers
of the limestone ball member (black shales containing limestone
spheroids) and phosphatic chert (Butterworth and Caulfield, 1997;
Goodfellow et al., 2010;Hulbert et al., 1992). The stratigraphic and
sedimentological sequence is similar throughout the black shale se-
quence in the Nick Property, Eagle Plain and Peel River occurrences
(Goodfellow et al., 2010). The mineralisation, principally consisting of
vaesite [NiS
2
], pyrite, marcasite [FeS
2
], sphalerite and wurtzite
[(Zn,Fe)S], varies in thickness from a few centimetres to 40 cm (Fig. 1),
but it covers a large area where different occurrences are located up to
400 km apart. The massive sulphides were formed as rhythmically la-
minated layers of sulphide grains and the sulphides are of both hy-
drothermal and biogenic origin (Hulbert et al., 1992;Orberger et al.,
2003). They contain high concentrations of Ni, Zn, PGE and Re and
anomalous levels of U, Mo, Ba, Se, As, V and P (Table 1;Butterworth
and Caulfield, 1997;Goodfellow et al., 2010;Hulbert et al., 1992;
Orberger et al., 2003). According to Orberger et al. (2003), the trace
element concentrations vary among different sulphide phases.
The thicker sequences of black shales in the Road River Group and
Canol Formation are not enriched in trace metals. However, in the
Canol Formation (including the Lower Earn Group), thin pyritic layers
or disseminated pyrite occur within the beds, generally with low con-
centrations of trace metals (Hulbert et al., 1992;Fraser and Hutchison,
2017). Furthermore, there are some exceptions, such as the sedimen-
tary exhalative massive sulphide (SEDEX) deposits, for instance the
Zn–Pb–Ag occurrence of Macmillan Pass hosted within the Lower Earn
Group rocks and Howards Pass hosted within the Road River Group
(Goodfellow, 2004).
Two models have been developed for the genesis of the sulphide-
rich layer:
1) Hydrothermal fluids originated from the underlying black shales,
discharged in the ocean floor and expanded laterally, interacting
with basinal brines. In the nutrient-rich ambient environment, the
deposition of the sulphide-rich layer was promoted by bacterial
activity. The mineralised layer exhibits a similar geochemical sig-
nature to the bitumen veins in the Road River Group, which re-
present the conduits for the hydrothermal fluids (Hulbert et al.,
1992).
2) Meteorite debris deposited on the sea floor is the source of the mi-
neralisation (Goodfellow et al., 2010).
5.3.1. Anthropogenic actions in the Yukon black shale occurrences
The Devonian Nick property mineralisation in Yukon Territory was
initially discovered as anomalous concentrations of Ni, Zn, PGE, U, Ba
and P in the adjacent stream sediments (Hulbert et al., 1992). Other
outcropping black shale and SEDEX deposits hosted in black shales in
Yukon Territory have produced acid rock drainage characterised by
high concentrations of trace metals in surface waters (Kwong et al.,
2009). The black shales in Yukon have not been exploited to date. The
Ni-rich sulphide mineralisation at the base of the Canol and equivalent
formations form a thin layer, though it has a vast extension and high Ni,
Mo and PGE concentrations (Table 1). The black shale-hosted SEDEX
massive sulphide deposits have been economically more attractive. For
example, MacMillan Pass has been a target of exploitation of Zn, Pb and
Ag.
Excavations, road cuts and other anthropogenic activities have
caused acid rock drainage in Yukon as sulphide-rich black shales have
exposed and consequently, the oxidation of sulphide minerals has oc-
curred. For instance, along the Dempster Highway near Eagle Plains
Lodge, acid rock drainage has caused damage to local vegetation.
Removal of the surficial sediments for a storage area during the con-
struction of the highway in the early 1970s exposed the underlying
sedimentary rocks and disturbed the permafrost leading to the circu-
lation of oxygenated water. The oxidation of pyrite-rich shales is
suggested to be promoted by Fe-oxidising bacteria, A. ferrooxidans, that
converts immobile elements into their mobile state, provoking locally
acid rock drainage with a pH of < 3.8 (Table 2;Lacelle et al., 2007;
Lacelle and Leveillé, 2010). In periglacial regions, the seasonal free-
ze–thaw cycles in the near-surface environment cause the temporary
storage of acidity, solutes and trace metals, whereas their abrupt release
can be observed during the spring thaw (Lacelle et al., 2007). Hence,
these studies emphasise that acid rock drainage may have a severe
impact in periglacial areas.
5.4. Enterprise deposit, Zambia
The Proterozoic Zambian Copperbelt is part of the Central African
Copperbelt, which hosts stratiform sedimentary Cu mineralisations. The
Ni-rich Enterprise deposit within the Zambian Copperbelt is estimated
to contain 40 Mt. of ore with 1.07% Ni on average. It is thought to
originally have been hosted by organic-rich, weakly argillaceous car-
bonate rocks that were silicified and then replaced by kyanite during a
complex metasomatic event (Capistrant et al., 2015). The highest Ni
concentrations are encountered in organic-poor and quartz-rich rocks in
close association with the organic-rich carbonaceous black shales. This
may imply that the organic-rich sedimentary rocks acted as a reductant.
As a result, Ni and Fe-Ni sulphides, including bravoite [(Fe,Ni)S
2
],
vaesite, millerite [NiS], pyrite, chalcopyrite, molybdenite, pyrrhotite
and carrollite [Cu(Co,Ni)
2
S
4
], precipitated in veins and as semimassive
replacements of host rocks. In places, the mineralisation exhibits high
concentrations of Ni and Cu (Table 1;Capistrant et al., 2015).
5.4.1. Acid mine drainage in the future
The black shale-associated Ni Enterprise deposit is currently under
development in order to open the mine in the future. Due to the current
development stage of the mine, the environmental impacts cannot yet
be estimated. However, it may be stated that the deposit has potential
to produce acid mine drainage characterised by Co, Cu, Mo and Ni
based on its ore mineral assemblage. The presence of carbonate mi-
nerals in the adjacent rock units (Capistrant et al., 2015) may provide
acid neutralisation potential.
6. Copper deposits
6.1. The Kupferschiefer, Poland and Germany
The ore deposits of the European Kupferschiefer are hosted by se-
diments deposited on a Permian sedimentary basin (Fig. 5). Lower
Zechstein beds cover the surface of about 1,000,000 km
2
. The Kup-
ferschiefer is organic-rich black shale, up to 1.2 m thick with an average
thickness of 0.3 m (Fig. 1). The shale exhibits a sharp contact with
underlying sandstone. In the boundary there is a discontinuous dolo-
mite layer. The contact between the Kupferschiefer and the Zechstein
limestone above is gradational (Pieczonka et al., 2015). The generally
stratiform Kupferschiefer mineralisation occurs in the footwall sand-
stones, in the Kupferschiefer black shale, and in the lowermost part of
the hanging-wall limestone (Kucha, 1990). The total thickness of the
mineralisation, including the footwall and hanging-wall ore, can range
from several decimetres to locally up to 50 m, and the vertical thickness
of the mineralisation gradually increases from west to east, being
thickest in Poland. The Kupferschiefer black shale has three miner-
alogical varieties: organic-rich pitchy shale, clayey shale and carbonate-
rich shale. The Kupferschiefer is composed of organic matter, illite,
glauconite, carbonate, sulphides, sulphates, quartz and feldspar (Kucha,
1990). The cut-off grade for Cu ores is 0.7% Cu. This mineralisation is
composed of a variety of sulphide minerals (Table 1).
There are many theories of the genesis of the Kupferschiefer de-
posits which Kulick et al. (1984),Vaughan et al. (1989),Oszczepalski
(1999) and Borg et al. (2012) have summarized. The original theory
was a syngenetic one. Early diagenetic formation as well as multistage
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
81
ore formation have also been suggested. Another view is that the sig-
nificant enrichment of Cu, Zn and Pb is related to an epigenetic process.
The brines were mobilised due to tectonic compression and magmatic
activity and redox reactions in the vicinity of the black shale assisted in
precipitating Cu, Zn and Pb from the brines (e.g.,Borg et al., 2012).
6.1.1. Environmental impacts of the Kupferschiefer deposits during and
after the mining activities
Cu ores have been mined from the Central European Kupferschiefer
for many centuries, but only three mines are currently active. The ac-
tive mines are located in Lubin, Polkowice-Sieroszowice and Rudna in
SW Poland. Stream water, stream sediments and soil are reported to
suffer Cu–Ni–Pb–Zn-rich pollution due to tailings areas and smelter
activity in the Lubin district (Table 2;Helios Rybicka et al., 1994;Sun,
1999;Sun et al., 2000), although no recent data are publically avail-
able. Bacteria-mediated weathering of organic matter and sulphide
minerals occurs in the underground workings of the Lubin mine where
microorganisms have affected the chemical composition of the Kup-
ferschiefer black shale (Matlakowska et al., 2012;Włodarczyk et al.,
2016). Physicochemical properties are changed through the bio-oxi-
dation of sulphide minerals and fossil organic matter, including
kerogen, and consequently, also the chemical characteristics of
groundwater changed (Bakowska et al., 2017). As a result of the de-
gradation of organic matter, a number of oxidised organic and in-
organic compounds are formed, including alcohols, organic acids, ke-
tones and aldehydes. Secondary inorganic compounds are formed as
well like biominerals such as sulphates. A study on a weathering profile
in a black shale quarry, in Poland, exhibits metal mobilisation in the
superficial layer of the profile due to the alteration of organic matter
and sulphides under oxic surface conditions (Marynowski et al., 2017).
Organic carbon, Cu, Mo, Pb, S, and U are depleted in the weathered
layers, whereas Cu, Pb, U and V are enriched in the interface between
the weathered and partially weathered layer.
In Germany, the Kupferschiefer deposits have had pronounced en-
vironmental impacts. In East-Thuringia decades of U mining and in
Saxony-Anhalt centuries of Cu mining have produced wastes. The
weathering of mine wastes have caused soil, surface water and
groundwater pollution (Table 2). In East-Thuringia, metal and U con-
centrations are elevated in soils. Secondary Fe and Mn oxides retain
these elements and hard pan cover is formed. The groundwater is
contaminated by F, Cl
−
, SO
42−
, Al, As, B, Cd, Cu, Fe, Mn, Ni, U and Zn
(Table 2;Carlsson and Büchel, 2005;Grawunder et al., 2009;Schäffner
et al., 2015). Similarly, in the Mansfeld district in Saxony-Anhalt, waste
rock, slag from Cu smelting and scrubber dust, known as the Theisen
sludge, were deposited on site without isolation from the soil or the
atmosphere (Daus and Weiß, 2001;Schubert et al., 2003, 2005;Schreck
et al., 2005). The Theisen sludge is considered as an important source of
pollution, as it is enriched in Zn, Pb, S and C (> 10 wt% each) and
additionally it contains elevated concentrations of As, Cd, Cu, Fe, Hg,
Rh, Sb and Ag (Table 2;Daus and Weiß, 2001;Wennrich et al., 2004).
The harmful elements are mobilised from the Theisen sludge through
mechanical erosion, by wind and heavy rains, by chemical weathering
and by leaching. The dispersion of polymetallic pollutants has been
reported along a 20-km watercourse, where elevated As, Cu, Pb, U and
Zn concentrations have been measured in the river sediments (Table 2;
Wennrich et al., 2004). Elevated concentrations of As, Cr, Cu, Ni, Pb, V
and Zn are reported in river sediments in the Mansfeld mining district
(Schreck et al., 2005). Müller et al. (2008) report differences in
weathering behaviour among black shale rock and historical slag heaps
of different ages in Mansfeld. The slag heaps generate a more readily
mobilised load of both radionuclides and chalcophile elements than the
black shale rock (Müller et al., 2008). Furthermore, Baborowski and
Bozau (2006) and Bozau et al. (2017) show that Zn, Cu, Pb, As and U
pollution occurred in the river systems during the mining activities in
the Mansfeld area (Table 2).
6.2. Black shales in the Zambian Copperbelt
The Proterozoic Zambian Copperbelt hosts stratiform Cu miner-
alisations. Two ore bodies occur within arcosic units: a lower ore body
is associated with overlying black shale and an upper ore body is as-
sociated with a shale sequence (McGowan et al., 2006). The miner-
alisation of the lower ore body originated from hydrothermal fluids
entering the arenites and coarser-grained layers of black shale where
the sulphides precipitated. The black shale unit may have facilitated the
lateral flow of the ore-forming hydrothermal fluids and has acted as a
sealing cap promoting sulphide precipitation. The lower ore body
contains abundant pyrite with bornite [Cu
5
FeS
4
], chalcopyrite, chal-
cocite [Cu
2
S], malachite [Cu
2
(CO
3
)(OH)
2
], chrysocolla
[(Cu,Al)₄H₄(OH)₈Si₄O₁₀·nH₂O,] and cuprite [Cu
2
O] (Table 1;McGowan
et al., 2006).
6.2.1. Environmental impacts of mining in the Zambian Copperbelt
Large-scale mining operations in the Central African Copperbelt for
almost a century have caused environmental pollution. Topsoil and
water pollution attributed to Cu smelting and acid mine drainage has
been described in the Kitwe (Nkana deposit) and Chambishi areas in
Zambia, respectively (Table 2;Ettler et al., 2011;Ntengwe and Maseka,
2006).
7. Uranium occurrences
7.1. Alum shales, Sweden
Black shale precursor sediments were deposited in the shallow-
marine environment of the Baltoscandian Platform over an extensive
period of tectonic stability during the Middle and Late Cambrian and
locally during the Early Ordovician (e.g.,Andersson et al., 1985). They
extend from Norway to Estonia, but here we focus on the black shale
deposits located in Sweden (Fig. 5).
The black shales, known as alum shales, occur in the entire lithos-
tratigraphic unit, although the name refers to particular parts of the
black shale formation from which the alum salt, KAl(S0
4
)
2
·12 H
2
0, was
extracted in the 17th century. The alum shale formations are in many
sites about 20 m thick, but they may reach thicknesses of up to 100 m in
southern Sweden. Folded formations occur in northern Sweden, re-
sulting in thicknesses of up to 185 m (Andersson et al., 1985). The
formation is underlain by Early Cambrian sandstones and overlain by
Ordovician limestones and/or grey shales, and the shale and limestone
dominated sequences continue through the Ordovician into the Silurian
(Anderson et al., 1985). The maturity of the alum shales varies due to
the subsequent depositional history. Generally, the alum shales in
central Sweden are less mature than those in the southern parts of the
country. Locally, in central Sweden, magmatic heating by the intrusion
of Permo-Carboniferous dolerites has matured the alum shales
(Buchardt et al., 1997;Sanei et al., 2014;Schovsbo, 2002). Alum shales
that have undergone low-grade metamorphism have experienced mo-
bilisation and enrichment of V, Ni, Zn and Ba, with organic matter and
pyrite acting as reductants. In central Sweden, in the more immature
alum shales, the organic carbon of algal origin is mainly present as
kerogen (99%) and < 1% is oily matter (Sanei et al., 2014). On the
other hand, in northern Sweden, the black shales underwent greenschist
metamorphism during the Caledonian orogeny, which affected their
maturity and organic carbon is present as graphite.
According to Schovsbo (2002), the U was preferentially enriched in
the near-shore environment, reflecting more vigorous bottom water
circulation that promoted higher rates of mass transfer and diffusion of
U across the sediment–water interface in comparison to the environ-
ment further offshore. The U concentrations may vary considerably.
Lecomte et al. (2017) described phosphate nodules containing 100 to
3000 ppm of U. By contrast, U is especially enriched in lenses, called
“kolm”, principally composed of organic matter, where the
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
82
concentrations may reach up to 1000–8000 ppm due to the resuspen-
sion of sediment in an anoxic water column enhancing diffusive ex-
change between suspended particles and the water column (Lecomte
et al., 2017;Schovsbo, 2002). The metamorphic grade also has an
impact on the mineralogy. On the one hand, the black shales of
southern Sweden with lower metamorphic grade do not present in-
dividual uranium minerals, but U rather appears associated with the
organic matter and in phosphate phases, such as in apatite (CaF[OH];
Lecomte et al., 2017). On the other hand, the temperature and pressure
gradient of the greenschist facies metamorphism caused remobilisation
of U and metals. This process led to the precipitation of uraninite and
other uraniferous phases associated with Ca, P, V and Si (Lecomte et al.,
2017). Different forms of pyrite have also been detected. First, fram-
boidal pyrite precipitated, followed by the formation of more massive
pyrite grains. With progressing diagenesis, pyrite crystallised in frac-
tures and, as a consequence of metamorphism, overgrowth occurred on
previously existing pyrite grains together with the precipitation of
chalcopyrite and sphalerite. Concentrations of trace elements in the
alum shale vary even within a deposit (Table 1).
7.1.1. Post-mining condition of the environment near the alum shale
occurrences
The alum shales are a source of acidity to the environment. The
oxidative dissolution of pyrite is principally responsible for acid pro-
duction, contributing to the release of elements and compounds and
forming secondary precipitates that are enriched in metals with respect
to those of the source black shales (Chi Fru et al., 2016). Organic carbon
is not trapped by the acid rock drainage precipitates; hence, it is re-
leased as a consequence of the weathering process (Chi Fru et al.,
2016). Field observations and laboratory weathering experiments have
demonstrated the acid generation and metal release from alum shales
(Falk et al., 2006;Lavergren et al., 2009b, b;Yu et al., 2014;Chi Fru
et al., 2016).
The environmental impacts are the consequence of the deposit type
and the exploitation history. In Degerhamn, SE Sweden, the burning
and processing of alum shales have intensified the mobility of sulphide-
associated elements and U (Falk et al., 2006; Lavergren et al., 2009 a,b).
Conditions on the sea shore promote the weathering of outcrops, waste
rocks and burnt alum shale residue. The Cd, Ni and Zn concentrations in
non-weathered alum shale are 6.9-, 2.3- and 6.8-fold higher, respec-
tively, than the concentrations in weathered or burnt shale, indicating
the liberation of metals during weathering or burning of the shale (Falk
et al., 2006). The sulphide oxidation from the deposits and the sub-
sequent mobility of the elements, controlled by pH, have an adverse
influence on the quality of local groundwater (Table 2). For instance,
Co, Cu, Fe, Mn, Mo, Ni, Sr, U and Zn concentrations are elevated.
Further, Åström et al. (2009) report elevated U concentrations in the
groundwater and stream waters in the black shale areas of Degerhamn
(Table 2). Additionally, leaching tests and humidity cell tests coupled
with sequential extraction procedures on processed black shale mate-
rials from Degerhamn corroborate the leaching behaviour of metals and
U and decreasing pH of the leachates (Falk et al., 2006;Yu et al., 2014).
Ni and U are abundantly leached into solution, whereas As is not as
mobile under acidic conditions, as it is retained by secondary oxy-
hydroxides and sulphates such as schwertmannite and jarosite. This
trend is also reflected in the local groundwater chemistry (Yu et al.,
2014). Additionally, during World War II, black shales were burnt for
oil production in Kvarntorp, central Sweden, leaving 28 Mt. of wastes
deposited in open pits and in a waste facility. The wastes are mainly
composed by ash, pyrite and kerogen, which are still burning. The ig-
nition of the organic matter generates temperatures above 500 °C. The
wastes are reported to drain acidic waters and the local groundwater
contains elevated concentrations of Ni, Mo and U (Table 2). The
leaching is predicted to increase when the deposits cool down
(Bäckström and Sartz, 2015). The water chemistry has changed after
neutralisation from pH 3 to around 7.5 in a pit lake filled with water in
Kvarntorp. Elements such as, Cd, Co, Cu, Ni, Pb and Zn have decreased
approx. > 10 to 60 times lower concentrations, whereas U remains
elevated and Mo has even increased from sub-microgram level to ap-
prox. 9 μg/L (Table 2;Allard et al., 2011). Furthermore, U was
exploited in Ranstad (Billingen area) in south-central Sweden in the
1960s. For remediation, the U mine tailings materials were mixed with
5% limestone and covered by a sequence of bentonite, till and top soil
layers (Allard et al., 1991). These measures have retarded the water and
oxygen influx, but some leachates contain elevated concentrations of
As, Cd, Cu, Mn, Ni, Mo, Zn and Pb (Table 2;Allard et al., 1991).
7.2. Okchon deposit, South Korea
The Okchon uraniferous black shales, also referred to as black slates
in the literature, in the central part of the Korean peninsula are of
Cambrian to Ordovician age and form part of the Guryongsan
Formation of the Okchon Group. The black shales extend over 100 km
and the uraniferous layer is several metres thick (Kim, 1989). These
rocks are considered as analogues to the Cambrian black shales in
China. Uranium exists in phases such as uranothorite [(Th,U)SiO
4
],
uraninite, brannerite [(U
4+
,Ca)(Ti,Fe
3+
)
2
O
6
], ekanite [Ca
2
ThSi
8
O
20
]
and thorutite [(Th,U,Ca)Ti
2
(O,OH)
6
] commonly disseminated in the
quartz and muscovite matrix (Shin et al., 2016). Uraninite and bran-
nerite have also been found to appear in association with pyrite and
monazite in the Okchon black shales (Lee et al., 1998). The formation of
uranium phases is associated with submarine hydrothermal processes.
Afterwards, they suffered regional metamorphism (Shin et al., 2016).
Besides of U, these rocks are enriched in As, Ba, Cu, Mo, V and Zn, and
the average content of total organic carbon, now present as graphite, is
as high as 21.2% (Table 1; Kim et al., 1989, 2015;Lee et al., 1998).
7.2.1. Natural weathering and the impact of mining activities in the Okchon
black shales
Several small-scale mines have operated locally to extract the highly
carbonaceous parts of the Okchon black shale deposit in South Korea.
However, there is very limited information on the environmental im-
pacts of these mines on the surrounding water bodies and soils. Woo
et al. (2002) reported high SO
4
and slightly elevated Mn, U and Zn
concentrations (Table 2) in groundwater near an abandoned U mine.
Research on the environmental impacts has focused on describing the
natural pollution in soils and accumulation in crop plants associated
with Okchon black shales, which outcrop in many places along the 100-
km formation (Table 2). Many investigations report on soil con-
tamination by a wide range of elements including As, Cd, Cu, Mn, Mo,
Ni, Pb, Se, Th, V, U and Zn (Kim and Thornton, 1993a,1993b;Lee et al.,
1998;Yi et al., 2003;Park et al., 2010).
7.3. Black shale occurrences with higher than average U concentrations in
Kentucky, USA
Extensive Devonian black shale sequences occur in the eastern USA
covering the Michigan basin, the Illinois basin and the Appalachian
basin. The Appalachian Basin extends from northern Alabama to
Newfoundland, but here the focus is on the black shales in Kentucky.
The precursor mudrocks include sections rich in organic carbon and
sulphur and some units contain higher than average concentrations of U
(Table 2;e.g.,Roen, 1984;Leventhal and Hosterman, 1982). Pyrite and
marcasite are the principal sulphide mineral phases.
7.3.1. Impact of road construction on the weathering of black shale
A road cut in the New Albany Shales in eastern Kentucky has pro-
moted weathering of the adjacent black shales. Under low pH (< 4),
Cd, Co, Ni and Zn are transported to surface water, though upon pH rise
they are absorbed and precipitated to the stream sediments. Mo, Pb, Sb
and Se seem to be enriched in Fe-oxyhydroxide coatings in the soil layer
covering the black shales, whereas Cu and U are retained by plant litter
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
83
in soil (Table 2;Tuttle et al., 2009). In another road cut in the Sunbury
Shale in north-eastern Kentucky, weathering of black shale has mobi-
lised Mn, Cd, Co, Cu, Mn, Ni, U and Zn (Perkins and Mason, 2015).
8. The country-wide black shale map and database in Finland
Black shales have been a target of extensive investigation in Finland
since the 1950's (e.g.,Peltola, 1960). Finland is the first country in the
world which has completed the nation-wide mapping of black shales
(Arkimaa et al., 1999, 2000;Airo et al., 2009;Loukola-Ruskeeniemi
et al., 2011;Hyvönen et al., 2013). The map was compiled by
correlating aeromagnetic and aeroelectromagnetic data (Fig. 6) since
earlier studies had shown that sulphide deposits and black shale units
can be classified by airborne magnetic and gamma-ray responses (Airo
and Loukola-Ruskeeniemi, 2004;Hautaniemi et al., 2005). The inferred
black shale units were verified with geochemical and petrophysical
studies of the samples selected from the deep drill cores. Statistical
analyses and interpretation of the airborne geophysical, geochemical
and rock petrophysical data were applied to characterise and classify
different types of black shales. The results are compiled in the black
shale database which is designed to be part of the digital geological
map database of Finland (DigiKP, Geological Survey of Finland, 2017).
Fig. 7. Min., mean and max. values of pH and the Cu, Ni, Zn and U concentrations in groundwater (G) and surface waters (S) in selected black shale locations affected
by mining or other anthropogenic activities. (Terrafame report, 2017; Terrafame, public monitoring data; Peng et al., 2009a;Kwong et al., 1997;Bozau et al., 2017;
Bobarowski and Bozau, 2006; Carlsson and Büchel, 2005;Lavergren et al., 2009b;Falk et al., 2006;Woo et al., 2002).
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
84
The black shale database provides background constraints for mineral
exploration, environmental studies and regional planning. The database
has been in active use since its release.
9. Discussion
9.1. Environmental concern
Wherever they occur, black shales are probable sources of acidity.
At the surface environment, chemical weathering of sulphide minerals
and organic matter or graphite is enhanced by bioweathering (Petsch
et al., 2005;Matlakowska et al., 2012;Nordstrom et al., 2015;
Włodarczyk et al., 2016;Chi Fru et al., 2016). Fig. 1 shows that the
environmental impacts of black shales may vary due to the volume of
sulphide-rich layers: the exceptionally metal-enriched parts may be
merely few centimetres thick, as in Yukon, or the low-grade ore may
exceed 300 m as in Talvivaara. However, in Yukon, even sulphide-poor,
but thicker black shale sequences have been shown to be a source of
acid rock drainage which may be especially harmful in periglacial re-
gions (Lacelle et al., 2007;Lacelle and Leveillé, 2010). In addition to
the chemical composition and mineralogy, the volume of the metal-rich
black shale and the chemical composition of the surrounding rocks are
important. Whether the adjacent rock units provide neutralisation ca-
pacity or not makes a difference. Especially, alternation of Ca-rich rocks
and black shale layers in the sequence favours acid neutralisation. The
Talvivaara black shale units are exceptional in thickness and Ca-rich
layers are scarce (Loukola-Ruskeeniemi and Heino, 1996), which
translates into a high potential for acid rock drainage. By contrast, in
the Lubin Cu-Ag mine in Poland, the black shale layer is less than one
metre thick and it occurs below a thick pile of limestones (Borg et al.,
2012).
The varying composition of the mineralisation within a deposit re-
sults in different pollution patterns. Generally speaking, the acid drai-
nage under natural and anthropogenic conditions is characteristically
contaminated by the main metal constituents of each black shale de-
posit, including Fe, Ni, Zn and Cu derived from pyrite and/or pyr-
rhotite, pentlandite/vaesite, sphalerite and chalcopyrite, respectively
(Fig. 7). Additionally, other elements such as As, Cd, Mo and U released
from, for example, arsenian pyrite, sphalerite, molybdenite and thu-
cholite/uraninite may have an adverse impact on the surrounding en-
vironment (Liu et al., 2017;Paikaray, 2012). Even the natural soils
formed on black shale bedrock display elevated concentrations of as-
sociated metals and serve as a source of contamination (Tables 3;
Loukola-Ruskeeniemi et al., 1998;Peng et al., 2004;Mäkinen et al.,
2010;Parviainen et al., 2014;Yu et al., 2012). Translocation to plants
from soil and water may consequently occur as described in China and
Korea (Kim and Thornton, 1993a,1993b;Lee et al., 1998;Fang et al.,
2002;Yi et al., 2003;Park et al., 2010). In Talvivaara, the natural
geochemical signature of the black shales is reflected as elevated con-
centrations of Cu, Mn, Ni, Pb and Zn in till, peat and stream sediments
evidencing multiple dispersion routes into the environment (Table 3).
The data in Table 2 illustrate the high variety of concentrations of
potentially harmful elements in the recipient environment, and also in
some cases show elevated concentrations as a consequence of mining
activities.
The impact of active mining at Talvivaara can be seen as elevated
concentrations of Cu, Ni, Zn and U in groundwater and elevated Ni and
Zn in surface water (Table 2;Fig. 7). In Hunan, U mining impacts the
surface waters where Cd, Co, Ni, Zn and U are recognised as major
pollutants (Peng et al., 2009a). On the one hand, in the Mansfeld dis-
trict in Germany, mining and processing of the Kupferschiefer ores have
produced waste with Cu, Pb and Zn as the main contaminants together
with As, Ni and U (Table 2;Daus and Weiß, 2001;Wennrich et al.,
2004;Schreck et al., 2005). Co, Mn, Ni, Zn and U exhibit high con-
centrations in river water and river sediments (Table 2;Carlsson and
Büchel, 2005;Grawunder et al., 2009;Schäffner et al., 2015).
Additionally, the mining of the Kupferschiefer ores in Poland release
Cu, Mo, Ni, Pb, U and Zn (Sun, 1999;Sun et al., 2000;Marynowski
et al., 2017). In Sweden, in Degerhamn, up to 200 μg/L of U is observed
in groundwater and up to 80 μg/L in surface water due to the burning
and processing of black shale (Åström et al., 2009).
Despite some tendencies in the pollution patterns, it is necessary to
highlight the polymetallic character and the heterogeneity of the black
shale occurrences (Tables 1 and 2). Consequently, the environmental
impacts vary even within a deposit which impedes generalisation of the
impact assessment. Hence, the potential for acid rock drainage and soil
pollution should be evaluated in each case.
Finally, a number of environmental factors affecting the behaviour
of elements in nature should be kept in mind when evaluating the
contamination risks. The geochemistry of the receiving water bodies is
controlled by pH as was observed in the closed mine sites of the alum
shale in Sweden (Falk et al., 2006;Allard et al., 2011). The divalent
cations are mobilised at low pH range whereas As, Mo and U tend to be
more mobile at pH > 7 although the behaviour of these elements is
complex. For example, the behaviour of U is related to pH, Eh, the
oxidation state of U and the abundance of CO
2−
3
ions (Cumberland
et al., 2016). In addition, microbial metabolism has the capability to
alter the solubility of U (Newsome et al., 2014).
9.2. Mining
Mining of black shales is without a doubt a major cause of pollution,
even though other anthropogenic activities are known to cause acid
rock drainage as well. The permitting procedures for mining projects in
most countries require the environmental impact assessment (e.g.,
Jantunen and Kauppila, 2015). In the active mining phase, a mining
company is obligated to monitor the quality of surface water, ground-
water, lake sediments, air and biosphere. For example, at Talvivaara,
attention has been paid into the water quality of the wells used for
drinking water (Karppinen et al., 2012;Terrafame report, 2017;
Terrafame report, 2018). In the Talvivaara black shale area, bioaccu-
mulation into crop is not the issue due to the scarcity of agricultural
activities, but the bioaccumulation studies have focused on fish, wild
mushrooms and berries that are habitually consumed by local residents.
The naturally elevated geochemical background levels for Ni and Zn
and low Ca values in glacial till seem to have been reflected in the low
Ca concentrations in the blood serum of local residents already well
before the large-scale mining activities began in the area (Kantola et al.,
2008,Kousa et al., 2011).
In historical black shale mining areas, such as the Kupferschiefer in
Germany, extensive research has revealed the major sources of pollu-
tion, the weathering mechanisms and the characterisation of the mo-
bilised harmful elements (Daus and Weiß, 2001;Schubert et al., 2003,
2005;Schreck et al., 2005). This allows the evaluation of the need for
remediation measures. In the Freiberg area in Saxony with a history of
800 years of mining activities, recommendations are updated annually
for the treatment of soils contaminated with As and Cd if they are used
for agriculture or gardening (Hertwig et al., 2010). These lessons learnt
may be useful for the other mining areas to manage the impacts of acid
mine drainage.
An important issue to consider in the prevention of harmful en-
vironmental impacts is the handling of wastes after mine closure.
Effective mine closure plans are crucial in order to prevent the release
of harmful elements and compounds from residual sulphides and sec-
ondary phases to the environment (e.g.,Heikkinen et al., 2008).
9.3. Land use planning for anthropogenic actions in black shale areas
The potential for acid rock drainage should be taken into con-
sideration in any anthropogenic actions causing disturbance to the soils
or bedrock and consequently causing acceleration of chemical weath-
ering. In the past, road cuts in black shale areas have caused
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
85
mobilisation of metals into groundwater, stream water and soil (Lacelle
et al., 2007;Tuttle et al., 2009;Lacelle and Leveillé, 2010Perkins and
Mason, 2015). Disturbance of the black shale bedrock during the con-
struction of the Halifax airport in Nova Scotia, Canada, contaminated
local surface water causing fish deaths (Lund et al., 1987;King and
Hart, 1990). In China and Korea, investigations have mainly focused on
the natural weathering profiles and the bioaccumulation of metals into
plants. Bioaccumulation by agricultural plants may pose a risk to hu-
mans and therefore land use planning is recommended in the agri-
cultural areas located in black shale areas.
In Finland, the authorities use the nation-wide black shale database
in regional planning. The information is used during road construction,
excavation and peat production activities as well. As black shales cause
acid rock drainage in Finland (e.g.,Loukola-Ruskeeniemi et al., 1998)
and drilling a well into black shale bedrock may provide drinking water
with poor quality, the database is also used in planning drinking water
supplies. For example, a map has been developed to show black shale
areas located in the areas with important groundwater resources
(Tarvainen et al., 2013). In addition, comparison of the chemical con-
centration of peat with the occurrence of black shales reveals that in
average the Al, As, Ba, Cd, Co, Cr, Cu, Fe, Mn, Ni, Pb, S, Ti, V, Zn and
ash contents of peat are elevated in black shale areas (Herranen, 2009;
Herranen and Toivonen, 2018).
An integrated approach is recommended for the evaluation of the
environmental impacts of black shale taking into account the geological
context, the natural constraints and the anthropogenic actions in each
site (Fig. 8). First, mapping of the outcropping black shales, and, if
possible, also the underground horizons at regional or national level is
recommended. This helps to locate the risk areas and allows a more
advanced land use planning to avoid exposing the sulphide-rich rocks.
Secondly, the main routes of metal mobilisation should be identified.
We highlight groundwater, surface water, river and stream sediments
and soil as the main recipients of the contamination derived from
chemical, biological and mechanical weathering of black shale. Soils
and biosphere in the proximity of mines may also be impacted by
dusting of mine residues. Furthermore, when evaluating the risk to
humans, the proximity of black shale deposits to the population centres
and to the wells for household use should be investigated, as well as the
bioaccumulation into edible plants and subsequent human intake. The
area affected by metal mobilisation is relatively local in most sites, for
instance concentrations of harmful elements in surface waters decrease
considerably after roughly one kilometre. Yet, the precipitated metals
in stream bed may become remobilized upon changes in the pH and Eh
conditions.
If prior information of the environmental impacts of a black shale
deposit is not available, examining the mineralogical and geochemical
data of the deposit itself is recommended to evaluate the metal dis-
tribution, i.e., where the potentially harmful elements sit to appraise
their potential mobilisation. Risk assessment and risk management may
control the environmental impact of black shale both under natural
conditions and during anthropogenic actions. For the black shale areas,
it is recommended to follow the risk management procedures devel-
oped for example for the naturally As-rich bedrock and soil (e.g.,
Lehtinen et al., 2007;Loukola-Ruskeeniemi et al., 2007;Parviainen
et al., 2015).
Future research would benefit of joint projects between different
scientific disciplines. Comparing the microbial processes and organic
geochemistry during the genesis of black shale with the low-tempera-
ture processes during weathering or during the bioleaching of black
shale ores, could bring novel ideas for the remediation actions and for
the prevention of pollution during and after mining activities.
10. Conclusions
Black shales are common in sedimentary sequences world-wide.
They are a noteworthy source of metal contamination if exposed, even
under natural conditions, due to their high sulphur, metal and organic
matter contents. As black shale forms a heterogeneous group of rocks,
i.e., the sulphide mineralogy, associated metal assemblages, host rocks,
thickness and extension vary, pollution risks have to be evaluated in
each case. Groundwater, surface water and soil are the main environ-
mental recipients. Additionally, human exposure may occur through
the bioaccumulation of metals into crop from contaminated agricultural
soil or water and through metal-rich groundwater. Anthropogenic ac-
tivities enhance the environmental impacts of black shale. The con-
taminated waters are characteristically polymetallic: elevated con-
centrations of Cu, Ni, Pb, U, Zn, As, Cd, Co, Mn and Mo may occur
depending on the deposit type. Hence, mining activities should be
Fig. 8. A scheme for an integrated research approach for the evaluation of the environmental impacts of black shales.
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
86
carried out under regulated monitoring. In regional planning, it is re-
commended to map the distribution of sulphide-rich black shale to
avoid acid rock drainage during civil engineering actions. Further, it is
not the best alternative to drill wells into the black shale bedrock or the
overlying soil.
Acknowledgements
We acknowledge the Geological Survey of Finland for continuous
long-term funding for the black shale research. Geophysicists Hilkka
Arkimaa, Eija Hyvönen, Jouni Lerssi and Meri-Liisa Airo have had a key
role in the country-wide black shale mapping in Finland. We thank
Kirsti Keskisaari and Harri Kutvonen from the Geological Survey of
Finland for drafting Figs. 1–6. We acknowledge the ‘Juan de la Cierva
-Incorporación’ (IJCI-2016-27412) Fellowship funded by the Spanish
Ministry of Economy and Competitiveness (MINECO). We also ac-
knowledge the reviewers whose constructive reviews helped to increase
the quality of the manuscript. Finally we thank Karsten Pedersen for
valuable comments and for the efficient editorial handling.
References
Airo, M.-L., Loukola-Ruskeeniemi, K., 2004. Characterization of sulfide deposits by air-
borne magnetic and gamma-ray responses in eastern Finland. Ore Geol. Rev 24,
67–84.
Airo, M.-L., Loukola-Ruskeeniemi, K., Hyvönen, E., 2009. Geophysical databases sup-
porting large- and detailed scale mineral exploration: black shale formations in
Finland. In: Smart science for exploration and mining: proceedings of the 10th
Biennial SGA Meeting, Townsville, Australia, 17th–20th August 2009, Townsville:
James Cook University, 781–783.
Allard, B., Arsenie, I., Håkansson, K., Karlsson, S., Ahlberg, A.-C., Lundgren, T., Collin, M.,
Rasmuson, A., Strandell, E., 1991. Effects of weathering on metal releases from an
engineered deposit for alum shale leaching residues. Water Air Soil Pollut. 57–58,
431–440.
Allard, B., Bäckström, M., Haller, S., Karlsson, S., Panova, E., Grawunder, A., 2011. Water
chemistry and trace metal concentrations in an acidic alum shale pit lake – effects of
liming. In: Rüde, T.R., Freund, A., Wolkersdorfer, A. (Eds.), Mine Water: Managing
the Challenges: Proceedings of the International Mine Water Association Congress.
RWTH, Aachen, pp. 503–508.
Andersson, A., Dahlman, B., Gee, D.G., Snäll, S., 1985. The Scandinavian Alum Shales.
Sveriges Geologiska Undersökning, Ser. Ca, No. 56 (Uppsala). 50 p.
Arkimaa, H., Hyvönen, E., Lerssi, J., Loukola-Ruskeeniemi, K., Vanne, J., 1999.
Compilation of maps of black shales in Finland: applications for exploration and
environmental studies. In: Autio, S. (Ed.), Geological Survey of Finland, Current
Research 1997–1998. Vol. 27. pp. 111–114 Geological Survey of Finland. Special
Paper.
Arkimaa, H., Hyvönen, E., Lerssi, J., Loukola-Ruskeeniemi, K., Vanne, J., 2000.
Proterozoic Black Shale Formations and Aeromagnetic Anomalies in Finland 1: 1 000
000. (Map and database. Geological Survey of Finland, Espoo: Geologian tutki-
muskeskus).
Åström, M.E., Peltola, P., Rönnback, P., Lavergren, U., Bergbäck, B., Tarvainen, T.,
Backman, B., Salminen, R., 2009. Uranium in surface and groundwaters in boreal
Europe. Geochem. Explor. Environ. Anal. 9, 51–62.
Baborowski, M., Bozau, E., 2006. Impact of mining on the uranium distribution in the
system of the river saale (Germany). Appl. Geochem. 21, 1073–1082.
Bäckström, M., Sartz, L., 2015. Uranium leaching from a burning black shale deposit –
present conditions and future scenarios. In: Merkel, B., Arab, A. (Eds.), Uranium - Past
and Future Challenges. Springer, Cham. https://doi.org/10.1007/978-3-319-11059-
2_5.
Bakowska, A., Karlicki, M., Wodarczyk, A., Matlakowska, R., 2017. Geomicrobiology of
subsurface mine of the fore-sudetic monocline. Biuletyn - Panstwowego Instytutu
Geologicznego 469, 201–216.
Blowes, D.W., Ptacek, C.J., Jambor, J.L., Weisener, C.G., 2003. The geochemistry of acid
mine drainage. In: Sherwood Lollar, B. (Ed.), Treatise on Geochemistry. 9. Elsevier,
0-08-043751-6, pp. 149–204.
Borg, G., Piestrynski, A., Bchmann, G.H., Püttmann, W., Walther, S., Fiedler, M., 2012.
Overview of the European Kupferschiefer deposits. In: Hedenquist, J.W., Harris, M.,
Camus, F. (Eds.), Geology and Genesis of Major Copper Deposits and Districts of the
World: A tribute to Richard H. Sillitoe. pp. 455–486 Society of Economic Geologists,
INC.Special Publication Number 16.
Bozau, E., Licha, T., Ließmann, W., 2017. Hydrogeochemical characteristics of mine
water in the Harz Mountains. Germany. Chem. der Erde 77 (4), 614–624. https://doi.
org/10.1016/j.chemer.2017.10.001.
Buchardt, B., Nielsen, A.T., Schovsbo, N.H., 1997. Alun Skiferen i Skandinavien. Geol.
Tidsskr. 3, 1–30 (In Danish).
Butterworth, B.P., Caulfield, D., 1997. Taiga property: A stratiform Ni-Zn-PGE target in
north-central Yukon. In: Roots, C.F. (Ed.), Yukon Exploration and Geology 1997,
Indian & Northern Affairs Canada/Department of Indian & Northern Development:
Exploration & Geological Services Division, 125–127.
Capistrant, P.L., Hitzman, M.W., Wood, D., Kelly, N.M., Williams, G., Zimba, M., Kuiper,
Y., Jack, D., Stein, H., 2015. Geology of the enterprise hydrothermal nickel deposit,
North-Western Province, Zambia. Econ. Geol. 110 (1), 9–38.
Carlsson, E., Büchel, G., 2005. Screening of residual contamination at a former uranium
heap leaching site, Thuringia, Germany. Chem. der Erde 65 (S1), 75–95.
Chi Fru, E., Hemmingsson, C., Callac, N., Perez, N., Panova, E.G., Broman, C., El Albani,
A., 2016. Atmospheric weathering of Scandinavian alum shales and the fractionation
of C, N and S isotopes. Appl. Geochem. 74, 94–108.
Chon H.-T., Oh, S.-Y., 2000. Hydrogeochemical characteristics and contamination of
surface water and groundwater in the middle Okchon zone, Korea. Geosystem Eng. 3
(3), 98–107.
Coveney, R.M., Chen, N., 1991. Ni-Mo-PGE-Au-rich ores in Chinese black shales and
speculations on possible analogues in the United States. Mineral. Deposita 26, 83–88.
Coveney, R.M., Glascock, M.D., 1989. A review of the origins of metal-rich Pennsylvanian
black shales, central U.S.A., with an inferred role for basinal brines. Appl. Geochem.
4, 347–367.
Coveney Jr., R.M., 2000. Metalliferous black shales and the role of organic matter with
examples from China, Poland and the United States. In: Giordano, T.H., Kettler, R.M.,
Wood, S.A. (Eds.), Ore Genesis and Exploration: The Roles of Organic Matter.
Reviews in Economic Geology 9. pp. 251–280 Society of Economic Geologists,
Littleton, Colorado, USA.
Coveney Jr., R.M., Grauch, R.I., Murowchick, J.B., 1994. Metals, phosphate and stone
coal in the Proterozoic and Cambrian of China: the geologic setting of precious metal-
bearing Ni-Mo ore beds. Soc. Econ. Geol. Newsletter 18, 6–11.
Coveney Jr., R.M., Pašava, J., 2004. Diverse connections between ores and organic
matter. Ore Geol. Rev. 24, 1–5.
Cumberland, S.A., Douglas, G., Grice, K., Moreau, J.W., 2016. Uranium mobility in or-
ganic matter-rich sediments: a review of geological and geochemical processes.
Earth-Sci. Rev. 159, 160–185.
Daus, B., Weiß, H. (Eds.), 2001. Fine-grained residues from copper smelting and their
environmental impacts – a case study from the Mansfeld District, Germany. UFZ-
Bericht 22/2001. UFZ Leipzig-Halle GmbH, Leipzig 113 p.
DigiKP, 2017. Digital map database of the bedrock of Finland. Geological Survey of
Finland. http://gtkdata.gtk.fi/Kalliopera/index.html (Accessed 11.1.2019).
Ervamaa, P., Heino, T., 1983. The Ni-Cu-Zn mineralization of Talvivaara, Sotkamo. In:
Exogenic Processes and Related Metallogeny in the Svecokarelian Geosynclinals
Complex. Geological Survey of Finland Guide 11. pp. 79–83.
Ettler, V., Mihaljevič, M., Kříbek, B., Majer, V., Šebek, O., 2011. Tracing the spatial dis-
tribution and mobility of metal/metalloid contaminants in Oxisols in the vicinity of
the Nkana copper smelter, Copperbelt province, Zambia. Geoderma 164, 73–84.
Falk, H., Lavergren, U., Bergbäck, B., 2006. Metal mobility in alum shale from Öland,
Sweden. J. Geochem. Explor. 90, 157–165.
Fan, D., 1983. Polyelements in the Lower Cambrian black shale series in southern China.
In: Augustithis, S.S. (Ed.), The Significance of Trace Metals in Solving Petrogenetic
Problems and Controversies. Theophrastus Publications S.A, Athens, pp. 447–474.
Fan, D., Zhang, T., Ye, J., Pašava, J., Kříbek, B., Dobeš, P., Vavřín, I., Žák, K., 2004.
Geochemistry and origin of tin-polymetallic sulfide deposits hosted by the Devonian
black shale series near Dachang, Guangxi, China. Ore Geol. Rev. 24 (1–2), 103–120.
Fang, W., Hu, R., Wu, P., 2002. Influence of black shales on soils and edible plants in the
Ankang area, Shaanxi province, P.R. of China. Environ. Geochem. Health 24, 35–46.
Fraser, T., 2014. Field descriptions of the Middle-Upper Devonian Canol formation on
Trail River, East Richardson Mountains, Yukon. In: MacFarlane, K.E., Nordling, M.G.,
Sack, P.J. (Eds.), Yukon Exploration and Geology, pp. 53–68 2013, Yukon Geological
Survey.
Fraser, T.A., Hutchison, M.P., 2017. Lithogeochemical characterization of the
Middle–Upper Devonian road river group and Canol and imperial formations on Trail
River, east Richardson mountains, Yukon: age constraints and a depositional model
for fine-grained strata in the Lower Paleozoic Richardson trough. Can. J. Earth Sci. 54
(7), 731–765. https://doi.org/10.1139/cjes-2016-0216.
Goodfellow, W.D., 2004. Geology, genesis, and exploration of SEDEX deposits, with
emphasis on the Selwyn Basin, Canada. In: Deb, M., Goodfellow, W.D. (Eds.),
Sediment Hosted Lead-Zinc Sulfide Deposits, New Delhi, India, ISBN 81-7319-520-X,
24–99.
Goodfellow, W.D., Geldsetzer, H., Gregoire, C., Orchard, M., Cordey, F., 2010.
Geochemistry and origin of geographically extensive Ni (Mo, Zn, U)-PGE sulfide
deposits hosted in Devonian black shales, Yukon. TGI-3 Workshop: Public geoscience
in support of base metal exploration programme and abstracts; by Geological
Association of Canada, Cordilleran Section; 15–18.
Grauch, R.I., Huyck, H.L.O., 1990. Metalliferous Black Shales and related ore deposits –
proceedings, 1989 United States working group meeting, International Geological
Correlation Program Project 254. U.S. Geological Survey Circular 1058, pp. 85.
Grawunder, A., Lonschinski, M., Merten, D., Büchel, G., 2009. Distribution and bonding of
residual contamination in glacial sediments at the former uranium mining leaching
heap of Gessen/Thuringia Germany. Chem. der Erde 69 (S2), 5–19.
Greenwood, P.F., Brocks, J.J., Grice, K., Schwark, L., Jaraula, C.M.B., Dick, J.M., Evans,
K.A., 2013. Organic geochemistry and mineralogy I. Characterisation of organic
matter associated with metal deposits. Ore Geol. Rev. 50, 1–27.
Gustavsson, N., Loukola-Ruskeeniemi, K., Tenhola, M., 2012. Evaluation of geochemical
background levels around sulfide mines – a new statistical procedure with beanplots.
Appl. Geochem. 27, 240–249.
Han, T., Fan, H., Zhu, X., Wen, H., Zhao, C., Xiao, F., 2017. Submarine hydrothermal
contribution for the extreme element accumulation during the early Cambrian. South
China. Ore Geol. Rev. 86, 297–308.
Hautaniemi, H., Kurimo, M., Multala, J., Leväniemi, H., Vironmäki, J., 2005. The ‘three in
one’ aerogeophysical concept of GTK in 2004. In: Airo, M.-L. (Ed.), Aerogeophysics in
Finland 1972–2004: Methods, System Characteristics and Applications, Geological
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
87
Survey of Finland, Special Paper 39, pp. 21–74.
Heikkinen, P., Noras, P., Salminen, R., 2008. Mine closure handbook. Geological Survey
of Finland. Special Publications 169 p. http://tupa.gtk.fi/julkaisu/erikoisjulkaisu/ej_
074.pdf.
Helios Rybicka, E., Wilson, M.J., McHardy, W.J., 1994. Chemical and mineralogical forms
and mobilization of copper and lead in soils from a Cu-smelting area in Poland. J.
Environ. Sci. Health. Part A: Environ. Sci. Engineer. Toxic. 29 (3), 531–546.
Herranen, T., 2009. Turpeen rikkipitoisuus Suomessa. Abstract: The sulfur concentration
of peat in Finland. In: Geological Survey of Finland, Report of Peat Investigation 398.
55 pages, 19 figures and 27 tables, Available at: http://tupa.gtk.fi/raportti/turve/
ttr_398.pdf.
Herranen, T., Toivonen, T., 2018. Turpeen alkuainemääritykset GTK:ssa. Summary: The
element assays of peat in the GTK. In: Geological Survey of Finland, Open File
Research Report 73/2018. 45 pages, 37 figures and 3 tables, Available at: http://
tupa.gtk.fi/raportti/arkisto/73_2018.pdf.
Hertwig, Th., Müller, I., Zeißler, K., 2010. Management of contaminated soils in urban
areas in the Ore Mountains (Germany). In: ConSoil 2010, 11th International
Conference on Management of Soil, Groundwater and Sediment, Salzburg, Austria,
22–24 September, pp. 2010.
Hulbert, L.J., Gregoire, D.C., Paktunc, D., Carne, R.C., 1992. Sedimentary nickel, zinc,
and platinum-group-element mineralization in Devonian black shales at the Nick
Property, Yukon, Canada: a new deposit type. Expl. Min. Geol. 1 (1), 39–62.
Huyck, H.L.O., 1990. Proposed definition of “black shale” and “metalliferous black shale”
for IGCP #254. In: Book of Abstracts, 8th IAGOD Symposium, A183–184, GSC,
Ottawa.
Hyvönen, E., Airo, M.-L., Arkimaa, H., Lerssi, J., Loukola-Ruskeeniemi, K., Vanne, J.,
Vuoriainen, S., 2013. Airborne geophysical, petrophysical and geochemical char-
acteristics of Palaeoproterozoic black shale units in Finland: applications for ex-
ploration and environmental studies. In: Geologian tutkimuskeskus, Tutkimusraportti
198 – Geological Survey of Finland, Report of Investigation 198, 2013 Current
Research − GTK Mineral Potential Workshop, Rauhalahti/Kuopio, May 2012, pp.
58–60.
Jambor, J.L., 1994. Mineralogy of sulfide-rich tailings and their oxidation products. In:
Blowes, D.W., Jambor, J.L. (Eds.), The Environmental Geochemistry of Sulfide Mine-
wastes, Short Course Handbook 22. Mineralogical Association of Canada, Waterloo,
Canada, pp. 59–102.
Guide: Environmental Impact Assessment Procedure for mining projects in Finland. In:
Jantunen, J., Kauppila, T. (Eds.), MEE guidelines and other publications 11/2015,
Ministry of Employment and the Economy, Finland, 102 p. http://en.gtk.fi/export/
sites/en/mineral_resources/EIA_guidelines_for_mining_projects_in_Finland_2015.pdf.
Jiang, S.Y., Yang, J.H., Ling, H.F., Chen, Y.Q., Feng, H.Z., Zhao, K.D., Ni, P., 2007.
Extreme enrichment of polymetallic Ni–Mo–PGE–Au In Lower Cambrian Black Shales
of South China: an Os isotope and PGE geochemical investigation. Palaeogeogr.
Palaeoclimatol. Palaeoecol. 254, 217–228.
Jin, L., Mathur, R., Rother, G., Cole, D., Bazilevskaya, E., Williams, J., Carone, A.,
Brantley, S., 2013. Evolution of porosity and geochemistry in Marcellus formation
black shale during weathering. Chem. Geol. 356, 50–63.
Kaiser, D., Konovalov, S., Schultz-Bull, D., Waniek, J.J., 2017. Organic matter along
longitudinal and vertical gradients in the Black Sea. Deep Sea Res Part I 129, 22–31.
Kantola, M., Loukola-Ruskeeniemi, K., Hatakka, T., 2008. Black shale bedrock is reflected
in trace element concentration profile of blood and hair of local residents. In: 33rd
International Geological Congress, 6–14 August 2008, Oslo, Norway: abstract CD-
ROM. 1 p. Optical disc (CD-ROM), (Electronic resource).
Kantola, M., Seppänen, K., Halonen, T., Henttonen, P., Loukola-Ruskeeniemi, K., 2000.
Mercury and some trace element levels in local residents living in black shale areas in
eastern Finland. In: Ympäristö, terveys ja turvallisuus kaivannaisteollisuudessa:
seminaari 31.10–1.11.2000 Haikon kartanossa: esitysten lyhennelmät.
Vuorimiesyhdistys. Sarja B 76. Vuorimiesyhdistys, pp. 30–32.
Kao, L.-S., Peacor, D.R., Coveney Jr., R.M., Zhao, G., Dungey, K.E., Curtis, M.D., Penner-
Hahn, J.E., 2001. A C/MoS2 mixed-layer phase (MoSC) occurring in metalliferous
black shales from southern China, and new data on jordisite. Amer. Mineral. 86
(7–8), 852–861.
Karppinen, H., Komulainen, H., Kousa, A., Nikkarinen, M., Tornivaara, A., 2012.
Haitalliset alkuaineet Kainuun kaivovesissä. Loppuraportti. (Harmful elements in
well waters in Kainuu. Final report) Kainuun maakunta-kuntayhtymä. ISSN 1795-
5661 (In Finnish). http://www.ely-keskus.fi/documents/10191/1145691/Haitalliset
+aineet+-loppuraportti.pdf/0a5f2c19-0504-4c9f-86f0-4fa6c2a0ad98 (Accessed
21.3.2018).
Kennedy, M.J., Pevear, D.R., Hill, R.J., 2002. Mineral surface control of organic carbon in
Black Shale. Science 295, 657–660.
Kim, J.W., 1989. Geochemistry and genesis of the Guryongsan (Ogcheon) uraniferous
black slate. J. Korean Inst. Min. Geol. 22, 35–63.
Kim, K.-W., Thornton, I., 1993a. Influence of uraniferous black shales on cadmium,
molybdenum and selenium in soils and crop plants in the Deog-Pyoung area of Korea.
Environ. Geochem. Health 15, 119–133.
Kim, K.-W., Thornton, I., 1993b. Influence of Ordovician uraniferous black shaleson the
trace element composition of soils and food crops, Korea. Appl. Geochem., Suppl
Issue 2, 249–255.
Kim, Y.J., Seo, J., Kang, S.-A., Choi, S.-G., Lee, Y.-J., 2015. Geochemistry and uranium
mineralogy of the black slate in the Okcheon Metamorphic belt, South Korea.
Geochem. J. 49 (4), 443–452.
King, M., Hart, B., 1990. Groundwater contribution to acid drainage from the Halifax
formation in Nova Scotia. Canad. Water Res. J. 15 (4), 357–365. https://doi.org/10.
4296/cwrj1504357.
Kontinen, A., 2012. F029 Talvivaara Ni-Zn-Cu. In: Eilu, P. (Ed.), Mineral Deposits and
Metallogeny of Fennoscandia, pp. 276–280 Geological Survey of Finland, Special
Paper 53.
Kontinen, A., Hanski, E., 2015. The Talvivaara Black Shale-Hosted Ni-Zn-Cu-Co deposit in
Eastern Finland. In: Maier, W., Lahtinen, R., O'Brien, H. (Eds.), Mineral Deposits of
Finland. Elsevier, pp. 557–612.
Kousa, A., Loukola-Ruskeeniemi, K., Nikkarinen, M., Havulinna, A.S., Karvonen, M.,
Moltchanova, E., Sorvari, J., Lehtinen, H., Rossi, E., Ruskeeniemi, T., Backman, B.,
Mäkelä-Kurtto, R., Kantola, M., Hatakka, T., Savolainen, H., 2011. Evaluation of the
relationship between the natural geological environment and certain chronic diseases
in Finland. In: Geological Survey of Finland, Special Paper 49, 247–262, 4 figures
and 2 tables.
Kucha, H., 1990. Geochemistry of the Kupferschiefer, Poland. Geol. Rundsch. 79 (2),
387–399.
Kulick, J., Leifeld, D., Meisl, S., Pöschl, W., Stellmacher, R., Strecker, G., Theuerjahr, A.-
K., Wolf, M., 1984. Petrofazielle und chemische Erkundung des Kupferschiefers der
Hessischen Senke und des Harz-Westrandes: Geologisches Jahrbuch, v. D68. pp. 223.
Kwong, Y.T.J., Roots, C.F., Roach, P., Kettley, W., 1997. Post-mine metal transport and
attenuation in the Keno Hill mining district, central Yukon. Canada. Environ. Geol.
30, 98–107.
Kwong, Y.T.J., Whitley, G., Roach, P., 2009. Natural acid rock drainage associated with
black shale in the Yukon Territory. Canad. Appl. Geochem. 24, 221–231.
Laatikainen, T., Seppänen, A., 2017. In: Seppänen, A. (Ed.), Epifyyttijäkälien alkuainea-
nalyysit (Elemental analysis of Lichen). 978-952-314-593-1, Kainuun
Bioindikaatoriselvitys. Elinvoimaa alueelle 3, pp. 115, Centre for Economic
Development, Transport and the Environment. (In Finnish).
Lacelle, D., Doucet, A., Lauriol, B., Clark, I.D., 2007. Acid drainage generation and sea-
sonal recycling in disturbed permafrost near Eagle Plains, northern Yukon Territory.
Canada. Chem. Geol. 243, 157–177.
Lacelle, D., Leveillé, R., 2010. Acid drainage generation and associated Ca–Fe–SO
4
mi-
nerals in a periglacial environment, Eagle Plains, Northern Yukon. Canada: A po-
tential analogue for low-temperature sulfate formation on Mars. Planet. Space Sci. 58,
509–521.
Lavergren, U., Åström, M.E., Bergbäck, B., Holmström, H., 2009a. Mobility of trace ele-
ments in black shale assessed by leaching tests and sequential chemical extraction.
Geochem. Explor. Environ. Anal. 9, 71–79.
Lavergren, U., Åström, M.E., Falk, H., Bergbäck, B., 2009b. Metal dispersion in ground-
water in an area with natural and processed black shale – Nationwide perspective and
comparison with acid sulfate soils. Appl. Geochem. 24, 359–369.
Lecomte, A., Cathelineau, M., Deloule, E., Brouand, M., Peiffert, C., Loukola-Ruskeeniemi,
K., Pohjolainen, E., Lahtinen, H., 2014. Uraniferous bitumen nodules in the
Talvivaara Ni–Zn–Cu–Co deposit (Finland): influence of metamorphism on uranium
mineralization in black shales. Mineral. Deposita 49, 513–533.
Lecomte, A., Cathelineau, M., Michels, R., Peiffert, C., Brouand, M., 2017. Uranium mi-
neralization in the Alum Shale formation (Sweden): evolution of a U-rich marine
black shale from sedimentation to Metamorphism. Ore Geol. Rev. 88, 71–98.
Lee, J.-S., Chon, H.-T., Kim, J.S., Kim, K.-W., Moon, H.-S., 1998. Enrichment of potentially
toxic elements in areas underlain by black shales and slates in Korea. Environ.
Geochem. Health 20, 135–147.
Lehmann, B., Frei, R., Xu, L., Mao, J., 2016. Early Cambrian Black Shale-hosted Mo-Ni
and V mineralization on the rifted margin of the Yangtze platform, China:
Reconnaissance chromium isotope data and a refined metallogenic model. Econ.
Geol. 111, 89–103.
Lehmann, B., Mao, J.W., Li, S.R., Zhang, G.D., Zeng, M.G., 2003. Re-Os dating of poly-
metallic Ni-Mo-PGE-Au mineralization in Lower Cambrian black shales of South
China and its geologic significance—A reply: Econ. Geology 98, 663–665.
Lehmann, B., Nägler, T.F., Holland, H.D., Wille, M., Mao, J., Pan, J., Ma, D., Dulski, P.,
2007. Highly metalliferous carbonaceous shale and early Cambrian seawater. Geol.
35 (5), 403–406.
Lehtinen, H., Sorvari, J., Pyy, O., 2007. Risk Management of Environmental Arsenic in
Finnish conditions - case Pirkanmaa region. Geological Survey of Finland,
Miscellaneous Publications, 66 p. http://projects.gtk.fi/export/sites/projects/ramas/
reports/RiskManagement2007.pdf.
Leppänen, J.J., Weckström, J., Korhola, A., 2017. Multiple mining impacts induce
widespread changes in ecosystem dynamics in a boreal lake. Sci. Rep 7, 10581.
https://doi.org/10.1038/s41598-017-11,421-8.
Leventhal, J., 1993. Metals in black shales. In: Engel, M.H., Macko, S.A. (Eds.), Organic
Geochemistry: Principles and Applications. Plenum Press, New York, pp. 581–607.
Leventhal, J.S., Hosterman, J.W., 1982. Chemical and mineralogical analysis of Devonian
black-shale samples from Martin County, Kentucky; Carroll and Washington counties,
Ohio; Wise County, Virginia; and Overton County, Tennessee, U.S.A. Chem. Geol. 37,
239–264.
Ling, S., Wu, X., Ren, Y., Sun, C., Liao, X., Li, X., Zhu, B., 2015. Geochemistry of trace and
rare earth elements during weathering of black shale profiles in Northeast
Chongqing, Southwestern China: their mobilization, redistribution, and fractionation.
Chem. der Erde 75, 403–417.
Liu, Y., Xiao, T., Perkins, R.B., Zhu, J., Zhu, Z., Xiong, Y., Ning, Z., 2017. Geogenic
cadmium pollution and potential health risks, with emphasis on black shale. J.
Geochem. Explor. 176, 42–49.
Lott, D.A., Coveney Jr., R.M., Murowchick, J.B., Grauch, R.I., 1999. Sedimentary ex-
halative nickel–molybdenum ores in south China. Econ. Geol. 94, 1051–1066.
Loukola-Ruskeeniemi, K., 1991. Geochemical evidence for the hydrothermal origin of
sulphur, base metals and gold in Proterozoic metamorphosed black shales, Kainuu
and Outokumpu areas, Finland. Mineral. Deposita 26 (2), 152–164.
Loukola-Ruskeeniemi, K., 1992. Geochemistry of Proterozoic metamorphosed black
shales in eastern Finland, with implications for exploration and environmental stu-
dies. In: Geological Survey of Finland, Erikoisjulkaisut - Special Publications 9, 86 p.
Loukola-Ruskeeniemi, K., 1995. Origin of the black-shale-hosted Ni-Cu-Zn deposit at
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
88
Talvivaara, Finland. Geological Survey of Finland, Special Paper 20. pp. 31–46.
Loukola-Ruskeeniemi, K., 1999. Origin of black shales and serpentinite-associated Cu-Zn-
Co ores at Outokumpu in Finland. Econ. Geol. 94, 1007–1028.
Loukola-Ruskeeniemi, K., Heino, T., 1996. Geochemistry and genesis of the black shale-
hosted Ni-Cu-Zn deposit at Talvivaara, Finland. Econ. Geol. 91 (1), 80–110.
Loukola-Ruskeeniemi, K., Heino, T., Talvitie, J., Vanne, J., 1991. Base-metal-rich meta-
morphosed black shales associated with Proterozoic ophiolites in the Kainuu schist
belt, Finland: a genetic link with the Outokumpu rock assemblage. Mineral. Deposita
26, 143–151.
Loukola-Ruskeeniemi, K., Hyvonen, E., Airo, M.-L., Arkimaa, H., Eskelinen, J., Lerssi, J.,
Vanne, J., Vuoriainen, S., 2011. Onko Suomessa uusia Talvivaara-tyyppisia malmeja?
Geofysikaalisiin ja geokemiallisiin tutkimuksiin perustuva Suomen mustaliuske-
kartta. Abstract: Evaluation of the ore potential of black shale units in Finland –
preliminary results of a geophysical and geochemical study. Geologi 63 (3), 68–79.
Loukola-Ruskeeniemi, K., Kantola, M., Seppänen, K., Henttonen, P., Kallio, E., Kurki, P.,
Savolainen, H., 2003. Mercury-bearing black shales and human Hg intake in eastern
Finland: impact and mechanisms. Environ. Geol. 43, 283–297.
Loukola-Ruskeeniemi, K., Lahtinen, H., 2013. Multiphase evolution in the black-shale-
hosted Ni–Cu–Zn–Co deposit at Talvivaara, Finland. Ore Geol. Rev. 52, 85–99.
Loukola-Ruskeeniemi, K., Ruskeeniemi, T., Parviainen, A., Backman, B., 2007. Arsenic in
the Pirkanmaa Region in Finland: Occurrence in the Environment, Risk Assessment
and Risk Management. Final results of the EU LIFE Environment Programme project
RAMAS. Helsinki University of Technology, Geoenvironmental Technology Section,
Special Publications. ISBN 978-951-22-9117-5 (In Finnish) Layman's report in
English: http://projects.gtk.fi/export/sites/projects/ramas/reports/
RAMASLaymansReport.pdf.
Loukola-Ruskeeniemi, K., Uutela, A., Tenhola, M., Paukola, T., 1998. Environmental
impact of metalliferous black shales at Talvivaara in Finland, with indication of lake
acidification 9000 years ago. J. Geochem. Explor. 64, 395–407.
Lund, O.P., Vaughan, J.G., Thirumurthi, D., 1987. Impact of acid drainage pollution from
mineralized slate at Halifax airport. Water Pollution Research Journal of Canada 222,
308–325.
Macquaker, J.H.S., Bohacs, K.M., 2007. On the accumulation of mud. Science 318, 1734.
https://doi.org/10.1126/science.1151980.
Mäkilä, M., Loukola-Ruskeeniemi, K., Säävuori, H., 2012. High Pre-mining Metal
Concentrations and Conductivity in Peat Around the Talvivaara Nickel Deposit,
Eastern Finland. Tutkimusraportti - Report of Investigation. Geological Survey of
Finland, 36 p.
Mäkinen, J., Kauppila, T., Loukola-Ruskeeniemi, K., Mattila, J., Miettinen, J., 2010.
Impacts of point source and diffuse metal and nutrient loading on three northern
boreal lakes. J. Geochem. Explor. 104, 47–60.
Makkonen, H.V., Halkoaho, T., Konnunaho, J., Rasilainen, K., Kontinen, A., Eilu, P., 2017.
Ni-(Cu-PGE) deposits in Finland – geology and exploration potential. Ore Geol. Rev.
90, 667–696.
Mao, J., Lehmann, B., Du, A., Zhang, G., Ma, D., Wand, Y., Zeng, M., Kerrich, R., 2002. Re-
Os dating of polymetallic Ni–Mo–PGE–Au mineralization in Lower Cambrian black
shales and its geologic significance. Econ. Geol. 47, 151–1061.
Marynowski, L., Pisarzowska, A., Derkowski, A., Rakociński, M., Szaniawski, R., Środoń,
J., Cohen, A.S., 2017. Influence of palaeoweathering on trace metal concentrations
and environmental proxies in black shales. Palaeogeogr. Palaeoclimatol. Palaeoecol.
472, 177–191.
Matlakowska, R., Skłodowska, A., Nejbert, N., 2012. Bioweathering of Kupferschiefer
black shale (Fore-Sudetic Monocline, SW Poland) by indigenous bacteria: implication
for dissolution and precipitation of minerals in deep underground mine. FEMS
Microbiol. Ecol. 81, 99–110.
McGowan, R.R., Roberts, S., Boyce, A.J., 2006. Origin of the Nchanga copper–cobalt
deposits of the Zambian Copperbelt. Mineral. Deposita 40, 617–638. https://doi.org/
10.1007/s00126-005-0032-8.
Melezhik, V.A., Fallick, A.E., Filippov, M.M., Larsen, O., 1999. Karelian shungite—an
indication of 2.0-Ga-old metamorphosed oil-shale and generation of petroleum:
geology, lithology and geochemistry. Earth-Sci. Rev. 47, 1–40.
Morel, F.M.M., Price, N.M., 2003. The biogeochemical cycles of trace metals in the
oceans. Science 300, 944. https://doi.org/10.1126/science.1083545.
Müller, N., Franke, K., Schreck, P., Hirsch, D., Kupsch, H., 2008. Georadiochemical evi-
dence to weathering of mining residues of the Mansfeld mining district. Germany.
Environ. Geol. 54, 869–877.
Newsome, L., Morris, K., Lloyd, J.R., 2014. The biogeochemistry and bioremediation of
uranium and other priority radionuclides. Chem. Geol. 363, 164–184.
Nordstrom, D.K., 2011. Hydrogeochemical processes governing the origin, transport and
fate of major and trace elements from mine wastes and mineralized rock to surface
waters. Appl. Geochem. 26, 1777–1791.
Nordstrom, D.K., Blowes, D.W., Ptacek, C.J., 2015. Hydrogeochemistry and microbiology
of mine drainage: An update. Appl. Geochem. 57, 3–16.
Ntengwe, F.W., Maseka, K.K., 2006. The impact of effluents containing zinc and nickel
metals on stream and river water bodies: the case of Chambishi and Mwambashi
streams in Zambia. Phys. Chem. Earth 31, 814–820.
Núñez, O., Fernández-Navarro, P., Martín-Méndez, I., Bel-Lan, A., Locutura, J.F., López-
Abente, G., 2016. Arsenic and chromium topsoil levels and cancer mortality in Spain.
Environ. Sci. Pollut. Res. Int. 23 (17), 17664–17675.
Och, L.M., Shields-Zhou, G.A., Poulton, S.W., Manning, C., Thirlwall, M.F., Li, D., Chen,
X., Ling, H.F., Osborn, T., Cremonese, L., 2013. Redox changes in early Cambrian
black shales at Xiaotan section, Yunnan province, South China. Precambrian Res.
225, 166–189.
Orberger, B., Pašava, J., Paul Gallien, J., Daudin, L., Trocellier, P., 2003. Se, As, Mo, Ag,
Cd, In, Sb, Pt, Au, Tl, Re traces in biogenic and abiogenic sulfides from Black Shales
(Selwyn Basin, Yukon territories, Canada): a nuclear microprobe study. Nucl.
Instrum. Methods Phys. Res., Sect. B 210, 441–448.
Oszczepalski, S., 1989. Kupferschiefer in southwestern Poland: Sedimentary environ-
ments, metal zoning, and ore controls. In: Boyle, R.W., Brown, A.C., Jefferson, G.F.,
Jowett, E.C., Kirkham, R.V. (Eds.), Sediment-hosted Stratiform Copper Deposits, pp.
571–600 Geol. Assoc. Can., Spec Pap 36.
Oszczepalski, S., 1999. Origin of the Kupferschiefer polymetallic mineralization in
Poland. Mineral. Deposita 34, 599–613.
Pagès, A., Barnes, S., Schmid, S., Coveney, R., Schwark, L., Liu, W., Grice, K., Fan, H.,
Wen, H., 2018. Geochemical investigation of the lower Cambrian mineralised black
shales of South China and the late Devonian Nick deposit. Canada. Ore Geol. Rev. 94,
396–413.
Paikaray, S., 2012. Environmetnal hazards of arsenic associated with black shales: a re-
view on geochemistry, enrichment and leaching mechanism. Rev. Environ. Sci.
Biotechnol. 11, 289–303.
Park, M., Chon, H.-T., Marton, L., 2010. Mobility and accumulation of selenium and its
relationship with other heavy metals in the system rocks/soils–crops in areas covered
by black shale in Korea. J. Geochem. Explor. 107, 161–168.
Parviainen, A., Loukola-Ruskeeniemi, K., Tarvainen, T., Hatakka, T., Härmä, P., Backman,
B., Ketola, T., Kuula, P., Lehtinen, H., Sorvari, J., Pyy, O., Ruskeeniemi, T., Luoma, S.,
2015. Arsenic in bedrock, soil and groundwater – the first arsenic guidelines for
aggregate production established in Finland. Earth-Sci. Rev. 150, 709–723.
Parviainen, A., Mäkilä, M., Loukola-Ruskeeniemi, K., 2014. Pre-mining acid rock drainage
in the Talvivaara Ni–Cu–Zn–Co deposit (Finland): Natural peat layers as a natural
analog to constructed wetlands. J. Geochem. Explor. 143, 84–95.
Pašava, J., 1993. Anoxic sediments – an important source for PGE; an overview. Ore Geol.
Rev. 8, 425–445.
Pašava, J., Gabriel, Z., 1988. IGCP begins work on metalliferous black shales. Episodes 11
(2), 140–141.
Pašava, J., Hladíková, J., Dobeš, P., 1996. Origin of Proterozoic metal-rich black shales
from the Bohemian Massif, Czech Republic. Econ. Geol. 91, 63–79.
Pašava, J., Kříbek, B., Vymazalová, A., Sýkorová, I., Žák, K., Orberger, B., 2008. Multiple
sources of metals of mineralization in Lower Cambrian black shales in south China:
evidence from geochemical and petrographic study. Resour. Geol. 58, 25–42.
Pašava, J., Kříbek, B., Žák, K., Li, C., Deng, H., Liu, J., Gao, Z., Luo, T., Zeng, M., 2003.
Environmental impacts of mining of Ni-Mo black shale-hosted deposits in the Zunyi
region, southern China: Preliminary results of the study of toxic metals in the system
rock–soil–plant. Bull. Geosci. 78 (3), 251–260.
Peltola, E., 1960. On the black schists in the Outokumpu region in eastern Finland.
Bulletin de la Commission géologique de Finlande 192 107 p.
Peng, B., Rate, A., Song, Z., Yu, C., Tang, X., Xie, S., Tu, X., Tan, C., 2014. Geochemistry of
major and trace elements and Pb-Sr isotopes of a weathering profile developed on the
Lower Cambrian black shales in central Hunan. China. Appl. Geochem. 51, 191–203.
Peng, B., Song, Z., Tu, X., Xiao, M., Wu, F., Lv, H., 2004. Release of heavy metals during
weathering of the Lower Cambrian Black Shales in western Hunan. China. Environ.
Geol. 45, 1137–1147.
Peng, B., Tang, X., Yu, C., Xie, S., Xiao, M., Song, Z., Tu, Z., 2009a. Heavy metal geo-
chemistry of the acid mine drainage discharged from the Hejiacun uranium mine in
central Hunan. China. Environ Geol. 57, 421–434.
Peng, B., Tang, X., Yu, C., Xu, L., Xie, S., Yang, G., Yin, C., Tu, X., 2009b. Geochemical
study of heavy metal contamination of soils derived from black shales at the HJC
uranium mine in central Hunan. China. Acta Geol. Sin. 83 (1), 89–106.
Pérez-Huerta, A., Coronadoc, I., Hegnad, T.A., 2018. Understanding biomineralization in
the fossil record. Earth-Sci. Rev. 179, 95–122.
Perkins, R.B., Mason, C.E., 2015. The relative mobility of trace elements from short-term
weathering of a black shale. Appl. Geochem. 56, 67–79.
Petsch, S.T., Edwards, K.J., Eglinton, T.I., 2005. Microbial transformations of organic
matter in black shales and implications for global biogeochemical cycles.
Palaeogeogr. Palaeoclimatol. Palaeoecol. 219, 157–170.
Pieczonka, J., Piestrzynski, A., Zygo, W., Cwiertnia, T., 2015. Geology of the European
Kupferschiefer Lubin, Poland. Excursion Guidebook. 13th SGA Biennial meeting
Nancy-France 24–27 August 2015, 50 p.
Piper, D.Z., Calvert, S.E., 2009. A marine biogeochemical perspective on black shale
deposition. Earth-Sci. Rev. 95, 63–96.
Polgári, M., Hein, J.R., Vígh, T., Szabó-Drubin, M., Fórizs, I., Bíró, L., Müller, A., Tóth, L.,
2012. Microbial processes and the origin of the Úrkút manganese deposit. Hungary.
Ore Geol. Rev. 47, 87–109.
Reynaud, J.Y., Trentesaux, A., El Albani, A., Aubineau, J., Ngombi-Pemba, L., Guiyeligou,
G., Bouton, P., Gauthier-Lafaye, F., Weber, F., 2018. Depositional setting of the 2·1 Ga
Francevillian macrobiota (Gabon): Rapid mud settling in a shallow basin swept by
high-density sand flows. Sedimentology 65 (3), 670–701.
Riekkola-Vanhanen, M., 2010. Talvivaara Sotkamo mine – bioleaching of a polymetallic
nickel ore in subarctic climate. Nova Biotechnol 10–1, 7–14.
Rimstidt, J.D., Chermak, J.A., Schreiber, M.E., 2017. Processes that control mineral and
element abundances in shales. Earth-Sci. Rev. 171, 383–399.
Roen, J.B., 1984. Geology of the Devonian black shales of the Appalachian Basin. Org.
Geochem. 5 (4), 241–254.
Salminen, R., Tarvainen, T., 1995. Geochemical mapping and databases in Finland. J.
Geochem. Explor. 55, 321–327.
Sanei, H., Petersen, H.I., Schovsbo, N.H., Jiang, C., Goodsite, M.E., 2014. Petrographic
and geochemical composition of kerogen in the Furongian (U. Cambrian) Alum Shale,
central Sweden: reflections on the petroleum generation potential. Int. J Coal Geol.
132, 158–169.
Schäffner, F., Merten, D., Pollok, K., Wagner, S., Knoblauch, S., Langenhorst, F., Büchel,
G., 2015. Fast formation of supergene Mn oxides/hydroxides under acidic conditions
in the oxic/anoxic transition zone of a shallow aquifer. Environ. Sci. Pollut. Res. 22,
19362–19375.
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
89
Schovsbo, H., 2002. Uranium enrichment shorewards in black shales: a case study from
the Scandinavian Alum Shale. GFF 124 (2), 107–115.
Schreck, P., Schubert, M., Freyer, K., Treutler, H.C., Weiss, H., 2005. Multi-metal con-
taminated stream sediment in the Mansfeld mining district: metal provenance and
source detection. Geochem. Explor. Environ. Anal. 5, 51–57.
Schubert, M., Morgenstern, P., Wennrich, R., Freyer, K., Weiß, H., 2003. The Weathering
behavious of heavy metals in ore processing residues (Mansfeld region, Germany).
Mine Water Environ. 22, 2–6.
Schubert, M., Wennrich, R., Weiß, H., Schreck, P., Zeller, T., Otto, H.H., Wolfram, H.,
2005. Formation of heavy metal bearing phases at a spring affected by the weathering
of ore processing residues. Eur. J. Mineral. 17, 119–128.
Seifert, A.G., Trumbore, S., Xu, X., Zhang, D., Kothe, E., Gleixner, G., 2011. Variable
effects of labile carbon on the carbon use of different microbial groups in black slate
degradation. Geochim. Cosmochim. Acta 75, 2557–2570.
Sethi, P., Schieber, J., 1998. Economica aspects of shales, an overview. In: Schieber, J.
(Ed.), Shales and mudstones, Vol. 2: Petrography, petrophysics, geochemistry and
economic geology. Stuttgart, Schweizerbart'sche Verlagsbuchhandlung, p. 237–253.
Shi, C., Caoa, J., Hua, K., Biana, L., Yao, S., Zhou, J., Han, S., 2014. New understandings
of Ni-Mo mineralization in early Cambrian black shales of South China: constraints
from variations in organic matter in metallic and non-metallic intervals. Ore Geol.
Rev. 59, 73–82.
Shin, I., Kim, S., Shin, D., 2016. Mineralogy and sulfur isotope compositions of the ur-
aniferous black slates in the Ogcheon Metamorphic Belt, South Korea. J. Geochem.
Explor. 169, 1–12.
Shpirt, M.Y., Punanova, S.A., Strizhakova, Y.A., 2007. Trace elements in black and oil
shales. Soil Fuel Chem. 41, 119–127.
Slack, J.F., Selby, D., Dumoulin, J.A., 2015. Hydrothermal, biogenic, and seawater
components in metalliferous Black Shales of the Brooks range, Alaska: synsedimen-
tary metal enrichment in a carbonate ramp setting. Econ. Geol. 110, 653–675.
Sracek, O., Mihaljevič, M., Kříbek, B., Majer, V., Veselovský, F., 2010. Geochemistry and
mineralogy of Cu and Co in mine tailings at the Copperbelt, Zambia. J. Afr. Earth Sci.
57, 14–30.
Stasiuk, R., Włodarczyk, A., Karcz, P., Janas, M., Skłodowska, A., Matlakowska, R., 2017.
Bacterial weathering of fossil organic matter and organic carbon mobilization from
subterrestrial Kupferschiefer black shale: long-term laboratory studies. Environ.
Microbiol. Rep. 9 (4), 459–466. https://doi.org/10.1111/1758-2229.12559.
Steiner, M., Wallis, E., Erdtmann, B.D., Zhao, Y.L., Yang, R.D., 2001. Submarine-hydro-
thermal exhalative ore layers in black shales from South China and associated fossils:
insights into a Lower Cambrian facies and bio-evolution. Palaeogeogr.
Palaeoclimatol. Palaeoecol. 169, 165–191.
Sun, Y.Z., 1999. Distribution of selected elements and PAH in freshly deposited sediments
of waste water streams from Lubin district, Southwest Poland. Environ. Geochem.
Health 21, 141–155.
Sun, Y.Z., Leischner, A., Göbbels, F.J., 2000. Residuary toxic elements and PAHs in se-
diments of the Zbiornik Gilow Tailings Pond and Zimmica Stream from Lubin District,
Southwest Poland. Environ. l Geochem Health 22, 249–261.
Swanson, V.E., 1961. Geology and geochemistry of uranium in marine black shales – a
review. U.S. Geol. Surv. Prof. Pap. 356-C, 67–112.
Talvivaara Mining Company, 2012. Talvivaara Mining Company Plc Home Page. http://
hugin.info/136227/R/1660955/538005.pdf accessed 9.3.2018.
Tang, X., Zhang, J., Liu, Y., Yang, C., Chen, Q., Dang, W., Zhao, P., 2018. Geochemistry of
organic matter and elements of black shale during weathering in Northern Guizhou,
Southwestern China: their mobilization and inter-connection. Chem. der Erde 78 (1),
140–151. https://doi.org/10.1016/j.chemer.2017.08.002.
Tarvainen, T., Backman, B., Hatakka, T., Loukola-Ruskeeniemi, K., Salminen, R.,
Savolainen, H., 2013. Geokemiallinen ympäristövaikutuskartta – Geokemisk
miljökonsekvens-karta 1:1000000: Mustaliuskeet pohjavesialueilla – Svartskiffer i
grundvattenområden. Geological Survey of Finland. Erikoiskartat – Special Maps,
Vol. 87. ISBN: 978-952-217-235-8.
Terrafame Oy, 2017. Selvitys pohjavesien pilaantuneisuudesta ja puhdistustarpeesta sekä
primääriliuotusalueen maaperään kohdistuvista päästöistä (Report on groundwater
pollution, cleanup needs and soil pollution derived from primary leaching stacks).
Report 101003524-001, pp. 89. (In Finnish)http://www.ymparisto.fi/download/
noname/%7B65C2CE13-0ABE-483D-85E7-7F0FF4452A04%7D/130067 (Accessed
21.3.2018).
Terrafame Oy, 2018. Terrafamen kaivoksen velvoitetarkkailu 2017 osa VIII: Pohjavedet.
(Required monitoring of the Terrafame mine 2017 Part VIII: groundwaters), pp. 56.
(in Finnish)https://www.terrafame.fi/ymparisto/ymparisto-vesien-hallinta/
ymparistotarkkailuraportit.html (Accessed 6.7.2018).
Tourtelot, H.A., 1979. Black shale – its deposition and diagenesis. Clays Clay Min. 27 (5),
313–321.
Tuovinen, H., Pelkonen, M., Lempinen, J., Pohjolainen, E., Read, D., Solatie, D., Lehto, J.,
2018. Behaviour of Metals during Bioheap Leaching at the Talvivaara Mine, Finland.
Geosciences 8, 66. https://doi.org/10.3390/geosciences8020066.
Tuovinen, H., Pohjolainen, E., Lempinen, J., Vesterbacka, D., Read, D., Solatie, D., Lehto,
J., 2016. Behaviour of radionuclides during microbially-induced mining of nickel at
Talvivaara, Eastern Finland. J. Environ. Radioact. 151, 105–113.
Tuttle, M.L.W., Breit, G.N., Goldhaber, M.B., 2009. Weathering of the New Albany Shale,
Kentucky: II. Redistribution of minor and trace elements. Appl. Geochem. 24,
1565–1578.
Vaughan, D.J., Sweeney, M.A., Friedrich, G., Diedel, R., Haranczyk, C., 1989. The
Kupferschiefer: an overview with an appraisal of the different types of mineralization.
Econ. Geol. 84, 1003–1027.
Wallis, E., 2007. Die Klima- und Umweltgeschichte der südchinesischen Yangtze
Plattform im Neoproterozoikum und frühen Kambrium: Hydrothermal aktive, Dichte-
stratifizierte Epikontinentalbecken, der Schlüssel zum Verständnis der Kambrischen
Explosion. Ph.D. thesis. Technical University of Berlin (227 p).
Wennrich, R., Mattusch, J., Morgenstern, P., Freyer, K., Treutler, H.C., Stärk, H.J.,
Brüggemann, L., Paschke, A., Daus, B., Weiss, H., 2004. Characterization of sediments
in an abandoned mining area; a case study of Mansfeld region. Germany. Environ.
Geol. 45, 818–833.
Wignall, P.B., 1994. Black Shales. Oxford University Press, New York 127 p.
Włodarczyk, A., Szymanska, A., Skłodowska, A., Matlakowska, A., 2016. Determination of
factors responsible for the bioweathering of copper minerals from organic-rich
copper-bearing Kupferschiefer black shale. Chemosphere 148, 416–425.
Woo, N.C., Choi, M.J., Lee, K.S., 2002. Assessment of groundwater quality and con-
tamination from Uranium-Bearing Black Shale in Goesan–Boeun areas. Korea.
Environ. Geochem. Health 24, 261–273.
Xu, L.G., Lehmann, B., Mao, J.W., Nägler, T.F., Neubert, N., Böttcher, M.E., Escher, P.,
2012. Mo isotope and trace element patterns of Lower Cambrian black shales in South
China: Multi-proxy constraints on the paleoenvironment. Chem. Geol. 318–319,
45–59.
Xu, J., Peng, B., Yu, C., Yang, G., Tang, X., Tan, C., Xie, S., Tu, X., Bao, Z., Quan, M., Xiao,
M., 2013. Geochemistry of soils derived from black shales in the Ganziping mine
area, western Hunan, China. Environ. Earth Sci. 70, 175–190.
Yi, J.M., Chon, H.T., Park, M., 2003. Migration and enrichment of Arsenic in the rock-soil-
crop plant system in areas covered with Black Shale, Korea. Sci. World J. 3, 194–205.
ISSN 1537-744X. https://doi.org/10.1100/tsw.2003.19.
Young, S., Loukola-Ruskeeniemi, K., Pratt, L., 2013. Reactions of hydrothermal solutions
with organic matter in Palaeoproterozoic black shales at Talvivaara, Finland: evi-
dence from multiple sulfur isotopes. Earth Planet. Sci. Lett. 367, 1–14.
Yu, C., Lavergren, U., Peltola, P., Drake, H., Bergbäck, B., Åström, M.E., 2014. Retention
and transport of arsenic, uranium and nickel in a black shale setting revealed by a
long-term humidity cell test and sequential chemical extractions. Chem. Geol. 363,
134–144.
Yu, C., Peng, B., Peltola, P., Tang, X., Xie, S., 2012. Effect of weathering on abundance
and release of potentially toxic elements in soils developed on Lower Cambrian black
shales, P. R. China. Environ. Geochem. Health 34, 375–390.
Zhao, Q., Wang, Y., Cao, Y., Chen, A., Ren, M., Ge, Y., Yu, Z., Wan, S., Hu, A., Bo, Q.,
Ruan, L., Chen, H., Qin, S., Chen, W., Hu, C., Tao, F., Xu, D., Xu, J., Wen, L., Li, L.,
2014. Potential health risks of heavy metals in cultivated topsoil and grain, including
correlations with human primary liver, lung and gastric cancer, in Anhui province,
eastern China. Sci. Total Environ. 470–471, 340–347.
Zhu, M.-Y., Zhang, J.-M., Steiner, M., Yang, A., Li, G.-X., Erdtmann, B.-D., 2003. Sinian-
Cambrian stratigraphic framework for shallow to deep-water environments of the
Yangtze Platform: an integrated approach. Prog. Nat. Sci. 13, 946–950.
Zou, C., 2013. Unconventional Petroleum Geology. Elsevier978-0-12-397162-3 384 p.
A. Parviainen and K. Loukola-Ruskeeniemi Earth-Science Reviews 192 (2019) 65–90
90