ArticlePDF Available

Review of uranium in the Jordanian phosphorites: distribution, genesis and industry

Authors:
JJEES Volume 4 (Special Publication, Number 2), Jan. 2011
ISSN 1995-6681
Pages 35- 45
Jordan Journal of Earth and Environmental Sciences
Review of Uranium in the Jordanian Phosphorites: Distribution,
Genesis and Industry
Abdulkader M. Abed*
Present address: Researcher, Deanship of Scientific Research
Jordan University of Science and Technology, Irbid
Abstract
Phosphorites are well-known, world wide, to accommodate a substantial amount of U relative to other sedimentary rocks.
This is due to a great extent to the crystal structure of apatite (carbonate flour apatite, francolite) where U substitutes Ca.
Phosphorites form in shallow marine environments through the accumulation of planktonic fossils debris on the sea floor in
areas characterized by upwelling currents. Once the organic-rich sediments are buried below the sea floor and the organic
matter is decomposed, PO4 is released to the interstitial solutions. These PO4-rich solutions either precipitate phosphorites
directly (authigenic) or the solutions react with pre existing sediments and transform them to phosphorites (early
diagenentic). The Jordanian deposits are dominantly of the former type. Consequently, phosphorites are also well known to
be associated with high percentage of organic matter which makes them a good source rock for petroleum.
In Jordan, phosphorites are wide spread from its extreme NW to the SE. Uranium contents are not the same in each locality.
It is not also the same in each bed in the same locality. In Al-Kora Basin, NW Jordan, U ranges between 60-379 ppm (parts
per million or g/ton) with an average of 153 ppm. In Ruseifa, just east of Amman, the range is 132-195 ppm and average of
123 ppm. Central Jordan phosphorites (Al-Abiad and Al-Hasa) have a lower U %. The range is 34-190 ppm with an average
of 105 ppm. Eshidiyya Basin has a much less U concentration. The range is 7-125 ppm with an average of 70 ppm.
However, a recent work on the uppermost phosphorite horizon, the A0, in Eshidiyya Basin proved the presence of a 3 m
thick bed with 242 ppm U. Furthermore, certain phosphorite horizons are known to have much more than the average U in
that locality; e.g. in Al-Kora Basin, the average U is 153 ppm while certain horizons have up to 379 ppm U, the uppermost
bed in Ruseifa has up to 195 ppm while the average is 123 ppm, and in Eshidiyya the A0 has 242 ppm while the average of
the lower horizon (A1-3) average is 70 ppm. Obviously, if the U is to be extracted from the phosphorite, the horizons with
higher concentrations of U should be explored and used.
Uranium sticks to francolite in its behavior from precipitation in the marine environment, through mining and beneficiation,
and the fertilizer industry. 1) There is a significant high correlation coefficient between CaO and P2O5 in the hundreds of
samples analyzed. 2) Fines (clays) produced when washing the ore have very little U and consequently, washing water for
the last 45 years has not contaminated groundwater with U in central Jordan. 3) Uranium is concentrated in the phosphoric
acid then into diammonium phosphate (DAP) in the fertilizer industry and not in the phosphogypsum.
© 2012 Jordan Journal of Earth and Environmental Sciences. All rights reserved
Keywords: Phosphorite; Upper Cretaceous; Uranium; Francolite; Jordan; Diammonium Phosphate; Phosphogypsum .
* Corresponding author. e-mail: aabed@ju.edu.jo.
1. Introduction
Phosphorites are the carrier of uranium, thus the study
of them is a prerequisite to the study of U. Phosphorites
are widespread in Jordan and cover a relatively large area
of the country from the extreme northwest to its south.
Low and high grade phosphorites are supposed to have
been deposited throughout the country, but subsequent
uplift and erosion, from the Oligocene onwards, had
removed it from various parts of the country; e.g. Ajlun
dome the western parts of the western mountain range, and
the extreme south (Bender, 1974; Powell, 1989, Abed,
2000). Furthermore, high grade phosphorites are also
present in the subsurface of the eastern desert of Jordan;
e.g. Zgaimat Al-Hasat (Abed and Amireh, 1999).
High grade phosphorites were discovered in Jordan in
1908 during the construction of the Hijaz Railway. In
1938, small scale mining from Ruseifa started and was
exported by mules through Haifa. The Jordan Phosphate
Mines Company (JPMC) commenced its work by 1953 in
Ruseifa, 1965 in Al-Hasa, 1979 in Al-Abiad, and 1988 in
Eshidiyya (Abed, 2000). Ruseifa mines were closed in
1988 despite the fact that there are several tens of millions
of high grade phosphorites south of Amman-Zarqa
Highway (Zarqa B area) (Abed, 1989). Central Jordan
© 2012 Jordan Journal of Earth and Environmental Sciences . All rights reserved - Volume 4, (Special Publication, Number 2)(ISSN 1995-6681)
36
phosphorites (Al-Hasa and Al-Abiad) are near depletion);
consequently, Eshidiyya is going to be the centre for
phosphorite mining and industry. However, several
hundred millions of high grade phosphorites are present in
Al-Kora Basin NW Jordan (Mikbel and Abed, 1985; Abed
and Al-Agha, 1989), likewise, huge, but not estimated,
subsurface deposits in the SE desert continuing into Saudi
Arabia are present (Abed and Amireh, 1999). In short,
there seems to be no shortage of high grade phosphorites
in Jordan even for the distant future.
On the other hand, uranium is well known to be
several folds enriched in phosphorites compared with other
sedimentary rocks such as sandstones, carbonates and
shales. Average U in phosphorite is 120 ppm compared
with 3.5, 0.5 and 2.2 ppm in shale, sandstone and
carbonate respectively; an enrichment factor of around 30
relative to shale (Altschuler, 1980; Slansky, 1986;
McArthur et al., 1987; Krauskopf and Bird, 1995).This is
basically due to the nature of the crystal structure of
francolite (carbonate flour apatite) where U is substituting
for calcium, both having a close radius (Ca+2= 0.99Å and
U= 0.97Å (e.g. Nathan, 1984; Mcclellan and Van
Kauwenberg, 1990). The enrichment of U in phosphorites
may also be due to the association of the latter with
abundant organic matter in the depositional environment
of the phosphorites (Barnett, 1990). Uranium, in the
Jordanian phosphorite, was studied by Coopens et al.,
(1977), Khalid and Abed (1982), Abed and Khalid (1985)
and Sadaqah et al., 2005 amongst others).
The aim of this paper is to discuss a) the distribution of
U in the various localities and the phosphorite horizons
within each locality; b) the genesis of phosphorite
formations in Jordan; and c) the behavior of U throughout
the processes of mining, upgrading and fertilizers industry.
2. GEOLOGICAL SETTING
The phosphorites of Jordan form part of the
Cretaceous-Eocene phosphorite episode which deposited
one of the most extensive deposits in the world in the
Eastern Mediterranean (Saudi Arabia, Iraq, Syria, Jordan
and Palestine) and North Africa (Egypt, Tunisia, Algeria,
Morocco, Mauritania, Senegal, …), and parts of the
Caribbean and NE S. America. It is second to the Miocene
North American episode (Sheldon 1981; Riggs, et al.,
1985; Notholt et al. 1989; Abed, 1994; Follmi 1996).
The phosphorites of Jordan are present within Al-Hisa
Phosphorite Formation (AHP). Fig. 1 shows the location of
high grade phosphorites in Jordan. The age of the AHP is
most probably early Maastrichtian, Uppermost Cretaceous
(Quennell, 1951; Hamam, 1977; Cappetta et al., 1996). In
general, the AHP consists of phosphorites, bedded chert,
limestones, oyster buildups, organic-rich marl (oil shale)
and other rock types (Powell, 1989; Abed, 2000).
However, the lithology and thickness of the AHP are
variable from one locality to the other. The total thickness
of the AHP is around 10 m in NW Jordan, 30 m in
Ruseifa, 40-60 m in central Jordan, 10-17 m in Eshidiyya
and 5-6 m in Zgaimat Al-Hasat in the SE desert ((Powell,
1989; Abed, 2000).
Fig. 1 Location map of the grade phosphorite deposits in Jordan.
The AHP consists of three formal members, from older
to younger Sulatani, Bahiyya and Qatrana. The three
members are well displayed throughout central Jordan
where the terminology was first used by El-Hiyari (1985)
(Table 1). In central Jordan, the Sultani Member consists
of alternating limestones, bedded chert and minor
phosphorites. The Bahiyya Member consists of oyster
banks or buildups up 30 m in thickness that are made of
oyster fragment in clinoforms dipping general to the SE
(Abed and Sadaqah, 1998). The Qatrana Member is the
host of the high grade phosphorites. The grade
phosphorites are friable with little calcareous cement. They
are present as lenses (small basins) with a diameter
ranging up to few kilometers and a thickness up 13 m.
Table 1. Nomenclature of the Upper Cretaceous rock units in
Jordan (Al-Hiyari, 1985)
In Eshidiyya Basin, the three formal members are
present only in the northern parts of the basin. The
Bahiyya coquina thins gradually towards the SE until it
disappears completely. The Sultani Member or the lower
© 2012 Jordan Journal of Earth and Environmental Sciences . All rights reserved - Volume 4, (Special Publication, Number 2)(ISSN 1995-6681) 37
member is hosting the main (high) grade phosphorite
deposits. It is divided by the miners into three producing
phosphate horizons from top to bottom; A1, A2 and A3,
separated by non phosphorite strata like; chert, porcelanite,
and minor beds of marl, limestone and dolomite.
(Sofremines, 1987; Abed et al., 2007). The A3, at the base,
is rich in quartz sand, while A2 is a friable, high grade
phosphorite. The A1 is a low grade indurated by calcite
cement (Fig. 2). The Qatrana Member or the upper
member is designated A0 and consists of up to 3 m of
friable high grade phosphorite in the north thinning to the
SE to few centimeters only.
Fig. 2 A columnar section describing the lithology of Eshidiyya phosphorites. A detailed section , right, shows the lithology of the A0
deposits in northern Eshidiyya.
Fig. 3 is a columnar section in central Jordan. The SE
desert deposits were discovered by Abed and Amireh
(1999). They are deep seated in the subsurface, but crop
out as windows at the cores of anticlines such as that of
Zgaimat Al-Hasat. Total thickness is 5-6 m consisting of
friable phosphorites with same quartz sand increasing to
the SE (Fig. 4). The age, most probably Turonian, seems
older than other phosphorites of Jordan (Khalifa and Abed,
2011).
Fig. 3 (left) Generalized lithological section in central Jordan.
Fig. 4 (right) Detailed columnar section of the Ajlun Group and the phosphorites of the SE desert at Zgaimat Al-Hasat (Khalifa and Abed,
2011). For legend, see Fig. 2.
© 2012 Jordan Journal of Earth and Environmental Sciences . All rights reserved - Volume 4, (Special Publication, Number 2)(ISSN 1995-6681)
38
The Ruseifa mines were closed in 1988, however, the
deposits consists of four phosphorite horizons designated
from bottom A1, A2, A3, and A4, separated by limestone,
trace fossils and some chert (Fig. 5). Al-Kora deposits,
NW Jordan, were discovered by Mikbel and Abed (1985)
with several million tons of grade phosphorites. The
deposits are up to 10 m thick, the lower part consists of
alternating chert and phosphorite caped by about 3 m thick
high grade, friable phosphorite horizon rich in organic
matter (Abed and Al-Agha, 1989). Fig. 6 shows the
lithology near the village of Tobna. These deposits have
not been mined yet.
Fig. 5 (left) columnar section describing the phosphorites at Ruseifa Basin.
Fig. 6 (right) Columnar section describing the phosphorites at Tobna, Al-Kora Basin, NW Jordan. For legend, see Fig. 2 .
3. Results and Discussion
3.1. Petrography and Mineralogy
The phosphorites of Jordan are granular in nature with
subordinate pristine (primary or not reworked) especially
in NW Jordan deposits. Grain types are phosphate
intraclasts, peloids, coprolites and skeletal fragments (bone
and teeth). The former two types are due mainly to
reworking of originally precipitated phosphate material, or
may partially be due to the nature of deposition between
the pore spaces of pre existing sediments (Riggs, 1979,
1980; Burnett, 1990; Abed and Fakhouri, 1996; Abed et
al., 2007 amongst many others).
Phosphate particles or grains, in marine phosphorites,
consist of the mineral carbonate flour apatite (francolite), a
variety of the mineral apatite. The phosphorites of Jordan
have the following chemical formula of francolite [Ca 9.86
Mg .005 Na .14] [PO4 4.93 CO3 1.07 F 2.06] (Abed and
Fakhouri, 1996). The crystal structure of francolite is
rather open, thus enhancing substitution (e.g. McClellan
and Van Kauwenberg, 1990). Uranium is one of many
elements that substitutes for Ca. Consequently, U is
present within the crystal structure of francolite and
behaves more similar to Ca than to P, simply because the
ionic radii are very close to each other. To shed more light
on this relationship, U ppm is correlated with each of
CaO%, P2O5%, U/ P2O5 and U/CaO in 184 samples from
the Jordanian phosphorites (Fig. 7). It is clear that a better
correlation is present between U and Ca than between U
and P, meaning that U is not only present within the
structure of francolite but it occupies same positions of Ca
in that structure.
Fig. 7 Correlation between U and A) P2O5%, B) CaO%, C)
U/P2O5 ratio, and D) U/CaO ratio. Solid line is the best fit while
dotted lines are for the 95% confidence. Number of samples is
184.
The yellowish-greenish U minerals found on the
surface of joints and fractures are due to leaching by
groundwater of the original U; i.e. secondary in origin.
Consequently, this type occurrence is very limited in
abundance, and thus, of minor importance and of no
economic value for U exploration.
© 2012 Jordan Journal of Earth and Environmental Sciences . All rights reserved - Volume 4, (Special Publication, Number 2)(ISSN 1995-6681) 39
3.2. Distribution of Uranium
In this title, the distribution of U in the various
localities in Jordan is discussed based on data acquired by
the author, his coworkers and students since 1980.
3.2.1. Eshidiyya Deposits
Phosphorites are present in two members, the lower
member (A1, A2 and A3 horizons) who accommodates the
main phosphorite reserves, and the upper member (A0
horizon) with limited reserves (Fig. 2). Tables 2 and 3
display the contents of U in both upper and lower members
respectively.
The U average of 60 samples analyzed from the lower
member is 51 ppm with a range 1 -175 ppm. If the samples
with less than 11 % P2O5 % are removed, because they are
not typical phosphorites, then the average U will be 70
ppm. The higher U concentrations up to 175 ppm can be
found in the calcareous, high grade A2 horizon. The A3
has the lowest U content, partially, because of the high
dilution with silica sand which can be present up to 50 %
in certain samples. The A1 is slightly higher than the A3
(Khaled and Abed, 1982). Upgrading the lower member
will certainly enhance the contents of U in these
phosphorites.
Table 2.Uranium, P2O5 and CaO contents and other ratios in the lower phosphorite member in Eshidiyya Basin (Khaled and Abed, 1982;
Abed and Khaled, 1985).
U ppm
P2O5% CaO%
U/
P2O5 CaO/
P2O5 U/
CaO
Uppm
P2O5
% CaO%
U/
P2O5 CaO/
P2O5 U/
CaO
1 45 23.96 41.23 1.88 1.72 1.09 31 63 5.83 39.53 10.81 6.78 1.59
2 96 21.93 37 3.14 1.69 2.59 32 5 9.88 32.05 0.51 3.24 0.16
3 143 33.13 53.12 4.23 1.60 2.69 33 5 10.27 39.35 0.49 3.83 0.13
4 146 23.98 53.3 4.43 2.22 2.74 34 5 6.55 9.17 0.76 1.40 0.55
5 44 27.22 40.4 1.62 1.48 1.09 35 50 30.88 48.04 1.62 1.56 1.04
6 44 23.16 40.78 1.75 1.76 1.08 36 35 31.74 45.96 1.10 1.45 0.76
7 45 25.68 39.04 1.75 1.52 1.15 37 50 22 38.47 2.27 1.75 1.30
8 45 22.18 36.49 2.02 1.65 1.23 38 75 32.84 49.08 2.28 1.49 1.53
9 44 27.63 44.1 1.39 1.60 1.00 39 85 33.96 49.94 2.50 1.47 1.70
10 141 32.63 48.8 4.32 1.50 2.89 40 100 33.76 49.93 2.96 1.48 2.00
11 111 25.33 44.3 4.38 1.75 2.51 41 88 26.6 43.65 3.31 1.64 2.02
12 72 19.99 42.91 3.6 2.15 1.68 42 85 18.27 27.87 4.65 1.53 3.05
13 45 26.14 42.06 1.72 1.61 1.07 43 100 20.7 33.79 4.83 1.63 2.96
14 44 33.82 30.25 1.30 0.89 1.45 44 46 28.25 50.63 1.63 1.79 0.91
15 18 7.45 24.94 2.42 3.35 0.72 45 45 20.05 44.57 2.24 2.22 1.01
16 7 4.1 7.26 1.71 1.77 0.96 46 63 22.92 40.46 2.75 1.77 1.56
17 16 7.7 25.96 2.08 3.37 0.62 47 8 33.13 44.52 0.24 1.34 0.18
18 29 24.15 37.82 1.20 1.57 0.77 48 75 29.91 48.33 2.51 1.62 1.55
19 43 35.91 55.09 1.20 1.53 0.78 49 125 29.7 39.63 4.21 1.33 3.15
20 30 24.21 38.22 1.24 1.58 0.78 50 1 0.83 2.82 1.20 3.40 0.35
21 35 27.78 43.86 1.26 1.58 0.80 51 2 10.89 18.19 0.18 1.67 0.11
22 51 33.66 54.18 1.52 1.61 0.94 52 1 1.5 3.14 0.67 2.09 0.32
23 69 30.34 47.19 2.27 1.56 1.46 53 5 11.38 17.59 0.44 1.55 0.28
24 34 10.68 16.8 3.18 1.57 2.02 54 2 8.8 18.11 0.23 2.06 0.11
25 125 26.71 55.35 4.68 2.07 2.26 55 25 10.1 16.93 2.48 1.68 1.48
26 89 17.07 54.58 5.21 3.20 1.63 65 3 13 20.36 0.23 1.57 0.15
27 175 27.32 52.24 6.41 1.91 3.35 57 1 11.07 16.56 0.09 1.50 0.06
28 34 23 40.84 1.48 1.78 0.83 58 1 7.78 30.47 0.13 3.92 0.03
29 46 33.01 50.55 1.39 1.53 0.91 59 1 2 22.79 0.50 11.40 0.04
30 34 23.94 36.28 1.42 1.52 0.94 60 1 1.66 25.56 0.60 15.40 0.04
Av 51 20.83 36.71 2.24 1.76 1.39 Av. of >11% P2O5 = 70 ppm
On the other hand, the upper phosphorite member (A0)
is much more enriched in U. Table 3 shows the analysis of
28 samples of the A0 with an average U of 133 ppm and a
range of 5-242 ppm. If the samples of less than 10% P2O5
% are removed, then the average U becomes 172 ppm and
the range 40-242 ppm, The data clearly indicates that U is
more than twice that of the lower member (Abed and
Sadaqah, 2011 in press). The A0 consists of two
phosphorite beds up to 3 m thick. The phosphorite beds are
calcareous with slight calcite cement, ammonites and trace
© 2012 Jordan Journal of Earth and Environmental Sciences . All rights reserved - Volume 4, (Special Publication, Number 2)(ISSN 1995-6681)
40
fossils. The A0 overlies the coquina member (middle
member) with subaerial unconformity. Abed and Sadaqah
(2011 in press ) concluded that the U enrichment is, at
least partially, due to slight leaching of the carbonates and
phosphorites at the unconformity. That is why the lower
bed of the A0 has more U compared with the upper bed
away from the unconformity.
Table 3.Uranium, P2O5 and CaO contents and other ratios in the upper phosphorite member (A0) in Eshidiyya Basin (Abed and Sadaqah,
2011 in press).
3.2.2. Central Jordan Deposits
Central Jordan phosphorites are represented by Al-Hasa
and Al-Abiad deposits and their surroundings. The
deposits are calcareous, granular, and slightly cemented by
calcite and are present in single isolated lenses. Several
hundred million tons or high grade phosphorites were
mined since 1965, and consequently these deposits are
near depletion. Table 4 shows the chemical results of 34
samples with an average of 105 ppm and a range 60-168
ppm (Abed and Khaled, 1985; Sadaqah et al., 2005; Abed
et al., 2008). However, there are no much deposits for
future U industry.
Table 4.Uranium, P2O5 and CaO contents and other ratios in central Jordan phosphorites, Al-Abiad and Al-Hasa (Abed and Khaled, 1985;
Abed et al., 2008).
© 2011 Jordan Journal of Earth and Environmental Sciences . All rights reserved - Volume 4, (Special Publication, Number 2)(ISSN 1995-6681)
41
However, grade phosphorites are present west of
Dhiban, near the village of Shgaig. The area is populated
and the deposits have not been mentioned by JPMC or
other workers (Abed and Khaled, 1985). Reserves are not
estimated and the U contents of 14 samples is 79 ppm with
a range of 4-118 ppm (Table 5)
Table 5. Uranium, P2O5 and CaO contents and other ratios in the
phosphorites of the Mujib Area (Abed and Khaled, 1985).
U
ppm P2O5% CaO%
U/
P2O5 CaO/
P2O5 U/
CaO
1 92 22.94 52.41 4.01 2.28 1.76
2 22 16.85 44.7 1.31 2.65 0.49
3 15 14.8 47.29 1.01 3.20 0.32
4 4 16.81 47.4 0.24 2.82 0.08
5 103 13.16 51.13 7.83 3.89 2.01
6 110 18.95 52.47 5.80 2.77 2.10
7 110 23.18 53.79 4.75 2.32 2.04
8 116 29.72 51.8 3.90 1.74 2.24
9 78 26.81 50.24 2.91 1.87 1.55
10 85 23.69 52.18 3.59 2.20 1.63
11 118 30.1 51.3 3.92 1.70 2.30
12 48 17.94 54.64 2.68 3.05 0.88
13 91 24.85 57.31 3.66 2.31 1.59
14 111 22.92 50.17 4.84 2.19 2.21
Av. 79 21.62 51.20 3.60 2.37 1.54
3.2.3. Ruseifa Deposits.
During 1988, Ruseifa mines were closed and the
remaining deposits south of the Amman-Zarqa highway
are becoming more and more urbanized. Thus, the
following lines are of historical value only. The Ruseifa
phosphorites consist of four horizons (A1, A2, A3, and A4
topmost) separated by carbonates and chert beds (Fig. 5).
Table 6 shows the analysis of 17 samples with an average
U of 123 ppm and a range of 57-184. Uranium content
increases upwards and the A4 horizon have the highest
content of more than 180 ppm (Abed and Khaled, 1985).
Table 6. Uranium, P2O5 and CaO contents and other ratios in the
phosphorites of Ruseifa (Abed and Khaled, 1985).
U
ppm P2O5% CaO% U/
P2O5 CaO/
P2O5 U/
CaO
1 123 17.25 52.61 7.02 3.05 2.34
2 121 26.66 55.26 4.55 2.07 2.19
3 113 13.26 52.7 8.52 3.97 2.14
4 56 24.67 42.58 2.27 1.73 1.32
5 86 24.59 42.97 3.5 1.75 2.00
6 57 25.31 43.62 2.25 1.72 1.31
7 183 29.79 52.65 6.11 1.77 3.48
8 184 28.94 50.33 6.53 1.74 3.66
9 127 26.31 46.71 4.82 1.78 2.72
10 73 21.69 43.68 3.37 2.01 1.67
11 119 20.96 53.17 5.68 2.54 2.24
12 117 17.31 49.45 6.76 2.86 2.37
13 161 30.5 53.8 5.28 1.76 2.99
14 162 31.11 53.75 5.21 1.73 3.01
15 181 21.75 53.82 8.32 2.47 3.36
16 117 16.54 54.2 7.07 3.28 2.16
17 117 17.93 52.1 6.53 2.91 2.25
Av. 123 23.21 50.20 5.52 2.16 2.46
3.2.4. Al-Kora Deposits, NW Jordan
The NW Jordan phosphorites are promising future
deposits for phosphorites and uranium. There are, at least,
several hundred million tons of high grade phosphorites
with high U content relative to other deposits discussed
earlier. However, the high population (many villages and
towns) and the green nature of the area (forests and
agriculture) might have been behind the decision not to
mine the deposits. Also, the deposits are present within a
folded belt which makes open pit mining rather difficult if
not impossible in certain localities (Mikbel and Abed
1985).
Table 7 shows the analysis of18 samples with an
average of 153 ppm U and a range of 59-379 ppm U (Abed
and Khaled, 1985; Sadaqah, 2000, Sadaqah et al., 2005).
The highest U content is around 6 times more than that of
Eshidiyya and might be of economic nature if extracted as
a byproduct through the fertilizers industry. Detailed field
and laboratory work can pinpoint the localities with the
high U content.
Table 7.Uranium, P2O5 and CaO contents and other ratios in the
phosphorites of Al-Kora Basin, NW Jordan.(Abed and Khaled,
1985; Sadaqah et al., 2005).
U
ppm P2O5% CaO%
U/
P2O5 CaO/
P2O5 U/
CaO
1 88 24.09 40.52 3.65 1.68 2.17
2 117 30.23 44.01 3.87 1.46 2.66
3 186 32.7 52.4 5.69 1.60 3.55
4 129 18 53.79 7.16 2.99 2.40
5 120 14.22 50.44 8.44 3.55 2.38
6 132 30.18 46.62 4.37 1.54 2.83
7 127 24.37 54.9 5.21 2.25 2.31
8 81 21.08 33.18 3.84 1.57 2.44
9 75 31.5 52.4 2.38 1.66 1.43
10 238 34.16 52.98 6.97 1.55 4.49
11 88 26.25 52.78 3.35 2.01 1.67
12 89 24.05 53.09 3.70 2.21 1.68
13 301 20.92 50.06 14.39 2.39 6.01
14 343 18.95 51.95 18.10 2.74 6.60
15 109 24.06 52.52 4.53 2.18 2.08
16 59 10.13 52.39 5.82 5.17 1.13
17 379 27.75 53.08 13.66 1.91 7.14
18 92 22 52.23 4.18 2.37 1.76
Av 153 24.15 49.96 6.63 2.07 3.06
3.2.5. Area Comparison
Absolute U content might not be the best way of
comparing its abundance in the various localities. For this
reason, the U/P2O5 ratio is used, meaning the amount of U
in ppm present for each 1% P2O5. The average U/P2O5
ratio for the samples in each locality (not the ratio of
averages) is shown in Table 8. It is clear from Table 8 that
the ratio increases northwards. It is 2.24 in Eshidiyya; i.e.
there is 2.24 ppm U in Eshidiyya deposits for each 1%
P2O5. It increases to 6.63 in NW Jordan, which clearly
shows that the NW Jordan phosphorites are three times
© 2011 Jordan Journal of Earth and Environmental Sciences . All rights reserved - Volume 4, (Special Publication, Number 2)(ISSN 1995-6681)
42
more enriched in U compared with Eshidiyya. The other
localities are intermediate between Eshidiyya and NW
Jordan. The possible reasons for this northwards increasing
trend are discussed further below.
Table 8The U/P2O5 sample averages in the phosphorite deposits of
Jordan, n = number of samples averaged.
U
ppm P2O5
% CaO
% U/
P2O5 n
Eshidiyya 51 20.83 36.7
1 2.24 60
Central
Jordan 105 27.46 48.5
2 3.81 44
Mujib 79 21.62 51.2 3.6 13
Ruseifa 123 23.21 50.2 5.52 16
NW Jordan 153 24.15 49.9
6 6.63 18
4. Behavior of U During Mining and Upgrading.
Mining and upgrading are the not discussed here
because they are not the subject of this paper. However,
mining and upgrading involve the removal of the
overburden, obtaining the phosphorite ore, crushing the
ore to liberate the suitable grain size, sieving to get a
roughly sand-size product, washing with fresh water to get
red of the fines especially clays, and finally drying.
Samples from all these stages were analyzed for their U
and other elements and the following conclusion was
reached (Al-Huwaiti et al., 2005; Abed et al., 2008).
Uranium behaves similar to the behavior of P and Ca (Fig.
8a); i.e. U stays fixed in the crystal structure of francolite,
as discussed in the mineralogy section.
Environmentally, the discharged washing water has not
contaminated the groundwater in Al-Hasa and Al-Abiad
mines area despite the fact that the washing process has
been ongoing since 1965 and 1979 respectively.
Groundwater samples from wells and springs in central
Jordan have a U content less than 2 ppb (parts per billion) (
Jiries et al., 2004; Abed et al., 2008). See Fig.8b.
Fig. 8 shows the behavior of U during mining, crushing, sieving
and washing (8a upper) and 8b shows the concentration in ppb in
the groundwater in central Jordan.
5. Behavior of U in the Fertilizers Industry.
In the fertilizers industry, the upgraded phosphorite ore
is reacted with sulphuric acid (H2SO4) to produce
phosphoric acid (H3PO4) and phosphogypsum. The crystal
structure of francolite is destroyed through this reaction, its
PO4 forms phosphoric acid, and the Ca of francolite forms
the phosphogypsum. Analysis of the two products, the
acid, and gypsum, clearly shows that U follows the PO4
group and is concentrated in the phosphoric acid. The
phosphogypsum has no more than 2 ppm U. The
phosphoric acid is then transferred to the fertilizer,
diammonium phosphate (DAP), and U is found to be
concentrated in the. DAP. Through these reactions, U is
found to be concentrated in the phosphoric acid and the
DAP by a factor of 1.5 (Fig. 9). For more details, see Abed
et al., (2008).
Fig. 9 Behavior of U in the fertilizers industry. Note that the U
and P2O5 are enriched by a factor of 1.5 in the DAP fertilizer and
the U in phosphogypsum is around 2 ppm.
6. Phosphogenesis and the Enrichment of U: Discussion
Why phosphorites have more U compared with other
sedimentary rocks? Why U contents increases northwards
in the Jordanian phosphorites? Following is an attempt to
answer these and others related to phosphogenesis.
Phosphorite, bedded chert, porcelanite and organic-rich
sediments, as an association, are known to deposit under
upwelling currents regimes (Sheldon, 1987; Iijima, et al.,
1994 and the papers within). This is well documented in
the recent and subrecent Earth history in shallow
continental shelf where upwelling currents are intense and
still ongoing; e.g. the coasts of SW Africa (Birch, 1980),
the western coast of South America up till California in the
north (Froelich et al., 1988; Burnett, 1990; Kolodny and
Garrison, 1994), and in the southeastern United States
where the Gulf Stream causes upwelling (Riggs et al.,
1998). Thus, the association of these facies is usually
taken as due to upwelling in ancient phosphorite deposits
(Baturin, 1982; Abed and Amireh, 1983; Alomogi-Labin
et al., 1993; Glenn et al., 1994; Follmi, 1996). However,
some authors had explained the formation of ancient
phosphorites without the need of upwelling (e.g. Heggie et
al., 1990; Glenn and Arthur, 1990)
Upwelling currents spread deep, cold marine water on
the surface of the relatively shallow continental shelves
adjacent to the continents (Fig. 10). Deep, cold water,
1000 m or more, is usually rich in nutrients, most
important of which are Si and P which are the basic food
© 2011 Jordan Journal of Earth and Environmental Sciences . All rights reserved - Volume 4, (Special Publication, Number 2)(ISSN 1995-6681)
43
for the phytoplanktons (diatoms and dinoflagellates); the
lowest step in marine food chains which inhabit the photic
zone or the upper 100-200 m of the sea water column.
Under such conditions, bioproductivity of the marine food
chain is highly increased. Rate of death of these organisms
is consequently increased to produce an oxygen minimum
zone (OMZ) extending few hundred metres below the
photic zone and might reach the shelf floor. The formation
of the OMZ reduces the amount of the oxidized soft tissues
of the descending organisms, thus giving more chances to
the soft tissues (organic matter, OM) to joint the sediments
accumulating at the shelf floor. A higher rate of
sedimentation will ensure a higher rate of burial for the
OM to escape long exposure and oxidation. Consequently,
a sediment- rich OM is formed (Baturin, 1982; Salansky,
1986; Glenn et al., 1994; Lucas and Prevot-Lucas, 1995).
Fig. 10 A schematic model for phosphorite deposition under an
upwelling regime.
Decomposition of the sediment organic matter below
the water/sediment interface by bacteria and fungi liberates
phosphorous to the interstitial pore fluids. Certain minerals
form from these fluids like calcite, dolomite and
palygorskite before the formation of apatite. This means
that P, most probably as PO4, is concentrated many folds
relative to sea water before its deposition (Riggs et al.,
1985). Finally, apatite is either precipitated directly from
these interstitial fluids (authigenic origin) or the fluids
react with pre existing sediments and transform them into
apatite (diagenetic origin) (Price and Calvert, 1978; Birch
et al., 1983; Froelich et al., 1983).
Jordanian phosphorites are postulated to have deposited
under upwelling cold, deep water from the Tethys Ocean
in the north onto its shallow epeiric shelf where Jordan
was situated during the upper most Cretaceous. This is
evident from the presence of phosphorite, bedded chert,
porcelanite (biogenic silica deposits), oil shale and a
pronounced negative cerium anomaly indicative of deep
oceanic water (Abed and Abu Murry, 1997). Furthermore,
Jordanian phosphorites are dominantly authigenic,
precipitated from the pore fluid solutions as phosphate
mud, which was then reworked into phosphate pellets and
intraclasts (Al-Agha, 1985; Abed and Al-Agha, 1989).
Recent pelletal phosphorites off the SW Africa and
Peru margins are carbonate fluorapatite (Baturin, 1971;
Burnett, 1977; Price and Calvert, 1978; Froelich et al.,
1983). The concentrations of P, C, and F in the upper few
tens of centimeters of the sediment pore water are high
enough to permit their direct chemical precipitation as
carbonate fluor apatite. On the other hand, the bone
material is made of dahlite; carbonate hydroxyapatite, with
F content far less than 1% or even hydroxyapatite which is
devoid of F (McConnell, 1973; Altschuler, 1973). Dahlite
is readily converted into carbonate fluorapatite through the
interaction with sediment pore water during early
diagenesis, incorporating F and possibly CO3 into its
structure. This conversion process takes place
contemporaneously with the direct precipitation of
pelletal/intraclast material as carbonate fluorapatite
(Froelich et al., 1983; Abed and Fakhouri, 1996).
Accordingly, there should be abundant organic matter
associated with the phosphorites in their depositional
environment. Organic matter is well known as a good
scavenger for U and many other trace metals. Most
probably, the newly deposited apatite would take up U into
its open crystal structure from the engulfing pore fluids
which were formed after organic matter decomposition.
Uranium, then, becomes part of the apatite occupying
some of the positions located for Ca in the apatite
structure. It should be emphasized that U in the ancient
phosphorites is present within the crystal structure of
apatite not adsorbed on the organic matter (Froelich et al.,
1983; Bernett, 1990).
In Jordan, the northwards increase of U may be due to
higher organic matter associated with the phosphorites in
north Jordan. The NW Jordan deposits have around 6%
organic matter (Abed and Al-Agha, 1989) with much
lower contents further south. The organic matter in the
other localities, other than NW Jordan, is to be seen within
the phosphate pellets and intraclasts. The dark colour of
some of these particles is indicator of organic matter. This
is evident from the leaching and oxidation of
pellets/intraclasts rims by percolating oxidizing
groundwater through these permeable phosphorite deposits
throughout the history of these deposits.
7. Conclusions
High grade phosphorites are wide spread in Jordan.
Despite the fact that Ruseifa mines were closed and the
central Jordan deposits and near depletion, huge reserves
are still existing in Eshidiyya Basin (in excess of 1000
million tons), Al-Kora Basin, NW Jordan (several hundred
million tons), and the subsurface of the eastern desert.
The phosphorites of Jordan are the main carrier of
Uranium in the country. Uranium is present within the
crystal structure of francolite (carbonate flour apatite)
substituting for Ca. Uranium content varies up 379 ppm in
certain phosphorite horizons.
Uranium content in the various localities increases
northwards. Eshidiyya deposits lower member (main
deposits) has around 70 ppm U while the upper member
(A0) 133 ppm, central Jordan 105 ppm, Ruseifa 123 ppm,
and NW Jordan 153 ppm. It should be emphasized that
certain phosphorite horizons have much more U compared
with these averages. Detailed field geology can pinpoint
such horizons if U is to be exploited. It is here postulated
that organic matter may be one of the main reason for the
northwards increase of U.
During mining and upgrading of the phosphorite ores,
U sticks to the mineral francolite and thus behaves similar
to Ca and P2O5. No groundwater contamination with U is
noticed in central Jordan despite the fact that fresh water
used in washing the upgraded ore has been discharged to
the local environment since more than 45 years.
© 2011 Jordan Journal of Earth and Environmental Sciences . All rights reserved - Volume 4, (Special Publication, Number 2)(ISSN 1995-6681)
44
Throughout the fertilizers industry, U follows P, first
into the phosphoric acid and then into the diammonium
phosphate (DAP) where it becomes concentrated by a
factor of 1.5 relative the fed ore. On the contrary, calcium
leaves francolite to phosphogypsum but U does not follow
suite, thus phosphogypsum is almost devoid of U with a
content up to 2 ppm.
REFERENCES
[1] Abed, A. M., 1989. On the genesis of the phosphorite-chert
association in the Amman Formation in Tel es Sur area,
Ruseifa, Jordan. Science Geologie Bulletin, Strasbourg 42,
141-153.
[2] Abed, A. M., 1994. Shallow marine phosphorite-chert-
palygorskite association, Upper Cretaceous Amman
Formation, Jordan. In: A. Iijima, A. M. Abed, and R.
Garrison, (Editors.), Siliceous, Phosphatic and Glauconitic
Sediments in the Tertiary and Mesozoic. Proc. 29th Intern.
Geol. Congr., Part C. VSP, The Netherlands, pp. 205-224.
[3] Abed, A. M., 2000. Geology of Jordan, Jordanian Geologists
Association, Amman (In Arabic).
[4] Abed, A. M. and Amireh, B., 1983. Petrography and
geochemistry of some Jordanian oil shale from north Jordan:
Journal of Petroleum Geology, 5, 261-274.
[5] Abed, A. M. and Khalid, H. 1985 Uranium distribution in the
Jordanian phosphorite. Dirasat, 12: 91-103.
[6] Abed, A. M. and Al-Agha, M. R., 1989. Petrography,
geochemistry and origin of the NW Jordan phosphorites. J.
Geol. Soc. London, 146: 499-506.
[7] Abed, A. M. & Fakhouri, K., 1996. On the chemical
variability of phosphatic particles from the Jordanian
phosphorites. Chemical Geology 131, 1-13.
[8] Abed , A. M. and Abu Murry, O., 1997. Rare earth element
geochemistry of the Jordanian Upper Cretaceous
phosphorites. Arab Gulf J. Scient. Res., 15, 41-61.
[9] Abed, A. M. and Sadaqah, R., 1998.. Role of Upper
Cretaceous oyster bioherms in the deposition and
accumulation of high- grade phosphorites in central Jordan.
Journal of Sedimentary Research, 68B, 1009-1020.
[10] Abed. A. M. & Amireh, B. S., 1999. Sedimentology,
geochemistry, economic potential and paleogeography of an
Upper Cretaceous phosphorite belt in the southeastern desert
of Jordan. Cretaceous Research, 20, 119-133.
[11] Abed, A. M., Arouri, K. and Boreham, C., 2005. Source rock
potential of the phosphorite-bituminous chalk-marl sequence
in Jordan. Marine and Petroleum Geology, 22, 413-425
[12] Abed, A. m., Sadaqah, R. and Al-Jazi, M., 2007 Sequence
stratigraphy and evolution of Eshidiyya phosphorite platform,
southern Jordan. Sedimentary Geology, 198, 209-219.
[13] Abed, A. M., Sadaqah, R., and Kuisi, M. A., 2008. Uranium
and potentially tozic metals during the mining, beneficiation
and processing of phosphorite and their effects on ground
water in Jordan. Mine water and the Environment, 27, 171-
182.
[14] Abed, A. M., Sadaqah, R., 2011. Enrichment of uranium in
the upper phosphorite horizon (A0) in Eashidiyya, south
Jordan. Cretaceous Research (in press).
[15] Al-Agha, M. R., 1985. Petrography, geochemistry and origin
of the NW Jordan phosphorites. Thesis. Univ. of Jordan
(unpubl).
[16] Al-Hwaiti, M., Matheis, G. and Saffarini, G., 2005,
Mobilization, redistribution and bioavailability of potentially
toxic elements in Shidiya phosphorites, Southern Jordan.
Environmental Geology 47, 431-444.
[17] Almogi-Labin, A., Bein, A. and Sass, E., 1993. Late
Cretaceous upwelling system along the southern Tethys
margin (Israel): Interrelationship between productivity,
bottom water environments, and organic matter preservation.
Paleoceanography, 8: 671-690.
[18] Altschuler, Z. S. 1973. The weathering of phosphatic
deposits: geochemical and environmental aspects. In: G.
Griffith, A. Becton, J. M. Spencer and D. T. Mitchell
(Editors.). Environmental Phosphorus Handbook. Willey,
New York, pp. 33-96.
[19] Altschuler, Z. S., 1980. The geochemistry of trace elements
in marine phosphorites, Part 2: Characteristics, abundances
and enrichment. SEPM Spec publ 29:19-30.
[20] Baturin, G. N. 1971. Recent authigenic phosphorite
formation on the SW African shelf. In: F. M. Delany
(Editor), The Geology of the East Atlantic Margin.
ICSU/SCOR Working Party 31 Symp. Cambridge. Inst.
Geol. Sci. Rep., 70/16, pp. 88-97.
[21] Baturin, G.N., 1982, Phosphorites on the sea floor..
Developments in Sedimentology 33, Elsevier, Amsterdam,
343 p.
[22] Bender, F. 1974. Geology of Jordan. Borntraeger, Berlin.
[23] Birch, G.F., 1980. A model for penecontemporaneous
phosphatization by diagenetic and authigenic mechanism
from the western margin of southern Africa. In Marine
phosphorites (ed Bentor, Y.), Society of Economic
Paleontologists and Mineralogists, Special Publications 29,
70-100
[24] Birch, G. F., Thomson, J., McArthur, J. and Burnett, W. C.,
1983. Pleistocene phosphorites off the west coast of South
Africa. Nature, 302: 601-603.
[25] Burnett, W. C. 1977. Geochemistry and origin of
phosphorites deposits from off Peru and Chile. Geol. Soc.
Amer. Bull., 88: 813-23.
[26] Burnett, W.C., 1990. Phosphorite growth and sediment
dynamics in the modern Peru shelf upwelling system, In
Phosphorite deposits of the world, v. 3, Neogene to modern
phosphorites (eds Burnett, W.C. and Riggs, S.), 62-74.
Cambridge University Press, Cambridge).
[27] Capetta, H., Pfeil, F. Schmidt-Kittler, N. and Martini, E.,
1996. New biostratigraphic data on the marine Upper
Cretaceous and Paleogene of Jordan. Unpublished report by
the, Jordan Phosphate Mining Co, Amman, Jordan.
[28] Coppens, R., Bashir, S. and Ritchard, P. 1977. Radioactivity
of El-Hasa phosphates, a preliminary study. Mineral
Deposita, 12, 189-196.
[29] El-Hiyari, M., 1985. The geology of Jabal Al-Mutarammil,
Map sheets No 3252, III. National Mapping Project, Bulletin
1, Natural Resources Authority, Amman, Jordan.
[30] Follmi, K. B., 1996. The phosphorus cycle, phosphogenesis
and marine phosphate-rich deposits. Earth Science Review
40, 55-124.
[31] Froelich, P. N., Kim, K. H., Jahnke, R., Burnett, W. C.,
Soutar, A. and Deakin, M. 1983. Pre water fluoride in Peru
continental margin sediments: uptake from seawater.
Geochim. Cosmochim. Acta, 47: 1905-12.
[32] Froelich, P., Arthur, M., Burnett, W., Deakin, M., Hensley,
V., Jahnke, R., Kaul, L., Kim, H., Soutar, A., and
Vathakanoan, C. 1988. Early diagenesis of organic matter in
Peru continental margin sediments- phosphorite precipitation:
Marine Geology, 80, 309-343.
[33] Glenn, C. R. and Arthur, M. A., 1990. Anatomy and origin
of a Cretaceous phosphorite-green sand giant, Egypt.
Sedimentology, 37: 123-148.
[34] Glenn, C. R.; Follmi, K. B.; Riggs, S. R.; Baturin, G. N.;
Grimm, K. A. Trappe, J.; Abed, A. M.; Galli-Oliver, C.;
Garrison, R. E.; Iiyin, A. V.; Jehl, C.; Rohrlich, V.; Sadaqah,
R. M.; Schidlowski, M.; Sheldon, R. E. and Siegmind, H.,
1994. Phosphorus and phosphorites: sedimentary and
environments of formation. Eclogae Geologicae Helvetica,
87, 747-788.
© 2011 Jordan Journal of Earth and Environmental Sciences . All rights reserved - Volume 4, (Special Publication, Number 2)(ISSN 1995-6681)
45
[35] Glikson, M., Gibson, D. L. and Philp, R. P., 1985. Organic
matter in Australian Cambrian oil shale and other Lower
Palaeozoic shales: Chemical Geology, 51, 175-191.
[36] Hamam, K. A., 1977. Forminifera from the Maastrichtian
bearing strata of Al-Hasa, Jordan. Jour. Foram. Res., 7: 34-
43.
[37] Heggie, D. T.; Skyring, G. W., O’Brian, G. W., Reimers, C.,
Herczef, A., MoriartyD. J. W., Burnett, W. C. and Milnes, A.
r., 1990. Organic carbon cycling and the modern phosphorite
formation on the East Australian continental margin: an
overview. In: A. J. G. Notholt and I. Jarvis (Eds.)
Phosphorite Research and Development. Geological Society
Special Publication 52, 87-118.
[38] Iijima, A., Abed, A. M. and Garrison, R., (Editors.),
Siliceous, Phosphatic and Glauconitic Sediments in the
Tertiary and Mesozoic. VSP, The Netherlands.
[39] Jarvis, I., 1992. Sedimentology, geochemistry and origin of
phosphatic chalk: The Upper Cretaceous of NW Europe.
Sedimentology, 39, 55-97.
[40] Jiries, A., El-Hasan, T., Al-Hiwati, M., and and Seiler, K. P.,
2004. Evaluation of the Effluent Water Quality Produced
from Phosphate Mines in Central Jordan. Mine Water and the
Environment. 23 (3): 133-137 Khalid, H., and Abed, A.
M., 1982. Petrography and geochemistry of Esh Shidiya
phosphates. Dirasat, 9: 81-102.
[41] Khalifa, M. Kh. And Abed, A. M., 2011. Lithostratigraphy
and microfacies analysis of the Ajlun Group (Cenomanian to
Turonian) in Wadi Sirhan Basin, SE Jordan. Jordan Journal
for Environmental and Earth Sciences (IN press)
[42] Kolodny, Y. and Garrison, R., 1994. Sedimentation and
diagenesis in paleo-upwelling zones of epeiric sea and
basinal settings: A comparison of the Cretaceous Mishash
Formation of Israel and the Miocene Monterey Formation of
California. In: A. Iijima, A. M. Abed, and R. Garrison,
(Editors.), Siliceous, Phosphatic and Glauconitic Sediments
in the Tertiary and Mesozoic. Proc. 29th Intern. Geol.
Congr., Part C. VSP, The Netherlands, pp. 133-158.
[43] Krauskopf, K. and Bird, D., 1995. Introduction to
geochemistry, 3rd Ed. McGraw Hill, Boston.
[44] Lucas, J. and Prevot-Lucas, L., 1995. Tethyan phosphates
and bioproductites. In: Nairn, A. E. et al. (eds), The ocean
basins and margins - The Tethys ocean. Plenum Press. 8,
367-391.
[45] McArthur, J. M., Hamilton, P. J., Greensmith, J. T., Walsh, J.
N., Boyce, A. B., Fallick, A. E., Birch, G., Benmore, R. A.
and Coleman, M. L., 1987. Francolite geochemistry -
meteoric alteration on a local scale. Chem. Geol. (Isotope
Geoscience), 65: 415-425.
[46] McClellan, G. H. and Kauwenbergh, S. J. V., 1990.
Mineralogy of sedimentary apatite. In: A. J. G. Notholt and I.
Jarvis (Editors), Phosphorite Research and Development.
Geol. Soc. Lond. Specl. Publ. no. 52, 23-31.
[47] McConnell, D., 1973. Apatite, crystal chemistry and
mineralogy. Springer, Berlin. 111pp.
[48] Mikbel, Sh. and Abed, A. M., 1985. Discovery of large
phosphate deposits in NW Jordan. Dirasat, 12, 125-124.
[49] Nathan, Y., 1984. The mineralogy and geochemistry of
phosphorites. In: Nriagu, J. O. and Moore, P. O. (eds),
Phosphate minerals. Springer, Berlin, 275-291.
[50] Notholt, A. J. G., Sheldon, R. P. and Davidson, D. F., 1989.
Phosphate deposits of the world, Volume 2, Phosphate rock
resources. Cambridge University Press, Cambridge.
[51] Powell, J. H., 1989. Stratigraphy and Sedimentation of the
Phanerozoic Rocks in Central and South Jordan, Part B:
Kurnub, Ajlun and Belqa Groups: Bulletin 11, NRA,
Amman.
[52] Quennell, A. M., 1951. The geology and mineral rsorces of
the (former) Transjordan. Colonial Geology and Mineral
Resources, London, 2, 85-115.
[53] Prevot, L. and Lucas, J., 1980. Behavior of some trace
elements in phosphatic sedimentary formations. In: Y. K.
Bentor (Editor), Marine phosphorites. SEPM Spec. Publ., 29:
31-41.
[54] Price, N. B. and Calvert, S. E., 1978. The geochemistry of
phosphorites from the Namibian shelf. Chem. Geol., 23: 151-
178.
[55] Riggs, S, R., 1979. Petrology of the Tertiary phosphorite
system of Florida., Econ. Geol., 74: 194-220.
[56] Riggs, S. R., 1980. Intraclast pellet phosphorite
sedimentation in the Miocene of Florida. Journal of the
Geological Society, London 137, 741-748.
[57] Riggs, S. R., Snyder, S., Hine, A., Snyder, Sc., Elligton, D.
and Mallette, P., 1985. Geologic framework of phosphate
resources in Onslow Bay, North Carolina continental shelf.
Econ. Geol., 80: 716-738.
[58] Riggs, S. R., Ambrose, W. G., Cook, J. W., Snyder, S. C., &
Snyder, St. W., 1998. Sediment production on sediment-
straved continental margins: The interrelationship between
hardbottoms, sedimentological and benthic community
processes, and storm dynamics. Journal of Sedimentary
Research, A68, 155-168.
[59] Sadaqah, R. M., 2000. Phosphogenesis, geochemistry, stable
isotopes and depositional sequence of the Upper Cretaceous
phosphorite formation in Jordan. Unpublished Ph. D thesis ,
Univ. of Jordan, Amman, 257p.
[60] Sadaqah, R. M., Abed, A.M., Grimm, K. A. and Pufahl, P.
K., 2005. The geochemistry of REE , Yttrium and Scandium
in some Upper Cretaceous Jordanian phosphorites, Dirasat
32, 32-47.
[61] Sheldon, R. P. 1981. Ancient marine phosphorites. Ann. Rev.
Earth planet. Sci. 9: 251-84.
[62] Sheldon, R. P., 1987. Association of phosphatic and siliceous
marine sedimentary deposits. In: J. R. Hein (Editor),
Siliceous sedimentary rock-hosted ores and petroleum.
Nostrand Reinhold, New York, pp. 56-79.
[63] Slansky, M., 1986. Geology of sedimentary phosphates, 210
pp. (Elsevier, Amsterdam).
[64] Soframines and Consortium Partners, 1987. Feasibility study
of Eshidiyya phosphate project. Unpubl. Jordan Phosphate
Mines Company, Amman, Jordan.
... Le gisement marin de Jordanie s'est formé au Crétacé lors du plus grand épisode de dépôt de phosphorites, notamment dans la zone Est de la Méditerranée et au Nord de l'Afrique. Les apatites sont contenues dans les phosphorites mais le gisement est également constitué de calcaires, de cherts, de concrétions d'huitres, de marnes riches en matière organique, etc. (Abed, 2011). ...
... In uranium ore samples, uranium concentrations vary from a few hundred µg/g to 20 wt.% and molybdenum concentrations vary from a few ng/g to more than 0.5 wt.% Thompson et al., 1978;UDEPO), with U/Mo ratios ranging between 1x10 -3 and 1 (CETAMA, 2011;. Uranium-bearing minerals are of two main types: (1) Primary minerals are U(IV) compounds such as uraninite (UO 2 ), pitchblende (collomorphous . ...
... The phosphorite from Jordan was formed in a marine environment, and uranium is hosted in apatite minerals. Phosphorites in the Eastern Mediterranean formed during a Cretaceous episode which deposited one of the most extensive deposits in the world (Abed, 2011). ...
Thesis
Au cours de ces dernières décennies, des études ont étés menées pour identifier plusieurs traceurs des matériaux du cycle du combustible nucléaire, dans le cadre de la lutte contre la prolifération nucléaire. Ces matériaux sont généralement collectés lors d’inspections dans des installations nucléaires, ou saisis lors de contrôles de trafics illicites. Les informations fournies par ces traceurs sont parcellaires et ne permettent pas de déterminer avec exactitude la provenance et l’historique industriel de ces matériaux.Le but de ce travail de thèse est de démontrer le potentiel de l’utilisation des isotopes du molybdène pour le traçage des matériaux du cycle du combustible nucléaire. Le choix s’est porté sur le molybdène car en raison de la similarité de leurs propriétés chimiques, le molybdène et l’uranium sont étroitement associés dans les minerais d’uranium et tout au long de la chaîne de purification de l’uranium. L’étude s’est focalisée sur une partie de l’amont du cycle du combustible, depuis l’extraction des minerais d’uranium jusqu’à la production des concentrés miniers d’uranium : divers procédés physiques et chimiques sont appliqués, à la fois pour purifier l’uranium et abaisser la concentration en molybdène.Au cours de cette étude, une nouvelle méthode de séparation du molybdène a été développée pour caractériser sa composition isotopique dans des minerais, minéraux et concentrés miniers d’uranium. La variabilité des compositions isotopiques du molybdène dans un gisement d’uranium est principalement due aux mécanismes d’adsorption et/ou de précipitation du molybdène. Les gisements magmatiques et sédimentaires ont des compositions isotopiques différentes, ce qui permet ainsi leurs distinctions. Les concentrés miniers d’uranium produits à partir de ces deux types de gisements ont des compositions isotopiques similaires aux minerais. Ces résultats soulignent ainsi le potentiel des isotopes du molybdène comme traceur des origines des concentrés miniers d’uranium. Cependant, un fractionnement des isotopes du molybdène a été établi lors de la production des concentrés miniers d’uranium pour deux usines au Niger. Les procédés de purification de l’uranium tels que la lixiviation, l’extraction par solvant et la précipitation ont été reproduits en laboratoire sur des échantillons réels pour expliquer le fractionnement isotopique du molybdène lors de la production des concentrés miniers. Au cours de ces procédés, le fractionnement peut être positif (lixiviation), négatif (extraction par solvant, précipitation à l’eau oxygénée) ou nul (précipitation à l’ammoniaque). Dans le cas des échantillons du Niger, la somme de ces procédés est négative, dans le sens des données expérimentales que nous avons obtenues, démontrant ainsi également le potentiel de l’utilisation des isotopes du molybdène comme traceur des procédés de transformations des matériaux du cycle du combustible nucléaire.
... The Mazıdağı phosphates were formed on a shallow marine platform located in the south of the Neo-Tethys Ocean between the African-Arabic and Eurasian plates during the Upper Cretaceous period. The Mazıdağı phosphate deposits from bottom to top occur The phosphate deposits of the eastern Mediterranean region are almost always associated with chert, porcelanite, and oyster-bearing limestone, in addition to minor marl, chalk, and sandstone [105][106][107][108][109]. In this study, within the Mazıdagı phosphate deposits, chert levels were observed with a thickness of 25-40 cm in certain places (Figure 4). ...
... The development of chert levels is thought to have come to an end during the precipitation of carbonates and phosphates, or otherwise to have been precipitated simultaneously with the carbonates. Chert levels have also been observed in Moroccan, Jordanian, Tunisian, and Syrian phosphate deposits on the same belt [20,106,[109][110][111][112]. Tlili et al. [112] claim that chert levels may be related to the bacterial activity observed in chert beds, which induces the dissolution of silica, which, in turn, saturates the depositional environment through silicic acid. ...
Article
Full-text available
The Tethyan phosphates were formed during the Upper Cretaceous and Eocene interval as a result of the collision of the African–Arabian and Eurasian plates and the closing of the Neo-Tethys Ocean. This study aimed to reveal the possible precipitation parameters of these phosphates by examining the main oxide, trace and rare earth element contents of the phosphates in the study region. The mean major oxide concentrations of the phosphates were found to be 51.6 wt.% CaO, 21.2 wt.% P2O5, 8.03 wt.% SiO2, 18.1 wt.% CO2, 0.51 wt.% K2O, 0.12 wt% Fe2O3, 0.05 wt% Al2O3, 0.18 wt% MgO and 0.02 wt.% MnO. The average trace element concentrations were 79 ppm Ba, 1087 ppm Sr, 0.23 ppm Rb, 14.7 ppm Ni, 108 ppm Cr, 262 ppm Zn, 27 ppm Cd, 21.6 ppm Y, 58 ppm V, 6.43 ppm As, 30.3 ppm Cu, 1.36 ppm Pb, 6.32 ppm Zr, 39 ppm U, 0.21 ppm Th and 1.33 ppm Co. The average trace element contents were 1742 ppm, with this indicating an enrichment assemblage of Sr, Cd, As, and Zn in comparison to PAAS (The Post-Archean Australian Shale). The total REE concentrations in the Mazıdağı phosphates varied from 3.30 to 43.1 ppm, with a mean of 22.1 ppm recorded. All phosphates showed heavy REE (HREE) enrichments and had similar REE patterns to PAAS. All samples had strongly negative Ce and positive Eu, Pr and Y anomalies. These anomalies indicate the existence of oxic and suboxic marine conditions during the formation of the phosphates. According to the proposed genetic model, the phosphates mostly formed in oxic and suboxic zones of the Tethys Ocean and were precipitated on slopes that depended on strong upwelling from an organic-rich basin in anoxic/suboxic conditions from deeper seawater. The Pb isotope data obtained also indicate the existence of a deep-sea hydrothermal contribution to this phosphate formation.
... It is generally assumed that uranium, as U 4+ , substitutes for Ca 2+ in the apatite lattice due to very similar ionic radii (e.g. Baturin and Kochenov, 2001;Soudry et al., 2002;Abed, 2011;Dar et al., 2014, and references therein). Conversely, the U 6+ content in phosphorites is thought to be the result of post-depositional direct oxidation as well as additional oxidation of the recoil product of the 238 U 4+ decay (e.g. ...
... Although trace elements are not necessarily located within the mineral structure itself but may be adsorbed onto crystals, as suggested for the Pb sorption onto CO 3 FAp (Prasad et al., 2002), it is well known that crystal structure of the apatite group minerals permits a wide variety of both cationic and anionic substitutions also involving a change in valence. In general, the PO 4 can be replaced by VO 4 , As 2 O 4 , SO 4 , or CO 3 ; the F may be partly or completely substituted for Cl or OH; small amounts of Mg, Mn, Sr, Pb, Na, U, rare earth elements and Y may substitute for Ca (Elliot, 1994;Abed, 2011). Blumenthal (1990), for instance, stated that cations of larger ionic radii than calcium, such as Sr and Pb, can be incorporated in the OHAp structure resulting in an expansion of the lattice parameters along the a and c axes. ...
Article
In the Tethyan realm the carbonate-dominated Meso-Cenozoic South Tethyan Phosphogenic Province is of considerable economic importance since it represents the greatest accumulation of sedimentary phosphorites. In southern Italy, in the Salentine peninsula (the southern part of the Apulian Carbonate Platform, ACP), is well documented the occurrence of Cenozoic P-rich levels consisting of nodules and pebbles and showing a large P2O5 fluctuation (4.07–22.07 wt%), due to variable calcite abundance. The mainly observed P-bearing minerals are hydroxyapatite and carbonate fluorapatite and U, together with Sr and Pb, preferentially acts as Ca substitutes in both lattices. Minor authigenic monazite (LREE-bearing phosphate) and xenotime (HREE+Y-bearing phosphate), likely formed during sediment burial compaction and diagenesis. The total REEs' abundances and the shape of the shale-normalized REE-patterns in the P-rich nodules are in the range of those typically observed in Paleocene-Eocene through Pleistocene-Recent P-rich sediments, supporting the idea of a broadly consistent ocean chemistry in this span of geological time. The (La/Yb)N proxy is within the modern seawater range, signifying early diagenetic adsorption played only a minor role in affecting the REEs' distribution. The Ce and Pr anomalies suggest some P-rich nodules experienced very localized hypoxic to anoxic conditions promoting Ce/Ce* spurious results in a generally oxic to hypoxic environment causing real negative Ce anomaly. This scenario is reinforced by the lack of the coupled uranium-vanadium enrichment typically observed in an anoxic environment. Since the Pliocene the eastern Mediterranean was variously affected by oxygenation and productivity and it is likely the ACP phosphate-rich sediments formed under low sedimentation rates and authigenesis in a bottom current-dominated regime, as also indicated by the glauconite occurrence. Transgressions and sea levels rising following major glaciations may have favored the deposition of phosphate sediments by creating new restricted basin configurations and increased nutrient input promoted by upwelling processes.
... All Belqa Gr. lithologies in Levant east of the Dead Sea-Jordanian Transform have abnormal enrichment in redox-sensitive elements (RSEs) that largely exceed the absolute contents in most of the organic-rich sediments worldwide. Trace elements are especially abundant and diverse in immature organic-rich siliceous chalk ('oil shale') of the Maastrichtian-Paleocene Muwaqqar Chalk Marl Fm., a unit of the Belqa Gr., in central Jordan ( Figure 2) [4,14,[18][19][20]30,[48][49][50][51]. The Muwaqqar Fm. 'oil shales' were deposited on the southern Neo-Tethys epicontinental shelf in a setting typical of pristine phosphorites, under anoxic/euxinic conditions during intense upwelling. ...
Article
Full-text available
Minerals of the Zn-Cd-S-Se system that formed by moderately reduced~800-850 • C combustion metamorphic (CM) alteration of marly sediments were found in marbles from central Jordan. Their precursor sediments contain Se-and Ni-enriched authigenic pyrite and ZnS modifications with high Cd enrichment (up to~10 wt%) and elevated concentrations of Cu, Sb, Ag, Mo, and Pb. The marbles are composed of calcite, carbonate-fluorapatite, spurrite, and brownmillerite and characterized by high P, Zn, Cd, U, and elevated Se, Ni, V, and Mo contents. Main accessories are either Zn-bearing oxides or sphalerite, greenockite, and CaFe -Ni-Cu-O-S-Se oxychalcogenides. CM alteration lead to compositional homogenization of metamorphic sphalerite, for which trace-element suites become less diverse than in the authigenic ZnS. The CM sphalerites contain up to~14 wt% Cd and~6.7 wt% Se but are poor in Fe (means 1.4-2.2 wt%), and bear 100-250 ppm Co, Ni, and Hg. Sphalerite (Zn,Cd,Fe)(S,O,Se) cub is a homogeneous solid solution with a unit cell smaller than in ZnS cub as a result of S 2− → O 2− substitution (a = 5.40852(12) Å, V = 158.211(6) Å 3). The amount of lattice-bound oxygen in the CM sphalerite is within the range for synthetic ZnS 1−x O x crystals (0 < x ≤ 0.05) growing at 900 • C.
... The Phosphorite Unit at Wadi Al-Abyad and at El-Hasa,120 km and 140 km south of Amman, is about 40 m thick (Bender 1968). The El-Shidiya Phosphorite Unit (8-10 m thick), 30 km S-SE of Ma'an, covers about 125 km 2 and is the most promising deposit (Abed 2011(Abed , 2013Abed et al. 2005Abed et al. , 2007Abed et al. , 2014. It underlies a coquina marker with TCP range between 65 and 72%. ...
Article
Mineral resources of Jordan were utilized by different civilizations since prehistoric time. The following review highlights important and strategic commodities that are found in a small country with a potential promising future. Jordan is rich in its diverse industrial rocks and mineral resources that are part of the whole stratigraphic sequence from the Precambrian basement complex to the Recent sediments. Such diversity of resources enables Jordan to be not only a host of some unlimited resources but also a natural geology museum. The known commodities include, among others, phosphates, bituminous marl (oil shale), tar sand, varicolored marbles, travertine, radioactive minerals, building and construction materials, clays and clay minerals, diatomaceous earth, porcelanite, Dead Sea brines, rock salt, chalk, limestone, gypsum, glass sand, basalt, pyroclastics, zeolites, granite, copper, manganese, gold, etc. Some of these commodities are unique and unusual in its mineralogy, chemistry, and origin. A novel geopolymerization process was developed in Jordan to produce green building materials (geopolymers) by using Jordanian kaolinite. Volkonskoites (Cr-rich smectite) of Jordan act as a sink for hazardous elements. Varicolored marbles of Jordan are considered a natural cement factory and are analogs of cementitious repositories with the highest alkaline circulating waters in the World. New minerals were reported for the first time, some of which are only known in meteorites. The unusual enrichment of the reduced sensitive elements (RSE) in the oil shale (similar to the source of Mississippi Valley-type deposits) makes it unique as a potential source of these strategic elements.
... Uranium has a minimum of 0.65 ppm in the carbonate and sandstone samples, and a maximum of 158 ppm in the phosphorite samples. The latter value is within the range of the other phosphorite deposits of Jordan (Sadaqah, 2000;Abed, 2011). Strontium follows Ca in its minerals. ...
Article
Full-text available
Representative samples from the Al-Hisa Phosphorite Formation (AHP) in the SE desert of Jordan, are selected from three sections: Batn El-Ghoul, Nagb Etayyeg, and Zgaimat Al-Hasah. The samples are investigated geochemically and petrographically. Geochemistry is discussed through the analysis of the major, trace and rare earth elements (REEs), loss on ignition (LOI), sulphur, and organic matter. The studied sections differ from the AHP in central Jordan by having highly reduced thicknesses, omission of underlying formations, abundant detrital quartz sand and age. The phosphate particles are of authigenic and biogenic origin. Early reworking from phosphate mud is evident from the shape and sorting of the phosphate particles. Major and trace elements are distributed into five factors: upwelling, provenance, redox potential, sea level changes, and resistates. The studied sediments were formed under oxic marine conditions because their REEs patterns preserve the REEs seawater linear pattern: depletion of LREEs, a negative Ce anomaly, and an enrichment of HREEs. This signature seems to have persisted through reworking and late diagenetic cementation, dolomitization and silicification. The SE desert phosphorite-siliciclastic sequence is better correlated with the Paleocene Jalamid Formation of northwestern Saudi Arabia because of similarities in age and lithology due to being both deposited on the western flanks of the Sirhan Paleohigh and the Rutba Paleohigh, respectively. The siliciclastics were delivered from a mixed mafic-felsic-older cycle siliciclastic provenance, and subjected to intensive chemical weathering under tectonic quiescence giving rise to mature quartzarenites.
Article
Full-text available
Minjingu phosphate ore is Tanzania's sole domestic supply of phosphorus (P). The ore contains medium to high concentrations of naturally occurring P 2 O 5 (20-35 %) and relevant concentrations of uranium and rare earth elements (REEs) are also suspected to be present. Currently, neither uranium nor REEs are recovered. They either end up in mine tailings or are spread across agricultural soils with fertilizer products. This work provides a first systematic review of the uranium and REE concentrations that can be expected in the different layers of Minjingu phosphate ore, the way the ore is presently processed, as well as a discussion on alternative processing pathways with uranium/REE recovery. The study analyzed ten distinct Minjingu phosphate ore layers, four mine tailings, and five intermediate and final mineral fertilizer products from the Minjingu mine and processing plant located in northern Tanzania. The results confirm that the uranium concentrations and to a lesser degree, the REE concentrations are indeed elevated if compared to concentrations in other phosphate ores. The study does not identify a significant risk resulting from this. The development of techno-economic solutions for more 2 comprehensive utilization of Minjingu ore is, however, strongly encouraged and suggestions on such processes are provided.
Article
Full-text available
Minjingu phosphate ore is Tanzania's sole domestic supply of phosphorus (P). The ore contains medium to high concentrations of naturally occurring P 2 O 5 (20-35 %) and relevant concentrations of uranium and rare earth elements (REEs) are also suspected to be present. Currently, neither uranium nor REEs are recovered. They either end up in mine tailings or are spread across agricultural soils with fertilizer products. This work provides a first systematic review of the uranium and REE concentrations that can be expected in the different layers of Minjingu phosphate ore, the way the ore is presently processed, as well as a discussion on alternative processing pathways with uranium/REE recovery. The study analyzed ten distinct Minjingu phosphate ore layers, four mine tailings, and five intermediate and final mineral fertilizer products from the Minjingu mine and processing plant located in northern Tanzania. The results confirm that the uranium concentrations and to a lesser degree, the REE concentrations are indeed elevated if compared to concentrations in other phosphate ores. The study does not identify a significant risk resulting from this. The development of techno-economic solutions for more 2 comprehensive utilization of Minjingu ore is, however, strongly encouraged and suggestions on such processes are provided.
Article
Full-text available
Introduction Phosphorites are marine sediments of biogenic origin containing 15–20 wt% P2O5 and between 50 and 120 ppm U (Boggs, 2009; Tzifas et al., 2014; Zarasvandi et al., 2019). The study of phosphorites, especially trace elements geochemistry, confirms the importance of deposition conditions and diagenesis on the elemental composition of phosphatic minerals. Even more importantly, marine phosphorites are considered to have an economic potential for elements such as REE, Sc, U and Th (Altschuler, 1980). Some trace elements, including Sr, Ba, Se, Mo, Ag, Pb, Zn, V, Cr, Ni, Cu, Cd, and U are commonly found in phosphorites and sediments rich in phosphorus related to the crystal structure of apatite and carrier organic ligands (Tzifas et al., 2014; Zarasvandi et al., 2019). In general, more than seven horizons with an extent of ∼400 to 100km have been delineated in the Zagros Mountains. The Zagros phosphorite horizon of Eocene-Oligocene age hosted by the Pabdeh Formation is located in the Zagros fold belt with NW-SE trend (Halalat and Bolourchi, 1994; Zarasvandi et al., 2019). The aim of this study is to investigate the geochemistry of trace elements in order to obtain the deposition and genesis conditions of these elements in the Zagros phosphorite horizon. Materials and methods A total of 29 samples were taken from phosphorite, hydrocarbon-bearing shale, phosphorite and limestone and oxide zone of the studied phosphorites of Zagros. Hence, based on stratigraphy, different samples from Zagros phosphorite horizons were collected from the phosphorites of Kuh-e-Sefid (n=9), Kuh-Rish (n=12) and Sheykh-Habil (n=8). Mineralogical and geochemical studies were carried out using ICP-MS analyses. 20 polished-thin sections were prepared. Mineralogy and petrography of the samples was determined and examined using polarizing-reflected light optical microscopy at the Shahid Chamran University of Ahvaz in Iran. Geochemical studies on mineralized and host rocks of Zagros phosphorite horizon were performed by the ICP-MS technique (Thermo Scientific- X Series II; DL= 0.001 ppb) at the Department of Earth Sciences, Pondicherry University in India. Results According to the petrographic studies, phosphorite components and non-phosphorite components mainly consist of pellets, Ooids, intraclasts, fish skeletal fragments, micro-fossils, glauconite, calcite, pyrite, iron-oxide and quartz. Several elements that substitute Ca including rare earth elements and trace elements are suitable for contribution in the carbonate-rich fluorapatite (francolite) crystalline structure. Thus, some oxo-anions such as VO4, As2O4, SO2, SO4 and CO3 can be substituted into PO4 structure in apatite group lattices (Tzifas et al., 2014; Zarasvandi et al., 2019). Consistently, the Zagros phosphorite horizon exhibits different concentrations of elements such as Sr, REE, Zn, V, Mo, Cr, Cd, Se, As and U. Trace element distribution patterns in the studied phosphorites are similar to phosphorite in Iran and worldwide, especially in terms of concentration of U, Se, and Cd that can be related to apatite group minerals crystal lattice (Tzifas et al., 2014; Zarasvandi et al., 2019). Due the low entrance rate of detrital components from continental to the basin, the most probable source for trace elements is hydrocarbon-bearing shale in the stratigraphic column as a result of activities of microorganisms. Discussion Field observation and microscopic studies showed that the phosphorite components occur as authigenic apatite with sparite cement, abundant pellets, ooids of symmetrical to elongated shape due to pressures caused by diagenesis, oval shape intraclasts, fish skeletal fragments and abundant microfossils. In additions to phosphorite and biogenic components, non-phosphorite minerals such as calcite, glauconite, pyrite, iron oxide, and microcrystalline quartz are present. There are many indications of change in conditions. They include bituminous shale in stratigraphic sequence, presence of abundant framboidal pyrites, PAAS-normalized patterns of REEs, negative Ce anomaly of all samples and positive Eu anomalies of all samples except bituminous shale sample of Kuh-e Rish phosphorite, the Ni/Co ratio and also the diagram of V/(V+Ni) vs. Ni/Co. These indicate changes in conditions from oxides during phosphate deposition into dysoxic to anoxic due to degradation and decomposition of organic compounds by microorganisms and the entry of trace elements such as uranium into the crystalline structure of apatite in the Zagros Basin. The significant economic potential of organometallic elements especially U and REE is observed in the Zagros phosphorite horizon due to favorable conditions of dysoxic to anoxic as a result of decomposition of organic compounds and then the entry of the elements into the apatite crystal structure. References Altschuler, Z.‌S., 1980. The Geochemistry of Trace Elements in Marine Phosphorites Part I. Characteristic Abundances and Enrichment. In: Y.K. Bentor (Editor), Marine Phosphorites-Geochemistry, Occurrence, Genesis. SEPM Society for Sedimentary Geology, Reston, pp. 19–30. https://doi.org/10.2110/pec.80.29.0019 Boggs, S., 2009. Petrology of Sedimentary Rocks. Cambridge University Press, England, 600 pp. Retrieved October 3, 2020 from https://www.researchgate.net/publication/281604561 Halalat, H. and Bolourchi, M., 1994. Geology of Iran: Phosphate. Geological Survey of Iran, Tehran, 362 pp. (in Persian with English abstract) Tzifas, I.Tr., Goldelitsas, A., Magganas, A., Anderoulakaki, E., Eleftheriond, G., Mertzimckis, T.J. and Perraki, M., 2014. Uranium-bearing phosphatized limestone of new Greece. Journal of Geochemical Exploration. 143: 62–73. https://doi.org/10.1016/j.gexplo.2014.03.009 Zarasvandi, A., Fereydouni, Z., Pourkaseb, H., Sadeghi, M., Mokhtari, B. and Alizadeh, B., 2019. Geochemistry of trace elements and their relations with organic matter in Kuh-e-Sefid phosphorite mineralization, Zagros Mountain, Iran. Ore Geology Reviews, 104: 72–87. https://doi.org/10.1016/j.oregeorev.2018.10.013
Article
Full-text available
Three major lithofacies have been described in Lajjun area, central Jordan, including the phosphorite and intercalated limestone of the Al‐Hisa Phosphorite Formation, as well as the Muwaqqar Chalk Marl Formation. ⁸⁷ Sr/ ⁸⁶ Sr isotopic data indicate a late Campanian age for the Al‐Hisa Phosphorite Formation, whereas the overlying chalk of the Muwaqqar Formation contains the lowest Maastrichtian planktic foraminiferal zone of Rugoglobigerina hexacamerata (CF8b). The phosphorites are of reworked origin as inferred from the homogeneous texture with lack of any concentric structure in the phosphatic pellets and the presence of bone fragments inside the phosphatic pellets, which are filled with phosphatic mud similar to the matrix of phosphatic pellet. They still reflect the original seawater rare‐earth pattern, as indicated from the similarity in their rare‐earth element (REE) concentration and patterns with seawater. Positive correlations between P 2 O 5 and Fe 2 O 3 , TiO 2 and K 2 O, relatively high rare‐earth contents, and similarity in (Nd/Yb) SN ratio to the modern shallow seawater suggest a continental marginal depositional environment for the limestone and chalk facies. δ ¹³ C values are lower and more variable compared with the estimated global range of upper Campanian–lower Maastrichtian δ ¹³ C values in shelf and oceanic basins. This indicates a diagenetic effect on the δ ¹³ C isotopes and suggests that a significant proportion of carbonate ions could have been resulted from the organic carbon oxidation. Three third‐order depositional sequences matching with the global and regional sea‐level patterns are identified in the studied section based on vertical facies changes. The basal transgressive surface of each sequence is constantly characterized by a phosphatic horizon.
Article
Full-text available
The interbedded phosphorites and cherts of the uppermost Amman Formation of the Tel es Sur (TES) area have been investigated and compared with the nearby economic phosphorites of Ruseifa. The strata are arranged in several shallowing upward cycles deposited in shallow subtidal, intertidal and supratidal (emerged) environments. -from Author
Article
Full-text available
A new large phosphate ore-body has been discovered in NW Jordan, and is herein described. It has almost the same stratigraphic position as other phosphate rocks in Central and S Jordan, i.e. the top of the Amman Formation (Phosphorite Member). Preservation of the ore is maintained by tectonic factors controlling the area (Wadi El-Yabis structure). The phosphate rocks are made of phosphatic faecal pellets, intraclasts, bones and teeth in a micritic limestone matrix. Estimated reserves are 370 million tons but such a preliminary estimate may be enhanced by more detailed work and drilling. The overburden consists of soft Chalk-Marl rocks, 0-30m thick. -from Authors
Article
Full-text available
A total of thirty-five samples of the economic grade Jordanian Upper Cretaceous phosphorites were analysed for certain major and rare earth elements (REE) with an overall objective of understanding phosphogenesis conditions. These samples represent pelletal (granular) phosphorites with traces of carbonates and marl. The deposits are of shallow subtidal environment of authigenic nature. The relative abundance of REE normalised to shale standard is indicative of proximal to onshore phosphorites. The distribution patterns of the REE show a depletion in cerium (Ce) and light REE (LREE) and enrichment in the heavy REE (HREE), indicating marine depositional environment. The highly negative Ce anomalies of these samples indicate phosphogensis under oxic marine condition which are interpreted as a signature of deep cold upwelling seawater on shallow epeiric shelf, cleanliness of these deposits from clay minerals and to the very limited diagenetic processes affecting them after deposition.
Article
“Crystal chemistry” is not now a widely accepted and integral part of introductory mineralogy in this country because it is commonly considered too difficult or complicated and because no suitable textbooks are available. However, crystal chemistry actually is simple and useful enough that it can and must be made use of at the introductory level. And with the aid of crystal chemistry, mineralogy courses can be greatly revitalized. One possible philosophy and line of approach is outlined in this paper whereby crystal chemistry can, with relative ease, be made a useful and integral part of an introductory mineralogy course. Some helpful reading material and laboratory work are also suggested.
Article
Among the periods known to have been more especially phosphogenic, the Late Cretaceous—Paleocene is one of the best known, because of the number and size of mined deposits such as those of Morocco and the Middle East. All deposits of that period are linked to Tethys and the bordering platforms, the evolution of which was responsible for the paleogeographical conditions needed to interrupt the biological cycle of phosphorus by mineralization of this element. Also needed is the occurrence of locations suitable for its preservation and concentration.