Sunteți pe pagina 1din 16

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/280079972

Extractive metallurgy of rhenium: A review


ARTICLE in MINERALS AND METALLURGICAL PROCESSING MARCH 2013
Impact Factor: 0.61

CITATION

READS

39

3 AUTHORS:
Caelen Anderson

Patrick R. Taylor

Haile Gold Mine

Colorado School of Mines

5 PUBLICATIONS 1 CITATION

25 PUBLICATIONS 128 CITATIONS

SEE PROFILE

SEE PROFILE

Corby Anderson
Colorado School of Mines
35 PUBLICATIONS 105 CITATIONS
SEE PROFILE

Available from: Caelen Anderson


Retrieved on: 01 October 2015

SPECIAL RARE-EARTH MINERALS ISSUE

Extractive metallurgy of rhenium:


a review
C.D. Anderson, P.R. Taylor and C.G. Anderson

PhD student, professor and professor, respectively, Kroll Institute for Extractive Metallurgy
Colorado School of Mines, Golden, CO

Abstract

A variety of processing technologies exist for recovery from both primary and secondary sources of rhenium.
Currently, there are no known primary rhenium deposits; thus, the method in which primary rhenium is produced
is dependent on the commodity of which it is a byproduct, e.g., copper, molybdenum, uranium, etc. In addition,
focus on the recovery of rhenium from secondary sources, such as alloy scraps and catalysts, is continually growing.
This paper presents a review of both primary and secondary processing technologies for the recovery of rhenium.
Minerals & Metallurgical Processing, 2013, Vol. 30, No. 1, pp. 59-73.
An official publication of the Society for Mining, Metallurgy, and Exploration, Inc.

Key words: Rare-earth minerals, Rhenium, Ammonium perrhenate, Extractive metallurgy

Introduction

lar, Chiles Molibdenos y Metales S.A. (Molymet), is the largest


molybdenum and rhenium producer in the world (Whittaker,
2012). The 2011 geographic breakdown of worldwide rhenium
reserves and production quantities are shown in Table 1.
Rhenium is typically sold on long-term contracts between
consumers and producers. Unfortunately, data available on
historical rhenium prices is mostly based on free market pricing and, thus, may not reflect the actual pricing of the rhenium
market to the desired extent. Nevertheless, Fig. 1 illustrates
a combination of historical free market rhenium prices from
a variety of sources, which demonstrate price fluctuations in
the market. Some of the major fluctuations may be attributed
to the following historical events (Polyak, 2011; Naumov,
2007; Blossom, 1998):

In recent times, the aerospace and petrochemical


industries have come to rely on a silver-white transition
metal, with a specific gravity of 21 and the secondhighest melting point (3,180 C) of any metal in the
periodic table. This metal is rhenium (Re), the last
natural element to be discovered, in 1925, by Mr. and
Mrs. Walter Noddack and Professor Otto Berg. The
element was named after the Rhine River of their native Germany (Millensifer, 2010). Rheniums unique
properties have made it a vital part of the superalloy
industry, most prominently in nickel superalloys used
in both aerospace and industrial gas-fired turbines. Its
second largest application is in the catalyst industry for
the production of unleaded gasoline (Polyak, 2011).
The concentration of rhenium in the earths crust
is rather small, at approximately 0.7-1.0 parts per
billion (Millensifer, 2010), although there is speculation that this number may be as high as 10 ppb
(Fleischer, 1959). The only documented occurrence
of rhenium as a mineral, rheniite (ReS2), was found
near the Russian Kudryavyi volcano (Korzhinsky
et al., 1994; Tessalina et al., 2008). Other than this
specific occurrence, rhenium is typically associated
with molybdenum-copper porphyry deposits in concentrations of up to 0.2% (Woolf, 1961). Currently,
there are no producers of primary rhenium, and almost
all rhenium, with few exceptions, is produced as a
byproduct of the molybdenum and copper industries.
In 2011, Chile produced the most rhenium (~25,000
kg) and contained the worlds largest rhenium reserve
(~1.3 M kg) (Polyak, 2011). One processor in particu-

1970: Use of rhenium in petrochemical catalysts begins.


1980: Amount of rhenium used in petrochemical catalysts doubles.
1991: Dissolution of U.S.S.R. leads to an increased supply of rhenium in the market.
1999, 2003, 2006: Appearance of new generation turbine
blades containing rhenium.
As of 2011, the U.S. Geological Survey (USGS) reported that
the free market price for 99.99% pure rhenium metal powder
is approximately $2,000/kg (Polyak, 2011).
Due to the lack of primary rhenium deposits, the method
with which it is processed is directly related to the method
with which the minerals it is associated with are produced,
typically copper and molybdenum. Characteristically, rhenium
found in mixed copper/molybdenum deposits is first separated
with the molybdenum from the copper, using conventional

Paper number MMP-12-077. Original manuscript submitted November 2012. Revised manuscript accepted for publication
December 2012. Discussion of this peer-reviewed and approved paper is invited and must be submitted to SME Publications
Dept. prior to August 31, 2013. Copyright 2013, Society for Mining, Metallurgy, and Exploration, Inc.
MINERALS & METALLURGICAL PROCESSING

59

Vol. 30 No. 1 February 2013

Figure 1 Historical rhenium metal power price from 1952-2012 (year 2012: Metal Pages 2012; years 2007-2011: Polyak,
2011; years 1999-2006: Naumov, 2007; years 1952-1998: Blossom, 1998).
Table 1 Worldwide rhenium reserves and production by
country. Source: Polyak, 2011.
Country

Reserves (kg)

Production (kg)

Chile

1,300,000

25,000

United States

390,000

6,000

Russia

310,000

1,500

Kazakhstan

190,000

2,500

Armenia

95,000

400

Peru

45,000

5,000

Canada

32,000

1,800

Poland

NA

4,500

Other countries

91,000

1,500

World total (rounded)

48,000

2,500,000

2ReS2 +7.5O2(g) = Re2O7(g)+ 4SO2(g)


(G298K = - 406.5 kcal)

It should be noted that rhenium heptmoxide is extremely


volatile (Pvap = 711 mmHg at 633K); thus, at the temperatures
used for molybdenum roasting (900-950 K) it is likely that
nearly all of the rhenium present is volatilized. This volatile
product exits the furnace with the flue gases and is subsequently
recovered as perrhenic acid (HReO4) after being scrubbed with
water via the following reaction.
Re2O7(g) + H2O = 2HReO4(aq)
(G298K = -15.16 kcal)

(2)

After scrubbing, the aqueous rhenium is typically recovered


using solvent extraction or ion-exchange processes. Generally, the end precursor product produced by these methods is
ammonium perrhenate (NH4ReO4) via crystallization (Millensifer, 2010).
In the following subsections, rhenium recovery methods
that rely on the pyrometallurgical volatilization of rhenium
and production of an ammonium perrhenate as an intermediate
product are discussed and broken into their respective branches
of concentration technology.

NA = not available

concentration technologies, typically froth flotation. After


this, rhenium is recovered as a byproduct from molybdenum
processing (Naumov, 2007).
The principal molybdenum end product, molybdenum trioxide (MoO3), is the basic raw material for most commercially
used products of molybdenum. Therefore, the method in which
rhenium is produced is dependent on the production method of
MoO3. Molybdenum trioxide is produced via pyrometallurgical roasting or hydrometallurgical pressure oxidation (Gupta,
1992; Ketcham et al., 2000).
In addition to being produced as a byproduct of the molybdenum/copper industry, focus on the recovery of rhenium
from secondary sources, such as alloy scraps and catalysts, is
continually growing.
Accordingly, the following sections are broken into both
the methods by which rhenium is recovered as a byproduct of
primary processing, as well as from secondary sources.

Ion exchange processes. The original Kennecott Process


for the recovery of rhenium involves the previously mentioned
scrubbing process, in which the exhaust gases of the molybdenum roasting circuit are washed with water to produce perrhenic acid. The pregnant solution is continually recirculated
through the scrubber circuit until the rhenium concentration
reaches approximately 100 mg/L. After this, the solution is
conditioned for 24 hours with caustic soda, soda ash and
oxidized with calcium hypochlorite. The pH of this solution
is brought up to 10 to precipitate any contaminants (primarily iron) remaining in solution and allow them to settle. The
solution is filtered and sent to an ion exchange circuit, where
the anionic resin preferentially adsorbs the aqueous rhenium
from the alkaline solution. After the loading stage is finished,
the rhenium is stripped by the addition of hydrochloric acid.
A caustic soda solution is then used to remove the remaining
adsorbed molybdenum, which is subsequently recovered as
calcium molybdate. Perchloric acid and hydrogen sulfide are
then added to this solution to precipitate rhenium as rhenium

Rhenium recovery from pyrometallurgical


effluent streams
In pyrometallurgical roasting of molybdenum concentrates
containing rhenium, rhenium present in the molybdenum
concentrate is oxidized to rhenium heptoxide (Re2O7) via the
following simplified reaction.
February 2013 Vol. 30 No. 1

(1)

60

MINERALS & METALLURGICAL PROCESSING

Figure 2 The original Kennecott Process (Sutulov, 1965).

sulfide (Re2S7). This precipitate is then redissolved in a solution


of ammonia and hydrogen peroxide, from which it crystallizes
as ammonium perrhenate (NH4ReO4). Figure 2 illustrates the
flowsheet for the original Kennecott Process.
Currently, KGHMs Glogow smelting facility (Glogow, Poland) is utilizing ion exchange technology (the Ecoren process)
for the recovery of aqueous rhenium, from copper concentrate
smelter flue gas. In this process, the rhenium within the copper
concentrate is volatilized as rhenium heptoxide and reports to
the sulfuric acid scrubbing plant. Like the Kennecott Process,
the rhenium heptoxide is scrubbed using water, and brought
into solution along with various other impurities, including
molybdate, sulfate and sulfite salts. The typical rhenium concentration in this solution is 0.02 g/L. This solution is sent
to a closed polypropylene filter press to remove any residual
silicon prior to being sent to the ion exchange columns. The
MINERALS & METALLURGICAL PROCESSING

solution is then sent to a series of ion exchange columns, where


rhenium is selectively adsorbed on the weakly basic anionic
resin as the perrhenate anion (ReO4-). After the loading phase,
the rhenium is eluted using an ammonia solution. The loaded
eluent is the sent to vacuum crystallization, where ammonium
perrhenate is produced. Typically, repeated recrystallization is
necessary to produce a high-grade (99.95% pure) ammonium
perrhenate product. The Glogow smelter facility has the capability to produce 4-5 t of ammonium perrhenate per year.
The Ecoren process flowsheet is shown in Fig. 3 (Chmielarz
and Litwinionek, 2010).
Solvent extraction processes. Some of the richest rheniumcontaining deposits in the world are found in Kazakhstan. In
the Zhezkhazgan deposit, copper concentrates produced can
contain up to 30 g/t of rhenium. Abisheva et al. have proposed
61

Vol. 30 No. 1 February 2013

Figure 3 KGHM Ecoren rhenium recovery at the Glogow smelter (Chmielarz and Litwinionek, 2010).

a process in which the copper concentrate is electrosmelted,


and the rhenium present is volatilized, as rhenium heptoxide
(Re2O7), and scrubbed with water/dilute sulfuric acid to form
perrhenic acid (HReO4) in concentrations of up to 0.25 g/L.
The aqueous rhenium solution is sent to a solvent extraction
step for the selective separation of rhenium from the impurity
elements present. The extractant used in this technology is trialklamine organic compound (TAA) in a kerosene diluent. The
loaded organic phase is stripped using ammonium hydroxide
to produce ammonium perrhenate. The barren organic phase
is then recycled for further use in solvent extraction, and the
impure ammonium perrhenate is continually dissolved and
recrystallized to produce a 98.5% pure ammonium perrhenate
product (Abisheva et al., 2011). A flowsheet of this process
is shown in Fig. 4.
The use of a tertiary amine rhenium extractant was conFebruary 2013 Vol. 30 No. 1

firmed by Singh et al., who ran a 250-L rhenium extraction


pilot plant using it. Results showed that 98% of the aqueous
rhenium was recoverable, and allowed for the production of a
98% pure rhenium precursor product using solvent extraction
(Singh et al., 1982).
As an alternative to the conventional method of ammonium
perrhenate precipitation, the USBM developed a solventextraction/electrowinning process for rhenium recovery from
molybdenite roasting. Once more, rhenium heptoxide present
in the flue gas was scrubbed using water. The aqueous rhenium
was oxidized using sodium chlorate and the pH of the solution was brought to 12 by caustic soda addition. Prior to the
solvent extraction step, the solution was filtered to remove any
precipitates that formed during pH adjustment. The pregnant
solution was then sent to a six-stage extraction circuit, which
used a combination of: 5% aliquat 336, 5% primary decyl
62

MINERALS & METALLURGICAL PROCESSING

Figure 4 Flowsheet for the recovery of rhenium from sulfuric acid scrubbing solution (Abisheva et al., 2011).

alcohol (PDA) and 90% kerosene as the extractant


phase. The loaded organic was then stripped using a
1-M perchloric acid/ammonium sulfate solution. After
stripping, the rhenium-rich electrolyte was sent for
electrowinning using a current density of 360 A/m2.
Results from the pilot plant experimentation show that
rhenium metal can be prepared from dilute impure solutions containing aqueous rhenium and molybdenum
(Churchward and Rosenbaum, 1963). An illustration
of the flowsheet for this process is shown in Fig. 5.

Alternative methods for rhenium recovery


In addition to the use of conventional ion exchange
or solvent extraction processes, there are a number of
processes that involve novel methods for the extraction
and production of rhenium. The following section
will describe each of these in their respective context.

Figure 5 USBM SX/EW process flowsheet for rhenium recovery


(Churchward and Rosenbaum, 1963).
MINERALS & METALLURGICAL PROCESSING

63

Recovery from Mo/Cu ores. As an alternative


to traditional roasting technology, Rio Tintos Kennecott facility at Bingham Canyon, UT has patented
(US# 6149883) an alkaline pressure oxidation process
for the recovery of molybdenum and rhenium from
molybdenite concentrates (molybdenum autoclave
process or MAP). In this process, the molybdenite
flotation concentrate is leached with either sodium
or potassium hydroxide at elevated temperature (150200 C) and pressure (517 - 1,400 kPa or 75-200 psig)
to form soluble molybdate (MoO42-).The aqueous
molybdate is then recovered using solvent extraction
and is stripped with an ammonium hydroxide eluent.
The rhenium present in the molybdenite concentrate
is primarily recovered in the solvent extraction step,
in the loaded strip solution. After this, it is recovered
using a selective ion exchange resin, which typically
contains quaternary amine functional groups (Ketcham
et al., 2000).
In 2009, Freeport-McMoRan applied for a patent
Vol. 30 No. 1 February 2013

Figure 6 Freeport pressure oxidation rhenium recovery process (Waterman et al., 2009).

(US# 0263490A1) for a proprietary method of recovering


rhenium as a byproduct of copper leaching/molybdenum oxidation. The application states that, the rhenium rich PLS can
originate from an active copper leach, stockpile copper leach,
acid blowdown stream, or a leach of molybdenite roaster flue
fumes and dusts. This loaded stream is sent to an activated
carbon column circuit for adsorption of the aqueous rhenium.
After adsorption, the loaded rhenium is stripped using an eluent solution containing approximately 2.5% sodium hydroxide
and 2.5% ammonium hydroxide at a pH of 7, and an operating temperature of 80-110 C. After elution, the rhenium-rich
stream is sent for rhenium recovery to produce a pure rhenium
product, and a rhenium lean eluate solution for optional reuse
in the circuit (Waterman et al., 2009). An example of Freeports
proposed rhenium recovery circuit is shown in Fig. 6.
In October 2012, the Australian Patent Office granted a
February 2013 Vol. 30 No. 1

patent (AUS# 2011229125) to Alexander Mining Plc. for the


MoReLeach process (Sutcliffe et al., 2012). Officially titled
Method of oxidative leaching of molybdenum-rhenium sulfide ores and/or concentrates, the process involves leaching
a molybdenum/rhenium concentrate in a closed reactor vessel
at atmospheric temperature and pressure. The lixiviant used is
an aqueous solution of chlorine-based oxidizing species, in
which the predominant chlorine-based oxidizing species are
hypochlorite (ClO-) ions. The molybdenum sulfide is oxidized
to the soluble molybdate anion (MoO42-), and the rhenium
present is oxidized to the perrhenate anion (ReO4-) via the
following proposed reactions (Gupta, 1992). Note: the second
reaction is proposed by the author.
MoS2 + 9NaClO + 3H2O = MoO42-(aq) + 9NaCl +
2SO42-(aq) + 6H+
64

(G298K = -506.9 kcal) (3)

MINERALS & METALLURGICAL PROCESSING

Figure 7 MoRe leach process (Sutcliffe et al., 2012).

ReS2+ 9.5NaClO + 2.5H2O = ReO4-(aq) + 9.5NaCl +


2SO42-(aq) + 5H+ (G298K = -535.7 kcal) (4)

sorbant, activated charcoal in the gold industry. But, in this


process, an ion exchange resin is added to a solution that has
already undergone leaching. Thus, after perrhenic acid is
produced by scrubbing with water, the ion exchange resin is
added to selectively remove the rhenium and molybdenum
from solution, via adsorption onto the resin. Experimental
results show that by using a strong basic anionic exchange
resin, rhenium can be selectively adsorbed from solution. Additionally, elution results have shown that molybdenum can be
selectively eluted from the resin by using ammonium chloride
(NH4Cl) as the eluent, leaving rhenium adsorbed. After this,
rhenium can then be stripped from the resin using dilute nitric
acid (Lan et al., 2006).
In 1947, Melaven and Bacon, of the University of Tennessee, patented (US# 2,414,965) a technology for the recovery of
rhenium from molybdenite roasting flue dust. In their process,
cyclone flue dust was recovered, and rhenium was leached
with water and compressed air. After lixiviation, potassium
chloride was added to the solution to recover rhenium as a pure
potassium perrhenate (KReO4) precipitate. Metallic rhenium
was then produced by roasting the potassium perrhenate in
a silver tube under a hydrogen atmosphere at 350 C. After
roasting, the residue is washed with hot distilled water, leaving a pure metallic rhenium product (Melaven and Bacon,

After dissolution, the pregnant solution is separated from


the undissolved residue and sent for a metal separation stage
such as solvent extraction, ion exchange, etc. Additionally, the
proposed process allows for the regeneration of hypochlorite
for subsequent reuse as a reagent. A flowsheet of the proposed
process is illustrated in Fig. 7.
In the mid-1990s, after discovering rheniite (ReS2) at the
Russian Kudryavyi volcano, Russian researchers found that the
high-temperature fumarole gases exiting the volcano contained
considerable (0.5-2.5 g/t) of rhenium in the form of gaseous
rhenium chlorides (ReCl5) and fluorides (ReF5). Currently,
researchers are working on using zeolites as the adsorption
medium for the rhenium rich fumarole gases. As of 2007, researchers are operating this process on a pilot scale, although
this process may prove rather difficult, due to the inherent
danger of the natural environment and lack of infrastructure
within the area (Naumov, 2007; Sinegribov et al., 2007).
Lan et al. have proposed a novel process for the recovery of
rhenium-containing flue dusts, using an application commonly
used in gold hydrometallurgy, known as the resin-in-pulp
process. The resin-in-pulp process relies on the addition of a
MINERALS & METALLURGICAL PROCESSING

65

Vol. 30 No. 1 February 2013

Figure 8 The Melaven process (Melaven and Bacon, 1947).

1947; Sutulov, 1965). The flowsheet for the Melaven process


is shown in Fig. 8.

The proposed flowsheet for this process is shown in Fig. 9.


In another case where rhenium is found in uranium leaching solutions, Chekmarev et al. have proposed the recovery
of rhenium using complexation and ultrafiltration methods
with the aid of water-soluble polyelectrolytes. Rhenium in the
pregnant leach solution is complexed using VA-type cationic
polyelectrolytes containing quaternary ammonium base groups.
These complexes are sent through ultrafiltration to selectively
remove the high molecular weight rhenium complex, leaving
a wash solution ready for subsequent rhenium processing
(Chekmarev et al., 2004).

Recovery from uranium leach liquors. After finding


rhenium adsorbed on their uranium ion-exchange columns in
Palangana, TX, efforts were made for the recovery of rhenium
as rhenium sulfide. Rhenium was present in the ammonium
carbonate leach liquor as the perrhenate anion, ReO4-, which
is selectively adsorbed on the anionic exchange resin. The
proposed process calls for the elution of this anion by way of
ammonium nitrate and precipitation of rhenium sulfide (Re2S7)
with the addition of hydrogen sulfide gas (Goddard, 1996).
February 2013 Vol. 30 No. 1

66

MINERALS & METALLURGICAL PROCESSING

Figure 9 Recovery of rhenium sulfide (Goddard, 1996).

Recycling of rhenium

high temperature properties of tungsten and rhenium, it is not


uncommon to find these metals in alloys together. Thus, W-Re
scrap may be recycled via an oxidative pyrometallurgical
roasting technique. Initially, the scrap is roasted at 1,000 C,
under an oxidizing atmosphere to produce rhenium heptoxide
(Re2O7), which is subsequently condensed in the cooler part
of the tube furnace (Fig. 10).
This material is then sent for digestion in water. The aque-

In addition to being produced as a byproduct of molybdenum, it is possible to recycle rhenium during processing, and
after its use within the industry. The following subsections will
illustrate a number of the current and proposed technologies
that might be used to recover rhenium from secondary sources.
Recycling from superalloys and alloy scraps. Due to the

Figure 10 Schematic of the W-Re tube furnace (Heshmatpour and McDonald, 1982).

MINERALS & METALLURGICAL PROCESSING

67

Vol. 30 No. 1 February 2013

Figure 11 The H.C. Starck process for superalloy recycling (Olbrich et al., 2009).

ous rhenium (ReO4-) is subsequently precipitated as potassium


perrhenate upon the addition of potassium chloride via the
following reaction.
KCl + ReO4- = KReO4 + Cl-(aq)
(G298K = -6.133 kcal)

Co, Ni, Fe, Mn and Cr from the leach liquor. Magnetic separation is then applied to the insoluble components for further
separation and concentration. The pregnant leach solution is
sent to an ion exchange step, where the aqueous rhenium is
selectively adsorbed and can be recovered using the methods
described in the previous sections (Olbrich et al., 2009). An
example of the flowsheet for this process is shown in Fig. 11.
Through the use of electrolytic decomposition of rhenium
superalloys, Stoller et al. have patented (US# 0110767) a
process that involves the use of titanium baskets as electrodes.
The baskets containing the superalloy scrap are fed to a polypropylene electrolysis cell containing a 18% HCl solution.
The electrolytic dissolution is carried out for 25 hours at a
frequency of 0.5 Hz, current of 50 amps, voltage of 3-4 V and
a temperature of 70 C. The remaining scrap is then filtered
from the pregnant solution and sent for further dissolution in
sodium hydroxide/peroxide solution. After completion, this
filtrate is sent to ion exchange for the recovery of rhenium and
molybdenum. Rhenium is recovered using the ion exchange
processes discussed previously (Stoller et al., 2008). An illustration of this process is shown in Fig. 12.

(5)

The potassium perrhenate is filtered and further purified


via continued dissolution and recrystallization. After purification, the salt is dried and sent for reduction under a hydrogen
atmosphere at approximately 350 C. Experimental results
show that 93.1% of the rhenium was recovered to produce a
99.98% pure Re product (Heshmatpour and McDonald, 1982).
H.C. Starck has applied for a patent (US # 0255372 A1) for
a process for the elevated temperature digestion and recycling
of rhenium-containing superalloys. Initially, the superalloys
are digested in a molten salt melt containing NaOH, Na2CO3,
and Na2SO4 at temperatures of 850-1,100 C in a directly fired
rotary kiln. In addition to this, oxidizing agents such as nitrates
and peroxides of the alkali metals are added. The melt from this
process is then cooled and sent to a comminution process for
size reduction. The material is then leached using water as the
lixiviant to dissolve the 6 and 7th group elements present in the
superalloy. This slurry is then filtered to separate the insoluble
February 2013 Vol. 30 No. 1

Recycling of spent Pt-Re catalysts


Petroleum-reforming catalysts containing rhenium and plati68

MINERALS & METALLURGICAL PROCESSING

Figure 12 Electrochemical method for recycling of superalloys (Stoller et al., 2008).

num on an alumina substrate are used in the refining industry


for the improvement of the octane level of fuels. After being
deactivated, an effective method for the recovery or rhenium
and other PGM metals is necessary. There are two basic methods by which this is achieved (Kasikov and Petrova, 2009):

solution, and cooled to allow for crystallization of ammonium


perrhenate. After continued redissolution and recrystallization,
a high-purity ammonium perrhenate precipitate is produced
(El Guindy, 1997). The flowsheet for this process is shown
in Fig. 13.
As an alternative to sulfuric acid, sodium bicarbonate may
also be used as a lixiviant. The proposed advantage of this
process is the complete removal of the ion exchange circuit
shown in Fig. 14. Experiments were performed on crushed
and uncrushed catalysts in both packed columns and agitated
leach vessels. Experimental results showed that that rhenium
is preferentially leached in the sodium bicarbonate solution.
Rhenium recovery reached 97% for crushed catalysts, and 87%
for uncrushed catalyst samples. After dissolution, the aqueous rhenium is crystallized via evaporative crystallization as
an ammonium perrhenate intermediate product (Angelidis et
al., 1999). The flowsheet for this process is shown in Fig. 14.

1. Complete dissolution of the alumina substrate.


2. Selective dissolution and recovery of rhenium and
platinum.
The following summaries will present examples of both of
these technologies; for further information, refer to Kasikovs
literature review Processing of deactivated platinum-rhenium
catalysts (Kasikov and Petrova, 2009).
Complete dissolution of the alumina substrate. During
complete dissolution of the alumina substrate, sulfuric acid
may be used for dissolution of alumina, rhenium and, to some
extent, platinum. The rhenium-rich solution is separated from
the platinum-containing residue and aqueous aluminum using ion exchange. Rhenium is eluted from the organic amine
resin by way of hydrochloric acid addition. After elution, the
rhenium-rich eluate is neutralized using ammonium hydroxide. This solution is then evaporated to form a super-saturated
MINERALS & METALLURGICAL PROCESSING

Selective leaching of rhenium and platinum. The methods


used to selectively recover platinum and rhenium from spent
catalysts without completely dissolving the alumina substrate
vary from calcination of the catalysts to selective leaching in
alkaline or acid conditions at ambient and elevated temperatures
(Kasikov and Petrova, 2009).
69

Vol. 30 No. 1 February 2013

Figure 13 Method for rhenium recovery from spent reforming catalysts (El Guindy, 1997).

February 2013 Vol. 30 No. 1

70

MINERALS & METALLURGICAL PROCESSING

Figure 14 - Rhenium recovery using sodium bicarbonate (Angelidis et al., 1999).

By calcining the catalyst at temperatures up to 1,150 C, the


-Al2O3 undergoes a phase transition to the chemically stable
-Al2O3 phase, lowering the dissolution of the alumina catalyst.
The platinum and rhenium can then be selectively leached
in concentrated (5 mol/L) sulfuric acid solutions containing
sodium chloride and a potassium persulfate oxidant (K2S2O8).
Kpumaneva reports that rhenium and platinum recoveries are as
high as 95.5% and 97%, respectively (Kpumaneva et al., 2001).
Additionally, a U.S. patent (US#: 5542957) has been granted
involving the selective leaching of platinum and rhenium at
elevated temperatures (50-300 C) and pressures (207-9,000 kPa
or 30-1,300 psig). In this process, a dilute solution of sulfuric
acid (0.001-1.0 mol/L) is used in the presence of ammonium

MINERALS & METALLURGICAL PROCESSING

iodide or bromide and oxygen to selectively leach Pt and Re,


while leaving behind the alumina substrate. At 160 C and an
oxygen overpressure of 800 kPa (116 psig), the authors report
a rhenium recovery of 98% (Han and Meng, 1996).

Production of metallic rhenium


Generally, metallic rhenium is not produced at the facility
at which it is concentrated and separated from other elements.
Instead, it is produced from ammonium perrhenate (APR)
using methods similar to those used in the molybdenum and
tungsten industries; i.e., a reductant such as carbon monoxide
or hydrogen is used. Although APR is the typical precursor
material that is reduced, potassium perrhenate may also be

71

Vol. 30 No. 1 February 2013

Figure 15 Production of rhenium from ammonium perrhenate (Millensifer, 2010).

used (Sutulov, 1965).


The two primary methods for the production of metallic
rhenium are:

two hours, and then the temperature is raised to 500 C for


an additional two hours. The reduction product is then cooled
slowly in an inert atmosphere to prevent oxidation and washed
with water to remove any residual alkaline material (Hurd
and Brimm, 1939). The proposed reaction for this technique
is shown below.

1. The reduction of ammonium perrhenate (NH4ReO4).


2. The reduction of potassium perrhenate (KReO4).

2KReO4 + 7H2 = 2Re + 2KOH + 6H2O


(G298K = -16.07 kcal)

Ammonium perrhenate (NH4ReO4) is the typical precursor


product used in the production of metallic rhenium and rhenium
compounds, including metallic rhenium powder and perrhenic
acid. Metallic rhenium is produced by being reduced by hydrogen gas at elevated temperature, T = 1,000 C (Hurd and
Brimm, 1939). Ammonium perrhenate (APR) is placed in boats
and subjected to countercurrent hydrogen gas flow. Depending
on the particle size of metallic rhenium powder product, the
reduction may be completed in single or multiple stages and
the APR may be ground prior to reduction (Millensifer, 2010).
The proposed reaction for this process is shown below. Figure
15 illustrates the typical production steps in making metallic
rhenium from ammonium perrhenate.

Conclusion
Most of the processes involved in the production of primary
and secondary rhenium involve the use of either elevated
temperatures, elevated pressures, large amounts of reagents
or a combination of the three. Thus, it is imperative that the
extraction and recovery of rhenium is as efficient as economically possible. Some possible research opportunities for the
development of higher efficiency processes are suggested in
the following paragraphs.
The roasting of molybdenum concentrates containing
rhenium is an elevated-temperature, exothermic process that
requires temperatures of 900-950 K. Early scrubbing attempts
of flue gases were relatively inefficient, capturing roughly 25%
of the rhenium present. Through innovation in the scrubbing
equipment used, recoveries have now increased to approximately 80% (Millensifer, 2010). Thus, increasing the recovery
of rhenium from roasting flue dusts presents itself as a viable
research opportunity.

2NH4ReO4(s) + 7H2 = 2Re + 2NH3 + 8H2O


(G298K = -37.16 kcal) (6)
In Hurd and Brimms process for the commercial reduction
of potassium perrhenate, the feed material is crushed to approximately 60 US mesh and dried at 175 C. The material is
then placed in a silver boat inside of a refractory tube furnace.
The boat is heated to 250 C under a hydrogen atmosphere for
February 2013 Vol. 30 No. 1

(7)

72

MINERALS & METALLURGICAL PROCESSING

B.D. Bryskin, ed., February 9-13, 1997, Orlando, FL, TMS, pp 89-97.
Fleischer, M., 1959, The geochemistry of rhenium, with special reference to
its occurrence in molybdenite, Economic Geology, Vol. 54, pp. 1406-1413.
Goddard, J.B., 1984, Recovery of rhenium from uranium in-situ leach liquor,
Society of Mining Engineers Transactions of AIME, Vol. 274, pp 1996-2000.
Gupta, C.K., 1992, Extractive Metallurgy of Molybdenum, 1st Edition, CRC
Press, pp 164-165.
Han, K.N., and Meng, X., 1996, Recovery of platinum group metals and rhenium from materials using halogen reagents, US Patent No. 5,542,957,
Washington, D.C., U.S. Patent and Trademark Office.
Heshmatpour, B., and McDonald, R.E., 1982, Recovery and refining of rhenium,
tungsten and molybdenum from W-Re, Mo-Re and other alloy scraps,
Journal of the Less Common Metals, Vol. 86, pp 121-128.
Hurd, L., and Brimm, E., 1939, Metallic rhenium, Inorganic Syntheses, Vol.
1, pp 175-178.
Kasikov, L., and Petrova, A., 2009, Processing of deactivated platinum-rhenium
catalysts, Theoretical Foundations of Chemical Engineering, Vol. 43, No.4,
pp 544-552.
Ketcham, V.J., Coltrinari, E.J., and Hazen, W.W., 2000, Pressure oxidation
process for the production of molybdenum trioxide from molybdenite, US
Patent No. 6,149,883, Washington, D.C., U.S. Patent and Trademark Office.
Korzhinsky, M.A., Tkachenko, S.I., Shmulovich, K.I., Taran, Y.A., and Steinberg,
G.S., 1994, Discovery of a pure rhenium mineral at Kudriavy volcano,
Nature, Vol. 369, No. 6475, pp 51-52.
Kpumaneva,, E., Rtveadze, V., and Igumnov, S., 2001, Platinum and rhenium
extraction with base material dilution, Khimicheskaya Technologiya, Vol.
5, pp 12-17.
Lan, X., Liang, S., and Song, Y., 2006, Recovery of rhenium from molybdenite
calcine by a resin-in-pulp process, Hydrometallurgy, Vol. 82, No. 3-4, pp.
133-136.
Melaven, A.D., and Bacon, J.A., 1947, Process for recovering rhenium, US
Patent No. 2,414,965, Washington, D.C., U.S. Patent and Trademark Office.
Millensifer, T.A., 2010, Rhenium and rhenium compounds, Kirk Othmer Encyclopedia of Chemical Technology, John Wiley and Sons, pp. 1-21.
Naumov, A., 2007, Rhythms of rhenium, Russian Journal of Non-Ferrous
Metals, Vol. 48, No. 6, pp. 418-423.
Olbrich, A., Meese-Marktscheffel, J., Jahn, M., Zertani, R., Stoller, V., Erb, M.,
Heine, K.H., and Kutzler, U., 2009, Recycling of superalloys with the aid of
an alkali metal salt bath, US Patent Application No. 0255372 A1,Washington,
D.C., U.S. Patent and Trademark Office.
Polyak, D.E., 2011, USGS Mineral Commodity Summary: Rhenium, Reston, VA:
U.S. Geological Survey.
Sinegribov, V., Sotskov, K., and Yudin. A., 2007, Recovery of valuable elements
from fumarole gases, Theoretical Foundations of Chemical Engineering,
Vol. 41, No. 5, pp. 593-598.
Singh, H., Rao, M.S., Rao, G.V., and Reddy, M.V., 1982, Development of a solvent extraction process for recovery of rhenium from impure scrub liquors,
SME Preprint, No. 82-99.
Stoller, V., Olbrich, A., Meese-Marktscheffel, J., Mathy, W., Erb, M., Nietfeld,
G., and Gille, G., 2008, Process for electrochemical decomposition of
superalloys, US Patent Application No. 0110767 A1, Washington, D.C.,
U.S. Patent and Trademark Office.
Sutcliffe, M.L., Johnston, G.M., and Welham, N.J., 2012. Method of oxidative
leaching of molybdenum: rhenium sulfide ores and/or concentrates, Australian Patent No. 2011229125, Canberra, Australia, Australian Patent Office.
Sutulov, A., 1965, Molybdenum Extractive Metallurgy, University of Concepcion,
pp. 167.
Tessalina, S.G., Yudovskaya, M.A., Chaplygin, I.V., Birck, J.L., and Capmas, F.,
2008, Sources of unique rhenium enrichment in fumaroles and sulphides
at kudryavy volcano, Geochimica Et Cosmochimica Acta, Vol. 72, No. 3,
pp. 889-909.
Waterman, B.T., Dixon, S.N., Morelli, T.L., Owusu, G., and Ormsby, S.T., Methods
and systems for recovering rhenium from a copper leach, US Patent Application No. 0263490 A1, Washington, D.C., U.S. Patent and Trademark Office.
Whittaker, P., 2012, Molymet to invest $390 million in us based rare earth
producer Molycorp, website article accessed September 18, 2012, http://
www.ipmd.net/news/001630.html.
Woolf, A.A., 1961, An outline of rhenium chemistry, Chemical Society: Quarterly
Reviews, Vol. 15, No. 3, pp. 372-391.

Conversely, another option for increased rhenium recovery


may involve the use of hydrometallurgical pressure oxidation
of molybdenum concentrates. Although this process involves
the use of elevated temperatures/pressures and additional reagents, the enhanced recoveries inherent to the process may
make this a viable alternative to traditional roasting processes.
Research of this technique, and other technologies involving
the use of low temperature hydrometallurgical oxidation is
currently ongoing in industry.
As a complementary process to both roasting and pressure
oxidation, the need for efficient separation of aqueous rhenium
from the process stream is essential. A variety of techniques
have been the subject of investigation, including ion exchange,
solvent extraction and activated carbon. With the continual
development of higher selectivity resins and extractants used
in both ion exchange and solvent extraction, the area presents
itself worthy of further research and development.
Relatively new to the rhenium industry is the processing of
rhenium-laden manufacturing scrap and end-of-life materials,
such as catalysts and super alloys. Since this is such a new
faction of the rhenium industry, the research possibilities are
almost endless, but should definitely include the utilization of
primary processing techniques, as well as the implementation
of end-of-life recycling programs.
As rhenium sources are depleted, the need for efficient and
economical extraction from both primary and secondary sources
is essential in maintaining the rhenium supply. This paper has
presented a review of both the current and former technologies used in the extractive metallurgy of rhenium, as well as
provided some areas for potential technological development.

Acknowledgments
Special recognition is due to Tom Millensifer for his assistance with this paper; in addition, the authors thank the Office
of Naval Research for the financial support that allowed this
research to occur.

References

Abisheva, Z.S., Zagorodnyaya, A.N., and Bekturganov, N.S., 2011, Review of


technologies for rhenium recovery from mineral raw materials in Kazakhstan,
Hydrometallurgy, Vol. 109, No. 1-2, pp. 1-8.
Angelidis, T.N., Rosopoulou, D.D., and Tzitzios, V., 1999, Selective rhenium
recovery from spent reforming catalysts, Industrial & Engineering Chemistry
Research, Vol. 38, No. 5, pp. 1830-1836.
Blossom, J.W., 1998, Rhenium: USGS Report on Prices, Reston, VA, U.S.
Geological Survey.
Chekmarev, A.M., Troshkina, I.D., Nesterov, Y.V., Maiboroda, A.B., Ushanova,
O.N., and Smirnov, N.S., 2004, Associated rhenium extraction in complex
processing of productive solutions of underground uranium leaching,
Chemistry for Sustainable Development, Vol. 12, pp. 113-117.
Chmielarz, A., and Litwinionek, K., 2010, Development of the technology for the
recovery of rhenium from polish copper smelters, Copper 2010, Hamburg,
DE, pp. 1803-1814.
Churchward, P.E., and Rosenbaum, J.B., 1963, Sources and recovery methods
for rhenium, United States Bureau of Mines, Report of Investigations: 6246.
El Guindy, M.I., 1997, Processing of spent platinum rhenium catalyst for rhenium
recovery, Proceedings of the Rhenium and Rhenium Alloys Symposium,

MINERALS & METALLURGICAL PROCESSING

73

Vol. 30 No. 1 February 2013

S-ar putea să vă placă și