You are on page 1of 9

Hydrometallurgy 147148 (2014) 5967

Contents lists available at ScienceDirect

Hydrometallurgy
journal homepage: www.elsevier.com/locate/hydromet

Zn-vapor pretreatment for acid leaching of platinum group metals from


automotive catalytic converters
Hideaki Sasaki , Masafumi Maeda
International Research Center for Sustainable Materials, Institute of Industrial Science, The University of Tokyo, 4-6-1 Komaba, Meguro-Ku, Tokyo 153-8505, Japan

a r t i c l e

i n f o

Article history:
Received 6 December 2013
Received in revised form 21 April 2014
Accepted 23 April 2014
Available online 4 May 2014
Keywords:
Precious metals
Automotive catalysts
Hydrometallurgy
Acid leaching

a b s t r a c t
To recover platinum group metals (PGMs) from spent automotive catalytic converters with a smaller environmental load, a Zn-vapor treatment prior to acid leaching was previously proposed. Because PGMs and Zn form
alloys which dissolve easily, the pretreatment is believed to reduce the use of both chemical agents and processing time. In this study, effects of this treatment on PGM leaching were examined using simulated catalysts.
Leaching tests in aqua regia demonstrated that Zn-vapor treatment enhanced the dissolution of PGMs, and
was especially effective on Rh. Pretreatments were also found to enhance PGM extractions in hydrochloric acid
depending on the conditions. Reactions on Zn-treated catalysts in the solution are discussed based on these
results.
2014 Elsevier B.V. All rights reserved.

Automotive catalysts account for large portions of the world consumption of platinum (Pt), palladium (Pd) and rhodium (Rh) (Butler,
2012). Because of their limited production and high prices, the recovery
of these platinum group metals (PGMs) is an important issue. A representative structure of an automotive catalytic converter is shown
in Fig. 1 (Gandhi et al., 2003; Heck and Farrauto, 2001; Jimenez de
Aberasturi et al., 2011; Matsumoto and Shinjoh, 2008). A monolithic
ceramic substrate with a honeycomb structure is coated with porous
oxides such as alumina (Al2O3), ceria (CeO2) and zirconia (ZrO2), to
which PGM particles several nanometer thick adhere. The PGMs work
as active sites for catalytic reactions such as the oxidation of carbon
monoxide and hydrocarbon. Porous oxides have a role in dispersing
the PGMs and maintaining their large surface area. CeO2 also acts as a
promoter of the catalytic reactions by storing and releasing oxygen
(Reaction (1)) depending on the situation (Gandhi et al., 2003; Heck
and Farrauto, 2001; Matsumoto and Shinjoh, 2008).

PGMs (Habashi, 1997; Suzuki et al., 2007; Yoo, 1998). In pyrometallurgical processes, spent catalysts are melted with metals such as copper
(Cu) to separate PGMs and ceramics into the metal phase and slag
phase, respectively. Smelting at high temperatures, however, requires
extensive equipment as well as a high amount of energy. In hydrometallurgical processes, PGMs are dissolved into an aqueous solution
and then recovered by various methods such as solvent extractions
(Marinho et al., 2010; Reddya et al., 2010) and ion-exchange methods
(Kononova et al., 2011; Sun et al., 2012), which continue to be under intensive investigation. Although the hydrometallurgical processes might
be effective in separating PGMs from ceramics because of the selectivity
of chemical agents on the metals, it requires strong acid and oxidizing
agents to dissolve PGMs. Dissolutions of Pt, Pd and Rh in an aqueous solution containing chloride ions are represented by Reactions (2)(4)
(Jimenez de Aberasturi et al., 2011). The standard potentials for the reactions, E, are high, and thus oxidizing agents such as nitric acid, Cl2 or
H2O2 are required (Habashi, 1997; Jimenez de Aberasturi et al., 2011;
Yoo, 1998).

CeO2 CeO2x x=2O2

Pt 6Cl PtCl6 

1. Introduction

Solid solutions of CeO2 and ZrO2 are widely used to improve the oxygen storage property (Gandhi et al., 2003; Matsumoto and Shinjoh,
2008).
Pyrometallurgical and hydrometallurgical processes combined with
physical separation have conventionally been used for the recovery of

Corresponding author. Tel.: +81 3 5452 6298; fax: +81 3 5452 6299.
E-mail address: hideakis@iis.u-tokyo.ac.jp (H. Sasaki).

http://dx.doi.org/10.1016/j.hydromet.2014.04.019
0304-386X/ 2014 Elsevier B.V. All rights reserved.

4e

E B 0:62V

E B 0:44V

Pd 4Cl PdCl4 

2e

3e

Rh 6Cl RhCl6 

E B 0:74V

The dissolution of Pt in aqua regia, for example, is represented by


Reaction (5) (Habashi, 1997), in which nitrate ion is an oxidant. Mixing
of HNO3 and HCl generates NOCl and nascent Cl2 (Reaction (6)), and

60

H. Sasaki, M. Maeda / Hydrometallurgy 147148 (2014) 5967

Fig. 1. Representative structure of automotive catalytic converter and presumed recovery process in this study.

these are also believed to oxidize PGMs (Angelidis and Skouraki, 1996;
Baghalha et al., 2009; Habashi, 1997).
Pt 2NO3

8H 8Cl PtCl6 

HNO3 3HClNOCl Cl2 2H2 O

4H2 O 2NOCl

5
6

Therefore, waste solutions contain toxic reagents and heavy metals,


and their disposal involves a high cost and large environmental load. In
addition, dissolutions of some PGMs, especially Rh, are slow, and require
higher temperatures (Habashi, 1997). Methods to dissolve PGMs are
thus studied using various leaching solutions as reviewed by Jha et al.
(2013). Extractions by pressure leaching (Chen and Huang, 2006), microwave (Jafarifara et al., 2005) or supercritical CO2 (Iwao et al., 2007)
were also examined, and a chlorination of PGMs prior to dissolution
has recently been proposed (Horike et al., 2012).
As a pretreatment for acid leaching of PGMs, the exposure to a metal
vapor was previously proposed (Kayanuma et al., 2004a,b; Okabe et al.,
2003a,b). Because PGM dissolution is promoted by alloying (Habashi,
1997), the pretreatment is believed to reduce the use of chemical agents
and environmental load involved in the recovery. Compared to other
metals such as calcium (Ca) and magnesium (Mg), vapor of zinc (Zn)
seems more appropriate because of its ease in handling. Moreover, Zn
might react with PGMs selectively, whereas Ca and Mg form compounds
with catalytic substrates containing Al2O3 and SiO2 (Kayanuma et al.,
2004a). Thus this study assumes the PGM recovery process shown in
Fig. 1. In previous research, the authors examined the dissolution of
PGMZn alloys in an acid solution using an electrochemical method to
observe the enhanced dissolution of PGMs quantitatively and discussed
it physicochemically (Sasaki and Maeda, 2012; Sasaki and Maeda, 2013;
Sasaki et al., 2010a). One possible reason for the enhancement is the
increase in surface area of PGMs caused by the preferential dissolution
of Zn from the alloys (Kayanuma et al., 2004a; Okabe et al., 2003a,b;
Sasaki and Maeda, 2012; Sasaki and Maeda, 2013; Sasaki et al., 2010a).
Furthermore, PGMs may be changed into chemically reactive states in

the process of the alloy dissolution (Habashi, 1997; Sasaki and Maeda,
2013).
The purpose of this study was to examine the effect of Zn-vapor
treatment on PGM extraction from catalysts. The use of actually spent
catalysts, however, presents the following difculties. Firstly, there are
many kinds of catalytic converters used industrially, and detailed information on their composition and structure is not available to the public.
Secondly, PGMs are distributed inhomogeneously in converters to optimize catalytic reactions (Lucena and Laserna, 2001). In addition, characteristics of spent catalysts depend on their usage environment, so that
PGM particle sizes and the degree of their contamination vary. To conduct a systematic investigation, therefore, this study used simulated catalysts specically prepared for the leaching tests.

Fig. 2. Experimental ow chart and appearance of the simulated catalyst prepared for this
study.

H. Sasaki, M. Maeda / Hydrometallurgy 147148 (2014) 5967

2. Experimental

61

PGMs, and PGMZn alloys were formed depending on the reaction


condition.

2.1. Procedure of leaching tests on simulated catalysts


The experimental ow is shown in Fig. 2. A number of simulated catalysts of a uniform composition were prepared with the cooperation of
Toyota Motor Corporation. Monolithic grid-like honeycomb substrates
were 30 mm in diameter, 50 mm in length, and about 19 g before coating. Each substrate was coated with about 5 g of a mixture containing a
small percent of PGMs (Pt, Pd, Rh), and oxides (CeO2, ZrO2, Al2O3, and
others) with a predened ratio. Chemical analysis of PGM contents in
the catalysts is explained in Section 2.2.
Some of the simulated catalysts were cut into cuboid shaped, 8 9
20 mm test pieces of about 1 g, and regarded as unused catalysts.
Spent catalysts were replicated by heating the simulated catalysts at
1100 C for 5 h in air. The coating layer, which was brown originally,
became gray with the heating. The weight of each simulated catalyst decreased by about 0.1 g, indicating the evaporation or falling off of some
components or adsorbed substances. After the simulated spent catalysts
were cut into test pieces, some of them were subjected to deoxidizing or
Zn-vapor treatment followed by a leaching test. The methods are described in detail in Sections 2.32.5.
2.2. Chemical analysis of PGM contents in simulated catalysts
A representative composition of test pieces was determined
by chemical analysis including alkali fusion and tellurium coprecipitation (Etoh and Tokumori, 1986). The outer region of a simulated
catalyst was cut off, and the inner part was crushed into powder.
Then 0.36 g of the powder was fused with 3.6 g of Na2O2 in an alumina crucible. The resulting mixture was dissolved by hydrochloric
acid (35%) and nitric acid (60%), and the solution then heated to remove
the nitric acid. After salt precipitated in the solution, it was again
dissolved by diluted hydrochloric acid. The undissolved residue,
which was assumed to be mainly SiO2, was removed by ltering. The
ltrate, which was about 200 mL, was mixed with 10 mL of solution containing tellurium (Te: 10 g/L). Then 30 mL of SnCl2 solution
(SnCl22H2O: 200 g/L) was added with heating. PGMs in the ltrate
were reduced by Sn2+, and coprecipitated with tellurium. The precipitated particles were gathered by ltration, and dissolved again by
aqua regia to determine their quantity by an inductively coupled plasma
atomic emission spectrometer (ICP-AES, SPS4000, Seiko Instruments
Inc., Japan).
Analysis revealed that PGM contents in the inner part were smaller
than predicted values, which were calculated from weights of the coating layer on the assumption that the substrate was coated homogeneously. Based on the cylindrical shape of the simulated catalysts, the
coating on the substrate might have been thicker near its periphery
than its inner part because the grid spacing was small. Representative
composition of the spent catalyst was also determined by the aforementioned chemical analysis, which revealed a 23% decrease of PGMs in
the heating.
2.3. Zn-vapor treatment
The Zn-vapor treatment was conducted in a similar manner to
earlier research (Sasaki and Maeda, 2012; Sasaki et al., 2010b), where
reactions between PGMs and Zn vapor were studied. Test pieces
were sealed in a quartz tube with Zn plates and a Ti sponge under
vacuum, and then heated in an electric furnace for a predened time.
Ti was used as a deoxidizing agent to stimulate the evaporation of Zn.
During the heating, the temperature of the Zn vapor source, T1, was
kept lower than any other part in the ampule in order to supply Zn of
a constant pressure. Test pieces were located a few centimeters away
from the Zn, and kept at temperature, T2, which was higher than T1 by
1520 C. Under the temperature gradient, Zn vapor reacted with

PGM ZngPGMZn alloy

The treatments were conducted under various conditions as shown


later in Table 1. Increase in the weight of each test piece was about
0.01 g, indicating that Zn vapor not only formed alloys with PGMs but
also reacted with oxides. From a thermodynamic viewpoint, Reaction (8) might have been possible (Barin, 1989; Huang et al., 2005).
CeO2 s xZngCeO2x s xZnOs

In a preliminary experiment, changes in PGMs caused by a Zn-vapor


treatment were observed using an actual spent catalyst obtained
from a recycling company in Japan. The spent catalyst was cut into
pieces, one of which was treated with Zn vapor for 1 h at T1 = 600 C
and T2 = 673 C. The coating layers of the pieces were observed by a
eld emission scanning electron microscope (FE-SEM, JSM-7001F, Jeol,
Japan) equipped with an energy dispersive X-ray spectrometer (EDX)
in cooperation with the Center for Nano Lithography & Analysis of The
University of Tokyo. Before the observation by FE-SEM, the pieces
were coated with ruthenium to make it electrically conductive. The observation results are shown in Section 3.1.
2.4. Deoxidizing
The deoxidizing treatments were conducted for comparative experiments to discuss the contribution of Zn to PGM extractions. Test pieces
were heated at 620 C for 1 h in a sealed tube with a Ti sponge heated at
600 C. Through the deoxidizing, the weight of the test pieces decreased
by about 0.01 g. Probable reactions were the reductions of CeO2 (Reaction (9)) and PGM oxides (Reaction (10)), which might have formed on
the catalytic surface.
CeO2 xTi CeO22x xTiO2

PtO2 Ti Pt TiO2

10

2.5. Leaching tests


After the various pretreatments, test pieces were immersed in 50 mL
of aqua regia or hydrochloric acid. The aqua regia was prepared by
mixing nitric acid (60%) and hydrochloric acid (35%, both acids from
Kanto Chemical Co., Inc., Japan) at a volume ratio of 1 to 3. Table 1
shows the experimental conditions. In Test I, extractions of PGMs in
aqua regia were examined at 60 C, and effects of the pretreatments
were discussed. Extractions in aqua regia at 17 C were examined in
Test II, where Zn-vapor treatments were conducted under various conditions. Leaching with diluted aqua regia was examined in Test III, and
the time variations of the extractions were also observed. Extractions
in hydrochloric acid (35%) were investigated in Test IV.
After immersion for a predened reaction time, the test piece was
removed from the solution, and rinsed carefully with water. The extraction solution was diluted with water, which included that used for rinsing the sample. After removing particles with a lter paper, the ltrate
was heated to vaporize nitrate ions and excessive hydrochloric acid,
which are undesirable in quantitative analysis of dissolving metals.
The resulting concentrate was diluted in a measuring ask with diluted
hydrochloric acid in order to adjust the content of each PGM below
40 ppm. The dissolved metals were evaluated by ICP-AES, and the extraction ratios were determined using the representative composition
of the test piece experimentally evaluated in Section 2.2.

62

H. Sasaki, M. Maeda / Hydrometallurgy 147148 (2014) 5967

Table 1
Experimental conditions.
#

Test I

Test II

Test III

Test IV

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26

Pretreatment

None (unused)
Heating in air (spent)
Deoxidizing
Zn-vapor
Unused
Spent
Deoxidizing
Zn-vapor
Zn-vapor
Zn-vapor
Zn-vapor
Spent
Zn-vapor
Spent
Spent
Zn-vapor
Zn-vapor
Spent
Deoxidizing
Zn-vapor
Unused
Spent
Deoxidizing
Zn-vapor
Zn-vapor
Zn-vapor

Leaching
T1/C

T2/C

Time/h

Extractant

T/C

Time/min

600

600
600
700
800

(Same as #10)

(Same as #10)
(Same as #10)

600

600
700
800

600

620
620
715
815

1
4
4
4

Aqua regia

60

30

Aqua regia

17

30

Diluted aqua regia

17

620

620
715
815

1
4
4

Hydrochloric acid

17

30
30
15
60
15
60
30

Hydrochloric acid

60

30

3. Results
3.1. Observation of a spent catalyst
A backscattered electron image of coating on an actual spent catalyst is shown in Fig. 3, in which white dots about of 100 nm can be
observed. These are believed to be PGM particles, and an EDX spectrum conrmed the existence of Pd and Rh. The observation conrmed a coarsening of PGM catalysts during use at high temperatures
(Matsumoto and Shinjoh, 2008).
Fig. 4 shows an image of the spent catalyst which was treated with
Zn vapor. Particles on the coating layer are larger than in Fig. 3, and
EDX analysis showed that Zn coexisted with Pd and Rh. The observation
demonstrated the reaction between the PGM catalysts and Zn vapor,
which was supplied through the gas phase.
3.2. Leaching tests on simulated catalysts
3.2.1. Leaching in aqua regia at 60 C (Test I)
Fig. 5(a) shows the extractions of PGMs after leaching in aqua regia
at 60 C for 30 min. Firstly, dissolutions of PGMs from an unused catalyst
(#1) were more difcult than from a simulated spent catalyst (#2).
Although further study is required to clarify the reason, the results indicated that PGMs might have been strongly bound or covered with oxide
or solvent in the coating layer before being used, and thus some heat
treatment would help the leaching of unused catalysts. Meanwhile,
Fig. 5(b) shows the extractions of oxides, which were lower after
heating than from the unused catalyst. The dissolution of the heattreated oxide is believed to be difcult due to its sintering at high temperatures (Graham et al., 1999).
Extractions of PGMs from spent catalysts were enhanced by
deoxidizing (#3), and that of Pt and Pd reached about 100%. (Here,
the excess of Pt extraction over 100% must have been caused by an
error involved in the experimental method, e.g. the PGM content in
each test piece estimated using a represented composition.) One possible reason for the enhanced extraction was a reduction of oxide layers
formed on PGM particles (Jimenez de Aberasturi et al, 2011). Because
surfaces of PGMs form oxides such as PtO2 and PdO in air (Barin,

1989; Glin and Primet, 2002) and the oxide layer makes PGMs passive
in a solution, their reduction to a metallic surface is assumed to enhance
the dissolution.
In addition, the deoxidization enhanced dissolutions of oxides as
shown by #3 in Fig. 5(b). Thus the shrinkage or dissolution of the oxides
also might have helped the extraction of PGMs by improving their exposure to the leaching solution. The effect of the deoxidizing on the dissolution of CeO2 will be discussed in Section 4.1.
Compared to deoxidizing (#3), Zn-vapor treatment (#4) was similarly effective for Pt and Pd, which attained the extraction of about
100%. The advantage of the treatment was obvious on Rh, of which
extraction was 60%. The alloying pretreatment is considered to be
more effective for Rh than Pt and Pd because its dissolution is intrinsically slow and passivation is likely to occur (Jimenez de Aberasturi
et al, 2011; Llopis et al., 1965).
3.2.2. Leaching in aqua regia at 17 C (Test II)
PGMs were extracted in aqua regia at 17 C in Test II, where Znvapor treatments were conducted at various temperatures (Table 1).
Fig. 6 shows the result; the extractions after leaching for 30 min were
generally lower at 17 C than at 60 C. For example, PGM extractions
from a spent catalyst (#6) were much lower compared to #2 in Test I
(Fig. 5). In regard to the effect of pretreatments, a similar tendency
was observed at 17 C and 60 C, i.e., deoxidizing (#7) and Zn-vapor
treatments (#8#11) enhanced the dissolution of PGMs. Extractions
of Pt and Pd from the Zn-treated catalysts were over 90%, and that of
Rh was about 60% except for #8. These achievements were comparable
to leaching at 60 C. In addition, extractions of PGMs were higher in #9
than #8, indicating that the prolonged Zn-vapor treatment was effective. The extraction of Rh was further improved in #10 and #11,
where Zn-vapor treatment was conducted at higher temperatures.
3.2.3. Time dependency of extraction in diluted aqua regia (Test III)
The extraction of PGMs was examined in diluted aqua regia, which
was prepared by mixing aqua regia with an equal volume of water. As
shown in Fig. 7, the Zn-vapor treatment promoted PGM extractions in
the diluted aqua regia (#12, #13) as was the case of undiluted aqua
regia (#6, #10). Furthermore, while the dilution slightly decreased

H. Sasaki, M. Maeda / Hydrometallurgy 147148 (2014) 5967

63

Fig. 3. Scanning electron micrograph of spent catalyst and EDX spectrum obtained from a region assumed to be a PGM particle.

PGM extractions without the pretreatment (#6, #12), leaching tests on


Zn-treated catalysts (#10, #13) were not affected negatively by the
dilution.
Time variations of the extraction were examined in the diluted aqua
regia by immersing untreated catalysts (#14, #15) and Zn-treated catalysts (#16, #17) for 15 or 60 min. As shown in Fig. 8, extractions of Pt
and Pd from spent catalysts were about 50% after 30 min without any
pretreatment, and then leveled off. An extraction of Rh was about 20%,
and it also plateaued. The uncompleted dissolution of PGMs might
have been caused by encapsulation, i.e. a part of the PGMs might
have been embedded in insoluble oxides (Graham et al., 1999). On the
other hand, extractions of Pt and Pd from Zn-treated catalysts reached
almost 100% in 15 min, and that of Rh reached 65%. These results demonstrated that Zn-vapor treatment on PGMs not only accelerated their
dissolutions but also improved eventual extractions.
3.2.4. Dissolution in hydrochloric acid (Test IV)
Extractions of PGMs in hydrochloric acid are shown in Fig. 9(a). At
17 C, extractions of Pt and Pd from the spent catalyst were lower
than 10%, and Rh was barely dissolved without pretreatment (#18).
The extractions were enhanced by deoxidizing (#19), however, Znvapor treatment inhibited dissolutions of Pt and Pd (#20). Dissolutions
of PGMs were more prominent at 60 C, and about 90% of Pt and Pd
were extracted from a spent catalyst without any pretreatment (#22).

Deoxidizing (#23) and Zn-vapor treatment at 600 and 700 C (#24,


#25) were effective in enhancing the dissolution of PGMs, especially
Rh. In contrast, Zn-vapor treatment at 800 C (#26) inhibited Pt and
Pd dissolutions.
Fig. 9(b) shows signicant dissolutions of Ce and Zr in hydrochloric
acid at 60 C. About 20% of Ce dissolved from spent catalyst (#22)
while only a few percent of Ce dissolved in aqua regia (Fig. 5(b)). The
dissolutions were notably promoted by Zn-vapor treatments as shown
by #24#26.
4. Discussion
4.1. Dissolution of oxides from catalysts
It is known that CeO2 is difcult to dissolve even in a strong acid and
at elevated temperature, in spite of the Pourbaix diagram showing that
Ce4+ becomes stable as hydrated ions such as Ce(OH)2+
under oxidiz2
ing conditions (Hayes et al., 2002; Yu et al., 2006). Some research has
shown reductive dissolutions of CeO2 by the formation of soluble Ce3+
(Virot et al., 2012). For example, CeO2 might be dissolved in hot concentrated hydrochloric acid accompanied by the oxidation of chloride ions
(Reaction (11)) (Lemont, 2008).
3

2CeO2 8HCl2Ce

4H2 O 6Cl Cl2

11

64

H. Sasaki, M. Maeda / Hydrometallurgy 147148 (2014) 5967

Fig. 4. Scanning electron micrograph of spent catalyst treated with Zn vapor and its EDX spectrum.

This presumption is consistent with the result that more CeO2 dissolved in hydrochloric acid than in aqua regia, which has an oxidizing
ability. In addition, PGMs on CeO2 probably promoted the dissolution
because they can provide active sites for the evolution of chlorine
(Yokoyama and Enyo, 1982). Thus the dissolution of CeO2 presumably
progresses through a mechanism of galvanic corrosion. For an example
of a similar experiment, Virot et al. (2012) reported that reductive dissolution of CeO2 was promoted by the oxidation of formic acid or
hydrazinium nitrate on PGMs.
Fig. 9(b) shows that deoxidizing promoted the dissolution of CeO2
(#23); a part of CeO2 was probably reduced (Reaction (9)), which improved its solubility. The dissolution of Ce was further promoted by
the Zn-vapor treatment indicating the reduction of CeO2 by Zn (Reaction (8)), and the extraction reached 70% in #26. The dissolution of
ZrO2 was also enhanced by the pretreatments in Fig. 9, presumably because ZrO2 existed as solid solutions with CeO2 (Matsumoto and
Shinjoh, 2008). Al2O3 was not much affected by these pretreatments.
4.2. Dissolution of PGMs in hydrochloric acid
Fig. 9(a) showed the dissolution of PGMs from simulated catalysts in
hydrochloric acid without adding an oxidizing agent. One presumable

oxidizer in the experiments is CeO2, of which reduction might be


expressed by Reaction (12) (Hayes et al., 2002; Yu et al., 2006).

CeO2 e 4H Ce

2H2 O

E B 1:66V

12

Because the standard potential of Reaction (12) is higher than PGM


dissolutions (Reactions (2)(4)), supposedly CeO2 is able to oxidize
PGMs to ions.
From that standpoint, PGM extractions might be affected negatively
by deoxidizing and Zn-vapor treatments, which reduce CeO2. In this
study, extractions of Pt and Pd from deoxidized catalyst (#19, #23)
were similar or greater than untreated spent catalysts (#18, #22), suggesting that promoting effects were dominant under these conditions.
On the other hand, the Zn-vapor treatment had a negative effect in
#20 and #26, where dissolutions of Pt and Pd were obviously inhibited.
The dissolution of Ce was especially promoted in #26, indicating that a
large part of CeO2 reacted with Zn vapor.
Additionally, another aspect of alloying should be referred to. When
a PGMZn alloy is immersed in an acid solution, the open circuit potential will be lower than pure PGM, and thus only Zn dissolves in some
cases (Sasaki and Maeda, 2013). This effect also might have inhibited
the extraction of PGMs in #20 and #26.

H. Sasaki, M. Maeda / Hydrometallurgy 147148 (2014) 5967

65

Fig. 5. Extraction of (a) PGMs and (b) oxides after immersion in aqua regia at 60 C for
30 min (Test I).

4.3. Applicability of Zn-vapor pretreatment


When aqua regia is used in the recovery process of precious metals,
the disposal of waste solution includes a denitration process, resulting
in high cost. Because PGMs supported on CeO2 could be extracted by hydrochloric acid as mentioned above, the recovery of PGM catalysts
might be less expensive. However, not all PGMs in scraps are supported
by the oxide, and thus extractants containing oxidizing agents are generally needed at this time. This study also indicated that diluted aqua
regia worked on Zn-treated PGMs as effectively as an undiluted one. If
PGMs are extracted with a weaker oxidizer due to pretreatments, the
cost and environmental load involved in the recovery will be lower. At
least, because reactive species in aqua regia and chlorine are degradable
or gaseous, the enhancement of PGM dissolution by the pretreatments
can reduce the amount of chemical agents used.
For a practical process, the methodology in this study can be modied. Although catalysts were exposed to Zn vapor under vacuum in
the experiments, the pretreatment will also be feasible in the atmosphere. In this instance, an introduction of inert gas (e.g., nitrogen) or
some kind of reductant is presumably needed to prevent surface oxidation of the Zn vapor source and enhance its evaporation. Furthermore,
crushing and grinding of the scrap before or after the pretreatment
will aid in improving the extraction of PGMs.

Fig. 6. Extraction of (a) PGMs and (b) oxides after immersion in aqua regia at 17 C for
30 min (Test II).

of Pt, Pd and Rh in aqua regia were promoted by deoxidizing. The Znvapor treatment was effective as well for Pt and Pd, and further
enhancement was observed on Rh. Time variation of the extractions
showed that the pretreatment not only accelerated the dissolution
of PGMs, but also increased their extraction ratios after prolonged
leaching. PGMs considerably dissolved also in hydrochloric acid, especially at an elevated temperature. The experimental results indicated

5. Conclusions
Extractions of PGMs from spent catalysts and the effects of various
pretreatments were examined using simulated catalysts. Extractions

Fig. 7. Inuence of dilution of aqua regia on PGM extractions.

66

H. Sasaki, M. Maeda / Hydrometallurgy 147148 (2014) 5967

Acknowledgments
The authors appreciate Toyota Motor Corporation for providing simulated catalysts, which were prepared specically for this study. A part
of this work was conducted in the Center for Nano Lithography & Analysis of The University of Tokyo, supported by the Ministry of Education,
Culture, Sports, Science and Technology (MEXT), Japan. This study was
nancially supported by the Ministry of the Environment in Japan
through a Grant-in-Aid for Scientic Research about Establishing a
Sound Material-Cycle Society (Grant K2120, K22089, K2346).

References

Fig. 8. Time variation of extraction of PGMs in diluted aqua regia (Test III). Circles indicate
results of extractions after Zn-vapor pretreatment, and triangles are results without the
pretreatment.

the feasibility of the leaching of PGMs without adding an oxidizing


agent, and the possible effect of CeO2 was discussed. Pretreatments
should be considered carefully in hydrochloric acid leaching because it
inhibits the dissolution of PGMs depending upon the conditions.

Fig. 9. Extraction of (a) PGMs and (b) oxides after immersion in hydrochloric acid for
30 min (Test IV).

Angelidis, T.N., Skouraki, E., 1996. Preliminary studies of platinum dissolution from a
spent industrial catalyst. Appl. Catal. A 142, 387395.
Baghalha, M., Gh, H.K., Mortaheb, H.R., 2009. Kinetics of platinum extraction from spent
reforming catalysts in aqua regia solutions. Hydrometallurgy 95, 247253.
Barin, I., 1989. Thermochemical Data of Pure Substances. VCH-Verlagsgesellschaft, mbH,
Germany.
Butler, J., 2012. Platinum 2012. Johnson Matthey, Hertfordshire, U.K.
Chen, J., Huang, K., 2006. A new technique for extraction of platinum group metals by
pressure cyanidation. Hydrometallurgy 82, 164171.
Etoh, M., Tokumori, H., 1986. Determination of platinum and rhodium in catalysts by inductively coupled plasma atomic emission spectrometry. Bunseki Kagaku 35, T39T42.
Gandhi, H.S., Graham, G.W., McCabe, R.W., 2003. Automotive exhaust catalysis. J. Catal.
216, 433442.
Glin, P., Primet, M., 2002. Complete oxidation of methane at low temperature over noble
metal based catalysts: a review. Appl. Catal. B 39, 137.
Graham, G.W., Jen, H.-W., Chun, W., McCabe, R.W., 1999. High-temperature-aging-induced encapsulation of metal particles by support materials: comparative results
for Pt, Pd, and Rh on ceriumzirconium mixed oxides. J. Catal. 182, 228233.
Habashi, F., 1997. Handbook of Extractive Metallurgy. Wiley-VCH, Weinheim, Germany.
Hayes, S.A., Yu, P., O'Keefe, T.J., O'Keefe, M.J., Stoffer, J.O., 2002. The phase stability of
cerium species in aqueous systems, I. EpH diagram for the CeHClO4H2O system.
J. Electrochem. Soc. 149, C623C630.
Heck, R.M., Farrauto, R.J., 2001. Automobile exhaust catalysts. Appl. Catal. A 221, 443457.
Horike, C., Morita, K., Okabe, T.H., 2012. Effective dissolution of platinum by using chloride
salts in recovery process. Metall. Mater. Trans. B 43, 13001307.
Huang, S., Li, L., Van der Biest, O., Vleugels, J., 2005. Inuence of the oxygen partial pressure on the reduction of CeO2 and CeO2ZrO2 ceramics. Solid State Sci. 7, 539544.
Iwao, S., El-Fatah, S.A., Furukawa, K., Seki, T., Sasaki, M., Goto, M., 2007. Recovery of palladium from spent catalyst with supercritical CO2 and chelating agent. J. Supercrit.
Fluids 42, 200204.
Jafarifara, D., Daryanavardb, M.R., Sheibani, S., 2005. Ultra fast microwave-assisted
leaching for recovery of platinum from spent catalyst. Hydrometallurgy 78, 166171.
Jha, M.K., Lee, J., Kim, M., Jeong, J., Kim, B., Kumar, V., 2013. Hydrometallurgical recovery/
recycling of platinum by the leaching of spent catalysts. Hydrometallurgy 133, 2332.
Jimenez de Aberasturi, D., Pinedo, R., Ruiz de Larramendi, I., Ruiz de Larramendi, J.I., Rojo,
T., 2011. Recovery by hydrometallurgical extraction of the platinum-group metals
from car catalytic converters. Miner. Eng. 24, 505513.
Kayanuma, Y., Okabe, T.H., Mitsuda, Y., Maeda, M., 2004a. New recovery process for rhodium using metal vapor. J. Alloys Compd. 365, 211220.
Kayanuma, Y., Okabe, T.H., Maeda, M., 2004b. Metal vapor treatment for enhancing the
dissolution of platinum group metals from automotive catalyst scrap. Metall. Mater.
Trans. B 35, 817824.
Kononova, O.N., Melnikov, A.M., Borisova, T.V., Krylov, A.S., 2011. Simultaneous ion exchange recovery of platinum and rhodium from chloride solutions. Hydrometallurgy
105, 341349.
Lemont, F., 2008. Promising optimization of the CeO2/CeCl3 cycle by reductive dissolution
of cerium (IV) oxide. Int. J. Hydrogen Energy 33, 73557360.
Llopis, J., Tordesillas, I.M., Muniz, M., 1965. Anodic corrosion of rhodium in hydrochloric
acid solutions. Electrochim. Acta 10, 10451055.
Lucena, P., Laserna, J.J., 2001. Three-dimensional distribution analysis of platinum, palladium and rhodium in auto catalytic converters using imaging-mode laser-induced
breakdown spectrometry. Spectrochim. Acta B 56, 177185.
Marinho, R.S., Afonsoa, J.C., Cunhab, J.W.S.D., 2010. Recovery of platinum from spent catalysts by liquidliquid extraction in chloride medium. J. Hazard. Mater. 179, 488494.
Matsumoto, S., Shinjoh, H., 2008. Dynamic behavior and characterization of automobile
catalysts. Adv. Chem. Eng. 33, 146.
Okabe, T.H., Yamamoto, S., Kayanuma, Y., Maeda, M., 2003a. Recovery of platinum using
magnesium vapor. J. Mater. Res. 18, 19601967.
Okabe, T.H., Kayanuma, Y., Yamamoto, S., Maeda, M., 2003b. Platinum recovery using calcium vapor treatment. Mater. Trans. JIM 44, 13861393.
Reddya, B.R., Rajua, B., Lee, J.Y., Park, H.K., 2010. Process for the separation and recovery of
palladium and platinum from spent automobile catalyst leach liquor using LIX 84I
and Alamine 336. J. Hazard. Mater. 180, 253258.
Sasaki, H., Maeda, M., 2012. Enhanced dissolution of Rh from RhZn3 formed through Zn
vapor pretreatment. Metall. Mater. Trans. B 43B, 443448.
Sasaki, H., Maeda, M., 2013. Enhanced dissolution of Pt from PtZn intermetallic compounds and underpotential dissolution from Zn-rich alloys. J. Phys. Chem. C 117,
1845718463.

H. Sasaki, M. Maeda / Hydrometallurgy 147148 (2014) 5967


Sasaki, H., Miyake, M., Maeda, M., 2010a. Enhanced dissolution rate of Pt from a PtZn
compound measured by channel ow double electrode. J. Electrochem. Soc. 157,
E82E87.
Sasaki, H., Nagai, T., Maeda, M., 2010b. Synthesis of PtZn and AuZn compounds by isopiestic method. J. Alloys Compd. 504, 475478.
Sun, P.P., Lee, J.Y., Lee, M.S., 2012. Separation of platinum(IV) and rhodium(III) from acidic
chloride solution by ion exchange with anion resins. Hydrometallurgy 113114,
200204.
Suzuki, S., Ogino, M., Matsumoto, T., 2007. Recovery of platinum group metals at Nippon
PGM Co., Ltd. J. MMIJ 123, 734736.

67

Virot, M., Chave, T., Horlait, D., Clavier, N., Dacheux, N., Ravaux, J., Nikitenko, S.I., 2012. Catalytic dissolution of ceria under mild conditions. J. Mater. Chem. 22, 1473414740.
Yokoyama, T., Enyo, M., 1982. On the mechanism of the chlorine evolution reaction on Pt
electrode. J. Electroanal. Chem. 136, 185190.
Yoo, J.S., 1998. Metal recovery and rejuvenation of metal-loaded spent catalysts. Catal.
Today 44, 2746.
Yu, P., Hayes, S.A., O'Keefe, T.J., O'Keefe, M.J., Stoffer, J.O., 2006. The phase stability of cerium species in aqueous systems II. The Ce(III/IV)H2OH2O2/O2 systems. Equilibrium
considerations and Pourbaix diagram calculations. J. Electrochem. Soc. 153, C74C79.

You might also like