ArticlePDF Available

Investigation of stainless steel pickling liquor as a precursor for high Capacity battery electrode materials

Royal Society of Chemistry
RSC Advances
Authors:

Abstract and Figures

Stainless steel and battery manufacturing industries can be united for mutual economic and environmental benefits by utilizing stainless steel pickling liquor waste product as a precursor for high energy iron fluoride based positive electrode materials in batteries. This study analyzes the feasibility, environmental, and economic cost of the ferric fluoride (FeF3) synthesis approach through the use of recycled pickling liquor from the stainless steel fabrication industry. This new synthesis method is determined as more environmentally friendly than the current methods of disposing spent stainless steel pickling liquor and producing ferric fluoride separately. X-ray diffraction analysis demonstrated the synthesis occurs in two steps: the conversion of spent pickling liquor to produce the crystallohydrate β-FeF3·3H2O followed by its dehydration into anhydrous FeF3. Materials obtained from pickling solutions were found electrochemically active with improved cycling stability and similar capacity compared to commercial FeF3. Pickling solutions synthesized with nickel and chromium to better replicate stainless steel industry spent pickling waste showed improved electrochemical performance.
Content may be subject to copyright.
Investigation of stainless steel pickling liquor as a
precursor for high Capacity battery electrode
materials
Sheel Sanghvi, Nathalie Pereira,*Anna Halajko and Glenn G. Amatucci
Stainless steel and battery manufacturing industries can be united for mutual economic and environmental
benets by utilizing stainless steel pickling liquor waste product as a precursor for high energy iron uoride
based positive electrode materials in batteries. This study analyzes the feasibility, environmental, and
economic cost of the ferric uoride (FeF
3
) synthesis approach through the use of recycled pickling liquor
from the stainless steel fabrication industry. This new synthesis method is determined as more
environmentally friendly than the current methods of disposing spent stainless steel pickling liquor and
producing ferric uoride separately. X-ray diraction analysis demonstrated the synthesis occurs in two
steps: the conversion of spent pickling liquor to produce the crystallohydrate b-FeF
3
$3H
2
O followed by
its dehydration into anhydrous FeF
3
. Materials obtained from pickling solutions were found
electrochemically active with improved cycling stability and similar capacity compared to commercial
FeF
3
. Pickling solutions synthesized with nickel and chromium to better replicate stainless steel industry
spent pickling waste showed improved electrochemical performance.
Introduction
Transportable power sources are enabling technologies for an
enormous breadth of devices that bring signicant benetto
society. Although almost all applications require high perfor-
mance, the sheer scale of battery production demands appro-
priate environmental responsibility and economic feasibility.
Metal uoride nanocomposites have been shown as an
appealing approach to positive electrode materials for primary
and secondary batteries based on their high energy densities
compared to the current state of the art.
1
More specically, iron-
based materials are desirable based on iron's low toxicity,
abundance and therefore potential lower cost. However, the
uoride component can contribute a fairly signicant process-
ing cost and possible impact to the environment especially if
processed in a non responsible manner.
The most common process of producing FeF
3
is shown by
the reaction in eqn (1). A constant stream of anhydrous
hydrogen uoride (HF) is passed over anhydrous ferric chloride
(FeCl
3
) in a heated, oxygen-free reactor.
2
FeCl
3
+HF/FeF
3
+ HCl (1)
Such method of production of iron uorides poses a signif-
icant environmental issue through an unnecessary, extra
exploitation of energy and precursor materials.
A possible alternative to produce FeF
3
is to utilize pickling
liquor, a waste product of the stainless steel industry. The
stainless steel manufacturing industry requires a nal process
known as pickling to etch the surface of as processed stainless
steel and restore the protective eects of chromium and other
surface oxides. Aer the pickling process is completed, the acid
is rinsed othe surface of the steel and stored in waste
containers as spent pickling liquor (SPL). As discussed in detail
below, such SPL typically contains nitric acid, hydrouoric acid,
iron and various other metals.
Large amounts of spent liquor need to be disposed of each
year by stainless steel manufacturers leading to a non insig-
nicant cost to the product. Although some stainless steel
manufacturers attempt to regenerate the acids or recover the
metals from the spent liquor, most processes are not fully
eective leading to appreciable environmental and nancial
costs. A frequently employed method to treat spent liquor is
through lime neutralization.
Both processes of producing new FeF
3
and disposing spent
liquor introduce certain environmental impacts. A quantitative
metric needs to be enacted to compare if the impact of these
process outweigh the impact of producing FeF
3
from SPL. This
is accomplished through the use of the Waste Reduction Algo-
rithm (WAR) GUI created by the United States Environmental
Protection Agency. The WAR GUI produces a numerical score
for a chemical process called the potential environmental
impact (PEI) to describe the impact a process will have. This is
dened in further detail below.
Energy Storage Research Group, Dept. of Materials Science and Engineering, Rutgers
University, North Brunswick, New Jersey 08902, US. E-mail: npereira@rci.rutgers.edu
Cite this: RSC Adv.,2014,4,57098
Received 30th September 2014
Accepted 23rd October 2014
DOI: 10.1039/c4ra11539b
www.rsc.org/advances
57098 |RSC Adv.,2014,4,5709857110 This journal is © The Royal Society of Chemistry 2014
RSC Advances
PAPER
The PEI values obtained from the WAR GUI for each of the
processes will be plugged into eqn (2) to estimate the environ-
mental impact of the fabrication of the FeF
3
from the spent
liquor compared to the combination of the spent liquor
disposal and the current commercial FeF
3
fabrication process.
Appropriate satisfaction of this equation would justify the
environmental benet of converting spent pickling liquor into
FeF
3
.
Env. Impact
New FeF
3
+ Env. Impact
SPL Disposal
>
Env. Impact
FeF
3
from SPL
(2)
However, before any environmental calculations can be
completed, a chemical pathway behind the conversion process
must be demonstrated. A possible pathway to convert SPL into
iron uorides is the precipitation of b-FeF
3
$3H
2
O from SPL
followed by dehydration into FeF
3
. Other research groups have
independently studied dierent points of this pathway. Tjus
et al. and Sartor et al. conrm that b-FeF
3
$3H
2
O crystals can be
separated from spent liquor.
3,4
Quite recently, Myung et al. and
Liu et al. produced the anhydrous form of ferric uoride from
the crystal hydrate.
5,6
This paper aims to unite all points to form
one complete method of an environmentally sound supply
stream to show that the waste pickling liquor from the stainless
steel industry can be transformed to high value added compo-
nent product of high energy density iron uoride for battery
applications. This study also encompasses the evaluation of the
electrochemical activity of the products synthesized from pick-
ling liquors with a comparison to a commercial FeF
3
material.
Finally, the economics of the synthesis of FeF
3
from spent
liquor are considered and compared to costs associated to the
disposal of spent liquor and the fabrication of FeF
3
using
current methods.
Experimental
Sample solutions of spent pickling liquor were synthesized by
dissolving iron nitrate hydrate Fe(NO
3
)
3
$9H
2
O (Alfa Aesar),
chromium nitrate hydrate Cr(NO
3
)
3
$9H
2
O (Sigma Aldrich), and/
or nickel nitrate hydrate Ni(NO
3
)
2
$6H
2
O (Sigma Aldrich) in a
solution of hydrouoric acid (HF) (48%, Sigma Aldrich) and
deionized water. Solutions were stored at room temperature for
one week to allow for crystallization to fully occur. The produced
crystals were rst ltered with lter paper (Whatman 542) then
washed with three 10 mL aliquots of ethanol (Sigma Aldrich).
Finally, the water adhering to the surface of the crystals was
removed by drying overnight at 70 C in air with <1% relative
humidity.
The dehydration of the crystals was completed in a Lindberg/
Blue-M tube furnace at 150400 C under owing argon for
various intervals of time. Additionally, dierential scanning
calorimetry (TA-Instruments Q10) was performed using pans
hermetically sealed in a helium environment subsequently
heated at a rate of 5 C min
1
from room temperature to 450 C.
X-ray diraction was completed using a Bruker D8 Advance
diractometer with a 0.022qstep size, a 1.9 s per step count, a
15602qangle span and a CuKaX-ray source. Furthermore,
using the soware TOPAS, X-ray diraction patterns were tted
by Rietveld Renement to estimate crystal size. Scanning elec-
tron microscopy was performed on a Zeiss Sigma Field Emis-
sion SEM with an Oxford INCA PentaFETx3 Energy Dispersive
Spectroscopy (EDS) attachment. The secondary electron
detector was used with a voltage of 5.00 kV and a working
distance of 7.9 millimeters.
In order to evaluate the electrochemical performance of the
materials obtained from pickling liquor and compare to
commercial FeF
3
(Advanced Research Chemicals), nano-
composites were fabricated by high-energy mechanomilling.
7
The iron uorides either synthesized from pickling liquor or
obtained from a commercial source were milled with 15 wt%
activated carbon (ASupra, Norit) for 1 h in helium.
Electrodes were fabricated according to the Bellcore devel-
oped process.
8
The active materials were mixed in acetone
(Aldrich) with poly(vinylidene uoride-co-hexauoropropylene)
binder (Kynar 2801, Elf Atochem), carbon black additive (Super
P, MMM) and dibutyl phthalate plasticizer (Aldrich). Aer
plasticizer extraction in anhydrous ether (Aldrich), the elec-
trodes typically consisted of 57.25% active material, 12% carbon
additive, and 31% binder. Two-electrode coin cells were
assembled using a lithium foil counter electrode and glass ber
separators (GF/D, Whatman) saturated with 1 M LiPF
6
in
ethylene carbonatedimethyl carbonate electrolyte (50 : 50 in
vol.%) from BASF. The cells were cycled at 60 C under a
constant current of 7.5 mA g
1
based on the weight of the
nanocomposite, between 1.5 and 4.5 V.
Various sources of information were consulted to determine
the material ow and energy consumption for each production
stream. Using such data, an environmental impact estimate was
calculated using the WAR GUI soware to generate a PEI score.
The PEI is composed of a sum of eight categories: human
toxicity potential by ingestion, human toxicity potential by
dermal contact or inhalation, aquatic toxicity potential, terres-
trial toxicity potential, global warming potential, ozone deple-
tion potential, photochemical oxidation potential, and
acidication potential. Each category is assigned a specic
weight depending on how important that environmental
concern is to the chemical process. For this study, all categories
were assumed to have equal weight. Furthermore, for each of
the eight categories, each chemical is assigned a normalized
value which is computed from toxicology information in the
program's database and literature. The normalized values are
added together to create the PEI. The specic PEI of interest for
this paper is calculated as a summation of the impact of
chemicals produced and energy consumed.
Many assumptions are considered for all PEI calculations in
this study. For example, the crystals produced from SPL by the
process described later in this paper are assumed to be
composed primarily of FeF
3
with nominal amounts of CrF
3
.
Additionally, except for the SPL composition, if the amount of
reactants or products in a specic chemical process is given as a
range of values, then the smallest amounts of each reactant or
product are used for calculations. This will minimize the envi-
ronmental impact of the chemical process and provide the best
case scenario. Another assumption is that non-descriptive items
This journal is © The Royal Society of Chemistry 2014 RSC Adv.,2014,4,5709857110 | 57099
Paper RSC Advances
such as solid waste or heavy metals are not included in PEI
calculations, which causes the mass balance to be slightly o.
Also, the impact of transportation between facilities was not
included in PEI calculations because of the multiple options
and distances for production facilities. Next, the energy impact
was combined to include all utilities such as heating, electricity,
refrigeration, etc. and gas was assumed to be the only source to
generate the energy. This provides a uniform basis for energy
calculations. Finally, the impact of process water was ignored
because water has an insignicant impact and can be used in
varying quantities for dierent processes.
Results
1. Disposal process of spent pickling liquor from the
stainless steel industry
Spent liquor typically has the composition of the Industrial
SPLshown in Table 1. The spent liquor is disposed of by lime
neutralization. Calcium hydroxide or slaked lime is introduced
into a tank containing the liquor to raise the pH of the solution
and precipitate ferric hydroxide and calcium uoride. The solid
waste from this reaction is removed by ltering and then, it is
pressed dry and landlled. The aqueous waste, which contains
calcium and nitrate ions, is either released into a stream or sent
to a wastewater treatment plant for further processing. Fig. 1
shows the inputs and outputs of this process.
2. Current commercial synthesis process of FeF
3
Anhydrous ferric uoride is rarely fabricated at industrial levels
because the hydrated, chloride form of the metal salt is
preferred. However, it is possible to manufacture anhydrous
ferric uoride by passing a stream of anhydrous hydrogen
uoride over anhydrous ferric chloride in an oxygen-free envi-
ronment.
2
Fig. 2 below shows this process along with the
production of the precursors.
The precursors for ferric uoride production are hydrogen
uoride and ferric chloride. Hydrogen uoride is synthesized by
reacting concentrated sulfuric acid with the mineral uorspar
in a heated reactor.
9
It is then dehydrated by condensation and
distillation. A co-product of this reaction is calcium sulfate,
which can enter other markets rather than being discarded as
waste.
10
The other precursor for this reaction, ferric chloride,
can be synthesized in its hydrated form by recycling spent steel
pickling liquor (contains hydrochloric acid) with scrap iron.
However, it is necessary to dehydrate the hydrated ferric chlo-
ride to provide a suitable precursor for ferric uoride
production. The dehydration methods known are not practical.
Dehydration by thermal treatment produces a low yield, while
dehydration with a dehydrating agent such as thionyl chloride
uses hazardous reactants and is costly. The issues associated
with both dehydration processes prevent the use of ferric
chloride produced by recycling steel pickling liquor. Instead,
anhydrous ferric chloride is produced by oxidizing red hot iron
with chlorine gas.
10
This process has a large yield and generates
hydrogen chloride as a waste product. The produced anhydrous
hydrogen uoride (HF) is passed over anhydrous ferric chloride
(FeCl
3
) in a heated, oxygen-free reactor to produce FeF
3
.
2
3. Synthesis of FeF
3
from Fe-only pickling liquor
Production of ferric uoride from spent pickling liquor can be
split into two steps: the synthesis of b-FeF
3
$3H
2
O from SPL and
the dehydration of b-FeF
3
$3H
2
O to FeF
3
. Two practices
employed in industry to prepare b-FeF
3
$3H
2
O are evaporation
or membrane ltration in conjunction with crystallization.
During evaporation, the spent liquor is heated to its boiling
point to remove excess acids and water while increasing the iron
concentration to 80120 kg m
3
of SPL.
11,12
This concentration
is required to form a super saturated solution that will later be
crystallized. Membrane ltration produces results similar to
evaporation. First, microltration eliminates large solids from
the solution and then, nanoltration separates metal ions from
the free acids. The portion retaining metal ions is recirculated
through the system multiple times to increase the iron
concentration.
3
The supersaturated solution is then held at
approximately 50 C to allow for the nucleation and growth of b-
FeF
3
$3H
2
O crystals. Additional hydrouoric acid can also be
added to supplement crystal growth.
Unlike evaporation or membrane ltration, low temperature
crystallization has not yet been employed by industry. Pickling
liquor is slowly cooled to about 40 C to precipitate ice and
FeF
3
$3H
2
O crystals, which can then be separated by
suction ltration.
4
These three possible routes to synthesize b-
FeF
3
$3H
2
O are highlighted in the upper portion of Fig. 3.
However, because the solutions in this paper are synthetically
prepared to specic concentrations, these techniques are not
studied in detail. They are only mentioned to describe the
conversion process from pickling liquor to FeF
3
in its entirety.
The b-FeF
3
$3H
2
O crystals are then converted to FeF
3
by
thermal decomposition as highlighted in the lower portion of
Fig. 3. The specic temperatures for decomposition are dis-
cussed herein. In this study, sample pickling liquors containing
iron as the only metallic species were prepared to establish a
Table 1 Spent pickling liquor compositions
Compositions of spent pickling liquor
Pickling liquor [Fe
3+
](gL
1
) [Cr
3+
](gL
1
) [Ni
2+
](gL
1
) [NO
3
](gL
1
)[F
](gL
1
)
Industrial SPL
22
3045 510 35 150180 6080
Fe SPL A 43.2 ——143.9 44.1
Fe SPL B
23
78.9 ——258.7 80.7
FeCrNi SPL
4
42.2 10.8 6 195 60
57100 |RSC Adv.,2014,4,5709857110 This journal is © The Royal Society of Chemistry 2014
RSC Advances Paper
comparative baseline. The solution was initially composed of
the composition Fe SPL A in Table 1, but this process resulted in
a very low yield of fabrication. The composition was therefore
altered to the composition Fe SPL B in Table 1 which improved
yield to about 27% mol of precipitated b-FeF
3
$3H
2
Oaer one
week of reaction based on original concentration of Fe in the
solution. The fabricated crystals were singled phase and iden-
tied as b-FeF
3
$3H
2
O by X-ray diraction analysis as shown in
Fig. 4.
DSC analysis was performed to study the characteristics of
decomposition during dehydration. Fig. 5 reveals the trans-
formation of b-FeF
3
$3H
2
O to FeF
3
occurs in a single, large
endothermic decomposition step at approximately 140.47 C
while the small bump at 97.32 C could be associated to the
evaporation of adsorbed water. As a result, heat treatments were
performed at various temperatures above 140 C under owing
Ar with the goal to dehydrate b-FeF
3
$3H
2
O into FeF
3
.Arst
sample was heat-treated at 200 C for two hours, which resulted
in the decomposition of b-FeF
3
$3H
2
O into a hydrate of lower
water content FeF
3
$0.33H
2
O and small amounts of anhydrous
FeF
3
(Fig. 6). A second sample was then heat-treated at higher
temperature. At 400 C, dehydration was driven further and the
sample consisted mostly of anhydrous FeF
3
with a very small
amount of FeF
3
$0.33H
2
O (Fig. 6). The residual FeF
3
$0.33H
2
O
phase could result from reaction of FeF
3
with ambient air upon
transfer from the tube furnace to a He-lled glove box. Finally, a
third sample was submitted to two successive 2 hour heat-
treatments at 150 C and 400 C leading to successful dehy-
dration with complete transformation into single phase FeF
3
(Fig. 7). Table 2 shows the lattice parameters of the FeF
3
materials synthesized from SPL B are consistent with JCPDS
values and close to that obtained from a commercial source. In
addition, the material we fabricated is of lower crystallinity than
FeF
3
obtained from a commercial source with 38 nm crystallite
size compared to 62 nm for the commercial sample.
Aer heat treatment at 400 C, the sample was observed to be
of an orange-brown color rather than the characteristic light
green color of commercial FeF
3
. It is possible an oxide layer has
formed on the surface of the powder in spite of the lack of
evidence by XRD analysis. The absence of oxide by XRD most
likely stems from amounts below detection limit. To further
identify if this oxide layer exists at the surface, temperature was
increased and a sample was submitted to three consecutive
heat-treatments at 150 C, 400 C and 450 C respectively. Fig. 7
Fig. 1 Material ow chart for spent pickling liquor neutralization.
Fig. 2 Material ow chart of ferric uoride production.
This journal is © The Royal Society of Chemistry 2014 RSC Adv.,2014,4,5709857110 | 57101
Paper RSC Advances
shows that heating from 400 C to 450 C results in the growth
of Fe
2
O
3
. Additionally, the color of the sample changes to a
slightly darker shade of orange-brown. Finally, because the
position of the FeF
3
peaks do not shiaer heat treatment,
there is indication that surface oxidation is occurring rather
than oxidation within the crystal lattice. Small amounts of
Fig. 3 Material ow chart of ferric uoride production from spent pickling liquor.
Fig. 4 XRD pattern of crystals precipitated from iron only pickling solution at room temperature. Peaks indexed are associated to b-FeF
3
$3H
2
O
(JCPDS 00-032-0464).
57102 |RSC Adv.,2014,4,5709857110 This journal is © The Royal Society of Chemistry 2014
RSC Advances Paper
surface oxidation are expected to be a benet to the ultimate
electrochemical processes.
4. Synthesis of FeF
3
and CrF
3
from FeCrNi pickling liquor
In order to more accurately replicate pickling liquors obtained
from the stainless steel fabrication process, chromium and
nickel metal species were added to generate pickling liquors of
composition FeCrNi SPL from Table 1. The composition is
derived from a metals ratio of 36 : 5:9 Fe : Ni : Cr and a metals
concentration of 60g L
1
.
4
The solution is prepared and stored
for one week to allow for crystallization to occur. Aer crystal-
lization was complete, a yield of nearly 29% was obtained,
assuming both FeF
3
$3H
2
O and CrF
3
$3H
2
O precipitate out of
solution. This value is similar to that obtained with pure iron
solutions.
X-ray diraction analysis (Fig. 8) of the crystals obtained
aer crystallization reveals the presence of b-FeF
3
$3H
2
O and,
CrF
3
$3H
2
Oora-FeF
3
$3H
2
O. Since CrF
3
$3H
2
O is isostructural to
the a-FeF
3
$3H
2
O phase, distinction between both phases is
dicult.
13
However, it should be noted that a-FeF
3
$3H
2
Oisa
disordered metastable phase whose crystal structure reorients
into b-FeF
3
$3H
2
O over an extended period of time or at elevated
temperatures and no alpha phase was identied in our previous
precipitations without the presence of Cr.
14
Therefore, we
believe it could be possible to distinguish between both phases
aer heat treatment.
DSC analysis, shown in Fig. 9, reveals that the decomposition
of the mixture of b-FeF
3
$3H
2
O with CrF
3
$3H
2
Oora-FeF
3
$3H
2
O
into FeF
3
occurs in two closely spaced endothermic decompo-
sition steps at approximately 146.96 C and 166.45 C respec-
tively. The rst step correlates to the decomposition of b-
FeF
3
$3H
2
O while the second step could correlate to the
decomposition of CrF
3
$3H
2
Oora-FeF
3
$3H
2
O. A small bump
before the main peak at 96.88 C can be associated to the
evaporation of adsorbed water.
Fig. 5 DSC scan of b-FeF
3
$3H
2
O heated in helium from RT to 450 C
at a rate of 5 C min
1
.
Fig. 6 XRD pattern of the initial b-FeF
3
$3H
2
O sample after 2 hour heat-treatment at 200 C (bottom), and at 400 C (top). Peaks indexed by stars
are consistent with FeF
3
(JCPDF 01-088-2023) and peaks indexed by diamonds are consistent with FeF
3
$0.33H
2
O (JCPDS 01-076-1262).
This journal is © The Royal Society of Chemistry 2014 RSC Adv.,2014,4,5709857110 | 57103
Paper RSC Advances
As performed with the pure b-FeF
3
$3H
2
O samples reported
above, samples were heated above the decomposition temper-
atures obtained by DSC. As with the optimized pure Fe uoride
hydrates, the samples were heat-treated at 150 C for 2 hours to
allow for the initial dehydration step, then at 400 C for 5 hours
to form the nal product. The resulting XRD scan reported in
Fig. 10 reveals single phase FeF
3
. Lattice parameters (Table 2)
are similar to that obtained with the Fe-only pickling solution
and consistent with JCPDS values for rhombohedral FeF
3
. The
synthesized material shows broader diraction peaks than the
ones obtained with the commercial FeF
3
consistent to a lower
crystal size of 41 nm obtained by Rietveld renement.
Since X-ray diraction does not present any evidence for any
distinct Cr-based phase, we can hypothesize on the possibility
of the formation of a chromiumiron solid solution during
crystallization. The later would be consistent with the Hume
Rothery rules determined from the similarity in atomic radii of
chromium (0.62 ˚
A) and iron (0.65 ˚
A) ions in the 3+ state.
15
This
would also be consistent with a previous report of Cr
0.5
Fe
0.5
F
3
solid solution uorides.
16
As shown in Table 3, quantitative EDS
analysis of the sample synthesized from FeCrNi pickling
liquor conrmed the material comprises iron, chromium,
uorine, and oxygen however no nickel was detected. The
Fe : Cr elemental ratio is calculated from the relative atomic
percents of Fe and Cr resulting in a value of approximately
3.747 : 1. This is incorporated into the chemical formula of
ferric uoride to give a composition of Fe
0.79
Cr
0.21
F
3
. Field
emission electron microscopy (FESEM) analysis of the nal
product revealed a wide distribution of crystal shape and sizes
(Fig. 11). Furthermore, even though X-ray analysis does not
indicate the presence of any oxide phase, through visual iden-
tication of the color change it can be concluded that an oxide
coating has formed on the surface of the ironchromium uo-
ride particles as similarly mentioned above for the pure FeF
3
sample.
5. Electrochemical performance
As demonstrated above, anhydrous iron uoride has been
successfully synthesized from iron based pickling liquors
through successive heat-treatments under argon. The samples
Fig. 7 XRD scan of the b-FeF
3
$3H
2
O sample submitted to two consecutive heat-treatments at 150 C and 400 C (bottom), and followed by a
third heat-treatment at 450 C for two hours (top). Peaks indexed by stars are associated to FeF
3
(JCPDS 01-071-3710) and peaks indexed by
open circles to Fe
2
O
3
(JCPDS 00-089-0599).
Table 2 Lattice parameters obtained by Reitveld renement using the
software TOPAS and compared to theoretical values
Source Crystal Structure a(
˚
A) c (˚
A)
Theoretical ( JCPDS 01-088-2023) Rhombohedral 5.1941(9) 13.334(9)
Commercial FeF
3
Rhombohedral 5.206(3) 13.287(3)
FeF
3
from Fe SPL B Rhombohedral 5.208(0) 13.352(8)
FeF
3
from FeCrNi SPL Rhombohedral 5.201(2) 13.314(2)
57104 |RSC Adv.,2014,4,5709857110 This journal is © The Royal Society of Chemistry 2014
RSC Advances Paper
tested for electrochemical activity were synthesized through a
succession of two heat-treatment at 150 C for two hours and at
400 C for ve hours. Both the materials obtained from Fe-only
and FeNiCr liquors were evaluated and compared to
commercial FeF
3
aer formation into nanocomposites. Fig. 12
presents the second cycle voltage proles obtained at 60 C with
a current of 7.5 mA g
1
based on the weight of the
nanocomposite. Both materials fabricated from pickling liquors
are electrochemically active and exhibit discharge proles
similar to that of commercial FeF
3
, except for the slightly lower
capacity of the FeCrNi sample. All capacities obtained excee-
ded 75% of theoretical values inclusive of the FeF
3
/Li
0.5
FeF
3
/LiFeF
3
insertion reaction above 3 V and the LiFeF
3
/3LiF +
Fe conversion reactions at approximately 2 V. No distinct elec-
trochemical signatures originating from CrF
3
or NiF
2
could be
abstracted from the voltage prole.
Discharge capacity per gram of FeF
3
was plotted as a func-
tion of cycle number in Fig. 13. Both samples fabricated from
pickling solution showed improved capacity retention
compared to the commercial material. Best results were
obtained with the sample derived from the FeCrNi solution.
6. Environmental calculations
6.1. Disposal of spent pickling liquor. The process of spent
pickling liquor disposal was summarized in Fig. 1 while Table 4
presents the quantities of input as well as the output materials
involved in the disposal process. The values in Table 4 are
calculated supposing stoichiometric reactions. However, an
excess of the hydroxide ion is included in this table to ensure
mass balance even though it forms water by neutralizing the
acids. Using these assumptions, the estimated environmental
Fig. 8 XRD pattern of crystals produced from the Fe, Ni, and Cr sample pickling liquors. Peaks marked by point-up triangles are associated to b-
FeF
3
$3H
2
O (JCPDS 00-032-0464) and peaks marked by point-down triangles are associated to CrF
3
$3H
2
O (JCPDS 00-017-0316) or a-FeF
3
-
$3H
2
O (JCPDS 00-024-0071).
Fig. 9 DSC scan of hydrated crystal sample from FeCrNi SPL
heated from RT to 450 C at a rate of 5 C min
1
.
This journal is © The Royal Society of Chemistry 2014 RSC Adv.,2014,4,5709857110 | 57105
Paper RSC Advances
impact for SPL disposal is 745 PEI per L of SPL. Since values for
this section are given in units of PEI per L of SPL and the theme
of units used throughout this paper is PEI per kg of FeF
3
, the
above value can be converted into 23 281 PEI per kg of FeF
3
.
This is accomplished with a conversation factor that 1 L of SPL
has the potential to yield 0.032 kg of FeF
3
(or (Fe,Cr)F
3
). The
derivation of this factor is explained in detail in Section 6.3
below. Major contributing elements to this calculation are
associated to the calcium and nitrate ions remaining in the
wastewater, which mainly have aquatic toxicity eects. The
energy contribution is assumed to be minimal because the
neutralization process does not require relatively demanding
mixing or ltering technology.
6.2. Production of anhydrous ferric uoride via current
commercial synthesis process. The commercial production of
anhydrous FeF
3
was presented in Fig. 2 while Table 5 shows the
material ow used to estimate the environmental impact of
production of ferric uoride. The estimated environmental
impact for ferric uoride production is 1.273 PEI per kg of FeF
3
.
Fig. 10 XRD patterns of crystals produced from the Fe- and the FeCrNi pickling liquors heated at 150 C for two hours and then at 400 C for
ve hours, compared to a commercial FeF
3
material. Peaks indexed indicate FeF
3
(JCPDS 01-071-3710) while system peak is present in the Fe
CrNi sample as shown by the star.
Table 3 Atomic percent information obtained from EDS of the nal
(Fe,Cr)F
3
product
Element Atomic percent
C 4.12
O 7.85
F 58.74
Cr 6.17
Fe 23.12
Fig. 11 SEM image of the nal (Fe,Cr)F
3
product at a 652
magnication.
57106 |RSC Adv.,2014,4,5709857110 This journal is © The Royal Society of Chemistry 2014
RSC Advances Paper
The chemical component, which has a value of 1.2662 PEI per
kg of FeF
3
, greatly outweighs the energy component, which has
a value of 0.0068 PEI kg
1
. This occurs because of the high
toxicity of the outputs, specically hydrogen chloride gas which
is released into the environment. Specic assumptions for this
PEI calculation are that uorspar is primarily calcium uoride
and the synthesis of ferric uoride from its precursors is stoi-
chiometric with a two percent loss of ferric chloride.
17
6.3. Production of (Fe,Cr)F
3
from spent pickling liquor. As
described above, production of (Fe,Cr)F
3
from spent pickling
liquor can be split into two steps: crystallization of the hydrate
from SPL and dehydration to (Fe,Cr)F
3
. The composition of SPL
used in this paper allows for crystallization to favorably occur
without any added chemicals or energy. The various techniques
of crystallization mentioned above incorporate an amount of
energy with an environmental cost comparatively negligible to
the rest of the system. Therefore, the only contributing factor to
environmental cost from the crystallization step is disposal of
the euent remaining aer crystallization. The exact compo-
sition of this liquid is unknown. However, by calculating
backwards from the estimated crystal yield and crystal compo-
sition, the remaining liquid composition is estimated to be 30.7
gL
1
[Fe
3+
], 7.7 g L
1
[Cr
3+
], 6 g L
1
[Ni
2+
], 43.6 g L
1
[F
], and
195 g L
1
[NO
3
]. Slight uctuations in this calculation should
not have a major eect on PEI. The worst case scenario for
disposing this euent is through lime neutralization. Table 6
below shows the stoichiometrically calculated material ow
used to estimate the PEI. The estimated environmental impact
for this method of crystallization euent disposal is 727.7 PEI
per L of SPL.
The dehydration step of the (Fe,Cr)F
3
crystallohydrate
follows a simple thermal decomposition process. Argon gas is
introduced into a furnace to remove ambient air and the
furnace is incrementally heated to 400 C for extended periods
of time. The main source of environmental impact for this step
is the energy required to heat the furnace. The environmental
impact of the process gas is assumed to be negligible because it
is oen a byproduct of other gas rening processes. The impact
of heating is estimated from the required enthalpy for trans-
formation given by the DSC shown in Fig. 9. The enthalpy is
approximately 16.313 kJ per gram of crystallohydrate. Thus, the
potential environmental impact for dehydration is 0.16 PEI per
kg of (Fe,Cr)F
3
.
Before the combined PEI for the entire conversion process
can be calculated, the units per L of SPL need to be adjusted into
is per kg of (Fe,Cr)F
3
. This is accomplished by calculating the
amount of metal uoride hydrates present in the proposed
solution and then adjusting the value according to the yield
obtained through experimental trials. Finally, the hydrated
form of the crystal must be converted to its dehydrated form.
Therefore, 1 L of SPL has the potential to yield 0.032 kg of
(Fe,Cr)F
3
. The new PEI for the crystallization step is 22 740 PEI
per kg of (Fe,Cr)F
3
. The combined PEI for the whole conversion
process is roughly the same value because thermal decompo-
sition does not contribute much to environmental impact.
Discussion
Crystallization from iron only pickling solutions proved that an
appreciable yield of nanocyrstalline b-FeF
3
$3H
2
O could rela-
tively rapidly be produced. The crystallization process eciency
could be further improved by optimization of temperature
control, crystallization time, crystal seed introduction, and HF
concentration. Forsberg and Rasmuson demonstrated that the
Fig. 12 Second cycle voltage prole of the samples synthesized from
pickling liquors compared to commercial FeF
3
performed at 60 C and
7.5 mA g
1
based on the weight of the nanocomposite.
Fig. 13 Discharge capacity (per gram of active FeF
3
)versus cycle
number of the samples synthesized from pickling liquors compared
commercial FeF
3
performed at 60 C and 7.5 mA g
1
based on the
weight of the nanocomposite.
Table 4 Inputs and outputs for SPL disposal used in the PEI calculation
Inputs Outputs
SPL (L) Ca(OH)
2
(g) Fe(OH)
3
(g) Cr(OH)
3
(g) Ni(OH)
2
(g) CaF
2
(g) Ca(NO
3
)
2
(g) OH
(g)
1 233.43 82.66 21.4 9.48 123.2 258.04 53.64
This journal is © The Royal Society of Chemistry 2014 RSC Adv.,2014,4,5709857110 | 57107
Paper RSC Advances
crystal growth of iron uoride trihydrate can be optimized at a
temperature of 50 C, which in turn results in a higher yield.
18
Additionally, basic crystal growth kinetics indicate that if the
sample pickling liquor was allowed to crystalize for a longer
time period, then a larger yield of b-FeF
3
$3H
2
O is produced
because there is more time to reach thermodynamic equilib-
rium. Furthermore, the same result can be achieved by the
addition of seed crystals which provide sites of reduced surface
energy for growth. Finally, we performed preliminary experi-
ments indicating adding additional concentrated HF to sample
pickling liquor would result in not only a dramatically increased
yield, but also a shorter crystallization time. The tradeoto this
approach is that precipitate composition could change depen-
dent on the altered solubility of the metal ions in solution.
Thermal treatment at 400 C was demonstrated to be eec-
tive in converting b-FeF
3
$3H
2
O into FeF
3
, but with a potential
slight oxidation of the surface. The synthesized FeF
3
material
successfully demonstrated electrochemical activity similar to
that of a commercial source and approached theoretical values.
In addition, the material obtained from prickling liquor also
provided improved capacity retention upon cycling compared to
the commercial sample. Such results may stem from the pres-
ence of an oxide surface layer based on previous results that
demonstrated the benets of the introduction of oxygen into
the iron uoride.
19
Using a similar method, nanoparticles of FeF
3
with a small
fraction of CrF
3
in solid solution were fabricated from Fe,Cr,Ni-
pickling liquors. Such material was observed to also be elec-
trochemically active, although of slightly lower capacity than
FeF
3
, but of further improved capacity retention. Further
optimization of this positive electrode material is possible
through renement of the techniques of Badway et al. and
Yabuuchi et al. which could result in a higher capacity battery
material with improved cycling capabilities.
20,21
The ability to form solid solutions with various contents of
Cr provides tuneability that is important to describing the
potential environmental impact of the conversion process as
the composition of the waste generated is directly related to the
composition of the material crystallized. If SPL is composed
mostly of a high concentration of chromium, then the potential
environmental impact for disposal or conversation would be
much higher because of chromium's much enhanced toxicity
compared to Fe. However, for the sake of this paper, the
composition of the waste generated by converting SPL into
(Fe,Cr)F
3
through crystallization is assumed to be 30.7 g L
1
[Fe
3+
], 7.7 g L
1
[Cr
3+
], 6 g L
1
[Ni
2+
], 43.6 g L
1
[F
], and 195 g
L
1
[NO
3
]. From this information the estimated environmental
impact for converting SPL into (Fe,Cr)F
3
can be calculated. It
should be noted that this calculation is solely based on the
disposal of the euents produced from a conversion process
which yields 29% of (Fe,Cr)F
3
. Any additional processing steps
are ignored because of the very little amount of energy or added
chemicals required to complete them.
The worst case scenario for disposal of the conversion
euent is through lime neutralization. Table 7 compares the
potential environmental impact of this process with the current
practices of SPL disposal and new FeF
3
production. The current
practices generate a slightly increased environmental cost,
indicating that the worst case scenario for converting SPL into
iron uorides is only marginally more environmentally sound.
Table 5 Inputs and outputs for ferric uoride production used in the PEI calculation
10,17,24
Inputs Amount (kg t
1
FeF
3
) Outputs Amount (kg t
1
FeF
3
)
Fluorspar 1117.21170.4 CaSO
4
(marketable coproduct) 1969.8
H
2
SO
4
1383.21436.4 CaSO
4
(nonmarketable
coproduct)
2.66226.62
Fe (scrap) 504.6 CaF
2
3.1937.24
Cl
2
962.4 SO
42
0.37210.64
Fluoride 0.03720.532
Si 0.05320.532
CO
2
2.93417.604
H
2
1.07119.071
Fe 0.073350.7.335
Zn 0.0073352.2005
Heavy metals <0.00073350.8802
Solid waste 7.33551.345
HCl 969.0
Total energy 6.1710.60 GJ t
1
FeF
3
FeF
3
, anhydrous 1000
Table 6 -Inputs and outputs for crystallization euent disposal used in the PEI calculation
Inputs Outputs
SPL (L) Ca(OH)
2
(g) Fe(OH)
3
(g) Cr(OH)
3
(g) Ni(OH)
2
(g) CaF
2
(g) Ca(NO
3
)
2
(g) OH
(g)
1 201.48 58.75 15.25 9.48 89.53 258.04 53.43
57108 |RSC Adv.,2014,4,5709857110 This journal is © The Royal Society of Chemistry 2014
RSC Advances Paper
However, this margin can be dramatically increased through
optimization of the conversion process as all our calculated
values were based on an assumption of <30% yield. Simple
adaption of the crystallization techniques mentioned above of
iron-only pickling solutions to t Fe,Cr,Ni-pickling liquors will
have a signicant impact on the yield. Table 8 shows the rela-
tionship between crystallization yield and environmental
impact. This table is created by following the procedure high-
lighted in Section 6.3 above and assumes additional environ-
mental costs associated with achieving the higher
crystallization yield are insignicant. It shows that the
increased eciency produces more (Fe,Cr)F
3
with less pickling
liquor to be discarded as waste, generating a lower PEI thereby
indicating a very signicant lower environmental impact. This
clearly demonstrates the signicant impact of yield on the
environmental and likely economic costs and should be the
focus of more detailed research. The optimization further
highlights how SPL conversion is the more environmentally-
friendly technology.
A best case scenario would be to eectively create a closed
loop process to recycle the conversion euent and minimize
environmental impact. One approach proposed by a group at
Complutense University of Madrid recommends the use of KF
and KOH as a reagent to selectively precipitate iron and chro-
mium in uoride salts. Followed by hydrolysis of the precipi-
tates and neutralization of the euent, this process produces
Fe(OH)
3
, Cr(OH)
3
, and Ni(OH)
2
as solids and a solution of K
+
,
F
, and NO
3
ions that can later be treated by membrane
ltration to recover valuable acids.
22
The focus of their process
is to recycle acids back towards pickling and recover nickel as a
marketable good. A slight modication of their process through
the use of HF as a reagent would shifocus towards the
production of FeF
3
$3H
2
O and CrF
3
$3H
2
O, precursors necessary
for battery material production. Subsequent dehydration of the
crystal hydrates would produce desired positive electrode
material. This process would remove much of the metal ions in
solution and produces a waste composed of hydrouoric and
nitric acid that needs to be reconstituted with new acids before
being returned into fresh pickling liquor stream. The only
contributing factors to the environmental impact would be the
production of the acids and energy, which have a negligible
impact when compared to what is saved from the lack of SPL
disposal.
While we demonstrated that spent pickling liquor similar to
that obtained by stainless steel manufacturers can be trans-
formed into a metal uoride material that can be used as a
positive electrode material for lithium batteries and that such
process is more environmentally friendly than current practices
for the disposal of the spent liquor and iron uoride fabrica-
tion, we are well aware that costs as well as demand will drive
the adoption of the proposed process. As a result we decided to
briey touch on the subject of costs without going into detail, as
it would fall out of the scope of this manuscript. From the
stainless steel industry perspective, even if the spent pickling
liquor were donated tremendous savings would be associated to
the elimination of the disposal process. From the battery
manufacturer, savings would be associated to lower costs for an
iron uoride material of potentially improved electrochemical
performance. Lower iron uoride fabrication costs would derive
from the elimination of the anhydrous HF-based process that
requires capital-intensive installations due to the corrosive HF
gas, as well as high utility and maintenance costs. Overall, the
proposed process would not only be more environmentally
sound but it would also bring about tremendous savings to both
industries. The battery manufacturers' economical benets
would also increase as the conversion process is optimized.
Finally, ultimate cost savings would be optimized by collocating
the facilities for the conversion of the SPL near stainless steel
plants where spent liquor is produced in order to reduce
transport costs, both economic and environmental.
Conclusion
It has been demonstrated that it is possible to convert spent
pickling liquor into a viable battery material through environ-
mentally benecial process of crystallization and dehydration.
Anhydrous FeF
3
of similar electrochemical performance to a
Table 7 Comparison of the potential environmental impact of each of scenario assuming a yield of 29% for the conversion of (Fe,Cr)F
3
$3H
2
Oto
(Fe,Cr)F
3
Scenario Process
Potential environmental
impact PEI per kg of battery material
Current industry practices SPL disposal 23 281
New FeF
3
production 1.27
Total 23 282.27
Worst case scenario Conversation of SPL into (Fe,Cr)F
3
22 740
Table 8 Eect of increasing crystallization yield on quantity of
material and environmental impact
Crystallization
yield (%)
kg of (Fe,Cr)F
3
/L
of SPL Total PEI
29% 0.032 22 740
40% 0.044 16 400
50% 0.055 13 010
60% 0.066 10 750
70% 0.077 9140
80% 0.088 7940
90% 0.099 6990
This journal is © The Royal Society of Chemistry 2014 RSC Adv.,2014,4,5709857110 | 57109
Paper RSC Advances
commercially derived FeF
3
has been established using compo-
sitions similar to the euent of the stainless steel pickling
process. Even though this material shows improved capacity
retention compared to commercial FeF
3
, much can be done to
further the performance of the material such as forming
nanocomposites with a conductive matrix in situ during the
crystallization process, incorporation of oxygen and optimiza-
tion of transition metal substitution, as demonstrated in the
literature. The use of the stainless steel euent as a source raw
material for the fabrication of iron based transition metal
uorides may be of great benets to both industries and the
environment in the future as the process represents a decrease
in overall environmental impact, and economic cost as a waste
product is transformed into a value added product. For the
stainless steel industry, savings would come in the form of
dramatically lower disposal costs, while the battery manufac-
turers' economical benets would stem from lower material
cost.
Acknowledgements
The authors would like to acknowledge the Rutgers Energy
Institute and the Rutgers University Energy Storage Research
Group for nancial support.
References
1 N. Pereira, F. Badway, M. Wartelsky, S. Gunn and
G. G. Amatucci, J. Electrochem. Soc., 2009, 156, A407A416.
2 T. Mahmood, C. Lindahl and R. Davis, US Pat., US4938945,
1988.
3 K. Tjus, R. Bergstrom, U. Fortkamp, K. Forsberg and
A. Rasmuson, Development of a Recovery System for Metals
and Acids from Pickling Baths using Nanoltration and
Crystallization, IVL Swedish Environmetnal Research
Institute Ltd., Stockholm, 2006.
4 M. Sartor, D. Buchloh, F. Rogener and T. Reichardt, Chem.
Eng. J., 2009, 153,5055.
5 S. Myung, S. Sakurada, H. Yashiro and Y. Sun, J. Power
Sources, 2013, 223,18.
6 L. Liu, H. Guo, M. Zhou, Q. Wei, Z. Yang, H. Shu, X. Yang,
J. Tan, Z. Yan and X. Wang, J. Power Sources, 2013, 238,
501515.
7 F. Badway, N. Pereira, F. Cosandey and G. G. Amatucci, J.
Electrochem. Soc., 2003, 150, A1209A1218.
8 A. Gozdz, C. Schmutz, M. Tarascon and P. Warren, US Pat.,
US5418091, 1995.
9 P. Patnaik, Handbook of Inorganic Chemicals, McGraw-Hill,
New York, 2003.
10 Large Volume Inorganic Chemicals Ammonia, Acids and
Fertilizers, European Commission Institute for Prospective
Technological Studies Sustainability in Industry, 2007.
11 A. Krepler, US Pat., US4252602, 1981.
12 J. B. Stephenson, J. C. Hogan and R. S. Kaplan, Environ. Prog.,
1984, 3,5053.
13 F. H. Herbstein, M. Kapon and G. M. Reisner, Z. Kristallogr.,
1958, 171, 209224.
14 D. G. Karraker and P. K. Smith, Inorg. Chem., 1991, 31, 1118
1120.
15 B. Adamczyk, A. Hess and E. Kemnitz, J. Mater. Chem., 1996,
6, 17311735.
16 I. Plitz, F. Badway, J. Al-Sharab, A. DuPasquier, F. Cosandey
and G. G. Amatucci, J. Electrochem. Soc., 2005, 152, A307
A315.
17 Commercial Source.
18 K. Forsberg and A. Rasmuson, J. Cryst. Growth, 2006, 296,
213220.
19 W. Zhang, L. Ma, H. Yue and Y. Yang, J. Mater. Chem., 2012,
22, 2476924775.
20 F. Badway, N. Pereira, F. Cosandey and G. G. Amatucci, J.
Electrochem. Soc., 2003, 150, A1318A1327.
21 N. Yabuuchi, M. Sugano, Y. Yamakawa, I. Nakai,
K. Sakamoto, H. Muramatsu and S. Komaba, J. Mater.
Chem., 2011, 21, 1003510041.
22 J. Hermoso, J. Dufor, J. L. Galvez, C. Negro and F. Lopez-
Mateos, Ind. Eng. Chem. Res., 2005, 44, 57505756.
23 Stainless Steel Manufacturer.
24 Additional Information submitted during information exchange
on Large Volume Inorganic Chemicals Solids and Other
Industry, European Commission Institute for Prospective
Technological Studies Sustainability in Industry, 2005.
57110 |RSC Adv.,2014,4,5709857110 This journal is © The Royal Society of Chemistry 2014
RSC Advances Paper
... At present, different pickling systems have been developed for different steel grades, mainly including HCl-based, H 2 SO 4 -based, HNO 4 -HF-based, and HCl-H 2 O 2 -based systems. The SPL is generally rich in metal cations, with a typical composition of 30-45 g/L Fe 3+ , 5-10 g/L Cr 3+ , 3-5 g/L Ni 2+ besides a large amount of residual acid [2][3][4]. A large amount of SPL is generated from the stainless-steel pickling and discharged without any treatment, which have caused serious pollution issues [5]. ...
... A Eh-pH diagram of the Fe-Cr-Ni-H 2 O system was drawn using FactSage 8.1, giving the result shown in Fig. 2. Fe 3+ , Cr 3+ , and Ni 2+ exist in the SPL for pH values < 2.6. Iron is most likely to appear in the form of Fe(OH) 3 for pH values > 2.6, and chromium and nickel remain in solution. Moreover, chromium exists in the form of Cr(OH) 3 for pH values > 4.1, while nickel remains solution. ...
... Iron is most likely to appear in the form of Fe(OH) 3 for pH values > 2.6, and chromium and nickel remain in solution. Moreover, chromium exists in the form of Cr(OH) 3 for pH values > 4.1, while nickel remains solution. As the alkalinity increases, the co-precipitation of Fe 3+ , Cr 3+ , and Ni 2+ may be promoted for pH > 9.5. ...
Article
Spent pickling liquor (SPL) from the process of the stainless-steel pickling and lithium-ion battery anode material have similar elemental compositions. Inspired by this fact, a process of preparing lithium-ion battery anode materials from SPL is proposed in this work. Key steps of the preparation of precursor by co-precipitation, preparation of composite transition metal oxide (CTMO) by high-temperature ball milling, and preparation and electrochemical performance of anode materials were investigated. The Fe, Cr, and Ni resources of SPL were recovered by co-precipitation at pH = 11, and the leachate met the effluent discharge standard. The temperature of ball milling was studied in detail, with results indicating that the decomposition of CTMO precursors was incomplete at 400 °C and 500 after 3 h of treatment. Raising the temperature increases the reaction rate. For the ball milling process, a temperature of 600 °C is suggested. An anode material MxOy/C, with M = Fe, Cr, and Ni was manufactured, and its electrochemical performance was found to be similar or even superior to that of a single oxide. The process proposed in this study is feasible and has been successfully demonstrated, which provides a new approach to the treatment of similar wastes in the metallurgical field.
... *Each extraction method was initiated with 10 mL of culture filterate (pH 9.0) containing 44 mg of rhamnolipid biosurfactant. Moreover, the use of FeCl 3 as a coagulant can be justified for larger scale recovery of anionic biosurfactants being least expensive as generated as a waste material from steel making operations [33,34]. ...
Article
Sequential fill-and-draw fermentation strategy provides an approach to increase the productivity by replenishing nutrients and minimizing the toxic effects of by-products. In the present work, the same strategy was adopted using lignocellulosic industrial rice-straw C6 hydrolysate stream to produce rhamnolipids from Achromobacter sp. (PS1) in a 6 L bioreactor with a working-volume of 2 L. The production results showed overall rhamnolipid production of 22.03 g/L in 15 days observed at par with 19.35 g/L obtained under shake flask conditions in 18 days. At each sequential feed (2% sugars), a rise in dissolved oxygen (D.O) concentration was observed in the range between 60-53% which declined to 47-39% with consecutive depletion in sugar concentration under no D.O control. For maximum extraction of rhamnolipids from culture broth, the synergistic effect of sweep floc-coagulation using FeCl3 at 0.4% (w/v) followed by its acidification and solvent extraction was adopted which resulted in maximum recovery of 97.5% compared to 89.05% recovery obtained in simply acidification followed by solvent extraction. The characterization of partially purified biosurfactant using tandem-MS revealed six-congeners, Rha-C10-C10 and Rha-Rha-C10-C10 being the most abundant. Oil recovery of 92.21% from motor-oil impregnated sand using crude rhamnolipid further added the value to the biosurfactant.
Article
Full-text available
The management of spent pickling acids (SPA) is an environmental challenge for the hot-dip galvanizing (HDG) industry. Bearing in mind its elevated content of iron and zinc, SPA can be regarded as a source of secondary materials in a circular economy approach. This work reports the pilot scale demonstration of non-dispersive solvent extraction (NDSX) in hollow fiber membrane contactors (HFMCs) to perform the selective zinc separation and SPA purification, so that the characteristics needed for use as a source of iron chloride are achieved. The operation of the NDSX pilot plant, which incorporates four HFMCs with a 80 m2 nominal membrane area, is carried out with SPA supplied by an industrial galvanizer, and consequently technology readiness level (TRL) 7 is reached. The purification of the SPA requires of a novel feed and purge strategy to operate the pilot plant in continuous mode. To facilitate the further implementation of the process, the extraction system is formed by tributyl phosphate as the organic extractant and tap water as the stripping agent, both easily available and cost-effective chemicals. The resulting iron chloride solution is successfully valorized as a hydrogen sulfide suppressor to purify the biogas generated in the anaerobic sludge treatment of a wastewater treatment plant. Additionally, we validate the NDSX mathematical model using pilot scale experimental data, providing a design tool for process scale-up and industrial implementation.
Article
Full-text available
A greener technology aiming at a smarter industrial waste treatment is proposed to produce chloride iron-zinc-aluminum layered double hydroxides (LDHs). Waste Pickling Acid (WPA) and sodium aluminate (NaAlO2) from secondary sources were meticulously mixed under mild experimental conditions using a sodium hydroxide solution as a pH-regulator. A set of characterization techniques (XRD, SEM, TGA, FTIR, AAS and adsorption-desorption of N2) indicated the formation of highly-dispersed nanoflake crystallites with textural characteristics and thermal stability similar to syntheses with high-quality chemicals. An interesting discussion on chemical composition and M2+/M3+ molar ratio is presented. Although the co-precipitation synthesis was conducted without control of environmental CO2, complete intercalation of the chloride anion was achieved, making these particles more favorable for further anion exchange applications. The experimental variables temperature of reaction and WPA/NaAlO2 volume ratio showed the strongest influence on the LDHs crystallinity and porosity. LDHs architected with iron and zinc have the potential to be applied in systems for removing sulfur gases for cleaner energy production, e.g. in the refining process of biogas to produce biomethane.
Article
This review considers key parameters for affordable Li-ion battery (LIB) – powered electric transportation, such as mineral abundance for active material synthesis, raw materials’ processing cost, cell performance characteristics, cell energy density, and the cost of cell manufacturing. We analyze the scarcity of cobalt (Co) and nickel (Ni) resources available for intercalation-type LIB cathode materials, estimate the demands for these metals by transportation and other industries and discuss risk factors for their price increase within the next two decades. We further contrast performance and estimates costs of LIBs based on intercalation materials, such as lithium nickel cobalt manganese oxide (NCM), lithium nickel cobalt aluminum oxide (NCA), lithium iron phosphate (LFP) and other oxide-based cathodes and carbonaceous anodes, with those of LIBs based on conversion-type active materials, such as lithium sulfide (Li2S) and lithium fluoride/iron (Fe) and copper (Cu)-based cathodes and silicon (Si)-based anodes. Our analyses of industry data suggest that in the long-term the LIB price will be dominated by cost of the cathode materials. In addition, the cost contributions of manufacturing, overhead and inactive materials will be reversely proportional to the cell energy density. As such, we expect that to-be developed energy-dense conversion-type LIBs should be able to reach the $30–40/kWh by around 2040–2050, while the intercalation-type LIBs will likely be 60% more expensive and sensitive to the Ni price variations. By analyzing the availability and costs of lithium (Li), sulfur (S), Si, fluorine (F), Fe and Cu we conclude that the lower cost, broader accessibility, much greater abundance, and improved health and safety aspects of employing conversion-type chemistries should warrant dedication of substantial efforts in their development. Furthermore, we predict that based on pure economics, the widespread introduction of zero carbon-emission transportation and sustainable energy sources is inevitable and independent on the winning LIB chemistry.
Article
Full-text available
A stainless-steel 321 dental substrate significantly corroded within Porphyromonas gingivalis growth culture in artificial saliva culture suspension, with and without NaF additive.
Article
Iron fluoride trihydrate can be used to prepare iron hydroxyfluoride with the hexagonal-tungsten-bronze (HTB) type structure, a potential cathode material for batteries. To understand this phase transformation, a structural description of β-FeF3·3H2O is first performed by means of DFT calculations and Mössbauer spectroscopy. The structure of this compound consists of infinite chains of [FeF6]n and [FeF2(H2O)4]n. The decomposition of FeF3·3H2O induces a collapse and condensation of these chains, which lead to the stabilization, under specific conditions, of a hydroxyfluoride network FeF3-x(OH)x with the HTB structure. The release of H2O and HF was monitored by thermal analysis and physical characterizations during the decomposition of FeF3·3H2O. An average distribution of FeF4(OH)2 distorted octahedra in HTB-FeF3-x(OH)x was obtained subsequent to the thermal hydrolysis/olation of equatorial anionic positions involving F(-) and H2O. This study provides a clear understanding of the structure and thermal properties of FeF3·3H2O, a material that can potentially bridge the recycling of pickling sludge from the steel industry by preparing battery electrodes.
Article
We have reinvestigated the crystal structure of the low-dimensional fluoride $\beta$-FeF$_3(H_2O)_2\cdot$H$_2$O using high resolution neutron and X-ray diffraction data. Moreover we have studied the magnetic behavior of this material combining medium resolution and high flux neutron powder diffraction together with magnetic susceptibility measurements. This fluoride compound exhibits vertex-shared 1D Fe$^{3+}$ octahedral chains, which are extended along the \textit{c}-axis. The magnetic interaction between adjacent chains involve super-superexchange interactions via an extensive network of hydrogen bonds. This interchain hydrogen bonding scheme is sufficiently strong to induce a long range magnetic order appearing below T$_N$ = 20(1) K. The magnetic order is characterized by the propagation vector \textbf{k}=$\left(0, 0, \frac{1}{2}\right)$, giving rise to a strictly antiferromagnetic structure where the Fe$^{3+}$ spins are lying with the \textit{ab}-plane. Magnetic exchange couplings extracted from magnetization measurements are found to be J$_{\parallel}$/k$_b$ = -18 K and J$_{\perp}$/k$_b$ = -3 K. These values are in good agreement with the neutron diffraction data, which show that the system became antiferromagnetically ordered at ca. TN = 20(1) K.
Article
The influence of hydrofluoric acid and nitric acid concentration on the growth rate of β-FeF3·3H2O crystals has been studied in different hydrofluoric acid (4.7–10.7 mol/(kg H2O)) and nitric acid (2.1–4.6 mol/(kg H2O)) mixtures at 50 °C. Seeded desupersaturation experiments were performed and the results were evaluated by considering the chemical speciation using two different speciation programs. The growth rate at 50 °C at a supersaturation ratio of 2, expressed in terms of free FeF3, was found to be in the range of (0.4–3.8)×10−11 m/s. The growth rate order was found to be two or higher in all experiments. The low growth rate and high growth rate order indicate that the growth rate is governed by the surface integration step. The growth rate was found to be independent of variations in acid concentrations; this is in accordance with the assumption of a surface integration controlled growth rate.
Article
Full-text available
The practical electroactivity of electrically insulating iron fluoride was enabled through the use of carbon-metal fluoride nanocomposites (CMFNCs). The nanocomposites were fabricated through the use of high energy mechanical milling and resulted in nanodomains of FeF, on the order of 1-20 nm encompassed in a matrix of carbon as characterized by transmission electron microscopy and X-ray diffraction (XRD) Electrochemical characterization of CMFNCs composed of 85/15 wt % FeF 3 /C resulted in a nanocomposite specific capacity as high as 200 mAh/g (235 mAh/(g of FeF 3 ) with the electrochemical activity associated with the Fe 3+ → Fe 2+ occurring in the region of 2.8-3.5 V. The CMFNCs revealed encouraging rate capability and cycle life with <10% fade after 50 cycles. Structural evolution during the first lithiation reaction was investigated with the use of ex situ and in situ XRD. Initial results suggest that x from 0 to 0.5 in Li x FeF 3 proceeds in a two-phase reaction resulting in a phase with significant redistribution of the Fe atoms within a structure very similar to the base FeF 3 . FeF 3 -based CMFNCs also exhibited a very high specific capacity of 600 mAh/g at 70°C due to a reversible reaction at approximately 2 V.
Article
The crystal structures of the green compounds CrF3 · 3H2O and CrF3 · 5H2O have been determined. The first is rhombohedral [aR = 5.668(4) Å, αR = 112.5(1)°, space group R[unk]m, Z = 1] and the second orthorhombic [a = 10.396(5), b = 8.060(5), c = 7.965(4) Å, space group Pbcn, Z = 4]. Both compounds contain the same octahedral molecules Cr[F3(H2O)3], but the nature of the isomers present could not be determined because of crystal disorder. There is extensive hydrogen bonding in both types of crystal. It has been found that violet Cr[H2O]6F3 · 3H2O decomposes thermally to green Cr[F3(H2O)3] in three stages, with violet Cr[H2O]6F3 identified as the product formed in the first stage.
Article
FeF3·3H2O, FeF3·0.33H2O, and FeF3 have been synthesized via a liquid-phase method followed by heat treatment at different temperatures. The structure and performance of these iron fluorides have been characterized by X-ray diffraction (XRD), scanning electron microscopy (SEM), high-resolution transmission electron microscopy (HRTEM), selected-area electron diffraction (SAED), charge–discharge test, cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS) and galvanostatic intermittent titration technique (GITT). Though FeF3·3H2O, FeF3·0.33H2O, and FeF3 have different crystalline structures, they can achieve the identical reversible electrochemical conversion reaction from Fe3+ to Fe0 in the wide voltage range (1.0–4.5 V vs. Li+/Li). Among these three iron fluorides, FeF3·0.33H2O shows the best electrochemical performance. Moreover, ball milling with acetylene black combined with limiting cut-off voltage can further improve its electrochemical performance. FeF3·0.33H2O/C composite delivers excellent electrochemical performance in the voltage range of 2.0–4.5 V. Its capacity retentions remain as high as 83.8% and 83.3% after 100 cycles at 0.1 and 5 C, respectively. This study suggests a potential feasibility to prepare the optimal crystal structure of iron fluorides as high-performance cathode material for lithium-ion batteries.
Article
Facile synthesis of rhombohedral type FeF3 introduced via two consecutive steps is introduced: i) acidic treatment of Fe2O3 followed by thermal evaporation at 80 °C resulting in hydrated β-FeF3·3H2O and ii) a simple thermal decomposition of the as-received β-FeF3·3H2O at 400 °C under an Ar atmosphere. A Rietveld refinement of x-ray diffraction data for the as-synthesized FeF3 indicates the formation of a highly crystalline FeF3 structure with a R3¯c space group. To overcome the high ionicity and improve the diffusivity, FeF3 is ball-milled with the aid of carbon (acetylene black). The electrochemical performance of nanosized FeF3 is not favored in voltage range of 1.5–4.5 V because the repetitive intercalation–conversion reaction accelerates the structural disruption within a few cycles, although a high capacity (518 mAh (g-fluoride)−1 at 20 mA g−1) is observed, assisted by the three-electron redox of Fe3+/0. Raising the lower cut-off voltage to 2 V, which allows only intercalation reaction, the FeF3 delivers a high capacity of 224 mAh g−1 with significantly improved capacity retention (71% at 100th cycle).
Article
Here, a novel architecture of a core–shell structured FeF3@Fe2O3 composite with particle size of 100–150 nm and tunable Fe2O3 content is synthesized by a simple heat treatment process utilizing FeF3 with fine network structure as precursor. The structure, morphology and electrochemical performance of the pristine FeF3 and the FeF3@Fe2O3 composites are studied by XRD, SEM, TEM and discharge–charge measurements. XRD results show that the Bragg peaks of the FeF3@Fe2O3 composites are well indexed to FeF3 and Fe2O3. SEM and TEM images reveal the core–shell structure of the composites. The comparison of the electrochemical performance between the pristine FeF3 and the FeF3@Fe2O3 composites reveals that the in situ Fe2O3 coating (even with small amount, 0.6–5.2 wt%) has great influence on the improvement of electrochemical performance.
Article
An FeF3 nanocomposite with carbon materials is prepared by a ball milling process. The effect of the strain induced by ball milling on the electrochemical performance is examined using a combination of synchrotron X-ray diffraction (SXRD) and X-ray absorption spectroscopy (XAS). The strain of the FeF3 particles in the nanocomposite is analyzed by applying the Williamson–Hall method to the SXRD patterns. Heat-treatmenting the FeF3 nanocomposite drastically relieves the strain induced by the ball milling. The electrochemical performance of the FeF3 nanocomposite is also significantly improved by the heat-treatment process at 350 °C. The FeF3 nanocomposite heat-treated at 350 °C delivers 200 mA h g−1 of reversible capacity with good capacity retention in a voltage range of 2.0–4.5 V. The heat-treatment process suppresses the increase in the poralization during the continuous cycling test. The power density of the heat-treated sample is superior to that of the ball milled sample without the heat treatment.
Article
Crystal growth of β-FeF3·3H2O has been investigated in mixtures of 3 mol kg-1 hydrofluoric acid and 3 mol kg-1 nitric acid at 30, 40 and 50 °C. Seeded isothermal desupersaturation experiments have been performed in the range: 1.3<S<3.6. Solution samples were analysed for total iron concentration with inductively coupled plasma atomic emission spectroscopy. The true supersaturation driving force was estimated by a proper speciation using the software SSPEC using appropriate stability constants. Growth rate parameters of the BCF surface diffusion growth rate equation and the empirical power-law equation have been estimated by fitting the supersaturation balance equation using a nonlinear optimization procedure. The results show that the growth rate is surface integration controlled. The growth rate at a supersaturation ratio of 2 was found to be 3.5×10-12 m s-1 at 30 °C, 7.4×10-12 m s-1 at 40 °C and 16×10-12 m s-1 at 50 °C. The activation energy of the rate constant of crystal growth was found to be 61 kJ mol-1.
Article
The calcination of α-CrF3·3H2O results in the formation of a chromium hydroxyfluoride with the pyrochlore structure. The stepwise replacement of chromium by iron and magnesium leads to considerable alterations in the structure and the surface properties of the calcination products, accompanied by significant changes in the catalytic activity. The dismutation of dichlorodifluoromethane and the dehydrochlorination of 1,1,1 -trichloroethane act as a probe reactions for Lewis acid sites. The syntheses of the catalysts were carried out by coprecipitation of mixed metal fluoride trihydrates and subsequent calcination procedures. The stepwise replacement with iron leads to a rebuilding of the lattice from the cubic pyrochlore structure of CrF3–x(OH)x into the pseudo-hexagonal tungsten bronze (HTB) structure of β-FeF3. The maximum catalytic activity towards CH3CCl3 dehydrochlorination was obtained for the 65% iron sample, which is accompanied by a maximum BET surface area and a maximal number of Lewis acid sites. CrF3–x(OH)x exhibits a dramatic loss of catalytic activity as well as BET surface area. Possible explanations are given by comparing the pseudo-HTB structure with the cubic pyrochlore structure with regard to the accessibility of the Lewis acid metal cations. Upon substitution of chromium by magnesium we obtained a maximum Lewis acidity for samples with 65–92% magnesium leading to a corresponding maximum in catalytic activity. The Brønsted acidity of both systems is predominantly weak. Bulk and surface hydroxy groups are distinguished.
Article
Utilizing a solid-state redox-driven conversion reaction enabled by mechanochemistry, conductive C:FeF3 nanocomposites were fabricated from insulative CF1 :FeF2 precursors. All reactions were characterized by X-ray diffraction and Fourier transform infrared spectroscopy. The latter provided insights to the progression of the CF and C phases and the metal fluoride during the course of the reaction. Such nanocomposites resulted in a four order of magnitude increase in electrical conductivity and enabled excellent specific capacity approaching 500 mAh/g vs. Li with good reversibility, although at slow rates. Utilizing the theoretical basis of the technique, other couples were examined to experimentally isolate the oxidative power of CF,. In the process, we have also shown that a composite of CF1 :CrF2 can be easily converted to C:CrF3. The resulting nanocomposite exhibited a specific capacity of 682 mAh/g at-an average voltage of approximately 1.9 V. The technique is also a powerful method for the fabrication of single phase metal fluoride solid solutions, as demonstrated with the fabrication of Cr0.5Fe0.5F3.