You are on page 1of 42

Subscriber access provided by - Access paid by the | UCSB Libraries

Article
Synthesis of Water Soluble and Catalytically
Active Gold-decorated Cobalt Ferrite Nanoparticles
Alessandro Silvestri, Sara Mondini, Marcello Marelli, Valentina Pifferi,
Luigi Falciola, Alessandro Ponti, Anna Maria Ferretti, and Laura Polito
Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.6b01266 • Publication Date (Web): 22 Jun 2016
Downloaded from http://pubs.acs.org on July 1, 2016

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Langmuir is published by the American Chemical Society. 1155 Sixteenth Street N.W.,
Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 41 Langmuir

1
2
3
4
5
6
7
8
Synthesis of Water Dispersible and Catalytically
9
10
11
12 Active Gold-decorated Cobalt Ferrite Nanoparticles
13
14
15
16
17 Alessandro Silvestri†‡, Sara Mondini†‡, Marcello Marelli†, Valentina Pifferi‡, Luigi Falciola‡,
18
19 Alessandro Ponti †, Anna Maria Ferretti*† and Laura Polito*†
20
21
22
23
† Laboratorio di Nanotecnologie, Istituto di Scienze e Tecnologie Molecolari, Consiglio
24 Nazionale delle Ricerche, Via G. Fantoli 16/15, 20138 Milan, Italy
25
26 ‡ Department of Chemistry, University of Milan, Via C. Golgi 19, 20133 Milan, Italy
27
28
29
30
31
32
33
ABSTRACT. Hetero-nanoparticles represent an important family of composite nanomaterials
34
35 that, in the last years, is attracting ever growing interest. Here, we report a new strategy for the
36
37 synthesis of water dispersible cobalt ferrite nanoparticles (CoxFe3-xO4 NPs) decorated with
38
39
40 ultrasmall (2-3 nm) gold nanoparticles (Au NPs). The synthetic procedure is based on the use of
41
42 2,3-meso-dimercapto succinic acid (DMSA), which plays a double role. Firstly, it transfers
43
44 cobalt ferrite NPs from the organic phase to aqueous media. Secondly, the DMSA reductive
45
46
47 power promotes the in-situ nucleation of gold NPs in proximity of the magnetic NP surface.
48
49 Following this procedure, we achieved a water dispersible nano-system (CoxFe3-xO4-DMSA-Au
50
51
NPs) which combines the cobalt ferrite magnetic properties with the catalytic features of
52
53
54 ultrasmall Au NPs. We showed that CoxFe3-xO4-DMSA-Au NPs act as an efficient nano-catalyst
55
56 to reduce 4-nitrophenol to 4-aminophenol, and that they can be magnetically recovered and
57
58
59
60
ACS Paragon Plus Environment 1
Langmuir Page 2 of 41

1
2
3
recycled. It is noteworthy that such nanosystem is more catalytically active than Au NPs with
4
5
6 equal size. Finally, a complete structural and chemical characterization of the hetero-NPs is
7
8 provided.
9
10
11
12 1. INTRODUCTION
13
14
15
16
One of the most compelling goals in the field of nanotechnology is the design and synthesis of
17
18 hetero-nanoparticles, combining in a single entity the physical and chemical properties of the
19
20 composing substances.1-5 In addition to this additive multi-functionality, hybrid nanostructures
21
22
23 can synergistically enhance the capabilities of the individual component due to proximity effects
24
25 or electronic coupling at the interface.6 By engineering at the nanoscale, it is possible to design
26
27 and create new materials with improved and fascinating properties, which can be employed in
28
29
30 many applications.1-5 Among a wide number of papers related to the production of hybrid nano-
31
32 structures,1-7 a large focus has been placed on the synthesis of hybrid magnetic nanoparticles
33
34 (NPs), mainly based on magnetite/maghemite (Fe3O4/γ-Fe2O3) and gold.8 These materials have
35
36
37 been extensively studied and synthesized with different types of morphologies like core-shell,9-11
38
39 inverted core-shell,12 and dumbbell.13-15 The reason for this huge interest resides in the large
40
41
42
number of applications in which gold-coated magnetic NPs can be used. In fact, tailoring their
43
44 features, these materials can be applied in the fields of nanomedicine,7,16 biosensors,16,17 cell
45
46 separation16,17 and catalysis.18,19 The optical and catalytic activity of gold can be modulated by
47
48
49 sizing and shaping the Au NP morphology or the hetero-nanoparticle geometry. For example, the
50
51 plasmon resonance peak of Au could be shifted to the NIR region by increasing size or
52
53 introducing branches on the NP surface.20 Furthermore an increase in the surface/volume ratio
54
55
56 plays a fundamental role in the catalytic activity of nanometric gold.21,22 On the other hand,
57
58
59
60
ACS Paragon Plus Environment 2
Page 3 of 41 Langmuir

1
2
3
magnetic properties can be fine tuned controlling the size and composition of ferrite NPs. For
4
5
6 instance, in comparison with magnetite/maghemite (Fe3O4/γ-Fe2O3), cobalt ferrite (CoFe2O4)
7
8 NPs present improved magnetic properties due to their high relaxivity and magnetic
9
10
11
anisotropy.23,24 Despite advantageous magnetic properties, few examples of hybrid systems
12
13 comprising cobalt ferrites and gold have been described.25-28 Moreover, an inherent difficulty in
14
15 the reproducible synthesis of these hetero-nanoparticles hamper their widespread exploitation.2
16
17
18 Herein, we report a reliable synthetic strategy to obtain water dispersible cobalt ferrite NPs
19
20 decorated with ultrasmall gold NPs (CoxFe3-xO4-DMSA-Au), by means of a facile protocol. The
21
22 multistep strategy is based on the solvothermal synthesis of cobalt ferrite nanoparticles,
23
24
25 stabilized by oleic acid and oleylamine (CoxFe3-xO4-OA) in organic solvent, followed by a ligand
26
27 exchange procedure to bring the NPs in aqueous environment, by using 2,3-meso-dimercapto
28
29
succinic acid (CoxFe3-xO4-DMSA). Next, the DMSA coating the magnetic NP surface was
30
31
32 exploited to reduce AuCl4- and to induce an in situ nucleation of gold NPs exclusively in the
33
34 proximity of the cobalt ferrite surface, without using any further reducing agents or preformed
35
36
37 gold seeds. The presence of a wide available surface area of noble metal onto a nanosized
38
39 magnetic support, makes this material a promising candidate to act as an efficient recyclable
40
41 catalyst. Therefore, taking advantage of their water solubility and magnetization, we studied the
42
43
44 catalytic reduction of 4-nitrophenol (4-NP) to 4-aminophenol (4-AP), monitoring the reaction
45
46 progress by UV-vis spectroscopy. The chance to reduce, at room temperature and in mild
47
48 conditions, water pollutants is presently of crucial interest.29,30 Nitroaromatics, in fact, are an
49
50
51 important class of environmentally toxic compounds widely employed in the production of
52
53 explosives, pesticides, plastics and pharmaceuticals.29 Noteworthy, the molecule resulting from
54
55
56
this reduction (4-AP) can find applications in many fields, underlying the opportunity to carry on
57
58
59
60
ACS Paragon Plus Environment 3
Langmuir Page 4 of 41

1
2
3
new studies to continuously improve the reduction efficiency.29,30 The morphological, elemental
4
5
6 and structural features of hetero-nanoparticles were evaluated by Transmission Electron
7
8 Microscopy (TEM), Inductively Coupled Plasma-Optical Emission Spectrometers (ICP-OES),
9
10
11
UV-Vis spectroscopy and Dynamic Light Scattering (DLS) analysis while the magnetic
12
13 characterization was performed by a Quantum Design MPMS XL-5 SQUID magnetometer.
14
15
16
17
18 2. EXPERIMENTAL SECTION
19
20
21 2.1 Materials. All reagents, unless cited otherwise, were purchased from Sigma-Aldrich,
22
23 and used without any further purification. HAuCl4•3H2O was stored as a 10 mM solution in
24
25
26 MilliQ water, at 4 °C and protected from light. The glassware used for gold NPs synthesis were
27
28 cleaned with aqua regia (HCl (37%) / HNO3 (65%) 3/1). Ultra-pure MilliQ water was used for
29
30
31
the preparation of all the aqueous solutions. The concentration of water dispersible NPs was
32
33 performed by means of Millipore Amicon® Ultra centrifugal filters with a 30 kDa cut-off for
34
35 CoxFe3-xO4-DMSA NPs and CoxFe3-xO4-DMSA-Au NPs, and with a cut-off of 3 kDa for Au-
36
37
38 DMSA NPs. The purification of NPs was performed by using dialysis tubes with a cut-off of 14
39
40 kDa. The dropped addition of HAuCl4 to the reaction mixture was performed with the aid of a
41
42 NE-1010 Higher Pressure Programmable Single Syringe Pump.
43
44
45
46 2.2. Synthesis of CoxFe3-xO4-OA NPs. CoxFe3-xO4-OA NPs were synthesized according to a
47
48 literature procedure.31 A solution of Fe(acac)3 (350 mg, 0.99 mmol), Co(acac)2 (128 mg, 0.495
49
50
mmol), oleic acid (1 mL, 3.15 mmol), oleylamine (1 mL, 3.04 mmol), 1,2- hexadecanediol (1.15
51
52
53 g, 4.45 mmol) in benzyl ether (25 mL) was warmed up to 120 °C under argon flow and kept at
54
55 this temperature for 1 h. The mixture was first heated to 210 °C with a rate of 8 °C/min and
56
57
58
maintained at this temperature for 2 h. Then, the temperature was raised to 300 °C with a rate of
59
60
ACS Paragon Plus Environment 4
Page 5 of 41 Langmuir

1
2
3
3 °C/min and maintained for 1 h. After cooling the reaction mixture to room temperature, the
4
5
6 obtained NPs were collected by centrifugation (6000 rpm) after the addition of a mixture of
7
8 ethanol/acetone 1/1 and then redissolved in hexane. This treatment was repeated two times. After
9
10
11
the last precipitation, CoxFe3-xO4-OA NPs were stored dissolved in hexane (20 mL) and stored at
12
13 room temperature.
14
15
16 2.3. Synthesis of water dispersible CoxFe3-xO4-DMSA NPs. The ligand exchange procedure
17
18
19 leading to the title product was performed in two steps. In the first step, 3 mL of hexane solution
20
21 of CoxFe3-xO4-OA were precipitated by centrifugation (after adding 10 mL of ethanol/acetone 1/1
22
23 mixture) and redissolved in 6 mL of toluene. A solution of DMSA (33 mg, 0.18 mmol) in DMSO
24
25
26 (2 mL) was dropped into a round-bottom flask (immersed in a ultrasonic bath) containing the
27
28 toluene dispersion of CoxFe3-xO4-OA NPs. The mixture was sonicated for 1 h and further
29
30
31
magnetically stirred for 2 days. Afterwards, 4 mL of toluene were added to aid the precipitation
32
33 of the CoxFe3-xO4-DMSA NPs. The NPs were magnetically recovered and washed consecutively
34
35 three times with ethanol and two times with acetone, and finally suspended in MilliQ water (10
36
37
38 mL). Then, to improve aqueous dispersion, the solution was basified to pH 10 (using aq. NaOH
39
40 solution, 0.1 M) and subsequently brought back to pH 7 (using aq. HCl solution, 0.1 M).26 In the
41
42 second step, a solution of DMSA (20 mg, 0.1 mmol) in water (10 mL) was added dropwise to the
43
44
45 NP dispersion, and the mixture was sonicated for 1 h. Then the mixture was magnetically stirred
46
47 for further 12 h and purified by dialyzing against MilliQ water for 72 h (changing water every 8-
48
49
10 h). Finally, the recovered CoxFe3-xO4-DMSA NPs were stored as diluted aqueous solution
50
51
52 (250 mL), with an iron concentration of 34 µg/mL. Such stock solution was stable for months at
53
54 room temperature.
55
56
57
58
59
60
ACS Paragon Plus Environment 5
Langmuir Page 6 of 41

1
2
3
2.4. Synthesis of CoxFe3-xO4-DMSA-Au NPs. In a round-bottom flask, 10 mL of the stock
4
5
6 solution of CoxFe3-xO4-DMSA NPs were stirred (750 rpm) under N2 atmosphere at room
7
8 temperature. An aqueous solution of HAuCl4 (10 mL, 0.2 mM) was slowly dropped into the
9
10
11
reaction mixture with the aid of a syringe pump (flow rate: 41.5 µL/min). After completing the
12
13 addition, the mixture was stirred for an additional hour. The NPs were concentrated, dialyzed
14
15 against MilliQ water for 48 h (changing water every 8-10 h) and stored as aqueous dispersion (2
16
17
18 mL) with an iron concentration of 140 µg/mL (2.51 mM) and a gold concentration of 86 µg/ml
19
20 (0.44 mM).
21
22
23 2.5. Synthesis of Au-DMSA NPs. Au-DMSA NPs were synthesized following the method
24
25
26 published by Negishi et al.32 DMSA (1.8 mg, 10 µmol) was dissolved in MilliQ water (0.5 mL)
27
28 and added to an aqueous solution (5 mL) of HAuCl4 (1.7 mg, 5 µmol). The reaction mixture was
29
30
31
stirred (750 rpm) for 10 min at room temperature, and the color of the solution changed from
32
33 yellow to light brown. The NPs were washed five times with MilliQ water by centrifugation,
34
35 using 3 kDa centrifugal filters units. The recovered Au-DMSA NPs were dissolved and stored in
36
37
38 aqueous solution (4 mL) at a gold concentration of 240 µg/mL.
39
40
41 2.6. Characterization of NPs. The morphology and size distribution of the NPs were evaluated
42
43 using transmission electron microscopy by means of a Zeiss LIBRA 200FE microscope,
44
45
46 equipped with a 200 kV FEG an in-column second-generation omega filter for energy selective
47
48 spectroscopy, and a STEM facility with HAADF detector. Both TEM and STEM imaging modes
49
50
were employed. Electron energy loss spectroscopy (EELS) and imaging (ESI), and energy
51
52
53 dispersive X-ray spectroscopy (EDX, OXFORD X-Stream 2 and INCA software) were used to
54
55 determine the elemental ratio and distribution inside the nanostructure. The samples for TEM
56
57
58
analysis were prepared depositing a drop of NP dispersion on a carbon-coated copper grid (300
59
60
ACS Paragon Plus Environment 6
Page 7 of 41 Langmuir

1
2
3
mesh) and evaporating the solvent. The particle size distribution was obtained by measuring at
4
5
6 least 500 nanoparticles, analyzing TEM images by the software Pebbles and Pebble Juggler.33
7
8 UV-Vis spectra were recorded by means of an Agilent 8453 instrument. DLS and ζ-potential
9
10
11
measurements were carried out by a 90 plus Particle Size Analyzer (Brookhaven Instrument
12
13 Corporation) equipped with a solid state He–Ne laser (wavelength = 661 nm). Experiments were
14
15 carried out at a scattering angle of 90° on diluted NP samples at 298 K. NP samples for DLS
16
17
18 were filtered through 0.45 µm cellulose acetate syringe filters before measuring. Each sample
19
20 was allowed to equilibrate for 3 min prior to starting the experiment. Ten independent
21
22 measurements of 60 s duration were performed for each sample. The NP hydrodynamic
23
24
25 diameters were calculated using Mie theory, considering absolute viscosity and refractive index
26
27 values of the medium to be 0.911 cP and 1.334, respectively. The ζ-potential was calculated from
28
29
the NP electrophoretic mobility using the Smoluchowski theory, assuming a medium viscosity of
30
31
32 0.891 cP, medium dielectric constant of 78.6, and Henry function of 1.5. FTIR spectra were
33
34 recorded by a Thermo Nicolet NEXUS 670 FTIR spectrometer, using KBr pellets. Gold, iron
35
36
37 and cobalt content of NPs was determined via Inductively Coupled Plasma-Optical Emission
38
39 Spectrometry (ICP-OES; iCAP 6300 Duo, Thermofisher). Samples were heated and acid
40
41 digested with aqua regia (500 µL for 5 times); the residues were recovered with HCl 37% (500
42
43
44 µL) and diluted with MilliQ water to a known volume. Multi-element standards containing the
45
46 three metals were used to create the external calibration line. The magnetic characterization was
47
48
49
performed by a Quantum Design MPMS XL-5 SQUID magnetometer. Weighted amounts of
50
51 powdered samples were packed in Teflon ribbon. The temperature dependence of the
52
53 magnetization was recorded in the temperature range T = 5 – 300 K using a measuring field
54
55
56 Hmeas = 10 Oe in both zero-field cooling (ZFC) and FC (Hcool = 10 Oe) modes. Magnetization
57
58
59
60
ACS Paragon Plus Environment 7
Langmuir Page 8 of 41

1
2
3
isotherms (hysteresis loops) were measured between +50 and –50 kOe at T = 5 K after cooling in
4
5
6 zero field. All data were corrected for diamagnetic contributions.
7
8
9 2.7. Catalytic reduction of 4-nitrophenol. The reaction was performed in a 3 mL quartz cell,
10
11
12 equipped with a magnetic stirrer and a cap. CoxFe3-xO4-DMSA-Au NPs ([Au]=0.44 mM) were
13
14 added to a solution of 4-nitrophenol (1.83 mL, 0.1 mM) in different mole percentages with
15
16 respect to the substrate (44, 22 or 9 µL of NP dispersion, corresponding to 10, 5 and 2%
17
18
19 Au:substrate molar ratio, respectively). Then, an aqueous solution of NaBH4 (330 µL, 500 mM)
20
21 was added to the mixture (initial reaction time). The transformation ratio was followed
22
23 monitoring the absorption at 400 nm and 550 nm (see SI) of the reaction mixture at intervals of 1
24
25
26 minute. After each catalytic cycle, the NP catalyst was magnetically recovered by removal of the
27
28 supernatant. The collected NPs were washed with water and fresh 4-nitrophenol and NaBH4
29
30
31
solutions were introduced in the cell to start a new catalytic cycle. For the control reactions, the
32
33 same procedure described above for the 10% molar ratio case was carried out, using CoxFe3-xO4-
34
35 DMSA NPs (180 µL, [Fe]=0.6 mM) and Au-DMSA NPs (14 µL, [Au]=1.2 mM, 10% molar
36
37
38 ratio) as catalyst.
39
40
41 2.8 Electrochemical reduction of 4-nitrophenol. Electrochemical investigations were
42
43 performed using an Autolab PG-Stat 12 (Ecochemie, The Netherlands) potentiostat/galvanostat,
44
45
46 in a conventional three electrodes cell with a saturated calomel reference electrode and a
47
48 platinum wire counter electrode. The working electrode was a glassy carbon one (AMEL, Italy, 3
49
50
mm diameter), modified by drop casting with a Kartell micropipette 20 µL of a solution of the
51
52
53 nanomaterials: CoxFe3-xO4-DMSA-Au NPs ([Au]=0.44 mM), CoxFe3-xO4-DMSA NPs ([Fe]=0.6
54
55 mM), Au-DMSA NPs ([Au]=0.4 mM). Current densities were determined dividing the
56
57
58
experimental currents by the electrode geometric area. Prior to modification, the GC support
59
60
ACS Paragon Plus Environment 8
Page 9 of 41 Langmuir

1
2
3
electrode surface was cleaned with synthetic diamond powder (Aldrich, diameter 1 µm, 99.9%)
4
5
6 on a Struers DP Nap wet cloth and rinsed in water. 0.1 M NaClO4 (Sigma Aldrich) was used as
7
8 supporting electrolyte in aqueous media (20 mL) and mixed with a saturated solution of 4-
9
10
11
nitrophenol (60 µL).
12
13
14 3. RESULTS AND DISCUSSION
15
16
17 3.1. Synthesis of water dispersible cobalt ferrite. To produce hetero-NPs we tuned a multi-
18
19
20 step protocol (Scheme 1).
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47 Scheme 1. Synthetic strategy for CoxFe3-xO4-DMSA-Au NPs.
48
49
50 The initial step consisted in the synthesis of organic dispersible cobalt ferrite (CoxFe3-xO4-OA
51
52
53 NPs), fabricated according to the procedure published by Zhu et al.31 by means of the
54
55 solvothermal decomposition of iron(III) and cobalt(II) acetylacetonate precursors. This
56
57
58
procedure allowed us to have an efficient control over the morphology of superparamagnetic NP
59
60
ACS Paragon Plus Environment 9
Langmuir Page 10 of 41

1
2
3
synthesis. We slightly modified the literature protocol by employing 1,2-hexadecanediol as a
4
5
6 mild reducing agent, that guarantees the monocrystallinity of cobalt ferrite nanocrystals.34 The
7
8 obtained NPs had spheroidal shape (median equivalent diameter = 7.8 ± 1.2 nm, Figure 1a and
9
10
11
Figure S1) and both the electron diffraction pattern and the HRTEM fringes (Figure S2)
12
13 correspond to a cubic ferrite.
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32 Figure 1. TEM images (scale bar = 50 nm) of a) CoxFe3-xO4-OA NPs, b) CoxFe3-xO4-DMSA NPs and c) CoxFe3-
33 xO4-DMSA-Au NPs with an expansion in the inset (bar 10 nm).
34
35
36
37
38 The chemical composition of the NPs has been investigated by recording EELS spectra
39
40
41 (Figure 2a) on a grid area of about 0.36 µm2. In the spectra the Fe-L2,3 edge (beginning at 708
42
43 eV) and the Co-L2,3 edge (beginning at 779 eV) can be observed. To calculate the molar ratio
44
45
46
between Fe and Co we fitted the spectra using the best-fit procedure implemented in the EELS
47
48 Model 3.3 software,35 obtaining an atomic Fe:Co ratio equal to 2.7±0.9 (x=0.8±0.2). By the ESI
49
50 maps (Figure 2c,d), which provide an image of the distribution of a specific element, it is evident
51
52
53 that both metal ions are present in each NPs. Thus, we established that the two metals were
54
55 homogeneously distributed in the sample. This finding, along with the diffraction data, showed
56
57 that we truly obtained cobalt ferrite NPs and not a mixture of iron oxide NPs and cobalt oxide
58
59
60
ACS Paragon Plus Environment 10
Page 11 of 41 Langmuir

1
2
3
NPs. Moreover, using the independent spectroscopic technique EDX (figure S3), we measured
4
5
6 the atomic percentage of Fe = 69.8% and Co = 30.2% that correspond to an atomic Fe:Co ratio
7
8 of 2.31 (x = 0.91). This experiment was performed on a grid area of about 2.5 µm2, larger than
9
10
11 that sampled in the EELS experiment. The Fe:Co ratio obtained by EELS and EDX are in a good
12
13 agreement and show the quasi stoichiometric composition of the cobalt ferrite NPs with x ≅ 0.9.
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39 Figure 2. a) EELS spectrum of CoxFe3-xO4-OA NPs where the two edges starting at 708eV and 779eV, correspond
40 to the FeL2,3 and CoL2,3 edges, respectively; b) Reference TEM image to be compared with the ESI maps of iron (c,
41 red) and cobalt (d, green). The bars correspond to 50 nm.
42
43
44
45
46
47
The second step was focused on the exchange of surfactants, to remove the long-chain
48
49 hydrophobic ligands (oleic acid, oleylamine) which coated the metal oxides NPs and to bring
50
51 them in water media. To accomplish the exchange, we chose to use 2,3-meso-dimercaptosuccinic
52
53
54 acid (DMSA), a ligand well known in literature for its ability to transfer ferrite NPs from organic
55
56 to aqueous media.36 The literature procedure was improved by adding a second exchange step.
57
58
59
60
ACS Paragon Plus Environment 11
Langmuir Page 12 of 41

1
2
3
First, a solution of DMSA in DMSO was dropped into a toluene dispersion of CoxFe3-xO4-OA
4
5
6 NPs. The substitution of OA with DMSA lets NPs to precipitate in toluene and then to be
7
8 dispersed in water. During the second step, the aqueous NP dispersion was treated with an excess
9
10
11
of DMSA, in order to improve its colloidal stability in water (Figure 1b and Figure S4-6). NPs
12
13 treated with an excess of DMSA resulted in a decreased aggregation on the TEM grid (Figure
14
15 S5), confirmed by the size distributions obtained by DLS (Figure S7). Figure S7, in fact, show a
16
17
18 decreased hydrodynamic size and distribution from the first to the second step of the ligand
19
20 exchange procedure, underlining the improvement in stability and dispersion. To evaluate the
21
22 colloidal stability before and after the treatment with the excess of DMSA, we measured the
23
24
25 variation of ζ-potential of NP solution, as a function of pH and ionic strength of the medium.37-39
26
27 At a fixed ionic strength ([KCl]=1 mM), the pH of the NP dispersion was tuned in a range
28
29
between 2 and 12 by adding aq. NaOH or aq. HCl (0.5 M) (Figure S8a), registering that the NP
30
31
32 dispersion, obtained after the second addition of DMSA, had a slight increment of negative ζ-
33
34 potential and thus an improved colloidal stability at pH > 3.5.37,39 At pH < 3.5, CoxFe3-xO4-
35
36
37 DMSA NPs treated with the excess of DMSA displayed a rapid increase of ζ-potential that could
38
39 be ascribed to the protonation of free carboxylic groups of DMSA (pK1 2.71, pK2 3.48). The
40
41 influence of the ionic strength on the CoxFe3-xO4-DMSA NPs colloidal stability was evaluated
42
43
44 increasing the concentration of KCl in a phosphate-buffered NP dispersion (PBS, pH=7). The
45
46 data (Figure S8b) showed lower ζ-potential values for higher ionic strengths,38 as expected for
47
48 charged NPs, and confirmed the improved stability of NPs obtained after the treatment with the
49
50
51 excess of DMSA in a large range of ionic force. FT-IR spectra (Figure S9) are in agreement with
52
53 the literature data,39 confirming the success of the ligand exchange procedure. By means of ICP-
54
55
56
OES measurements (Figure S10), the atomic Fe:Co ratio in CoxFe3-xO4-DMSA NPs was
57
58
59
60
ACS Paragon Plus Environment 12
Page 13 of 41 Langmuir

1
2
3
established to be 71:29 (x = 0.87), confirming the result obtaining by EELS and EDX
4
5
6 experiments on CoxFe3-xO4-OA NPs. Therefore, the “two-steps” ligand exchange procedure did
7
8 not alter the chemical composition of the ferrite NPs.
9
10
11
12 3.2. Synthesis of gold decorated cobalt ferrite. The obtained water dispersible CoxFe3-xO4-
13
14 DMSA NPs were used as a substratum for the localized growth of ultrasmall gold NPs,
15
16 exploiting the known reducing ability of DMSA for the production of nanometer sized gold
17
18
19 nanoparticles.32 Recently, Carlà et al.26 described the production of core-shell CoFe2O4/Au
20
21 composites, by using DMSA as stabilizing agent for CoFe2O4 NPs and a sonochemical process
22
23 to achieve the reduction of gold salts. In another paper, Odio et al.40 took advantages of the
24
25
26 ability of magnetite-DMSA NPs to adsorb gold ions from aqueous environments, recording the
27
28 formation of reduced gold. Drawing inspiration from literature, we exploited the double role that
29
30
31
DMSA can play to produce controlled and water dispersible cobalt ferrite/gold hetero-structured
32
33 NPs (Scheme 1), by slowly dropping HAuCl4 in a CoxFe3-xO4-DMSA NP solution. In order to
34
35 optimize the decoration step, we evaluated the influence of two main parameters: the Au3+
36
37
38 amount and the rate of its addition. A solution of HAuCl4 (0.2 mM) was dropped (83 µL/min) in
39
40 5 mL of CoxFe3-xO4-DMSA NP water dispersion ([Fe] = 34 µg/mL), supplying three different
41
42 quantities (10-3 mmol, 2·10-3 mmol, 4·10-3 mmol). The addition of 10-3 mmol of Au3+ resulted in
43
44
45 a low decoration level (1.8±1.2 Au NPs/ CoxFe3-xO4-DMSA NPs) with Au NPs having
46
47 equivalent diameter < 3 nm (Figure S11a,d). The injection of 4·10-3 mmol of Au precursor led to
48
49
a good decoration level (3.1±1.7 Au NPs/ CoxFe3-xO4-DMSA NPs) but also to the formation of
50
51
52 undesired large gold NPs having equivalent diameter >20 nm (Figure S11c,f). The best
53
54 compromise was achieved using 2·10-3 mmol of HAuCl4 resulted in 2.4±1.3 Au NPs/ CoxFe3-
55
56
57 xO4-DMSA NPs with Au NPs having equivalent diameter < 3 nm (Figure S10b,e). To evaluate
58
59
60
ACS Paragon Plus Environment 13
Langmuir Page 14 of 41

1
2
3
the role of the dropping rate in the decoration step, the optimized quantity of Au3+ (2·10-3 mmol)
4
5
6 was slower added (flow rate = 41.5 µL/min) or rapid injected. By decreasing the Au3+ addition
7
8 rate, the instantaneous DMSA/Au ratio in the mixture incremented, promoting the formation of
9
10
11 small Au NPs (median diameter 2.7 ± 0.5 nm) with respect to the growth of the larger ones,41
12
13 achieving a decoration of 2.9±1.7 Au NPs per CoxFe3-xO4-DMSA NP (Figure 4, S11g and S13-
14
15
14). While, the sample obtained by the single injection present a low decoration degree (1.4±1.7
16
17
18 Au NPs per CoxFe3-xO4-DMSA NP) and the presence of larger Au NPs with dmedian=28 ± 8 nm
19
20 (Figure 4c and Figure S12).41 Moreover, comparing UV-vis spectra (Figure 3a), the surface
21
22
23
plasmonic resonance (SPR) peak at 550 nm, correlated to the presence of large Au NPs, is
24
25 evident only in the sample obtained by single injection.20,42 SPR peak is barely visible in the
26
27 CoxFe3-xO4-DMSA-Au NP solution obtained by the dropped addiction, confirming the scarce
28
29
30 presence of large Au NPs.
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46 Figure 3. a) UV-Vis spectra of the CoxFe3-xO4-DMSA NPs and CoxFe3-xO4-DMSA-Au NPs obtained by a single
47
injection or by a slow dropping (41.5 µL/min); b) STEM Image of CoxFe3-xO4-DMSA-Au NPs obtained by slow
48 dropping; c) STEM Image of CoxFe3-xO4-DMSA-Au NPs obtained by single injection. The scale bars correspond to
49 20 nm.
50
51
52
53
TEM and STEM images confirmed the information obtained from UV-vis spectra (Figure 1c
54
55
56 and Figure 3b,c). STEM images show the high degree and uniformity of decoration that can be
57
58
59
60
ACS Paragon Plus Environment 14
Page 15 of 41 Langmuir

1
2
3
obtained following the optimized synthetic strategy (Figure 3b and Figure S15). As expected, we
4
5
6 identified in the electron diffraction of the CoxFe3-xO4-DMSA-Au NPs (Figure S16) the two
7
8 patterns corresponding to Au and CoxFe3-xO4. The atomic ratio between Fe, Co and Au was
9
10
11
quantified by ICP-OES measurement as Fe:Co:Au = 72:13:14. After gold decoration, a variation
12
13 in the ratio between Fe and Co was detected, corresponding to a decrease in cobalt content26 (x ≅
14
15
0.5) (Figure S17). This result is in agreement with the elemental composition measured using
16
17
18 EDX spectroscopy (Figure S18). In summary, the best synthetic outcome – extensive decoration
19
20 and minimization of the presence of large gold NPs – is achieved by slow injection (flow rate =
21
22
23
41.5 µL/min) of 0.33 eq of gold precursor, calculated with respect to the iron content of the
24
25 ferrite NPs.
26
27 3.3. Magnetic properties of cobalt ferrite NPs.
28
29
30 We investigated the magnetic properties of the cobalt ferrite nanocrystals before and after the
31
32 gold decoration in order to study the effects of the decrease of the cobalt content (Figure 4 and
33
34 Figure S19-S20). The main magnetic parameters are reported in Table 1.
35
36
37
38
39 Table 1. Magnetic parameters of CoxFe3-xO4-DMSA and CoxFe3-xO4-DMSA-Au NPs
40 Sample Hc Ms Mr Ms/Mr Tmax Tder
41
42
(kOe) (emu/g) (emu/g) (K) (K)
43 CoxFe3-xO4 DMSA 14.9 142 113 0.79 240 200
44 CoxFe3-xO4 DMSA-Au 14.5 107 84 0.78 270 230
45
46
47
48
The isothermal magnetization M(H) of both samples presents a broad hysteresis loop with
49
50 large coercive field typical of magnetically hard cobalt ferrite (Figure 4). The coercive field Hc of
51
52 both samples is typical of cobalt ferrite NPs and smaller than the bulk value of 25.2 kOe.43 Hc is
53
54
55 only slightly smaller after gold decoration despite the substantial Co loss (x decreases from 0.87
56
57 to 0.46). It was however recently observed44 that in sub-stoichiometric CoxFe3-xO4 NPs the
58
59
60
ACS Paragon Plus Environment 15
Langmuir Page 16 of 41

1
2
3
magnetic properties quickly change when x is small and level off about x = 0.5, in accordance
4
5
6 with the present observation. The recorded hysteresis loops are almost, but not exactly, major
7
8 loops and the saturation magnetization Ms values might therefore be slightly underestimated. MS
9
10
11
decreases from 142 to 107 emu/g upon gold decoration. Both values are larger than the MS bulk
12
13 value, (93 emu/g at 5K).43
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Figure 4. Hysteresis loops of the cobalt ferrite samples before and after gold decoration. Hollow circles: CoxFe3-
35 xO4-DMSA; solid circles: CoxFe3-xO4-DMSA-Au.The magnetization is given relative to the mass of cobalt ferrite.
36
37
38
39
40 Such large values are difficult to interpret since electron diffraction excluded the presence of
41
42 non-ferritic phases with larger magnetization. A similar effect has been observed in small (3 nm)
43
44 stoichiometric cobalt ferrite NPs diluted in a silica matrix and ascribed to frozen canted spins
45
46
47 located in a 0.3 nm thick surface shell of the ferrite nanocrystal.45 However, it is uncertain if this
48
49 conclusion can also be drawn for our 7.8 nm sub-stoichiometric undiluted cobalt ferrite
50
51
nanocrystals. The squareness ratio Mr/Ms is 0.79 and 0.78 for CoxFe3-xO4 DMSA and CoxFe3-xO4
52
53
54 DMSA-Au respectively, close to the value predicted for non-interacting randomly-oriented
55
56 Stoner-Wohlfahrt nanocrystals having cubic magnetocrystalline anisotropy.46,47 Such high values
57
58
59
60
ACS Paragon Plus Environment 16
Page 17 of 41 Langmuir

1
2
3
have already been observed in nanocomposites containing cobalt-ferrite NPs in a polyaniline
4
5
6 matrix.48 We also measured the zero field cooling (ZFC) and field cooling (FC) magnetization
7
8 curves of both samples as a function of the temperature (from 5 to 300 K) using a measuring
9
10
11
field of 10 Oe (Figure S19-S20). These curves are typical of magnetic nanoparticles and clearly
12
13 display the transition to the superparamagnetic regime at high temperature. In general, the
14
15 behavior of both samples agrees with the prediction that cobalt ferrite nanoparticles having
16
17
18 diameter < 10 nm are superparamagnetic at 300 K.49 Several proposals have been made to
19
20 estimate the average blocking temperature from ZFC/FC curves. A popular choice is the
21
22 temperature Tmax where MZFC reaches its maximum value. We observed Tmax = 240 and 270 K in
23
24
25 the case of CoxFe3-xO4-DMSA and CoxFe3-xO4-DMSA-Au NPs, respectively. The latter value is
26
27 in agreement with a literature value for ferrite NPs with similar cobalt content.44 A deeper insight
28
29
into the blocking/unblocking behavior of the NPs can be achieved by computing the derivative
30
31
32 −d(MFC −MZFC)/dT, which is an approximate representation of the distribution of the magnetic
33
34 anisotropy energy barriers (Figure S19-S20).50 The most probable barrier is at Tder = 200 and 230
35
36
37 K for CoxFe3-xO4-DMSA and CoxFe3-xO4-DMSA-Au NPs, respectively. Both estimates thus
38
39 indicate that the blocking temperature increases by about 30 K upon loss of cobalt caused by the
40
41 gold decoration. This is unexpected since (i) the cobalt ferrite size does not significantly change
42
43
44 upon gold decoration and (ii) cobalt ions largely add to the global magnetic anisotropy due to the
45
46 their strong single-ion anisotropy. At the moment we can only suggest that decoration with gold
47
48 NPs modifies the interparticle interaction (e.g. the dipole-dipole interaction) in such a way that
49
50
51 the cobalt ferrite nanocrystals are subject to a larger global magnetic anisotropy translating into a
52
53 higher blocking temperature.
54
55
56
57
58
59
60
ACS Paragon Plus Environment 17
Langmuir Page 18 of 41

1
2
3
3.4. Catalytic application of gold decorated cobalt ferrite. Taking advantage of the
4
5
6 magnetic properties of the nanometric ferrite support and the catalytic potential of Au NPs, the
7
8 produced CoxFe3-xO4-DMSA-Au NPs could be used as quasi-homogeneous and magnetically
9
10
11
recyclable catalyst for red-ox reactions.19,21,22 In the present study the catalytic activity of
12
13 optimally-decorated CoxFe3-xO4-DMSA-Au NPs has been investigated performing the reduction
14
15 of 4-NP to 4-AP, monitoring the reaction by UV-vis spectroscopy and using NaBH4 as reducing
16
17
18 agent (Figure 5).29 In the presence of an excess of NaBH4, 4-NP is converted into its sodium salt
19
20 (4-nitrophenolate) which shows an intense absorption peak at 400 nm. Through monitoring the
21
22 decreasing of the absorbance at 400 nm during the time, it is possible to follow the kinetics of the
23
24
25 reaction (Figure 5).
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50 Figure 5. UV-vis spectra of 4-nitrophenol, sodium 4-nitrophenolate and sodium 4-aminophenolate
51
52
53 To evaluate the catalytic efficiency (Figure 6a), three different CoxFe3-xO4-DMSA-Au NPs
54
55 amounts have been tested (2, 5 and 10 % molar with respect to Au, as measured by ICP-OES).
56
57
58 Furthermore, as negative control, the red-ox reactions were performed without addition of
59
60
ACS Paragon Plus Environment 18
Page 19 of 41 Langmuir

1
2
3
catalyst or by adding the colloidal water dispersion of cobalt ferrite (CoxFe3-xO4-DMSA NPs).
4
5
6 Either without catalyst or with CoxFe3-xO4-DMSA NPs, no reaction occurred as expected,51
7
8 indicating that the presence of the noble metal is required for catalysis. Instead, in the presence
9
10
11
of CoxFe3-xO4-DMSA-Au NPs, the reaction was completed in a few minutes after an induction
12
13 time dependent on the amount of the catalyst. According to the Langmuir–Hinshelwood (LH)29,52
14
15 model (Figure S21) which describes the kinetics of this reaction, both reactants (4-NP and
16
17
18 hydride) need to be adsorbed on the surface of the NPs to react. The observed induction period is
19
20 then related to the slow restructuring of NP surfaces by the reactants,52,53 before the reaction
21
22 starts. To confirm this hypothesis we performed two catalyzed reactions pre-incubating NPs with
23
24
25 NaBH4 or with 4-NP: in both cases we still observed an induction time even if shorter (2-4 min
26
27 vs 6-8 min, Figure S21-S22). The catalytic activity of the hetero-nanoparticles is excellent
28
29
(Figure 6b and Table S24). For instance, when the catalyst is used at 2% mol (Au), the reaction
30
31
32 proceeds with one of the highest rate reported in literature (knorm = 4.4 ·106 s-1mol-1) by using
33
34 nanosized catalyst.19 Finally, to compare our catalytic system with a non-magnetically supported
35
36
37 one, we synthesized Au-DMSA NPs with a median diameter of 2.1±0.4 nm following a literature
38
39 procedure (Figure S25a-S26).32 Using 10% molar of Au-DMSA NPs with respect to 4-NP, the
40
41 initial activation time was shorter but the reaction rate was drastically slower. In fact, kobs for Au-
42
43
44 DMSA NPs was 6 times lower than that observed for the corresponding amount of CoxFe3-xO4-
45
46 DMSA-Au NPs (Figure 6b and Table S24).
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment 19
Langmuir Page 20 of 41

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22 Figure 6. a) The reduction efficiency of 4-NP by: ■ No catalyst; ● CoxFe3-xO4-DMSA NPs; ∆ 10% Au-DMSA NPs;
23 ▲ 10% CoxFe3-xO4-DMSA-Au NPs; □ 5% CoxFe3-xO4-DMSA-Au NPs; ○ 2% CoxFe3-xO4-DMSA-Au NPs.
24 Remaining Ratio is defined as the residual percentage of 4-nitrophenolate in the reaction solution, calculated from
25 the absorbance values as explained in SI. b) The observed kinetic constant kobs for 10% Au-DMSA NPs and 2%,
26 5% and 10% CoxFe3-xO4-DMSA-Au NPs.
27
28 The higher catalytic activity of CoxFe3-xO4-DMSA-Au NPs respect to Au-DMSA NPs can be
29
30
31
ascribed to the unique electronic features of hetero-nanoparticles. Even if the electronic behavior
32
33 at the metallic interface is still unclear, in literature it is hypothesized that an electron flow from
34
35 gold to ferrite NPs drives the remarkable improvement in catalytic activity of these kinds of
36
37
38 hetero-nanoparticles.6,54,55 To deeper investigate their catalytic behavior, we carried out an
39
40 electrochemical reduction of 4-NP. An electrocatalytic effect (100 mV peak shift towards less
41
42 negative potentials) is observed for all the modified electrodes (Figure 7), testifying the catalytic
43
44
45 properties of the investigated materials in the 4-NP reduction reaction. Moreover, in the case of
46
47 CoxFe3-xO4 modified electrodes (with or without Au NPs) an increase in peak current densities
48
49
50
(i.e. electrochemical reaction rate) is evident, ascribed to an increase of the surface area and thus
51
52 of the available reactive sites of these materials. This result is in agreement with the previous
53
54 experiment where kobs calculated for Au-DMSA NPs resulted lower than that observed for
55
56
57 CoxFe3-xO4-DMSA-Au NPs (Figure 6).
58
59
60
ACS Paragon Plus Environment 20
Page 21 of 41 Langmuir

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 Figure 7. Cyclovoltammetric reduction patterns (0.1 V s-1 scan rate) of 4-NP (60 µL of a saturated aqueous solution
26 in 20 ml of 0.1 M NaClO4 aqueous solution) on the GC support and on three different modified electrodes.
27
28
29 Interestingly, from the electrochemical results, it seems that Au NPs does not provide an extra
30
31
32 contribute in the direct electrocatalytic reduction of 4-NP, since CoxFe3-xO4-DMSA NPs can
33
34 efficiently catalyze the reaction. On the contrary, in our model reaction condition which exploits
35
36
the use of NaBH4, CoxFe3-xO4-DMSA NPs are inactive while Au NPs presence is crucial to
37
38
39 promote on their surface the absorption and oxidation of sodium borohydride.56 After this first
40
41 step, the synergistic effect between Au NPs and cobalt ferrite transfer efficiently the electrons
42
43
44
required for the reduction.52,57 This aspect, together with the increased surface active areas, can
45
46 explain the better performance registered for CoxFe3-xO4-DMSA-Au NPs with respect to Au-
47
48 DMSA NPs.
49
50
51 Due to their magnetization, the hybrid NPs can be rapidly (few seconds) and easily
52
53 recovered using a neodymium-based magnet (Figure S27) and re-dispersed in MilliQ water for
54
55 further catalytic cycles. To verify the recycling efficiency of the CoxFe3-xO4-DMSA-Au NPs, we
56
57
58 performed 5 consecutive catalytic cycles (Figure 8a). In the first cycle the induction time is not
59
60
ACS Paragon Plus Environment 21
Langmuir Page 22 of 41

1
2
3
negligible, but it slowly diminishes with the repetition of the cycles, confirming the LH
4
5
6 model.29,52,53
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26 Figure 8. Reduction efficiency of 4-NP in the consecutive catalytic cycles (a) and the corresponding kobs calculated
27 for each cycle (b).
28
29
30 The observed rate constant was calculated for each recycle (Figure 8b). After the first cycle
31
32
33
(kobs = 2.7·10-2 s-1), the kobs halved and maintained the value of 1.2·10-2 s-1 during the next four
34
35 cycles. The high performance of the nano-catalyst was also proved by the fact that, even after 5
36
37 catalytic cycles, the reduction of 4-NP was complete in less than 10 minutes.
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
Figure 9. The concentration of Fe, Co and Au in: digestion blank, catalytic reaction supernatant, recovered
57
catalyst after one catalytic cycle, and fresh catalyst.
58
59
60
ACS Paragon Plus Environment 22
Page 23 of 41 Langmuir

1
2
3
After the first cycle’s completion, the amount of the metals was quantified by ICP-OES as in
4
5
6 the recovered catalyst as in the supernatant solution and it was compared with those measured in
7
8 the fresh catalyst and in the digestion blank (Figure 9). Noteworthy, no leaching was detected as
9
10
11
the metal composition and concentration remained constant in the recovered and fresh catalyst.
12
13 Moreover the metals abundance in the supernatant was comparable with that of the digestion
14
15 blank. We performed an accurate characterization of the catalyst after the first cycle using
16
17
18 electron microscopy. TEM (Figure 10 and Figure S28) and STEM (Figure S28) images show
19
20 that recovered CoxFe3-xO4-DMSA-Au NPs formed aggregates probably included in an
21
22 amorphous material, justifying the decreased reaction rate observed after the first cycle.19
23
24
25 Nevertheless, CoxFe3-xO4-DMSA-Au NPs are not sintered, maintaining their single crystal
26
27 nature. By EDX analysis (Figure S29) we calculated the atomic percentage of iron, cobalt and
28
29
gold (79.1%, 10.1% and 10.4%, respectively; x = 0.34). This result is in agreement with the
30
31
32 concentration ratio measured by ICP-OES and, qualitatively, also with the EELS spectrum
33
34 (Figure S30). Finally, ESI images (Figure S31) confirm the homogeneous distribution of iron
35
36
37 and cobalt in the ferrite NPs of the recovered catalyst.
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
Figure 10. TEM image of the recovered catalyst after one cycle. The bar corresponds to 20 nm.
57
58
59
60
ACS Paragon Plus Environment 23
Langmuir Page 24 of 41

1
2
3
4. CONCLUSIONS
4
5
6
7 In this paper we reported an effective synthetic procedure to obtain water dispersible hybrid
8
9 hetero-structured NPs, composed by a cobalt ferrite nanocore decorated by ultrasmall gold NPs.
10
11
12 The designed protocol exploits the presence of 2,3-meso-dimercapto succinic acid acting both as
13
14 coordinate ligand for water dispersible cobalt ferrite and as reducing agent for Au3+ ions. Thanks
15
16 to its versatility, the procedure can be potentially applied to any type of spinel ferrite NPs,
17
18
19 MxFe3-xO4 (M = Fe, Zn, Mn, Cu, Ni). A full characterization of each step has been provided,
20
21 highlighting the factors that play a crucial role in the decoration step. Moreover, taking
22
23 advantage of the ultrasmall size of Au NPs, we demonstrated that CoxFe3-xO4-DMSA-Au NPs act
24
25
26 as efficient catalyst for the reduction of 4-NP to 4-AP showing a full conversion down to 2%
27
28 molar of catalyst with an excellent knorm = 4.4·106s-1mol-1. Finally, by exploiting the
29
30
31
magnetization of cobalt ferrite NPs, the catalyst was magnetically recovered and recycled up to 5
32
33 times, maintaining a very high kobs over the cycles.
34
35
36
37
38
39
40 ASSOCIATED CONTENT
41
42
43 Supporting Information. Synthesis of CoxFe3-xO4-OA NPs. Figure S1: TEM and statistical
44
45 distribution. Figure S2: HRTEM and diffraction pattern. Figure S3: EDX Spectrum. Ligand
46
47
48 exchange procedure (CoxFe3-xO4-DMSA NPs). Figure S4: picture of NPs before and after ligand
49
50 exchange. Figure S5: TEM images of CoxFe3-xO4-OA NPs and CoxFe3-xO4-DMSA NPs. Figure
51
52
53
S6: statistical distribution. Figure S7: DLS size distribution of CoxFe3-xO4-OA NPs and CoxFe3-
54
55 xO4-DMSA NPs. Figure S8: variation of the ζ-potential in function of pH and ionic strength.
56
57 Figure S9: FTIR spectra. Figure S10: ICP-OES analysis. Synthesis of CoxFe3-xO4-DMSA-Au
58
59
60
ACS Paragon Plus Environment 24
Page 25 of 41 Langmuir

1
2
3
NPs. Figure S11: optimization process. Figure S12: STEM image of CoxFe3-xO4-DMSA-Au NPs
4
5
6 obtained by single injection. Figure S13: statistical size distribution of Au NPs. Figure S14:
7
8 multimodal size distribution and ζ-potential value. Figure S15: STEM images of CoxFe3-xO4-
9
10
11
DMSA-Au NPs. Figure S16: diffraction pattern. Figure S17: ICP-OES. Figure S18: EDX
12
13 Spectrum. Magnetic properties of cobalt ferrite NPs. Figure S19: ZFC and FC magnetization of
14
15 CoxFe3-xO4-DMSA NPs. Figure S20: ZFC and FC magnetization of CoxFe3-xO4-DMSA-Au NPs.
16
17
18 Catalytic activity of CoxFe3-xO4-DMSA-Au NPs. Figure S21. Schematic representation of LH
19
20 model. Figure S22: reduction efficiency after incubation with NaBH4. Figure S23: reduction
21
22 efficiency after incubation with 4-NP. Table S24: kobs and knorm. Figure S25: TEM images of Au-
23
24
25 DMSA NPs before and after catalytic reaction. Figure S26: UV-vis spectrum of Au-DMSA NPs.
26
27 Figure S27: picture of the recovering of catalyst. Figure S28: TEM and STEM images of CoxFe3-
28
29
xO4-DMSA-Au NPs after 1st cycle. Figure S29: EDX of CoxFe3-xO4-DMSA-Au NPs after 1st
30
31
32 cycle. Figure S30: EELS of CoxFe3-xO4-DMSA-Au NPs after 1st cycle. Figure S31: TEM and
33
34 ESI of CoxFe3-xO4-DMSA-Au NPs after 1st cycle. This material is available free of charge via the
35
36
37 Internet at http://pubs.acs.org.
38
39
40 AUTHOR INFORMATION
41
42
43 Corresponding Author
44
45
*E-mails: laura.polito@istm.cnr.it; anna.ferretti@istm.cnr.it
46
47
48
49 Notes
50
51
52
The authors declare no competing financial interest.
53
54 Author Contributions
55
56
57
58
59
60
ACS Paragon Plus Environment 25
Langmuir Page 26 of 41

1
2
3
The manuscript was written through contributions of all authors. All authors have given approval
4
5
6 to the final version of the manuscript.
7
8
9
10
11
12 ACKNOWLEDGMENT
13
14 This work has been financially supported by Regione Lombardia (Italy) through the project
15
16 RSPPTECH. LP thanks MIUR (Italy) for (PRIN 2010-2011: contract 2010JMAZML_003).
17
18
19
20
21 REFERENCES
22
23 (1) Song, H. Metal hybrid nanoparticles for catalytic organic and photochemical
24
25 transformations. Acc. Chem. Res. 2015, 48, 491−499.
26 (2) Sankar, M.; Dimitratos, N.; Miedziak, P.J.; Wells, P.P.; Kielye, C.J.;Hutchings, G.J.
27 Designing bimetallic catalysts for a green and sustainable future. Chem. Soc. Rev. 2012, 41,
28 8099–8139.
29 (3) Wu, X.; Gao, Y.; Dong, C.-M. Polymer/gold hybrid nanoparticles: from synthesis to
30
31
cancer theranostic applications. RSC Adv. 2015, 5, 13787-13795.
32 (4) Banin, U.; Ben-Shahar, Y.; Vinokurov, K. Hybrid semiconductor−metal nanoparticles:
33 from architecture to function. Chem. Mater. 2014, 26, 97−110.
34 (5) Wang, D. Li, Y. Bimetallic nanocrystals: liquid-phase synthesis and catalytic
35 applications. Adv. Mater. 2011, 23, 1044–1060.
36 (6) Carbone, L.; Cozzoli, P. D. Colloidal heterostructured nanocrystals: Synthesis and
37
38 growth mechanisms. Nano Today 2010, 5, 449–493.
39 (7) Leung, K.C.-F.; Xuan, S.; Zhu, X.; Wang, D.; Chak, C.-P.; Lee, S.-F.; Hob, W.K.-W.;
40 Chung B.C.-T. Gold and iron oxide hybrid nanocomposite materials. Chem. Soc. Rev. 2012, 41,
41 1911–1928.
42 (8) Goon, I.Y.; Lai, L.M.H.; Lim, M.; Munroe, P.; Gooding, J.J. Amal, R. Fabrication and
43
44
dispersion of gold-shell-protected magnetite nanoparticles: systematic control using
45 polyethyleneimine. Chem. Mater. 2009, 21, 673–681.
46 (9) Park, H.-Y.; Schadt, M.J.; Wang, L.; Lim, I-I. S.; Njoki, P.N.; Kim, S. H.; Jang, M.-Y.;
47 Luo, J.; Zhong, C.-J. Fabrication of magnetic core@shell Fe oxide@Au nanoparticles for
48 interfacial bioactivity and bio-separation. Langmuir 2007, 23, 9050-9056.
49
(10) Lyon, J.L.; Fleming, D.A.; Stone, M.B.; Schiffer, P.; Williams, M.E. Synthesis of Fe
50
51 oxide core/Au shell nanoparticles by iterative hydroxylamine seeding. Nano Lett. 2004, 4, 719-
52 723.
53 (11) Tintoré, M.; Mazzini, S.; Polito, L.; Marelli, M.; Latorre, A.; Somoza, Á.; Aviñó, A.;
54 Fàbrega, C.; Eritja, R. Gold-coated superparamagnetic nanoparticles for single methyl
55 discrimination in DNA aptamers. Int. J. Mol. Sci. 2015, 16, 27625–27639.
56
57
58
59
60
ACS Paragon Plus Environment 26
Page 27 of 41 Langmuir

1
2
3
(12) Pineider, F.; de Juliàn Fernàndez, C.; Videtta, V.; Carlino, E.; al Hourani, A.; Wilhelm,
4
5 F.; Rogalev, A.; Cozzoli, P. D.; Ghigna, P.; Sangregorio, C. Spin-Polarization transfer in
6 colloidal magnetic-plasmonic Au/iron oxide hetero-nanocrystals. ACS Nano 2013, 7, 857–866.
7 (13) Wang, C.; Yin, H.; Dai, S.; Sun, S. A general approach to noble metal-metal oxide
8 dumbbell nanoparticles and their catalytic application for CO oxidation. Chem. Mater. 2010, 22,
9 3277–3282.
10
11
(14) Xu, C.; Xie, J.; Ho, D.; Wang, C.; Kohler, N.; Walsh, E.G.; Morgan, J.R.; Chin, Y.E.;
12 Sun, S. Au–Fe3O4 dumbbell nanoparticles as dual-functional probes. Angew. Chem. Int. Ed.
13 2008, 47, 173 –176.
14 (15) Yu, H.; Chen, M.; Rice, P.M.; Wang, S.X.; White, R.L.; Sun, S. Dumbbell-like
15 bifunctional Au-Fe3O4 nanoparticles. Nano Lett. 2005, 5, 379-382.
16
(16) Fratila, R.M.; Mitchell, S.G.; del Pino, P.; Grazu, V.; de la Fuente, J.M. Strategies for
17
18 the biofunctionalization of gold and iron oxide nanoparticles. Langmuir 2014, 30, 15057−15071.
19 (17) Bagga, K.; Brougham, D.F.; Keyesde, T.E.; Brabazon, D. Magnetic and noble metal
20 nanocomposites for peparation and optical detection of biological species. Phys. Chem. Chem.
21 Phys.2015, 17, 27968-27980.
22 (18) Wang, D.; Astruc, D. Fast-growing field of magnetically recyclable nanocatalysts.
23
24
Chem. Rev. 2014, 114, 6949−6985.
25 (19) Walker, J.M.; Zaleski, J.M. A simple route to diverse noble metal-decorated iron oxide
26 nanoparticles for catalysis. Nanoscale 2016, 8, 1535–1544.
27 (20) Yang, X.; Yang, M.; Pang, B.; Vara, M.; Xia, Y. Gold Nanomaterials at Work in
28 Biomedicine. Chem. Rev. 2015, 115, 10410−10488.
29
(21) Taketoshi, A.; Haruta, M. Size- and structure-specificity in catalysis by Gold Clusters.
30
31 Chem. Lett. 2014, 43, 380–387.
32 (22) Haruta, M.; Kobayashi, T.; Sano, H.; Yamada, N. Novel gold catalysts for the oxidation
33 of carbon monoxide at a temperature far below 0° C. Chem. Lett. 1987, 405–408.
34 (23) Limaye, M.V.; Singh, S.B.; Date, S.K.; Kothari, D.; Reddy, V.R.; Gupta, A.; Sathe, V.;
35 Choudhary, R.J.; Kulkarni, S.K. High coercivity of oleic acid capped CoFe2O4 nanoparticles at
36
37 room temperature. J. Phys. Chem. B 2009, 113, 9070–9076.
38 (24) Munjal, S.; Khare, N.; Nehate, C.; Koul, V. Water dispersible CoFe2O4 nanoparticles
39 with improved colloidal stability for biomedical applications. J. Magn. Magn. Mater. 2016, 404,
40 166–169.
41 (25) Mikalauskaitė, A.; Kondrotas, R.; Niaura, G.; Jagminas, A.Gold-Coated Cobalt Ferrite
42
43
Nanoparticles via Methionine-Induced Reduction. J. Phys. Chem. C 2015, 119, 17398−17407.
44 (26) Carlà, F.; G.; Campo, Sangregorio, C.; Caneschi, A.; de Juliàn Fernàndez, C.; Cabrera,
45 L.I. Electrochemical characterization of core@shell CoFe2O4/Au composite. J. Nanopart Res.
46 2013, 15, 1813-1829.
47 (27) Pita, M.; Abad, J.M.; Vaz-Dominguez, C.; Briones, C.; Mateo-Martí, E.; Martín-Gago,
48 J.A.; del Puerto Morales, M.; Fernández, V.M. Synthesis of cobalt ferrite core/metallic shell
49
50 nanoparticles for the development of a specific PNA/DNA biosensor. J. Colloid Interface Sci.
51 2008, 321, 484–492.
52 (28) Gallo, J.; Garcia, I.; Padro, D.; Arnaiz, B.; Penades, S. Water-soluble magnetic
53 glyconanoparticles based on metal-doped ferrites coated with gold: Synthesis and
54 characterization. J. Mater. Chem. 2010, 20, 10010–10020.
55
56
57
58
59
60
ACS Paragon Plus Environment 27
Langmuir Page 28 of 41

1
2
3
(29) Zhao, P.; Feng, X.; Huang, D.; Yang, G.; Astrucc, D. Basic concepts and recent
4
5 advances in nitrophenol reduction by gold- and other transition metal nanoparticles. Coord.
6 Chem. Rev. 2015, 287, 114–136.
7 (30) Hu, H.; Xin, J.H.; Hu, H.; Wang, X.; Miao, D.; Liu, Y. Synthesis and stabilization of
8 metal nanocatalysts for reduction reactions – a review. J. Mater. Chem. A 2015, 3, 11157-11182.
9 (31) Zhu, H.; Zhang, S.; Huang, Y.-X.; Wu, L.; Sun, S. Monodisperse MxFe3–xO4 (M = Fe,
10
11
Cu, Co, Mn) nanoparticles and their electrocatalysis for oxygen reduction reaction. Nano Letters
12 2013, 13, 2947-2951.
13 (32) Negishi, Y.; Tsukuda, T. One-pot preparation of subnanometer-sized gold clusters via
14 reduction and stabilization by meso-2,3-dimercaptosuccinic acid. J.Am.Chem.Soc. 2003, 125,
15 4046-4047.
16
(33) Mondini, S.; Ferretti, A.M.; Puglisi, A.; Ponti, A. PEBBLES and PEBBLEJUGGLER:
17
18 software for accurate, unbiased, and fast measurement and analysis of nanoparticle morphology
19 from transmission electron microscopy (TEM) micrographs. Nanoscale 2012, 4, 5356-5372.
20 Software freely available from the http://pebbles.istm.cnr.it.
21 (34) Moya, C.; del Puerto Morales, M.; Batlle, X.; Labarta, A. Tuning the magnetic
22 properties of Co-ferrite nanoparticles through the 1,2 hexadecanediol concentration in the
23
24
reaction mixture. Phys. Chem. Chem. Phys. 2015,17, 13143-13149.
25 (35) Verbeeck, J.; Van Aert, S. Model based quantification of EELS spectra.
26 Ultramicroscopy 2004, 101, 207–224.
27 (36) Miguel-Sancho, N.; Bomatí-Miguel, O.; Colom, G.; Salvador, J.-P.; Marco, M.-P.;
28 Santamaría, J. Development of stable, water-dispersible, and biofunctionalizable
29
superparamagnetic iron oxide nanoparticles. Chem. Mater. 2011, 23, 2795–2802.
30
31 (37) Palma, S.I.C.J.; Marciell, M.; Carvalho, A.; Veintemillas-Verdaguer, S.; del Puerto
32 Morales, M.; Roque, A.C.A. Effects of phase transfer ligands on monodisperse iron oxide
33 magnetic nanoparticles. J. Colloid Interface Sci. 2015, 437, 147–155.
34 (38) Irigoyen, J.; Arekalyan, V.B.; Navoyan, Z.; Iturri, J.; Moya, S.E.; Donath, E. Spherical
35 polyelectrolyte brushes' constant zeta potential with varying ionic strength: an electrophoretic
36
37 study using a hairy layer approach. Soft Matter 2013, 9, 11609-11617.
38 (39) Chen, Z.P.; Zhang, Y.; Zhang, S.; Xia, J.G.; Liu, J.W.; Xu, K.; Gu, N. Preparation and
39 characterization of water-soluble monodisperse magnetic iron oxide nanoparticles via surface
40 double-exchange with DMSA. Colloid. Surface A 2008, 316, 210-216.
41 (40) Odio, O.F.; Lartundo-Rojas, L.; Santiago-Jacinto, P.; Martínez, R.; Reguera, E. Sorption
42
43
of gold by naked and thiol-capped magnetite nanoparticles: an XPS approach. J. Phys. Chem. C
44 2014, 118, 2776–2791.
45 (41) Hostetler, M.J.; Wingate, J.E.; Zhong, C.-J.; Harris, J.E.; Vachet, R.W.; Clark, M.R.;
46 Londono, J.D.; Green, S.J.; Stokes, J.J.; Wignall, G.D.; Glish, G.L.; Porter, M.D.; Evans, N.D.;
47 Murray, R.W. Alkanethiolate Gold Cluster Molecules with Core Diameters from 1.5 to 5.2 nm:
48 Core and Monolayer Properties as a Function of Core Size. Langmuir 1998, 14, 17-30.
49
50 (42) Malola, S.; Lehtovaara, L.; Enkovaara, J.; Hakkinen, H. Birth of the localized surface
51 plasmon resonance in monolayer-protected gold nanoclusters. ACS Nano 2013, 7, 10263–10270.
52 (43) Manova, E.; Kunev, B.; Paneva, D.; Mitov, I.; Petrov, L.; Estournès, C.; D’Orlèans,
53 C.; Rehspringer, J.L.; Kurmoo, M. Mechano-Synthesis, Characterization, and Magnetic
54 Properties of Nanoparticles of Cobalt Ferrite, CoFe2O4. Chem .Mat. 2004,16,5689-5696.
55
56
57
58
59
60
ACS Paragon Plus Environment 28
Page 29 of 41 Langmuir

1
2
3
(44) Li, D.; Diroll, Doan-nguyen, T.V.V.T.; Kikkawa, J.M.; Murray, C.B. Synthesis and
4
5 Size-Selective Precipitation of Monodisperse Nonstoichiometric MxFe3–xO4 (M=Mn, Co)
6 Nanocrystals and Their DC and AC Magnetic Properties. Chem. Mater. 2016, 28, 480−489.
7 (45) Peddis, D.; Cannas, C.; Piccaluga, G.; Agostinelli, E.; Fiorani, D. Spin-glass-like
8 freezing and enhanced magnetization in ultra small CoFe2O4 nanoparticles. Nanotechnology,
9 2010,21,125705-125714.
10
11
(46) Usov, N.A.; Peschany, S.E. Theoretical hysteresis loops for single-domain particles
12 with cubic anisotropy. J.Magn.Magn.Mater. 1997, 174, 247-260.
13 (47) Walkert, M.; Mayof, P.I.; O’Gradyt, K.; Charlest, S.W.; Chantrell, R.W. The magnetic
14 properties of single-domain particles with cubic anisotropy: I. Hysteresis loops. J. Phys:
15 Condens. Matter 1993, 5, 2779-2792.
16
(48) Della Pina, C.; Ferretti, A.M.; Ponti, A.; Falletta, E. A green approach to magnetically-
17
18 hard electrically –conducting polyaniline/CoFe2O4 nanocomposites. Composites Science and
19 technology 2015, 110,138-144.
20 (49) Coey, J.M.D. Magnetism and magnetic materials. Cambridge university press 2009.
21 (50) Del Bianco, L.; Hernando A, Fiorani D in Surface Effects in Magnetic Nanoparticles,
22 ed. Fiorani D., Springer, New York. 217-238.
23
24
(51) Singh, C.; Goyal, A.; Singhal, S. Nickel-doped cobalt ferrite nanoparticles: efficient
25 catalysts for the reduction of nitroaromatic compounds and photo-oxidative degradation of toxic
26 dyes. Nanoscale, 2014, 6, 7959-7970.
27 (52) Wunder, S.; Polzer, F.; Lu, Y.; Mei, Y.; Ballauff, M. Kinetic analysis of catalytic
28 reduction of 4-nitrophenol by metallic nanoparticles immobilized in spherical polyelectrolyte
29
brushes. J. Phys. Chem. C 2010, 114, 8814–8820.
30
31 (53) Cao, J.; Mei, S.; Ott, A.; Ballauff, M.; Lu, Y. In situ synthesis of catalytic active Au
32 nanoparticles onto gibbsite-polydopamine core-shell nanoplates. Langmuir 2015, 31, 9483–9491.
33 (54) Lin, F.; Doong, R. Hygly efficient reduction of 4-nitrophenol by heterostructured gold-
34 magnetite nanocatalysts. Appl. Catal. A-Gen. 2014, 486, 32-41.
35 (55) Lee, Y.; Garcia, M.A.; Huls, N.A.F.; Sun, S. Synthetic tuning of the catalytic properties
36
37 of Au-Fe3O4 nanoparticles. Angew. Chem. Int. Ed. 2010, 49, 1271 –1274.
38 (56) Pei, F.; Wang, Y.; Wanga, X.; He, P.; Chen, Q.; Wang, X.; Wang, H.; Yi, L.; Guo, J.
39 Performance of supported Au–Co alloy as the anode catalyst of direct borohydride-hydrogen
40 peroxide fuel cell. Int.J.Hydrogen Energ. 2010, 35, 8136-8142.
41 (57) Goyal, A.; Bansal S.; Kumar, V.; Singh, J.; Singhal, S. Mn substituted cobalt ferrites
42
43
(CoMnxFe2−xO4(x = 0.0, 0.2, 0.4, 0.6, 0.8,1.0)): As magnetically separable heterogeneous
44 nanocatalyst for the reduction of nitrophenols. Applied Surface Science 2015, 324, 877–889.
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment 29
Langmuir Page 30 of 41

1
2
3
Insert Table of Contents Graphic and Synopsis Here
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment 30
Page 31 of 41 Langmuir

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 Scheme 1. Synthetic strategy for CoxFe3-xO4-DMSA-Au NPs.
27 139x77mm (300 x 300 DPI)
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Langmuir Page 32 of 41

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 Figure 1. TEM images (scale bar = 50 nm) of a) CoxFe3-xO4-OA NPs, b) CoxFe3-xO4-DMSA NPs and c) CoxFe3-
20 xO4-DMSA-Au NPs with an expansion in the inset (bar 10 nm).
180x60mm (150 x 150 DPI)
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 33 of 41 Langmuir

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
27
28
29
30
31
32
33
34
35
36
37
38
39
Figure 2. a) EELS spectrum of CoxFe3-xO4-OA NPs where the two edges starting at 708eV and 779eV,
40
correspond to the FeL2,3 and CoL2,3 edges, respectively; b) Reference TEM image to be compared with
41 the ESI maps of iron (c, red) and cobalt (d, green). The bars correspond to 50 nm.
42 84x82mm (600 x 600 DPI)
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Langmuir Page 34 of 41

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18 Figure 3. a) UV-Vis spectra of the CoxFe3-xO4-DMSA NPs and CoxFe3-xO4-DMSA-Au NPs obtained by a single
19 injection or by a slow dropping (41.5µL/min); b) STEM Image of CoxFe3-xO4-DMSA-Au NPs obtained by slow
20 dropping; c) STEM Image of CoxFe3-xO4-DMSA-Au NPs obtained by single injection. The scale bars
21 correspond to 20 nm.
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 35 of 41 Langmuir

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
27
28
29
30
31
32
Hysteresis loops of the cobalt ferrite samples before and after gold decoration. Hollow circles: CoxFe3-xO4-
33
DMSA; solid circles: CoxFe3-xO4-DMSA-Au.The magnetization is given relative to the mass of cobalt ferrite.
34 64x48mm (300 x 300 DPI)
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Langmuir Page 36 of 41

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
27
28
29
30
31
32
33
Figure 5. UV-vis spectra of 4-nitrophenol, sodium 4-nitrophenolate and sodium 4-aminophenolate
34 202x160mm (96 x 96 DPI)
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 37 of 41 Langmuir

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23 Figure 6. a) The reduction efficiency of 4-NP by: ■ No catalyst; ● CoxFe3-xO4-DMSA NPs; ∆ 10% Au-DMSA
24 NPs; ▲ 10% CoxFe3-xO4-DMSA-Au NPs; □ 5% CoxFe3-xO4-DMSA-Au NPs; ○ 2% CoxFe3-xO4-DMSA-Au NPs.
Remaining Ratio is defined as the residual percentage of 4-nitrophenolate in the reaction solution, calculated
25
from the absorbance values as explained in SI. b) The observed kinetic constant kobs for 10% Au-DMSA
26 NPs and 2%, 5% and 10% CoxFe3-xO4-DMSA-Au NPs.
27 116x53mm (300 x 300 DPI)
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Langmuir Page 38 of 41

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
27
28
29
30
31
32
33
34
35
36
37
38
39
40 Figure 7. Cyclovoltammetric reduction patterns (0.1 V s-1 scan rate) of 4-NP (60 µL of a saturated aqueous
41 solution in 20 mL of 0.1 M NaClO4 aqueous solution) on the GC support and on three different modified
42 electrodes.
43 84x84mm (300 x 300 DPI)
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 39 of 41 Langmuir

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 Figure 8. Reduction efficiency of 4-NP in the consecutive catalytic cycles (a) and the corresponding kobs
22 calculated for each cycle (b).
23 104x43mm (300 x 300 DPI)
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Langmuir Page 40 of 41

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
27
28
29
30
31 Figure 9. The concentration of Fe, Co and Au in: digestion blank, catalytic reaction supernatant, recovered
32 catalyst after one catalytic cycle, and fresh catalyst.
33 129x92mm (300 x 300 DPI)
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 41 of 41 Langmuir

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
27
28
29
30
31
32
33
34
35
36
37
38
39
40 Figure 10. TEM image of the recovered catalyst after one cycle. The bar corresponds to 20 nm.
41 84x84mm (150 x 150 DPI)
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment

You might also like