Download as pdf or txt
Download as pdf or txt
You are on page 1of 76

Subscriber access provided by Warwick University Library

Article
Kinetics of Oxygen Electroreduction on Me-N-C (Me = Fe, Co, Cu)
Catalysts in Acidic Medium. Insights on the Effect of the Transition Metal
Luigi Osmieri, Alessandro H. A. Monteverde Videla, Pilar Ocón, and Stefania Specchia
J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.7b02455 • Publication Date (Web): 01 Aug 2017
Downloaded from http://pubs.acs.org on August 3, 2017

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.

The Journal of Physical Chemistry C 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 75 The Journal of Physical Chemistry

1
2
3 1 Kinetics of Oxygen Electroreduction on Me-N-C (Me = Fe, Co, Cu)
4
5
6 2 Catalysts in Acidic Medium. Insights on the Effect of the Transition
7
8
9 3 Metal.
10
11
12 4
13
14 5 Luigi Osmieri a,b *, Alessandro H. A. Monteverde Videla a, Pilar Ocón b, Stefania Specchia a *
15
16 6
17
18 a
7 Politecnico di Torino, Dipartimento di Scienza Applicata e Tecnologia, Corso Duca degli Abruzzi
19
20
21 8 24, 10129, Torino, Italy.
22
b
23 9 Universidad Autónoma de Madrid, Departamento de Química Física Aplicada, C/Francisco
24
25 10 Tomás y Valiente 7, 28049, Madrid, Spain.
26
27 11 * Corresponding Authors:
28
29
30 12 luigi.osmieri@polito.it, ligiosmi@gmail.com
31
32 13 stefania.specchia@polito.it
33
34 14
35
36 15 Abstract
37
38
16 The influence of three different transition metals (Me = Fe, Co, Cu) on the oxygen reduction
39
40
41 17 reaction (ORR) kinetics in acidic medium of Me-N-C catalysts synthesized using Me(II)-
42
43 18 phthalocyanine as precursors is investigated in this work. Through a detailed electrochemical
44
45 19 characterization using cyclic voltammetry and rotating ring-disk electrode, several kinetics
46
47 20 parameters such as Tafel slope, reaction order for oxygen and proton, apparent activation energy,
48
49
50 21 selectivity towards hydrogen peroxide production, and kinetics of reduction of adsorbed oxygen
51
52 22 were determined. The behavior of these three catalysts is analyzed in detail. A comparison between
53
54 23 each other of the catalysts, and with a Pt-based catalyst is done. The results obtained provide clear
55
56 24 evidence of the important role played by each transition metal in the formation of more or less
57
58
25 effective active sites. The ORR kinetics behavior can be well interpreted according to the
59
60 1
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 2 of 75

1
2
3 1 occurrence of a redox-mediated coverage of the active sites at low overpotentials (close to the ORR
4
5 2 onset), which has influence on the Tafel slope, as well as on the oxygen adsorption and activation
6
7 3 energy of the process. The results clearly show that, among the other transition metals considered,
8
9
10 4 Fe is the best performing one in carrying out the ORR.
11
12 5
13
14 6 1. Introduction
15
16 7 To make the commercialization of proton exchange membrane fuel cells (PEMFC) viable in the
17
18
8 next future, a fundamental point will be to lower as much as possible their platinum content.1 In
19
20
21 9 facts, Pt is used as the preferred catalyst in these devices at both anode and cathode, due to its
22
23 10 enhanced activity towards both hydrogen oxidation reaction (HOR) and oxygen reduction reaction
24
25 11 (ORR),2 but it suffers from two main problems: high cost and scarcity.3 The kinetics of ORR
26
27 12 occurring at the cathode is several orders of magnitude slower compared to the anodic HOR, and
28
29
30 13 consequently most of the Pt is located at the cathode of PEMFC rather than at the anode.4 In this
31
32 14 regard, many research efforts have been pursued in the last decades to develop Pt-free ORR
33
34 15 catalysts, using several types of precursor materials.5 Since from the discovery in the early ’60 that
35
36 16 macrocyclic molecules containing coordinated transition metals (mainly Fe and Co) are active
37
38
17 towards ORR,6 and then with the discovery that the activity and the stability can be considerably
39
40
41 18 improved after pyrolyzing the precursors, several types of Me-N-C materials have been synthesized
42
43 19 and tested as ORR catalysts.7–9 In recent years, the electroactivity of some of these materials, which
44
45 20 was originally considerably lower compared to Pt-based catalysts in acidic medium, has also shown
46
47 21 remarkable and encouraging improvements.10–14
48
49
50 22 The ORR kinetics in acidic conditions on Pt-based catalysts under different forms, i.e. single crystal
51
52 23 Pt electrodes,15 polycrystalline Pt,16 and Pt catalysts supported on various carbon-based materials
53
54 24 such as Vulcan XC72R,17 have been extensively studied in the past decades. These studies led to the
55
56 25 determination among other things, of the apparent activation energy, reaction order for O2,
57
58
26 predominant reaction mechanism, O2 adsorption mechanisms and rate determining steps of the
59
60 2
ACS Paragon Plus Environment
Page 3 of 75 The Journal of Physical Chemistry

1
2
3 1 ORR.18–21 On the other hand, the literature is scarce about analogous studies conducted on Pt-free
4
5 2 ORR catalysts, and in particular for Me-N-C catalysts. An exhaustive study was conducted by
6
7 3 Chlistunoff22 on a Fe-N-C catalyst synthesized using polyaniline as N precursor. Jaouen et al.
8
9
10 4 performed an activation energy calculation on a Fe-N-C catalyst for ORR.23 However, no
11
12 5 comparative studies have been carried out so far investigating the effects of different transition
13
14 6 metals such as Fe, Co, and Cu on the behavior of these Me-N-C catalysts on such kinetic
15
16 7 parameters like activation energy and reaction order. The role of the transition metal present in the
17
18
8 precursor(s) used for the synthesis of pyrolyzed NNM catalysts for ORR has been discussed in the
19
20
21 9 literature from the point of view of the ORR activity and selectivity induced in the final catalysts in
22
23 10 some other works.24–28
24
25 11 In spite of the debate about the fact that the transition metal is really part of the ORR active site, or
26
27 12 it just serves for its formation during the pyrolysis is still going on,29 it is commonly assumed that
28
29
30 13 this transition metal center coordinated with N atoms is the ORR active site in unpyrolyzed
31
32 14 macrocyclic molecules like porphyrins, corroles and phthalocyanines.30–34 These molecular
33
34 15 structures are commonly found in living beings (i.e. in hemoglobin and cytochrome-C-oxidase),
35
36 16 which use molecular dioxygen to carry out cellular respiration. These molecules are proteins, which
37
38
17 have the “ability” to bind with O2, and transport it from one site to another through the body, or to
39
40
41 18 act as proper catalysts for the 4-electrons O2 reduction to H2O. In nature, the metal centers involved,
42
43 19 for example in the transport of O2 in blood, mono- or di-oxygenation involving oxygen atom
44
45 20 incorporation, and substrate oxidation (dehydrogenation or removal of electrons), are Fe and
46
47 21 Cu.35,36 Interestingly, Co ion is not present in such proteins, and the substitution of Fe with Co in
48
49
50 22 human adult hemoglobin (HbA) has shown a reduction in oxygen affinity by over a factor of 10, an
51
52 23 oxygen affinity too low for use as a blood substitute.37
53
54 24 In this work, our purpose is to investigate in detail the behavior of a series of Fe-N-C, Co-N-C, and
55
56 25 Cu-N-C catalysts towards ORR in acidic medium in a three-electrode cell configuration. This study
57
58
26 provides a deeper insight to better understand the influence of the different transition metals (Fe,
59
60 3
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 4 of 75

1
2
3 1 Co, and Cu) on the ORR kinetics in acidic medium, in particular regarding the apparent activation
4
5 2 energy, reaction orders for O2 and H+, behavior towards the reduction of the adsorbed O2 via cyclic
6
7 3 voltammetry, and main reaction pathway. This would enable to compare the behavior of these Me-
8
9
10 4 N-C catalysts among each other, and with Pt-based catalysts, which have been more extensively
11
12 5 described in the literature.
13
14 6 Our findings reveal good agreement with the proposed theory of potential-dependent ORR active
15
16 7 sites coverage and consequent blocking effect, which is in strict relation to the red-ox potential of
17
18
8 the metal-based active sites, and at the origin of the “asymmetric volcano plot” for these Me-N-C
19
20
21 9 catalysts, as recently proposed in the literature.38–40 This study provides new insights supported by
22
23 10 several experimental pieces of evidence about the effect of using Fe, Co, or Cu as the transition
24
25 11 metal precursor, and their influence on the ORR performance of the final catalyst.
26
27 12
28
29
30 13
31
32 14 2. Experimental Section
33
34 15 2.1. Catalysts synthesis.
35
36 16 The catalysts were synthesized as described in our previous work.41 Briefly, Me(II)-phthalocyanine
37
38
17 (Me = Fe ,Co, Cu) were dissolved in a proper solvent and impregnated on SBA-15 ordered
39
40
41 18 mesoporous silica used as a sacrificial template. After complete solvent evaporation, a pyrolysis
42
43 19 under inert atmosphere (flowing N2) was carried out in a tubular quartz furnace for 1 h at 800 °C.
44
45 20 Then the SBA-15 silica was removed by washing in 5 wt. % HF solution.
46
47 21 2.2. Ink formulation for RDE tests.
48
49
50 22 For RDE tests of Me-N-C catalysts, the ink was prepared by dispersing a given mass of catalyst
51
52 23 (mcat, typically around 10 mg) in a solution obtained mixing known volumes of isopropanol,
53
54 24 deionized water, and 5 wt. % Nafion® ionomer hydro-alcoholic solution. An ink formulation is
55
56 25 characterized by its mass ratio of Nafion ionomer to catalyst, or Nafion-to-catalyst-ratio (NCR).
57
58
59
60 4
ACS Paragon Plus Environment
Page 5 of 75 The Journal of Physical Chemistry

1
2
3 1 The volumes (in µL) of Nafion solution, deionized water and isopropanol to use to prepare the ink
4
5 2 are calculated as follows:
6
7  ∙
 ∙ 
8  =
9 0.05
10
11  = 15 ∙

12
13  = 50 ∙
 −   − 
14
15
3 Where mcat is expressed in mg and ρnaf is the density of the Nafion 5 wt. % solution expressed in g
16
17
18 4 mL−1. With this formulation, the catalyst density in the ink is 0.02 mg µL−1. The ink is kept under
19
20 5 sonication (130W, Soltec 2200M3S sonicator) for 30 min to achieve a good dispersion. To have the
21
22 6 desired catalyst loading, a proportional volume of ink is pipetted on the RDE electrode surface.
23
24 7 For comparison, a commercial Pt-based catalyst (20 wt. % Pt on Vulcan XC-72 QuinTech,
25
26
27 8 Germany) was also tested. For this catalyst, the ink was prepared by dispersing 10 mg of catalyst,
28
29 9 33 µL of 5 wt. % Nafion® solution, 734 µL of isopropanol, and 20 µL of deionized water. This led
30
31 10 to having 2.5 µg of Pt per µL of ink. After 30 minutes under sonication, a proportional volume of
32
33 11 ink was pipetted to have a Pt loading on the electrode of 38 µg cm−2.
34
35
12
36
37
38 13 2.3. Electrochemical tests.
39
40 14 The electrochemical tests were conducted in a conventional three-electrodes electrochemical cell
41
42 15 configuration, using a rotating disk electrode equipment (RRDE-3A ALS, Japan) and a multi-
43
44 16 potentiostat (Bio-Logic SP-150, France). The cell was equipped with a glassy carbon disk (4 mm
45
46
47 17 diameter) – Pt ring (7 and 5 mm outer and inner diameter, respectively) working electrode, a Pt
48
49 18 helical wire counter electrode, and a saturated calomel reference electrode (SCE). For Me-N-C
50
51 19 catalysts, the electrolyte was an aqueous solution of H2SO4 (in different concentrations depending
52
53 20 on the type of test), while for Pt/C catalyst was 0.1 M HClO4 solution was used (to avoid the
54
55
56
21 detrimental effect of SO42− and HSO4− ions adsorption on Pt, being HClO4 a non-specifically
57
58 22 adsorbing electrolyte for Pt42). The electrolyte was saturated with pure N2, pure O2 or with mixtures
59
60 5
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 6 of 75

1
2
3 1 of both gases in different proportions, depending on the type of test, by direct gas bubbling into the
4
5 2 solution. For N2–O2 mixtures, the gas flows were controlled using two mass flowmeters
6
7 3 (Bronkhorst ELFLOW series, Netherlands), maintaining constant the total flow rate at 150 NmL
8
9
10 4 min−1. For RRDE measurements, the ring potential was kept at 1.2 V vs. RHE. At this potential, the
11
12 5 H2O2 oxidation reaction is under diffusion control.42 The RRDE tests were conducted using a bi-
13
14 6 potentiostat (CH Instruments Mod. CH760E USA). The ring background current measured at
15
16 7 potentials higher than the reaction onset was subtracted from the ring signal throughout the whole
17
18
8 scanned potential range.43
19
20
21 9 Before starting the tests, 50 cyclic voltammetry (CV) cycles at 100 mV s–1 scan rate were
22
23 10 performed in the potential window 0.0–1.2 V vs. RHE in N2 saturated electrolyte, to obtain a clean
24
25 11 and stable working electrode surface.44
26
27 12 For the ORR activity measurements, staircase voltammetries (SV) were recorded from 1.2 to 0.0 V
28
29
30 13 vs RHE in O2-saturated electrolyte with a potential step of 0.01 V and a holding time at each
31
32 14 potential of 30 s. In this way, the background capacitive current had passed, and a steady-state value
33
34 15 of the faradaic current was measured, enabling to obtain steady-state polarization curves.45 The
35
36 16 RDE rotation speed was set at 900 rpm.
37
38
17 For the adsorbed oxygen electroreduction tests, CV cycles at different scan rates (within the range
39
40
41 18 5– 500 mV s−1) were recorded first under N2-saturated electrolyte and then in O2-saturated
42
43 19 electrolyte, in the potential window from 1.2 to 0.0 V vs RHE. For the CV experiments in the
44
45 20 presence of O2, at least two consecutive full CV cycles were recorded, to enable the subtraction of
46
47 21 the second cycle from the first one, as done for the calculation of the differential voltammograms
48
49
50 22 (to eliminate the contribution of the reduction of diffused O2). In addition, between one CV
51
52 23 experiment at a certain scan rate and the subsequent one, to allow O2 to adsorb onto the catalyst
53
54 24 surface completely, O2 was left bubbling into the electrolyte for at least 15 minutes. This time was
55
56 25 verified to be enough to assure a saturation of the catalyst surface with O2 (no difference in the
57
58
26 ORR current peaks was observed with higher saturation times).
59
60 6
ACS Paragon Plus Environment
Page 7 of 75 The Journal of Physical Chemistry

1
2
3 1 For the Koutecky-Levich experiment, linear sweep voltammetries (LSV) were recorded at different
4
5 2 rotation speeds (200–500–900–1600–2500–3600 rpm) and 5 mV s–1 scan rate.
6
7 3 For activation energy calculation experiments, LSV were recorded at 5 mV s–1 scan rate and with a
8
9
10 4 RDE rotation speed of 900 rpm at different temperatures (10–15–20–25–32–40–50–60 °C) by
11
12 5 placing the cell in a thermostatic bath, which temperature was regulated and controlled by a water
13
14 6 heating-cooling system (Mod. CRIOTERM 190, I.S.CO. s.r.l., Italy). These measurements were
15
16 7 carried out in isothermal conditions, being the reference and working electrode both placed inside
17
18
8 the thermostatic cell.
19
20
21 9 For the Rotating Ring-Disk Electrode (RRDE) test, the ring potential was fixed at 1.2 V vs RHE,
22
23 10 the disk potential was scanned at 5 mV s−1, and the electrode rotation speed was set to 900 rpm.
24
25 11 At the end of each test, an electrochemical impedance spectroscopy measurement was done at the
26
27 12 open circuit voltage (OCV), with a wave amplitude of 10 mV and frequencies in the range of 10
28
29
30 13 kHz–1 Hz. The high-frequency resistance value was used to subtract the ohmic drop contribution
31
32 14 from SV and LSV curves.42,46
33
34 15 Hereafter, the electrode potentials were corrected and referred to the reversible hydrogen electrode
35
36 16 (RHE), and the current densities were normalized to the geometric area of the glassy carbon disk
37
38
17 electrode.
39
40
41 18
42
43 19
44
45 20 3. Results and Discussion
46
47 21 3.1. Preliminary electrode optimization: catalyst loading and Nafion-to-catalyst ratio.
48
49
50 22 A preliminary test on the preparation of the working electrode for RDE tests was performed using
51
52 23 the Fe-N-C catalyst (which is the most active catalyst, as shown afterward). The first part of the
53
54 24 optimization consisted in testing different catalyst loadings deposited on the glassy carbon surface
55
56 25 of the RDE. The chosen amounts were determined from the ink density, that is, the catalyst mass
57
58
26 content per volume of ink (mg mL−1). According to the ink formulation described in Section 2.3,
59
60 7
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 8 of 75

1
2
3 1 five different electrodes were prepared, by pipetting 1–2–3–4–5 µL of ink on the RDE. These ink
4
5 2 volumes result in 0.16–0.32–0.48–0.64–0.80 mg of catalyst per cm2 geometric electrode area,
6
7 3 respectively. Then, polarization curves were recorded in the O2-saturated electrolyte with a rotation
8
9
10 4 speed of 900 rpm. As shown in Figure 1a, better results were obtained increasing the catalyst
11
12 5 loading from 0.16 to 0.64 mg cm−2. No improvement was obtained with a further increase to 0.80
13
14 6 mg cm−2. The iL values also tend to increase as the catalyst loading increase.
15
16 7 To eliminate the effects of the mass transport limitations, the kinetic currents were calculated by
17
18
8 Equation (1):47
19
20
 ∙
21 9  = −   (1)

22
23
24 10 Where: ik is mass transport-corrected current density, i is the measured current density, and iL is the
25
26 11 limiting current density measured in the plateau region of the polarization curve at high
27
28 12 overpotential. A correction for the ohmic drop was also done, based on the resistance values
29
30 13 obtained by the EIS measurement as described in Section 2.3. The as-calculated ik values can be
31
32
33
14 transformed in specific mass current densities (im, A g−1), simply dividing by the catalyst loading on
34
35 15 the electrode (mg cm−2). Figure 1b shows the as calculated potential vs. logarithm of specific mass
36
37 16 current density plot.
38
39 17
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 8
ACS Paragon Plus Environment
Page 9 of 75 The Journal of Physical Chemistry

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 1
38
39 2 Figure 1. (a) SV of Fe-N-C catalyst in O2-saturated 0.5 M H2SO4 at 900 rpm with different catalyst
40 3 loadings and NCR = 0.2. (b) Tafel plot obtained from (a) after mass-transport correction and
41
4 normalization per mass of catalyst on the electrode.
42
43 5
44
45 6 However, the differences found in the ORR measurements with different catalyst loadings (Figure
46
47 7 1a) were reduced after the mass transport correction and the current density normalization per mass
48
49 8 of catalyst (Figure 1b). The results with 0.64 mg cm−2 catalyst loading are still slightly better than
50
51
52 9 the others, especially at higher overpotentials. Based on these results, if not differently stated, we
53
54 10 decided to fix the catalyst loading used hereafter for the RDE experiments to 0.64 mg cm−2.
55
56 11 In the second part of the preliminary optimization study, the effect of NCR was investigated,
57
58 12 keeping fixed at 0.64 mg cm−2 the catalyst loading. Different inks were prepared as described in
59
60 9
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 10 of 75

1
2
3 1 Section 2.2, with four different NCR: 0.10–0.20–0.33–0.50. As shown in Figure 2a, in spite of the
4
5 2 constant catalyst loading, the capacitive currents vary changing the NCR, having a maximum for
6
7 3 NCR = 0.2. The influence of Nafion on the capacitive current could be linked to pseudo-capacitance
8
9
10 4 phenomena, and to modifications of the contact interface between the catalyst surface and the
11
12 5 electrolyte.23,48–50 However, the Nafion ionomer not only provides proton conductivity but, more
13
14 6 importantly in the RDE system, also acts as a binder for the catalytic particles, causing
15
16 7 modifications of the percolating network for electrons and protons within the catalyst layer.23
17
18
8 Therefore, the decrease in the capacitive current density with the increase of NCR can be explained
19
20
21 9 by the formation of a thick Nafion film, which encapsulates some of the catalytic particles, causing
22
23 10 a loss of electrical contact with the electrolyte. Similar results were also found in the literature.49,51
24
25 11
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 10
ACS Paragon Plus Environment
Page 11 of 75 The Journal of Physical Chemistry

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
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55 1
56 2 Figure 2. (a) CV of Fe-N-C catalyst with an ink prepared with different NCR recorded in N2-
57
58 3 saturated 0.5 M H2SO4 solutions at 10 mV s−1 with a catalyst loading of 0.64 mg cm−2. (b) SV of
59
60 11
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 12 of 75

1
2
3 1 Fe-N-C catalyst with an ink prepared with different NCR in O2-saturated 0.5 M H2SO4 at 900 rpm.
4 2 (c) Tafel plot obtained from polarization curves in (b) after mass-transport correction.
5
3
6
7
8 4 The ORR tests in Figure 2b show that different NCR cause also variations in the ORR activity of
9
10 5 the Fe-N-C catalyst. The best performance was obtained with NCR = 0.20. The mass-transport
11
12 6 corrected kinetic current density was also calculated using Equation (1) as described before. Figure
13
14 7 2c shows that the differences among the different NCR are less evident in the kinetically controlled
15
16
17 8 region (low overpotentials), but they become more evident in the mixed kinetic-diffusion controlled
18
19 9 region and in the diffusion-limited current density region at higher overpotentials. By increasing the
20
21 10 NCR to 0.33 and 0.50, the diffusion-limited current density slightly decreases. This effect could be
22
23 11 explained by the formation of a thicker Nafion film around catalytic particles, which could limit the
24
25
12 transport of oxygen to the catalyst active sites, leading to a decrease of the limiting current.52 Also
26
27
28 13 with a lower NCR of 0.10, the performance was worse. This fact could be related to the binding-
29
30 14 effect between Nafion and the catalyst particles in the catalyst layer. In the literature, the effect of
31
32 15 NCR on ink formulation has been studied, with the NCR mass ratio varying generally from 0.10 to
33
34 16 1.50,23,51,53–55 but also with considerably lower values (between 0.014 and 0.087).49 The optimum
35
36
37 17 NCR for ORR activity depends on the nature of the different catalysts characteristics (i.e. active
38
39 18 sites, surface area, pore size distribution, hydrophobicity), which play a crucial role in the final
40
41 19 activity. In particular, as demonstrated in recent works,48,49 Nafion could form self-assemblies onto
42
43 20 the surface of a catalyst depending on the hydrophobicity/hydrophilicity of its carbonaceous
44
45
21 surface. This points out the importance of performing such this type of NCR optimization before
46
47
48 22 testing a new electrocatalyst for PEMFC applications.
49
50 23 Since the highest activity for this Fe-N-C catalyst was obtained with an NCR = 0.2, this NCR value
51
52 24 was used for all the tests in the present work, considering that the other Me-N-C catalysts have
53
54 25 similar morphological and superficial composition characteristics.41
55
56
57 26
58
59 27 3.2. ORR activity.
60 12
ACS Paragon Plus Environment
Page 13 of 75 The Journal of Physical Chemistry

1
2
3 1 Figure 3a shows the ORR steady-state polarization curves of the Me-N-C catalysts synthesized
4
5 2 using different Me(II)-phthalocyanine precursors (Me = Fe, Co, Cu, Zn) and for the catalyst H-N-C,
6
7 3 prepared using the unmetallized phthalocyanine (H-Pc), as described in our previous work.41
8
9
10 4
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 5
47 6
48 7 Figure 3. (a) SV of the Me-N-C catalysts (loading 0.64 mg cm−2) measured in O2-saturated 0.5 M
49
50 8 H2SO4 at 900 rpm. The SV of a commercial Pt/C catalyst measured in 0.1 M HClO4 in both
51 9 cathodic and anodic scan directions is also shown for comparison. (b) Plot of the linear zones of the
52 10 potential vs logarithm of the mass-transport corrected current densities from (a).
53
11
54
55
56
57
58
59
60 13
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 14 of 75

1
2
3 1 Tafel slopes were calculated after correction for mass-transport limitation, starting from the data in
4
5 2 Figure 3a. Cathodic transfer coefficients and exchange current densities were also calculated, as
6
7 3 summarized in Table 1. The cathodic transfer coefficient αc is defined by Equation (2):56,57
8
9
 " $%|' |
10
4  = − (2)
11 ! "(
12
13
14 5 where: E is the applied potential, R is the gas constant, T is the absolute temperature, F is the
15
16
17 6 Faraday constant and ic is the cathodic current density. Since the symbol ln|ic| implies that the
18
19 7 argument of the logarithm is of dimension one, the ic value is ideally divided by the corresponding
20
21 8 unit, e.g., ln(|ic| / mA cm–2). Therefore, in practice, αc is defined as the reciprocal of the
22
23 9 corresponding Tafel slope, –dE/dln|ic|, made dimensionless by the multiplying factor RT/F.56,57
24
25
10 Other characteristic parameters which define the performance of an ORR catalyst are the onset
26
27
28 11 potential (Eon) and the half-wave potential (E1/2). In this paper, Eon has been defined as the potential
29
30 12 required to generate a current density of 0.1 mA cm−2 in a steady-state RDE experiment. This
31
32 13 definition is arbitrary, and we assumed it following other literature works, in order to minimize the
33
34 14 effect of residual currents in the staircase voltammetry RDE measurement.12 Thus, it has to be
35
36
37 15 intended as an experimental “screening factor” that can be used for example to discriminate whether
38
39 16 an ORR electrocatalyst shows a potentially good activity to deserve to be tested in a PEMFC
40
41 17 device. E1/2 is the potential required to have half the maximum current density in the polarization
42
43 18 curve.
44
45
46 19 Table 1. ORR parameters calculated from the data in Figure 3 for all the Me-N-C catalysts and the
47
48 20 commercial Pt/C catalyst.
49 Low η region High η region
50 Eon E1/2 Tafel slope i0 Tafel slope i0
Catalyst αc αc
51 [V vs RHE] [V vs RHE] [mV dec−1] [mA cm−2] [mV dec−1] [mA cm−2]
52 Fe-N-C 0.83 0.72 64.0 0.924 5.14 · 10−8 140.8 0.420 8.67 · 10−4
53 Co-N-C 0.79 0.71 53.8 1.099 1.00 · 10−9 143.3 0.413 7.15 · 10−4
Cu-N-C 0.68 0.50 - - - 123.2 0.480 3.66 · 10−6
54
Zn-N-C 0.50 0.33 - - - 131.8 0.449 3.26 · 10−7
55
H-N-C 0.50 0.32 - - - 140.1 0.422 7.45 · 10−7
56
Pt cath 0.87 0.76 60.5 0.977 1.41 · 10−7 202.2 0.292 3.63 · 10−2
57
Pt anod 0.93 0.82 62.6 0.944 2.30 · 10−6 125.7 0.470 3.93 · 10−3
58
21
59
60 14
ACS Paragon Plus Environment
Page 15 of 75 The Journal of Physical Chemistry

1
2
3 1 Analyzing the data shown in Table 1, Fe-N-C is the most active catalyst. It exhibits 100 mV
4
5 2 negative shift in both Eon and E1/2 in comparison with Pt/C (considering the anodic scan direction).
6
7 3 Moreover, its diffusion limiting current almost corresponds to the limiting current observed for
8
9
10 4 Pt/C. Co-N-C has a slightly lower activity, and a lower diffusion limiting current, with a not well-
11
12 5 developed plateau region. Cu-N-C is considerably less active than the two previous catalysts,
13
14 6 having about 250 mV lower Eon and 320 mV lower E1/2 than the Pt/C. Zn-N-C and H-N-C have a
15
16 7 very poor activity in acid medium, both showing 430 mV lower Eon and about 500 mV lower E1/2
17
18
8 compared to Pt/C.
19
20
21 9 Remarkably, that Zn-N-C and H-N-C exhibit almost the same ORR activity, being their polarization
22
23 10 curves practically superimposed (see Figure 3a-b). This suggests that even if the Zn presence leads
24
25 11 to considerable modifications in the final catalyst chemical-physical properties (e.g. surface area
26
27 12 and pore size distribution, total amount of N incorporated),41 this will have practically no influence
28
29
30 13 on the ORR activity in acidic medium. Similar results were found for tests in alkaline medium in
31
32 14 our previous work.41 This finding confirms the important role played by the most effective
33
34 15 transition metals (Fe and Co) in the formation of the ORR active sites during the heat treatment. Cu
35
36 16 also has certain effectiveness in the formation of active sites, but considerably lower than Co and
37
38
17 Fe.
39
40
41 18 In the Tafel plots of both Fe-N-C and Co-N-C two different Tafel slope zones have been identified
42
43 19 (see Figure 3b), a first one at low overpotentials (low η), approximately between 0.85 and 0.75 V
44
45 20 vs. RHE, and a second one at high overpotentials (high η), approximately between 0.75 and 0.60 V
46
47 21 vs. RHE. At low η the Tafel slope is lower, being approximately of 60 mV dec−1. At high η the
48
49
50 22 slope increases, and the linear trend is less evident, suggesting that a change in the reaction
51
52 23 mechanism is occurring.22 A linearization was also made in this zone, in spite of the lower
53
54 24 coefficient of determination (R2), and the calculated Tafel slopes are around 140 mV dec−1. An
55
56 25 analogous behavior was found by many other authors for similar Me-N-C catalysts.22,23,26,58,59
57
58
26 Considering the exchange current densities in the low η region, Fe-N-C has an almost double i0
59
60 15
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 16 of 75

1
2
3 1 value than Co-N-C, whereas Pt/C shows an almost one and two orders of magnitude higher i0 value
4
5 2 compared to Fe-N-C in the cathodic and anodic scan direction, respectively.44,60 In the high η
6
7 3 region, the i0 values for Fe-N-C and Co-N-C are closer, being again almost one and two orders of
8
9
10 4 magnitude lower than for Pt/C catalyst in the cathodic and anodic scan direction, respectively.
11
12 5 Cu-N-C catalyst does not exhibit the double slope behavior, and a single linear region with a slope
13
14 6 value of about 123 mV dec−1 is observed along all the potential range from 0.68 to 0.55 V vs. RHE.
15
16 7 The i0 of Cu-N-C in the high η region is almost two orders of magnitude lower compared to Fe-N-C
17
18
8 and Co-N-C.
19
20
21 9 Zn-N-C and H-N-C catalysts also show a single-slope behavior in the potential range 0.5–0.3 V vs.
22
23 10 RHE, with a Tafel slope of 132 and 140 mV dec−1, respectively, and i0 values are one order of
24
25 11 magnitude lower compared to Cu-N-C. Since the ORR activity of Zn-N-C and H-N-C in acidic
26
27 12 medium is very poor and thus they are not interesting for potential application as cathodic catalysts
28
29
30 13 for PEMFC, hereafter we will not consider these two catalysts for the investigation of the other
31
32 14 kinetic aspects of the ORR.
33
34 15 To investigate more in detail the effect of the catalyst loading on the ORR kinetics, we recorded the
35
36 16 polarization curve in RDE at 900 rpm rotation speed with two different ink quantities deposited on
37
38
17 the electrode for Fe-N-C, Co-N-C, and Cu-N-C catalysts. Figure 4 shows the Tafel plots of the
39
40
41 18 respective SV polarization curves in Figure 3a for all of the catalysts with 0.64 mg cm−2 loading,
42
43 19 and in Figure 1a for Fe-N-C with 0.16 mg cm−2 loading. The SV polarization curves for Co-N-C
44
45 20 and Cu-N-C with 0.32 mg cm−2 loading are not reported here.
46
47 21
48
49
50
51
52
53
54
55
56
57
58
59
60 16
ACS Paragon Plus Environment
Page 17 of 75 The Journal of Physical Chemistry

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 1
37 2
38 3 Figure 4. Mass transport corrected Tafel plots for Fe-N-C, Co-N-C, and Cu-N-C catalysts: (a)
39 4 Current density referred to the geometric area of the electrode; (b) Current density referred to the
40
5 mass of catalyst deposited on the electrode.
41
42 6
43
44 7 For all of the catalysts, when the current densities are referred to the geometric area of the electrode
45
46 8 (Figure 4a), the apparent ORR kinetics appears to be enhanced with a higher catalyst loading.
47
48 9 However, when the current densities are normalized per unit mass of catalyst deposited on the
49
50
10 electrode surface (Figure 4b), the plots with low catalyst loading undergo a vertical shift, making
51
52
53 11 them almost overlap the plots with high catalyst loading in the whole potential range considered.
54
55 12 Therefore, for all the catalysts, we obtained a good agreement between mass specific kinetic current
56
57 13 densities (A g−1) after the mass-transport correction for both high and low catalyst loadings. In
58
59
60 17
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 18 of 75

1
2
3 1 particular, for Fe-N-C and Co-N-C, this indicates that the curvature of the E vs. log|i| plots (with a
4
5 2 consequent changing in the Tafel slope), does not originate from incomplete catalyst utilization in
6
7 3 the catalytic layer61 or uncompensated resistance,46 but it is the consequence of changes in the
8
9
10 4 intrinsic ORR kinetics, which is potential-dependent.22
11
12 5 The behavior of the two most active catalysts (Fe-N-C and Co-N-C) in terms of observation of two
13
14 6 different Tafel slope regions at different overpotentials, can be compared to the behavior of Pt-
15
16 7 based catalysts, as also reported in our previous work.60 In particular, Pt-based catalysts also show a
17
18
8 change in the Tafel slope from ~60 mV dec−1 to ~120 mV dec−1 with the increase of η. For Pt-based
19
20
21 9 catalysts, this fact is attributed to a potential-dependent coverage and consequent blocking of ORR
22
23 10 active Pt surface by chemisorbed oxygen-species (i.e. thin Pt oxide layer). This occurrence causes a
24
25 11 change in the O2 adsorption conditions with the potential.39,62 In particular, at low η, where the Pt
26
27 12 surface is highly covered by oxide species, the O2 adsorption will be similar to the situation
28
29
30 13 described by a Temkin isotherm. At higher η instead, the Pt surface is practically free from oxide
31
32 14 coverage, and thus the O2 adsorption will take place in conditions similar to what described by a
33
34 15 Langmuir isotherm (that is, virtually no interaction between adsorbed molecules). However, this
35
36 16 does not involve a change in the rate determining step of the reaction, that remains the first electron
37
38
17 transfer.62
39
40
41 18
42
43 19 3.3. Electroreduction of adsorbed oxygen.
44
45 20 3.3.1. First method
46
47 21 The first analytical method we used to study the behavior of our catalysts towards the reduction of
48
49
50 22 the adsorbed O2 via CV, consists in subtracting the CV cycle recorded in N2-saturated electrolyte
51
52 23 from the first CV cycle recorded in O2-saturated electrolyte. Thus, in the residual current of the
53
54 24 differential curve (that is, the blue CV curve in Figure 5), the pure capacitive (electric double layer)
55
56 25 and pseudo-capacitive (surface functionalization red-ox processes) effects63 have been eliminated.
57
58
26 Thus, the differential curve contains the contribution of the reduction of both the oxygen adsorbed
59
60 18
ACS Paragon Plus Environment
Page 19 of 75 The Journal of Physical Chemistry

1
2
3 1 on the catalyst surface, and the oxygen diffusing from the bulk of the solution. At lower scan rates,
4
5 2 the contribution of the diffused oxygen is likely to be higher, because the lower scan rate provides
6
7 3 more time to the oxygen diffusing from the bulk of the electrolyte to reach the catalyst surface.
8
9
10 4 Figure 5 shows the results obtained for Fe-N-C, Co-N-C and Cu-N-C catalysts. Comparing them,
11
12 5 we can see that the intensity of the current density peak, Ip, in the differential CV is considerably
13
14 6 higher for Fe-N-C and Cu-N-C than for Co-N-C, as evidenced by the CV plots recorded at 10 mV
15
16 7 s−1. These differential CV show only one direct reduction peak, without the presence of any reverse
17
18
8 anodic peak, suggesting that the process is irreversible.64 In accordance with the theory of the
19
20
21 9 potential sweep techniques, Ip varies linearly with the square root of the scan rate, ν1/2, in a process
22
23 10 totally governed by diffusion.22,64 On the other hand, for a diffusionless process, totally ascribed to
24
25 11 reactions of species adsorbed on the electrode surface, the variation of Ip with ν1/2 is quadratic.22,65
26
27 12 For Fe-N-C the Ip values in the scan rate range considered (5–500 mV s−1) vary approximately
28
29
30 13 between 2.4 and 16.6 mA cm−2, as evident in the Ip vs. ν1/2 plot in Figure 5a, which shows an
31
32 14 almost quadratic trend (with a coefficient of determination, R2 = 0.968). However, the curve trend
33
34 15 could also be fitted with a straight line (but with a slightly lower R2 = 0.946). This suggests that the
35
36 16 ORR phenomena occurring have both the contribution of diffusion and surface adsorption.
37
38
17 For totally irreversible processes, it is possible to determine the cathodic transfer coefficient, αc,
39
40
41 18 defined as in Equation (2), by using the slope of the plot of the peak potential, Ep, of a CV vs. the
42
43 19 logarithm of the scan rate (ν), using Equation (3) for a diffusionless process, and Equation (4) for a
44
45 20 diffusion-controlled process:57
46
47 ).*
48 21  = − +$,-.∙! (3)
49
).*
 = − +$,-.∙)! (4)
50
51
22
52
53 23 These equations are less accurate when reactants and products are specifically adsorbed and the
54
55 24 degree of surface coverage approaches unity.57
56
57
58
59
60 19
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 20 of 75

1
2
3 1 However, looking at the Ep vs. log(ν) plot in Figure 5a, it seems that the slope is changing with the
4
5 2 potential, thus no linear trend can be found. As previously discussed, we can hypothesize ORR to
6
7 3 be a diffusionless process, and thus use Equation (3) to calculate αc. If we consider the slope as a
8
9
10 4 derivative, and we divide the plot into three different zones (corresponding to low, intermediate,
11
12 5 and high ORR overpotentials), we can find a quite good linear trend, with different slopes, and
13
14 6 consequently different αc. This lets to deduce that the reaction mechanism on Fe-N-C catalyst is
15
16 7 intrinsically changing with overpotential, similarly to what previously found for RDE tests
17
18
8 described in Section 3.2, and by Chlistunoff.22 Table 2 summarizes these values.
19
20
21 9
22
23 10 Table 2. Parameters for the calculation of αc derived from data in Figure 5.
24
25 E zone E range (V vs. RHE) Slope R2 αc
26 Fe-N-C Low η 0.70 – 0.60 –0.116 0.982 0.51
27 Intermediate η 0.60 – 0.45 –0.248 0.994 0.24
28 High η 0.40 – 0.10 –0.641 0.938 0.09
29 Co-N-C – – –0.148 0.957 0.20
30 Cu-N-C Intermediate η 0.50 – 0.30 –0.201 0.988 0.29
31
32
High η 0.30 – 0.00 –0.445 0.985 0.13
33 11
34
35 12 As discussed before, the Ip values of Co-N-C in the ν range considered (10–500 mV s−1) are
36
37 13 considerably lower compared to Fe-N-C, varying approximately between 0.7 and 5.5 mA cm−2. If
38
39 14 we consider the Ip vs. ν1/2 plot in Figure 5b, we observe that it shows an almost linear trend
40
41
42 15 (although showing a low R2 = 0.88, due to the poorly defined current peaks observed for this
43
44 16 catalyst), suggesting that the ORR is mainly governed by diffusion contribution. This fact well
45
46 17 agrees with the lower Ip found for Co-N-C. Concerning the Ep vs. log(ν) plot in Figure 5b, in the
47
48 18 scan rate range 10–200 mV s−1, where the ORR is almost controlled by diffusion, we got an
49
50
19 approximately linear trend (with R2 = 0.957). This corresponds to αc = 0.20, which similar to the
51
52
53 20 one found for Fe-N-C in the intermediate η region.
54
55 21 For Cu-N-C the Ip vs. ν1/2 plot in Figure 5c is much better fitted with a parabolic curve (R2 = 0.999),
56
57 22 than with a straight line (R2 = 0.946), suggesting that the process could be well approximated by the
58
59
60 20
ACS Paragon Plus Environment
Page 21 of 75 The Journal of Physical Chemistry

1
2
3 1 reduction of the O2 adsorbed on the catalyst surface. The difference between the behavior of Cu-N-
4
5 2 C compared to Fe-N-C catalyst can be ascribed to differences on the catalysts surface features,
6
7 3 which could influence the amount of surface-confined O2 within the catalytic layer. This can be
8
9
10 4 caused for example by a different Nafion agglomerates distribution, leading to a difference in the
11
12 5 relative contribution of the reduction of adsorbed and diffused O2 on within the two catalyst layers
13
14 6 of Fe-N-C and Cu-N-C. The Ip values of Cu-N-C in the ν range considered (10–350 mV s−1) are
15
16 7 comparable to the values found for Fe-N-C, varying approximately between 1.5 and 26.2 mA cm−2.
17
18
8 From the Ep vs. log(ν) plot in Figure 5c, considering that in this case we are under conditions of a
19
20
21 9 diffusionless process (with predominant reduction of adsorbed O2), the slope is changing with the
22
23 10 potential, and a single linear trend cannot be found, similarly to Fe-N-C. However, by dividing the
24
25 11 plot into two zones, a better linear trend with different slopes was obtained, suggesting a consequent
26
27 12 change of αc and reaction mechanism, like in the case of Fe-N-C.
28
29
30 13 The different behavior of Co-N-C compared to Fe-N-C and Cu-N-C, could suggest that O2 adsorbs
31
32 14 much more weakly on Co-based active sites than on Fe- and Cu-based sites. These results and
33
34 15 considerations open an insight on the role that the different transition metals play in ORR, and on
35
36 16 their effectiveness during the pyrolysis in the formation of good ORR active sites.
37
38
17 In addition, these findings could indicate that the Fe-based and the Cu-based active sites exhibit a
39
40
41 18 similar behavior toward the adsorption of O2 and the reduction of the adsorbed O2, except that Fe-
42
43 19 sites perform their role at more positive potentials, resulting in a better ORR catalytic performance
44
45 20 for PEMFC applications. All these results are in accordance with the theory of redox-mediated
46
47 21 formation of the active sites proposed in the literature.38–40,66 Thus, the redox potential (Eredox) of Fe-
48
49
50 22 N-C active sites is more close to the top of the “volcano plot”. For Cu-N-C, Eredox could be located
51
52 23 more negatively on the left side of the asymmetric “volcano plot”. The active sites of Co-N-C show
53
54 24 higher ORR Eon compared to Cu-N-C (almost comparable to Fe-N-C), but their ability to adsorb
55
56 25 and subsequently reduce O2 is remarkably lower. These results could be a further confirmation of
57
58
59
60 21
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 22 of 75

1
2
3 1 the fact that the active sites formation Eredox potential for Co-N-C is located in more positively than
4
5 2 Fe-N-C, on the right side of the asymmetric “volcano plot”.
6
7 3
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 4
40 5
41 6 Figure 5. CV recorded at 10 mV s−1 in electrolyte saturated with O2 (black curve) and N2 (red
42
43 7 curve), and their subtraction (blue curve), peak current density of the subtracted CV in function of
44 8 the square root of scan sate, and peak potential of the subtracted CV in function of the logarithm of
45 9 scan rate. (a) Fe-N-C; (b) Co-N-C; (c) Cu-N-C.
46
10
47
48
49 11
50
51 12 3.3.2. Second method
52
53 13 The second analytical method consisted in the determination of the differential CV by subtracting
54
55 14 the second CV cycle from the first CV cycle recorded in the O2-saturated electrolyte. Even though a
56
57
58 15 complete elimination is never possible, this would reduce as much as possible the contribution of
59
60 22
ACS Paragon Plus Environment
Page 23 of 75 The Journal of Physical Chemistry

1
2
3 1 the reduction of the diffused O2 to the peak current density.22 Figure 6a-b-c shows the differential
4
5 2 CV peaks obtained for Fe-N-C, Co-N-C, and Cu-N-C catalysts.
6
7 3 For Fe-N-C, almost symmetrical full differential CV peaks were obtained (see Figure 6a) in all the ν
8
9
10 4 range considered (10–350 mV s−1), with Ip values between 0.7 and 13 mA cm−2. Consequently, the
11
12 5 calculation of the full width at half maximum (FWHM) was possible for all of the peaks recorded at
13
14 6 different ν. In addition, the FWHM values, ∆E1/2, were plotted vs. the ν1/2, and a linear trend was
15
16 7 obtained. This αc value close to unity is similar to the value found in RDE experiment in Section 3.2
17
18
8 in the low overpotential region (see Table 1). The third plot in Figure 6a represents the variation of
19
20
21 9 αc (calculated as indicated by Laviron in the case of diffusionless electrochemical systems:65 αc =
22
23 10 62.5 mV / ∆E1/2 (mV)) with the peak potential Ep, which in turns varies with ν. The results show
24
25 11 that there is a decrease of αc with decreasing Ep, suggesting a potential-dependent change in the
26
27 12 reaction mechanism, as found in Section 3.3.1, as well as in Section 3.2. The αc values range
28
29
30 13 approximately between 0.6 and 0.2. In the RDE experiments (Section 3.2, Table 1) the αc value in
31
32 14 the high η region (between 0.7 and 0.6 V vs RHE) was 0.42, which very well matches with the αc
33
34 15 values in this potential range in Figure 6a.
35
36 16 For Co-N-C the situation is considerably different. Observing the differential CV in Figure 6b, a
37
38
17 peak trend is only slightly evident for all the scan rates considered, and the peaks show very low Ip
39
40
41 18 values compared to both Fe-N-C and Cu-N-C. Moreover, the peaks do not follow the trend of
42
43 19 increasing Ip and ∆E1/2 with the increase of the scan rate. On the contrary, for ν values between 50
44
45 20 and 200 mV s−1, Ip seems to decrease with the increase of ν. This result appears to be a further
46
47 21 confirmation of the finding of Section 3.3.1, about the fact that the reduction of adsorbed O2 is
48
49
50 22 much less important on Co-N-C catalyst, being the diffusion contribution dominant, in comparison
51
52 23 with Fe-N-C and Cu-N-C. As a consequence, for Co-N-C it was not possible to find a linear
53
54 24 correlation in the plot of ∆E1/2 vs. ν1/2, and consequently neither to calculate the αc extrapolated at
55
56 25 scan rate = 0, nor to find a trend of variation for the αc with Ep was possible for this catalyst (see
57
58
26 Figure 6b-c). In Figure 6c, the calculated αc values for Co-N-C oscillate between 1.04 and 0.73.
59
60 23
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 24 of 75

1
2
3 1 However, the reliability of these values is too poor, due to the very low contribution of the adsorbed
4
5 2 O2 reduction on this catalyst. Nevertheless, comparing these results with the αc value obtained for
6
7 3 Co-N-C in RDE experiment (αc = 1.01, see Table 1) we again obtain a good agreement.
8
9
10 4 For Cu-N-C the situation is more similar to Fe-N-C, with the difference that the ORR starts at
11
12 5 considerably lower potentials. Almost symmetrical differential CV peaks are evident in Figure 6c,
13
14 6 with Ip values between 0.6 and 12 mA cm−2. However, at high scan rates (150 and 200 mV s−1) the
15
16 7 peaks appear cut, because their Ep is too low and close to the lower limit of the CV potential
17
18
8 window. Consequently, for the calculation of the ∆E1/2 we considered only the scan rate values
19
20
21 9 between 10 and 100 mV s−1. The ∆E1/2 vs. ν1/2 plot in Figure 6b shows a linear trend, which enables
22
23 10 the calculation of the αc value extrapolated at ν = 0. This value is equal to 0.48, which is almost
24
25 11 one-half of the value found for Fe-N-C, and perfectly matches the αc value found for Cu-N-C
26
27 12 catalyst in RDE polarization curve in Section 3.2 (Table 1). Concerning the αc vs. Ep plot in Figure
28
29
30 13 6c, similarly to Fe-N-C, there is a decrease of αc with decreasing Ep, varying approximately
31
32 14 between 0.40 and 0.25.
33
34 15
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 24
ACS Paragon Plus Environment
Page 25 of 75 The Journal of Physical Chemistry

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 1
34 2
35 3 Figure 6. Differential CV recorded in O2-saturated 0.5 M H2SO4 at different scan rates; FWHM of
36
4 the differential CV peaks vs. the square root of the scan rate; cathodic transfer coefficient for the
37
38 5 reduction of adsorbed O2 calculated from the FWHM of the differential CV peaks widths for (a) Fe-
39 6 N-C, (b) Co-N-C, and (c) Cu-N-C catalysts.
40 7
41
42
43
8 As in Section 3.3.1, also these results indicate that Fe-N-C is the catalyst that better performs
44
45 9 towards adsorbed oxygen reduction, approaching more the top of the asymmetric “volcano plot”
46
47 10 described in the literature regarding O2 binding energy on its active sites, and metal-based active
48
49 38,66
11 sites Eredox. The results obtained for Co-N-C are a further evidence of the poor O2 adsorption
50
51 12 affinity on the surface of this catalyst. Also for Cu-N-C the results confirm the finding of the
52
53
54 13 previous section.
55
56 14
57
58 15 3.4. Reaction order for O2
59
60 25
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 26 of 75

1
2
3 1 For the determination of the reaction order for O2, staircase voltammetry experiments were
4
5 2 performed in 0.5 M H2SO4 solution saturated with O2–N2 mixtures in four different proportions, to
6
7 3 have O2 at different partial pressures. Figure 7 shows the experimental results, where the O2 partial
8
9
10 4 pressures used in each experiment are also reported.
11
12 5 For all the catalysts, the limiting current density decreases as the O2 partial pressure in the gas flow
13
14 6 bubbling into the electrolyte solution decreases. A similar trend of results was obtained by Paulus et
15
16 7 al. for a Pt/C catalyst.42
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 26
ACS Paragon Plus Environment
Page 27 of 75 The Journal of Physical Chemistry

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
41
42
43
44 1
45 2
46 3 Figure 7. Tafel plot in the kinetic zone of the polarization curve recorded in 0.5 M H2SO4 at 900
47
48 4 rpm with different O2 partial pressures at 25°C, and double-logarithmic plots of mass transport
49 5 corrected current density at different potentials in function of O2 concentration for (a) Fe-N-C, (b)
50 6 Co-N-C, and (c) Cu-N-C catalysts.
51 7
52
53
54 8 The O2 solubility in moderately concentrated H2SO4 aqueous solutions follows the Henry’s law,67,68
55
56 9 thus it decreases linearly with the O2 partial pressure.
57
58 10 The reaction order for oxygen at constant electrode potential can be defined by Equation (5):
59
60 27
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 28 of 75

1
2
0$,1 |2 |
3 1  = / 4 (5)
4 0$,13 (, ,
5
6 2 Where: RO2 is the reaction order for oxygen, im is the mass-transport corrected current density, C is
7
8 3 the O2 concentration in the electrolyte, E is the electrode potential, T is the temperature.
9
10
11 4 According to Equation (5), the reaction order can be calculated from the slope of a double-
12
13 5 logarithmic plot of im vs. C*, where C* is defined as the O2 concentration normalized to the O2
14
15 6 saturation concentration in the electrolyte for pure O2 at 1 bar, C0 (being C0 = 1.05 mM in 0.5 M
16
17 7 H2SO4 aqueous solution). As shown in the double logarithmic plots in Figure 7, and summarized in
18
19
20
8 Table 3, for all the three catalysts RO2 in the kinetic control potential range is not constant. In
21
22 9 particular, for potentials closer to the onset potential (Eon), RO2 is about 0.5. Then, as η increases,
23
24 10 RO2 increases as well. Considering the most active catalysts, that is Fe-N-C, the reaction order is
25
26 11 ~0.5 close to Eon, and increases up to ~0.85 at 0.7 V. For Co-N-C RO2 is also ~0.5 close to Eon, but
27
28 12 in the same potential range its increase is more marked, reaching ~1 at 0.7 V. Cu-N-C catalyst
29
30
31 13 behaves similarly to Fe-N-C, in a 100 mV more negative potential range. Some studies about the
32
33 14 reaction order for O2 have been done for Pt-based catalysts in RDE and PEMFC in the literature,
34
35 15 some of them showing values close to unity in all the potential ranges considered,19,42,69 some others
36
37 16 showing values lower than unity, (between 0.7 and 0.85).70–72
38
39
40 17
41
42 18 Table 3. Reaction orders for O2 calculated at different potentials in the kinetic control zone for Fe-
43 19 N-C, Co-N-C and Cu-N-C catalysts.
44 Catalyst
45
46 E vs RHE [V] Fe-N-C Co-N-C Cu-N-C
47 0.80 0.54 0.48 -
48 0.77 0.65 0.76 -
49
50 0.75 0.69 0.84 -
51 0.72 0.76 0.92 -
52 0.70 0.84 0.98 0.46
53
54 0.67 - - 0.58
55 0.65 - - 0.66
56 0.62 - - 0.75
57
58
0.60 - - 0.80
59 20
60 28
ACS Paragon Plus Environment
Page 29 of 75 The Journal of Physical Chemistry

1
2
3 1 Unfortunately, in the literature only few studies report the reaction order for O2 for similar types of
4
5 2 Me-N-C catalysts in acidic conditions. Gojkovic et al. reported a reaction order for O2 equal to 1,
6
7 3 but these authors measured it using a different method, and at lower potentials compared to our
8
9
10 4 measurements.62 Chlistunoff found a reaction order of 1 at all the considered potentials for a
11
12 5 pyrolyzed Fe-Polyaniline catalyst which activity was tested in oxygen and air-saturated acidic
13
14 6 electrolyte.22
15
16 7 The RO2 values that we obtained, showing an increase from ≈ 0.5 at potentials close to Eon, to values
17
18
8 approaching 1 at higher η, in our opinion could be an indirect confirmation of the redox-mediated
19
20
21 9 formation of the active sites occurring on our catalysts with the increase of η. This reasoning could
22
23 10 be envisioned as an “incomplete catalyst utilization” occurring at low η, where the active sites are
24
25 11 going to be formed due to the redox-mediated mechanism, and the active sites coverage is
26
27 12 decreasing from θ ≈ 1 to θ ≈ 0. For the above-mentioned Pt catalysts showing RO2 < 1, an analogous
28
29
30 13 phenomenon could take place, with the surface coverage with oxidized species, which causes the Pt
31
32 14 surface to not be fully available to carry out the ORR.
33
34 15 3.5. Reaction order for H+
35
36 16 H+ ion, in addition to O2, is the other reactant of the ORR in acidic conditions. For the determination
37
38
17 of the reaction order for H+, ORR activity experiments have to be performed at different H+
39
40
41 18 concentrations, that is, at different electrolyte solution pH. This is in contrast to the fact that the
42
43 19 determination of the order of an electrochemical reaction for one of its reactants has to be
44
45 20 performed at a constant electrode potential. In fact, if we use the RHE as a potential reference scale,
46
47 21 we have to consider that this reference varies with the electrolyte solution pH of about 60 mV per
48
49
50 22 pH unit.
51
52 23 However, the reaction order for H+ (RH+) can be approximately expressed by Equation (6):22
53
54 24
55
56 ;$,1< D F I@
57 25 R 78 ≅ :;$,1= ? + /).*GH
E
4 :I$,1=
A>B
? (6)
>8 @A>B >8
58
59
60 29
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 30 of 75

1
2
3 1
4
5 2 where: i is the current density, CH+ is the molar concentration of H+ ion, ERHE is the potential
6
7 3 measured vs. RHE, αc is the cathodic transfer coefficient, R is the gas constant and T is the absolute
8
9
10 4 temperature. The SV with three different H2SO4 concentrations (0.10–0.25–0.50 M, which
11
12 5 respectively pH values are 1–0.6–0.3) were measured and the results are shown in Figure 8.
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 6
57 7
58
59
60 30
ACS Paragon Plus Environment
Page 31 of 75 The Journal of Physical Chemistry

1
2
3 1 Figure 8. SV recorded at 900 rpm in O2-saturated H2SO4 solutions at different pH, and the
4 2 respective Tafel plots after mass-transport correction for (a) Fe-N-C, (b) Co-N-C, and (c) Cu-N-C
5
3 catalysts.
6
7 4
8
9 5 For Fe-N-C and Co-N-C, in the low η range (between 0.8 and 0.7 V vs. RHE) the Tafel plot is
10
11 6 almost linear (see Figure 8a-b) and the slope is equal to ~ 60 mV dec−1. In addition, the plots at
12
13
7 different pH values are almost completely overlapping in this potential range. Thus, at a certain
14
15
16 8 potential in this range, we have:
17
;$,1<
18 9 :;$,1= ? ≅ 0 (7)
19 >8 @A>B
20
21 10
22
23 D F J J
24 11 /).*GH
E
4 = HKL.$ M$,-. ≅ NO PQ I.RST (8)
25
26
27 12
28
I@
29 13 :I$,1=
A>B
? ≅ 60 mV dec J (9)
30 >8
31
32 14 Substituting expressions (7), (8), and (9) into Equation (6), the reaction order for H+ is
33
34 15 approximately 1. In the same way, considering approximately no variations of the plots with pH, the
35
36
37
16 reaction order for H+ suffers a decrease as the cathodic overpotential increase, due to the changes in
38
39 17 the Tafel slope (see Table 1). As for the reaction order for O2, also in this case the literature studies
40
41 18 are scarce. Again, Chlistunoff performed a similar study for a pyrolyzed Fe-Polyaniline catalyst,
42
43 19 which exhibited a similar behavior.22 For Cu-N-C, the same considerations can be made. However,
44
45 20 this catalyst exhibits a constant Tafel slope of about 120 mV dec−1 in the potential range between
46
47
48 21 0.60 and 0.40 V vs RHE (see Figure 8c). Thus, in this potential range, the reaction order for H+
49
50 22 should be lower than unity.
51
52 23 Regarding the determination of RH+ in these experimental conditions, a consideration deserves to be
53
54 24 made: the variation of H2SO4 concentration causes a non-negligible variation of O2 solubility.68,73
55
56
57
25 This affects the conditions of validity of determination of the reaction order for H+, that is, the
58
59 26 concentration of one of the other reactants taking part in the reaction (in this case O2) is not
60 31
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 32 of 75

1
2
3 1 constant. In fact, for a rigorous determination of the reaction order for H+, only the H+ concentration
4
5 2 must vary, while all the other parameters (i.e., temperature, pressure, potential, O2 concentration)
6
7 3 must be constant. Even in the small pH range considered in our experiments (from 1.0 to 0.3, that
8
9
10 4 is, for an H2SO4 concentration varying from 0.1 to 0.5 M) the variation of O2 concentration in the
11
12 5 electrolyte solution due to the increase of H2SO4 concentration is already considerable. In fact,
13
14 6 looking at the polarization curves in Figure 8 in the high η region, where the contribution of the
15
16 7 mass transport in higher, the diffusion limiting currents values recorded at lower pH are lower. This
17
18
8 is due to the decrease of O2 solubility with decreasing pH. A correction for O2 solubility in
19
20
21 9 solutions with different pH should be operated to eliminate these discrepancies. For this reason, we
22
23 10 limited our experiments to this small pH range, assuming valid our calculation of the reaction order
24
25 11 for H+ only in the low η region (kinetic limitation).
26
27 12
28
29
30 13 3.6. Activation Energy calculation for ORR
31
32 14 The activation energy (Ea) was evaluated at different fixed potentials using the Arrhenius Equation
33
34 15 in the temperature range 10–60 °C. In performing the RDE experiments at different temperatures,
35
36 16 two different effects must be taken into account. The first is the ORR kinetic increase with
37
38
17 increasing temperature, and the second is the decrease in O2 concentration with increasing
39
40
41 18 temperature. In particular, the O2 concentration in the electrolyte solution varies from 1.38 mM at
42
43 19 20 °C to 0.61 mM at 60 °C.74 Thus, to take into account this effect, the current density measured at
44
45 20 different temperatures has been corrected according to the Equation (10):23
46
47 21
48
49 O.ab
JP_
50 22 i∗P = iP ∙ exp : = ? (10)
51 `
52
53 23
54
55 24 Where: im is the measured mass-transport corrected current density, CO2 is the actual O2
56
57
58
25 concentration in the liquid electrolyte at the temperature at which the measurement was done, and
59
60 32
ACS Paragon Plus Environment
Page 33 of 75 The Journal of Physical Chemistry

1
2
3 1 im* is the corresponding current density for an oxygen concentration of 1 mM. Then, the Arrhenius
4
5 2 plot (logarithm of current density vs. the inverse of the absolute temperature) can be drawn, and Ea
6
7 3 can be calculated from the linearization of the Arrhenius law:
8
9
10 4
11
@
12 5 i∗P cTe = i∗P cTfe ∙ exp /).*GH
g
4 (11)
13
14
15 6
16
17 7 where T is the temperature at which the measure was done, and im*(T∞) is a constant value. The
18
19 8 slope of the Arrhenius plots is equal to –Ea / (2.3 R).
20
21 9 Figure 9 shows the Tafel plots at different temperatures with the above mentioned corrections, and
22
23
24 10 the respective Arrhenius plots, for Fe-N-C, Co-N-C and Cu-N-C catalysts.
25
26 11 The Arrhenius plots were calculated at different potentials in the regions under kinetic and mixed
27
28 12 kinetic-diffusion control for the three different catalysts. Table 4 summarizes the corresponding Ea
29
30 13 values, which are comparable with the values reported in the literature for ORR for Pt-based
31
32
14 catalysts, which vary between 27 and 11 kJ mol−1 at potentials in the range 0.70–0.95 V.18,20,21,23,42
33
34
35 15 Unfortunately, as for the reaction orders, also for Ea there is a lack in the literature for this type of
36
37 16 Me-N-C electrocatalysts. Jaouen et al.23 obtained values around 9 kJ mol−1 at 0.9 V vs RHE for a
38
39 17 series of pyrolyzed Fe-N-C catalysts synthesized using carbon black, iron acetate, and
40
41 18 phenanthroline.
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 33
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 34 of 75

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
41
42
43
1
44
45
2 Figure 9. Effect of the temperature on ORR activity in RDE experiments in O2-saturated 0.5 M
46 3 H2SO4 at 900 rpm. Tafel plot after correction for mass-transport and O2 solubility at different
47 4 temperatures, and the respective Arrhenius plot at different potentials for (a) Fe-N-C, (b) Co-N-C,
48 5 and (c) Cu-N-C catalysts.
49
50 6
51
52 7 For Fe-N-C and Co-N-C catalysts a decrease of Ea with the increase of η is observed. In particular,
53
54 8 in the low η region, where the Tafel slope is of about 60 mV dec−1, the Ea has values close to 20 kJ
55
56 9 mol−1. Then, moving to higher η, where the Tafel slope has almost a double value, the Ea values
57
58
59
10 tend to decrease, approaching 14–16 kJ mol−1 at 0.65 V vs. RHE and even lower values of 12–13 kJ
60 34
ACS Paragon Plus Environment
Page 35 of 75 The Journal of Physical Chemistry

1
2
3 1 mol−1 at 0.6 V vs. RHE. A similar trend of decreasing Ea with increasing overpotential
4
5 2 (simultaneously with the changing in the Tafel slope) was found in the literature for Pt/C, Pt-Ni/C
6
7 3 and Pt-Co/C catalysts in acidic conditions.18 Unlike the two previous catalysts, for Cu-N-C the
8
9
10 4 activation energy was calculated in the potential range between 0.60 and 0.45 V vs. RHE, having
11
12 5 this catalyst a 150 mV lower Eon. The Ea of Cu-N-C, similarly to the Tafel slope (see Figure 3b),
13
14 6 does not significantly change with η. In fact, for Cu-N-C catalysts, Ea oscillates between 21.3 and
15
16 7 23.6 kJ mol−1 at all the potential considered.
17
18
8 Thus, Fe-N-C and Co-N-C catalysts show the expected trend of Ea decrease with the increase of η.
19
20
21 9 On the other hand, the Cu-N-C catalyst shows an almost constant Ea value with η. This anomalous
22
23 10 behavior is not trivial to be interpreted at this point, and a deeper study, which is out of the scope of
24
25 11 this work, would be necessary to better understand the reason. A possible reason for this behavior is
26
27 12 a potential dependent variation of the pre-exponential factor of the Arrhenius equation (e.g.
28
29
30 13 including the entropic term), which could compensate the corresponding decrease of the apparent
31
32 14 activation energy with η. A similar behavior of constant activation enthalpies with the increase of η
33
34 15 was reported by Paulus et al. for Pt and Pt-alloys catalysts.21
35
36 16 Paulus et al.42 determined the activation energy for a Pt catalyst supported on carbon in the low
37
38
17 overpotential region using the Arrhenius method in both H2SO4 and HClO4 electrolytes, obtaining
39
40
41 18 values of 26–28 kJ mol−1. Similar values were obtained for low index Pt single crystal surfaces and
42
43 19 polycrystalline Pt surfaces.15,16,18,21 Neyerlin et al. reported a series of activation energy values for
44
45 20 Pt/C catalysts measured in a fuel cell. These values are in the range between 42 and 96 kJ mol−1.19
46
47 21 The reason of these significantly higher Ea in comparison with the values found in three-electrode
48
49
50 22 cell configuration with a liquid electrolyte, are related to the higher complexity of the fuel cell
51
52 23 system, where phenomena like water flooding effects in the cathode catalyst layer may occur.42,75
53
54 24 However, it must be considered that the calculation of Ea at the same η (as done in this work), is
55
56 25 only an estimation of the Ea value. Thus, Ea is an “apparent” activation energy value.21 In the case
57
58
26 of Pt in fact, the im* in Equations (10) and (11) is strongly dependent on the quantity of oxides on
59
60 35
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 36 of 75

1
2
3 1 the Pt surface which, in turn, is temperature-dependent, even though the oxide coverage is
4
5 2 considered as a pre-exponential term in Equation (11).21
6
7 3
8
9
10 4 Table 4. Activation energies calculated at different potentials for Fe-N-C, Co-N-C and Cu-N-C
11 5 catalysts (from Figure 9).
12 Activation Energy [kJ mol–1]
13
14
E vs RHE [V] Fe-N-C Co-N-C Cu-N-C
15 0.75 20.91 22.82 –
16 0.70 19.57 18.50 –
17
0.65 16.34 14.56 –
18
19 0.60 13.93 12.67 21.71
20 0.55 – – 23.61
21
0.50 – – 23.60
22
23 0.45 – – 21.33
24 6
25
26 7 Trying to make an analogy between Pt-based catalysts and our most active catalysts (i.e., Fe-N-C
27
28
8 and Co-N-C), we can speculate that the reaction mechanism could be the same: the first step is the
29
30
31 9 adsorption of O2 on the catalyst active site (O2  O2*), followed by the first electron transfer,
32
33 10 which is the rate determining step: O2*  O2*−. Surface coverage effects have been proposed as the
34
35 11 reason for changes in ORR kinetic behavior on Pt surfaces, where changes in the surface coverage
36
37 12 of adsorbed oxygen-containing groups occur.76–78 These oxygen-containing groups, such as −OH
38
39
40 13 species, can originate from both H2O and O2,78 and can control the availability of the O2 adsorption
41
42 14 sites.21 A clear overlap between the onset potential of ORR and the removal of these oxygen-
43
44 15 containing species from the Pt surface is observed,78 considering the CV of Pt in de-aerated
45
46 16 solutions and the ORR experiments in RDE. In this region, the Tafel slope change also occurs.
47
48
17 Trying to make a parallelism between Pt-based catalysts and our NNM catalysts (e.g. Fe-N-C), the
49
50
51 18 presence of a redox peak in the CV, i.e., related to the Fe2+/Fe3+ couple,22,79 could be compared to
52
53 19 the Pt-oxides reduction peak in the CV of Pt surface. The onset of ORR on Pt surface is associated
54
55 20 with the appearance of the Pt-oxides reduction peak in the CV recorded in the absence of O2, as
56
57 21 discussed in the literature.78 Thus, the onset of the ORR on our NNM catalysts could be associated
58
59
60 36
ACS Paragon Plus Environment
Page 37 of 75 The Journal of Physical Chemistry

1
2
3 1 with the potential at which the redox peak starts to appear.22,79,80 However, in our catalysts, this
4
5 2 redox peak is not so clearly evident (see Figure 2a) as in other NNM catalyst reported in the
6
7 3 literature, or as the Pt-oxides reduction peak in the CV of Pt-based catalysts. In particular, in our
8
9
10 4 catalysts, the presence of these red-ox peaks related to transition metal-based red-ox centers, could
11
12 5 be “hidden” by the much high capacitive currents (e.g., in comparison with Vulcan which is the
13
14 6 usual carbon support for Pt-based catalysts),79 which are due to the high specific surface area and
15
16 7 surface functionalization (pseudo-capacitive effects due to the presence of oxygenated functional
17
18
8 groups like quinone/hydroquinone on the surface of our catalysts).50,63,80,81 The presence of a broad
19
20
21 9 peak in the potential range between 0.8 and 0.4 V vs RHE is detected in our catalysts (see Figure
22
23 10 2a and Figure 5). We could state this is a way to justify the analogies between the behavior of these
24
25 11 NNM catalysts and the Pt-based ones.
26
27 12
28
29
30 13 3.7. Koutecky-Levich analysis.
31
32 14 In order to have some indication about the ORR reaction pathway, i.e., if O2 is reduced directly to
33
34 15 H2O via a direct 4 e− mechanism, or via a two-steps 2 e− + 2 e− mechanism with the formation of a
35
36 16 peroxide intermediate, or only partially to H2O2 via a 2 e− mechanism, a series of electrochemical
37
38
17 tests were performed. The first test was the Koutecky–Levich (K–L) analysis. Figure 10 shows the
39
40
41 18 results of the LSV recorded in the O2 saturated electrolyte at different RDE rotation speed for Fe-N-
42
43 19 C, Co-N-C, and Cu-N-C catalysts. For all the three catalysts the limiting current densities are lower
44
45 20 than those predicted by the Levich equation.26 The variation of the current density with the rotation
46
47 21 speed is a direct means to investigate if the reaction is under mass-transport control of reactants
48
49
50 22 diffusing from the bulk solution or under kinetic control.26 As expected, for all of the catalysts there
51
52 23 is a noticeable increase in the diffusion-limited current density with the RDE speed. However, no
53
54 24 one of the catalysts exhibits a perfect “plateau” in the diffusion-limited region at high η, and this is
55
56 25 more evident for Co-N-C catalyst. A linear trend with an almost parallel slope at different
57
58
59
60 37
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 38 of 75

1
2
3 1 potentials is observed in the K–L plots for all of the catalysts, enabling the calculation of the
4
5 2 number of electrons involved in the ORR, as summarized in Table 5.
6
7 3 Another informative aspect of the K–L plot is the intercept of the straight line, which should in
8
9
10 4 principle be 0 if the current is fully controlled by O2 diffusion in solution. This is not the case for of
11
12 5 our Me-N-C catalysts, which all exhibit a non-zero intercept (see K–L plots in Figure 10),
13
14 6 suggesting the presence of an additional transport limitation, which could be related to the presence
15
16 7 of a Nafion film between the solution and the electrode surface. This phenomenon has been
17
18
8 reported in the literature for both Pt-based42 and Pt-free catalysts.82 The presence of this Nafion film
19
20
21 9 in our experiments is likely, since a relatively high amount of Nafion was used to prepare the
22
23 10 catalyst ink (NCR = 0.2). The related additional transport resistance value can be calculated from
24
25 11 the intercept of the K–L plots. However, these values are considerably lower than transport
26
27 12 resistance caused by the diffusion in the solution, and can be neglected in the calculation of the
28
29
30 13 mass transport corrected current densities (the calculation is not reported here).
31
32 14 Analyzing the results, it must be considered that the K–L theory was developed for a smooth and
33
34 15 thin electrode surface.79 However, when the catalyst loading on the electrode is on the order of
35
36 16 hundred µg cm−2 (as typically used in the literature for NNM catalysts, and also in this work), we
37
38
17 cannot affirm that the catalyst film deposited on the electrode is so thin to be under the hypothesis
39
40
41 18 of the K–L theory.22,83 Among others, a recent publication points out how the K–L method is not
42
43 19 suitable to determine number of electrons involved in ORR. In facts, ORR is not a single-step one-
44
45 20 way reaction, and it is not always first order with respect to oxygen concentration (as moreover
46
47 21 found in our results, see Section 3.4).84 Thus, the ORR does not fulfill the assumptions of the K–L
48
49
50 22 method. In the same work of it is demonstrated that the number of electrons calculated using K–L
51
52 23 method is always significantly different from the one calculated using RRDE.84
53
54 24
55
56
57
58
59
60 38
ACS Paragon Plus Environment
Page 39 of 75 The Journal of Physical Chemistry

1
2 1
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
41
42
43
44 2
45 3
46 4 Figure 10. LSV recorded at 5 mV s−1 scan rate and different rotation speeds in O2-saturated 0.5 M
47 5 H2SO4 after background capacitive current subtraction, and the corresponding K−L plots at
48
6 different potentials in the diffusion-limited region for (a) Fe-N-C, (b) Co-N-C, and (c) Cu-N-C
49
50 7 catalysts.
51 8
52
53 9 Considering the results in Table 5 as a qualitative comparison among our catalysts, Fe-N-C is the
54
55 10 catalyst with the better performance in terms of selectivity towards a complete 4 e− ORR. Co-N-C
56
57 11 catalyst has a considerably lower selectivity, despite having an Eon very close to the Fe-N-C one. A
58
59
60 39
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 40 of 75

1
2
3 1 higher number of electrons was obtained for Cu-N-C compared to Co-N-C, but the former has a
4
5 2 remarkably lower Eon than the latter.
6
7 3
8
9
10 4 Table 5. Total number of electrons involved in the ORR for Fe-N-C, Co-N-C and Cu-N-C catalysts
11 5 resulting from the K-L analysis in Figure 10.
12 6
13 Catalyst # of e−
14 0.05 V vs. RHE 0.1 V vs. RHE 0.2 V vs. RHE
15 Fe-N-C 3.73 3.70 3.64
16
17 Co-N-C 2.71 2.75 2.87
18 Cu-N-C 3.09 3.05 2.97
19 7
20
21 8
22
23
24 9 3.8. Hydrogen peroxide reduction.
25
26 10 To deeper investigate which is the predominant pathway for ORR, the catalysts’ activity toward the
27
28 11 H2O2 reduction reaction (HPRR) in the absence of O2 was assessed. To this purpose, LSV were
29
30 12 measured in N2-saturated solution after the addition of H2O2 in a concentration of 1 mM, which is
31
32
13 very close to the concentration of O2 in saturated 0.5 M H2SO4.60,73 Moreover, this concentration is
33
34
35 14 about the maximum H2O2 concentration that can be found during ORR experiments in RDE.85 As
36
37 15 shown in Figure 11a, in spite of almost the same concentration of H2O2 and O2 in the solution, the
38
39 16 currents due to ORR are almost 3 times higher for Fe-N-C and Cu-N-C and almost 2 times higher
40
41 17 for Co-N-C in comparison to those due to HPRR. Moreover, the increase of HPRR current with
42
43
44 18 cathodic potential is almost linear, and the maximum reduction current is smaller than the diffusion
45
46 19 limited currents expected for 1 mM H2O2 from the Levich equation. Similar results have been
47
48 20 reported in the literature for heat-treated Fe-N/C catalysts.22,26,82 This suggests that HPRR is under
49
50 21 kinetic control within all the scanned potential range. The kinetics of HPRR is sluggish compared to
51
52
22 ORR, since it is never fast enough to completely reduce all the available flux of H2O2, even at high
53
54
55 23 η. It must be considered that part of the reduction current measured during the experiment carried
56
57 24 out in the presence of 1 mM H2O2 could also be originated from H2O2 chemical
58
59
60 40
ACS Paragon Plus Environment
Page 41 of 75 The Journal of Physical Chemistry

1
2
3 1 disproportionation.26,86 In facts, the H2O2 generated during ORR on a catalytic site can be reduced
4
5 2 to H2O on the same site or on another site either through electroreduction or through
6
7 3 disproportionation.26 In the latter case, the O2 originating from the disproportionation reaction could
8
9
10 4 be electrochemically reduced to H2O2 once again. If the latter case occurs, it will result in an
11
12 5 apparent 4 electrons reduction of O2 even if the electroreduction of H2O2 to H2O never occurs, or it
13
14 6 occurs in minimum part, as demonstrated for our catalysts (see results in Fig. 11a). In the past,
15
16 7 Jaouen and Dodelet investigated the relative importance of H2O2 disproportionation on ORR
17
18
8 pathway and they found it to be minor.26 On the other hand, the work of Masa et al. shows that the
19
20
21 9 H2O2 disproportionation occurs fast on a Fe/N/C catalyst in alkaline conditions, and cannot be
22
23 10 neglected.87 In this work the measurement of the rate of chemical disproportionation of H2O2 was
24
25 11 not performed because, even if occurring, it will result in an apparent 4 electrons ORR pathway, as
26
27 12 it can be indirectly inferred from the results in Fig 11 a-b.
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 41
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 42 of 75

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 1
38 2 Figure 11. (a) LSV at 1600 rpm in N2-saturated 0.5 M H2SO4 + 1 mM H2O2 and in O2-saturated 0.5
39
3 M H2SO4 for Fe-N-C, Co-N-C and Cu-N-C catalysts. (b) H2O2 molar generation calculated from
40
41
4 RRDE test performed in 0.5 M H2SO4 for Fe-N-C, Co-N-C, and Cu-N-C and in 0.1 M HClO4 for
42 5 the Pt/C catalyst.
43 6
44
45 7 3.9. RRDE test.
46
47 8 As a further means to investigate the ORR pathway, RRDE test was performed. Equation (12) was
48
49
9 used to calculate the percentage of H2O2 produced during the ORR:
50
51
52 10
53
)kl /n
54 11 % H) O) = 100 ∙ q (12)
55 ko p/ rl 4
56
57 12
58
59
60 42
ACS Paragon Plus Environment
Page 43 of 75 The Journal of Physical Chemistry

1
2
3 1 where Id is the current at the disk, Ir is the current at the ring and N is the ring collection efficiency
4
5 2 (44%).27 Figure 11b shows the results, evidencing that Co-N-C is the catalyst producing the highest
6
7 3 H2O2 amount, which reaches a maximum value of about 15 % at 0.4 V vs. RHE at 900 rpm. Cu-N-
8
9
10 4 C, leads to the formation of a considerably lower amount of H2O2, with a maximum of 3.5 molar %
11
12 5 at 0.5 V vs RHE at 900 rpm, notwithstanding the lower electroactivity in terms of onset and half-
13
14 6 wave potentials. Fe-N-C is the catalyst that shows the lowest H2O2 production (between 0.5 and 2
15
16 7 % at 900 rpm), confirming its better performance not only in terms of activity but also in terms of
17
18
8 selectivity. Similar H2O2 production values were measured by other groups for similar Fe-N/C
19
20
21 9 pyrolyzed catalysts in acidic conditions.22,26,79 A commercial Pt/C catalyst was also analyzed for
22
23 10 comparison. For this catalyst, the H2O2 generation at lower η is considerably lower than for the Me-
24
25 11 N-C catalysts, especially in the anodic potential scan direction. In the low η region (approximately
26
27 12 between 0.85 and 0.60 V vs. RHE) the H2O2 generation is higher in the cathodic scan direction, due
28
29
30 13 to the presence of oxides on the Pt surface at these potentials, which could partially hinder the
31
32 14 complete 4 e− O2 reduction. In addition, a significant increase in the ring current for Pt/C catalyst
33
34 15 was detected approaching the hydrogen underpotential deposition region. This behavior is typical of
35
36 16 Pt in acidic conditions.42,46,79
37
38
17 The H2O2 amount detected by RRDE technique are considerably lower than the values calculated
39
40
41 18 by the K–L analysis in Section 3.7. These discrepancies are a further confirmation that the
42
43 19 conditions of validity of the K–L model are not verified in these ORR measurements, as pointed out
44
45 20 in Section 3.7.
46
47 21 From a rigorous point of view, the RRDE experiments must be carried out at different rotation
48
49
50 22 speeds, to investigate the effect of the rotation speed on the H2O2 production.84,88 Thus, the results
51
52 23 in Figure 11b should be considered more from a qualitative point of view, in terms of comparison
53
54 24 between the different behavior induced by the different transition metals (Fe, Co, Cu) in the three
55
56 25 different catalysts with respect to the ORR selectivity, rather than as an extremely rigorous
57
58
26 quantitative determination of H2O2 production. The results in Figure 11b show an evident
59
60 43
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 44 of 75

1
2
3 1 qualitative behavior in agreement with the proposed theory of the “asymmetric volcano plot” and
4
5 2 the redox-mediated formation of the active sites, with other experimental and theoretical studies
6
7 3 reported in the literature, and with the other results shown in this work.
8
9
10 4 By concluding, from the results of the tests described in Sections 3.7, 3.8, and 3.9, it can be deduced
11
12 5 that all the three Me-N-C catalysts are able to reduce O2 (to H2O and/or H2O2), but much less to
13
14 6 reduce H2O2 to H2O, as discussed in our previous work for Fe-N-C catalyst under acidic
15
16 7 conditions.60 However, on Co-N-C catalyst, the partial 2e− O2 reduction to H2O2 proceeds in a
17
18
8 greater extent than on Fe-N-C and Cu-N-C. This behavior can be related to the findings of Section
19
20
21 9 3.2. In particular, a weaker adsorption of O2 on Co-based active sites can explain both the lower
22
23 10 peak current densities observed in the tests of Section 3.2, as well as a higher detection of H2O2 in
24
25 11 RRDE experiments. O2 is adsorbed weakly on Co-based sites, resulting in an easier desorption after
26
27 12 the partial 2 e- reduction, before the O–O bond breaking occurs.25,89,90
28
29
30 13
31
32 14 3.10. Resume of experimental evidence and discussion about the ORR kinetics.
33
34 15 Analyzing and trying to interpret the results of Sections 3.3.1 and 3.3.2, we can consider some
35
36 16 experimental evidence reported in other literature studies about biological molecules where
37
38
17 adsorption, transport and reduction of O2 on Fe-, Co-, and Cu-macrocyclic units is involved.
39
40
41 18 As a background and confirmation of our experimental results, the work of Chen et al.25 about
42
43 19 experimental and theoretical (DFT) studies on Fe-phthalocyanine and Co-phthalocyanine based
44
45 20 catalysts for ORR, shows that the O2 adsorption energy on the metallized center is related to the
46
47 21 catalytic activity. In particular, the Fe-based catalyst showed almost three times higher O2
48
49
50 22 adsorption energy (in absolute value) compared to the Co-based one, showing at the same time 100
51
52 23 mV more positive Eon and E1/2, thus suggesting that the O2 adsorption energy is a key parameter for
53
54 24 enhancing the ORR kinetics. In addition, they found that the selectivity towards 4 e− reduction is
55
56 25 associated with the adsorption energy and the adsorption configuration of the peroxide intermediate
57
58
26 on the catalytic site. In particular, the O–O bond cleavage during the adsorption is more favored on
59
60 44
ACS Paragon Plus Environment
Page 45 of 75 The Journal of Physical Chemistry

1
2
3 1 the Fe-based sites, which also show higher H2O2 adsorption energy, and led to the formation of a
4
5 2 lower amount of peroxide during ORR. On the contrary, on the Co-based sites, the lower H2O2
6
7 3 adsorption energy hampers the O–O bond breaking, increasing the relative significance of the 2 e−
8
9
10 4 pathway producing H2O2 on this catalyst. All these considerations are confirmed by our
11
12 5 experimental results (see Section 3.9).
13
14 6 Molecular dynamics simulations conducted on Co-substituted proteins responsible for O2
15
16 7 adsorption and transport in living beings (e.g. hemoglobin and myoglobin),90 which contain Fe-
17
18
8 based centers for the oxygen bonding, have shown that the oxygen ligand rotational motion around
19
20
21 9 the metal-oxygen bond is considerably faster than for -Co−O2 than for Fe−O2. Consequently, the
22
23 10 Fe-complex shows more localized ligand sites, while for Co-complex several configurations are
24
25 11 possible with a higher mobility of -Co−O2 compared to Fe−O2, supporting the previous results.
26
27 12 Even though we cannot demonstrate the existence of such these porphyrin- and phthalocyanine-like
28
29
30 13 structures in our catalysts after the pyrolysis, based on some literature studies, the formation of Me-
31
32 14 Nx clusters embedded in graphene, and even more connecting the edges of different graphene
33
34 15 segments (with the presence of Fe–N4 active sites bridged between graphitic micropores), was
35
36 16 found to be energetically favorable.91,92 This is in agreement with experimental evidence reported in
37
38
17 the literature,93,94 as well as in this work, considering the high microporosity of these catalysts.41
39
40
41 18 The ORR catalytic activity is related to the binding energy of O2 on the catalyst surfaces, thus, a
42
43 19 good ORR catalyst must have an optimal (neither too strong nor too weak) binding energy of O2 on
44
45 20 its surface.38,91,95,96 In particular, an optimal Me-N-C catalyst for ORR should bind the O2 molecule
46
47 21 to the transition metal-based active site neither too strongly nor too weakly. In this regard, the series
48
49
50 22 of 3d-type transition metals (from Cr to Zn) is expected to exhibit a volcano shape as a function of
51
52 23 the O2 adsorption energy.
53
54 24 The dependence of the rate of an electrocatalytic reaction on electrode potential is set by two
55
56 25 complementary effects of the potential: (i) the potential-dependent formation of the active sites via a
57
58
59
60 45
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 46 of 75

1
2
3 1 redox-mediated mechanism, and (ii) the lowering of the Ea of the processes (involving electron
4
5 2 transfer) occurring on the active sites.39
6
7 3 The turnover frequency (TOF), which is related (inversely proportional) to Ea, and Eredox of the
8
9
10 4 active center, are in turns strictly related to each other, being determined by the intrinsic nature of
11
12 5 the active sites. If the active sites possess a high Eredox, their site-blocking effect is lowered, but at
13
14 6 the same time, they will have a low O2 binding energy, which causes a low TOF (high Ea barrier to
15
16 7 be overcome). As described by J. Li et al.,38,66 considering both Ea and Eredox, the optimal Eredox
17
18
8 value should be between 300 and 400 mV lower than the ORR thermodynamic potential (E0 =
19
20
21 9 +1.23 V vs. SHE).
22
23 10 According to the kinetic equation describing the relation between current (i) and electrode applied
24
25 11 potential (E):38,39
26
27 (y (( z
28 12 cse ∝   ∙ c1 − ue ∙ vwx /− 4 ∙ vwx /− 4 (13)
 {
29
J J
1−u = =
30
13 ~€S€‚ƒ„…†
(14)
31 Jp|
32 Jp} ‡ˆ

33
34 14 assuming that αc = 0.5, a Tafel slope of 120 mV dec−1 must be observed for ORR. Nevertheless,
35
36 15 both Pt-based catalyst and several highly active Me-N-C catalysts reported in the literature, show
37
38 16 curvature in the Tafel slope, which is potential-dependent. In particular, as for our Fe-N-C (and
39
40 17 partially also for Co-N-C) catalyst, the Tafel slope in the low η region is ~ 60 mV dec−1, increasing
41
42
43 18 to ~ 120 mV dec−1 at high η. This behavior is well explained by the theory of the redox surface
44
45 19 mediated generation of active sites at low η.38–40
46
47 20 Here we will try to resume all the experimental evidence found in this work, and to propose an
48
49 21 explanation of the behavior of the three different metals (Fe, Co, Cu) according to the site-blocking
50
51
22 effects and the redox potential mediated theory of the active sites formation.38
52
53
54 23 Fe-N-C
55
56 24 The behavior of Fe-N-C catalyst is analogous to the behavior of Pt/C catalysts: it exhibits a double
57
58 25 Tafel slope behavior with a 60 mV dec−1 slope at low η, becoming gradually 120 mV dec−1 at
59
60 46
ACS Paragon Plus Environment
Page 47 of 75 The Journal of Physical Chemistry

1
2
3 1 higher η. This behavior is in accordance to the theory of the site-blocking of a redox-mediated
4
5 2 mechanism of formation of the active sites. Moreover, according to what proposed in the
6
7 3 literature,38 the Eredox (and consequently the O2 adsorption energy) of Fe-N-C active sites should be
8
9
10 4 located more in proximity of the top of the asymmetric “volcano plot” compared to Co-N-C (likely
11
12 5 located on the right side branch) and Cu-N-C (likely located on the left side branch).
13
14 6 Regarding the apparent activation energy, Ea, which is related to the TOF according to what stated
15
16 7 by Mukerjee and co-workers, the value is higher at high E, and it decreases with the decrease of E.
17
18
8 The RO2 at potentials close to Eon (low η) is < 1, increasing from about 0.5 moving towards 1 with
19
20
21 9 the decrease of E. The calculation of the RH+, even if less accurate, shows a decrease with the
22
23 10 potential.
24
25 11 According to Equation (14), at potentials E >> Eredox the coverage (θ) approaches unity, and thus no
26
27 12 active sites are free for the ORR, and the ORR faradaic current is null. At E << Eredox the situation is
28
29
30 13 exactly the opposite, that is, θ ≈ 0, with the maximum quantity of active sites available, and the
31
32 14 maximum current theoretically achievable. Considering this, both the decrease of Ea and the
33
34 15 increase of RO2 with the decrease of potential are reasonable results.
35
36 16 The change in Tafel slope is also in accordance with these facts, since the 60 mV dec−1 observed at
37
38
17 low η is given by the Nernstian behavior of the Tafel slope according to Equation (15). In facts,
39
40
41 18 when the number of active sites available for ORR is the limiting factor (at low η, when θ → 1),
42
43 19 Equation (13) becomes:
44
J J J
45
20 cse ≅ ≈ = ~€S€‚ƒ„… †
(15)
46 Jp| |
} ‡ˆ
47
48
49
21 From the Nernst Equation, we have:
50
 „…
51 22 s − sŠ}"‹ = ln / 4 (16)
! ‚ƒ
52
53
54 23 where aox and ared are respectively the concentration of the oxidized and the reduced form of the
55
56 24 active site on the surface of the catalyst. Substituting in Equation (15) and passing to logarithms, we
57
58 25 obtain:
59
60 47
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 48 of 75

1
2

3 1 s= ln  − sŠ}"‹ (17)
4 !
5
6 2 and the experimental evidence of a Nernstian Tafel slope is explained:
7
"( 
8 3 =− = −60 mVdec J (18)
9 "$,1 ).*O*!
10
11 4 Instead, when all the active sites for ORR become available (θ → 0), the behavior of the Tafel slope
12
13 5 becomes to be governed by the faradaic ORR process, showing 120 mV dec−1, which is accordance
14
15 6 to αc ≈ 0.5 (see Table 1).
16
17
7 Regarding the behavior towards the reduction of adsorbed O2 it shows well defined ORR peaks
18
19
20 8 with high Ip, and Ep shifting to more negative values with the increase of the scan rate. The αc
21
22 9 values calculated with the two differential CV methods in the potential regions between 0.7 and 0.6
23
24 10 V vs. RHE (where the 120 mV dec−1 Tafel slope was measured in RDE, see Table 1 and Figure
25
26 11 3b) and resumed in Table 2 and Figure 6a, approach 0.5. These findings indicate that in the case of
27
28
29 12 the reduction of adsorbed O2 via CV, where we are not in static conditions as in SV measured in
30
31 13 RDE, the reduction of O2 already adsorbed on the catalyst surface takes place when the active sites
32
33 14 formation is already occurred, as well as the O2 adsorption on them. Thus the active sites are no
34
35 15 more the limiting factor for the ORR, and the typical Tafel slope for the first electron transfer is
36
37
38
16 observed.
39
40 17 Regarding the production of peroxide (RRDE test), it is low for this catalyst (maximum % H2O2 ≈
41
42 18 2%, see Figure 11b), confirming a good adsorption of the peroxide intermediate on the active sites.
43
44 19 Co-N-C
45
46 20 The behavior of Co-N-C catalyst is somehow different compared to Fe-N-C, even though it shows
47
48
49 21 some analogies. The analogies are: similar potential-dependent double Tafel slope, same trend of Ea
50
51 22 and RO2 with potential, and similar values of RH+ in the two different zones of the Tafel plot.
52
53 23 The differences between the two catalysts are however marked in the CV of the adsorbed O2
54
55 24 reduction (Section 3.2). Moreover, a considerably higher (almost 10 times) H2O2 was detected in
56
57
58
25 the RRDE test, and the curvature of the Tafel slope is much more marked at high η (as evident in
59
60 48
ACS Paragon Plus Environment
Page 49 of 75 The Journal of Physical Chemistry

1
2
3 1 Figures 3-4-8-9). Taking into account the phenomena mentioned above about the relation between
4
5 2 the redox mediated formation of active sites and Ea, we could assign to the Eredox of the active sites
6
7 3 of the Co-N-C catalyst a higher value compared to the ones of Fe-N-C, also according to the
8
9
10 4 literature,91,95,96 with a consequent weaker O2-site binding energy.89,90 This is well in accordance to
11
12 5 the following experimental evidence: RO2 of Co-N-C > RO2 of Fe-N-C at low overpotentials (higher
13
14 6 amount of “free” active sites already available, that is, lower θ); Ea of Co-N-C > Ea of Fe-N-C at
15
16 7 high overpotentials (lower O2 binding energy); steeper decrease of Ea with potential for Co-N-C
17
18
8 compared to Fe-N-C; higher H2O2 production (related to the lower site-O2 binding energy, which
19
20
21 9 results in a faster desorption of the peroxide intermediate in the bulk of the electrolyte before the O–
22
23 10 O bond cleavage; lower ORR peak currents in the differential CV for the reduction of adsorbed O2
24
25 11 (confirming that O2 is only weakly adsorbed on the catalyst surface, being adsorption much
26
27 12 hindered than in Fe-N-C).
28
29
30 13 By concluding, for Co-N-C, all these experimental evidence suggest that the Eredox is located in the
31
32 14 right side branch of the asymmetric volcano plot postulated by Mukerjee et al., and the Ea of the
33
34 15 rate determining step of the faradaic ORR is higher than for Fe-N-C. In addition, the steeper change
35
36 16 in the Tafel slope at high η for Co-N-C seems to be more driven by the achievement of the
37
38
17 maximum number of active sites available on the catalyst surface,40 being this the factor limiting the
39
40
41 18 current density, more than the kinetics of the ORR rate determining step.
42
43 19 Cu-N-C
44
45 20 Compared to Fe-N-C and Co-N-C its Eon is remarkably lower, and the most evident difference that
46
47 21 it shows compared to the former catalysts is the single Tafel slope of 120 mV dec−1. It also shows
48
49
50 22 other differences, having an almost constant Ea at different potentials. The analogy is the increase of
51
52 23 the RO2 with the increase of η in the kinetic control potential range. Analogously to Fe-N-C catalyst,
53
54 24 Cu-N-C exhibits well developed adsorbed ORR current peaks (located at lower E). The reason why
55
56 25 a single 120 mV dec−1 Tafel slope is observed for Cu-N-C could be attributed to the fact that for its
57
58
26 active sites the Eredox potential is situated on the left-side branch of the asymmetric “volcano plot”,
59
60 49
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 50 of 75

1
2
3 1 corresponding to a high O2 adsorption energy, which somehow hampers the ORR at low η. The
4
5 2 reason why the 60 mV dec−1 Nernstian Tafel slope is not observed for this catalyst could be
6
7 3 attributed to the fact that the site blocking species are released at a potential higher than the Eon of
8
9
10 4 the ORR. Thus, due to the too high “site−O2” binding energy, the ORR occurs at lower potentials,
11
12 5 when all the available active sites are “free” from blocking species (θ ≈ 0). The ORR activity of the
13
14 6 Cu-N-C catalyst is too low, in consequence of its low Eon and too strong O2 binding energy on the
15
16 7 active sites.96 These considerations also explain the almost constant value of Ea (no ORR limited by
17
18
8 the “Nernstian” generation of active sites), and possibly also the RO2 lower than 1 (between 0.5 and
19
20
21 9 0.8) in the kinetic zone of the polarization curve.
22
23 10
24
25 11 4. Conclusions
26
27 12 A set of electrochemical tests in the three-electrodes cell has been proposed to investigate the
28
29
30 13 kinetics of the ORR in acidic medium on a series of Me-N-C (Me = Fe, Co, Cu) via CV and RDE-
31
32 14 RRDE. These tests enabled to determine the following parameters: Tafel slope, reduction of
33
34 15 adsorbed O2, reaction orders for O2 and H+, apparent activation energy, and H2O2 selectivity. The
35
36 16 results obtained for the three different catalysts evidenced a clear influence of the transition metal in
37
38
17 the ORR activity and kinetics.
39
40
41 18 Fe-N-C was found to be the most performing catalyst, showing a potential dependent double Tafel
42
43 19 slope behavior, according to the surface redox-mediated coverage of active sites. Compared to the
44
45 20 other two catalysts, its Ea and Eredox should be located more in the proximity to the top of the
46
47 21 asymmetric “volcano plot”, as proposed in the literature. Indeed, its behavior is more similar to the
48
49
50 22 one of a Pt-based catalyst.
51
52 23 Co-N-C also shows an apparently potential dependent double Tafel slope behavior, but it performs
53
54 24 slightly worse compared to Fe-N-C in terms of Eon. Its adsorbed O2 reduction Ip were found to be
55
56 25 considerably lower, and it produces much more H2O2 during ORR. These results indicate that its O2
57
58
26 adsorption energy is weaker, suggesting that the Eredox of its active sites should be located more
59
60 50
ACS Paragon Plus Environment
Page 51 of 75 The Journal of Physical Chemistry

1
2
3 1 positively compared to Fe-N-C, causing an enhancement in the Ea of the ORR. Co-N-C should be
4
5 2 located more on the right-side branch of the asymmetric “volcano-plot” compared to Fe-N-C.
6
7 3 Cu-N-C shows a considerably lower Eon, a single Tafel slope behavior, and high adsorbed O2 Ip
8
9
10 4 similarly to Fe-N-C, but more displaced to negative potentials. This behavior suggests that the Eredox
11
12 5 for the active sites of Cu-N-C should be more negative, causing an enhancement of Ea due to the
13
14 6 stronger O2 adsorption energy. Thus, Cu-N-C cannot be considered a suitable ORR catalyst, since it
15
16 7 is located too negatively in the left-side branch of the asymmetric “volcano plot”.
17
18
8
19
20
21 9 Acknowledgments
22
23 10 Funding by the Italian Ministry of Education, University and Research [PRIN NAMEDPEM,
24
25 11 “Advanced nanocomposite membranes and innovative electrocatalysts for durable polymer
26
27 12 electrolyte membrane fuel cells”, grant n. 2010CYTWAW] is gratefully acknowledged. Funding
28
29
30 13 from the Spanish Ministry of Economy project ENE2016-77055-C3-1-R and from the Madrid
31
32 14 Regional Research Council (CAM) project S2013/MAE-2882 (RESTOENE-2) are acknowledged
33
34 15 as well.
35
36 16
37
38
39
17 References
40
41 18 (1) Chen, Z.; Higgins, D.; Yu, A.; Zhang, L.; Zhang, J. A Review on Non-Precious Metal
42
43 19 Electrocatalysts for PEM Fuel Cells. Energy Environ. Sci. 2011, 4, 3167–3192.
44
45 20 (2) Banham, D.; Ye, S.; Pei, K.; Ozaki, J.; Kishimoto, T.; Imashiro, Y. A Review of the Stability
46
47 21 and Durability of Non-Precious Metal Catalysts for the Oxygen Reduction Reaction in
48
49
50 22 Proton Exchange Membrane Fuel Cells. J. Power Sources 2015, 285, 334–348.
51
52 23 (3) Jeon, I.-Y.; Yu, D.; Bae, S.-Y.; Choi, H.-J.; Chang, D. W.; Dai, L.; Baek, J.-B. Formation of
53
54 24 Large-Area Nitrogen-Doped Graphene Film Prepared from Simple Solution Casting of Edge-
55
56 25 Selectively Functionalized Graphite and Its Electrocatalytic Activity. Chem. Mater. 2011, 23,
57
58
26 3987–3992.
59
60 51
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 52 of 75

1
2
3 1 (4) Jaouen, F.; Proietti, E.; Lefèvre, M.; Chenitz, R.; Dodelet, J.-P.; Wu, G.; Chung, H. T.;
4
5 2 Johnston, C. M.; Zelenay, P. Recent Advances in Non-Precious Metal Catalysis for Oxygen-
6
7 3 Reduction Reaction in Polymer Electrolyte Fuel Cells. Energy Environ. Sci. 2011, 4 , 114–
8
9
10 4 130.
11
12 5 (5) Othman, R.; Dicks, A. L.; Zhu, Z. Non Precious Metal Catalysts for the PEM Fuel Cell
13
14 6 Cathode. Int. J. Hydrogen Energy 2012, 37, 357–372.
15
16 7 (6) Jasinski, R. A New Fuel Cell Cathode Catalyst. Nature 1964, 201, 1212–1213.
17
18
8 (7) Monteverde Videla, A. H. A.; Osmieri, L.; Specchia, S. Non-Noble Metal (NNM) Catalysts
19
20
21 9 for Fuel Cells: Tuning the Activity by a Rational Step-by-Step Single Variable Evolution. In
22
23 10 Electrochemistry of N4 Macrocyclic Metal Complexes - Volume 1: Energy; Zagal, J. H.,
24
25 11 Bedioui, F., Eds.; 2016; pp 69–102.
26
27 12 (8) Vignarooban, K.; Lin, J.; Arvay, A.; Kolli, S.; Kruusenberg, I.; Tammeveski, K.; Munukutla,
28
29
30 13 L.; Kannan, A. M. Nano-Electrocatalyst Materials for Low Temperature Fuel Cells: A
31
32 14 Review. Chinese J. Catal. 2015, 36, 458–472.
33
34 15 (9) Wu, G.; Santandreu, A.; Kellogg, W.; Gupta, S.; Ogoke, O.; Zhang, H.; Wang, H. L.; Dai, L.
35
36 16 Carbon Nanocomposite Catalysts for Oxygen Reduction and Evolution Reactions: From
37
38
17 Nitrogen Doping to Transition-Metal Addition. Nano Energy 2016, 29, 83–110.
39
40
41 18 (10) Choi, C. H.; Baldizzone, C.; Polymeros, G.; Pizzutilo, E.; Kasian, O.; Schuppert, A. K.;
42
43 19 Ranjbar Sahraie, N.; Sougrati, M.-T.; Mayrhofer, K. J. J.; Jaouen, F. Minimizing Operando
44
45 20 Demetallation of Fe-N-C Electrocatalysts in Acidic Medium. ACS Catal. 2016, 6, 3136–
46
47 21 3146.
48
49
50 22 (11) Wang, X.; Wang, B.; Zhong, J.; Zhao, F.; Han, N.; Huang, W.; Zeng, M.; Fan, J.; Li, Y. Iron
51
52 23 Polyphthalocyanine Sheathed Multiwalled Carbon Nanotubes: A High-Performance
53
54 24 Electrocatalyst for Oxygen Reduction Reaction. Nano Res. 2016, 9 , 1497–1506.
55
56 25 (12) Wu, G.; More, K. L.; Johnston, C. M.; Zelenay, P. High-Performance Electrocatalysts for
57
58
26 Oxygen Reduction Derived from Polyaniline, Iron, and Cobalt. Science 2011, 332 , 443–448.
59
60 52
ACS Paragon Plus Environment
Page 53 of 75 The Journal of Physical Chemistry

1
2
3 1 (13) Cheon, J. Y.; Kim, T.; Choi, Y.; Jeong, H. Y.; Kim, M. G.; Sa, Y. J.; Kim, J.; Lee, Z.; Yang,
4
5 2 T.-H.; Kwon, K.; et al. Ordered Mesoporous Porphyrinic Carbons with Very High
6
7 3 Electrocatalytic Activity for the Oxygen Reduction Reaction. Sci. Rep. 2013, 3, 2715.
8
9
10 4 (14) Wang, X.; Zhang, H.; Lin, H.; Gupta, S.; Wang, C.; Tao, Z.; Fu, H.; Wang, T.; Zheng, J.;
11
12 5 Wu, G.; et al. Directly Converting Fe−doped Metal-Organic Frameworks into Highly Active
13
14 6 and Stable Fe−N-C Catalysts for Oxygen Reduction in Acid. Nano Energy 2016, 510, 426–
15
16 7 436.
17
18
8 (15) Grgur, B. N.; Marković, N. M.; Ross, P. N. Temperature-Dependent Oxygen
19
20
21 9 Electrochemistry on Platinum Low-Index Single Crystal Surfaces in Acid Solutions. Can. J.
22
23 10 Chem. 1997, 75, 1465–1471.
24
25 11 (16) Damjanovic, A.; Sepa, D. B. An Analysis of the pH Dependence of Enthalpies and Gibbs
26
27 12 Energies of Activation for O2 Reduction at Pt Electrodes in Acid Solutions. Electrochim.
28
29
30 13 Acta 1990, 35, 1157–1162.
31
32 14 (17) Bett, J.; Lundquist, J.; Washington, E.; Stonehart, P. Platinum Crystallite Size Considerations
33
34 15 for Electrocatalytic Oxygen Reduction. Electrochim. Acta 1973, 18, 343–348.
35
36 16 (18) Anderson, A. B.; Roques, J.; Mukerjee, S.; Murthi, V. S.; Markovic, N. M.; Stamenkovic, V.
37
38
17 Activation Energies for Oxygen Reduction on Platinum Alloys: Theory and Experiment. J.
39
40
41 18 Phys. Chem. B 2005, 109, 1198–1203.
42
43 19 (19) Neyerlin, K. C.; Gu, W.; Jorne, J.; Gasteiger, H. A. Determination of Catalyst Unique
44
45 20 Parameters for the Oxygen Reduction Reaction in a PEMFC. J. Electrochem. Soc. 2006, 153,
46
47 21 A1955-A1963.
48
49
50 22 (20) Stamenković, V.; Schmidt, T. J.; Ross, P. N.; Marković, N. M. Surface Composition Effects
51
52 23 in Electrocatalysis: Kinetics of Oxygen Reduction on Well-Defined Pt 3 Ni and Pt 3 Co
53
54 24 Alloy Surfaces. J. Phys. Chem. B 2002, 106, 11970–11979.
55
56 25 (21) Paulus, U. A.; Wokaun, A.; Scherer, G.; Schmidt, T. J.; Stamenkovic, V.; Radmilovic, V.;
57
58
26 Markovic, N. M.; Ross, P. N. Oxygen Reduction on Carbon-Supported Pt - Ni and Pt - Co
59
60 53
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 54 of 75

1
2
3 1 Alloy Catalysts. J. Phys. Chem. B. 2002, 106, 4181–4191.
4
5 2 (22) Chlistunoff, J. RRDE and Voltammetric Study of ORR on Pyrolyzed Fe / Polyaniline
6
7 3 Catalyst . On the Origins of Variable Tafel Slopes. J. Phys. Chem. C 2011, 115, 6496–6507.
8
9
10 4 (23) Jaouen, F.; Goellner, V.; Lefèvre, M.; Herranz, J.; Proietti, E.; Dodelet, J. P. Oxygen
11
12 5 Reduction Activities Compared in Rotating-Disk Electrode and Proton Exchange Membrane
13
14 6 Fuel Cells for Highly Active FeNC Catalysts. Electrochim. Acta 2013, 87, 619–628.
15
16 7 (24) Tributsch, H.; Koslowski, U. I.; Dorbandt, I. Experimental and Theoretical Modeling of Fe-,
17
18
8 Co-, Cu-, Mn-Based Electrocatalysts for Oxygen Reduction. Electrochim. Acta 2008, 53,
19
20
21 9 2198–2209.
22
23 10 (25) Chen, R.; Li, H.; Chu, D.; Wang, G. Unraveling Oxygen Reduction Reaction Mechanisms on
24
25 11 Carbon-Supported Fe-Phthalocyanine and Co-Phthalocyanine Catalysts in Alkaline Solutions
26
27 12 Report Documentation Page Unraveling Oxygen Reduction Reaction Mechanisms on
28
29
30 13 Richard G . Lugar Center for Renewable. J. Phys. Chem. C 2009, 113, 20689–20697.
31
32 14 (26) Jaouen, F.; Dodelet, J.-P. O2 Reduction Mechanism on Non-Noble Metal Catalysts for PEM
33
34 15 Fuel Cells. Part I: Experimental Rates of O2 Electroreduction, H2O2 Electroreduction, and
35
36 16 H2O2 Disproportionation. J. Phys. Chem. C 2009, 113, 15422–15432.
37
38
17 (27) Xu, P.; Chen, W.; Wang, Q.; Zhu, T.; Wu, M.; Qiao, J. Effects of Transition Metal
39
40
41 18 Precursors (Co, Fe, Cu, Mn, or Ni) on Pyrolyzed Carbon Supported Metal- Aminopyrine
42
43 19 Electrocatalysts for Oxygen Reduction Reaction. RSC Adv. 2014, 5, 6195–6206.
44
45 20 (28) Domínguez, C.; Pérez-Alonso, F. J.; Abdel Salam, M.; Gómez de la Fuente, J. L.; Al-
46
47 21 Thabaiti, S. a.; Basahel, S. N.; Peña, M. a.; Fierro, J. L. G.; Rojas, S. Effect of Transition
48
49
50 22 Metal (M: Fe, Co or Mn) for the Oxygen Reduction Reaction with Non-Precious Metal
51
52 23 Catalysts in Acid Medium. Int. J. Hydrogen Energy 2014, 39, 5309–5318.
53
54 24 (29) Dodelet, J. The Controversial Role of the Metal in Fe- or Co-Based Electrocatalysts for the
55
56 25 Oxygen Reduction Reaction in Acid Medium. In Electrocatalysis in Fuel Cells; Shao, M.,
57
58
26 Ed.; Lecture Notes in Energy; Springer London: London, 2013; Vol. 9, pp 271–338.
59
60 54
ACS Paragon Plus Environment
Page 55 of 75 The Journal of Physical Chemistry

1
2
3 1 (30) Zagal, J. H.; Maritza, A. P.; Silva, J. F. Fundamental Aspects on the Catalytic Activity of
4
5 2 Metallomacrocyclics for the Electrochemical Reduction of O2. In N4-Macrocyclic Metal
6
7 3 Complexes; Zagal, J. H., Bedioui, F., Dodelet, J.-P., Eds.; Springer, 2006; pp 41–82.
8
9
10 4 (31) Isaacs, F.; Dehaen, W.; Maes, W.; Ngo, T. H.; León, D. R.-. Direct 4-Electron Reduction of
11
12 5 Molecular Oxygen to Water Modified Electrodes. Int. J. Electrochem. Sci. 2013, 8, 3406–
13
14 6 3418.
15
16 7 (32) Smith, T. D.; Pilbrow, J. R. Recent Developments in the Studies of Molecular Oxygen
17
18
8 Adducts of Cobalt (II) Compounds and Related Systems. Coord. Chem. Rev. 1981, 39, 295–
19
20
21 9 383.
22
23 10 (33) Isaacs, M.; Aguirre, M. J.; Toro-Labbé, A.; Costamagna, J.; Páez, M.; Zagal, J. H.
24
25 11 Comparative Study of the Electrocatalytic Activity of Cobalt Phthalocyanine and Cobalt
26
27 12 Naphthalocyanine for the Reduction of Oxygen and the Oxidation of Hydrazine.
28
29
30 13 Electrochim. Acta 1998, 43, 1821–1827.
31
32 14 (34) Liu, H.; Zhang, L.; Zhang, J.; Ghosh, D.; Jung, J.; Downing, B. W.; Whittemore, E.
33
34 15 Electrocatalytic Reduction of O2 and H2O2 by Adsorbed Cobalt Tetramethoxyphenyl
35
36 16 Porphyrin and Its Application for Fuel Cell Cathodes. J. Power Sources 2006, 161, 743–752.
37
38
17 (35) Kim, E.; Helton, M. E.; Wasser, I. M.; Karlin, K. D.; Lu, S.; Huang, H.-W.; Moënne-
39
40
41 18 Loccoz, P.; Incarvito, C. D.; Rheingold, A. L.; Honecker, M.; et al. Superoxo, M-Peroxo, and
42
43 19 (M-Oxo Complexes from heme/O2 and Heme-Cu/O2 Reactivity: Copper Ligand Influences
44
45 20 in Cytochrome c Oxidase Models. In Proc. Natl. Acad. Sci.; 2003; pp 3623–3628.
46
47 21 (36) Monteverde Videla, A. H. a.; Osmieri, L.; Armandi, M.; Specchia, S. Varying the
48
49
50 22 Morphology of Fe-N-C Electrocatalysts by Templating Iron Phthalocyanine Precursor with
51
52 23 Different Porous SiO2 to Promote the Oxygen Reduction Reaction. Electrochim. Acta 2015,
53
54 24 177, 43–50.
55
56 25 (37) Marden, M. C.; Kiger, L.; Poyart, C.; Rashid, A. K.; Kister, J.; Stetzkowski-Marden, F.;
57
58
26 Caron, G.; Haque, M.; Moens, L. Modulation of the Oxygen Affinity of Cobalt-Porphyrin by
59
60 55
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 56 of 75

1
2
3 1 Globin. FEBS Lett. 2000, 472, 221–224.
4
5 2 (38) Li, J.; Alsudairi, A.; Ma, Z.-F.; Mukerjee, S.; Jia, Q. Asymmetric Volcano Trend in Oxygen
6
7 3 Reduction Activity of Pt and Non-Pt Catalysts: In Situ Identification of the Site-Blocking
8
9
10 4 Effect. J. Am. Chem. Soc. 2017, 139, 1384–1387.
11
12 5 (39) Gottesfeld, S. Generation of Active Sites by Potential-Driven Surface Processes: A Central
13
14 6 Aspect of Electrocatalysis. ECS Trans. 2014, 61, 1–13.
15
16 7 (40) Zagal, J. H.; Koper, M. T. M. Reactivity Descriptors for the Activity of Molecular MN4
17
18
8 Catalysts for the Oxygen Reduction Reaction. Angew. Chemie - Int. Ed. 2016, 55, 14510–
19
20
21 9 14521.
22
23 10 (41) Osmieri, L.; Monteverde Videla, A. H. A.; Armandi, M.; Specchia, S. Influence of Different
24
25 11 Transition Metals on the Properties of Me-N-C (Me = Fe, Co, Cu, Zn) Catalysts Synthesized
26
27 12 Using SBA-15 as Tubular Nano-Silica Reactor for Oxygen Reduction Reaction. Int. J.
28
29
30 13 Hydrogen Energy 2016, 41, 22570–22588.
31
32 14 (42) Paulus, U. A.; Schmidt, T. J.; Gasteiger, H. A.; Behm, R. J. Oxygen Reduction on a High-
33
34 15 Surface Area Pt/Vulcan Carbon Catalyst: A Thin-Film Rotating Ring-Disk Electrode Study.
35
36 16 J. Electroanal. Chem. 2001, 495, 134–145.
37
38
17 (43) Serov, A.; Tylus, U.; Artyushkova, K.; Mukerjee, S.; Atanassov, P. Mechanistic Studies of
39
40
41 18 Oxygen Reduction on Fe-PEI Derived Non-PGM Electrocatalysts. Appl. Catal. B Environ.
42
43 19 2014, 150–151, 179–186.
44
45 20 (44) Muthukrishnan, A.; Nabae, Y.; Hayakawa, T.; Okajima, T.; Ohsaka, T. Fe-Containing
46
47 21 Polyimide-Based High-Performance ORR Catalysts in Acidic Medium: A Kinetic Approach
48
49
50 22 to Study the Durability of Catalysts. Catal. Sci. Technol. 2015, 5, 475–483.
51
52 23 (45) Ferrandon, M.; Kropf, A. J.; Myers, D. J.; Kramm, U.; Bogdano, P.; Wu, G.; Johnston, C.
53
54 24 M.; Zelenay, P. Multitechnique Characterization of a Polyaniline − Iron − Carbon Oxygen
55
56 25 Reduction Catalyst. J. Phys. Chem. C 2012, 116, 16001–16013.
57
58
26 (46) van der Vliet, D.; Strmcnik, D. S.; Wang, C.; Stamenkovic, V. R.; Markovic, N. M.; Koper,
59
60 56
ACS Paragon Plus Environment
Page 57 of 75 The Journal of Physical Chemistry

1
2
3 1 M. T. M. On the Importance of Correcting for the Uncompensated Ohmic Resistance in
4
5 2 Model Experiments of the Oxygen Reduction Reaction. J. Electroanal. Chem. 2010, 647,
6
7 3 29–34.
8
9
10 4 (47) Monteverde Videla, A. H. a; Zhang, L.; Kim, J.; Zeng, J.; Francia, C.; Zhang, J.; Specchia, S.
11
12 5 Mesoporous Carbons Supported Non-Noble Metal Fe-N X Electrocatalysts for PEM Fuel
13
14 6 Cell Oxygen Reduction Reaction. J. Appl. Electrochem. 2013, 43, 159–169.
15
16 7 (48) Chlistunoff, J.; Sansiñena, J.-M. Nafion Induced Surface Confinement of Oxygen in Carbon-
17
18
8 Supported Oxygen Reduction Catalysts. J. Phys. Chem. C 2016, 120, 28038–28048.
19
20
21 9 (49) Chlistunoff, J.; Sansiñena, J.-M. On the Use of Nafion® in Electrochemical Studies of
22
23 10 Carbon Supported Oxygen Reduction Catalysts in Aqueous Media. J. Electroanal. Chem.
24
25 11 2016, 780, 134–146.
26
27 12 (50) Frackowiak, E.; Béguin, F. Carbon Materials for the Electrochemical Storage of Energy in
28
29
30 13 Capacitors. Carbon N. Y. 2001, 39, 937–950.
31
32 14 (51) Monteverde Videla, A. H. a.; Ban, S.; Specchia, S.; Zhang, L.; Zhang, J. Non-Noble Fe–NX
33
34 15 Electrocatalysts Supported on the Reduced Graphene Oxide for Oxygen Reduction Reaction.
35
36 16 Carbon N. Y. 2014, 76, 386–400.
37
38
17 (52) Shinozaki, K.; Morimoto, Y.; Pivovar, B. S.; Kocha, S. S. Suppression of Oxygen Reduction
39
40
41 18 Reaction Activity on Pt-Based Electrocatalysts from Ionomer Incorporation. J. Power
42
43 19 Sources 2016, 325, 745–751.
44
45 20 (53) Byon, H. R.; Suntivich, J.; Shao-Horn, Y. Graphene-Based Non-Noble-Metal Catalysts for
46
47 21 Oxygen Reduction Reaction in Acid. Chem. Mater. 2011, 23, 3421–3428.
48
49
50 22 (54) Tsai, C.-W.; Tu, M.-H.; Chen, C.-J.; Hung, T.-F.; Liu, R.-S.; Liu, W.-R.; Lo, M.-Y.; Peng,
51
52 23 Y.-M.; Zhang, L.; Zhang, J.; et al. Nitrogen-Doped Graphene Nanosheet-Supported Non-
53
54 24 Precious Iron Nitride Nanoparticles as an Efficient Electrocatalyst for Oxygen Reduction.
55
56 25 RSC Adv. 2011, 1, 1349.
57
58
26 (55) Zhang, L.; Kim, J.; Dy, E.; Ban, S.; Tsay, K. C.; Kawai, H.; Shi, Z.; Zhang, J. Synthesis of
59
60 57
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 58 of 75

1
2
3 1 Novel Mesoporous Carbon Spheres and Their Supported Fe-Based Electrocatalysts for PEM
4
5 2 Fuel Cell Oxygen Reduction Reaction. Electrochim. Acta 2013, 108, 480–485.
6
7 3 (56) Guidelli, R.; Compton, R. G.; Feliu, J. M.; Gileadi, E.; Lipkowski, J.; Schmickler, W.;
8
9
10 4 Trasatti, S. Defining the Transfer Coefficient in Electrochemistry: An Assessment (IUPAC
11
12 5 Recommendations 2014). Pure Appl. Chem. 2014, 86, 259–262.
13
14 6 (57) Guidelli, R.; Compton, R. G.; Feliu, J. M.; Gileadi, E.; Lipkowski, J.; Schmickler, W.;
15
16 7 Trasatti, S. Defining the Transfer Coefficient in Electrochemistry: An Assessment (IUPAC
17
18
8 Technical Report). Pure Appl. Chem. 2014, 86, 245–258.
19
20
21 9 (58) Meng, H.; Larouche, N.; Lefvre, M.; Jaouen, F.; Stansfield, B.; Dodelet, J. P. Iron Porphyrin-
22
23 10 Based Cathode Catalysts for Polymer Electrolyte Membrane Fuel Cells: Effect of NH3 and
24
25 11 Ar Mixtures as Pyrolysis Gases on Catalytic Activity and Stability. Electrochim. Acta 2010,
26
27 12 55, 6450–6461.
28
29
30 13 (59) Jaouen, F.; Herranz, J.; Lefèvre, M.; Dodelet, J.-P.; Kramm, U. I.; Herrmann, I.; Bogdanoff,
31
32 14 P.; Maruyama, J.; Nagaoka, T.; Garsuch, A.; et al. Cross-Laboratory Experimental Study of
33
34 15 Non-Noble-Metal Electrocatalysts for the Oxygen Reduction Reaction. ACS Appl. Mater.
35
36 16 Interfaces 2009, 1, 1623–1639.
37
38
17 (60) Osmieri, L.; Escudero-cid, R.; Monteverde Videla, A. H. A.; Ocón, P.; Specchia, S.
39
40
41 18 Performance of a Fe-N-C Catalyst for the Oxygen Reduction Reaction in Direct Methanol
42
43 19 Fuel Cell : Cathode Formulation Optimization and Short-Term Durability. Appl. Catal. B
44
45 20 Environ. 2017, 201, 253–265.
46
47 21 (61) Perry, M. L. Mass Transport in Gas-Diffusion Electrodes: A Diagnostic Tool for Fuel-Cell
48
49
50 22 Cathodes. J. Electrochem. Soc. 1998, 145, 5–15.
51
52 23 (62) Gojkovic, S. L.; Zecevic, S. K.; Savinell, R. F. O2 Reduction on an Ink-Type Rotating Disk
53
54 24 Electrode Using Pt Supported on High-Area Carbons. J. Electrochem. Soc. 1998, 145, 3713–
55
56 25 3720.
57
58
26 (63) Osmieri, L.; Monteverde Videla, A. H. A.; Specchia, S. The Use of Different Types of
59
60 58
ACS Paragon Plus Environment
Page 59 of 75 The Journal of Physical Chemistry

1
2
3 1 Reduced Graphene Oxide in the Preparation of Fe-N-C Electrocatalysts: Capacitive Behavior
4
5 2 and Oxygen Reduction Reaction Activity in Alkaline Medium. J. Solid State Electrochem.
6
7 3 2016, 20, 3507–3523.
8
9
10 4 (64) Greef, R.; Peat, R.; Peter, L. M.; Pletcher, D.; Robinson, J. Instrumental Methods in
11
12 5 Electrochemistry; Ellis Horwood Limited: Chichester, West Sussex, England, 1985.
13
14 6 (65) Laviron, E. General Expression of the Linear Potential Sweep Voltammogram in the Case of
15
16 7 Diffusionless Electrochemical Systems. J. Electroanal. Chem. 1979, 101, 19–28.
17
18
8 (66) Li, J.; Ghoshal, S.; Liang, W.; Sougrati, M.-T.; Jaouen, F.; Halevi, B.; McKinney, S.;
19
20
21 9 McCool, G.; Ma, C.; Yuan, X.; et al. Structural and Mechanistic Basis for the High Activity
22
23 10 of Fe–N–C Catalysts toward Oxygen Reduction. Energy Environ. Sci. 2016, 9, 2418–2432.
24
25 11 (67) Kaskiala, T. Determination of Oxygen Solubility in Aqueous Sulphuric Acid Media. Miner.
26
27 12 Eng. 2002, 15, 853–857.
28
29
30 13 (68) Cai, W.; Zhao, X.; Liu, C.; Xing, W.; Zhang, J. Electrode Kinetics of Electron-Transfer and
31
32 14 Reactant Transport in Electrolyte Solution. In Rotating Electrode Methods and Oxygen
33
34 15 Reduction Electrocatalysis; Xing, W., Yin, G., Zhang, J., Eds.; Elsevier; pp 33–65.
35
36 16 (69) Damjanovic, A.; Genshaw, M. A.; Bockris, J. O. The Role of Hydrogen Peroxide in Oxygen
37
38
17 Reduction at Platinum in H,SO, Solution. J. Electrochem. Soc. 1967, 114, 466–472.
39
40
41 18 (70) Kocha, S. S. Principles of MEA Preparation. In Handbook of Fuel Cells – Fundamentals,
42
43 19 Technology and Applications; Vielstich, W., Lamm, A., Gasteiger, H. A., Eds.; John Wiley
44
45 20 & Sons, Ltd,: Chichester, 2003; Vol. 3, pp 538–565.
46
47 21 (71) Takahashi, I.; Kocha, S. S. Examination of the Activity and Durability of PEMFC Catalysts
48
49
50 22 in Liquid Electrolytes. J. Power Sources 2010, 195, 6312–6322.
51
52 23 (72) Alia, S. M.; Pylypenko, S.; Dameron, A.; Neyerlin, K. C.; Kocha, S. S.; Pivovar, B. S.
53
54 24 Oxidation of Platinum Nickel Nanowires to Improve Durability of Oxygen-Reducing
55
56 25 Electrocatalysts. J. Electrochem. Soc. 2016, 163, F296–F301.
57
58
26 (73) Gubbins, K. E.; Walker, R. D. J. The Solubility and Diffusivity of Oxygen in Electrolytic
59
60 59
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 60 of 75

1
2
3 1 Solutions. J. Electrochem. Soc. 1965, 112, 469–471.
4
5 2 (74) Wakabayashi, N.; Takeichi, M.; Itagaki, M.; Uchida, H.; Watanabe, M. Temperature-
6
7 3 Dependence of Oxygen Reduction Activity at a Platinum Electrode in an Acidic Electrolyte
8
9
10 4 Solution Investigated with a Channel Flow Double Electrode. J. Electroanal. Chem. 2005,
11
12 5 574, 339–346.
13
14 6 (75) Monteverde Videla, A. H. A.; Sebastian, D.; Vasile, N. S.; Osmieri, L.; Aricò, A. S.; Baglio,
15
16 7 V.; Specchia, S. Performance Analysis of Fe-N-C Catalyst for DMFC Cathodes: Effect of
17
18
8 Water Saturation in the Cathodic Catalyst Layer. Int. J. Hydrogen Energy 2016, 41, 22605–
19
20
21 9 22618.
22
23 10 (76) Strmcnik, D. S.; Rebec, P.; Gaberscek, M.; Tripkovic, D.; Stamenkovic, V.; Lucas, C.;
24
25 11 Markovic, N. M. Relationship between the Surface Coverage of Spectator Species and the
26
27 12 Rate of Electrocatalytic Reactions. J. Phys. Chem. C 2007, 111, 18672–18678.
28
29
30 13 (77) Shinagawa, T.; Garcia-Esparza, A. T.; Takanabe, K. Insight on Tafel Slopes from a
31
32 14 Microkinetic Analysis of Aqueous Electrocatalysis for Energy Conversion. Sci. Rep. 2015, 5,
33
34 15 13801.
35
36 16 (78) Holewinski, A.; Linic, S. Elementary Mechanisms in Electrocatalysis: Revisiting the ORR
37
38
17 Tafel Slope. J. Electrochem. Soc. 2012, 159, H864–H870.
39
40
41 18 (79) Ramaswamy, N.; Mukerjee, S. Fundamental Mechanistic Understanding of Electrocatalysis
42
43 19 of Oxygen Reduction on Pt and Non-Pt Surfaces: Acid versus Alkaline Media. Adv. Phys.
44
45 20 Chem. 2012, 2012, 1–17.
46
47 21 (80) Ramaswamy, N.; Mukerjee, S. Influence of Inner- and Outer-Sphere Electron Transfer
48
49
50 22 Mechanisms during Electrocatalysis of Oxygen Reduction in Alkaline Media. J. Phys. Chem.
51
52 23 C 2011, 115, 18015–18026.
53
54 24 (81) Osmieri, L.; Monteverde Videla, A. H. A.; Specchia, S. Activity of Co–N Multi Walled
55
56 25 Carbon Nanotubes Electrocatalysts for Oxygen Reduction Reaction in Acid Conditions. J.
57
58
26 Power Sources 2015, 278, 296–307.
59
60 60
ACS Paragon Plus Environment
Page 61 of 75 The Journal of Physical Chemistry

1
2
3 1 (82) Gojković, S. L.; Gupta, S.; Savinell, R. F. Heat-Treated iron(III) Tetramethoxyphenyl
4
5 2 Porphyrin Chloride Supported on High-Area Carbon as an Electrocatalyst for Oxygen
6
7 3 Reduction: Part III. Detection of Hydrogen-Peroxide during Oxygen Reduction. J.
8
9
10 4 Electroanal. Chem. 1999, 462, 63–72.
11
12 5 (83) Lai, L.; Potts, J. R.; Zhan, D.; Wang, L.; Poh, C. K.; Tang, H.; Gong, H.; Shen, Z.; Lin, J.;
13
14 6 Ruoff, R. S. Exploration of the Active Center Structure of Nitrogen-Doped Graphene-Based
15
16 7 Catalysts for Oxygen Reduction Reaction. Energy Environ. Sci. 2015, 5, 7936–7942.
17
18
8 (84) Zhou, R.; Zheng, Y.; Jaroniec, M.; Qiao, S.-Z. Determination of the Electron Transfer
19
20
21 9 Number for the Oxygen Reduction Reaction: From Theory to Experiment. ACS Catal. 2016,
22
23 10 6, 4720–4728.
24
25 11 (85) Jaouen, F. O2 Reduction Mechanism on Non-Noble Metal Catalysts for PEM Fuel Cells.
26
27 12 Part II: A Porous-Electrode Model To Predict the Quantity of H2O2 Detected by Rotating
28
29
30 13 Ring-Disk Electrode. J. Phys. Chem. C 2009, 113, 15433–15443.
31
32 14 (86) Hsueh, K.-L.; Chin, D.-T.; Srinivasan, S. Electrode Kinetics of Oxygen Reduction: A
33
34 15 Theoretical and Experimental Analysis of the Rotating Ring-Disc Electrode Method. J.
35
36 16 Electroanal. Chem. 1983, 153, 79–75.
37
38
17 (87) Masa, J.; Zhao, A.; Wei, X.; Muhler, M.; Schuhmann, W. Metal-Free Catalysts for Oxygen
39
40
41 18 Reduction in Alkaline Electrolytes: Influence of the Presence of Co, Fe, Mn and Ni
42
43 19 Inclusions. Electrochim. Acta 2014, 128, 271–278.
44
45 20 (88) Georgescu, N. S.; Jebaraj, a. J. J.; Scherson, D. A Critical Assessment of Formula as a
46
47 21 Figure of Merit for Oxygen Reduction Electrocatalysts in Aqueous Electrolytes. ECS
48
49
50 22 Electrochem. Lett. 2015, 4, F39–F42.
51
52 23 (89) Jameson, G. B.; Ibers, J. A. Biological and Synthetic Dioxygen Carriers. In Bioinorganic
53
54 24 Chemistry; 1994; pp 167–252.
55
56 25 (90) Degtyarenko, I.; Nieminen, R. M.; Rovira, C. Structure and Dynamics of Dioxygen Bound to
57
58
26 Cobalt and Iron Heme. Biophys. J. 2006, 91, 2024–2034.
59
60 61
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 62 of 75

1
2
3 1 (91) Kattel, S.; Wang, G. A Density Functional Theory Study of Oxygen Reduction Reaction on
4
5 2 Me–N4 (Me = Fe, Co, or Ni) Clusters between Graphitic Pores. J. Mater. Chem. A 2013, 1,
6
7 3 10790–10797.
8
9
10 4 (92) Holby, E. F.; Zelenay, P. Linking Structure to Function: The Search for Active Sites in Non-
11
12 5 Platinum Group Metal Oxygen Reduction Reaction Catalysts. Nano Energy 2016, 29, 54–64.
13
14 6 (93) Jaouen, F.; Lefèvre, M.; Dodelet, J.-P.; Cai, M. Heat-Treated Fe/N/C Catalysts for O2
15
16 7 Electroreduction: Are Active Sites Hosted in Micropores? J. Phys. Chem. B 2006, 110,
17
18
8 5553–5558.
19
20
21 9 (94) Lefèvre, M.; Proietti, E.; Jaouen, F.; Dodelet, J.-P. Iron-Based Catalysts with Improved
22
23 10 Oxygen Reduction Activity in Polymer Electrolyte Fuel Cells. Science 2009, 324, 71–74.
24
25 11 (95) Liu, K.; Lei, Y.; Wang, G. Correlation between Oxygen Adsorption Energy and Electronic
26
27 12 Structure of Transition Metal Macrocyclic Complexes. J. Chem. Phys. 2013, 139, 204306.
28
29
30 13 (96) Zhang, Z.; Yang, S.; Dou, M.; Liu, H.; Gu, L.; Wang, F. Systematic Study of Transition-
31
32 14 Metal (Fe, Co, Ni, Cu) Phthalocyanines as Electrocatalysts for Oxygen Reduction and Their
33
34 15 Evaluation by DFT. RSC Adv. 2016, 6, 67049–67056.
35
36 16
37
38
17
39
40
41 18
42
43 19
44
45 20
46
47 21
48
49
50 22
51
52 23
53
54 24 Table of Contents Image
55
56
57
58
59
60 62
ACS Paragon Plus Environment
Page 63 of 75 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 1
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 63
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 64 of 75

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
41
42
43
44
45
46
Figure 1. (a) SV of Fe-N-C catalyst in O2-saturated 0.5 M H2SO4 at 900 rpm with different catalyst loadings
47
and NCR = 0.2. (b) Tafel plot obtained from (a) after mass-transport correction and normalization per mass
48 of catalyst on the electrode.
49
50 105x152mm (300 x 300 DPI)
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 65 of 75 The Journal of Physical Chemistry

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
41
42
43
44
45
46
Figure 2. (a) CV of Fe-N-C catalyst with an ink prepared with different NCR recorded in N2-saturated 0.5 M
47
H2SO4 solutions at 10 mV s−1 with a catalyst loading of 0.64 mg cm−2. (b) SV of Fe-N-C catalyst with an
48 ink prepared with different NCR in O2-saturated 0.5 M H2SO4 at 900 rpm. (c) Tafel plot obtained from
49 polarization curves in (b) after mass-transport correction.
50
51 105x233mm (300 x 300 DPI)
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 66 of 75

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
41
42
43
44
45
46
Figure 3. (a) SV of the Me-N-C catalysts (loading 0.64 mg cm−2) measured in O2-saturated 0.5 M H2SO4 at
47
900 rpm. The SV of a commercial Pt/C catalyst measured in 0.1 M HClO4 in both cathodic and anodic scan
48 directions is also shown for comparison. (b) Plot of the linear zones of the potential vs logarithm of the
49 mass-transport corrected current densities from (a).
50
51 105x153mm (300 x 300 DPI)
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 67 of 75 The Journal of Physical Chemistry

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
41
42
43
44
45
46
Figure 4. Mass transport corrected Tafel plots for Fe-N-C, Co-N-C, and Cu-N-C catalysts: (a) Current density
47
referred to the geometric area of the electrode; (b) Current density referred to the mass of catalyst
48 deposited on the electrode.
49
50 104x149mm (300 x 300 DPI)
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 68 of 75

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. CV recorded at 10 mV s−1 in electrolyte saturated with O2 (black curve) and N2 (red curve), and
34 their subtraction (blue curve), peak current density of the subtracted CV in function of the square root of
35 scan sate, and peak potential of the subtracted CV in function of the logarithm of scan rate. (a) Fe-N-C; (b)
36 Co-N-C; (c) Cu-N-C.
37
38 210x164mm (300 x 300 DPI)
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 69 of 75 The Journal of Physical Chemistry

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 6. Differential CV recorded in O2-saturated 0.5 M H2SO4 at different scan rates; FWHM of the
34 differential CV peaks vs. the square root of the scan rate; cathodic transfer coefficient for the reduction of
35 adsorbed O2 calculated from the FWHM of the differential CV peaks widths for (a) Fe-N-C, (b) Co-N-C, and
36 (c) Cu-N-C catalysts.
37
38 210x166mm (300 x 300 DPI)
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
The Journal of Physical Chemistry Page 70 of 75

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
41
42
43 Figure 7. Tafel plot in the kinetic zone of the polarization curve recorded in 0.5 M H2SO4 at 900 rpm with
44 different O2 partial pressures at 25°C, and double-logarithmic plots of mass transport corrected current
density at different potentials in function of O2 concentration for (a) Fe-N-C, (b) Co-N-C, and (c) Cu-N-C
45 catalysts.
46
47 210x226mm (300 x 300 DPI)
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 71 of 75 The Journal of Physical Chemistry

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
41
42
43 Figure 8. SV recorded at 900 rpm in O2-saturated H2SO4 solutions at different pH, and the respective Tafel
44 plots after mass-transport correction for (a) Fe-N-C, (b) Co-N-C, and (c) Cu-N-C catalysts.
45
46 210x230mm (300 x 300 DPI)
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 72 of 75

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
41
42 Figure 9. Effect of the temperature on ORR activity in RDE experiments in O2-saturated 0.5 M H2SO4 at 900
43 rpm. Tafel plot after correction for mass-transport and O2 solubility at different temperatures, and the
44 respective Arrhenius plot at different potentials for (a) Fe-N-C, (b) Co-N-C, and (c) Cu-N-C catalysts.
45
46 210x224mm (300 x 300 DPI)
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 73 of 75 The Journal of Physical Chemistry

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
41
42
43 Figure 10. LSV recorded at 5 mV s−1 scan rate and different rotation speeds in O2-saturated 0.5 M H2SO4
44 after background capacitive current subtraction, and the corresponding K−L plots at different potentials in
45 the diffusion-limited region for (a) Fe-N-C, (b) Co-N-C, and (c) Cu-N-C catalysts.
46
209x227mm (300 x 300 DPI)
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 74 of 75

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
41
42
43
44
45
46
Figure 11. (a) LSV at 1600 rpm in N2-saturated 0.5 M H2SO4 + 1 mM H2O2 and in O2-saturated 0.5 M
47
H2SO4 for Fe-N-C, Co-N-C and Cu-N-C catalysts. (b) H2O2 molar generation calculated from RRDE test
48 performed in 0.5 M H2SO4 for Fe-N-C, Co-N-C, and Cu-N-C and in 0.1 M HClO4 for the Pt/C catalyst.
49
50 105x152mm (300 x 300 DPI)
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 75 of 75 The Journal of Physical Chemistry

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 57x47mm (300 x 300 DPI)
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

You might also like