Subscriber access provided by NAGOYA UNIV
Remediation and Control Technologies
Defect Sites in Ultrathin Pd Nanowires Facilitate the Highly Efficient Electrochemical Hydrodechlorination of Pollutants by H
*ads
Rui Liu, Huachao Zhao, Xiaoyu Zhao, Zuoliang He, yujian Lai, Wanyu Shan, Deribachew Bekana, Gang Li, and Jing-fu Liu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b02740 • Publication Date (Web): 01 Aug 2018 Downloaded from http://pubs.acs.org on August 2, 2018
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 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 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.
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 32
Environmental Science & Technology
1
Defect Sites in Ultrathin Pd Nanowires Facilitate the
2
Highly Efficient Electrochemical
3
Hydrodechlorination of Pollutants by H*ads
4
Rui Liu,*a Huachao Zhao,a,b Xiaoyu Zhao,b Zuoliang He,a,c Yujian Lai, a,c Wanyu Shan, a, c
5
Deribachew Bekana, a, c Gang Li,a and Jingfu Liua
6
a
7
Eco-Environmental Sciences, Chinese Academy of Sciences, Beijing 100085, China
8
b
9
Technology, Tianjin 300457, China
10
c
State Key Laboratory of Environmental Chemistry and Ecotoxicology, Research Center for
College of Chemical Engineering and Materials Science, Tianjin University of Science and
University of Chinese Academy of Sciences, Beijing 100049, China
11 12
ACS Paragon Plus Environment
1
Environmental Science & Technology
13
Page 2 of 32
ABSTRACT
14
Adsorbed atomic H (H*ads) facilitates indirect pathways play a major role in the
15
electrochemical removal of various priority pollutants. It is crucial to identify the atomic sites
16
responsible for the provision of H*ads. Herein, through a systematic study of the distribution of
17
H*ads on Pd nanocatalysts with different sizes and, more importantly, deliberately controlled
18
relative abundance of surface defects, we uncovered the central role of defects in the provision of
19
H*ads. Specifically, the H*ads generated on Pd in an electrochemical process increased markedly
20
upon introducing defect sites by changing the morphology to ultrathin polycrystalline Pd
21
nanowires (NWs), while dramatically reduced upon decreasing the number of surface defects
22
through an annealing treatment. Benefiting from a proportion of H*ads up to 40% of the total H*
23
species, the Pd NWs showed an electrochemical active surface area normalized rate constant of
24
13.8 ± 0.8 h-1 m-2, which is 8-9 times higher than its Pd/C counterparts. The pivotal role of
25
defect sites for the generation of H*ads was further verified by blocking such sites with Rh and Pt
26
atoms, while theoretical calculation also confirms that the adsorption energy of H*ads on these
27
sites is much higher than that on the Pd{111} facet.
28
KEYWORDS:
29
Nanowire Catalyst, Hydrogen Evolution Reaction, Active Center, Indirect Pathway, Halogenated
30
Organic Pollutants.
31
ACS Paragon Plus Environment
2
Page 3 of 32
32
Environmental Science & Technology
INTRODUCTION
33
The removal of environmental priority pollutants and disinfection byproducts with high
34
efficiency under mild conditions through electrocatalytic reduction is an ever-growing area of
35
research.1-11 Unlike the chemical or biological reduction process in which a reductive substance,
36
e.g., H2, functions as the electron donator, the electrocatalytic reduction process is reagent free,
37
and the electron itself reduces the target pollutant at the cathode surface through either a direct or
38
an indirect pathway. Direct reduction occurs by electron tunneling or the formation of a
39
chemisorption complex between the pollutant and cathode material, while in the indirect
40
pathway, the electron primarily reduces a proton to form a surface-adsorbed atomic H (H*)
41
species in a step known as the Volmer process.12 The indirect pathway usually occurs at a low
42
overpotential, which is crucial for effectively suppressing side processes such as the hydrogen
43
evolution reaction (HER), and, therefore, shows higher electron utilization and Faradaic
44
efficiency over the direct pathway. Naturally, the electrochemical reduction process dominated
45
or enhanced by the indirect pathway has drawn increasing research focus.4, 6, 8 In this regard, Pd
46
is the most favorable electrocatalyst for indirect electrochemical reduction,13 not only because of
47
its high efficiency in taking up protons to generate H* species at low overpotentials but also for
48
its ability to retain the H* species via adsorption onto the Pd surface (H*ads) and absorption to the
49
Pd atoms through the formation of Pd-H bonds (H*abs).14 However, in addition to the large
50
portion of formed H* species that are converted to hydrogen molecules through the Heyrovsky
51
(formation of H2 molecule by the reaction between electron, proton and H*ads) or Tafel
52
(combination of two H*ads into one H2 molecule) step,12 which lowers the decontamination
53
efficiency, recent work by Jiang et al. revealed that H*ads, which only accounts for less than 10%
54
of the atomic H, is the sole active species for the reduction of halogenated pollutants.14
ACS Paragon Plus Environment
3
Environmental Science & Technology
Page 4 of 32
55
Nevertheless, highly efficient indirect reduction of a pollutant can be achieved either by the
56
judicious selection of the operating potential14 or by using other materials that show an improved
57
surface/subsurface binding energy for H*ads.6 Thus, further identification of atomic Pd sites for
58
the generation/retainment of H*ads15 and development of new materials/strategies that effectively
59
remove hazardous pollutants electrochemically under mild conditions are highly desirable.
60
Recently, two separate studies highlighted the influence of defect sites on the
61
generation/retainment of atomic H* and hydrogen diffusion behavior.6, 16 Liu et al. reported that
62
the formation of H2 was effectively suppressed by introducing defects into TiO2, and, as a result,
63
highly efficient electrochemical reduction of nitrophenol was achieved.6 In addition, dislocations
64
in Pd crystals were observed to effectively trap H atoms.16 Thus, we propose that defect sites are
65
responsible for the stabilization of H*ads and that the performance of electrochemical
66
decontamination mediated by the indirect pathway can be enhanced by increasing the density of
67
such sites in Pd and related catalysts. To verify this hypothesis, we synthesized Pd/C catalysts
68
with different sizes (from 2 to 6 nm) but similar degrees of defects and deliberately reduced the
69
amount of defects by annealing Pd/C at elevated temperature. Moreover, as a strategy for
70
maximizing the relative abundance of defects, the Pd nanoclusters were allowed to interconnect
71
to form polycrystalline ultrathin nanowires (NWs), during which a large portion of the Pd facets
72
were transformed into defect structures, such as stack faults (SFs, interruption of the normal
73
stacking sequence of the atomic planes) and twin boundaries (TBs, atoms on either side of a
74
plane are mirror images of each other with a 141° angle for the {111} facets).17-19 The cyclic
75
voltammetric (CV) behavior of the synthesized Pd catalysts in Na2SO4 and 2,4-dichlorophenol
76
(2,4-D) provided clear evidence that the relative abundance of defect sites in the Pd catalyst was
77
the key factor that determines their ability to suppress the HER process and the amount of H*ads
ACS Paragon Plus Environment
4
Page 5 of 32
Environmental Science & Technology
78
retained. Benefiting from the excellent H*ads generation/retainment capacity, the Pd NWs showed
79
a higher 2,4-D removal efficiency with a six-times-lower Pd loading. Furthermore, we elegantly
80
deposited Pt and Rh atoms, which exhibit high catalytic activity at defective sites but have low
81
H*ads stabilizing capacity. The dramatically decreased amount of H*ads and the markedly reduced
82
2,4-D removal efficiency unambiguously revealed the central role of the defect sites rather than
83
the intrinsic catalytic activity of Pd NWs providing the primary contribution to the 2,4-D
84
electrochemical reduction process. Finally, we calculated the adsorption energy diagram for H*ads
85
on the Pd {111} facet and defect site and found that H*ads adsorbed on defect sites are much
86
more stable over its counterparts on the Pd {111} facet in good agreement with our experiment
87
data.
88
EXPERIMENTAL SECTION
89
Synthesis of Pd/C catalysts. Pd/C catalysts with different diameters were synthesized by
90
following the reported method with necessary modification.20-21 Specifically, the 10 wt% Pd/C
91
catalysts were synthesized in the presence of ethylenediaminetetraacetic acid (EDTA), 2-fold, 5-
92
fold or no NaOH, and referred to as Pd/C-EDTA, Pd/C-NaOH-2, Pd/C-NaOH-5 and Pd/C-
93
NaOH-0, respectively. Pd/C-NaOH-2 was annealed at 400°C for 2 hours under a continuous N2
94
flow to increase the crystallinity, and the resulting material was denoted Pd/C-NaOH-2-400°C.
95
For the preparation of Pd/C electrocatalyst ink, Pd/C catalyst was dispersed in a mixture of
96
deionized water, isopropyl alcohol, and Nafion (v/v/v = 4/1/0.05) under sonication for 15 min,
97
with a final Pd concentration of 5.35 mg mL−1.
98
Synthesis of Pd, Pd@Rh and Pd@Pt NWs. Pd NWs with a diameter of ~2 nm were
99
synthesized using our previously developed protocol, which referred to the reduction of
100
Pd(NO3)3 by a proper amount of KBH4 with Triton X-114 as a stabilizer and structure director.17
ACS Paragon Plus Environment
5
Environmental Science & Technology
Page 6 of 32
101
For the preparation of Pd@Rh or Pd@Pt core–shell NWs, 0.1, 0.2, 0.5, 1.0, or 2.0 mL of ice-cold
102
(stored in ice bath for 30 min before use) 1.0 mM RhCl3 (or H2PtCl6) solution was added
103
dropwise into 10.0 mL of freshly synthesized Pd NWs under stirring.22-23
104
Electrode preparation. The working electrode was prepared by loading the desired amount
105
of Pd catalyst onto a glassy carbon electrode (GCE, 5.0 mm inner diameter, id) or carbon paper
106
(Toray 090, with a thickness of 280 µm and a porosity of 0.78, pretreated according to the
107
reported method14). A 50 µl aliquot of a Pd NW (0.05 µmol Pd) or Pd@Pt/Pd@Rh NW
108
dispersion containing 0.05 µmol Pd was drop cast onto a GCE as a supportless electrocatalyst.
109
This was followed by adding 10 µl of a Nafion solution (0.5% w/v in ethanol, DuPont). For Pd/C
110
electrocatalyst, a 10.0-µl aliquot of Pd/C ink was modified on the GCE. For electrochemical
111
hydrodechlorination (EHDC) of 2,4-D, the Pd catalysts were modified on carbon paper, and the
112
amount of Pd in Pd/C was increased to 2.0 mg. Whereas for Pd NWs, the amount of Pd loaded
113
was 0.36 and 2.0 mg, with the Pd NWs-modified carbon paper washed three times with acetone
114
to remove the residual Triton X-114 before the addition of Nafion. The prepared electrode was
115
activated in a N2-saturated 0.1 M HClO4 solution under cyclic voltammetry at −0.25 to 1.0 V. Pt
116
wire and a Ag/AgCl electrode were used as the counter and reference electrodes, respectively. To
117
determine the mass content of the Pd catalyst, the catalyst dispersion was digested with aqua
118
regia, filtered and then analyzed by inductively coupled plasma-mass spectrometry (ICP-MS,
119
Agilent 7700).
120
CO-stripping experiment. The electrochemical active surface area (EASA) for the Pd
121
catalysts was estimated by a CO-stripping experiment, which was also utilized to demonstrate
122
the presence of Pt/Rh atoms on Pd NWs. After CO was chemisorbed onto the surface metal
123
atoms, it was then stripped at 50 mV·s-1, with the Pd surface area (SAPd) calculated by:21
ACS Paragon Plus Environment
6
Page 7 of 32
Environmental Science & Technology
124
PdEASA=QCO-stripping/420 (mC cm−2)
125
EHDC experiments. The EHDC of 2,4-D was performed in a two-compartment
126
electrochemical cell separated by a proton-exchange membrane (Nafion-117). Ag/AgCl (3.0 M
127
KCl) and Pt foil were utilized as the reference and counter electrodes, respectively. During the
128
EHDC process, 50 mg·L−1 2,4-D (0.31 mM) in N2-saturated 50 mM Na2SO4 was electrolyzed at
129
a constant potential, the residual 2,4-D and the generated 2-chlorophenol (2-CP)/4-chlorophenol
130
(4-CP) and phenol were analyzed at specific time points by high-performance liquid
131
chromatography (HPLC). The EHDC experiment was also performed in the presence of acetic
132
acid/sodium acetate buffer with different initial pH to evaluate the effect of the solution pH on
133
the EHDC process. To investigate the 2,4-D removal performance of the Pd NWs under more
134
environmentally relevant conditions, 2.5 and 1.0 mg L-1 2,4-D were spiked into five real water
135
samples collected from the Yellow Sea, the Olympic Green park, the Qunyu River, Qunming
136
Lake and Beijing tap water.
137
All reported data are mean values from at least three parallel experiments. The experimental
138
uncertainties, including instrumental errors and relative standard deviations, and blank sorption
139
were assessed in the absence of catalyst. The results showed that the total uncertainty was less
140
than 3.0% for 2,4-D with high concentration, which slightly increased to 5~7% with the decrease
141
of their initial concentration to 2.5 and 1.0 mg L-1.
142
Simulation methodology. First-principles density functional theory calculations are carried
143
out using the Vienna ab initio simulation package (VASP v.5.4.1) to examine the adsorption
144
energy of the H atom on the free surface, vacancy, stack fault, and twin boundary. Throughout
145
the calculation, the generalized gradient approximation (GGA) and the projector augmented
146
wave (PAW) pseudopotentials with the exchange and correlation in the Perdew-Burke-Ernzerhof
ACS Paragon Plus Environment
7
Environmental Science & Technology
Page 8 of 32
147
(PBE) are employed to calculate the total energy.24-25 A cut-off energy of 300 eV is used for the
148
plane-wave basis. The bulk of Pd is first built and optimized using the Monkhorst-Pack 10 × 10
149
× 10 k-pointing mesh. The lattice parameter for the optimized Pd unit cell is 3.94 Å. Based on
150
the optimized Pd unit cell, the perfect plane, the plane with vacancy, the plane with stack fault,
151
and the plane with twin boundary are built respectively using the slab model with a 2 nm thick
152
vacuum layer added along the Z direction. During all calculations for the Pd planes, the
153
Monkhorst-Pack 3 × 5 × 1 k-pointing mesh is used. The adsorption energy of H on the plane is
154
calculated as:
155
Eads = E(H) + EPd-site – EPd-site + H26
156
Characterization. X-ray diffraction (XRD) patterns were recorded on an X-ray diffractometer
157
with Cu Kα radiation in the 2ϴ range of 30-90°. Transmission electron microscopy (TEM) and
158
high-resolution TEM (HRTEM) images were obtained on a JEM-2100F instrument, while
159
spherical aberration corrected scanning transmission electron microscopy (Cs-STEM) and
160
energy-dispersive X-ray spectroscopy (EDS) mapping were performed on a JEM-ARM200F
161
system (JEOL, Japan). Electrochemistry experiments were carried out with an electrochemical
162
workstation (CHI 852C, Chenhua Co., China).
163
RESULTS AND DISCUSSION
164
Pd catalyst characterization. Figure 1a displays TEM images of the Pd catalysts, which
165
revealed varying sizes for the Pd NPs that are well-dispersed on the surface of carbon black and
166
Pd NWs. Although we cannot quantitatively describe the amount of surface defects on the Pd/C
167
catalysts, it is reasonable to assume that they have similar degrees of defects since they were
168
synthesized using the same method. Note that although large NPs with diameters of up to 15 nm
ACS Paragon Plus Environment
8
Page 9 of 32
Environmental Science & Technology
169
can be observed in Pd/C-NaOH-2-400°C, which is attributed to the sintering of neighboring Pd
170
particles during the annealing process, the majority of NPs have diameters of less than 5 nm
171
(Figure S1-6). The XRD patterns for the Pd samples shown in Figure 1b (marked with a red star)
172
are consistent with the Pd standard (PDF#46-1043). The peak broadening analysis (based on the
173
Scherrer equation) and the TEM observations show that the sizes of the Pd NWs, Pd/C-EDTA,
174
Pd/C-NaOH-2, Pd/C-NaOH-5 and Pd/C-NaOH-0 samples are 2.38 ± 0.44, 2.24 ± 0.41, 2.51 ±
175
0.37, 4.47 ± 0.90 and 6.27 ± 0.80 nm (Figure 1c), respectively, while the average size of Pd/C-
176
NaOH-2-400°C is approximately 4.85 ± 1.48 nm. In accordance with the change in the size of
177
the Pd catalysts, their surface area also changed. Estimated from CO-stripping, the EASA of
178
Pd/C-EDTA, Pd/C-NaOH-2, Pd/C-NaOH-5 and Pd/C-NaOH-0 are 114.3, 93.9, 75.1 and 53.1 m2.
179
g-1, respectively. For the partial sintering of Pd particles during the annealing process, the EASA
180
of Pd/C-NaOH-2-400°C is largely reduced to 65.8 m2. g-1. However, this value is still 20% larger
181
than that of Pd/C-NaOH-0, which is in good agreement with the smaller diameter of Pd/C-
182
NaOH-2-400°C. On the other hand, the EASA of Pd NWs is also as large as 107.4 m2. g-1.
183
The HRTEM image shown in Figure S1-6 displays a lattice spacing of ∼0.22 nm for the
184
synthesized Pd nanocatalysts, which matches well with the {111} atomic planes of face-
185
centered-cubic (fcc) Pd. Moreover, since the Pd/C catalysts were synthesized using chemical
186
reduction in an ice bath, rich defect sites such as SFs and TBs were frequently observed in their
187
HRTEM images (Figure S1-4). However, after annealing at 400°C for two hours, the majority of
188
the defect sites were removed, and almost all of the Pd NPs became single crystalline (Figure S5).
189
On the other hand, dense defect sites were distributed on the ultrathin Pd NWs, which formed by
190
oriented attachment of nascent Pd clusters,17 with SFs (blue arrow), TBs (red arrow) as well as
191
lattice defects (green arrow) observed across the whole NW (Figure 1d). Interestingly, the
ACS Paragon Plus Environment
9
Environmental Science & Technology
Page 10 of 32
192
nonionic surfactant Triton X-114 utilized in the synthesis of the Pd NWs also facilitated the
193
uniform distribution of the Pd NWs on carbon paper as a support-free electrocatalyst (Figure 1e).
194
Enhanced H*ads provision capacity of Pd NWs. The H*ads provision capacities of the
195
different Pd catalysts were first evaluated by CV in 50 mM Na2SO4 by varying starting potentials
196
(from -0.65 to -1.15 V) but with a fixed ending potential of 0.45 V.14 As in the previous study,14
197
upon negatively shifting the starting potential, broad peaks associated with the oxidation of
198
molecular H2 and H*ads successively emerged at approximately -0.80 to -0.60 V and -0.10 to -
199
0.00 V (Figure S8-13). For the Pd/C catalysts, when the starting potential was lowered to -0.95 V,
200
the oxidation peak for H*abs emerged at ~-0.2 V, with its height rapidly increasing at the expense
201
of the fast attenuation of the first two peaks. However, for the Pd NWs, besides the largely
202
suppressed H2 oxidation peak, no apparent H*abs oxidation peak was observed at any potential
203
except for a single broad oxidation peak at ~0 V, which is a direct result of the enhanced H*ads
204
generation/storage capacity of the Pd NWs. The distribution of the different H* species on the Pd
205
electrocatalysts at a starting potential of -1.05 V is highlighted in Figure 2a. Unlike the dominant
206
oxidation peak for the H*ads observed for the Pd NWs at approximately 0 V, for all the Pd/C
207
catalysts, regardless of their diameter, the main peak was located at ~-0.2 V with a small
208
shoulder at ~0 V. In addition to the pronounced H*ads oxidation peak observed for the Pd NWs,
209
the hydrogen evolution capacity of the Pd NWs was substantially weakened. Both the low
210
cathodic current observed in Na2SO4 (Figure S14) and the high overpotential of -0.326 V (vs
211
reversible hydrogen electrode, RHE) in 0.5 M H2SO4 (Figure S15-16) indicated an elevated
212
activation energy for the HER process on the Pd NWs and the high binding affinity for H*ads on
213
the Pd NWs. Moreover, the evident reduction peak in the backward scan reveals the reversibility
214
of the H*ads reduction/oxidation cycle on defect sites.
ACS Paragon Plus Environment
10
Page 11 of 32
Environmental Science & Technology
215
To verify that this observed H oxidation peak is associated with H*ads and that this H* species is
216
active in the reduction of 2,4-D, the CV experiments were also performed in the presence of 50
217
mg·L-1 2,4-D. As expected, the H*ads oxidation peak was completed quenched by 2,4-D, while
218
the H2 and H*abs peaks remained nearly unchanged (Figure 2b). Moreover, from the backward
219
scan data shown in Figure 2b and S12, the H* reduction peak was almost unaffected following
220
the addition of 2,4-D. If the disappearance of H*ads is the direct result of coadsorption of 2,4-D on
221
the H*ads adsorption site, we should have observed a significant decrease in the H* reduction peak.
222
Otherwise, the unchanged H* reduction peak with the addition of 2,4-D rules out the possibility
223
of 2,4-D coadsorption with H*ads on Pd sites, which is very crucial for Pd NWs to continuously
224
and effectively provide H*ads for the electroreduction of 2,4-D.
225
Based on the different oxidation behavior, the amount of the above-mentioned hydrogen
226
species was estimated from the oxidation charge. Note the partial overlap of the H*ads and H*abs
227
oxidation peak – the amount of H*abs is deduced from the oxidation peak between -0.4 to 0.2 V in
228
the presence of 2,4-D, while that of H*ads is the difference in the oxidation peak in the presence
229
and absence of 2,4-D.14 As is summarized in Figure 2c, the amount of H*ads increased with
230
decreasing starting potential, with the largest amount generated at -1.05 V. Accordingly, the
231
relative abundance of H*ads increased from approximately 10% at -0.7 V to 40% at -1.05 V but
232
was lowered to ~30% with the further decrease of start potential to -1.10 V. In contrast, the
233
relative abundance of H*ads for the Pd/C catalysts was in the range of 8% to 14% and further
234
decreased to 5% for Pd/C-NaOH-2-400°C, strongly supporting the pronounced H* storage
235
capacity of Pd NWs.
236
Pronounced 2,4-D removal capacity of Pd NWs. The dramatically increased amount of H*ads
237
on the Pd NWs is very meaningful for its usage as an advantageous electrocatalyst for the EHDC
ACS Paragon Plus Environment
11
Environmental Science & Technology
Page 12 of 32
238
of 2,4-D. Although showed comparable activity with Pd/C catalysts in the reduction of 2,4-D
239
with H2 as electron donor (Figure S17), Pd NWs displayed superior electrochemical
240
decontamination performance in the constant-potential electrolysis of 50 mg· L-1 2,4-D.
241
Catalyzed by 0.36 mg of Pd NWs for 6 hours, the removal ratio of 2,4-D increased stepwise from
242
5.5 ± 1.4% at -0.65 V to 81.5 ± 4.1% at -1.05 V and largely decreased to 65.4 ± 3.6% with a
243
further negative shift in the applied potential to -1.15 V (Figure 2d). Detailed analysis of the
244
kinetics showed that the reduction was a pseudo-first-order process (Figure 2e), with the largest
245
rate constant of 0.25 h-1 achieved at -1.05 V. Meanwhile, HPLC analysis revealed that phenol is
246
the major product with a good mass balance of 94-108% (Figure S18), indicating the removal of
247
2,4-D mainly occurs via the electrochemical reduction pathway. In contrast to the high 2,4-D
248
removal efficiency of the Pd NWs, the activities of its Pd/C counterparts were much lower; even
249
when the Pd loading was increased to ~2.0 mg, the highest removal ratio was only 66.6 ± 5.6%,
250
which was achieved by the Pd/C-EDTA catalyst with the smallest diameter (Figure 2f and S19).
251
With the gradually increasing size of the Pd NPs, this value decreased to 48.5 ± 4.5% for Pd/C-
252
NaOH-0, which had the largest size of 6.0 nm. Importantly, although the EASA is 20% higher
253
than that of Pd/C-NaOH-0, the 2,4-D removal rate was further lowered to 43.5 ± 5.7% for Pd/C-
254
NaOH-2-400°C.
255
To highlight how the catalysis performance changes with Pd microstructure, the kinetics data
256
for the Pd catalysts in the EHDC process were normalized with EASA (Figure 2g). Interestingly,
257
for all the defect-containing Pd/C catalysts, the normalized rate constants are independent of the
258
sizes of the Pd NPs, but almost remain constant in the range of 1.6~1.8 h-1·m-2. However, for
259
defect-free Pd/C-NaOH-2-400°C, this value drops to 1.03 ± 0.13 h-1·m-2. On the other hand, the
260
normalized rate constant for Pd NWs is as high as 13.8 ± 0.8 h-1·m-2. This 8~9-fold increase in
ACS Paragon Plus Environment
12
Page 13 of 32
Environmental Science & Technology
261
the activity of Pd NWs over defect containing Pd/C and ~14 times more catalytically active than
262
defect-free Pd/C strongly support the superiority of defective sites in the EHDC process, as well
263
as the effectiveness of introducing such sites by interconnecting Pd clusters into nanowires.
264
Meanwhile, besides showing the best EHDC performance, Pd NWs also showed a high current
265
efficiency (CE) of 35.0 ± 2.9% during the whole EHDC process. In comparison, the CE for
266
Pd/C-EDTA, Pd/C-NaOH-2, Pd/C-NaOH-5 and Pd/C-NaOH-0 were decreased to 13.5 ± 1.2,
267
14.9 ± 1.0, 16.6 ± 1.5 and 12.5 ± 1.2%, respectively, with this value further lowered to 8.5 ± 1.0%
268
for Pd/C-NaOH-2-400°C. Furthermore, the CEs from different Pd nanostructures displayed a
269
good linear dependence on their relative abundance of H*ads compared to the total amount of H
270
species (Figure S20), which supports the hypothesis that an H*ads mediated indirect pathway is
271
the main route for the EHDC process. Moreover, since the H*ads is generated from different Pd
272
electrocatalysts, the similarity in their reactivity infers that the studied Pd catalysts showed
273
comparable activity in the EHDC of 2,4-D by H*ads.
274
The critical role of H*ads in the EHDC of 2,4-D. To further demonstrate the critical role of
275
H*ads in the EHDC of 2,4-D, we first confirmed its presence by a 5,5-dimethyl-pyrroline-l-oxide
276
(DMPO) trapping and electron spin resonance (ESR) experiment (Figure 3a).8 Moreover, the
277
apparent linear correlation between the rate constant and the amount of H*ads generated (Figure
278
3a) verifies that H*ads (Figure 3b) is the essential reactive species in the removal of 2,4-D. In
279
addition, the linear relationship also reflects the effective suppression of H2 formation, which
280
adversely affects the reduction of 2,4-D by inhibiting the mass transfer of 2,4-D to the reactive
281
sites.14 This was also verified by introducing tert-butanol (TBA), a previously reported effective
282
H*ads scavenger, into the reaction solution.8 With the addition of, approximately 5.0 mM TBA
283
into the Na2SO4 electrolyte solution, an immediate quenching of the H*ads oxidation peak was
ACS Paragon Plus Environment
13
Environmental Science & Technology
Page 14 of 32
284
observed (Figure 3c). This suppressed H*ads generation markedly changed the EHDC
285
performance, and the rate constant dropped by 38% and 81% in the presence of 1.0 and 10.0 mM
286
TBA, respectively (Figure 3d-e). Meanwhile, the presence of O2, which completes H*ads and
287
reactive sites with 2,4-D also has obvious influence on the EHDC performance of Pd NWs. The
288
EHDC efficiency dropped from 75.6 ± 1.7% for the N2 saturated solution to 35.2 ± 3.2% for the
289
air-saturated solution, and further decreased to 21.6 ± 2.0% when the solution is presaturated
290
with O2 (Figure 3d-e). The above results support the hypothesis that the H*ads-mediated indirect
291
pathway is the dominant mechanism during the EHDC of 2,4-D catalyzed by Pd NWs.
292
On the other hand, if we saturate the solution with H2, the 2,4-D removal rate, especially the
293
initial rate, can be largely increased, both in Pd NWs and Pd/C catalysts (Figure 3 f and g).
294
Specifically, on Pd NWs, the enhanced EHDC appears to be contributed by direct H2 reduction
295
because the 2,4-D removed by the EHDC process in the presence of H2 equals the amount of 2,4-
296
D removed by H2 reduction plus 2,4-D removed by the EHDC process in N2-saturated solution
297
(P> 0.05). Intriguingly, for catalyzed Pd/C-NaOH-2, the increased amount of 2,4-D removed
298
cannot be simply explained by H2 reduction, or the H2 molecule accelerated EHDC process itself
299
(P< 0.05), which again demonstrates that the EHDC of 2,4-D is hindered by the formation of H2
300
in Pd/C.
301
Notably, with the gradual uptake of H+ from the electrolysis solution, the solution pH was
302
quickly increased from 5.86 to more than 11 within 10 min of electrolysis (Figure 3h, i and S21).
303
Therefore, it is imperative to study how the solution pH influences the EHDC process, and more
304
importantly, whether the concentration of H+ is a crucial factor in determining the EHDC
305
kinetics.27 This was done by introducing an acetic acid/sodium acetate buffer into the electrolysis
306
solution. When the initial pH was adjusted to 3.72, 4.77, 6.09, and 7.10 by a buffer solution, the
ACS Paragon Plus Environment
14
Page 15 of 32
Environmental Science & Technology
307
first-order rate constant for the Pd NW catalyzed EHDC process was slightly decreased from
308
0.30 ± 0.02 to 0.27 ± 0.01, 0.24 ± 0.02 and 0.18 ± 0.01 h-1, respectively. This weak pH
309
dependency reflects the fact that the EHDC of 2,4-D with Pd NWs can be performed under a
310
wide pH range. Moreover, this signifies an important factor that the stability of the generated H*
311
rather than the concentration of H+ is the rate-determining step in this process. In contrast,
312
catalyzed by Pd/C, i.e., Pd/C-NaOH-2, the rate constant decreased tenfold with an increase of the
313
initial pH from 3.72 to 7.10. This different pH-dependence of the EHDC kinetics on Pd NWs and
314
Pd/C is attributed to their different H+ or H* utilization efficiency. For the high H* utilization
315
efficiency of Pd NWs (up to 40%), less H* is needed compared to the case of Pd/C, where only
316
10% of H* take part in the EHDC process. It is worth pointing out that unlike the effective
317
control of the solution pH by the buffer, when the initial pH was adjusted to 7.10 with pure
318
sodium acetate, the solution pH also quickly increased to 11.3, similar to the case without buffer,
319
but showed much lower activity. We attributed this low activity to the acetate ion, which
320
competes with the binding site for 2,4-D on Pd nanostructures.
321
Blockage of Pd defect sites by deposition of Pt or Rh atoms. Moreover, to further highlight
322
the central role of defect sites in the EHDC process and to rule out the possibility that the
323
enhanced 2,4-D EHDC efficiency of Pd NWs stems from its intrinsic high catalytic
324
performance,28 we took advantage of the high surface energy of defective sites and selectively
325
deposited Pt and Rh atoms on them.22-23 As demonstrated in our previous works,22-23, 29 unlike the
326
formation of an atomic layer of Ag, Pd or Pt overlayer on Au NWs with a smooth surface and
327
low surface energy,29 the deposition of metal atoms on defect-rich Pd/Pt NWs initially occurs at
328
the high-energy sites, e.g., defective sites and the {100}/{110}facet,30 which results in the
329
formation of nanoislands (Figure S22-24).22 On the other hand, the deposited Pt/Rh atoms are
ACS Paragon Plus Environment
15
Environmental Science & Technology
Page 16 of 32
330
highly active for the reductive cleavage of carbon-halogen bonds (Figure S17), including C-F
331
bonds with extraordinarily high energies,31 but are weak H*ads stabilizers, which can be
332
speculated from the low overpotential for the HER process on Pt and Rh electrocatalysts.26, 32-33
333
Therefore, if the enhanced 2,4-D EHDC performance of Pd NWs is a result of the superior
334
catalysis performance of Pd NWs, the deposited Pt/Rh atoms would increase the 2,4-D removal
335
rate, and vice versa, if the presence of highly catalytic active Rh/Pt results in the decreased
336
EHDC performance, the only explanation is that these atoms lowered the H*ads provision capacity.
337
As is shown in Figure 4a, EDX elemental mapping demonstrated that Pt atoms were enriched
338
on some parts of the Pd@Pt NWs, while benefiting from the large Z-contrast between Pd (Z=46)
339
and Pt atoms (Z=78), the atomic-resolution Cs-STEM image combined with intensity profile
340
analysis (Figure S25) revealed that the Pt atoms were enriched at the defect sites. Furthermore,
341
the appearance of a new CO stripping peak (Figure S26-28) at very low potential, approximately
342
0.25 V (vs Ag/AgCl) reveals the presence of highly catalytic active Pt/Rh atoms on the Pd NWs.
343
Since the CO-stripping peak from Pt NWs is located at ~0.70 V,22 the most plausible explanation
344
for the extremely high activity of Rh/Pt atoms on Pd NWs is that these atoms are located at
345
highly active sites such as defective sites.34-35 In addition, with the stepwise addition of Pt/Rh
346
adatoms, the peak associated with H*ads rapidly decreased and shifted to more negative values
347
(Figure 4b and c). At the same time, evident H2 and H*abs oxidation peaks were observed for
348
Pt/Rh deposition amounts of 5.0% or above. These results again demonstrated that these atoms
349
are primarily located at the defective sites and that their presence released the H*ads species and
350
promoted the HER process. In addition, since Pd@Pt and Pd@Rh NWs displayed the same
351
morphology and were stabilized by the same surfactant (TX-114) with Pd NWs, such large
352
negative-shifts of the oxidation peak upon the deposition of Pt/Rh atom on Pd NWs rule out the
ACS Paragon Plus Environment
16
Page 17 of 32
Environmental Science & Technology
353
possibility that these factors, i.e., morphology or surfactant, are the main parameters influencing
354
the atomic H oxidation potential. Note that as a secondary contribution, the deposited Pt/Rh
355
atoms may also influence the electronic structure of the Pd atoms and change their affinity for
356
H*ads. In accordance with the decreased amount of H*ads, an immediate drop in the EHDC
357
efficiency was also observed. Importantly, as Pt atoms show better HER activity (or a lower
358
H*ads affinity) than Rh atoms, a much faster decrease in the EHDC efficiency was observed for
359
the Pt-modified catalyst (blue line in Figure 4d, e). When the amount of Pt/Rh relative to Pd
360
increased to 20%, the EHDC efficiency drops from 79.5 ± 7.1 to 23.8 ± 3.3 or 28.6 ± 4.1%, again
361
showing the importance of the Pd defect sites in the EHDC process. The presence of highly
362
catalytically active Pt/Rh but markedly decreased EHDC efficiency strongly supports the
363
hypothesis that the superior catalysis performance of Pd NWs is primarily associated with its
364
enhanced H*ads provision capacity.
365
Adsorption energy of H*ads (Eads) on Pd sites. The theory that defective sites possess
366
enhanced H*ads generation/retainment capacity was further supported by a density function
367
theory (DFT) calculation. Since the energy barrier for the Volmer process on Pd surface is very
368
low (down to 0.2 eV on Pd {111} facets)36, we ignore the difference in the amount of H*ads
369
generated on different Pd sites, and focus on the adsorption energy of this specie on Pd{111} and
370
three defective Pd sites, or the retainment capacity of H*ads on these sites. As shown in Figure 5a,
371
on the Pd{111}facet, both the atop- and fcc hollow-adsorption configuration are stable with an
372
Eads of 2.23 and 2.78 eV, while for lattice defect sites, where only the atop configuration is
373
permissible, the Eads increased to 2.45 eV, which is 0.22 eV higher than its counterpart on
374
Pd{111}. The situation is even more evident in the case of the TB and SF sites, where only the
375
fcc hollow-adsorption is stable with an Eads of 2.72 and 2.77 eV, respectively. Therefore,
ACS Paragon Plus Environment
17
Environmental Science & Technology
Page 18 of 32
376
compared with high stability the H*ads on defective sites (especially on TB and SF site), only a
377
portion of H*ads on the Pd{111}, c.a., hollow-adsorbed H*ads, is stable enough and available for
378
the EHDC process.
379
Proposed mechanism. All the above results collectively point out that the defect sites in the Pd
380
NWs enhanced the stability of the generated H*ads, which in turn largely facilities its usage as an
381
electrocatalyst for the EHDC of 2,4-D. Polycrystalline Pd NWs have also been reported to
382
display superior H2 adsorption and dissociation/activation performance over that of Pd
383
cuboctahedra,28 which was attributed to the exposure of more Pd atoms at TB sites in the NWs.
384
In addition, these TB sites also have a high surface energy for the adsorption and catalytic
385
conversion of reactants. On the other hand, the defective-rich TiO2-x with abundant oxygen
386
vacancies or ≡Ti(III) was also shown to enhance electroreduction performance by increasing the
387
H* generation capacity. Meanwhile, these sites also increase the catalytic activity by ameliorating
388
the electron transfer kinetics and reducing electrode polarization.6 Herein, the enhanced EHDC
389
performance of Pd NWs neither result from its high activity (Figure 4, S16), nor the improved
390
electron transfer capacity of Pd NWs. In fact, our electrochemical impedance spectroscopy (EIS,
391
Figure S29) result showed that a Pd NW-modified electrode exhibits a much larger arc radius
392
than the Pd/C-modified electrode, revealing the increased interface impedance and the slowed
393
electron transfer in Pd NWs. Meanwhile, a four-point probe resistance measurement also
394
excluded possible interference from different resistances (Table S1). Therefore, the primary
395
result for the increased EHDC performance of Pd NWs is its superior H*ads generation/retainment
396
capacity.
397
We hypothesize that for the surface Pd atoms with “weak” binding energies, the H*ads generated
398
through the Volmer mechanism are highly mobile, i.e., the atop-adsorbed H*ads on Pd{111} sites
ACS Paragon Plus Environment
18
Page 19 of 32
Environmental Science & Technology
399
quick diffuse into the Pd lattice to form a stable H*abs species or the formation of H2 (HER) by
400
the Tafel or Heyrovsky process. This results in a very low EHDC rate or even a long induction
401
time in which no EHDC occurs (Figure S30). Indeed, an induction time up to 20 min has been
402
observed by Sun et al. in a parallel study during which H*ads was converted into H*abs to saturate
403
the Pd crystalline rather than take part in the EHDC process,37 which results in a low current
404
efficiency of 3.7-6.7%. However, on the defect-rich Pd NWs, the high adsorption energy of H*ads
405
on such sites largely increased the stability of this crucial reactive specie for the effective
406
reduction of 2,4-D by suppressing the formation of H*abs (Figure 2a-c, 5b) and H2 (reflected in
407
the increased onset potential for the HER process(Figure S15).
408
Mechanistically, the reaction kinetic is controlled by both the binding energy of
409
reactants/intermediates/products on the reactive centers and the energy-barrier of the rate-
410
determining step. From this point, both the adsorption energy of the key reactant, e.g., H*ads and
411
the energy barrier of the rate-determining step for the cleavage of C-Cl bond, herein, influences
412
the EHDC performance. However, the presented data, especially the good linearity dependence
413
of the EC and amount of H*ads reveals that the amount of H*ads plays a pivotal role in determining
414
the EHDC process, with the effect of the adsorption energy on the reactivity of H*ads almost
415
insensible in the current study. This is because unlike the chemical reduction process, where a
416
sufficient number of electrons are supplied by H2 or other reductant, in the EHDC process, for
417
the insufficiency of the provided H*ads, the amount of H*ads becomes a crucial factor in
418
determining the reaction path. This is the reason why we focused our primary attention to the
419
stability of H*ads on different Pd atomic sites. From the viewpoint of the whole EHDC process, a
420
detailed calculation for how the adsorption configuration and adsorption energy of H*ads
ACS Paragon Plus Environment
19
Environmental Science & Technology
Page 20 of 32
421
influences the reaction process is needed to obtain a more comprehensive conclusion on the
422
EHDC process on different Pd sites in the future.
423
Environmental implications. The defective Pd NWs are applicable to the decontamination of
424
2,4-D-polluted water through the EHDC process. This is demonstrated by the removal of 2.5 or
425
1.0 mg L-1 of 2,4-D from sea, river, lake or tap water. As presented in Table 1, up to 76.9% of
426
2,4-D in tap water was removed after six hours of electrochemical treatment. Even for the water
427
sample with a complex matrix, i.e., water sampled from the Beijing Olympic Green Park, which
428
is irrigated by recycled sewage, the EHDC efficiency was higher than 50%. It is worth noting
429
that compared with the satisfactory EHDC performance of Pd NWs in freshwater, the removal
430
rate of 2,4-D in sea water is much lower (