You are on page 1of 24

ll

OPEN ACCESS

Article
18.1% single palladium atom catalysts on
mesoporous covalent organic framework for
gas phase hydrogenation of ethylene
Chun-Te Kuo, Yubing Lu,
Pezhman Arab, K. Shamara
Weeraratne, Hani El-Kaderi,
Ayman M. Karim

helkaderi@vcu.edu (H.E.-K.)
amkarim@vt.edu (A.M.K.)

Highlights
18.1 wt % palladium single atoms
are supported on mesoporous
covalent organic framework

Isolated palladium single atoms


are coordinated to identical imine
sites in the framework

Palladium single atoms are stable


and active for gas phase ethylene
hydrogenation

Noble metal single-atom catalysts maximize metal utilization and offer


opportunities to design heterogeneous catalysts at the molecular scale. Kuo et al.
demonstrate the use of mesoporous covalent organic framework to anchor high
loading of palladium as single atoms and showcase their stability and activity for
gas phase ethylene hydrogenation.

Kuo et al., Cell Reports Physical Science 2,


100495
July 21, 2021 ª 2021 The Author(s).
https://doi.org/10.1016/j.xcrp.2021.100495
ll
OPEN ACCESS

Article
18.1% single palladium atom catalysts
on mesoporous covalent organic framework
for gas phase hydrogenation of ethylene
Chun-Te Kuo,1 Yubing Lu,1 Pezhman Arab,2 K. Shamara Weeraratne,2 Hani El-Kaderi,2,*
and Ayman M. Karim1,3,*

SUMMARY
Noble metal single-atom catalysts maximize metal utilization and
offer opportunities to design heterogeneous catalysts at the molec-
ular scale. Mesoporous covalent organic frameworks provide an
ideal support to stabilize metal single atoms with specific ligand
configuration similar to a homogeneous catalyst. In this work, a
high loading of single Pd atoms, 18.1 wt %, on mesoporous imine-
linked covalent organic framework was synthesized, characterized,
and evaluated for ethylene hydrogenation. X-ray photoelectron
spectroscopy, X-ray absorption spectroscopy, and diffuse-reflec-
tance infrared Fourier transform spectroscopy of adsorbed CO
demonstrate that the Pd is atomically dispersed with a highly homo-
geneous local coordination. The Pd single atoms are active for hy-
drogenation of ethylene to ethane at room temperature. The study
demonstrates that mesoporous COFs provide a large number of
identical metal binding sites that are good candidates for immobiliz-
ing metal single atoms and their use in gas-phase catalytic applica-
tions.

INTRODUCTION
Supported metal single atoms,1–6 dimers,7–10 and subnanometer clusters11–15 have
recently drawn significant interest because they exhibit extraordinary catalytic per-
formance due to their unique geometric structure and electronic properties
compared with nanoparticles. Hülsey et al.16 showed that Rh single atoms supported
on phosphotungstic acid have extraordinary activity for CO oxidation even at around
50 C (323 K). We recently reported that Pt single atoms and subnanometer clusters
exhibit remarkably high selectivity for hydrogenation of acetylene to ethylene in
contrast to the low selectivity on large Pt nanoparticles.17 Because of the importance
of palladium as a catalyst for various reactions such as (selective) hydrogenation,18–21
dehydrogenation,22–25 and cross-coupling reactions,26–29 several strategies for syn-
thesizing Pd with controlled nuclearity have been reported. Pei et al.30 prepared Cu-
alloyed Pd single-atom catalyst and reported high selectivity, 85%, toward ethylene 1Department of Chemical Engineering, Virginia
at 100% acetylene conversion. Yan et al.31 fabricated atomically dispersed Pd on gra- Polytechnic Institute and State University,
phene using atomic layer deposition and reported 100% selectivity to butenes at 95% Blacksburg, VA 24060, USA
2Department of Chemistry, Virginia
conversion during selective hydrogenation of 1,3-butadiene.
Commonwealth University, Richmond, VA
23284-2006, USA
Covalent organic frameworks (COFs) are crystalline porous organic solids con- 3Lead contact
structed from linking organic building blocks together by strong covalent *Correspondence: helkaderi@vcu.edu (H.E.-K.),
bonds.32–34 With the aid of reticular chemistry, COFs can be designed with a precise amkarim@vt.edu (A.M.K.)
control over their textural properties (i.e., pore size and shape, surface area) while https://doi.org/10.1016/j.xcrp.2021.100495

Cell Reports Physical Science 2, 100495, July 21, 2021 ª 2021 The Author(s). 1
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
ll
OPEN ACCESS
Article

the chemical nature of the building blocks dictates the chemical functionality of the
pores. Given their high surface areas and diverse chemical functionalities, COFs
have attracted considerable interest in heterogeneous catalysis because of their
ability to host catalytic metals in the form of both metal nanoparticles and single
atoms coordinated to the walls of the frameworks.35–40 The latter is of great interest
to metal single-atom catalytic processes because the periodic structure of COFs en-
ables high loading and rapid access of catalytic centers without compromising mass
transport rates of the reactants and products. Metal loading in COFs is typically car-
ried out by either pre- or post-synthetic modification strategies.38,41 Chelated metal
ions can be part of the building blocks prior to polymerization or metal ions can be
introduced after framework synthesis which gives more flexibility in terms of metal
loading level and metal type. For example, imine-linked COFs have been explored
as metal support scaffolds because of their Schiff-base nature, which proved to be an
effective strategy for accessing recyclable catalysts with remarkable activity in many
catalytic processes, especially Suzuki cross-coupling reactions.38,42–47 Surprisingly,
COFs use in gas-phase catalytic transformation remains very scarce.

In this work, we synthesize a high-surface area (SABET = 1,640 m2 g1) PdCl2-func-


tionalized imine-linked COF (PdCl2@ILCOF-1) with 18.1 wt % atomically dispersed
Pd, which, to the best of our knowledge, is the highest reported weight loading of
a Pd single-atom catalyst. We characterize the PdCl2@ILCOF-1 catalyst using various
spectroscopy techniques, including X-ray absorption near-edge structure (XANES),
extended X-ray absorption fine structure (EXAFS), X-ray photoelectron spectros-
copy (XPS), and Fourier transform infrared spectroscopy (FTIR) of adsorbed CO,
and test its catalytic performance for ethylene hydrogenation. The results show
that Pd atoms are isolated and bound to the COF through N bonds. The ethylene
hydrogenation kinetics are consistent with isolated Pd atoms and show that ethylene
adsorption is not limiting, while hydrogen is formed on Pd through dissociative
chemisorption.

RESULTS AND DISCUSSION


Physical properties of ILCOF-1 and PdCl2@ILCOF-1
Because high surface area and wide channels decorated with single metal binding
sites are vital for catalyst accessibility and catalytic activity, we selected ILCOF-148
for Pd(II) immobilization and subsequent use in gas-phase hydrogenation of
ethylene. ILCOF-1 was prepared by condensation reaction between tetrakis(p-for-
mylphenyl)pyrene and 1,4-p-phenylenediamine in 1,4-dioxane using acetic acid as
a catalyst according to our reported method.48 ILCOF-1 is mesoporous (pore size
 2.3 nm) and has a two-dimensional (2D) eclipsed layered structure (Figure 1)
and accessible nitrogen sites readily available for complexing metal ions. PdCl2@IL-
COF-1 was prepared by post-synthetic modification of ILCOF-1 with PdCl2 using a
solution of bis(benzonitrile)palladium(II) chloride (Figure 1). Treatment of ILCOF-1
with bis(benzonitrile)palladium(II) chloride at ambient temperature resulted in an im-
mediate color change from yellow to dark brown as a result of PdCl2 coordination to
the imine sites of the framework (Figure 1). If each pair of nitrogen atoms of the imine
sites in the framework are coordinated to a Pd atom, the theoretical Pd content
would be 19.0 wt %. The palladium content of PdCl2@ILCOF-1 was analyzed by
ICP and found to be 18.1 wt %, which is very close to the theoretical limit.

Porosity parameters of pristine ILCOF-1 and PdCl2@ILCOF-1 were obtained


from argon sorption/desorption isotherms at 87 K (Figures 2A and 2B). ILCOF-1
has a very high surface area (SABET = 2,620 m2 g1) and large pores about 2.3 nm

2 Cell Reports Physical Science 2, 100495, July 21, 2021


ll
OPEN ACCESS
Article

Figure 1. Schematic of the PdCl2@ILCOF-1 catalyst structure


Schematic representation of immobilization of PdCl 2 into ILCOF-1 and its space-filling representation of the resulting 2D network. Gray, carbon; blue,
nitrogen; hydrogen is omitted for clarity.

in diameter. In comparison, the surface area of PdCl2@ILCOF-1 decreased to


1,640 m2 g1, while the pore size distribution (PSD) is somewhat broader and domi-
nated by smaller micropores (Figure 2B). To study the impact of PdCl2 complexation
with ILCOF-1 on its crystallinity, we collected powder X-ray diffraction (XRD) patterns
for ILCOF-1 and PdCl2@ILCOF-1, which show that the crystallinity of the pristine
framework is maintained after functionalization with PdCl2 (Figure 2C). Maintaining
high crystallinity and similar pore size are both very important in terms of catalytic
applications in order to maintain full accessibility to the palladium catalytic centers
aligned along the channels of the framework and to allow rapid mass transport of re-
actants and products.

Local coordination and electronic properties of Pd


To gain insights into the coordination environment of the palladium center, we per-
formed XPS and X-ray absorption spectroscopy (XAS). The characteristic peaks for
Pd and Cl ions in the XPS survey spectrum of PdCl2@ILCOF-1 are shown in Figure 3A.
The successful coordination of Pd(II) to the nitrogen atoms was evidenced by a 1 eV
shift to higher binding energy in the N1s XPS spectrum (Figure 3B). To study the
chemical environment around PdCl2, the XPS spectra of PdCl2@ILCOF-1 for Pd3d
and N1s were compared with that of 1,10-phenanthrolin-dichloropalladium
(PdCl2[phen], purchased from Sigma-Aldrich, used as received) as shown in Figures
3C and 3D. Our studies show that the binding energy values for Pd and N atoms of
PdCl2@ILCOF-1 and PdCl2(phen) are very similar (Figures 3C and 3D). Therefore, it
can be concluded that PdCl2 appears to be coordinated to the framework by two ni-
trogen atoms, which is consistent with the reported crystal structure of
PdCl2(phen).49 This is consistent with the 1 eV shift to higher binding energy in

Cell Reports Physical Science 2, 100495, July 21, 2021 3


ll
OPEN ACCESS
Article

A B C

Figure 2. Physical properties of the ILCOF-1 and PdCl2@ILCOF-1 catalyst


(A–C) Argon uptake isotherms at 87 K (A), pore size distributions (B), and powder X-ray diffraction studies (C) for ILCOF-1 and PdCl 2 @ILCOF-1. Filled
and open symbols represent adsorption and desorption, respectively.

N1s XPS spectrum for all the nitrogen atoms of the framework upon complexation
with PdCl2 (Figure 3B), indicating binding of PdCl2 to the imine groups.47

Figure 4A shows the XANES spectra at the Pd K-edge of dried PdCl2@ILCOF-1 cata-
lyst and Pd standard samples, PdCl2 and PdO. PdO has the highest white-line inten-
sity, while PdCl2@ILCOF-1 and PdCl2 are similar but with different spectral patterns.
As shown by the EXAFS results (Figures 4B and 4C; Figure S1), the PdCl2@ILCOF-1
EXAFS spectrum has peaks with some similarities to PdO and PdCl2. The EXAFS
modeling results (Figure 5; Table 1) show that Pd, on average, is bound to two Cl
and two N atoms at 2.03 and 2.31 Å, respectively (see detailed fit in Table S1 and
Figure S3). The results show that PdCl2 remains intact during the synthesis and binds
to the framework through two N atoms. This is consistent with the XPS results, which
showed similar spectra for the PdCl2@ILCOF-1 catalyst and PdCl2(phen). We can
conclude from the EXAFS and XPS results that Pd is stabilized as isolated single
atoms within the pores of ILCOF-1. We note that the Pd remained isolated and
the Pd coordination was unaffected by pretreatment in He at 120 C, as confirmed
by the almost identical XANES and EXAFS spectra (Figure S2). To determine the ho-
mogeneity of the local environment around the Pd atoms in the PdCl2@ILCOF-1
catalyst, CO adsorption was studied at 40 C by diffuse-reflectance infrared Fourier
transform spectroscopy (DRIFTS). A symmetrical and relatively narrow band at
2,130 cm1 (Figure 4D; full width at half maximum [FWHM] of 20 cm1) was
observed. No other CO bands were present between 1,800 and 2,200 cm1, where
the typical vibrational frequencies for linear, bridge, and hollow CO bound on Pd are
reported,50–54 confirming that Pd is present as isolated single atoms. The CO band
position on Pd atoms in the PdCl2@ILCOF-1 catalyst (2,130 cm1) is very close to that
reported on Cu-alloyed Pd single atoms (2,133 cm1).30 The relatively narrow
2,130 cm1 CO band in DRIFTS indicates that the local environment of the Pd single
atoms in the PdCl2@ILCOF-1 catalyst is similar.

Catalytic activity and kinetics of ethylene hydrogenation


Ethylene hydrogenation is an important reaction, can be performed at mild condi-
tions, and has been extensively used as a probe reaction for extended metal surfaces
and supported metal particles such as Pt, Ir, Rh, and Au.55–64 We tested the single-
atom PdCl2@ILCOF-1 catalyst for ethylene hydrogenation at atmospheric pressure
and a constant temperature of 20 C. We note that the catalyst activity remained
almost identical after 6 h time on stream (Figure S4), indicating the stability of Pd un-
der reaction conditions. The catalytic activities are reported per Pd atom assuming
100% dispersion, which is justified by the results from EXAFS and CO adsorption in

4 Cell Reports Physical Science 2, 100495, July 21, 2021


ll
OPEN ACCESS
Article

A B

C D

Figure 3. Electronic properties of the PdCl2@ILCOF-1 catalyst


(A–D) XPS survey spectra of PdCl2 @ILCOF-1 (A), comparison of N1s XPS spectra of ILCOF-1 and
PdCl 2 @ILCOF-1 (B), comparison of Pd3d XPS spectra of PdCl 2 @ILCOF-1 and PdCl 2 (phen) (phen =
1,10-phenanthroline) (C), and comparison of N1s XPS spectra of PdCl 2 @ILCOF-1 and PdCl 2 (phen)
(Sigma-Aldrich) (D).

DRIFTS showing that Pd remained mostly present as single atoms in the spent cata-
lyst (see Table S1; Note S1; Figures S5–S7). We note that although the Pd is present
mostly as single atoms in the spent catalyst, the DRIFTS and EXAFS results indicate
the presence of a minority Pd clusters, and possibly dimers, which contribute to the
measured reactivity.

Figure 6 shows the dependence of the turnover frequency (TOF) on the hydrogen and
ethylene partial pressures. The slopes of best linear fit to the kinetic data under
steady-state ethylene hydrogenation indicate that the reaction order in H2 was
0.59 G 0.01 and the reaction order in C2H4 was 0.01 G 0.01 (Figure 6). Kinetic studies
of ethylene hydrogenation have been reported on supported metal single atoms,63
subnanometer clusters,61,65–67 and nanoparticles.68,69 For example, the Gates group
studied ethylene hydrogenation on supported Ir65,66,70,71 and Rh60,61,67,72 single
atoms, pair sites, and small clusters, and the catalysts were found to be active at
353 K and atmospheric pressure. On MgO-supported Ir pair-site and Ir6 catalysts,
the reaction order in H2 and C2H4 were found to be 0.5 and 0, respectively. In
another study, Guzman and Gates63 characterized and tested mononuclear gold
complexes supported on MgO for ethylene hydrogenation at atmospheric pressure
and 353 K. The reaction order in H2 (Phydrogen = 4.7–60.0 kPa at Pethylene = 5.3 kPa) was
0.5 and that the reaction order in ethylene (Pethylene = 4.7–40.0 kPa at Phydrogen =
5.3 kPa) was 0.0 on mononuclear gold complex catalyst supported on MgO. More-
over, Molero et al.69 measured the kinetics of ethylene hydrogenation on Pd(111)

Cell Reports Physical Science 2, 100495, July 21, 2021 5


ll
OPEN ACCESS
Article

A B

C D

Figure 4. Local coordination and electronic properties of the PdCl2@ILCOF-1 catalyst


Characterization Pd K-edge X-ray absorption spectra of the dried PdCl 2 @ILCOF-1 catalyst, PdCl 2 ,
and PdO.
(A–C) XANES (A), (B) EXAFS magnitude, and (C) imaginary part of the of the Fourier transformed k 2 -
weighted c(k) data, Dk = 3–12.0 Å1 .
(D) DRIFTS spectra of CO adsorption on the PdCl 2 @ILCOF-1catalyst at 40  C after N 2
pretreatment for 12 h at 120  C.

at 320 K. The reaction orders in ethylene (Pethylene = 6.7–40.0 kPa at Phydrogen =


13.3 kPa) and hydrogen (Pethylene = 6.0–80.0 kPa at Phydrogen = 13.3 kPa) were 1.05
and 0.22, respectively. Additionally, on 4.2 nm Pd nanoparticles supported on sil-
ica, Binder et al.68 reported reaction orders in ethylene and hydrogen at 298 K of
0.39 G 0.01 and 1.04 G 0.07, respectively. Our results suggest that PdCl2@ILCOF-1
catalyzes ethylene hydrogenation by a similar reaction mechanism as Ir pair-site and
Ir6 catalysts and mononuclear gold complex catalyst supported on MgO, where
ethylene adsorption is not limiting while hydrogen forms on Pd through dissociative
adsorption. The results suggest that Pd single atoms in PdCl2@ILCOF-1 are active for
hydrogenation of ethylene, but the nature of the active site and the role of Cl need
further study. We note that other single-atom catalysts with high metal loading
have been reported recently.73–78 For example, Babucci et al.78 recently reported
a 14.8 wt % atomically dispersed Ir on reduced graphene aerogel. Wang et al.75 syn-
thesized atomically dispersed Pt, Pd, Ru, Rh, and Ir on mesoporous sulfur-doped car-
bon with 10 wt % loading. Thus, to the best of our knowledge, the 18.1 wt % atomi-
cally dispersed Pd (PdCl2@ILCOF-1) presented in this work is the highest reported
weight loading of a Pd single-atom catalyst. More important, our results show that
despite the high Pd loading, all the metal binding sites are chemically very similar
as indicated by CO DRIFTS. Therefore, our work demonstrates that COFs provide
a large number of identical metal binding sites that can be used to immobilize and
study metal single-atom catalysts for gas-phase reactions.

6 Cell Reports Physical Science 2, 100495, July 21, 2021


ll
OPEN ACCESS
Article

A B

Figure 5. Structure of the PdCl2@ILCOF-1 catalyst


(A and B) K-edge EXAFS spectra and fit for the PdCl 2 @ILCOF-1 catalyst dried by N 2 treatment for
12 h at 120  C. (A) EXAFS magnitude and (B) imaginary part of the of the Fourier transformed k 2 -
weighted c(k) data, Dk = 3–12.0 Å1

In summary, we report atomically dispersed Pd supported on imine-linked COF (IL-


COF-1) at a palladium loading of 18.1 wt %. XPS, XAS, and infrared (IR) spectroscopy
were used to investigate the COF-supported Pd catalyst. The results provide evi-
dence of the atomic dispersion of Pd atoms, which are stabilized by covalent bonds
to N from the COF. The Pd atoms are shown to be catalytically active for ethylene
hydrogenation and appear to be stable during catalysis. The kinetic measurements
suggest that ethylene adsorption is not limiting, while hydrogen is formed on Pd
through dissociative chemisorption. This study demonstrates that COFs are excel-
lent candidates for immobilizing metal single-atom catalysts and their use in gas-
phase catalytic reactions.

EXPERIMENTAL PROCEDURES
Resource availability
Lead contact
Further information and requests for resources and data should be directed to and
will be fulfilled by the lead contact, Ayman M. Karim (amkarim@vt.edu).

Materials availability
Unless otherwise noted, reagents obtained from commercial suppliers were used
without further purification. Solvents were technical grade and were used without
further purification. See catalyst preparation section for full details of synthetic
procedures.

Data and code availability


The authors declare that data supporting the findings of this study are available
within the paper and the supplemental information. All other data are available
from the lead contact upon reasonable request.

Catalyst preparation
ILCOF-1 was synthesized according to our previously reported procedure.48 In brief,
a Pyrex tube, with outer diameter (OD) of 12 mm and inner diameter (ID) of 10 mm,
was charged with the monomers tetrakis(p-formylphenyl)pyrene (20 mg, 32 mmol)
and 1,4-p-phenylenediamine (7.0 mg, 64 mmol). Then, the monomers were sus-
pended in a solution obtained by mixing 3.0 mL 1,4-dioxane and 0.6 mL 3.0 M acetic
acid. The mixture was sonicated for 10 min, and 6 M acetic acid (0.6 mL) was added
slowly. The tube was degassed by three freeze-pump-thaw cycles and flash-frozen at

Cell Reports Physical Science 2, 100495, July 21, 2021 7


ll
OPEN ACCESS
Article

Table 1. Summary of EXAFS modeling of the PdCl2@ILCOF-1 catalyst after different


pretreatments

Sample Fresh Dried


Pretreatment helium at 30  C helium at 120 C
NPd-N 1.9 G 0.3 1.7 G 0.3
NPd-Cl 2.2 G 0.4 2.2 G 0.4
RPd-N (Å) 2.02 G 0.01 2.03 G 0.01
RPd-Cl (Å) 2.31 G 0.01 2.31 G 0.01
All spectra were collected at 30 C in the same gas used for the pretreatment. Full fit details are provided
in Table S1. N, coordination number of absorber-backscatterer pair; R, radial absorber-backscatterer dis-
tance.

77 K using a liquid N2 bath and then flame-sealed under vacuum. The mixture was
heated at 120 C for 3 days. The resulting mixture was filtered and washed with anhy-
drous acetone followed by tetrahydrofuran (THF) to afford ILCOF-1 as a yellow pow-
der (20 mg).

PdCl2@ILCOF-1 was synthesized using the ILCOF-1 and PdCl2 as follows. A dry
25 mL Schlenk flask with a stir bar was charged with ILCOF-1 (70 mg, 0.09 mmol) un-
der an inert atmosphere. A freshly prepared solution of bis(benzonitrile)palladium(II)
chloride (PdCl2[CNPh]2, 70 mg) in anhydrous chloroform (15 mL) was added using a
syringe. The resulting brown mixture was stirred at room temperature for 2 h under a
N2 atmosphere. The product was isolated by filtration and washed with anhydrous
chloroform via Soxhlet under N2 atmosphere overnight. The solid was dried under
vacuum at 120 C to obtain PdCl2@ILCOF-1 (97 mg, 95% yield) as a dark brown pow-
der. The palladium content of the sample was determined using inductively coupled
plasma optical emission spectroscopy (ICP-OES) analysis using a Vista MPX CCD
Simultaneous ICP-OES, and the Pd content was found to be 18.1 wt %.

Catalyst characterization
For porosity analysis of the samples, argon sorption measurements were carried out
using a Quantachrome Autosorb iQ volumetric analyzer using ultrahigh purity (UHP)
grade argon at 87 K. The samples were outgassed at 110 C under vacuum for 24 h
before collecting argon sorption isotherms. The specific surface area of the samples
was calculated from the argon adsorption branch on the basis of the Brunauer-Em-
mett-Teller (BET) model. PSD was calculated from adsorption argon isotherms by
nonlocal density functional theory (NLDFT) using a cylindrical pore (zeolite) model.
Powder XRD patterns were collected on a Panalytical X’Pert pro multipurpose
diffractometer (MPD) with a copper Ka radiation.

The XPS analysis was conducted using a Thermo Fisher Scientific ESCALAB 250 in-
strument with a monochromatized Al Ka X-ray source. For preparation of the sam-
ples for XPS, the powder was pressed into a piece of double-stick carbon tape, which
was mounted on the sample holder. For the purpose of charge compensation during
XPS analysis, a combination of a low-energy electron flood gun and an argon ion
flood gun was used. The binding energy scale of the XPS spectrum was calibrated
by setting the C1s peak at 285.0 eV. The XPS data were analyzed using Thermo
Avantage software (version 4.84).

DRIFTS of CO adsorption was performed on a Thermo Fisher Scientific Nicolet iS50


FTIR spectrometer equipped with a liquid nitrogen-cooled mercury-cadmium-tellu-
ride (MCT) infrared detector. The catalyst was loaded into the sample cup of an in

8 Cell Reports Physical Science 2, 100495, July 21, 2021


ll
OPEN ACCESS
Article

A B

Figure 6. Kinetics of ethylene hydrogenation catalyzed by PdCl2@ILCOF-1


(A) Dependence of TOF on H 2 partial pressure at a constant ethylene partial pressure of 0.5 kPa (the
reaction order in H 2 is 0.6).
(B) Dependence of TOF on C2 H 4 partial pressure at a constant H 2 partial pressure of 5 kPa (the
reaction order in ethylene is 0.01).
The symbols and error bars represent the average and SD from repeat experiments. All
measurements were taken at 293 K.

situ Praying Mantis diffuse reflection cell (Harrick Scientific Products). Prior to CO
chemisorption, the catalyst was first dried in 100 sccm N2 flow at 120 C for 12 h. After
the N2 treatment, the catalyst was cooled down to 40 C using liquid nitrogen un-
der 40 sccm N2. For the spent catalyst, the catalyst was first exposed to a flow of
0.5 kPa C2H4 and 5 kPa H2 (balance N2, 160 sccm total flowrate). A background spec-
trum was collected at 40 C under N2 and subtracted from the series of subsequent
spectra. A 1 kPa CO gas (balance N2) flowed through the sample for 20 min and then
purged with 40 sccm N2 to remove the gas phase CO on the PdCl2@ILCOF-1 cata-
lyst, while a series of CO spectra were collected continuously. Each spectrum is an
average of 32 scans with data spacing of 0.482 cm1. The CO (5%, balance N2),
H2, and C2H4 (20%, balance N2) were obtained from Airgas, and N2 was obtained
from in-house liquid N2 boil-off. N2 and H2 were purified by high-capacity moisture
and oxygen traps (Restek catalog #21997 and #20601). CO was purified with a cor-
rosive gas purifier (Metal-X Purification Medium, Nanochem).

XAS measurements at the Pd K-edge for the PdCl2@ILCOF-1 samples were per-
formed at the Stanford Synchrotron Radiation Light Source (SSRL) at beamline 9-3
using an in-house-built cell with a 4 mm ID glassy carbon tube.79 Beamline 9-3 is a
16-pole, 2 T wiggler side station with vertically collimating mirror for harmonic rejec-
tion and a cylindrically bend mirror for focusing. The photon energy was selected us-
ing a liquid nitrogen-cooled, double-crystal Si (220) f = 90 monochromator. The
samples were scanned simultaneously in transmission and fluorescence detection
modes using ion chambers (filled with pure N2) and a 100-element solid-state Ge
monolith detector (Canberra). A Pd standard foil was scanned simultaneously with
each sample for energy calibration. Step-scanning X-ray absorption spectra were
measured from 24,150 to 25,244 eV, corresponding to photoelectron wave number
k = 15.2 Å1. XANES and EXAFS data processing and analysis were performed using
Athena and Artemis programs of the Demeter data analysis package.80,81 For each
sample, four scans were collected and merged after alignment. c(k) was obtained by
subtracting smooth atomic background from the normalized absorption coefficient
using the AUTOBK code. The theoretical EXAFS signal was constructed using the
FEFF6 code82 and fitted to the data in R-space using the Artemis program. The theo-
retical EXAFS signals for Pd-N and Pd-Cl scattering paths were constructed using the
FEFF6 code using the crystal structure of Pd(OAc)2TOP2 and replacing the O atom
bound to Pd from the acetate with N and the phosphorous bound to Pd in trioctyl-
phosphine with Cl. For the Pd-Pd scattering path, we used an fcc Pd structure. The
spectra were fitted by varying the coordination number of the single scattering

Cell Reports Physical Science 2, 100495, July 21, 2021 9


ll
OPEN ACCESS
Article

paths, Pd-Pd, Pd-N, and Pd-Cl, the effective scattering lengths, the bond length dis-
order of each path, and the correction to the threshold energy, DE0 (DE0 was the
same for Pd-N and Pd-Cl because these paths were calculated using the same
model). Because of the strong correlation between the parameters of the Pd-N
and Pd-Cl scattering paths, especially coordination numbers and bond disorders,
we used one common parameter for the bond length disorder of both paths. This
allowed us to obtain smaller uncertainties in the model fit parameters for comparison
purposes between samples. S02 (the passive electron reduction factor) was obtained
by first analyzing the spectrum for a Pd foil, and the best fit value (0.79) was fixed dur-
ing the fitting. The k range used for Fourier transform of the c(k) was 3–12.3 Å1, and
the r range for fitting was 1.2–2.3 Å for fitting Pd-N and Pd-Cl scattering paths only
and 1.2–2.9 Å when Pd-Pd needed to be included in the fit. The best parameters fit
using a k weight of 2 in Artemis are reported, but the results were similar to those
using k weights of 1, 2, and 3.

Catalytic activity and kinetics measurements


The PdCl2@ILCOF-1 catalyst was tested for ethylene hydrogenation in a quartz
packed-bed reactor.
+ H2
C2 H4 / C2 H6 ð  DH+ = 137 kJ = molÞ
The 18.1 wt % PdCl2@ILCOF-1 catalyst was diluted (intra-pellet) with silica gel (SiO2;
0.075–0.250 mm, 150 A, ACROS Organics calcined in air at 800 C before use) such
that the mixture contains 0.1% Pd by weight in order to minimize transport limita-
tions. The mixture of PdCl2@ILCOF-1 and SiO2 were ground using an agate mortar
and pestle and then pelletized and sieved into 180–300 mm pellets, and 0.02 g of the
diluted mixture was loaded into a 0.75 inch OD glass packed-bed reactor. The cata-
lyst was first dried at 120 C for 12 h in He flow at 100 sccm, then cooled to 20 C prior
to catalytic measurements. All the activity measurements were carried out at atmo-
spheric pressure, and the gas composition from the reactor outlet was analyzed us-
ing an online gas chromatograph (Micro GC Fusion Gas Analyzer, Inficon). C2H4, and
C2H6 were separated using an Rt-Alumina Na2SO4 packed column and then quanti-
fied using a thermal conductivity detector (TCD). The feed gases H2 (UHP, 99.999%),
20% C2H4 (certified gas, Airgas), and He (UHP, 99.999%) were obtained from Airgas.
H2 and He were purified by high-capacity moisture and oxygen traps (Restek catalog
#20600). All gas flows were controlled by mass flow controllers (5850EM, Brooks
Instrument).

The conversion and TOF (rate of reaction per Pd atom per second) of ethylene hydro-
genation were determined from differential conversion and calculated as shown
below. We note that no products other than C2H4 and C2H6 were detected by the
Micro GC and that the carbon balance (outlet/inlet) was close to 100% under all
conditions.

moles of C2 H6 ðoutletÞ
Conversion =
moles of C2 H4 ðinletÞ


 rate of C2 H6 formation moleC2 H6 g1
cat s
1
TOF s1 = 1

moles of surface Pd molePd gcat

Catalytic hydrogenation of ethylene was carried out with an ethylene (C2H4) partial
pressure of 0.5 kPa and hydrogen (H2) partial pressure of 5 kPa at 20 C. Kinetics ex-
periments were performed by varying the C2H4 and H2 partial pressures in the
ranges 0.5–2.5 kPa and 1–5 kPa, respectively. The PdCl2@ILCOF-1 catalyst was

10 Cell Reports Physical Science 2, 100495, July 21, 2021


ll
OPEN ACCESS
Article

found to be very stable, with less than 5% change in activity over a 6 h period (the
typical time for the kinetic measurement). The only product detected was ethane,
and there was no measurable conversion of C2H4 and H2 in the absence of the
catalyst.

SUPPLEMENTAL INFORMATION
Supplemental information can be found online at https://doi.org/10.1016/j.xcrp.
2021.100495.

ACKNOWLEDGMENTS
This research was sponsored primarily by the American Chemical Society Petroleum
Research Fund (PRF) and was accomplished under award number 55575-ND5. Use
of the SSRL (beamline 9-3, user proposal 4938) was supported by the U.S. Depart-
ment of Energy, Office of Basic Energy Sciences, under contract DE-AC02-
76SF00515.

AUTHOR CONTRIBUTIONS
C.-T.K. conducted the ethylene hydrogenation kinetics, DRIFTS measurements (and
helped with XAFS), performed the data analysis, and wrote the first draft of the
manuscript. Y.L. performed the XAFS and some of the DRIFTS experiments. P.A.
and K.S.W. synthesized the catalyst and performed the surface area, XRD, and
XPS measurements. A.M.K. and H.E.-K. conceived the idea and directed the project.
C.-T.K., Y.L., H.E.-K., and A.M.K. cowrote the paper. All authors discussed the results
and commented on the paper.

DECLARATION OF INTERESTS
The authors declare no competing interests.

Received: February 3, 2021


Revised: May 26, 2021
Accepted: June 16, 2021
Published: July 21, 2021

REFERENCES
1. Fu, J., Lym, J., Zheng, W., Alexopoulos, K., 5. Qi, J., Finzel, J., Robatjazi, H., Xu, M., Hoffman, 9. Wang, H., Liu, J.-X., Allard, L.F., Lee, S., Liu, J.,
Mironenko, A.V., Li, N., Boscoboinik, J.A., Su, A.S., Bare, S.R., Pan, X., and Christopher, P. Li, H., Wang, J., Wang, J., Oh, S.H., Li, W., et al.
D., Weber, R.T., and Vlachos, D.G. (2020). C–O (2020). Selective methanol carbonylation to (2019). Surpassing the single-atom catalytic
bond activation using ultralow loading of noble acetic acid on heterogeneous atomically activity limit through paired Pt-O-Pt ensemble
metal catalysts on moderately reducible dispersed ReO4/SiO2 catalysts. J. Am. Chem. built from isolated Pt1 atoms. Nat. Commun.
oxides. Nat. Catal. 3, 446–453. Soc. 142, 14178–14189. 10, 3808.

2. Zang, W., Yang, T., Zou, H., Xi, S., Zhang, H., 6. Wei, Y.S., Zhang, M., Zou, R., and Xu, Q. (2020). 10. Zhao, Y., Yang, K.R., Wang, Z., Yan, X., Cao,
Liu, X., Kou, Z., Du, Y., Feng, Y.P., and Shen, L. Metal-organic framework-based catalysts with S., Ye, Y., Dong, Q., Zhang, X., Thorne, J.E.,
(2019). Copper single atoms anchored in single metal sites. Chem. Rev. 120, 12089– Jin, L., et al. (2018). Stable iridium dinuclear
porous nitrogen-doped carbon as efficient pH- 12174. heterogeneous catalysts supported on
universal catalysts for the nitrogen reduction metal-oxide substrate for solar water
reaction. ACS Catal. 9, 10166–10173. 7. Domestici, C., Tensi, L., Zaccaria, F., Kissimina, oxidation. Proc. Natl. Acad. Sci. U S A 115,
N., Valentini, M., D’Amato, R., Costantino, F., 2902–2907.
3. Lu, Y., Wang, J., Yu, L., Kovarik, L., Zhang, X., Zuccaccia, C., and Macchioni, A. (2020). 11. Fernández, E., and Boronat, M. (2019). Sub
Hoffman, A.S., Gallo, A., Bare, S.R., Sokaras, D., Molecular and heterogenized dinuclear Ir-Cp* nanometer clusters in catalysis. J. Phys.
Kroll, T., et al. (2019). Identification of the active water oxidation catalysts bearing EDTA or Condens. Matter 31, 013002.
complex for CO oxidation over single-atom Ir- EDTMP as bridging and anchoring ligands. Sci.
on-MgAl2O4 catalysts. Nat. Catal. 2, 149–156. Bull. (Beijing) 65, 1614–1625. 12. Lee, S., Halder, A., Ferguson, G.A., Seifert, S.,
Winans, R.E., Teschner, D., Schlögl, R.,
4. Shan, J., Li, M., Allard, L.F., Lee, S., and Flytzani- 8. Guan, E., Ciston, J., Bare, S.R., Runnebaum, Papaefthimiou, V., Greeley, J., Curtiss, L.A.,
Stephanopoulos, M. (2017). Mild oxidation of R.C., Katz, A., Kulkarni, A., Kronawitter, and Vajda, S. (2019). Subnanometer cobalt
methane to methanol or acetic acid on C.X., and Gates, B.C. (2020). Supported oxide clusters as selective low temperature
supported isolated rhodium catalysts. Nature metal pair-site catalysts. ACS Catal. 10, oxidative dehydrogenation catalysts. Nat.
551, 605–608. 9065–9085. Commun. 10, 954.

Cell Reports Physical Science 2, 100495, July 21, 2021 11


ll
OPEN ACCESS
Article

13. Negreiros, F.R., Halder, A., Yin, C., Singh, A., 25. Bi, Q.Y., Lin, J.D., Liu, Y.M., He, H.Y., Huang, Recent development of covalent organic
Barcaro, G., Sementa, L., Tyo, E.C., Pellin, M.J., F.Q., and Cao, Y. (2016). Dehydrogenation of frameworks (COFs): synthesis and catalytic
Bartling, S., Meiwes-Broer, K.H., et al. (2018). formic acid at room temperature: boosting (organic-electro-photo) applications. Mater.
Bimetallic Ag-Pt sub-nanometer supported palladium nanoparticle efficiency by coupling Horiz. 7, 411–454.
clusters as highly efficient and robust oxidation with pyridinic-nitrogen-doped carbon. Angew.
catalysts. Angew. Chem. Int. Ed. Engl. 57, Chem. Int. Ed. Engl. 55, 11849–11853. 40. Song, Y., Sun, Q., Aguila, B., and Ma, S. (2018).
1209–1213. Opportunities of covalent organic frameworks
26. Cotugno, P., Casiello, M., Nacci, A., Mastrorilli, for advanced applications. Adv. Sci. (Weinh.) 6,
14. Yao, S., Lin, L., Liao, W., Rui, N., Li, N., Liu, Z., P., Dell’Anna, M.M., and Monopoli, A. (2014). 1801410.
Cen, J., Zhang, F., Li, X., and Song, L. (2019). Suzuki coupling of iodo and bromoarenes
Exploring metal–support interactions to catalyzed by chitosan-supported Pd- 41. Lin, S., Diercks, C.S., Zhang, Y.-B., Kornienko,
immobilize subnanometer Co clusters on g– nanoparticles in ionic liquids. J. Organomet. N., Nichols, E.M., Zhao, Y., Paris, A.R., Kim, D.,
Mo2N: A highly selective and stable catalyst for Chem. 752, 1–5. Yang, P., Yaghi, O.M., and Chang, C.J. (2015).
CO2 activation. ACS Catal. 9, 9087–9097. Covalent organic frameworks comprising
27. Elazab, H.A., Siamaki, A.R., Moussa, S., cobalt porphyrins for catalytic CO2 reduction in
15. Zandkarimi, B., and Alexandrova, A.N. (2019). Gupton, B.F., and El-Shall, M.S. (2015). Highly water. Science 349, 1208–1213.
Dynamics of subnanometer pt clusters can efficient and magnetically recyclable
graphene-supported Pd/Fe3O4 nanoparticle 42. Ding, S.-Y., Gao, J., Wang, Q., Zhang, Y., Song,
break the scaling relationships in catalysis.
catalysts for Suzuki and Heck cross-coupling W.-G., Su, C.-Y., and Wang, W. (2011).
J. Phys. Chem. Lett. 10, 460–467.
reactions. Appl. Catal. A Gen. 491, 58–69. Construction of covalent organic framework for
catalysis: Pd/COF-LZU1 in Suzuki-Miyaura
16. Hülsey, M.J., Zhang, B., Ma, Z., Asakura, H., Do,
28. Narayanan, R., and El-Sayed, M.A. (2005). coupling reaction. J. Am. Chem. Soc. 133,
D.A., Chen, W., Tanaka, T., Zhang, P., Wu, Z.,
Carbon-supported spherical palladium 19816–19822.
and Yan, N. (2019). In situ spectroscopy-guided
engineering of rhodium single-atom catalysts nanoparticles as potential recyclable catalysts
for the Suzuki reaction. J. Catal. 234, 348–355. 43. Gonçalves, R.S., de Oliveira, A.B., Sindra, H.C.,
for CO oxidation. Nat. Commun. 10, 1330. Archanjo, B.S., Mendoza, M.E., Carneiro, L.S.,
29. Zhang, Z., and Wang, Z. (2006). Diatomite- Buarque, C.D., and Esteves, P.M. (2016).
17. Kuo, C.-T., Lu, Y., Kovarik, L., Engelhard, M., Heterogeneous catalysis by covalent organic
supported Pd nanoparticles: an efficient
and Karim, A.M. (2019). Structure sensitivity of frameworks (COF): Pd (OAc) 2@ COF-300 in
catalyst for Heck and Suzuki reactions. J. Org.
acetylene semi-hydrogenation on Pt single cross-coupling reactions. ChemCatChem 8,
Chem. 71, 7485–7487.
atoms and subnanometer clusters. ACS Catal. 743–750.
9, 11030–11041. 30. Pei, G.X., Liu, X.Y., Yang, X., Zhang, L., Wang,
A., Li, L., Wang, H., Wang, X., and Zhang, T. 44. Yan, Q., Xu, H., Jing, X., Hu, H., Wang, S., Zeng,
18. Chan, C.W.A., Mahadi, A.H., Li, M.M.-J., (2017). Performance of Cu-alloyed Pd single- C., and Gao, Y. (2020). Post-synthetic
Corbos, E.C., Tang, C., Jones, G., Kuo, W.C.H., atom catalyst for semihydrogenation of modification of imine linkages of a covalent
Cookson, J., Brown, C.M., Bishop, P.T., and acetylene under simulated front-end organic framework for its catalysis application.
Tsang, S.C. (2014). Interstitial modification of conditions. ACS Catal. 7, 1491–1500. RSC Advances 10, 17396–17403.
palladium nanoparticles with boron atoms as a
green catalyst for selective hydrogenation. 31. Yan, H., Cheng, H., Yi, H., Lin, Y., Yao, T., Wang, 45. Yang, J., Wu, Y., Wu, X., Liu, W., Wang, Y., and
Nat. Commun. 5, 5787. C., Li, J., Wei, S., and Lu, J. (2015). Single-atom Wang, J. (2019). An N-heterocyclic carbene-
Pd1/graphene catalyst achieved by atomic functionalised covalent organic framework with
19. Choudary, B., Kumar, K.R., and Kantam, M.L. layer deposition: remarkable performance in atomically dispersed palladium for coupling
(1991). Synthesis and catalytic activity in selective hydrogenation of 1, 3-butadiene. reactions under mild conditions. Green Chem.
selective hydrogenation of palladium J. Am. Chem. Soc. 137, 10484–10487. 21, 5267–5273.
complexes anchored in montmorillonite.
J. Catal. 130, 41–51. 32. Geng, K., He, T., Liu, R., Dalapati, S., Tan, K.T., 46. Mullangi, D., Nandi, S., Shalini, S., Sreedhala,
Li, Z., Tao, S., Gong, Y., Jiang, Q., and Jiang, D. S., Vinod, C.P., and Vaidhyanathan, R. (2015).
20. Al-Shammary, A., Caga, I., Winterbottom, J., (2020). Covalent organic frameworks: design, Pd loaded amphiphilic COF as catalyst for
Tata, A., and Harris, I. (1991). Palladium-based synthesis, and functions. Chem. Rev. 120, 8814– multi-fold Heck reactions, C-C couplings and
diffusion membranes as catalysts in ethylene 8933. CO oxidation. Sci. Rep. 5, 10876.
hydrogenation. J. Chem. Technol. Biotechnol.
52, 571–585. 33. Kandambeth, S., Dey, K., and Banerjee, R. 47. Leng, W., Peng, Y., Zhang, J., Lu, H., Feng, X.,
(2019). Covalent organic frameworks: chemistry Ge, R., Dong, B., Wang, B., Hu, X., and Gao, Y.
21. Ciebien, J.F., Cohen, R.E., and Duran, A. (1998). beyond the structure. J. Am. Chem. Soc. 141, (2016). Sophisticated design of covalent
Catalytic properties of palladium nanoclusters 1807–1822. organic frameworks with controllable
synthesized within diblock copolymer films: bimetallic docking for a cascade reaction.
hydrogenation of ethylene and propylene. 34. Waller, P.J., Gándara, F., and Yaghi, O.M. Chemistry 22, 9087–9091.
Supramol. Sci. 5, 31–39. (2015). Chemistry of covalent organic
frameworks. Acc. Chem. Res. 48, 3053–3063. [48]. Rabbani, M.G., Sekizkardes, A.K., Kahveci, Z.,
Reich, T.E., Ding, R., and El-Kaderi, H.M.
22. Chen, Y., Wang, H., Liu, C.-J., Zeng, Z., Zhang,
35. Yusran, Y., Li, H., Guan, X., Fang, Q., and Qiu, S. (2013). A 2D mesoporous imine-linked
H., Zhou, C., Jia, X., and Yang, Y. (2012).
(2020). Covalent organic frameworks for covalent organic framework for high pressure
Formation of monometallic Au and Pd and
catalysis. EnergyChem 2, 100035. gas storage applications. Chem. Eur. J. 19,
bimetallic Au–Pd nanoparticles confined in
3324–3328.
mesopores via Ar glow-discharge plasma
36. Guo, J., and Jiang, D. (2020). Covalent organic
reduction and their catalytic applications in 49. Ha, K. (2009). Dichlorido(1,10-phenanthroline-
frameworks for heterogeneous catalysis:
aerobic oxidation of benzyl alcohol. J. Catal. kN,N0 )palladium(II). Acta Crystallogr. Sect. E
principle, current status, and challenges. ACS
289, 105–117. Struct. Rep. Online 66, m38.
Cent. Sci. 6, 869–879.
23. Deng, W., Chen, J., Kang, J., Zhang, Q., and 37. Liu, J., Wang, N., and Ma, L. (2020). Recent 50. Borchert, H., Jürgens, B., Zielasek, V.,
Wang, Y. (2016). Carbon nanotube-supported advances in covalent organic frameworks for Rupprechter, G., Giorgio, S., Henry, C., and
Au-Pd alloy with cooperative effect of metal catalysis. Chem. Asian J. 15, 338–351. Bäumer, M. (2007). Pd nanoparticles with highly
nanoparticles and organic ketone/quinone defined structure on MgO as model catalysts:
groups as a highly efficient catalyst for aerobic 38. Ma, D., Wang, Y., Liu, A., Li, S., Lu, C., and an FTIR study of the interaction with CO, O2,
oxidation of amines. Chem. Commun. (Camb.) Chen, C. (2018). Covalent organic frameworks: and H2 under ambient conditions. J. Catal. 247,
52, 6805–6808. Promising materials as heterogeneous 145–154.
catalysts for CC bond formations. Catalysts 8,
24. Akbayrak, S., Tonbul, Y., and Özkar, S. (2017). 404. 51. Boronin, A., Slavinskaya, E., Danilova, I.,
Nanoceria supported palladium (0) Gulyaev, R., Amosov, Y.I., Kuznetsov, P.,
nanoparticles: superb catalyst in 39. Sharma, R.K., Yadav, P., Yadav, M., Gupta, R., Polukhina, I., Koscheev, S., Zaikovskii, V., and
dehydrogenation of formic acid at room Rana, P., Srivastava, A., Zboril, R., Varma, R.S., Noskov, A. (2009). Investigation of palladium
temperature. Appl. Catal. B 206, 384–392. Antonietti, M., and Gawande, M.B. (2020). interaction with cerium oxide and its state in

12 Cell Reports Physical Science 2, 100495, July 21, 2021


ll
OPEN ACCESS
Article

catalysts for low-temperature CO oxidation. 62. Fierro-Gonzalez, J.C., and Gates, B.C. (2008). 73. Li, H., Wang, L., Dai, Y., Pu, Z., Lao, Z., Chen, Y.,
Catal. Today 144, 201–211. Catalysis by gold dispersed on supports: the Wang, M., Zheng, X., Zhu, J., Zhang, W., et al.
importance of cationic gold. Chem. Soc. Rev. (2018). Synergetic interaction between
52. Di Gregorio, F., Bisson, L., Armaroli, T., Verdon, 37, 2127–2134. neighbouring platinum monomers in CO2
C., Lemaitre, L., and Thomazeau, C. (2009). hydrogenation. Nat. Nanotechnol. 13,
Characterization of well faceted palladium 63. Guzman, J., and Gates, B.C. (2004). A 411–417.
nanoparticles supported on alumina by mononuclear gold complex catalyst supported
transmission electron microscopy and FT-IR on MgO: spectroscopic characterization 74. Shi, R., Tian, C., Zhu, X., Peng, C.-Y., Mei, B., He,
spectroscopy of CO adsorption. Appl. Catal. A during ethylene hydrogenation catalysis. L., Du, X.-L., Jiang, Z., Chen, Y., and Dai, S.
Gen. 352, 50–60. J. Catal. 226, 111–119. (2019). Achieving an exceptionally high loading
of isolated cobalt single atoms on a porous
53. Pérez, Y., Ruiz-González, M.L., González- 64. Cortright, R.D., Goddard, S.A., Rekoske, J.E., carbon matrix for efficient visible-light-driven
Calbet, J.M., Concepción, P., Boronat, M., and and Dumesic, J.A. (1991). Kinetic study of photocatalytic hydrogen production. Chem.
Corma, A. (2012). Shape-dependent catalytic ethylene hydrogenation. J. Catal. 127, 342–353. Sci. (Camb.) 10, 2585–2591.
activity of palladium nanoparticles embedded
in SiO2 and TiO2. Catal. Today 180, 59–67. 65. Guan, E., Debefve, L., Vasiliu, M., Zhang, S.,
75. Wang, L., Chen, M.X., Yan, Q.Q., Xu, S.L., Chu,
Dixon, D.A., and Gates, B.C. (2019). MgO-
54. Wang, H., and Liu, C.-j. (2011). Preparation and S.Q., Chen, P., Lin, Y., and Liang, H.W. (2019). A
Supported iridium metal pair-site catalysts are
characterization of SBA-15 supported Pd sulfur-tethering synthesis strategy toward high-
more active and resistant to CO poisoning than
catalyst for CO oxidation. Appl. Catal. B 106, analogous single-site catalysts for ethylene
loading atomically dispersed noble metal
672–680. catalysts. Sci. Adv. 5, eaax6322.
hydrogenation and hydrogen–deuterium
exchange. ACS Catal. 9, 9545–9553.
55. Cremer, P.S., Su, X., Shen, Y.R., and Somorjai, 76. Zhang, B.W., Jiao, Y., Chao, D.L., Ye, C., Wang,
G.A. (1996). Ethylene hydrogenation on Pt (111) 66. Schöttle, C., Guan, E., Okrut, A., Grosso- Y.X., Davey, K., Liu, H.K., Dou, S.X., and Qiao,
monitored in situ at high pressures using sum Giordano, N.A., Palermo, A., Solovyov, A., S.Z. (2019). Targeted synergy between
frequency generation. J. Am. Chem. Soc. 118, Gates, B.C., and Katz, A. (2019). Bulky adjacent Co atoms on graphene oxide as an
2942–2949. calixarene ligands stabilize supported iridium efficient new electrocatalyst for Li–CO2
pair-site catalysts. J. Am. Chem. Soc. 141, batteries. Adv. Funct. Mater. 29, 1904206.
56. Zaera, F. (1996). On the mechanism for the
hydrogenation of olefins on transition-metal 4010–4015.
77. Zhao, L., Zhang, Y., Huang, L.B., Liu, X.Z.,
surfaces: the chemistry of ethylene on Pt (111).
67. Yang, D., Xu, P., Guan, E., Browning, N.D., and Zhang, Q.H., He, C., Wu, Z.Y., Zhang, L.J., Wu,
Langmuir 12, 88–94.
Gates, B.C. (2016). Rhodium pair-sites on J., Yang, W., et al. (2019). Cascade anchoring
57. Argo, A.M., Odzak, J.F., Goellner, J.F., Lai, F.S., magnesium oxide: synthesis, characterization, strategy for general mass production of high-
Xiao, F.-S., and Gates, B.C. (2006). Catalysis by and catalysis of ethylene hydrogenation. loading single-atomic metal-nitrogen catalysts.
oxide-supported clusters of iridium and J. Catal. 338, 12–20. Nat. Commun. 10, 1278.
rhodium: hydrogenation of ethene, propene,
and toluene. J. Phys. Chem. B 110, 1775–1786. 68. Binder, A., Seipenbusch, M., Muhler, M., and 78. Babucci, M., Sarac Oztuna, F.E., Debefve, L.M.,
Kasper, G. (2009). Kinetics and particle size Boubnov, A., Bare, S.R., Gates, B.C., Unal, U.,
58. Bianchini, C., Farnetti, E., Graziani, M., Kaspar, effects in ethene hydrogenation over and Uzun, A. (2019). Atomically dispersed
J., and Vizza, F. (1993). Molecular solid-state supported palladium catalysts at atmospheric reduced graphene aerogel-supported iridium
organometallic chemistry of tripodal pressure. J. Catal. 268, 150–155. catalyst with an iridium loading of 14.8 wt%.
(polyphosphine) metal complexes. Catalytic ACS Catal. 9, 9905–9913.
hydrogenation of ethylene at iridium. J. Am. 69. Molero, H., Stacchiola, D., and Tysoe, W.
Chem. Soc. 115, 1753–1759. (2005). The kinetics of ethylene hydrogenation 79. Karim, A.M., Howard, C., Roberts, B., Kovarik,
catalyzed by metallic palladium. Catal. Lett. L., Zhang, L., King, D.L., and Wang, Y. (2012). In
59. Uzun, A., and Gates, B.C. (2009). Dynamic 101, 145–149. HSitu X-ray absorption fine structure studies on
structural changes in a molecular zeolite- the effect of pH on Pt electronic density during
supported iridium catalyst for ethene 70. Argo, A.M., and Gates, B.C. (2003). MgO-
aqueous phase reforming of glycerol. ACS
hydrogenation. J. Am. Chem. Soc. 131, 15887– supported Rh6 and Ir6: structural
Catal. 2, 2387–2394.
15894. characterization during the catalysis of ethene
hydrogenation. J. Phys. Chem. B 107, 5519–
xlü, G.,
60. Bernales, V., Yang, D., Yu, J., Gümüs 5528. 80. Ravel, B., and Newville, M. (2005).
Cramer, C.J., Gates, B.C., and Gagliardi, L. Athena, Artemis, Hephaestus: data analysis for
(2017). Molecular rhodium complexes 71. Lu, J., Serna, P., and Gates, B.C. (2011). Zeolite- X-ray absorption spectroscopy using IFEFFIT.
supported on the metal-oxide-like nodes of and MgO-supported molecular iridium J. Synchrotron Radiat. 12, 537–541.
metal organic frameworks and on zeolite HY: complexes: support and ligand effects in
catalysts for ethylene hydrogenation and catalysis of ethene hydrogenation and H–D 81. Newville, M. (2001). IFEFFIT: interactive XAFS
dimerization. ACS Appl. Mater. Interfaces 9, exchange in the conversion of H2 + D2. ACS analysis and FEFF fitting. J. Synchrotron Radiat.
33511–33520. Catal. 1, 1549–1561. 8, 322–324.

61. Guan, E., and Gates, B.C. (2018). Stable 72. Martinez-Macias, C., Serna, P., and Gates, B.C. 82. Zabinsky, S.I., Rehr, J.J., Ankudinov, A., Albers,
rhodium pair sites on MgO: influence of ligands (2015). lsostructural zeolite-supported rhodium R.C., and Eller, M.J. (1995). Multiple-
and rhodium nuclearity on catalysis of ethylene and iridium complexes: tuning catalytic activity scattering calculations of x-ray-absorption
hydrogenation and H–D exchange in the and selectivity by ligand modification. ACS spectra. Phys. Rev. B Condens. Matter 52,
reaction of H2 with D2. ACS Catal. 8, 482–487. Catal. 5, 5647–5656. 2995–3009.

Cell Reports Physical Science 2, 100495, July 21, 2021 13


Cell Reports Physical Science, Volume 2

Supplemental information

18.1% single palladium atom catalysts


on mesoporous covalent organic framework
for gas phase hydrogenation of ethylene
Chun-Te Kuo, Yubing Lu, Pezhman Arab, K. Shamara Weeraratne, Hani El-
Kaderi, and Ayman M. Karim
Table S1. Summary of EXAFS modeling of the Pd-COF sample after different pretreatments. All spectra
were collected at 30 °C in the same gas used for the pretreatment.

Sample Fresh Dried Reduced

He at
Pretreatment He at 30 °C H2 at 30 °C
110 °C

NPd-N 1.9±0.3 1.7±0.3 1.9 ±0.1

NPd-Cl 2.2±0.4 2.2±0.4 1.7±0.1

NPd-Pd - - 0.8±0.2

RPd-N (Å) 2.02±0.01 2.03±0.01 2.02±0.01

RPd-Cl (Å) 2.31±0.01 2.31±0.01 2.33±0.01

RPd-Pd (Å) - - 2.80±0.01


3 2
σ2 Pd-N ×10 (Å ) 3±2 4±1 4±1

3 2
σ2 Pd-Cl × 10 (Å ) 3±2 4±1 4±1
3 2
σ2 Pd-Pd × 10 (Å ) - - 8±2
a b b b
∆E0 Pd-N/Cl (eV) 12 12 12

∆E0 Pd-Pd (eV) - - 5±2

Reduced χ2 57 73 36

R-factor 0.0017 0.0007 0.0002


a
ΔE0 was the same for Pd-N and Pd-Cl since these paths were calculated using the same model.
b
value was fixed during the fit. The edge energy from the XANES spectra of the 3 samples was almost
identical.

Notation: N, coordination number of absorber-backscatterer pair; R, radial absorber-backscatterer distance;


σ2 , the mean square displacement of the half-path length and represents the stiffness of the bond for a
single scattering path, ∆E0, correction to the threshold energy.
(a)

(b)

(c)

2
Figure S1. Pd K-edge k -weighted EXAFS data in k-space for (a) PdCl2@ILCOF-1 dried, (b) PdCl2 (c)
PdO.
(a)

(b)

(c)

Figure S2. Pd K-edge X-ray absorption spectroscopy of PdCl2@ILCOF-1 catalysts (fresh and dried).
2
XANES (a), and EXAFS magnitude (b) and imaginary (c) part of the of the Fourier transformed k -
-1
weighted χ(k) data, Δk=3-12.0 Å .
(a)

(b)

Figure S3. EXAFS fit to PdCl2@ILCOF-1 dried. Contribution of different scattering paths. (N path =Pd-N
and Cl path = Pt-Cl). (a) r-space magnitude part. (b) r-space imaginary part.
Figure S4. stability test of the PdCl2@ILCOF-1. Dependence of TOF on C2H4 partial pressure was
measured at a constant H2 partial pressure of 5 kPa. The stability test was performed at the initial
condition after ~6 h measurement.
Note S1 related to state of the catalyst in H2 and after reaction

In order to understand the catalyst structure after reaction condition, DRIFTS of CO adsorption after in-situ
ethylene hydrogenation, 0.5 kPa C2H4 and 5 kPa H2 in N2 balance at 20 °C, was conducted and is shown
-1
in the orange spectrum in Figure S5. The linearly bound CO band red-shifted from 2130 to 2122 cm , while
-1
a smaller portion did not shift (see shoulder band at 2130 cm ). It is plausible that the red-shift is due to a
change in the electronic property of Pd after exposure to H2 and ethylene (e.g. presence of H2 and/or
1-6 −1
ethylene on Pd and or the support). Additionally, a CO band at 1920 cm which can be assigned to bridge
7-12
adsorbed CO on Pd clusters that formed during the exposure to H2 flow. However, we did not observe
-1
a linear CO band in the 2000-2100 cm range. Our DRIFTS results suggest that under reaction conditions
a small percentage of Pd single atoms agglomerate into clusters, likely smaller than 1 nm. This is consistent
with the EXAFS results showing that Pd was still mostly present as single atoms (EXAFS after H2 treatment,
Table S1, Figure S5 and S6), where the Pd-N and Pd-Cl coordination numbers were still close to those
before reaction. Additionally, reduction of the PdCl2@ILCOF-1 catalyst at 120 °C resulted in almost
-1
complete reduction of the Pd and formation of nanoparticles as indicated by the metallic linear (2067 cm )
-1
and bridge (1919 cm ) CO adsorption peaks in DRIFTS (green spectrum in Figure S5). From the
-1
comparison between the two spectra in Figure S5, we could see that the small peak at 1922 cm only
represents a very small proportion of the Pd, and the large majority of the Pd is still present as single atoms
-1
(CO peak at 2122 cm ) after ethylene hydrogenation.

Figure S5. DRIFTS spectra of CO adsorption on the PdCl2@ILCOF-1 catalyst (1) at -40 °C after in-situ
ethylene hydrogenation reaction (0.5 kPa C2H4 and 5 kPa H2 in N2 balance) at 20 °C (orange) and (2) at
25 °C after H2 treatment at 120 °C (green).
(a)

(b)

(c)

Figure S6. Pd K-edge X-ray absorption spectroscopy of PdCl2@ILCOF-1 catalysts (dried and H2 flow at
2
30 °C). XANES (a), and EXAFS magnitude (b) and imaginary (c) part of the of the Fourier transformed k -
-1
weighted χ(k) data, Δk=3-12.0 Å .
(a)

(b)

2
Figure S7. Pd L3-edge EXAFS spectra and fit of the Fourier transformed k -weighted χ(k) data (∆k= 3-
-1
12.0 Å ) for the PdCl2@ILCOF-1catalyst after H2 flow at 30 °C. (a) EXAFS magnitude and (b)imaginary
2 -1
part of the of the Fourier transformed k -weighted χ(k) data, Δk=3-12.0 Å
Supplemental References

1. Bertolini, J.; Delichere, P.; Khanra, B.; Massardier, J.; Noupa, C.; Tardy, B., Electronic properties of
supported Pd aggregates in relation with their reactivity for 1, 3-butadiene hydrogenation. Catal. Lett.
1990, 6 (2), 215-223.
2. Chen, L.-Y.; Hunter, G. W.; Neudeck, P. G.; Bansal, G.; Petit, J. B.; Knight, D., Comparison of
interfacial and electronic properties of annealed Pd/SiC and Pd/SiO2/SiC Schottky diode sensors. J.
Vacuum Sci. Technol. A: Vacuum, Surfaces, and Films 1997, 15 (3), 1228-1234.
3. Chou, P.; Vannice, M. A., Calorimetric heat of adsorption measurements on palladium: I. Influence of
crystallite size and support on hydrogen adsorption. J Catal 1987, 104 (1), 1-16.
4. Louie, S. G., Hydrogen on Pd (111): Self-Consistent Electronic Structure, Chemical Bonding, and
Photoemission Spectra. Phys. Rev. Lett. 1979, 42 (7), 476.
5. Switendick, A., The change in electronic properties on hydrogen alloying and hydride formation. In
Hydrogen in Metals I, Springer: 1978; pp 101-129.
6. Sykes, E. C. H.; Fernández-Torres, L. C.; Nanayakkara, S. U.; Mantooth, B. A.; Nevin, R. M.; Weiss,
P. S., Observation and manipulation of subsurface hydride in Pd {111} and its effect on surface
chemical, physical, and electronic properties. Proc. Natl. Acad. Sci. 2005, 102 (50), 17907-17911.
7. Cabilla, G. C.; Bonivardi, A. L.; Baltanás, M. A., Characterization by CO/FTIR spectroscopy of
Pd/silica catalysts and its correlation with syngas conversion. Catal. Lett. 1998, 55 (3-4), 147-156.
8. Gelin, P.; Siedle, A. R.; Yates Jr, J. T., Stoichiometric adsorbate species interconversion processes in
the chemisorbed layer. An infrared study of the carbon monoxide/palladium system. J. Phys. Chem.
1984, 88 (14), 2978-2985.
9. Marchesini, F.; Gutierrez, L.; Querini, C.; Miró, E., Pt, In and Pd, In catalysts for the hydrogenation of
nitrates and nitrites in water. FTIR characterization and reaction studies. Chem. Eng. J. 2010, 159 (1-
3), 203-211.
10. Palazov, A.; Chang, C.; Kokes, R., The infrared spectrum of carbon monoxide on reduced and
oxidized palladium. J. Catal. 1975, 36 (3), 338-350.
11. Sheu, L.; Knözinger, H.; Sachtler, W., Ship-in-a-bottle formation of Pd13(CO)x clusters in zeolite NaY.
Catal. Lett. 1989, 2 (3), 129-137
12. Soma-Noto, Y.; Sachtler, W., Infrared spectra of carbon monoxide adsorbed on supported palladium
and palladium-silver alloys. J. Catal. 1974, 32 (2), 315-324.

You might also like