You are on page 1of 10

Molecular Catalysis 495 (2020) 111157

Contents lists available at ScienceDirect

Molecular Catalysis
journal homepage: www.elsevier.com/locate/mcat

The selective hydrogenation of nitroarenes and alkenes catalyzed by Pd@ T


MOFs: The role of electronic interactions between Pd nanoparticles and
MOFs on the reaction
Jiaxian Xua, Fei Chenb,c,**, Xuran Xua, Guo-Ping Lua,*
a
School of Chemical Engineering, Nanjing University of Science & Technology, Xiaolingwei 200, Nanjing, 210094, PR China
b
Nanjing Institute of Environmental Sciences, Ministry of Ecology and Environment, Jiangwangmiao 8, Nanjing, 210042, Jiangsu, China
c
Jiangsu Collaborative Innovation Center of Atmospheric Environment and Equipment Technology (CICAEET), Nanjing University of Information Science & Technology,
Ningliu 219, Nanjing, 210044, Jiangsu, China

ARTICLE INFO ABSTRACT

Keywords: The electronic metal-support interactions that are well-known in metal-loaded metal oxides, are also disclosed in
the electronic metal-support interactions Pd@MOFs. According to the experimental and calculation results, the superior catalytic performance of Pd@
palladium nanoparticles MIL-125(Ti) is mainly attributed to the higher ratio of Pd0/Pd2+ and electronic density of Pd nanoparticles
metal-organic frameworks which are triggered by stronger Lewis base sites and less Lewis acid sites of MIL-125(Ti) respectively. Therefore,
MIL-125(Ti)
the Pd@MIL-125(Ti) catalyzed hydrogenation of nitroarenes and alkenes can be carried out in water using low
the hydrogenation of nitroarenes and alkenes
Pd usage (0.12 mol%) with excellent selectivity under mild conditions (room temperature, 1 atm H2). Moreover,
the aqueous catalytic system still stays in reactor by simple extraction, and can be reused without further
treatment, thus resulting in low environmental factor.

1. Introduction transition states, but also can interaction with MNPs by electronic ef-
fects [11–14]. If this is the case, it can be expected that the similar
Metal-organic frameworks (MOFs) have become a hot research area electronic interactions between metal-oxygen clusters of MOFs and
owing to their highly tunable porosity, high density of accessible metal MNPs also exist since metal-oxygen clusters in MOFs have similar
sites, high specific surface area and good designability, so they are properties and structures with metal oxides. Nevertheless, few attempts
suitable for various applications, especially in catalysis [1–4]. In most address this potentially crucial observation [15,16]. More recently,
related reports, the synergistic effects between MOFs and metal nano- Yamauchi’s group has disclosed a charge-transfer interaction between
particles (MNPs) for enhance catalysis are generalized that (1) MOFs’ Pt NPs and MOFs [15]. The electronic interaction between M-HKUST-1
ordered, stable and nano pores enable their inherent superiority in and Pd NPs has also been demonstrated by Zeng’s group [16]. Ac-
confining MNPs, for improved catalytic performance [1,5–7]; (2) The cordingly, we envisage that control of the electronic effects between
functional groups of organic linkers and metal nodes can also absorb MNPs and MOFs by screening different MOFs may lead to more efficient
and activate the reactants and transition states as the cocatalysts heterogeneous catalysts (M@MOFs).
[1,8–10]. However, all of that represent the simple functional synergy On the other hand, catalytic hydrogenations are one of the most
between these two components, which offers limited promotion for the useful methodologies for the reduction of organic substances, such as
development of the related research. Therefore, it is still appealing and nitroarenes and alkenes [17–20]. A large number of heterogeneous
significant for deeper understanding and insight into the synergistic catalysts have been developed for the hydrogenations of nitroarenes
effects between MNPs with MOFs. and alkenes [21–26]. Among of these catalysts, M@MOFs exhibit ex-
It has been well known that the metal oxides, a classical kind of cellent performance on these transformations [27]. Nevertheless, all of
carriers for MNPs, generally possess both Lewis acid and base sites, these works only focus on employing MOFs as the porous carriers to
which not only provide the active sites to activate the reactants and stabilize and disperse MNPs [28–32]. No report addresses the role of the

Corresponding author.

Corresponding author at: Nanjing Institute of Environmental Sciences, Ministry of Ecology and Environment, Jiangwangmiao 8, Nanjing, 210042, Jiangsu,
⁎⁎

China.
E-mail addresses: chenfei@nies.org (F. Chen), glu@njust.edu.cn (G.-P. Lu).

https://doi.org/10.1016/j.mcat.2020.111157
Received 21 April 2020; Received in revised form 20 May 2020; Accepted 29 July 2020
Available online 11 August 2020
2468-8231/ © 2020 Elsevier B.V. All rights reserved.
J. Xu, et al. Molecular Catalysis 495 (2020) 111157

electronic MNPs-MOFs interactions in the hydrogenations of nitroar- Table 1


enes and alkenes. Meanwhile, there is still room for improvement of the The catalytic performance of Pd@MOFs on the hydrogenation of 1aa.
substrate scope and catalytic performance.
Guided by the principles of green chemistry, the hydrogenation of
nitroarenes and alkenes would be carried out using a recyclable catalyst
in water as the only medium at room temperature under atmospheric
hydrogen with high yields and selectivity. Based on these results and
our interesting in exploring M@MOFs for catalysis [33–36], we disclose Entry Catalyst t (h) Yield (%)b Selectivity (%)b

that Pd@MIL-125(Ti) has superior catalytic performance than other 1 Pd@UiO-66(Zr) 3 15 99


Pd@MOFs for the hydrogenations of nitroarenes and alkenes in water 2 Pd@MIL-101(Fe) 3 69 99
under mild conditions (rt, 1 atm H2), which may be attribute to the 3 Pd@MIL-125(Ti) 3 84 99
electronic effects between Pd nanoparticles and MIL-125(Ti). 4 Pd@MIL-125(Ti) 6 99 99
5 Pd@MOF-74(Cu) 3 nr /
6 Pd@MOF-74(Zn) 3 nr /
2. Experimental 7 Pd@MOF-5(Zn) 3 19 99
8 Pd@MOF-74(Ni) 3 5 99
2.1. Materials and Methods 9 MIL-125(Ti) 3 nr /
10 Pd@TiO2 3 29 99
11 Pd NPsc 3 18 /
All chemical reagents are obtained from commercial suppliers and 12 Pd@MIL-125(Ti)d 1 94 99
used without further purification. GC-MS was performed on an ISQ 13 Pd@MIL-101(Fe)d 1 76 99
Trace 1300 in the electron ionization (EI) mode. GC analyses are per- 14 Pd@MOF-74(Zn)d 1 8 99
formed on an Agilent 7890A instrument (Column: Agilent 19091J-413: a
Conditions: 0.5 wt.% Pd@MOF 0.12 mol% (∼5 mg), 1a 0.2 mmol, H2O
30 m × 320 μm × 0.25 μm, carrier gas: H2, FID detection. The crystal
1.5 mL, 1 atm H2, rt.
structure of the synthesized catalysts was recorded by X-ray diffraction b
The yields and selectivity were determined by GC using anisole as the in-
(XRD) using a D8ADVANCED X-ray diffractometer, employing a scan-
ternal standard.
ning rate of 0.1 s-1. Scanning electron microscopy (SEM) spectras were c
Pd NPs 0.5 mol% was used.
taken using a Hitachi S-4800 apparatus on a sample powder previously d
2 wt.% Pd@MOF 0.5 mol% (∼5 mg) was used.
dried and sputter-coated with a thin layer of gold. High Resolution
Transmission electron microscopy (HRTEM) was performed on a
Philips-FEI Tecnai G2 F20 operating at 300 kv. X-ray photoelectron
spectroscopy (XPS) was performed on an ESCALAB 250Xi spectrometer,
using an Al Kα X-ray source (1350 eV of photons). Inductively coupled
plasma mass spectrometry (ICP-MS) was performed on an Optima 7300
DV. Raman spectra were recorded on an Aramis with a wavelength of
532 nm. Temperature-programmed desorption (TPD) of NH3 and CO2-
TPD were conducted on a Quantachrome TPRWin v3.52 instrument.
The samples were pretreated in He flow at 150 °C with a rate of 15 mL
min−1 for 30 min and cooled to 50 °C, and then swept in CO2 (NH3)
flow with a rate of 15 mL min-1 for 40 min. After treatment in He flow
for 50 min to remove physical adsorption, the samples were raised at a
heating rate of 10 °C min-1 to 500 °C, and the signals were monitored by
using a TCD detector. BET surface areas were recorded with N2 ad-
sorption/desorption isotherms at 77 K on a Micromeritics ASAP 2920
instrument. Before measurements, the samples were degassed at 150 °C
for 12 h. The Raman spectra were obtained using confocal Raman
spectroscopy (inVia-Reflex) employing 785 nm radiation (3 mW).

2.2. Computational Details

Spin-polarized DFT calculations as implemented in the Vienna ab


initio simulation package were performed using the plane-wave pseu-
dopotential basis set [37]. The ion-electron interactions were treated
with the projected augmented wave pseudopotentials [38], and the
generalized gradient approximation parametrized by Perdew, Burke,
and Ernzerhof was used to describe the electronic exchange-correlation
energy [39]. The outer-shell electrons, i.e., 1 s1 of H, 2s22p2 of C, 2s22p4
of O, 3s23p63d24s2 of Ti, 3d64s2 of Fe, 3d104s2 of Zn, 4s24p64d25s2 of
Zr, and 4d10 of Pd were explicitly calculated. An effective Hubbard
parameter of 3.5 eV was added for Ti, Fe, Zn, and Zr to mitigate the self-
interaction errors. The plane-wave basis was expanded up to a cut-off Fig. 1. XPS spectra of (a) Pd 3d of 0.5 wt.% Pd@MIL-125(Ti) and 0.5 wt%
energy of 400 eV. All structures were fully relaxed by the conjugate Pd@TiO2; (b) Pd 3d of 2 wt.% Pd@MIL-125(Ti), 2 wt.% Pd@MIL-101(Fe) and
gradient method until the force component on each atom was less than 2 wt.% Pd@MOF-74(Zn).
0.02 eV/Å, and the convergence criteria of total energy in the self-
consistent field method was set to 10-5 eV. Typical MOF clusters con- where the Brillouin zone integration was performed using the Γ point
sisted of Me (Me = Ti, Fe, Zn, or Zr), O, H, and carboxyl was adopted to with a Gaussian smearing width of 0.05 eV. The MOF-74(Zn) was in the
simulate the supporting substrates. The MIL-125(Ti), MIL-101(Fe) or form of one-dimensional chain and two units (6 Zn atoms and 13.424 Å
UiO-66(Zr) cluster was put in a cubic box with a side length of 30 Å,

2
J. Xu, et al. Molecular Catalysis 495 (2020) 111157

Fig. 2. XPS spectra of O 1s of MIL-125(Ti) and 0.5 wt.% Pd@MIL-125(Ti).

Fig. 4. The adsorption energies of Pd4 on different SBUs of MOFs (a) MIL-
125(Ti); (b) MIL-101(Fe); (c) MOF-74(Zn) (Brown is C; red is O; white is H; light
blue is Ti; gold is Fe; dark blue is Zn).

Fig. 5. The adsorption free energy of H atom (H*) on Pd(111) with different
Fig. 3. (a) CO2-TPD and (b) NH3-TPD profiles. bader charge.

in length, 3 k-points adopted) were included in the super cell. The divided into two parts. Each part was respectively added into 1.8 mL
adsorption energy (ΔEads) was calculated by: titanium(IV) butoxide (TBOT) and 1.65 g H2BDC. After ultrasonication
and dissolution, the two parts were combined in a 50 mL Teflon-lined
ΔEads = E [substrate + adsorbate] –E [substrate] ‒E [adsorbate], stainless steel autoclave, which was heated to 150 ℃ for 48 h. After
where E [substrate + adsorbate] and E [substrate] are the DFT energies of the cooling down to room temperature, wash with methanol several times,
system with and without adsorbate, respectively, and E [adsorbate] is the and dry.
DFT energy for the adsorbate in a vacuum. The adsorption free energy MIL-101(Fe) [42]: 200 mg of FeCl3·6H2O and 116 mg of H2BDC
(ΔG) was further corrected by zero-point energy and entropy. Bader were dissolved in 15 mL of DMF solvent, then move it into Teflon-lined
charge analysis was used to determine the charge distribution [40]. stainless steel autoclave, which was heated to 120 ℃ for 24 h.After
cooling down to room temperature, wash with methanol several times,
and dry.
2.3. The procedure for the synthesis of MOFs
MOF-74(Zn) [43]: 125 mg of 2,5-dihydroxyterephthalic acid
(DHTA) and 565 mg of Zn(NO3)2·6H2O were dissolved in 30 mL of DMF
MIL-125(Ti) (MIL means Materials of Institute Lavoisier who first
by ultrasound applied for 10 min. Then 2 mL of water was added, which
synthesizes this kind of MOFs, the digital stands for different structure
were capped tightly and placed in a 120℃ oven for 24 h. After cooling
in MIL series) [41]: a mixture of DMF (27 mL) and CH3OH (3 mL) was

3
J. Xu, et al. Molecular Catalysis 495 (2020) 111157

Fig. 6. SEM images of (a) MIL-125(Ti), (b) 0.5 wt.% Pd@MIL-125(Ti); TEM images of (c) 0.5 wt.% Pd@MIL-125(Ti), (f) 0.5 wt.% Pd@TiO2; (d), (e) HRTEM images of
0.5 wt.% Pd@MIL-125(Ti).

down to room temperature,wash with methanol several times, and dry. of MOFs during the synthesis of Pd@MOF. Low temperature and the use
of as little NaBH4 as possible are recommend to exclude the effect of
2.4. The general procedure for the synthesis of 0.5 wt.% Pd@MOF NaBH4 on the crystal structure destruction.

500 mg of MOF was added to 15 mL of water and sonicated. 0.5 mL 2.5. The general procedure for the selective hydrogenation of nitroarenes or
of a sodium tetrachloropalladate solution with a palladium content of alkenes
5 mg/mL was added to the solution, stirring at room temperature for
2 h. Then, the solution was cooled to 0 °C, and sodium borohydride A mixture of nitroarene 1 (or alkene 3) 0.2 mmol, Pd@MIL-125(Ti)
aqueous solution (2 mg in 2 mL H2O) slowly dorpped to the mixture, 0.12 mol% (∼5 mg) were added in water (1.5 mL), which was stirred
and stirred at 0 °C for 2 h. Finally, the resultant precipitate was sepa- under 1 atm H2 at room temperature for 6 h. After the reaction was
rated by centrifugation and washed repeatedly with CH3OH. The ob- completed, the mixture was extracted with ethyl acetate (1 mL × 3).
tained powder sample was dried in a vacuum oven at 80 °C for 12 h. The organic layer was collected by syringes and filtered through a bed
Because NaBH4 has strong basicity, it will destroy the crystal structure of silica gel layered over Celite. The product were analyzed by GC-MS.

4
J. Xu, et al. Molecular Catalysis 495 (2020) 111157

Fig. 7. TEM images and their corresponding element mapping images of Pd@MIL-125(Ti).

filtered through a bed of silica gel layered over Celite. The volatiles
were removed in vacuo to afford the crude product 2a. The aqueous
solution after extracted, containing Pd@MIL-125(Ti), was submitted to
another reaction cycle. To the aqueous phase, 1a 0.2 mmol was added
and the reaction stirred under 1 atm H2 for 6 h at room temperature.
The extraction cycle was then repeated for the separation of 2a and
aqueous catalytic system.

3. Results and discussion

To verify the feasibility of our proposed assumption, several 0.5 wt.


% Pd@MOFs (see Table 1) were prepared by the classical impregnation
reduction method (see Experimental section 2.4). The hydrogenation of
4-nitroanisole 1a was selected as the model reaction to test the catalytic
performance of these Pd@MOFs at room temperature under 1 atom H2.
Only Pd@MIL-125(Ti) and Pd@MIL-101(Fe) could provide satisfactory
results (entries 2, 3). 99% yield of 2a could be obtained using Pd@MIL-
125(Ti) as the catalyst by prolonging the reaction time to 6 h (entry 4).
No product was found in the presence of MIL-125(Ti), indicating that
Pd sites is the main catalytic active sites. As a comparison, Pd@TiO2
were also tested, resulting in a poor yield of 2a (29%, entry 10). Low
yield of product 2a was obtained when naked Pd NPs was used, in-
dicating the crucial role of MOFs on the improvement of Pd’s perfor-
mance (entry 11). For both 0.5 wt.% and 2 wt.% Pd@MOFs, the order
of catalytic performance is the same: Pd@MIL-125(Ti) > Pd@MIL-
101(Fe) > Pd@MOF-74(Zn) (entries 12-14).
According to previous reports [15,16], the XPS tests were performed
to confirm the electronic interactions between MOFs and Pd NPs
Fig. 8. (a) Nitrogen adsorption-desorption isotherms of Pd@MIL-125(Ti) and (Fig. 1). No obvious Pd 3d peak is found in 0.5 wt.% Pd@MIL-125(Ti)
(b) the corresponding pore-size distribution. which may be attributed to the encapsulation of Pd NPs in the porous
structure of MIL-125(Ti) (Fig. 1a). As a comparison, Pd 3d peaks of
0.5 wt.% Pd@TiO2 are observed because the Pd NPs should be loaded
The yield and selectivity were determined by GC with 4-anisole as an
on the surface of TiO2. Although the electronic states of Pd in 2 wt.%
internal standard. The volatiles were removed in vacuo to afford the
Pd@MOFs don't directly represent those in 0.5 wt.% Pd@MOFs, but we
crude product. Further column chromatography on silica gel was re-
believe that the order of the Pd electronic state is the same owing to the
quired to afford the pure desired products.
same order of catalytic performance (Table 1). Moreover, no obvious Pd
3d peak is found in 0.5 wt.% Pd@MIL-125(Ti) which makes it im-
2.6. Recycling aqueous catalytic system possible to study the electronic states of Pd. Therefore, three 2 wt.%
Pd@MOFs instead of 0.5 wt.% Pd@MOFs were selected to determine
After reaction completion, the mixture was then extracted with the electronic states of Pd (Fig. 1b). The order of the Pd°/Pd2+ ratio is
EtOAc (3 × 1 mL). The organic layer was collected by syringes and Pd@MIL-125(Ti) > Pd@MIL-101(Fe) > Pd@MOF-74(Zn) (Table S2),

5
J. Xu, et al. Molecular Catalysis 495 (2020) 111157

Scheme 1. The Pd@MIL-125(Ti)-catalyzed selective hydrogenation of nitroarenes.a,b.


a
Conditions: nitroarene 1 0.2 mmol, Pd@MIL-125(Ti) 0.12 mol%, H2O 1.5 mL, H2 1 atm, rt, 6 h; Y and S meant the yield and selectivity of 2, which was determined
by GC using anisole as the internal standard. b The bold parts of all products 2 were the hydrogenation groups. c The reaction time was 12 h.

Scheme 2. Molecules 5-7 with biological activities from 2d, 2i, 2 l respectively.

6
J. Xu, et al. Molecular Catalysis 495 (2020) 111157

Scheme 3. The Pd@MIL-125(Ti)-catalyzed selective


hydrogenation of alkenes.a,b.
a
Conditions: alkenes 0.2 mmol, Pd@MIL-125(Ti)
0.12 mol%, H2O 1.5 mL, H2 1 atm, rt, 6 h; Y and S
meant the yield and selectivity of 4, which was de-
termined by GC using anisole as the internal standard.
b
The bold parts of all products 4 were the hydro-
genation groups.

Table 2
Comparison of Pd@MIL-125(Ti) catalyst with reported nanocatalysts for selective hydrogenation of 3a to 4a.
Entry Catalyst Conditions Y/S (%)a Refs.

1 PtFeNi ZNWs + AlCl3 70 °C, 3.5 h, 1 atm H2, EtOH 89.7/94.9 [57]
1 mol%
2 Pt3Fe/CNT 60 °C, 0.5 h, 20 atm H2, H2O 60.3/97.2 [58]
1 mol%
3 Pt3Ni@Ni32Cu(OH)2 50 °C, 3 h, 30 atm H2, H2O/EtOH 87.9/89.2 [59]
-ZNWs 0.68 mol%
4 Cu/SiO2 0.11 mol% 100 °C, 15 h, 25 atm H2, toluene 99/90 [60]
5 Pd/LaFeO3 0.1 mol% 80 °C, 10 atm H2, 3 h, cyclohexane 90.6/100 [61]
6 (SiO2-ALD)-Pt/Al2O3 0.2 mol% rt, 30 atm H2, 8 h, isopropanol 95/65 [62]
7 Pd-0.3Ag/MCM-41 0.1 mol% 110 °C, 10 atm H2, 2 h, cyclohexane 99/98.6 [63]
8 Pd@MIL-125(Ti) rt, 1 atm H2, 6 h, H2O 93/95 This work
0.12 mol%

a
Y/S means Yield/Selectivity.

meanwhile the order of Pd NPs electronic density is the same, thus of acid and base sites [33,47,48]. The acid sites of Pd@MOFs include
suggesting the electronic interactions between MOFs and Pd NPs Pd2+ and Mn+ (Mn+ = Ti4+, Fe3+, Zn2+) sites. As shown in Fig. 3a,
[15,16,36]. The naked Pd NPs has lower ratio of Pd0/Pd2+ and elec- the order of numbers of acid sites is Pd@MOF-74(Zn) > Pd@MIL-
tronic density of Pd0 than Pd@MIL-125(Ti), which proves the stabi- 101(Fe) > Pd@MIL-125(Ti) (Table S1). The basic sites of Pd@MOFs are
lizing of MIL-125(Ti) on Pd NPs and the electronic interactions between O sites including Ti-O and C-O. According to the CO2-TPD results
MIL-125(Ti) and Pd NPs. (Fig. 3b and Table S1), the order of numbers of Lewis basic sites is Pd@
According to O 1s spectrum of MIL-125(Ti) (Fig. 2), the peaks at MOF-74(Zn) ≈ Pd@MIL-125(Ti) > Pd@MIL-101(Fe), and the Lewis
binding energies of about 530.2 eV arises from lattice oxygen species, base sites of Pd@MIL-125(Ti) are stronger than Pd@MOF-74(Zn).
and oxygen components in the BDC (terephthalic acid) may give a peak For a solid catalyst, it can simultaneously have more basic and acid
with a binding energy of 531.8 eV [33]. An obvious decrease of oxygen sites than other solid catalysts [33,47,48], because the basic and acidic
single of the BDC in the O 1s region implies that the structural collapse sites are stationary, which will not neutralize each other, and the
takes place during the synthesis process of Pd@MIL-125(Ti), which also amount of basic and acidic sites mainly depends on its surface area and
is certified by the results of XRD and Raman results (Figs. S2, S3). structure. Therefore, it is reasonable that Pd@MOF-74(Zn) has most
Therefore, the crystal structure of MIL-125(Ti) has been destroyed and amount of basic sites and acidic sites.
turned into an amorphous compound after loading Pd NPs. Based on TPD and XPS results, some points can be concluded as
It is well known that Lewis acid and base sites of carriers can in- follow. There are synergy effects between basic and acidic sites for the
teraction with MNPs by electronic effects [11–14,16]. The CO2-TPD and catalytic performance of Pd@MOFs. (I) Stronger Lewis base sites (O
NH3-TPD (Fig. 3) of Pd@MOF-74(Zn), Pd@MIL-125(Ti) and Pd@MIL- sites) of MIL-125(Ti) may enhance the absorption and stability of Pd
101(Fe) were also employed to investigate the potential electronic ef- NPs in MIL-125(Ti), thus leading to higher ratio of Pd0/Pd2+. (II) Less
fects of between Lewis acid and base sites of MOFs with Pd NPs. The Lewis acid sites (Ti4+ and Pd2+ sites) of the catalyst may decrease the
structure of MOF-74, MIL-125 and MIL-101 are decomposed around interaction of Pd NPs with Ti4+ and Pd2+ sites, resulting in higher
400 °C [44–46], so the desorption peaks after 400 °C should result from electronic density of Pd NPs. (III) The superior catalytic performance of
the decomposition of MOFs, and only the TPD profiles from 50 °C to Pd@MIL-125(Ti) is ascribed to the higher ratio of Pd0/Pd2+ and elec-
400 °C are meaningful to the acid-base properties of all three catalysts. tronic density of Pd NPs that are beneficial to the activation of H2 (the
The TPD tests are performed under the same conditions, so the peak rate-determining step of the hydrogenation of nitroarenes) [21].
area of TPD spectra can qualitatively indicate the order of the number The crystal structure of MOF has been destroyed and turned into an

7
J. Xu, et al. Molecular Catalysis 495 (2020) 111157

the adsorption models between Pd4 and MOFs (Fig. 4), which supports
the conclusion (I) and (II).
The adsorption free energy of H atom (H*) on Pd(111) with dif-
ferent bader charge were also calculated to verify conclusion (III)
(Fig. 5). The results indicate that lower bader charge of Pd atom can
lead to higher adsorption free energy of H atom, which is consistent
with conclusion (III).
The morphology of the MIL-125(Ti) particles and Pd@MIL-125(Ti)
was assessed by SEM measurements (Fig. 6). MIL-125(Ti) possesses a
uniform diamond-like trustum of a pyramid structure (Fig. 6a). Com-
pared with MIL-125(Ti), the surface morphology of Pd@MIL-125(Ti)
(Fig. 6b) appears to be rougher owing to the destruction structure of
MIL-125(Ti) [49]. As depicted in Fig. 6c, it can be found that Pd@MIL-
125(Ti) exhibits a sheet-like structure, which coincides with the image
obtained by SEM, but no obvious Pd NPs are observed, suggesting that
Pd NPs are embedded in MIL-125(Ti). As a comparison, lots of Pd NPs
(∼5 nm) are observed on the surface of TiO2. These outcomes can ex-
plain the results of Pd 3d XPS spectra (Fig. 1a).
To determine the distribution of Pd species on this catalyst, the
representative high-resolution TEM (HRTEM) images were shown in
Fig. 6d, e. The nanoparticles of Pd are uniformly distributed in MIL-
125(Ti), which have clear crystalline fringe patterns. Moreover, the
latticefringe distance is about 0.227 nm which is close to the char-
acteristics of a face centered cubic Pd(111) (0.224 nm) indicating that
Pd is attached to MIL-125(Ti) [50]. The results of TEM-EDS indicate
that Pd@MIL-125(Ti) mainly contains C, O, Ti and Pd elements, and the
relatively uniform distribution of Pd was observed in MIL-125(Ti)
(Fig. 7).
The porous nature of Pd@MIL-125(Ti) was also investigated by N2
adsorption-desorption isotherms. The N2 adsorption-desorption iso-
therms for Pd@MIL-125(Ti) and MIL-125(Ti) display a type-IV hyster-
esis loop, characteristic of micro/mesoporous materials (Fig. 8). The
BET surface area of Pd@MIL-125(Ti) is 747 m2/g and the pore volume
is 0.55 cm3/g, which is lower than those of MIL-125(Ti) (1046 m2/g,
0.63 cm3/g). After loading Pd NPs, the surface area and pore volume
decreases, owing to the destruction of the crystal structure of MIL-
125(Ti). However, the catalytic efficiency of Pd@MIL-125(Ti) is im-
proved because Pd NPs adhere to the pores of MIL-125(Ti), providing
more active sites and facilitating the reaction of the catalyst with the
Fig. 9. The reusability and heterogeneous nature of the catalyst: (a) recycle substrate.
studies; (b) filtration experiments. In order to explore the substrate range and the general applicability
of 0.5 wt.% Pd@MIL-125(Ti), we explored the hydrogenation reduction
of various nitroarenes (Scheme 1). The sensitive functional groups, such
as -CHO, keto, cyano, amido and -Cl, were not substantially reduced in
yield, and the corresponding anilines were still obtained in higher yield
and selectivity (2f-2 j). Several useful aromatic amines were also syn-
thesized from corresponding nitroamines. For example, m-phenylene-
diamine (2d) can be used as the analytical reagent, resin curing agent,
polymerization inhibitor and raw material of pharmaceuticals (Scheme
2, Bruton's Tyrosine Kinase Inhibitor 5) [51,52]. 3,4-Dichloroaniline
Fig. 10. E factor studies. (2 j) is often used as the dye intermediate, pesticide intermediate and
biological component intermediate (cytotoxic agents 6) [53,54]. 8-
amorphous compound after loading Pd NPs, so the Pd@MOFs are Amino-6-methoxyquinoline (2 l) can also be used as an intermediate for
complex composites, and it is hard to give their real formula of the antimalarial drugs (compound 7) [55,56]. Notably, both nitro group
surface. However, we provide three simple models (Pd4 on second and C-C double bond were reduced in the cases of 2m-2o.
building units (SBUs) of MOFs) (Fig. 4) to calculate the absorption sites Considering the excellent catalytic performance of the catalyst on
and energies of Pd NPs in MOFs [2]. Notably, the geometric structure of the C-C double bond, some compounds containing both C = C and
MOFs will influence the absorption sites of Pd4 in SBUs (steric effects), C = O groups including -CHO, keto, carboxyl and ester groups was
which may affect the electronic interactions between Pd nanoparticles selected (Scheme 3). These hydrogenation still provided moderate to
(NPs) and MOFs. The order of the adsorption energies (MIL- excellent yields with good to excellent selectivity (4a-4d, 4f). The furan
125(Ti) > MOF-74(Zn) > MIL-125(Fe)) is consistent with the order of ring was also hydrogenated in the cases of 4e. On comparison of the
numbers of Lewis base sites (O2+ sites), which supports the conclusion catalytic activity of Pd@MIL-125(Ti) with other reported nanocatalysts
(I). Compared with MIL-101(Fe) and MOF-74(Zn), Pd4 has weaker in- for the synthesis of 4a from cinnamaldehyde 3a (Table 2), it can be
teraction with metal ions (acid sites) of MIL-125(Ti) and stronger in- observed that the present catalyst exhibit the best catalytic performance
teraction with O sites (basic sites) of MIL-125(Ti) in accordance with for this transformation.
Reusability is one of the key factors for heterogeneous catalysts. In

8
J. Xu, et al. Molecular Catalysis 495 (2020) 111157

order to investigate the recyclability of the prepared catalyst, we con- Declaration of Competing Interest
ducted the reusability experiment for Pd@MIL-125(Ti) in the model
reaction (Fig. 9a). After completion of reaction, the aqueous phase was The authors declare that they have no known competing financial
extracted several times with ethyl acetate to remove the reactants and interests or personal relationships that could have appeared to influ-
products from the aqueous phase, while the catalyst remains in the ence the work reported in this paper.
aqueous phase for the next reaction. After seven runs, there was no
significant change in catalytic efficiency. Acknowledgements
Most of MOFs are weakly coordinated, so they are slowly hydro-
lyzed in the presence of moisture to cause framework collapse [64], We gratefully acknowledge the Fundamental Research Funds for the
which is the one of the main reasons for the application of MOFs in Central Universities (30920021120), Key Laboratory of Biomass Energy
industrial catalysis. After nine runs, an obvious O-C = O single were and Material, Jiangsu Province (JSBEM201912), the National Natural
appeared at 292.4 (Fig. S4a) that was belong to the BDC [65], in- Science Foundation of China (21905089) and the Chinese Postdoctoral
dicating that the further structural damage took place during the re- Science Foundation (2019M662775) for financial support. This work
action and recycling process. Meanwhile the intensity of Pd 3d peaks was a project funded by the Priority Academic Program development of
was enhanced, implying the agglomeration and bareness of Pd NPs. The Jiangsu Higher Education Institution. We also thank Analysis and Test
obvious aggregation and deformation of MIL-125(Ti) particles observed Center Nanjing University & Technology for the help in obtaining the
by SEM and TEM (Fig. S5) also suggest the structural collapse of the Raman and XRD data.
catalyst. Furthermore, the loss of Pd was ignored after recycling nine
times based on the ICP results (Table S2). Therefore, the structural Appendix A. Supplementary data
collapse of Pd@MIL-125(Ti) and the agglomeration and bareness of Pd
NPs may be the two main factor for the deactivation of this material. Supplementary material related to this article can be found, in the
Although MOF-5(Zn) and UiO-66(Zr) have better stability than MIL- online version, at doi:https://doi.org/10.1016/j.mcat.2020.111157.
125(Ti) in water, but the performance of Pd@MOF-5(Zn) and Pd@UiO-
66(Zr) is unsatisfactory (Table 1, entries 1,7). References
The heterogeneous nature of the catalyst was also demonstrated by
a filtration experiment (Fig. 9b). The catalyst was filtered off after 1 h at [1] Q. Yang, Q. Xu, H.L. Jiang, Chem. Soc. Rev. 46 (2017) 4774–4808.
room temperature, and the isolated solution was allowed to react for a [2] Y.-Z. Chen, R. Zhang, L. Jiao, H.-L. Jiang, Coordin. Chem. Rev. 362 (2017) 1–23.
further 5 h under identical conditions. No further increase in yield was [3] V. Pascanu, G. Gonzalez Miera, A.K. Inge, B. Martin-Matute, J. Am. Chem. Soc. 141
(2019) 7223–7234.
observed. The E factor is a measure of the waste generated of a chemical [4] L. Zeng, X. Guo, C. He, C. Duan, ACS Catal. 6 (2016) 7935–7947.
process, and a lower E factor indicates higher efficiency and less pol- [5] G. Lu, S. Li, Z. Guo, O.K. Farha, B.G. Hauser, X. Qi, Y. Wang, X. Wang, S. Han,
lution of the process to the environment [66]. Hydrogenation of nitro X. Liu, J.S. DuChene, H. Zhang, Q. Zhang, X. Chen, J. Ma, S.C. Loo, W.D. Wei,
Y. Yang, J.T. Hupp, F. Huo, Nat. Chem. 4 (2012) 310–316.
group and C-C double bond catalyzed by a recyclable catalyst in water [6] Y. Liu, Y. Shen, W. Zhang, J. Weng, M. Zhao, T. Zhu, Y.R. Chi, Y. Yang, H. Zhang,
as the only medium can significantly reduce E factor. Based on 15% F. Huo, Chem. Commun. 55 (2019) 11770–11773.
organic solvent loss in the solvent recovery process of the model system [7] W. Zhang, B. Zheng, W. Shi, X. Chen, Z. Xu, S. Li, Y.R. Chi, Y. Yang, J. Lu, W. Huang,
F. Huo, Adv. Mater. 30 (2018) e1800643.
(Fig. 10), the E factor of the system (1 mmol scale) was determined to [8] M. hao, K. Yuan, Y. Wang, G. Li, J. Guo, L. Gu, W. Hu, H. Zhao, Z. Tang, Nature 539
be only 3.44 (for detailed procedures, see the supplementary data). (2016) 76–80.
[9] K.M. Choi, K. Na, G.A. Somorjai, O.M. Yaghi, J. Am. Chem. Soc. 137 (2015)
7810–7816.
4. Conclusions
[10] Y.-Z. Chen, Y.-X. Zhou, H. Wang, J. Lu, T. Uchida, Q. Xu, S.-H. Yu, H.-L. Jiang, ACS
Catal. 5 (2015) 2062–2069.
In summary, an efficient, recyclable catalyst Pd@MIL-125(Ti) by [11] P. Hu, Z. Huang, Z. Amghouz, M. Makkee, F. Xu, F. Kapteijn, A. Dikhtiarenko,
which the hydrogenation of nitroarenes and alkenes can be performed Y. Chen, X. Gu, X. Tang, Angew. Chem. 53 (2014) 3418–3421.
[12] T. Komanoya, T. Kinemura, Y. Kita, K. Kamata, M. Hara, J. Am. Chem. Soc. 139
in water using low Pd usage (0.12 mol%) with excellent selectivity and (2017) 11493–11499.
low E-factor under mild conditions (room temperature, 1 atm H2), has [13] C.T. Campbell, Nat. Chem. 4 (2012) 597–598.
been disclosed. The excellent catalytic performance of Pd@MIL-125(Ti) [14] J. An, Y. Wang, J. Lu, J. Zhang, Z. Zhang, S. Xu, X. Liu, T. Zhang, M. Gocyla,
M. Heggen, R.E. Dunin-Borkowski, P. Fornasiero, F. Wang, J. Am. Chem. Soc. 140
is owing to the synergistic effects between Pd NPs with MIL-125(Ti). (2018) 4172–4181.
According to the experimental and calculation results, it has been found [15] S. Yoshimaru, M. Sadakiyo, A. Staykov, K. Kato, M. Yamauchi, Chem. Commun. 53
that the synergistic effects between Pd NPs and MOFs are not only well- (2017) 6720–6723.
[16] Y.C. Tan, H.C. Zeng, Nat. Commun. 9 (2018) 4326.
known steric effects which can restrict the growth of Pd NPs and afford [17] W. Li, X. Cui, K. Junge, A.-E. Surkus, C. Kreyenschulte, S. Bartling, M. Beller, ACS
more active metal sites, but also the electronic effects that can regulate Catal. 9 (2019) 4302–4307.
the electronic states of Pd NPs. In the case of MIL-125(Ti), stronger [18] M.K. Karunananda, N.P. Mankad, ACS Catal. 7 (2017) 6110–6119.
[19] A. Corma, P. Serna, Science 313 (2006) 332–334.
Lewis base sites may enhance the absorption and stability of Pd NPs,
[20] J. Lam, K.M. Szkop, E. Mosaferi, D.W. Stephan, Chem. Soc. Rev. 48 (2019)
and less Lewis acid sites may decrease the interaction between Pd NPs 3592–3612.
with acid sites. Thus, Pd@MIL-125(Ti) has the higher ratio of Pd0/Pd2+ [21] J. Song, Z.-F. Huang, L. Pan, K. Li, X. Zhang, L. Wang, J.-J. Zou, Appl. Catal. B-
Environ. 227 (2018) 386–408.
and electronic density of Pd NPs than other Pd@MOFs, which is the
[22] I. Favier, D. Pla, M. Gomez, Chem. Rev. 120 (2020) 1146–1183.
main reason for the superior catalytic performance of Pd@MIL-125(Ti). [23] J. Yang, Y. Zhu, M. Fan, X. Sun, W.D. Wang, Z.J. Dong, J. Colloid Interface Sci. 554
In addition, we believe that the electronic interactions between Pd NPs (2019) 157–165.
and MOFs elucidated by this work may open a new access for the ex- [24] F. Yang, J. Tang, R. Ou, Z. Guo, S. Gao, Y. Wang, X. Wang, L. Chen, A. Yuan,
Applied Catal. B: Environmental 256 (2019) 117786.
ploration of metal-loaded MOFs (M@MOFs) for catalysis. [25] F. Yang, J. Wang, S. Gao, S. Zhou, Y. Kong, Mole. Catal. 486 (2020) 110873.
[26] F. Yang, S. Ding, H. Song, N. Yan, Sci. China Mater. 63 (2020) 982–992.
CRediT authorship contribution statement [27] S. Yang, L. Peng, S. Bulut, W.L. Queen, Chem. Eur. J. 25 (2019) 2161–2178.
[28] H. Liu, L. Chang, C. Bai, L. Chen, R. Luque, Y. Li, Angew. Chem. 128 (2016)
5103–5107.
Jiaxian Xu: Investigation, Data curation, Methodology, Writing - [29] X. Wang, M. Li, C. Cao, C. Liu, J. Liu, Y. Zhu, S. Zhang, W. Song, ChemCatChem 8
original draft. Fei Chen: Funding acquisition, Resources, Software, (2016) 3224–3228.
[30] L. Chen, X. Chen, H. Liu, Y. Li, Small 11 (2015) 2642–2648.
Writing - review & editing. Xuran Xu: Writing - review & editing, [31] L. Chen, W. Huang, X. Wang, Z. Chen, X. Yang, R. Luque, Y. Li, Chem. Commun. 53
Formal analysis. Guo-Ping Lu: Conceptualization, Funding acquisition, (2017) 1184–1187.
Investigation, Methodology, Supervision, Writing - review & editing. [32] L. Chen, X. Chen, H. Liu, C. Bai, Y. Li, J. Mater. Chem. A 3 (2015) 15259–15264.

9
J. Xu, et al. Molecular Catalysis 495 (2020) 111157

[33] G.-P. Lu, X. Li, L. Zhong, S. Li, F. Chen, Green. Chem. 21 (2019) 5386–5393. (2019) 172–186.
[34] J.-w. Zhang, G.-p. Lu, C. Cai, Green. Chem. 19 (2017) 4538–4543. [52] R.D. Caldwell, H. Qiu, B.C. Askew, A.T. Bender, N. Brugger, M. Camps,
[35] J.-W. Zhang, D.-D. Li, G.-P. Lu, T. Deng, C. Cai, ChemCatChem 10 (2018) M. Dhanabal, V. Dutt, T. Eichhorn, A.S. Gardberg, A. Goutopoulos, R. Grenningloh,
4258–4263. J. Head, B. Healey, B.L. Hodous, B.R. Huck, T.L. Johnson, C. Jones, R.C. Jones,
[36] J.-W. Zhang, D.-D. Li, G.-P. Lu, T. Deng, C. Cai, ChemCatChem 10 (2018) I. Mochalkin, F. Morandi, N. Nguyen, M. Meyring, J.R. Potnick, D.C. Santos,
4966–4972. R. Schmidt, B. Sherer, A. Shutes, K. Urbahns, A.C. Follis, A.A. Wegener,
[37] G. Kresse, J. Furthmuller, Phys. Rev. B 54 (1996) 11169–11186. S.C. Zimmerli, L. Liu-Bujalski, J. Med. Chem. 62 (2019) 7643–7655.
[38] P.E. Blochl, Phys. Rev. B 50 (1994) 17953–17979. [53] M. Toolabi, S. Moghimi, T.O. Bakhshaiesh, S. Salarinejad, A. Aghcheli,
[39] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 3865–3868. Z. Hasanvand, E. Nazeri, A. Khalaj, R. Esmaeili, A. Foroumadi, Eur. J. Med. Chem.
[40] W. Tang, E. Sanville, G.A. Henkelman, J. Phys.: Condens. Matter. 21 (2009) 185 (2020) 111786.
084204. [54] F. Liu, H. Wang, S. Li, G.A.L. Bare, X. Chen, C. Wang, J.E. Moses, P. Wu,
[41] Z. Liu, Y. Wu, J. Chen, Y. Li, J. Zhao, K. Gao, P. Na, Catal. Sci. Techol. 8 (2018) K.B. Sharpless, Angew. Chem. Int. Ed. 58 (2019) 8029–8033.
1936–1944. [55] C.J. Ribeiro, M. Espadinha, M. Machado, J. Gut, L.M. Goncalves, P.J. Rosenthal,
[42] V.R. Bakuru, S.B. Kalidindi, Chem. Eur. J. 23 (2017) 16456–16459. M. Prudencio, R. Moreira, M.M. Santos, Bioorg. Med. Chem. 24 (2016) 1786–1792.
[43] Z.X. Zhang, Y. Xiao, M.F. Cui, J.H. Tang, Z.Y. Fei, Q. Liu, X. Chen, X. Qiao, Dalton [56] K. Stenzel, A. Hamacher, F.K. Hansen, C.G.W. Gertzen, J. Senger, V. Marquardt,
Trans. 48 (2019) 14971–14974. L. Marek, M. Marek, C. Romier, M. Remke, M. Jung, H. Gohlke, M.U. Kassack,
[44] L. Zhang, L.L. Wang, L. Gong le, X.F. Feng, M.B. Luo, F. Luo, J. Hazard. Mater. 311 T. Kurz, J. Med. Chem. 60 (2017) 5334–5348.
(2016) 30–36. [57] S. Bai, L. Bu, Q. Shao, X. Zhu, X. Huang, J. Am. Chem. Soc. 140 (2018) 8384–8387.
[45] T.B. Yang, L.X. Sun, F. Xu, Z.Q. Wang, Y.J. Zou, H.L. Chu, Angew. Chem. Int. Ed. [58] Y. Dai, X. Gao, X. Chu, C. Jiang, Y. Yao, Z. Guo, C. Zhou, C. Wang, H. Wang,
727 (2017) 683–687. Y. Yang, J. Catal. 364 (2018) 192–203.
[46] D. Chen, S. Chen, Y. Jiang, S. Xie, H. Quan, L. Hua, X. Luo, L. Guo, RSC Adv. 7 [59] P. Wang, Q. Shao, X. Cui, X. Zhu, X. Huang, Adv. Func. Mater. 28 (2018) 1705918.
(2017) 49024–49030. [60] J. Mendes-Burak, B. Ghaffari, C. Coperet, Chem. Commun. 55 (2019) 179–181.
[47] Y. Lin, G.-P. Lu, X. Zhao, X. Cao, L. Yang, B. Zhou, Q. Zhong, Z. Chen, Mole. Catal. [61] S. Bewana, M.J. Ndolomingo, E. Carleschi, B.P. Doyle, R. Meijboom, N. Bingwa,
482 (2020) 110695. ACS Appl. Mater. Interfaces. 11 (2019) 32994–33005.
[48] Y.X. Liu, H.H. Wang, T.J. Zhao, B. Zhang, H. Su, Z.H. Xue, X.H. Li, J.S. Chen, [62] Z.H. Wang, F. Zaera, ACS Catal. 8 (2018) 8513–8524.
Journal of the American Chemical Society 141 (1) (2019) 38–41. [63] R. Li, W. Yao, Y. Jin, W. Jia, X. Chen, J. Chen, J. Zheng, Y. Hu, D. Han, J. Zhao,
[49] S. Smolders, T. Willhammar, A. Krajnc, K. Sentosun, M.T. Wharmby, Chem. Eng. J. 351 (2018) 995–1005.
K.A. Lomachenko, S. Bals, G. Mali, M.B.J. Roeffaers, D.E. De Vos, B. Bueken, [64] X.-W. Zhu, X.-P. Zhou, D. Li, Chem. Commun. 52 (2016) 6513–6516.
Angew. Chem. 58 (2019) 9160–9165. [65] K. Kanamura, S. Shiraishi, H. Takezawa, Z.-i. Takehara, Chem. Mater. 9 (1997)
[50] C. Ge, G. Fang, X. Shen, Y. Chong, W.G. Wamer, X. Gao, Z. Chai, C. Chen, J.-J. Yin, 1797–1804.
ACS Nano 10 (2016) 10436–10445. [66] B.H. Lipshutz, N.A. Isley, J.C. Fennewald, E.D. Slack, Angew. Chem. Int. Ed. 52
[51] T. Hao, Y. Li, S. Fan, W. Li, S. Wang, S. Li, R. Cao, W. Zhong, Eur. J. Med. Chem. 175 (2013) 10952–10958.

10

You might also like