Вы находитесь на странице: 1из 10

Journal of

Materials Chemistry A
View Article Online
PAPER View Journal | View Issue

Systematic variation of the optical bandgap in


Cite this: J. Mater. Chem. A, 2017, 5,
titanium based isoreticular metalorganic
Published on 07 April 2017. Downloaded by University Novi Sad on 24/09/2017 16:26:30.

11854 frameworks for photocatalytic reduction of CO2


under blue light
Matthew W. Logan,a Suliman Ayad,b Jeremy D. Adamson,a Tristan Dilbeck,b
Kenneth Hanson b and Fernando J. Uribe-Romo *a

A series of metalorganic frameworks isoreticular to MIL-125-NH2 were prepared, where the 2-amino-
terephthalate organic links feature N-alkyl groups of increasing chain length (from methyl to heptyl) and
varying connectivity (primary and secondary). The prepared materials display reduced optical bandgaps
correlated with the inductive donor ability of the alkyl substituent as well as high photocatalytic activity
towards the reduction of carbon dioxide under blue illumination operating over 120 h. Secondary N-alkyl
substitution (isopropyl, cyclopentyl and cyclohexyl) exhibits larger apparent quantum yields than the primary
N-alkyl analogs directly related to their longer lived excited-state lifetime. In particular, MIL-125-NHCyp
Received 13th January 2017
Accepted 7th April 2017
(Cyp cyclopentyl) exhibits a small bandgap (Eg 2.30 eV), a long-lived excited-state (s 68.8 ns) and
a larger apparent quantum yield (Fapp 1.80%) compared to the parent MIL-125-NH2 (Eg 2.56 eV, Fapp
DOI: 10.1039/c7ta00437k
0.31%, s 12.8 ns), making it a promising candidate for the next generation of photocatalysts for solar fuel
rsc.li/materials-a production based on earth-abundant elements.

limiting their applicability on environmentally relevant scales.


Introduction Thus, materials that are able to reduce CO2 with the use of
The accumulation of carbon dioxide in the atmosphere as visible light are an imminent and currently unfullled need.
a result of human activity is one of the largest contributing Metalorganic frameworks (MOFs)5 are a class of porous
factors in the gradual warming of the Earth.1 Our planet has materials composed of metal-oxide clusters connected through
a natural system of negative feedback processes to balance organic links forming highly crystalline open structures that oer
these accumulations, primarily by carbon xation through a promising alternative to metal-oxide semiconductors for pho-
photosynthesis. The amount of carbon dioxide already in the tocatalysis. Research in the design and synthesis of new MOFs
atmosphere is far too large for natural photosynthesis to reduce (reticular chemistry6) has allowed for the creation of functional
its accumulation in the short term. Therefore, it is urgent that materials with unprecedented high porosity (surface areas up to
technologies be developed that are capable of xing large 10 000 m2 g1)7 and exceptional chemical and thermal stability,
quantities of CO2 into organic matter, which are both energet- with applications in gas storage,8 gas separation,9 heterogeneous
ically ecient and capable of creating a usable feedstock of catalysis,10 payload release,11 and sensing,12 among others. MOFs
reduced carbon (so-called solar fuel).2 These new technologies are materials with extraordinary chemical and functional versa-
require the use of a naturally abundant and sustainable source tility, because both the organic and inorganic components can be
of energy. Early transition metal oxides, in particular titanium systematically varied towards targeted applications in a system-
oxides, are capable of reducing CO2 when excited by UV light.3 atic manner.13
Using sunlight to reduce CO2 directly would be ideal; however, Titanium based MOFs, in particular, combine the photo-
the majority of the sun's rays are in the visible-IR range,4 which catalytic activity of titanium oxide clusters with the light
is too low energy to induce photoreduction with these materials, absorption properties of organic linkers, producing materials
that can be photocatalytically active under UV-visible light.14
a
These catalytic reactions are facilitated through either oxidati-
Department of Chemistry, University of Central Florida, 4111 Libra Dr. Rm. 251 PSB,
Orlando, FL 32816-2366, USA. E-mail: fernando@ucf.edu
ve,14a,15 or reductive16 photoinduced electron transfer within the
b
Department of Chemistry, Florida State University, 95 Chiean Way Rm. 118 DLC, Ti-MOF under irradiation. MIL-125-NH2 (MIL Materials Insti-
Tallahassee, FL 32306-4390, USA tut Lavoisier) is a titanium-based MOF constructed from 2-ami-
Electronic supplementary information (ESI) available. CCDC 15273031527308. noterephthalate and a Ti8O12 ring-shaped cluster (Fig. 1) that has
For ESI and crystallographic data in CIF or other electronic format see DOI: shown activity towards photocatalytic reduction of carbon
10.1039/c7ta00437k

11854 | J. Mater. Chem. A, 2017, 5, 1185411863 This journal is The Royal Society of Chemistry 2017
View Article Online

Paper Journal of Materials Chemistry A

incorporating 2,5-diaminoterephthalate and 2-amino-


terephthalate into the MIL-125 framework eectively decreases
the optical bandgap. Other eorts include post-synthetic modi-
cation of MIL-125-NH2 via diazotization of the amine followed
by addition of an electron-rich aniline forming diazo-based
groups on the surface of the MOF particles, which exhibit
enhanced photooxidation of benzyl alcohols.15b
Here we report the synthesis of a family of titanium MOFs
isoreticular to MIL-125-NH2, where the amine functionality is
decorated with alkyl chains of varying length and connectivity,
i.e., methyl, ethyl, isopropyl, n-butyl, cyclopentyl, cyclohexyl, and
Published on 07 April 2017. Downloaded by University Novi Sad on 24/09/2017 16:26:30.

n-heptyl (Fig. 1). Our rationale for this functionalization strategy


relies on the hypothesis that sequential addition of carbon atoms
to the N-alkyl amino group will result in increased electron
density in the terephthalate ring as a consequence of subtle
inductive eects. These are well-known electronic eects that
result from the polarization of a chemical species through s-
bonds due to dierences in electronegativity between a hetero-
atom (in this case nitrogen) and the carbon atoms in the alkyl
chains. Moreover, these eects should be more evident in the
secondary and tertiary alkyls.20 Thus, increasing the number of s-
bonds in the chain should result in larger electron density
around the nitrogen atom in 2-aminoterephthalate, causing an
increased density in the aromatic ring, and therefore destabiliz-
ing the top of the valence band. By successively destabilizing the
valence band in small steps, the bandgap of the MOFs should
decrease, and higher photocatalytic activity can be achieved
under visible light irradiation. We studied the eect of the N-alkyl
substitution on the CO2 adsorption, as well as the photophysical
and photocatalytic properties of the materials, particularly the
absorption onset, and excited state lifetimes. These parameters
were also studied to understand the photocatalytic activity
Fig. 1 Top: Crystal structure of the parent MOF MIL-125 utilized in this towards reduction of CO2 utilizing visible light only, in the form
study. Bottom: Molecular structure of the organic link in MIL-125-NH2 of blue LEDs.
and the N-substituted isoreticular MOFs prepared in this work. Gray
spheres carbon, white hydrogen, red oxygen, blue polyhedral
titanium, and the yellow sphere represents the size of the pore in the Experimental
MOF.
All starting materials and solvents, unless otherwise specied,
were obtained from commercial sources (Aldrich, Fisher, VWR)
dioxide.17 The biggest challenge in increasing and optimizing and used without further purication. All reactions were per-
this activity for operating in the visible range relies on tuning the formed under ambient laboratory conditions, and no precau-
bandgap (or HOMOLUMO gap) of this material (Eg 2.56 eV). tions were taken to exclude oxygen or atmospheric moisture
de Miguel et al. established that photoirradiation of MIL-125- unless otherwise specied. Anhydrous N,N-dimethylformamide
NH2 with UV-visible light produces a long-lived charge-separated (DMF) and tetrahydrofuran (THF) were puried using a custom-
state whose lifetime is directly related to photocatalytic activity.18 built alumina-column based solvent purication system (Inno-
Moreover, it has been demonstrated through high-level DFT vative Technology). Anhydrous methanol was obtained from
calculations that this charge-separated excited state is accessible Aldrich (Sureseal). Mesitylene and hexamethylphosphoramide
due to the dierence in the electronic distribution of the frontier (HMPA) were dried over activated 4 A molecular sieves.
bands/orbitals.19 Hendon et al. propose that in this MOF, the top Deuterated solvents (CDCl3, DMSO, D2O, CD3CN, and NaOD
of the valence bandor HOMOis localized in the aromatic ring 40% in D2O) were obtained from Cambridge Isotope Lab.
of the 2-aminoterephthalate, and the bottom of the conduction
bandor LUMOis localized in the d-states of the titanium ions
favoring charge separation upon photoexcitation.19a They also Instrumentation
propose that a strategy to decrease the bandgap of this MOF (and High-resolution 1H and 13C nuclear magnetic resonance (NMR)
thus absorbing more visible light) is by destabilizing the valence spectra were collected using a Bruker AVANCE-III 400 MHz
band by increasing the electron density in the p-cloud of the spectrometer. The chemical shis are reported relative to the
terephthalate. It has been previously demonstrated that solvent residual signal. NMR data analysis was performed using

This journal is The Royal Society of Chemistry 2017 J. Mater. Chem. A, 2017, 5, 1185411863 | 11855
View Article Online

Journal of Materials Chemistry A Paper

MestReNova package (v. 10.0.2). Liquid chromatography-mass sphere was passed through a single grating (1800 l mm1, 500
spectra (LC-MS) were recorded using an Agilent 6230 TOF nm blaze) CzernyTurner monochromator and nally detected
coupled with an Agilent Zorbax SB-C18 analytical column. by a Peltier-cooled Hamamatsu R928 photomultiplier tube.
Fourier-transform infrared spectra were recorded using a Per- Synchronous spectral scans were performed with both excita-
kin Elmer Spectrum ONE Universal FT-IR ATR. 32 scans were tion and emission monochromators, with a zero wavelength
collected for each sample at a resolution of 120 000 cm1 from oset, stepping through the pre-set spectral range. Absorbance
4000650 cm1. Scanning electron microscopy (SEM) was con- was then calculated using Edinburgh's F900 soware package.
ducted using a JEOL JSM 6480 microscope under low-vacuum Tauc plots were obtained assuming a(hn) (Ahn)2.
with an accelerating voltage of 20 kV. Sample preparation
involved the dispersion of samples upon carbon tape attached Transient absorption (TA) measurements
to a stainless steel sample holder.
Published on 07 April 2017. Downloaded by University Novi Sad on 24/09/2017 16:26:30.

The samples for TA were prepared by suspending MIL-125-NHR


in 3.0 mL of MeCN in a 1 cm  1 cm quartz cuvette (A z 0.2
Powder X-ray diraction O.D.). Measurements were performed on a spectrometer
Powder X-ray diraction (PXRD) data were collected using composed of a Continuum Surelite EX Nd:YAG laser combined
a Rigaku MiniFlex 600 q2q diractometer in BraggBrentano with a Continuum Horizon OPO (lex 405 nm, 57 ns, operated
geometry with a 300 mm goniometer diameter, Ni-ltered CuKa at 1 Hz, beam diameter 0.5 cm, 5 mJ per pulse) integrated into
radiation (l 1.5418 A) at 600 W power (40 kV, 15 mA), a commercially available Edinburgh LP980 laser ash photol-
equipped with a NaI(Tl) SC-70 scintillation detector, 5.0 inci- ysis spectrometer. White light probe pulses generated by
dent and receiving Soller slits, a 0.625 divergent slit, a 1.25 a pulsed 150 W Xe lamp were passed through the sample,
scattering slit, and a 0.3 mm receiving slit and a Ni-CuKb lter. focused into the monochromator, and then, for full spectrum
Samples were analyzed from 3 to 40 2q-degrees with 0.02 per data, were detected using an intensied Andor iStar CCD
step and a scan rate of 0.25 2q-degrees min1. The samples were camera. Single wavelength kinetic absorption at 500 nm was
prepared by dropping the powder sample into a glass sample detected by a photomultiplier tube with a 435 nm long pass
holder and pressing the powder with a razor blade spatula lter placed before the detector to reject unwanted scattered
forming a smooth surface. Crystal models were created using light. Detector outputs were processed using a Tektronix
Materials Studio modeling suite21 starting from the published TDS3012C Digital Phosphor Oscilloscope interfaced to a PC.
crystal of MIL-125 (CCDC code RUPRUQ)14a by adding the cor- Single wavelength kinetic data were the result of averaging 50
responding N-alkyl chains. Rietveld renements were per- laser shots. Detector outputs were processed using the Edin-
formed with the modeled crystals and the experimental burgh's L900 (version 8.2.3, build 0) soware package. Time-
patterns using GSAS-II (see ESI for more details).22 resolved absorption data were t with the exponential function
in eqn (1) where s is the lifetime.
Gas adsorption isotherm analysis
I Ioet/s (1)
Gas adsorption isotherm analysis was performed using
a Micromeritics ASAP 2020 surface area and porosimetry
analyzer. Measurements were performed at 77 K (liquid N2 Synthesis of N-alkyl-2-amino-terephthalate dimethyl ester
bath) for N2 (g), and at 273, 283 and 298 K (water circulator bath) (1ad, 1g). 2-Amino-terephthalate dimethyl ester (2.00 g, 9.56
for CO2 (g) on thermally activated samples. BrunauerEmmett mmol) was suspended in anhydrous DMF (20 mL) and stirred
Teller (BET) surface areas were obtained by performing a Rou- until fully dissolved (HMPA was used instead of DMF for 1g).
querol analysis over the linear isotherm to determine the upper K2CO3 (5.30 g, 38.2 mmol) was added, followed by dropwise
limits of the BET model from the N2 isotherms. A linear least addition of alkyl iodide (14.3 mmol), and the mixture was stir-
squares tting over the BET plot provided the parameter for the red for 18 h at 100  C. The reaction mixture was then cooled to
volume of the monolayer, BET surface area and C-constant were room temperature and quenched with 2 M HCl (aq.) to pH 3.
obtained following the recommendation by Snurr et al. for The mixture was extracted with EtOAc (3  50 mL), and the
determination of the surface area in MOFs.23 Dierential combined organic extracts were washed with brine (3  50 mL),
enthalpies of CO2 adsorption were obtained by tting the dried over anhydrous MgSO4, ltered through celite and the
isotherms at the three measured temperatures into a virial solvent was removed using a rotary evaporator. The obtained
equation using the Micromeritics data processing soware. crude was puried using ash chromatography (SiO2, 15% v/v
EtOAc : hexanes, dry loading).
Solid-state absorption spectra Synthesis of N-cycloalkyl-2-amino-terephthalate dimethyl
Solid-state absorption spectra were collected at room tempera- ester (1e, 1f). Under N2 (g) in a 25 mL two-neck round bottom
ture using an Edinburgh FLS980 spectrometer with an inte- ask with a magnetic stir bar, 2-aminoterephthalate dimethyl
grating sphere accessory. The light output from a housed 450 W ester (1.00 g, 4.78 mmol) was suspended in anhydrous DMF (3.2
Xe lamp was passed through a single grating (1800 l mm1, 250 mL) and stirred until fully dissolved. TMSCl (1.30 g, 11.95
nm blaze) CzernyTurner monochromator and then into the mmol) and the corresponding cyclic ketone (5.25 mmol) were
integrating sphere containing the powder MIL-125-NHR sample added to the solution and the mixture was cooled to 0  C in an
or scattering reference (BaSO4). The output from the integrating ice bath. NaBH4 (0.18 g, 4.78 mmol) was added cautiously over 5

11856 | J. Mater. Chem. A, 2017, 5, 1185411863 This journal is The Royal Society of Chemistry 2017
View Article Online

Paper Journal of Materials Chemistry A

min. The mixture was then allowed to warm up to room mounted on top of a magnetic stirrer. The temperature inside
temperature and monitored by TLC until the disappearance of the photoreactor was kept at 25  C with the aid of an attached
the starting material. The mixture was quenched with saturated fan and a thermometer. The photon ux of the photoreactor was
NaHCO3 (aq.) followed by extraction with EtOAc (3  10 mL). determined using iron oxalate actinometry at 466 nm following
The combined organic fractions were then washed with brine (3 the protocol by Parker et al.24 obtaining a volumetric photon ux
 10 mL) and dried over MgSO4, ltered through celite and the of Ro,l 1.92  103 mol (photons) L1 h1. The apparent
solvent was removed using a rotary evaporator. The obtained quantum yield, Fapp (also named photonic eciency), was
crude was puried using ash chromatography (SiO2, 510% v/v determined according to the IUPAC recommended protocol by
EtOAc : hexanes, dry loading). Serpone et al.25 using the formula:
Synthesis of N-alkyl-2-amino-terephthalic acid (2af). The
Rin
alkylated diester 1af (0.50 g) was dissolved in THF (25 mL), Fapp (2)
Published on 07 April 2017. Downloaded by University Novi Sad on 24/09/2017 16:26:30.

Ro;l
followed by addition of 1 M NaOH (aq., 12.8 mL). The solution
was heated to 70  C and stirred for 8 h. The mixture was where Rin corresponds to the initial rate of the reaction (see
concentrated in a rotary evaporator at 45  C to remove the below) and Ro,l is the volumetric photon ux.
excess THF. The mixture was cooled to room temperature fol-
lowed by addition of 1 M HCl (aq.) until pH 3. The observed
Photochemical reduction of carbon dioxide to formate
precipitate was isolated by ltration, rinsed with water, and
dried in air at room temperature for 6 h. Photochemical reduction of carbon dioxide to formate was
Synthesis of MIL-125-NHR. Compounds 2af (0.41 mmol) performed under ambient conditions in the photoreactor
were placed in 20 mL scintillation vials with anhydrous DMF described above. MIL-125-NHR (35 mg) was loaded into a 3
(4.90 mL) and anhydrous MeOH (0.35 mL). The mixture was mL borosilicate vial with a magnetic stirrer. 1.5 mL of a CO2-
mixed thoroughly by immersing the vial in an ultra-sonic bath bubbled stock solution containing triethanolamine (TEOA,
for 1 min and then transferred into a borosilicate glass tube (see 0.30 mol L1) and mesitylene (1.5 mmol L1) in MeCN-d3 was
ESI for more details) and bubbled with N2 (g) for 5 min. added to the vial and tightly capped with a PTFE/rubber
Ti(OiPr)4 (0.031 mL, 0.106 mmol) was added to the tube via septum. A ow of CO2 (g) was bubbled into the vial with the aid
syringe, immediately ash frozen in liquid N2, and ame sealed of a needle with an input positive pressure of 1 psi for 30 min
as described in the ESI. The sealed tube was heated to 150  C previous to photoirradiation. During irradiation, 40 mL
for 48 h. Aer cooling to room temperature, the formed powder aliquots were sampled every 12 h with the aid of a syringe, the
was isolated by ltration and rinsed with DMF (3) and CHCl3 aliquot was ltered through a 0.2 mm PTFE syringe membrane
(3). The powder was immersed in CHCl3 and stored for 3 d in into an NMR tube. The ltrate was diluted with CDCl3 (0.6
a desiccator, replacing the solvent during this time (8). The mL) and a total of 256 transients were measured for each time
solvent was removed by decantation and the solvent-wet powder point. Kinetic plots were obtained in both concentration and
was dried under dynamic vacuum (10 mtorr) for 48 h at room turnover number in the ordinate axis. The concentration was
temperature. The degassed yellow solid was stored under N2 in obtained by addition of the intensity of the photoreduced-CO2
a desiccator. signals (formate + HEF + BHEF) and normalizing with respect
to the internal standard, and the turnover number was deter-
mined as the total number of moles of photoreduced-CO2 (mol
Blue LED photochemical reactor setup and calibration CO2 + mol HEF + mol BHEF) per mol of the photocatalyst (with
Photocatalytic reactions were carried out in a home-built reactor respect to the organic linker). The initial rate of the reaction
using a 60 W blue LED strip (5 m long, MEILI, product # CNMX- was obtained by tting a third order polynomial function
Hardlines-464321) with l 466 nm (fwhm 20 nm), the strip using a linear least squares tting and then evaluating its rst
was coiled and glued inside a 20 cm diameter tin can and derivative at t 0.

Scheme 1 Synthesis of isoreticular MIL-125-NHR MOFs. Conditions: (i-a) (R Me, Et, iPr, Bu, hep) RI, K2CO3, DMF/HPMA, 100  C, 18 h. (i-b) (R
Cyp, Cy) cycloketone, TMSCl, NaBH4, DMF, N2 (g), 0  C, 15 min. (ii) NaOH, THF, 70  C, 8 h. (iii) Ti(OiPr)4, DMF, MeOH, 150  C, 48 h.

This journal is The Royal Society of Chemistry 2017 J. Mater. Chem. A, 2017, 5, 1185411863 | 11857
View Article Online

Journal of Materials Chemistry A Paper

Results and discussion


The isoreticular family of MOFs was prepared via a SN2 reaction
of 2-amino terephthalate dimethyl ester (Scheme 1) with the
corresponding alkyl iodide under basic conditions forming
intermediates 1ad and 1g in moderate yields. Intermediates
1ef were synthesized via reductive amination with cyclo-
pentanone or cyclohexanone in the presence of NaBH4/TMSCl
in cold DMF in high yields. Hydrolysis of the dimethyl esters in
NaOH/THF aorded linkers 2ag, and solvothermal MOF crys-
tallization with Ti(OiPr)4 in DMF/MeOH in ame sealed boro-
Published on 07 April 2017. Downloaded by University Novi Sad on 24/09/2017 16:26:30.

silicate tubes produced the desired MOFs in high yields (see


ESI). The molecular integrity of the linkers within the MOFs
was determined using 1H NMR by dissolving the synthesized
MOFs in 0.1 M NaOD in D2O (Fig. S1). These measurements
allowed identifying any loss of alkyl chains at the N-substitu-
tion, as it has been observed by others26 that solvothermal
conditions using tetravalent early transition metal ions (such as
Zr+4) in DMF can induce dealkylations at the N-substitution. No
loss of N-alkyl chains was observed in any of the prepared
MOFs, indicating a robust and clean crystallization. Attempts to
prepare a double methylated aminoterephthalate MIL-125-
NMe2 using linker 3 were unsuccessful (Scheme 2) due to the
decomposition of the linker into the monomethylated species Fig. 2 Powder X-ray diraction patterns of the isoreticular family of
2a, with 80% conversion while forming non-porous amorphous MIL-125-NHR compared to the simulated pattern of the parent MIL-
solids. We deduced that the steric constrain imposed by the 125-NH2 (Rietveld renement parameters in the ESI).
double substitution impedes the crystallization of the MOF,
while Ti+4 induces N-dealkylation, similar to Zr-based MOFs,
and thus only single N-alkylated MOFs were successfully the library of suitable models for crystal structure solution with
synthesized. the Rietveld method. Renements of the experimental patterns
The MOFs were obtained as microcrystalline powders that vs. the theoretical crystals allowed the solution of the crystal
were analyzed by scanning electron microscopy (SEM) and structures in the high-symmetry tetragonal I4/mmm space group
powder X-ray diraction (PXRD) crystallography. SEM images with low residuals (see ESI), where the NR groups were
(Fig. S10S17) displayed samples with uniform particle shape assumed to be disordered around the four possible positions in
and texture, suggesting homogeneous materials. All the mate- the terephthalate ring and inside the pore.
rials display sharp diraction peaks (Fig. 2) starting at 6.8 2q, The porosity of the MOFs was determined using N2 gas
evidencing a high degree of crystallinity with large unit cells, as adsorption at 77 K. All isotherms display IUPAC Type II
expected in a MOF. All patterns were indexed by comparison to isotherm behavior that corresponds to microporous materials
the simulated pattern of MIL-125-NH2,17 suggesting the forma- with fully reversible adsorption behavior (Fig. S18S47).
tion of frameworks that are isoreticular to MIL-125. Crystal Application of the BrunauerEmmettTeller (BET) model in the
simulation using Materials Studio Modeling Suite21 produced low-pressure region of the isotherm resulted in BET surface
areas between 200 and 1000 m2 g1. A progressive decrease is
observed in the order of MIL-125-NH2 (SBET 1200 m2 g1) >
MIL-125-NHMe (SBET 1000 m2 g1) > MIL-125-NHEt (SBET
750 m2 g1) > MIL-125-NHBu (SBET 690 m2 g1) > MIL-125-
NHCyp (SBET 580 m2 g1) > MIL-125-NHiPr (SBET 490 m2
g1) > MIL-125-NHCy (SBET 420 m2 g1) > MIL-125-NHhep
(SBET 230 m2 g1). The total cumulative pore volume, ob-
tained from non-local density functional theory (NLDFT) tting
of the N2 isotherms, resulted in decreasing pore volumes
consistent with increased substitution (Fig. S48) decreasing
from 0.73 cm3 (STP) g1 in MIL-125-NH2 to 0.09 cm3 (STP) g1 in
MIL-125-NHhep. The CO2 adsorption isotherms at 273 (Fig. 3a),
283 and 298 K (Fig. S49S56) exhibit a similar trend in uptake
Scheme 2 Attempted synthesis of MIL-125-NMe2. Note: attempts to
where increased N-alkyl substitution results in lower uptakes
prepare this MOF under the same conditions as in mono N-substituted
MOFs resulted in N-demethylation of the linker with formation of from 221 cm3 g1 in MIL-125-NH2, to 19.1 cm3 g1 in MIL-125-
amorphous non-porous samples. NHhep. These trends are consistent with: (1) the increased

11858 | J. Mater. Chem. A, 2017, 5, 1185411863 This journal is The Royal Society of Chemistry 2017
View Article Online

Paper Journal of Materials Chemistry A


Published on 07 April 2017. Downloaded by University Novi Sad on 24/09/2017 16:26:30.

Fig. 3 (a) CO2 gas adsorption isotherms (273 K) and (b) dierential (or isosteric) enthalpy of adsorption vs. loading plots of MIL-125-NHR with R
H (brown), Me (purple), Et (blue), iPr (green), Bu (yellow), Cyp (orange), Cy (red), and hep (pink). CO2 adsorption points at high pressure in MIL-125-
NH2 have been measured elsewhere17 and are not shown for clarity.

amount of weight of the MOF by addition of extra atoms in the the alkylated derivatives; presumably the shi can be attributed
N-alkyl chain, and (2) the decreased accessibility to the pores as to the alkyl donor group destabilizing the top of the valence
a result of clogging of the pore and pore apertures as the alkyl band (HOMO) of the amino terephthalate linker. Application of
chains occupy the free space of the pores. We also observed that the Tauc model to the photoabsorption plots (Fig. S57S64)
although the functionalized MOFs display lower CO2 uptakes allowed for the determination of the optical bandgap (or
than the parent MIL-125-NH2 at room temperature (especially HOMOLUMO gap). A decrease in the bandgap of the materials
MIL-125-NHCyp, MIL-125-NHCy, and MIL-125-NHhep), their is observed (Fig. 4b) from MIL-125-NH2 (Eg 2.56 eV) to MIL-
adsorption capacity has little eect on their CO2 photoreduction 125-NHCy (Eg 2.29 eV) in a stepwise mode with increased alkyl
rates (see below), which only suggests that mass transport chain substitution, as expected from the inductive eects of the
within the pores is not the rate determining step in the corresponding alkyl chain. Of the series, the n-heptyl
photoreduction. Determination of the dierential enthalpy of substituted MOF displayed a larger bandgap than the cyclo-
adsorption Dh_ ads (Fig. 3b) allows understanding of the ther- pentyl and cyclohexyl, and is close in value to the n-butyl
modynamic interactions between CO2 and the MOFs. We substituted MOF. This behavior is due to the linear nature of the
observed no particular trend given by the systematic N-alkyl heptyl chain. Inductive eects decrease exponentially as the C
substitution across the uptake ranges, with all MOFs observing C s-bonds are further away from the heteroatom, expecting very
similar Dh_ ads to MIL-125-NH2, ranging from 20 kJ mol1 little induction beyond the fourth s-bond, as observed in MIL-
(MIL-125-NHEt) to 27 kJ mol1 (MIL-125-NHMe) when loading 125-NHBu and MIL-125-NHhep. In contrast, cyclic substituents
1 CO2 molecule per unit cell. This lack of trend in adsorption retain the s-bonds in close proximity to the amine group,
enthalpy evidences the convoluted eect of: (1) the inherent resulting in stronger electronic induction and therefore lower
porosity of the parent MIL-125 framework, (2) the electron rich bandgaps. We expect that tertiary carbon substitutions (e.g.,
nature of the amino nitrogen lone pair acting as a Lewis basic tert-butyl), or N,N-dialkyls (e.g. MIL-125-NMe2) would result in
site, and (3) the steric hindrance imposed by the bulky alkyl even lower bandgaps.
substituents. While the enthalpy of adsorption is a combination The photocatalytic rates and eciencies of MIL-125-NHR
of all these eects, the undiluted electronic eects are better MOFs are known to be dependent on the lifetime of the pho-
understood by the determination of the bandgap of the MOFs. togenerated charge-separated state.18 To gain insights into the
The UV-visible absorption spectra for crystalline powder of excited state dynamics, nanosecond transient absorption spec-
the MIL-125-NHR series were acquired using an integrating troscopy was performed on suspensions of selected MIL-125-
sphere and the results can be seen in Fig. 4a. The parent MIL- NHR in MeCN and the results can be seen in Fig. 4c and
125-NH2 exhibits a low energy absorption peak at 375 nm. As Fig. S66S69. Upon excitation at 405 nm the parent material,
expected, a gradual red shi in the absorption is observed for MIL-125-NH2, exhibits a bleach from 390 nm to 530 nm and

This journal is The Royal Society of Chemistry 2017 J. Mater. Chem. A, 2017, 5, 1185411863 | 11859
View Article Online

Journal of Materials Chemistry A Paper


Published on 07 April 2017. Downloaded by University Novi Sad on 24/09/2017 16:26:30.

Fig. 4 (a) Solid-state absorption spectra of MIL-125-NHR MOFs, (b) measured bandgap for MIL-125-NHR MOFs using the Tauc model (error
bands correspond to one standard deviation from the linear tting), (c) transient absorption spectra of MIL-125-NHMe, and (d) time-resolved
absorption traces at 500 nm for MIL-125-NHR suspended in MeCN (lex 405 nm).

positive DA features at l < 390 nm and l > 530 nm (Fig. S65). of longer lifetimes with greater donor ability of the alkyl substit-
These transient spectroscopic features are indicative of the uent. Presumably, the enhanced stabilization of the oxidized
charge-separated state with a hole on the aminoterephthalate terephthalate by the secondary alkyl chain is responsible for the
linker and the electron on the titanium-oxo cluster.18,27 Similarly, long lifetime observed for MIL-125-NHiPr, MIL-125-NHCyp and
absorption features are observed for the alkylated derivatives with MIL-125-NHCy.
example spectra of MIL-125-NHMe shown in Fig. 4c and all other We explored the photocatalytic activity of the isoreticular
spectra in the ESI. As was observed with the ground state MOFs towards CO2 photoreduction using blue irradiation
absorption, the alkylated complexes' DA features are red shied by (Scheme 3). A LED light source was utilized because of its
30 nm compared to MIL-125-NH2. The lifetime of the charge- accessibility, low cost, high irradiance, low power needs and
separated state for the measured set of MIL-125-NHR materials narrow emission prole, compared to Xe-based lamps.28 More-
was monitored at 500 nm (lex 405 nm) and the results can be over, using a narrow band aids in circumventing photocatalysis
seen in Fig. 4d. The lack of the signal rise time indicates that the by any residual titanium oxide formed during the MOF
charge-separated state is generated within the instrument synthesis, as these phases require light in the UV region (l < 350
response time (<10 ns). The charge-separated lifetime increases in nm).3 Irradiating MIL-125-NHR suspensions in MeCN-d3 under
the order of MIL-125-NH2 (s 12.8 ns) < MIL-125-NHMe (s 38.2 a ow of CO2 and in the presence of triethanolamine, TEOA (a
ns) < MIL-125-NHBu (s 52.1 ns) < MIL-125-NHEt (s 60.63 ns) < reducing agent) and mesitylene (an internal standard) with blue
MIL-125-NHCyp (s 68.9 ns) < MIL-125-NHCyp (s 68.8 ns) < light (l 466 nm) resulted in the formation of formate as the
MIL-125-NHhep (s 69.2 ns) < MIL-125-NHiPr (s 75.11 ns) < product of CO2 reduction, which displays a 1H NMR signal at
MIL-125-NHCy (s 91.35 ns) generally corresponding to a trend 8.45 ppm (Fig. 5). Two other signals were also observed at 8.08

11860 | J. Mater. Chem. A, 2017, 5, 1185411863 This journal is The Royal Society of Chemistry 2017
View Article Online

Paper Journal of Materials Chemistry A

Scheme 3 Photocatalytic reduction of carbon dioxide. Note: HEF and BHEF appear as formate reacts with the hydrolyzed products of oxidized
triethanolamine.29

measured the concentration of formate with our photoreactor


Published on 07 April 2017. Downloaded by University Novi Sad on 24/09/2017 16:26:30.

([formate] 0.165 mmol L1 at t 12 h) and compared it to the


observations of Li et al. ([formate] 0.136 mmol L1 at t 10
h),17 resulting in very similar values, providing a validation to
our approach. Fitting polynomial functions to the concentra-
tion-based kinetic plot, followed by evaluation at t 0 h of their
rst derivative, allowed us to obtain the initial rate of the
photoreduction. The initial rates for all the MOFs were utilized
to obtain an apparent quantum yield (Fapp, also referred to as
photonic eciency25) using eqn (2). In this work we use the term
apparent quantum yield for two reasons: rst, the ratio of light
absorption/scattering by the MOF is not accounted for (i.e. we
assume complete light absorption), and second, because the
kinetic plots account for two side reactions (amide formation)
that occur aer the photocatalytic step. As shown in Fig. 6b,
secondary N-alkyls, isopropyl, cyclopentyl, and cyclohexyl
(Fig. 6b blue histograms) display a larger Fapp (1.51.8%) than
the primary N-alkyls (0.300.40%, Fig. 6b orange histograms);
especially MIL-125-NHCyp, which displayed the largest
apparent quantum yield Fapp 1.80%. The increased catalytic
activity is consistent with the increased excited-state lifetime
and decreased bandgap for the secondary alkyl substituent
species.
It is noteworthy to emphasize that we observed no trend or
correlation between the photocatalytic rates/eciency and the
gas adsorption properties of the MOFs. We believe that elec-
Fig. 5 H NMR traces (400 MHz, CDCl3, 25  C) at varying times of the
1
tronic eects from N-alkyl substitution determine the photo-
photoreduction of CO2 under blue LED illumination using MIL-125-
NHCyp. Signals for formate, HEF, BHEF, residual solvent, and internal
catalytic eciency with little eect from mass transport. This
standard are indicated. assessment is consistent with the reaction mechanism
proposed by Fu et al.,17 who propose that upon photoexcitation,
an electron is promoted from the p-cloud in terephthalate to
a Ti d-state and is further transferred to CO2, reducing it to
and 8.04 ppm, which correspond to N-(2-hydroxyethyl)form- formate. The MOF catalyst is then regenerated by oxidizing
amide (HEF) and N,N-bis(2-hydroxyethyl)formamide (BHEF), TEOA (the latter also serving as a proton source). Our observed
respectively (Fig. S73). These two species appear as result of correlation between the excited-state lifetime and apparent
formate reacting with the hydrolyzed products of oxidized TEOA quantum yield suggests that the rate-limiting step of the
(Scheme 3), as the photoreduction operates under non-anhy- reduction is the electron transfer from Ti+3 to CO2 competing
drous conditions.29 Integration of the NMR signals and with charge recombination of the MOF. The chemical stability
normalization with respect to the mesitylene internal standard of the MIL-125-NHR MOFs was assessed by PXRD and 1H NMR
provided kinetic plots (Fig. S74S81) for the three CO2-photo- aer 120 h of photoreaction. No changes in the PXRD patterns
reduced species. Fig. 6a displays the kinetic plot of the total were observed, suggesting the stability of the crystalline
CO2-reduced products over 120 h utilizing MIL-125-NHMe and framework (Fig. S9). Moreover, only marginal linker decom-
MIL-125-NHhep. Similar to the trend in photoabsorption and position was observed in the 1H NMR spectra of the base-
the bandgap, the increased number of carbon atoms in the N- digested MOFs. The reusability of the MOF photocatalysts was
alkyl chain results in increased photocatalytic activity, with assessed by re-exposing MIL-125-NHCyp for a second photo-
turnover numbers at 120 h between 0.284 for MIL-125-NHMe catalytic run for another 120 h, observing a less than 5% vari-
and 1.151 for MIL-125-NHhep. For comparison purposes, we ation in the concentration of reduced CO2 species, indicating

This journal is The Royal Society of Chemistry 2017 J. Mater. Chem. A, 2017, 5, 1185411863 | 11861
View Article Online

Journal of Materials Chemistry A Paper


Published on 07 April 2017. Downloaded by University Novi Sad on 24/09/2017 16:26:30.

Fig. 6 (a) Kinetic plot of the CO2-photoreduced products (formate, HEF and BHEF) over time utilizing MIL-125-NHMe (purple) and MIL-125-
NHhep (pink). The turnover number determined as total moles of reduced CO2 per mole of catalyst (based on linker). (b) CO2 photoreduction
apparent quantum yield of the isoreticular MIL-125-NHR MOFs, the histogram color indicates the nature of the alkyl chain.

photochemical stability aer reuse (Fig. S83). Future work Acknowledgements


includes studying the eect of other reductants (e.g. alcohols,
hydrogen) and more complex substitutions, such as including Transient absorption measurements were performed on
tertiary N-alkyls and p-donors that will increase the lifetime of a spectrometer supported by the National Science Foundation
the excited state also induce smaller bandgaps. under Grant No. CHE-1531629. We thank Mr Hugh Hayes and
Prof. Gang Chen (UCF) for experimental assistance.

Conclusion
References
We prepared an isoreticular series of titanium-based MOF
photocatalysts by increasing the N-alkyl substitution in MIL- 1 R. Quadrelli and S. Peterson, Energy Policy, 2007, 35, 59385952.
125-NHR, where R varies from methyl, to ethyl, isopropyl, n- 2 (a) W. Tu, Y. Zhou and Z. Zou, Adv. Mater., 2014, 26, 4607
butyl, cyclopentyl, cyclohexyl, and n-heptyl. The isoreticular 4626; (b) C. Wang, R. L. Thompson, J. Baltrus and
nature of their crystal structure was studied by powder X-ray C. Matranga, J. Phys. Chem. Lett., 2010, 1, 4853; (c)
crystallography and gas adsorption. Their photophysical and T. W. Woolerton, S. Sheard, E. Reisner, E. Pierce,
photocatalytic properties were studied displaying a gradual S. W. Ragsdale and F. A. Armstrong, J. Am. Chem. Soc.,
decrease in the optical bandgap from 2.56 eV in MIL-125-NH2 to 2010, 132, 21322133; (d) P. Li, Y. Zhou, Z. Zhao, Q. Xu,
2.29 eV in MIL-125-NHCy, consistent with increased electron X. Wang, M. Xiao and Z. Zou, J. Am. Chem. Soc., 2015, 137,
density around the organic linker via inductive eects of the N- 95479550; (e) R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki
alkyl chain. We explored the photocatalytic eciency of the and Y. Taga, Science, 2001, 293, 269271.
MOFs towards the reduction of carbon dioxide under blue LED 3 S. N. Habisreutinger, L. Schmidt-Mende and J. K. Stolarczyk,
light, displaying increased reaction rates and quantum yields Angew. Chem., Int. Ed., 2013, 52, 73727408.
consistent with the increased alkyl substitution. We observed 4 D. M. Schultz and T. P. Yoon, Science, 2014, 343, 1239176.
that the CO2 photoreduction quantum yield is higher for MOFs 5 H. Furukawa, K. E. Cordova, M. O'Keee and O. M. Yaghi,
with increased excited-state lifetime, particularly with Science, 2013, 341, 1230444.
secondary N-alkyl substituents, and concluded that from the 6 O. M. Yaghi, M. O'Keee, N. W. Ockwig, H. K. Chae,
current proposed mechanism, the electron transfer from Ti+3 to M. Eddaoudi and J. Kim, Nature, 2003, 423, 705714.
CO2 is the rate-limiting step in the photoreduction. This study 7 H. Furukawa, N. Ko, Y. B. Go, N. Aratani, S. B. Choi, E. Choi,
allowed us to understand how small variations in the organic A. O. Yazaydin, R. Q. Snurr, M. O'Keee, J. Kim and
linker of the MOF can result in tuning of the photophysical and O. M. Yaghi, Science, 2010, 329, 424428.
photocatalytic properties of the materials, solely by selection of 8 L. J. Murray, M. Dinca and J. R. Long, Chem. Soc. Rev., 2009,
specic substituents in the organic links. 38, 12941314.

11862 | J. Mater. Chem. A, 2017, 5, 1185411863 This journal is The Royal Society of Chemistry 2017
View Article Online

Paper Journal of Materials Chemistry A

9 (a) K. Sumida, D. L. Rogow, J. A. Mason, T. M. McDonald, and L. Wu, Inorg. Chem., 2015, 54, 11911193; (c) H. Wang,
E. D. Bloch, Z. R. Herm, T.-H. Bae and J. R. Long, Chem. X. Yuan, Y. Wu, G. Zeng, X. Chen, L. Leng, Z. Wu, L. Jiang
Rev., 2012, 112, 724781; (b) J.-R. Li, R. J. Kuppler and and H. Li, J. Hazard. Mater., 2015, 286, 187194.
H.-C. Zhou, Chem. Soc. Rev., 2009, 38, 14771504. 17 Y. Fu, D. Sun, Y. Chen, R. Huang, Z. Ding, X. Fu and Z. Li,
10 (a) L. Zeng, X. Guo, C. He and C. Duan, ACS Catal., 2016, 6, Angew. Chem., Int. Ed., 2012, 51, 33643367.
79357947; (b) H. He, J. A. Perman, G. Zhu and S. Ma, Small, 18 M. de Miguel, F. Ragon, T. Devic, C. Serre, P. Horcajada and
2016, 12, 63096324. H. Garca, ChemPhysChem, 2012, 13, 36513654.
11 C.-Y. Sun, C. Qin, X.-L. Wang and Z.-M. Su, Expert Opin. Drug 19 (a) C. H. Hendon, D. Tiana, M. Fontecave, C. Sanchez,
Delivery, 2013, 10, 89101. L. D'arras, C. Sassoye, L. Rozes, C. Mellot-Draznieks and
12 L. E. Kreno, K. Leong, O. K. Farha, M. Allendorf, R. P. Van A. Walsh, J. Am. Chem. Soc., 2013, 135, 1094210945; (b)
Duyne and J. T. Hupp, Chem. Rev., 2012, 112, 11051125. C.-K. Lin, D. Zhao, W.-Y. Gao, Z. Yang, J. Ye, T. Xu, Q. Ge,
Published on 07 April 2017. Downloaded by University Novi Sad on 24/09/2017 16:26:30.

13 (a) O. M. Yaghi, J. Am. Chem. Soc., 2016, 138, 1550715509; S. Ma and D.-J. Liu, Inorg. Chem., 2012, 51, 90399044; (c)
(b) C. Wang, D. Liu and W. Lin, J. Am. Chem. Soc., 2013, P. Sippel, D. Denysenko, A. Loidl, P. Lukenheimer,
135, 1322213234. G. Sastre and D. Volkmer, Adv. Funct. Mater., 2014, 24,
14 (a) M. Dan-Hardi, C. Serre, T. Frot, L. Rozes, G. Maurin, 38853896; (d) H. Q. Pham, T. Mai and N.-N. Pham-Tran, J.
C. Sanchez and G. Ferey, J. Am. Chem. Soc., 2009, 131, Phys. Chem. C, 2014, 118, 45674577.
1085710859; (b) S. Yuan, T.-F. Liu, D. Feng, J. Tian, 20 E. V. Anslyn, and D. A. Dougherty, Modern Physical Organic
K. Wang, J. Qin, Q. Zhang, T.-P. Chen, M. Bosch, L. Zou, Chemistry, University Science Books, 2006.
S. Teat, S. Dalgarno and H.-C. Zhou, Chem. Sci., 2015, 6, 21 Materials Studio, version 8.0, BIOVIA Soware Inc., San
39263930; (c) B. Bueken, F. Vermoortele, Diego, CA, 2014.
D. E. P. Vanpoucke, H. Reinsch, C.-C. Tsou, P. Valvekens, 22 B. H. Toby and R. B. Von Dreele, J. Appl. Crystallogr., 2013,
T. De Baerdemaeker, R. Ameloot, C. E. A. Kirschhock, 46, 544549.
V. Van Speybroeck, J. M. Mayer and D. De Vos, Angew. 23 K. S. Walton and R. Q. Snurr, J. Am. Chem. Soc., 2007, 129,
Chem., Int. Ed., 2015, 54, 1391213917; (d) L. Zou, D. Feng, 85528556.
T.-F. Liu, Y.-P. Chen, S. Yuan, K. Wang, X. Wang, 24 C. Hatchard and C. A. Parker, Proc. R. Soc. London, Ser. A,
S. Fordham and H.-C. Zhou, Chem. Sci., 2016, 7, 1063 1956, 235, 518536.
1069; (e) H. Assi, L. C. Pardo Perez, G. Mouchaham, 25 N. Serpone and A. Salinaro, Pure Appl. Chem., 1999, 71, 303
F. Ragon, M. Nasalevich, N. Guillou, C. Martineau, 320.
H. Chevreau, F. Kapteijn, J. Gascon, P. Fertey, E. Elkaim, 26 H. Hahm, H. Ha, S. Kim, B. Jung, M. H. Park, Y. Kim, J. Heo
C. Serre and T. Devic, Inorg. Chem., 2016, 55, 71927199; (f) and M. Kim, CrystEngComm, 2015, 17, 56445650.
H. L. Nguyen, F. Gandara, H. Furukawa, T. L. H. Doan, 27 J. G. Santaclara, M. A. Nasalevich, S. Castellanos,
K. E. Cordova and O. M. Yaghi, J. Am. Chem. Soc., 2016, W. H. Evers, F. Spoor, K. Rock, L. D. Siebbeles, F. Kapteijn,
138, 43304333; (g) C. Serre, J. A. Groves, P. Lightfoot, F. Grozema and A. Houtepen, ChemSusChem, 2016, 9, 388
A. M. Z. Slawin, P. A. Wright, N. Stock, T. Bein, M. Haouas, 395.
F. Taulelle and G. Ferey, Chem. Mater., 2006, 18, 1451; (h) 28 S. Pimputkar, J. S. Speck, S. P. DenBaars and S. Nakamura,
J. Gao, J. Miao, P.-Z. Li, W. Y. Teng, L. Yang, Y. Zhao, Nat. Photonics, 2009, 3, 180182.
B. Liu and Q. Zhang, Chem. Commun., 2014, 50, 37863788. 29 (a) A. Tills, E. Blondiaux, X. Frogneux and T. Cantat, Green
15 (a) M. W. Logan, Y. A. Lau, Y. Zheng, E. A. Hall, Chem., 2015, 17, 157168 and references therein; (b)
M. A. Hettinger, R. P. Marks, M. L Hoster, F. M. Rossi, H. Hishida, H. Tanaka, K. Tanaka and T. Tanaka, Chem.
Y. Yuan and F. J. Uribe-Romo, Catal. Sci. Technol., 2016, 6, Lett., 1987, 597600; (c) K. Kobayashi, T. Kikuchi,
56475655; (b) M. A. Nasalevich, M. G. Goesten, S. Kitagawa and K. Tanaka, Angew. Chem., Int. Ed., 2014,
T. J. Savenije, F. Kapteijn and J. Gascon, Chem. Commun., 53, 1181311817; (d) L. Zhang, Z. Han, X. Zhao, Z. Wang
2013, 49, 1057510577; (c) D. Sun, L. Ye and Z. Li, Appl. and K. Ding, Angew. Chem., Int. Ed., 2015, 54, 61866189;
Catal., B, 2015, 164, 428432. (e) Y. Pellegrin and F. Odobel, C. R. Chim., 2017, 20, 283
16 (a) T. Toyao, M. Saito, Y. Horiuchi, K. Mochizuki, M. Iwata, 295; (f) U. B. Patil, A. S. Singh and J. M. Nagarkar, Chem.
H. Higashimura and M. Matsuoka, Catal. Sci. Technol., Lett., 2013, 42, 524526.
2013, 3, 20922097; (b) L. Shen, M. Luo, L. Huang, P. Feng

This journal is The Royal Society of Chemistry 2017 J. Mater. Chem. A, 2017, 5, 1185411863 | 11863

Вам также может понравиться