You are on page 1of 12

Article

pubs.acs.org/JPCC

Fluorescence of A100 MOF and Adsorption of Water, Indole, and


Naphthalene on A100 by the Spectroscopic, Kinetic, and DFT Studies
Jun Dai,† Michael L. McKee,‡ and Alexander Samokhvalov*,†

Department of Chemistry, Rutgers University, Camden, New Jersey 08102, United States

Department of Chemistry and Biochemistry, Auburn University, Auburn, Alabama 36849, United States
*
S Supporting Information

ABSTRACT: Metal−organic frameworks (MOFs) are promising materials for


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

adsorption and separations. It is important to understand the details of chemical bonding


between the adsorbate and structural units in the MOFs. In A100 MOF, the near-UV−
Downloaded via CENTRAL SOUTH UNIV on February 28, 2022 at 14:29:18 (UTC).

visible fluorescence is found to be the intralinker fluorescence. Naphthalene and indole


form the stoichiometric “host-guest” π−π adsorption complexes with A100 that contain
one adsorbate molecule per two BDC linkers, and adsorption of indole causes a strong
quenching of the intralinker fluorescence. The excitation wavelength dependent steady-
state fluorescence spectra, the nanosecond time-resolved fluorescence spectra, and DFT
calculations indicate the strong π−π interactions between adsorbed indole and
naphthalene and aromatic ring of the BDC linker, as well as hydrogen bonding between
adsorbed indole and COO group of the linker. Activated A100 adsorbs up to four water molecules per BDC linker. Kinetic study
of adsorption of naphthalene and indole from n-alkane on hydrated A100 yields the preferential adsorption of indole as
determined by the in-situ time-dependent fluorescence spectroscopy and complementary ex-situ UV−vis absorption
spectroscopy.

1. INTRODUCTION The high microporosity of MIL-53(Al) results in the Langmuir


Aromatic N-heterocyclic compounds are abundantly present in surface area up to 1600 m2/g. In the lp form, micropores are
petroleum1 and other fossils, in refinery streams,2 and in certain free of adsorbate and have the dimensions ca. 8.5 × 8.5 Å. The
commercial fuels.3 Upon combustion of nitrogen-containing lp form of MIL-53(Al) can adsorb water vapor at room
fuels, toxic nitrogen oxides are released that cause air pollution. temperature, yielding the narrow pore (np) form12 with
The adsorption of aromatic N-heterocyclic compounds from formula Al(OH)[O2C−C6H4−CO2] H2O. The hydrogen
liquid fuels was studied on zeolites, activated carbons,4 and bonds are formed between the H atoms of adsorbed water
novel sorbents such as ionic liquids.5 Metal−organic frame- molecule and the O atoms of the COO groups of the linker.
works (MOFs) have a very high surface area and pore volume The np MIL-53(Al) can reversibly lose adsorbed water upon
and are well suited for adsorption. Adsorption of many heating, and the lp form is formed. The reversible change
industrially important gases such as hydrogen,6 acetylene,7 between two forms of MIL-53 aka “breathing” has been utilized
and carbon dioxide8 on the MOFs has been studied quite for selective separations, e.g., the np MIL-53(Cr) features
extensively, and an excellent review has recently been published selective adsorption of CO2 in the presence of methane.13 MIL-
on the topic,9 while adsorption in the liquid phase was studied 53(Al) has a commercially available equivalent A100 Basolite
much less. Many MOFs have a moderate to high stability to MOF18 with specific surface area 1084 m2/g, total pore volume
elevated temperatures and oxygen, but the majority of the 0.51 cm3/g, and similar XRD pattern.19
MOFs undergo hydrolysis, e.g., ref 10. Many Al-containing Indole (IND) is the smallest molecule with fused
MOFs11 with linkers of aromatic acids are water stable heteroaromatic and aromatic rings; IND and alkyl-substituted
including microporous MIL-53(M), where M = Al,12 Cr,13 indoles are present in petroleum, tar sands and bitumen,
Fe,14 In,15 Sc.16 The 3D framework of MIL-53(Al) is built by streams of petroleum refineries,2 and commercial liquid fossil
the infinite 1D chains of repeating corner-sharing octahedral fuels. The functional group of indole is present in amino acid
AlO4(OH)2 units, and the chains are interconnected by the tryptophan and proteins containing tryptophan and in many
COO groups of organic linker, the anion of 1,4-benzenedi- biologically active natural and synthetic compounds: metabo-
carboxylic acid (BDC).17 MIL-53(Al) is synthesized hydro- lites, e.g., tryptamine, natural pigments, e.g., eumelanin,20
thermally, and in its as-synthesized form, rhombic-shaped medicinal drugs for treatment of many diseases including HIV21
micropores ca. 7.3 × 7.7 Å are filled with an excess of 1,4- and cancer,22 in food supplements, e.g., melatonin, etc.
benzenedicarboxylic acid (BDCH2) used in the synthesis.12
Upon thermal “activation”, MIL-53(Al) desorbs the BDCH2 Received: October 11, 2014
impurity and water and changes to its large-pore (lp) form aka Revised: January 9, 2015
the high-temperature (ht) form Al(OH)[O2C−C6H4−CO2]. Published: January 9, 2015

© 2015 American Chemical Society 2491 DOI: 10.1021/jp510272s


J. Phys. Chem. C 2015, 119, 2491−2502
The Journal of Physical Chemistry C Article

Naphthalene (NAP) is the smallest hydrocarbon with a fused- spectroscopic characterization of the host−guest adsorption
ring aromatic system, and fused-ring aromatic hydrocarbons are complexes of IND, NAP, and water with activated A100 using
the common chemical components of all fossils and liquid fossil the complementary wavelength-dependent and nanosecond
fuels. Adsorption of IND, pyridine, pyrrole, and quinoline was time-resolved fluorescence spectroscopy; (iv) complementary
studied from n-octane23 on activated MIL-53(Al), but the DFT calculations of interactions of water, IND, and NAP with
adsorbed amount of IND was not reported.23 Adsorption of binding sites in the cluster of MIL-53(Al) due to the π−π
IND and quinoline on activated MIL-53(Al) was tentatively interactions and hydrogen bonding, and (v) the spectroscopic
assigned to the π−π interactions, but no experimental proof study of kinetics and stoichiometry of adsorption of IND vs
was provided.23 To our knowledge, no direct spectroscopic NAP on hydrated A100 by the complementary in-situ time-
characterization of the nature of chemical bonds between the dependent fluorescence spectroscopy and ex-situ UV−vis
aromatic and the heteroaromatic adsorbate and adsorption sites spectroscopy.
in MIL-53 or A100 MOF was reported. The mechanism of
adsorption of the smallest fused-ring aromatic hydrocarbon 2. EXPERIMENTAL AND THEORETICAL METHODS
naphthalene on mesoporous F300 MOF was reported by our
group.24 Adsorption of naphthalene on microporous MOF 2.1. Chemicals. Naphthalene (NAP), indole (IND), 1,4-
[Cu2(BDC)2(dabco)] was reported,25 but its mechanism was benzenedicarboxylic acid (BDCH2), and A100 Basolite MOF
not determined. Adsorption of the fused-ring hydrocarbons on were from Sigma; DMF and n-heptane of spectroscopic purity
MIL-53(Al) and A100 MOF was not reported, to our were from Fisher.
knowledge. 2.2. Activation and Hydration of A100. The as-received
Fluorescence spectroscopy is a convenient nondestructive A100 Basolite termed asA100 was activated at 150 °C and <1 ×
method for studies of adsorption.26 However, there are very few 10−3 Torr for 24 h36 and kept isolated from moisture in
papers on characterization of adsorption of aromatic and ambient air; the obtained activated A100 (actA100) has the Hill
heterocyclic compounds in solution27 or titration of adsorbate formula C8H5AlO5. Upon activation, actA100 was placed inside
in solution with the use of fluorescence spectroscopy.28 Direct the desiccator with liquid water to be in contact with water
experimental determination of the type of chemical bonding in vapor at relative humidity (RH) ≈ 100% at 25 °C. The
the “host−guest” adsorption complexes of the MOFs and desiccator was kept closed for 2 or 7 days, and obtained
aromatic hydrocarbons or aromatic N-heterocyclic compounds hydrated A100 MOFs were termed 2hydA100 and 7hydA100,
was not reported by fluorescence spectroscopy, to our respectively.
knowledge. Further, fluorescence spectroscopy is well suited 2.3. Preparation of the Stoichiometric Adsorption
to study the MOFs29,30 and mechanisms of adsorption on the Complexes of A100 with IND or NAP. The sample of 0.024
MOFs, but the origin of the fluorescence from many MOFs is g asA100 was thermally activated in a 0.5 cc quartz fluorescence
not known. The origin of the fluorescence from MIL-53(Al) cuvette, and 0.5 cc of 0.08 M solution of IND or NAP in n-
was tentatively assigned to emission due to the LMCT,31 while heptane was added under argon. The cuvette was promptly
it was noted in a recent critical review that the LMCT closed with a PTFE (polytetrafluoroethylene) stopper, shaken
fluorescence of MIL-53(Al) is unlikely for thermodynamic for 1 h on a rotary shaker (Roto-Shake Genie), opened under
reasons.32 argon, and placed into the high-vacuum system, and n-heptane
Density functional theory (DFT) is an ab initio technique of solvent was slowly evaporated at 100−1000 mTorr. The turbo
quantum chemistry that is commonly used to calculate the pump was turned on until pressure < 1 mTorr; then the
energies and geometries of adsorption sites and adsorption vacuum system was vented with argon; the cuvette was closed
complexes. Due to a high requirement on the CPU, DFT with a PTFE stopper and sealed with Parafilm tape.
calculations are usually performed on the small clusters 2.4. Steady-State Fluorescence Spectroscopy. Steady-
containing one metal site and one or a few linker units.15 state fluorescence spectroscopy experiments were conducted
The DFT calculations of the small clusters of HKUST-1 are using a Cary Eclipse spectrometer equipped with a custom-
rather abundant,33,34 while calculations of adsorption on MIL- designed angular fluorescence accessory from Quantum
53 are rare. Small clusters of MIL-53(Ga) and MIL-53(Al) Northwest Inc. The angle of incidence of the excitation beam
containing one metal site and two BDC linkers were utilized in was set at 30° from the normal to the surface of the cuvette to
the quantum chemical study of “breathing” upon adsorption minimize spectral artifacts due to primary and secondary
and desorption of water.15 To our knowledge, calculations of reabsorption of light.37 Visualization of the cross-section of the
the structure and binding energy in the host−guest adsorption photoexcitation beam reaching the surface of the cuvette
complexes formed by either aromatic or heteroaromatic allowed the conclusion that the excitation beam was completely
compounds with the cluster of MIL-53 were not reported. absorbed by the specimen. Wavelength-dependent steady-state
Recently, we reported a study of reactive adsorption of H2S fluorescence measurements were conducted at λexc = 260, 280,
on Cu(II) sites in Cu−ZnO/SiO2 sorbents by the comple- 310, and 372 nm, and bandwidths of excitation and emission
mentary UV−vis diffuse reflectance spectroscopy (UV−vis slits were chosen at 5 and 5 nm to avoid spectral saturation.
DRS), electron spin resonance (ESR) spectroscopy, and DFT 2.5. Nanosecond Time-Resolved Fluorescence Spec-
computations.35 Most recently, we reported the mechanism of troscopy. Nanosecond time-resolved fluorescence spectrosco-
adsorption of aromatic sulfur compounds on C300 MOF33 and py measurements were conducted using the Fluorolog-3
mechanisms of adsorption of N-heterocyclic compound IND vs spectrometer from Horiba Scientific. The sample in the
fused-ring aromatic hydrocarbon NAP on F300 MOF.24 fluorescence cuvette was oriented at 30° from the normal to
Herein, we report the following: (i) direct experimental proof the front surface of the cuvette. The light source was pulsed
of the origin of the near-UV and visible fluorescence from NanoLED with λexc = 366 nm and pulse duration < 1 ns. Time-
A100; (ii) the stoichiometry of adsorption complexes formed resolved emission was collected at the center wavelength 440
by activated A100 with water, IND, and NAP; (iii) nm and bandwidth 8 nm.
2492 DOI: 10.1021/jp510272s
J. Phys. Chem. C 2015, 119, 2491−2502
The Journal of Physical Chemistry C Article

2.6. Kinetics of Adsorption by in-Situ Time-Depend- Scheme 1. Changes of Mass and Chemical Formula of
ent Fluorescence Spectroscopy in Suspension. A Cary Activated and Hydrated A100 MOF
Eclipse fluorescence spectrometer was equipped with a custom-
designed angular fluorescence accessory that had a ministirrer
that can rotate the PTFE-coated magnetic ministir bar inside
the sealed fluorescence cuvette. This accessory allowed
maintaining the 3.5 cc quartz fluorescence cuvette at a constant
temperature of 20 °C. The sample of 0.144 g of hydrated
2hydA100 was placed in the cuvette, the PTFE-coated
magnetic ministir bar was added under argon, 3.5 cc of 0.08
M solution of IND or NAP in n-heptane was added under
argon, and the cuvette was closed. The suspension in the
cuvette was continuously stirred, and the time-dependent
fluorescence spectra were continuously measured at λexc = 310
nm.
2.7. Kinetics of Adsorption by ex-Situ UV−Vis
Absorption Spectroscopy. Chemical analysis of the liquid
phase of the suspension with A100 was conducted in parallel
with the in-situ time-dependent fluorescence spectroscopic
measurements (above). Periodically, the stirrer was stopped,
the MOF was allowed to settle, an aliquot of 6 μL of liquid 3.2. Fluorescence Spectra of A100 MOF. The
phase was collected from the top of the cuvette, and a clear fluorescence spectra obtained with an empty quartz cuvette
supernatant was diluted with n-heptane by a factor of 500 and using angular accessory (see Experimental and Theoretical
analyzed by UV−vis spectroscopy. Molar concentrations were Methods) had no peaks as expected. Figure 1A shows the
determined by the Bougert−Lambert−Beer method. normalized fluorescence spectra of actA100 (solid black line)
2.8. Quantum Chemical Calculations.38,39 All calcu- and 7hydA100 (solid blue line) at λexc = 260 nm.
lations were conducted with the Gaussian 09 program,40 and
geometries were calculated at the M06-2X/LANL2DZ level.

3. RESULTS AND DISCUSSION


3.1. Preparation of hydA100 by Water Adsorption.
Prior to spectroscopic measurements, we determined the
change of mass of asA100 after its thermal activation to form
actA100; the decrease of mass was 4.8 wt % due to the loss of
“free” BDCH2 and adsorbed water. Next, we determined the
increase of the mass of actA100 due to subsequent adsorption
of water after prolonged exposure (up to 7 days) in water vapor
at RH ≈ 100% and 25 °C (see Experimental and Theoretical
Methods); 7hydA100 was kept in a closed vessel while
weighing. Each structural unit of MIL-53(Al) and A100
contains four BDC linkers; each linker is shared between two
neighboring units; thus, each unit contains two BDC linkers,
(AlOH)2−[BDC]2. The 7hydA100 obtained by us has the
formula (AlOH)2−[BDC]2−(H2O)8, Scheme 1.
Next, we thermally activated 7hydA100 following the same
procedure as for the activation of asA100 (see Experimental
and Theoretical Methods). The decrease of mass of 7hydA100 Figure 1. Fluorescence spectra of actA100 (solid black line) and
7hydA100 (solid blue line): (A) at λexc = 260 nm and (B) at λexc = 310
due to the loss of water was 24 wt %; thus, within the error, nm.
actA100 with formula (AlOH)2−[BDC]2 was formed again.
When hydration of actA100 in water vapors was conducted for
2 days, the increase of mass of actA100 corresponds to the
formula (AlOH)2−[BDC]2−(H2O)3 of obtained 2hydA100, The spectra are nearly of the same shape and intensity; thus,
Scheme 1. hydration of actA100 does not cause quenching or increase the
Adsorbed amounts of water on A100 (Scheme 1) are higher fluorescence. For both actA100 and 7hydA100 MOFs, there is a
than those reported12 for MIL-53(Al). We note the following large fluorescence peak centered at ca. 380 nm and a much
major differences. First, this is a controlled high relative smaller peak at 320 nm. The fluorescence spectra obtained at
humidity (RH ≈ 100%) in our experiments (see Experimental λexc = 280 nm (not shown) show the same spectral position and
and Theoretical Methods). Second, this is a long adsorption shapes of the peaks. Figure 1B shows the fluorescence spectra
time in our experiments (2−7 days) vs brief exposure of MIL- of actA100 (solid black line) and 7hydA100 (solid blue line) at
53(Al) to air in reported study.12 We believe that studies of λexc = 310 nm; there is only one fluorescence peak at ca. 380
water adsorption under controlled high humidity and nm. The spectra of actA100 and 7hydA100 are nearly of the
prolonged adsorption times are useful for understanding the same shape and intensity; thus, hydration of actA100 does not
behavior of MOFs. cause any quenching or increase of the fluorescence at λexc =
2493 DOI: 10.1021/jp510272s
J. Phys. Chem. C 2015, 119, 2491−2502
The Journal of Physical Chemistry C Article

310 nm. The fluorescence spectra in Figure 1 imply the energy


level diagram as in Scheme 2.

Scheme 2. Proposed Energy Levels and Excitation and


Fluorescence Transitions in A100 MOF

Figure 2. UV absorption spectra of 0.0007 M solution of BDCH2 in


aqueous 0.1 M HCl (dashed line) and 0.002 M solution of BDCH2 in
phosphate buffer at pH = 8.5 (solid line).

fluorescence), (iii) due to energy transfer from one linker to


another (the interlinker fluorescence) aka the ligand-to-ligand
charge transfer, LLCT, (iv) the ligand-to-metal charge transfer
The proposed energy level diagram is consistent with the (LMCT) fluorescence due to the transfer of electron density
Kasha rule:41 photoexcitation at 260 nm proceeds from the S0 from the linker to the metal cation in the MOF, (v) from
to the S2 level, and then a weak emission at 320 nm occurs due cations of f metals in the MOFs. It was speculated that the
to the transition from the S2 to the S0 level. However, the main fluorescence from MIL-53(Al) centered at ca. 410 nm may
relaxation pathway is via internal conversion (IC) from the S2 originate from the LMCT;31 however, no experimental or
to the S1 level and the subsequent high intensity emission from computational evidence was provided. To our knowledge, no
the S1 to the S0 level with the peak at 380 nm. Both transitions experimental proof of the origin of the fluorescence spectra of
shown in Scheme 2 can be observed in Figure 1A as Al-MOFs was reported.
fluorescence bands centered at 320 and 380 nm. The excitation We decided to determine the origin of the fluorescence from
at 310 nm causes an excitation from the S0 to the S1 level and electronic levels in A100 MOF shown in Scheme 2. Electronic
the subsequent high-intensity emission from S1 to S0 centered transitions in solids are determined by UV−vis diffuse
at 380 nm, Scheme 2. reflectance spectroscopy (UV−vis DRS). The UV-DRS of
The proposed scheme of radiative transitions (Scheme 2) is A100 MOF at ∼260 nm could not be obtained using the
also consistent with reported fluorescence spectra of certain available instrument; hence, we investigated the model system:
nonsubstituted and substituted aromatic hydrocarbons. For a solution of the neutral form of aromatic linker in A100, 1,4-
example, fluorescence spectra of naphthalene and several 1- benzenedicarboxylic acid BDCH2 in water. For BDCH2, acidity
substituted naphthalenes in solution at room temperature were constants44 are pKa1 = 3.51 and pKa2 = 4.82. Therefore, at pH =
reported to contain both S2 → S0 and S1 → S0 transitions,42 8.5, both carboxylic groups are dissociated to form (BDC)2−
with rate constants for radiative decay and internal conversion. anion, while at pH = 1 both carboxylic groups are not
Interestingly, both S2 → S0 and S1 → S0 transitions in dissociated to form the “free” acid BDCH2. The solubility of
naphthalene and associated internal conversion were also BDCH2 in 0.1 M HCl to obtain pH = 1 is significantly lower
reported in the fluorescence spectra of naphthalene vapor.43 than that in phosphate buffer to obtain pH = 8.5. Figure 2
This indicates that the basic energy level diagram and radiative shows the UV spectra of 0.0007 M solution of BDCH2 in
transitions may be the same for certain aromatic compounds in aqueous 0.1 M HCl (dashed line) and 0.002 M solution of
the vapor and condensed phase. Further, the appearance of BDCH2 present as anion (BDC)2− in the buffer at pH = 8.5
both S2 → S0 and S1 → S0 transitions in fluorescence spectra of (solid line).
naphthalene and substituted naphthalenes was supported by Buffer solution at pH = 8.5 containing (BDC)2− is a model
the UV−vis absorption spectra in solution, where two system for A100 MOF where BDC linker is present as the
respective absorption bands S0 → S2 and S0 → S1 have been (BDC)2− anion connected to Al(OH)2+ cationic units. In
found.42 The BDC linker in A100 MOF contains an aromatic Figure 2, two absorption bands are present: at 240 and 295 nm
benzene ring that has structural similarity with the aromatic that correspond to transitions S0 → S2 and S0 → S1 (HOMO
system of fused-ring aromatic hydrocarbon naphthalene. The → LUMO), respectively, in neutral form BDCH2 and anionic
origin of two fluorescence transitions in A100 MOF from form BDC2−. Thus, for both BDCH2 and its anion BDC2− the
excited S1 and S2 states (Scheme 2) was investigated by us energy gap between the HOMO and the LUMO electronic
using the UV−vis absorption spectra of solution of the linker of levels is ca. 295 nm or 4.2 eV. This value of the HOMO−
A100 MOF, see Figure 2. LUMO energy gap is close to that at 4.1 eV or 302 nm reported
3.3. Origin of the Fluorescence from A100 MOF. recently for the adsorbed monolayer of terephthalic acid.45
Recent reviews on the luminescent MOFs29,30 summarize the One can see in Figure 2 that the UV spectra of BDC in
origin of the known mechanisms of the fluorescence: (i) from neutral BDCH2 form and anionic form (BDC)2− with both
the adsorbate, (ii) from the linker in the MOF (the intralinker COOH groups dissociated are very similar that implies the
2494 DOI: 10.1021/jp510272s
J. Phys. Chem. C 2015, 119, 2491−2502
The Journal of Physical Chemistry C Article

excitation of the π → π* transitions involving molecular orbitals excitation to the higher vibrational states. The first mechanism
(MOs) of aromatic system of benzene ring. On the other hand, of excitation to the higher vibrational states can be realized via
the spectral position and intensity of the n → π* transition vibrational overtones in the anharmonic oscillator with
involving the lone electron pair of the O atom of the COO selection rule Δv = ±2 for vibrational quantum number v,
group in BDCH2 would be dependent on dissociation of the compared to the selection rule for harmonic oscillator Δv = ±1.
COO group that is not the case. Thus, the fluorescence from The second mechanism for a large red shift of the fluorescence
A100 MOF (Scheme 2) is due to an intralinker emission from is via exciting combination frequencies. In our recent paper,24
the aromatic system of benzene ring of the BDC linker. the set of combination frequencies due to vibrations in the
Further, it is known that adsorbed water in MIL-53(Al) is H naphthalene molecule was detected as vibronic shoulders in the
bonded to the COO group of the BDC linker.12 A100 MOFs is fluorescence spectrum of naphthalene in diluted solution. In
similar19 to MIL-53(Al), and fluorescence spectra of actA100 addition, the reported experimental FTIR spectrum of model
(without adsorbed water) and hydA100 (with adsorbed water compound 2-aminoterephthalic acid46 that is structurally
coordinated to COO group) are essentially the same, Figure 1. similar to BDCH2 contains a very intense set of overtone and
This finding further indicates that the fluorescence from A100 combination modes. Therefore, the strong red shift of the
MOF is independent of chemical bonds involving the COO fluorescence from the monomer of BDC in Figure 3A is
group, i.e., the fluorescence in Figure 1 originates from the consistent with the excitation and emission from electronic
MOs of benzene ring. We checked the origin of the states of the aromatic system of the benzene ring in BDC
fluorescence of A100 MOF by comparing the fluorescence involving vibrational overtones and/or combination frequen-
spectra from A100 MOF and solutions of model compound cies. Further, the fluorescence spectrum of diluted 5 × 10−4 M
BDCH2 in water at different pH. water solution of another model compound, benzoic acid,47
Figure 3A shows the normalized fluorescence spectra at λexc = appears at 310−390 nm when excited with λexc = 250 nm. Thus,
260 nm obtained from BDCH2 in its anionic (BDC)2− form as both benzoic acid47 and terephthalic acid BDCH2 (Figure 3A)
in diluted solution undergo a significant red shift of the
fluorescence when excited close to the maximum of absorption,
and hence, the origin of the fluorescence is the same, namely,
excitation of aromatic system of benzene ring.
Figure 3B shows the fluorescence spectra at λexc = 260 nm
from A100 MOF that contains BDC linker in its anionic
(BDC)2− form (solid line) vs neat solid BDCH2 present as
molecular form (dotted line). One can see that the presence of
Al(III) cation in A100 MOF does not significantly change the
fluorescence spectrum compared to neat BDCH2. Therefore,
the fluorescence from A100 in Figure 3B is unlikely to be the
LMCT fluorescence involving Al(III) cation. This conclusion is
consistent with the recent critical review:32 “given the
extraordinary energetic difficulty of accomplishing one electron
reduction of Al(III)... Presumably emission is instead occurring
from a linker-localized excited state”. Further, positions of the
fluorescence spectral maxima and spectral widths in Figure 3
are about the same for solid A100 MOF, solid BDCH2,
(BDC)2− in diluted 0.002 M solution, and BDCH2 in diluted
0.0007 M solution. This is a standard method in fluorescence
spectroscopy to use diluted solutions with concentration
Figure 3. Normalized fluorescence spectra at λexc = 260 nm of (A)
∼<10−3 M to observe only the emission from the monomers
0.002 M solution of BDCH2 in phosphate buffer at pH = 8.5 (solid
line) and 0.0007 M solution of BDCH2 in 0.1 M HCl (dotted line) but not from excimers or complexes. Indeed, in 0.002 M
and (B) A100 MOF (solid line) and neat BDCH2 (dotted line). solution of (BDC)2− at pH = 8.5 (Figure 3) there are 0.002 mol
of anion (BDC)2− present per 55.6 mol of water solvent, i.e.,
one anion (BDC)2− is surrounded with 27 800 water molecules,
0.002 M solution in phosphate buffer at pH = 8.5 (solid line) vs on average. In such diluted solution, the observed fluorescence
neutral BDCH2 as 0.0007 M solution in 0.1 M HCl (dotted (Figure 3A) is the intramolecular fluorescence from electronic
line). states of (BDC)2− monomer rather than the fluorescence from
The spectra have nearly the same position of spectral any molecular associate involving more than one molecular unit
maximum and shape that indicates that the fluorescence of BDC. In addition, electrically charged (BDC)2− anions in
emission is not dependent on whether the COOH group is solution at pH = 8.5 would electrostatically repel each other,
dissociated. These data are consistent with absorption spectra thus making formation of any excimers or other clusters highly
of BDCH2 vs (BDC)2− in solution (Figure 2) that are also unlikely. Two conclusions can be made by comparing Figure
independent of the pH. Therefore, the fluorescence from BDC 3A and 3B. First, the fluorescence spectra are essentially the
originates from electronic states of the aromatic system of the same for BDC monomers in diluted solutions in Figure 3A and
phenyl ring in the BDC linker. The fluorescence from either BDC linker in A100 MOF (Figure 3B). Therefore, in all cases
BDC2− or BDCH2 in Figure 3A appears as a broad peak the fluorescence originates from the monomer of the BDC unit
centered at 380 nm when excited at λexc = 260 nm. Thus, the rather than from some associated form of BDC units, e.g.,
fluorescence from BDC2− or BDCH2 is strongly “red shifted” excimers or the LLCT excited state in A100. This conclusion is
from the excitation wavelength that can be explained by the consistent with the relative position of neighboring BDC linker
2495 DOI: 10.1021/jp510272s
J. Phys. Chem. C 2015, 119, 2491−2502
The Journal of Physical Chemistry C Article

units in activated A100 and MIL-53(Al): BDC linkers are green line) and indole (AlOH)2−[BDC]2−[IND]1 (solid red
positioned at ca. 90° to each other;12 therefore, any interlinker line).
energy transfer (LLCT) from one BDC unit to another is There is a decrease of the fluorescence in the complex
unlikely for sterical reasons. Second, in all cases, the (AlOH)2−[BDC]2−[NAP]1 by ca. 50% in the spectral range
fluorescence at 380 nm originates from the S1 → S0 or ca. 350−400 nm after adsorption of naphthalene, compared to
LUMO → HOMO transition within the aromatic system in the A100, Figure 4A. This emission range surrounds the emission
BDC monomer. In summary, as shown by combination of the maximum of A100 MOF at ca. 380 nm, see also Figure 1A.
wavelength-dependent fluorescence spectra in the solid phase Therefore, quenching of emission from A100 occurs after
and diluted solutions of model systems and the complementary adsorption of naphthalene. Further, there is an increase of
UV−vis absorption spectroscopy in solution, the fluorescence emission at ca. 410−480 nm after adsorption of naphthalene,
from A100 MOF originates in the intralinker transitions of Figure 4A. Such an “extra” fluorescence signal corresponds to
BDC. Therefore, one can expect that binding of the adsorbate emission from adsorbed highly fluorescent49 NAP molecule in
molecule to the aromatic system of the BDC linker would the complex with A100 that appears in the spectrum in addition
modify the fluorescence. We test this hypothesis below by to the intralinker fluorescence from the BDC linker. This is
comparing the fluorescence spectra of A100 MOF and its further commented in the discussion of Figure 4B below. On
host−guest adsorption complexes with indole and naphthalene. the other hand, the fluorescence intensity in the spectrum of
3.4. Fluorescence Spectra of the Host−Guest Adsorp- adsorption complex with indole (AlOH)2−[BDC]2−[IND]1 is
tion Complexes of NAP and IND with actA100 at λexc = a factor of ∼10 lower than that of actA100 (AlOH)2−[BDC]2.
260 and 310 nm. Prior to spectroscopic measurements, we Therefore, there is a strong quenching of the fluorescence from
determined the increase of mass of actA100 after the formation the BDC linker in A100 MOF upon the formation of the host−
of the host−guest adsorption complex with naphthalene (NAP) guest adsorption complex with indole. The absence of emission
or indole (IND). The increase of mass corresponds to the from adsorbed indole is attributed to an efficient intersystem
formulas of the host−guest adsorption complexes crossing to the triplet state due to the significant spin−orbit
(AlOH)2−[BDC]2−[NAP]1 and (AlOH)2−[BDC]2−[IND]1, coupling in the aromatic heterocyclic compounds, compared to
respectively. In other words, one NAP or IND molecule is aromatic hydrocarbons.50 We propose that in the host−guest
adsorbed per structural unit in A100 MOF. The kinetic adsorption complex (AlOH)2−[BDC]2−[IND]1 an adsorbed
diameter of indole at 5.4 Å23 is comparable to the micropore indole molecule accepts the energy of the photoexcited states of
size36 in A100 MOF at 5.3 Å. On the other hand, the kinetic the BDC linker and transfers it to the triplet state and/or
diameter of the naphthalene molecule at 6.57 Å48 is larger than vibrations. Such mechanism is consistent with the proposed
the micropore size. We speculate that “breathing” of A100 interlinker fluorescence of BDC in A100 MOF, namely, in the
occurs to admit one naphthalene molecule to the structural unit adsorption complex with indole, the excitation energy of BDC
(AlOH)2−[BDC]2 to form the stoichiometric host−guest linker is transferred not to another BDC linker but to the
adsorption complex (AlOH)2−[BDC]2−[NAP]1. weakly fluorescent adsorbed molecules of indole that causes the
Figure 4A shows the fluorescence spectra at λexc = 260 nm observed quenching of the fluorescence in the adsorption
obtained with actA100 with formula (AlOH)2−[BDC]2 (solid complex. In solution, the fluorescence quenching of the excited
black line) and host−guest adsorption complexes formed by state of the π−π complex containing indole and benzene has
actA100 with naphthalene (AlOH)2−[BDC]2−[NAP]1 (solid been reported.51 We propose that in the host−guest adsorption
complex of indole and A100 MOF, the bonding is between the
π system of adsorbed indole and the π system of the BDC
linker. We will further investigate this interaction using the
DFT calculations, section 3.8.
Figure 4B shows the fluorescence spectra of (AlOH)2−
[BDC]2 (solid black line) and the host−guest adsorption
complexes with naphthalene (AlOH)2 −[BDC]2−[NAP]1
(solid green line) and indole (AlOH)2−[BDC]2−[IND]1
(solid red line). λexc = 310 nm was chosen to excite S1 → S0
fluorescence transition, and the same observations are made as
with λexc = 260 nm: there is an increase of the fluorescence in
the range ca. 400−480 nm in (AlOH)2−[BDC]2−[NAP]1 vs
(AlOH)2−[BDC]2. This indicates that the origin of that “extra”
emission at ca. 400−480 nm is an emission from adsorbed
naphthalene, Figure 4A and 4B. Therefore, there are two kinds
of spectral changes in Figure 4 after adsorption of naphthalene:
(1) quenching of emission from A100 at ca. 350−400 nm as
observed with λexc = 260 nm (Figure 4A) and (2) increase of
emission due to adsorbed naphthalene at ca. 400−480 nm as
observed with both λexc = 260 (Figure 4A) and 310 nm (Figure
4B). On the other hand, a complete quenching of emission
from the BDC linker in A100 occurs in the adsorption complex
Figure 4. Fluorescence spectra of (AlOH)2−[BDC]2 (black line), with indole (AlOH)2−[BDC]2−[IND]1.
complex (AlOH)2−[BDC]2−[NAP]1 (green line), and (AlOH)2− 3.5. Fluorescence Spectra of BDCH2 and IND in
[BDC]2−[IND]1 (red line) at (A) λexc = 260 nm and (B) λexc = 310 Solution at λexc = 310 nm. We have shown that the
nm. fluorescence from A100 MOF originates from the aromatic
2496 DOI: 10.1021/jp510272s
J. Phys. Chem. C 2015, 119, 2491−2502
The Journal of Physical Chemistry C Article

system of the BDC linker, and fluorescence quenching by (Figure 5) is similar to quenching of the fluorescence from the
indole in the adsorption complex is due to the π−π interaction BDC linker in A100 by IND in the adsorption complex at λexc =
between adsorbed indole and the aromatic system of the linker. 310 nm, Figure 4B. Quenching of the fluorescence in Figure 5A
We decided to measure the fluorescence spectrum of a solution occurs in the absence of Al(III) cation; hence, the fluorescence
of BDCH2 vs an equimolar solution of BDCH2 and IND. quenching does not depend on whether the Al(III) cation is
BDCH2 is insoluble in the nonpolar solvents such as hexane; present. Therefore, this is the fluorescence from the BDC linker
therefore, polar DMF was used. Figure 5A shows the in A100 MOF, rather than the LMCT fluorescence that is
quenched by the IND adsorbate.
The fluorescence spectra of the host−guest adsorption
complex (AlOH)2−[BDC]2−[NAP]1 in Figure 4A and 4B are
also expected to originate mostly from the excited states of the
BDC linker. However, when excited at λexc = 310 nm (Figure
4B), the complex shows the significant spectral contribution
from adsorbed NAP. To distinguish between the contribution
to the total fluorescence spectrum due to the emission from the
BDC linker in A100 MOF (that may be quenched by the IND
adsorbate) and any possible emission from either NAP or
perhaps IND adsorbate we performed a complementary
experiment with an excitation wavelength of 372 nm.
3.6. Fluorescence Spectra of the Host−Guest Adsorp-
tion Complexes ( AlOH) 2 −[BDC] 2 −[NAP ] 1 an d
(AlOH)2−[BDC]2−[IND]1 at λexc = 372 nm. Figure 6 shows

Figure 5. (A) Fluorescence spectra of 0.08 M solution of BDCH2 in


DMF (solid line) and equimolar 0.08 M solution of BDCH2 and IND
in DMF (dashed line) at λexc = 310 nm. (B) Absorption spectra of
solution of BDCH2 in DMF (solid line) and equimolar solution of
BDCH2 and IND in DMF (dashed line).

fluorescence spectra of a 0.08 M solution of BDCH2 in DMF


(solid line) and equimolar 0.08 M solution of BDCH2 and IND
in DMF (dashed line). To interpret the intensity of the two Figure 6. Fluorescence spectra of (AlOH)2−[BDC]2 (black line),
fluorescence spectra in Figure 5A, the number of photons complex (AlOH)2−[BDC]2−[NAP]1 (green line), and complex
absorbed at the excitation wavelength λexc = 310 nm needs to (AlOH)2−[BDC]2−[IND]1 (red line) at λexc = 372 nm.
be the same in a solution of BDCH2 and in solution containing
both BDCH2 and IND. We measured the UV−vis absorption
spectra of a solution of BDCH2 in DMF and a binary equimolar the fluorescence spectra of actA100 (AlOH)2−[BDC]2, the
solution of BDCH2 and indole in DMF. Since these 0.08 M host−guest adsorption complex with naphthalene (AlOH)2−
solutions have too strong absorbance at the wavelength of [BDC]2−[NAP]1, and host−guest adsorption complex with
interest 310 nm, we had to dilute these solutions with DMF to indole (AlOH)2−[BDC]2−[IND]1 at λexc = 372 nm. As was
decrease absorbance to a measurable value. Figure 5B shows checked in the reference experiments with IND and NAP in
the absorbance spectra of a 0.08 M solution of BDCH2 in DMF solutions and the solid phase (data not shown), λexc = 372 nm
and a 0.08 M equimolar binary solution of BDCH2 and IND in would photoexcite only electronic states in A100 MOF.
DMF after dilution with pure DMF by a factor of 750. One can For the complex (AlOH)2−[BDC]2−[IND]1, there is a
see that absorbance at 310 nm for the binary solution complete quenching of the fluorescence from the BDC linker.
containing equimolar amounts of BDCH2 and indole is only This is consistent with fluorescence quenching by IND
10% higher than that for the solution with BDCH2 only. adsorbate at λexc = 310 nm, Figure 4B. In Figure 6, there is
Therefore, the number of absorbed photons is about the same some decrease of the fluorescence from (AlOH)2−[BDC]2−
(∼10% difference) for a solution of BDCH2 alone and solution (NAP)1; therefore, when NAP in (AlOH)2−[BDC]2−(NAP)1
of BDCH2 in the presence of IND. complex is not photoexcited, quenching of the fluorescence
The fluorescence spectra in Figure 5A were measured at λexc from the BDC linker occurs. Fluorescence quenching by the
= 310 nm to excite the S1 → S0 transition in BDCH2 to observe NAP adsorbate is much weaker than that by the IND in the
fluorescence quenching of the BDC linker by adsorbed indole. complex (AlOH)2−[BDC]2−(IND)1 as expected given the
One can see a strong quenching of the fluorescence (by factor above-mentioned much more efficient intersystem crossing to
∼6) from BDCH2 in the presence of an equimolar amount of the nonemitting states in the excited molecule of IND50
IND. Fluorescence quenching of BDCH2 by IND in solution compared to NAP. Those data imply that NAP and IND have
2497 DOI: 10.1021/jp510272s
J. Phys. Chem. C 2015, 119, 2491−2502
The Journal of Physical Chemistry C Article

Table 1. Nanosecond Time-Resolved Fluorescence Data from A100 and Its Complexes
(AlOH)2−[BDC]2 (AlOH)2−[BDC]2−[NAP]1 2hydA100, (AlOH)2−[BDC]2−(H2O)3 (AlOH)2−[BDC]2−[IND]1
τ 1 (ns) 3.56 ± 0.06 3.22 ± 0.07 2.91 ± 0.04 3.81 ± 0.05
B1 (au) 895.97 939.72 1431.54 705.54
τ2 (ns) 19.79 ± 0.37 16.25 ± 0.24 15.51 ± 0.24 25.74 ± 0.31
B2 (au) 208.75 264.30 261.89 116.67
ratio B1/B2 4.29 3.55 5.47 6.04

the same mechanism of adsorption on actA100 that is due to (bottom) that are π−π bonded to the aromatic system of the
the formation of the π−π adsorption complexes with the BDC linker.
aromatic system of the BDC linker. In order to learn more The relatively small ratio B1/B2 for the complex with
about the origin of the fluorescence in the adsorption naphthalene (AlOH)2−[BDC]2−[NAP]1 is close to that of
complexes of A100 with NAP and IND, we decided to use actA100 (AlOH)2−[BDC]2. On the other hand, the larger ratio
the complementary nanosecond time-resolved fluorescence B1/B2 for the adsorption complex with indole
spectroscopy. (AlOH)2−[BDC]2−[IND]1 is close to that of hydA100
3.7. Nanosecond Time-Resolved Fluorescence Spec- (AlOH)2−[BDC]2−(H2O)3. Those relationships imply the
troscopy of Adsorption Complexes. The nanosecond time- similar kinds of bonding of the adsorbates with similar
resolved fluorescence decay curves were obtained from actA100 functional groups, i.e., that adsorbed IND may be bonded to
and complexes of actA100 with IND, NAP, and water at λexc = A100 in a manner similar to water, by hydrogen bonding. In
366 nm that is close to λexc = 372 nm (Figure 6) at which no order to investigate the relative contributions from the π−π
excitation of the adsorbate occurs. Therefore, electronic states interactions and H bonding between adsorbed IND or NAP
of the BDC linker in A100, but not IND or NAP, were and A100, we use the complementary DFT calculations, section
photoexcited at λexc = 366 nm. The intensity of the nanosecond 3.8; complementary experimental mechanistic research is in
time-resolved fluorescence was numerically fitted with one-,
progress.
two-, and three-exponential decay functions. The smallest value
3.8. DFT Computations of Geometry, Binding Energy,
of chi-square parameter was obtained in fitting with a
and Electronic Spectra of Adsorption Complexes of
biexponential decay: f(x) = A + B1 exp(−x/τ1) + B2 exp(−x/
Naphthalene and Indole with Activated A100 MOF. The
τ2). Table 1 shows the obtained numeric values for time
constants τ1 and τ2 and their contributions B1 and B2, M06-2X exchange/correlation density functional has been
respectively. thoroughly tested52−57 and found to incorporate some
All samples show two time constants τ1 and τ2 with similar medium-range dispersion interactions. Calculations on the
numeric values, which indicates the emission from the same model complex NAP-MIL-53(Al) with a naphthalene molecule
electronic states, i.e., from aromatic ring of the BDC linker as required 250 atoms and 1758 basis functions with the
expected for λexc = 366 nm. Scheme 3 shows the proposed LANL2DZ basis set. Since LANL2DZ is a rather small basis
energy diagram of the fluorescence from BDC in A100 MOF in set, the basis set superposition errors (BSSE) in the binding
the presence of IND adsorbate (top) and NAP adsorbate energies of both NAP and IND were quite large (>5 kcal/mol).
For that reason, the binding energies of both NAP and IND
Scheme 3. Proposed Energy Diagram of Electronic Levels were corrected for BSSE by using the counterpoise method.58
and Transitions in the Adsorption Complexes of NAP and In order to match the conditions of the adsorption
IND with Activated A100 MOF experiment and the DFT computations, the “low-coverage
limit” with either IND or NAP was chosen, namely, in the
stoichiometr ic host−guest adsorpt ion com plexes
(AlOH)2−[BDC]2−[NAP]1 and (AlOH)2−[BDC]2−[IND]1
there is, on average, one adsorbate molecule per structural
unit of the MOF. The second adsorbed NAP or IND molecule
would be located inside the next unit; thus, adsorbed NAP or
IND molecules would not interact with each other. Therefore,
in the low-coverage limit as pertinent to the choice of the
relevant cluster for the DFT calculations, only the interactions
between the adsorbate molecule and the adsorption site (BDC
linker in this case) are to be taken into account. The cluster
with more than one adsorbed molecule per cavity would
correspond to the higher adsorption capacities that are to be
found at the adsorption equilibrium in solution. Those
structures would not properly model the stoichiometric
host−guest adsorption complexes of A100 MOF that were
maintained in vacuum (see Experimental and Theoretical
Methods) to remove any weakly bound NAP or IND. Thus, the
clusters with more than one adsorbate molecule per cavity were
not further considered by us in this study, and we used clusters
with one molecule per cavity.
2498 DOI: 10.1021/jp510272s
J. Phys. Chem. C 2015, 119, 2491−2502
The Journal of Physical Chemistry C Article

Figure 7A shows the optimized structure of the cluster with apparently to accommodate the hydrogen bond. The IND
one adsorbed NAP molecule, and Figure 7B shows that with molecule is situated above one BDC linker to allow the
one adsorbed IND molecule. hydrogen atom in the NH group to interact with an oxygen
atom in the COO group of the adjacent BDC linker. Hydrogen
bonding is detected between the NH hydrogen of the IND
molecule and an oxygen atom of the BDC linker, with the
calculated NH---O distance of 2.59 Å. The recent review of the
typical geometry of the hydrogen-bonded NHO groups60
provides the length distribution with a center at 2.9 Å.
Therefore, the calculated length of the NH---O bond in the
cluster of MIL-53(Al) with indole corresponds to rather short
and strong hydrogen bonding.
A critical geometrical analysis of the π−π stacking in the
complexes of aromatic nitrogen-containing ligands concluded61
that the typical π−π interaction occurs via an offset or slipped
stacking arrangement, i.e., the interacting rings are parallel
displaced. Further, in complexes with typical π−π stacking,61
the ring normal and the vector between the ring centroids
would form an angle of about 20° up to centroid−centroid
distances of 3.8 Å. This is observed in Figure 7B for adsorbed
IND: the benzene aromatic ring of IND is located over the
benzene ring of the BDC linker, with a ring−ring distance of ca.
3.25 Å. Therefore, the DFT computations indicate the strong
π−π stacking interaction between the aromatic system of
adsorbed IND and the aromatic system of the BDC linker. This
π−π stacking interaction takes place simultaneously with the H
bonding of adsorbed IND with BDC linker as discussed above.
Table 2 shows the calculated binding energies (without and

Table 2. Binding Energies of IND and NAP to MIL-53(Al)


Cluster
binding energy binding energy BSSE correcteda
cluster (kcal/mol) (kcal/mol)
MIL- −17.9 −10.9
53(Al)+NAP
MIL- −20.5 −13.8
53(Al)+IND
a
BSSE represents basis set superposition errors.

Figure 7. Optimized structure of model cluster of MIL-53(Al) with with counterpoise correction) for adsorption of one NAP and
(A) naphthalene and (B) indole: black atoms, carbon; white atoms, IND molecule into the cavity of the MIL-53(Al) model cluster
aluminum; small white atoms, hydrogen; blue atom, nitrogen. computed at the M06-2X/LANL2DZ level of theory.
For the cluster with one adsorbed IND molecule, IND
The Cartesian coordinates of the MIL-53(Al) model cluster molecule binds 2.9 kcal/mol more strongly than one adsorbed
and the model complex with one molecule of NAP or IND NAP molecule. The enhanced binding can be attributed to
optimized at the M06-2X/LANL2DZ level are available in hydrogen bonding between the NH hydrogen of the IND
Table S1, Supporting Information. In Figure 7A, the NAP molecule and an oxygen atom of the BDC linker. In addition,
molecule binds parallel with the bottom of the cavity in the gap there are two short CH---O distances between IND (Figure
between two BDC linkers. In this orientation, the NAP 7B) and different COO groups of the MIL-53(Al) model, as
molecule has a “parallel displaced” orientation59 with the 6- well as two short CH---O distances between NAP and different
membered π system of the BDC linker. The parallel orientation COO groups of the MIL-53(Al) model (Figure 7A). In both
of the NAP molecule suggests the interactions between the cases, the short distances are between the CH hydrogen atoms
aromatic π system of adsorbed NAP and that of the BDC of the adsorbate molecule and the COO group on one linker
linker. The distance between the NAP plane and the BDC forming the end of the cavity of MIL-53(Al) cluster. The
linker plane is 3.08 Å, Figure 7A. The spatial limit to the π−π distances are 2.24 and 2.45 Å in the cluster with adsorbed IND
interactions between two atoms is the sum of their van der and 2.57 and 2.62 Å in the cluster with adsorbed NAP. In the
Waals radii rvdW. For carbon,60 the rvdW = 1.77 Å yields the long- case the cluster with IND, the CH---O interactions are short
range limit for the π−π interaction at twice the rvdW or 3.54 Å. enough to suggest that they contribute to binding. Overall, our
Therefore, the distance between the plane of the NAP molecule DFT data on the energy and geometry of binding of IND and
and the plane of the BDC linker in Figure 7A at 3.08 Å is within NAP to the model cluster of MIL-53(Al), Figure 7A and 7B
the typical range of the π−π interactions. In contrast to the and Table 2, indicate a strong π−π interaction between
NAP molecule, the molecular plane of IND is somewhat tilted adsorbed IND and NAP and BDC linker of MIL-53(Al).
with respect to the bottom of the MIL-53(Al) cavity, Figure 7B, Therefore, our DFT data are consistent with quenching of
2499 DOI: 10.1021/jp510272s
J. Phys. Chem. C 2015, 119, 2491−2502
The Journal of Physical Chemistry C Article

emission from the BDC linker in A100 after the formation of structural unit. Figure 8B shows the kinetics of adsorption of
the host−guest adsorption complexes with IND and NAP, IND on 2hydA100 in the same suspension as measured in
Figure 6. Yet another important conclusion of our DFT study is parallel with the in-situ fluorescence spectra by the comple-
that in the adsorption complex with IND, the NH---O mentary ex-situ UV−vis spectroscopy of the liquid phase of the
hydrogen bond with a NH---O distance of 2.59 Å is rather suspension. In ca. 60 min, adsorption equilibrium is achieved,
short and strong. In order to investigate the possibility of which is consistent with the plateau of the in-situ time-
preferential binding of IND vs NAP to A100 MOF, one needs dependent fluorescence in Figure 8A. The equilibrium of
to change the number and/or strength of hydrogen bonds. The adsorption of indole on the activated MIL-53(Al) in n-octane23
obvious method is to study the adsorption of IND vs NAP on was achieved in less than 1 h. The adsorption kinetics on
A100 MOF with a large amount of preadsorbed water hydrated A100 MOF, Figure 8B, proceeds on a similar time
molecules that can form hydrogen bonds with IND but not scale, which may indicate the diffusion-limited kinetics.
NAP adsorbate. The amount of IND adsorbed on 2hydA100 in equilibrium
3.9. Adsorption of IND and NAP on Hydrated A100 in was determined by UV−vis absorption spectroscopy (Figure
Suspension. The kinetics of adsorption of IND or NAP on 8B), and the corresponding adsorption complex has the
hydrated A100 or hydrated MIL-53(Al) in solution and the formula (AlOH)2−[BDC]2−[IND]0.6−(H2O)x. Experimental
adsorbed amounts were not reported, to our knowledge. Figure techniques available to us do not allow answering the question
8A shows the in-situ time-dependent fluorescence measure- whether the amount of preadsorbed water in 2hydA100
ments at λexc = 310 nm in suspension of 2hydA100 in 0.08 M changes as adsorption of indole on 2hydA100 in suspension
solution of IND in n-heptane. proceeds. The similar kinetic experiment on the adsorption of
naphthalene from solution in n-heptane on 2hydA100 was
conducted by us (data not shown), and the adsorption
equilibrium was achieved in about 40 min. The amount of
NAP adsorbed onto 2hydA100 in equilibrium was determined
by a similar procedure using UV−vis absorption spectroscopy.
Scheme 4 shows the stoichiometry of adsorption of NAP and

Scheme 4. Stoichiometry of Adsorption of NAP and IND on


Hydrated A100 in n-Heptane

Figure 8. (A) Area of the in-situ fluorescence spectra of 2hydA100 in


0.08 M solution of IND in n-heptane at λexc = 310 nm vs adsorption
time. (B) Adsorption kinetics of IND on 2hydA100 by UV−vis
spectroscopy. IND on 2-hydA100 in the suspension in n-heptane, and the
adsorbed amount of NAP is twice smaller on a molar basis than
that of IND.
The Y axis shows the integrated area of the fluorescence Therefore, one can conclude that the preferential adsorption
spectrum, with the Raleigh excitation peak excluded. It was of IND over NAP on hydrated A100 MOF occurs. The
determined in the reference experiments that a 0.08 M solution hydrated A100 prepared by us contains higher amounts of
of IND in n-heptane shows the negligible fluorescence at the adsorbed water than the reported12 hydrated MIL-53(Al) with
experimental settings used. Therefore, the integrated fluo- one water molecule per structural unit. The presence of the
rescence in Figure 8A originates from the electronic states of large amounts of preadsorbed polar molecules other than water
the π system of the BDC linker in A100 MOF. In Figure 8A, in the micropores of A100 and MIL-53(Al) MOFs is expected
one can see the progressive decrease of the fluorescence of to lead to the higher selectivity of adsorption of polar aromatic
A100 as a result of quenching of the fluorescence of the BDC N-heterocyclic compounds vs nonpolar aromatic hydrocarbons
linker by adsorbed IND, until the plateau after ca. 60 min. due to hydrogen bonding. Experiments are in progress to
Therefore, the adsorption of indole proceeds on hydrated investigate the selectivity of adsorption of aromatic and
2hydA100 that contains three preadsorbed water molecules per heterocyclic adsorbates on Al-MOFs with preadsorbed polar
2500 DOI: 10.1021/jp510272s
J. Phys. Chem. C 2015, 119, 2491−2502
The Journal of Physical Chemistry C Article

molecules. Another conclusion of this study is that the (4) Almarri, M.; Ma, X.; Song, C. Role of Surface Oxygen-Containing
molecules of heterocyclic aromatic adsorbate can be used as a Functional Groups in Liquid-Phase Adsorption of Nitrogen
fluorescent “self-marker” to learn about the π−π interactions Compounds on Carbon-Based Adsorbents. Energy Fuels 2009, 23,
between the linker in the MOF and the adsorbate molecule in 3940−3947.
the host−guest adsorption complexes. In this approach, the (5) Huh, E. S.; Zazybin, A.; Palgunadi, J.; Ahn, S.; Hong, J.; Kim, H.
linker in the MOF does not need to show a strong fluorescence S.; Cheong, M.; Ahn, B. S. Zn-Containing Ionic Liquids for the
(e.g., in heteroaromatic linker) if the quenching of the Extractive Denitrogenation of a Model Oil: A Mechanistic
Consideration. Energy Fuels 2009, 23, 3032−3038.
adsorbate fluorescence occurs; experiments are in progress to
(6) Saha, D.; Deng, S. G.; Yang, Z. G. Hydrogen Adsorption on
test this hypothesis in the broader context. Metal-Organic Framework (MOF-5) Synthesized by DMF Approach.
J. Porous Mater. 2009, 16, 141−149.
4. CONCLUSIONS (7) Xiang, S.-C.; Zhang, Z.; Zhao, C.-G.; Hong, K.; Zhao, X.; Ding,
The origin of the fluorescence from A100 MOF is the D.-R.; Xie, M.-H.; Wu, C.-D.; Das, M. C.; Gill, R.; et al. Rationally
intralinker fluorescence, i.e., emission from the π system of Tuned Micropores within Enantiopure Metal-Organic Frameworks for
benzene ring of the BDC linker. Activated A100 forms the Highly Selective Separation of Acetylene and Ethylene. Nat. Commun.
stoichiometric host−g ue s t adso r p ti o n co mp l ex es 2011, 2, 204.
(AlOH)2−[BDC]2−[NAP]1 and (AlOH)2−[BDC]2−[IND]1 (8) Keskin, S.; van Heest, T. M.; Sholl, D. S. Can Metal−Organic
with naphthalene and indole, and either adsorbed molecule is Framework Materials Play a Useful Role in Large-Scale Carbon
π bonded to the aromatic system of the BDC linker. The origin Dioxide Separations? ChemSusChem 2010, 3, 879−891.
(9) Li, B.; Wen, H.-M.; Zhou, W.; Chen, B. Porous Metal−Organic
of quenching of the intralinker fluorescence in A100 MOF is
Frameworks for Gas Storage and Separation: What, How, and Why? J.
due to heteroaromatic indole adsorbate and, to a much lesser Phys. Chem. Lett. 2014, 5, 3468−3479.
extent, aromatic naphthalene adsorbate. The DFT calculations (10) Bezverkhyy, I.; Ortiz, G.; Chaplais, G.; Marichal, C.; Weber, G.;
reveal the strong interactions due to π−π stacking between Bellat, J.-P. MIL-53(Al) under Reflux in Water: Formation of γ-
adsorbed NAP or IND and the aromatic system of the BDC AlO(OH) Shell and H2BDC Molecules Intercalated into the Pores.
linker as well as a strong H bonding between the NH group of Microporous Mesoporous Mater. 2014, 183, 156−161.
adsorbed IND and the oxygen atom in the COO group of the (11) Gaab, M.; Trukhan, N.; Maurer, S.; Gummaraju, R.; Müller, U.
BDC linker. The highest amount of water adsorbed on A100 The Progression of Al-Based Metal-Organic Frameworks − from
MOF corresponds to (AlOH)2−[BDC]2−(H2O)8. The in-situ Academic Research to Industrial Production and Applications.
time-dependent fluorescence spectroscopic study of kinetics Microporous Mesoporous Mater. 2012, 157, 131−136.
and capacity of adsorption on hydrated A100 in n-heptane (12) Loiseau, T.; Serre, C.; Huguenard, C.; Fink, G.; Taulelle, F.;
yields the preferential adsorption of indole vs naphthalene. Henry, M.; Bataille, T.; Férey, G. A Rationale for the Large Breathing


of the Porous Aluminum Terephthalate (MIL-53) Upon Hydration.
ASSOCIATED CONTENT Chem.Eur. J. 2004, 10, 1373−1382.
(13) Llewellyn, P. L.; Bourrelly, S.; Serre, C.; Filinchuk, Y.; Ferey, G.
*
S Supporting Information
How Hydration Drastically Improves Adsorption Selectivity for CO2
Cartesian coordinates of the MIL-53(Al) model cluster and over CH4 in the Flexible Chromium Terephthalate MIL-53. Angew.
model clusters with one molecule of NAP or IND optimized at Chem., Int. Ed. 2006, 45, 7751−7754.
the M06-2X/LANL2DZ level. This material is available free of (14) Gordon, J.; Kazemian, H.; Rohani, S. Rapid and Efficient
charge via the Internet at http://pubs.acs.org. Crystallization of MIL-53(Fe) by Ultrasound and Microwave

■ AUTHOR INFORMATION
Corresponding Author
Irradiation. Microporous Mesoporous Mater. 2012, 162, 36−43.
(15) Coudert, F.-X.; Ortiz, A. U.; Haigis, V.; Bousquet, D.; Fuchs, A.
H.; Ballandras, A.; Weber, G.; Bezverkhyy, I.; Geoffroy, N.; Bellat, J.-
P.; et al. Water Adsorption in Flexible Gallium-Based MIL-53 Metal−
*Phone: +1-856-225-6282. E-mail: alexsam@camden.rutgers. Organic Framework. J. Phys. Chem. C 2014, 118, 5397−5405.
edu. (16) Chen, L.; Mowat, J. P. S.; Fairen-Jimenez, D.; Morrison, C. A.;
Notes Thompson, S. P.; Wright, P. A.; Düren, T. Elucidating the Breathing of
The authors declare no competing financial interest. the Metal−Organic Framework MIL-53(Sc) with Ab Initio Molecular


Dynamics Simulations and in Situ X-Ray Powder Diffraction
ACKNOWLEDGMENTS Experiments. J. Am. Chem. Soc. 2013, 135, 15763−15773.
(17) Ferey, G.; Latroche, M.; Serre, C.; Millange, F.; Loiseau, T.;
A.S. thanks Rutgers University for his Research Council Grant Percheron-Guegan, A. Hydrogen Adsorption in the Nanoporous
no. 202221 and Dr. Jim Mattheis from Horiba Scientific for Metal-Benzenedicarboxylate M(OH)(O2C-C6H4-CO2) (M = Al3+,
access to the Fluorolog-3 spectrometer. M.M. acknowledges the Cr3+), MIL-53. Chem. Commun. 2003, 2976−2977.
Alabama Supercomputer Center for a generous allocation of (18) Yilmaz, B.; Trukhan, N.; Muller, U. Industrial Outlook on
computer time. Zeolites and Metal Organic Frameworks. Chin. J. Catal. 2012, 33, 3−

■ REFERENCES
(1) Snyder, L. R. Petroleum Nitrogen Compounds and Oxygen
10.
(19) Möllmer, J.; Lange, M.; Möller, A.; Patzschke, C.; Stein, K.;
Lässig, D.; Lincke, J.; Glaser, R.; Krautscheid, H.; Staudt, R. Pure and
Compounds. Acc. Chem. Res. 1970, 3, 290−299. Mixed Gas Adsorption of CH4 and N2 on the Metal-Organic
(2) Singh, D.; Chopra, A.; Patel, M. B.; Sarpal, A. S. A Comparative Framework Basolite ® A100 and a Novel Copper-Based 1,2,4-
Evaluation of Nitrogen Compounds in Petroleum Distillates. Triazolyl Isophthalate MOF. J. Mater. Chem. 2012, 22, 10274−10286.
Chromatographia 2011, 74, 121−126. (20) Samokhvalov, A.; Liu, Y.; Simon, J. D. Characterization of the
(3) Kim, J. H.; Ma, X.; Zhou, A.; Song, C. Ultra-Deep Desulfurization Fe(III)-Binding Site in Sepia Eumelanin by Resonance Raman
and Denitrogenation of Diesel Fuel by Selective Adsorption over Confocal Microspectroscopy. Photochem. Photobiol. 2004, 80, 84−88.
Three Different Adsorbents: A Study on Adsorptive Selectivity and (21) Xu, H.; Lv, M. Developments of Indoles as Anti-HIV-1
Mechanism. Catal. Today 2006, 111, 74−83. Inhibitors. Curr. Pharm. Design 2009, 15, 2120−2148.

2501 DOI: 10.1021/jp510272s


J. Phys. Chem. C 2015, 119, 2491−2502
The Journal of Physical Chemistry C Article

(22) Brancale, A.; Silvestri, R. Indole, a Core Nucleus for Potent (43) Watts, R. J.; Strickler, S. J. Fluorescence and Internal Conversion
Inhibitors of Tubulin Polymerization. Med. Res. Rev. 2007, 27, 209− in Naphthalene Vapor. J. Chem. Phys. 1966, 44, 2423−2426.
238. (44) Braude, E. A.; Nachod, F. C. Determination of Organic Structures
(23) Wang, Z. Y.; Sun, Z. G.; Kong, L. H.; Li, G. Adsorptive Removal by Physical Methods; Academic Press: New York, 1955.
of Nitrogen-Containing Compounds from Fuel by Metal-Organic (45) Zhang, W.; Cao, L.; Wan, L.; Liu, L.; Xu, F. A Photoelectron
Frameworks. J. Energy Chem. 2013, 22, 869−875. Spectroscopy Study on the Interfacial Chemistry and Electronic
(24) Dai, J.; McKee, M.; Samokhvalov, A. Adsorption of Naphthalene Structure of Terephthalic Acid Adsorption on TiO2(110)-(1 × 1)
and Indole on F300 MOF in Liquid Phase by the Complementary Surface. J. Phys. Chem. C 2013, 117, 21351−21358.
Spectroscopic, Kinetic and DFT Studies. J. Porous Mater. 2014, 21, (46) Karabacak, M.; Cinar, M.; Unal, Z.; Kurt, M.; FT-IR, U. V.
709−727. Spectroscopic and DFT Quantum Chemical Study on the Molecular
(25) Maes, M.; Schouteden, S.; Hirai, K.; Furukawa, S.; Kitagawa, S.; Conformation, Vibrational and Electronic Transitions of 2-Amino-
De Vos, D. E. Liquid Phase Separation of Polyaromatics on terephthalic Acid. J. Mol. Struct. 2010, 982, 22−27.
[Cu2(BDC)2(Dabco)]. Langmuir 2011, 27, 9083−9087. (47) Martin, R.; Clarke, G. A. Fluorescence of Benzoic Acid in
(26) Samokhvalov, A.; Tatarchuk, B. J. Review of Experimental Aqueous Acidic Media. J. Phys. Chem. 1978, 82, 81−86.
Characterization of Active Sites and Determination of Molecular (48) Greathouse, J. A.; Ockwig, N. W.; Criscenti, L. J.; Guilinger, T.
Mechanisms of Adsorption, Desorption and Regeneration of the Deep R.; Pohl, P.; Allendorf, M. D. Computational Screening of Metal-
and Ultradeep Desulfurization Sorbents for Liquid Fuels. Catal. Rev.- Organic Frameworks for Large-Molecule Chemical Sensing. Phys.
Sci. Eng. 2010, 52, 381−410. Chem. Chem. Phys. 2010, 12, 12621−12629.
(27) Shan, G.; Liu, H.; Xing, J.; Zhang, G.; Wang, K. Separation of (49) Berlman, I. B. Handbook of Fluorescence Spectra of Aromatic
Molecules; Academic Press: New York, 1971; p 258.
Polycyclic Aromatic Compounds from Model Gasoline by Magnetic
(50) Nijegorodov, N.; Luhanga, P. V. C.; Nkoma, J. S.; Winkoun, D.
Alumina Sorbent Based on pi-Complexation. Ind. Eng. Chem. Res.
P. The Influence of Planarity, Rigidity and Internal Heavy Atom Upon
2004, 43, 758−761.
Fluorescence Parameters and the Intersystem Crossing Rate Constant
(28) Anand, R.; Borghi, F.; Manoli, F.; Manet, I.; Agostoni, V.;
in Molecules with the Biphenyl Basis. Spectrochim. Acta, Part A 2006,
Reschiglian, P.; Gref, R.; Monti, S. Host−Guest Interactions in 64, 1−5.
Fe(III)-Trimesate MOF Nanoparticles Loaded with Doxorubicin. J. (51) Borsarelli, C. D.; Bertolotti, S. G.; Previtali, C. M. Exciplex-Type
Phys. Chem. B 2014, 118, 8532−8539. Behavior and Partition of 3-Substituted Indole Derivatives in Reverse
(29) Allendorf, M. D.; Bauer, C. A.; Bhakta, R. K.; Houk, R. J. T. Micelles Made with Benzylhexadecyldimethylammonium Chloride,
Luminescent Metal-Organic Frameworks. Chem. Soc. Rev. 2009, 38, Water and Benzene. J. Photochem. Photobiol. 2001, 73, 97−104.
1330−1352. (52) Zhao, Y.; Schultz, N. E.; Truhlar, D. G. Design of Density
(30) Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Luminescent Functional Functionals by Combining the Method of Constraint Satisfaction with
Metal−Organic Frameworks. Chem. Rev. 2012, 112, 1126−1162. Parametrization for Thermochemistry, Thermochemical Kinetics, and
(31) Yang, C.-X.; Ren, H.-B.; Yan, X.-P. Fluorescent Metal−Organic Noncovalent Interactions. J. Chem. Theory Comput. 2006, 2, 364−382.
Framework MIL-53(Al) for Highly Selective and Sensitive Detection (53) Zhao, Y.; Truhlar, D. G. A New Local Density Functional for
of Fe3+ in Aqueous Solution. Anal. Chem. 2013, 85, 7441−7446. Main-Group Thermochemistry, Transition Metal Bonding, Thermo-
(32) Deria, P.; Mondloch, J. E.; Karagiaridi, O.; Bury, W.; Hupp, J. chemical Kinetics, and Noncovalent Interactions. J. Chem. Phys. 2006,
T.; Farha, O. K. Beyond Post-Synthesis Modification: Evolution of 125, 194101.
Metal-Organic Frameworks Via Building Block Replacement. Chem. (54) Zhao, Y.; Truhlar, D. G. Comparative DFT Study of Van Der
Soc. Rev. 2014, 43, 5896−5912. Waals Complexes: Rare-Gas Dimers, Alkaline-Earth Dimers, Zinc
(33) Demir, M.; McKee, M.; Samokhvalov, A. Interactions of Dimer, and Zinc-Rare-Gas Dimers. J. Phys. Chem. A 2006, 110, 5121−
Thiophenes with C300 Basolite MOF in Solution by the Temperature- 5129.
Programmed Adsorption and Desorption, Spectroscopy and Simu- (55) Zhao, Y.; Truhlar, D. G. Density Functional for Spectroscopy:
lations. Adsorption 2014, 20, 829−842. No Long-Range Self-Interaction Error, Good Performance for
(34) Nicholson, T. M.; Bhatia, S. K. Electrostatically Mediated Rydberg and Charge-Transfer States, and Better Performance on
Specific Adsorption of Small Molecules in Metallo-Organic Frame- Average Than B3LYP for Ground States. J. Phys. Chem. A 2006, 110,
works. J. Phys. Chem. B 2006, 110, 24834−24836. 13126−13130.
(35) Dhage, P.; Samokhvalov, A.; McKee, M. L.; Duin, E. C.; (56) Zhao, Y.; Truhlar, D. G. The M06 Suite of Density Functionals
Tatarchuk, B. J. Reactive Adsorption of Hydrogen Sulfide by for Main Group Thermochemistry, Thermochemical Kinetics, Non-
Promoted Sorbents Cu-ZnO/SiO2: Active Sites by Experiment and covalent Interactions, Excited States, and Transition Elements: Two
Simulation. Surf. Interface Anal. 2013, 45, 865−872. New Functionals and Systematic Testing of Four M06-Class
(36) Blanco-Brieva, G.; Campos-Martin, J. M.; Al-Zahrani, S. M.; Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120,
Fierro, J. L. G. Effectiveness of Metal-Organic Frameworks for 215−241.
Removal of Refractory Organo-Sulfur Compound Present in Liquid (57) Mackie, I. D.; DiLabio, G. A. Accurate Dispersion Interactions
Fuels. Fuel 2011, 90, 190−197. from Standard Density-Functional Theory Methods with Small Basis
(37) Lakowicz, J. R. Principles of Fluorescence Spectroscopy; Springer: Sets. Phys. Chem. Chem. Phys. 2010, 12, 6092−6098.
New York, 2006. (58) van Duijneveldt, F. B.; van Duijneveldt-van de Rijdt, J. G. C. M.;
(38) Cramer, C. J. Essentials of Computational Chemistry: Theories and van Lenthe, J. H. State of the Art in Counterpoise Theory. Chem. Rev.
Models; Wiley: New York, 2004. 1994, 94, 1873−1885.
(39) Koch, W.; Holthausen, M. C. A Chemist’s Guide to Density (59) Sinnokrot, M. O.; Sherrill, C. D. Highly Accurate Coupled
Functional Theory, 2nd ed.; Wiley-VCH: Weinheim, Germany, 2001. Cluster Potential Energy Curves for the Benzene Dimer: Sandwich, T-
(40) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Shaped, and Parallel-Displaced Configurations. J. Phys. Chem. A 2004,
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, 108, 10200−10207.
(60) Alvarez, S. A Cartography of the Van Der Waals Territories.
B.; Petersson, G. A. et al. Gaussian 09, Revision D. 01; Gaussian:
Dalton Trans. 2013, 42, 8617−8636.
Wallingford, CT, 2009.
(61) Janiak, C. A Critical Account on π-π Stacking in Metal
(41) Kasha, M. Characterization of Electronic Transitions in
Complexes with Aromatic Nitrogen-Containing Ligands. J. Chem. Soc.,
Complex Molecules. Discuss. Faraday Soc. 1950, 9, 14−19.
Dalton Trans. 2000, 3885−3896.
(42) Privalova, A. I.; Morozova, J. P.; Kashapova, E. R.; Artyukhov, V.
J. Spectral and Luminescent Properties of 1-Substituted Naphthalenes.
J. Appl. Spectrosc. 2011, 78, 309−317.

2502 DOI: 10.1021/jp510272s


J. Phys. Chem. C 2015, 119, 2491−2502

You might also like