Open Access Article
Lubna Alrawashdeh
*a,
Shrouq Almarabeha,
Khaleel I. Assaf
*b,
Anthony I. Day
c,
Lynne Wallace
c and
Suhair A. Bani-Atta
d
aDepartment of Chemistry, Faculty of Science, The Hashemite University, P.O. Box 330127, Zarqa 13133, Jordan. E-mail: lubna.reyad@hu.edu.jo; shrouqalmarabeh@gmail.com
bDepartment of Chemistry, Faculty of Science, Al-Balqa Applied University, Al-Salt 19117, Jordan. E-mail: khaleel.assaf@bau.edu.jo
cSchool of Physical, Environmental and Mathematical Sciences, UNSW Australia, Australian Defence Force Academy, Canberra, Australia. E-mail: a.day@unsw.edu.au; lynne.wallace@unsw.edu.au
dAnalytical Chemistry Research Laboratory, Department of Chemistry, Faculty of Science, University of Tabuk, Tabuk 71491, Saudi Arabia. E-mail: s_bantatta@ut.edu.sa
First published on 20th August 2025
The supramolecular host–guest interaction between heteroleptic iridium(III) complexes and cucurbit[10]uril (Q[10]) in an aqueous medium was investigated in this work. Both studied iridium complexes, [Ir(ppy)2(bpy-(CHO)2)]+ (complex 1) and [Ir(ppy)2(bpy-(COOH)2)]+ (complex 2), possessed two phenylpyridine ligands and a single R-bipyridine ligand. The formation of the encapsulated species (Q[10]·1 and Q[10]·2) was demonstrated by 1H NMR and luminescence studies. A significant improvement was observed in the luminescence properties of both iridium complexes (emission intensity, quantum yield and lifetime) upon the addition of Q[10] in an aqueous medium. Results suggested a major effect of the hydrophobic cavity in the destabilization of the 3MLCT state (lowest excited state) of the iridium complexes. The binding study of both complexes with Q[10] revealed the formation of 1
:
1 and 1
:
2 host–guest species, with the binary complex dominating the emission behavior. The equilibrium between the emitting species was significantly influenced by the temperature, wherein the 1
:
1 inclusion complex was less favorable at elevated temperatures. The effect of pH on the emission profiles of the free and encapsulated iridium complexes was also investigated in this study. Density functional theory (DFT) calculations showed that introducing different substituent groups (CHO and COOH) on the bpy ligand had a negligible effect on the orientation of these complexes within the Q[10] cavity.
Previous studies have shown that several cyclometalated iridium(III) complexes have serious limitations as sensors for biological analytes because of their poor water solubility and weak emission intensity in aqueous media.12,13 Based on this, various studies have been undertaken to develop water-soluble luminescent cyclometalated iridium(III) complexes by changing the attached substituent groups (more hydrophilic) on the ligands.14,15 Recently, our group proposed a novel method to overcome the limitations by changing the environment around iridium complexes. This was obtained by applying host–guest chemistry to selected metal complexes using cucurbit[n]urils (Q[n]) as host molecules. The aqueous solubility and emission properties were significantly improved after encapsulating the iridium complexes within the hydrophobic cavity of Q[n].16
Q[n] are a prominent family of water-soluble macrocyclic host molecules that have garnered considerable attention in the field of supramolecular host–guest chemistry.17,18 They consist of n glycoluril units (n = 5–8, 10) connected by methylene bridges and have portal dimensions of 2.4–11.0 Å.17 Q[n] are pumpkin-shaped, highly symmetrical macrocycles with a hydrophobic cavity suitable for accommodating neutral/non-polar guest molecules,19 and carbonyl rims offer binding sites for non-covalent interactions (such as hydrogen bond and ion–dipole interactions).20 Q[n] have advantages over other host molecules such as cyclodextrins and calixarenes because of their high selectivity21 and affinity20 toward guest molecules. Q[n] have been used in many applications, including sensing,22 separation,23 molecular recognition,24 drug delivery and biomedical systems.25–27
Q[10] is the member that possesses the largest accessible cavity (870 Å3 and portal diameter = 11 Å) among the commonly used members of the family (Q[5]–Q[10]).28 Recently, larger Q[n] homologues (n = 13–15) have also been isolated; however, their “twisted” conformations limit the available space in the cavity.29,30 Thus, Q[10] can encapsulate large-sized guests or two small guest molecules to form ternary complexes.31,32 Moreover, special attention has been given to encapsulating large transition metal complexes for various applications. For example, Q[10] has been used as a drug vehicle for di-ruthenium and di-platinum complexes,33 while various encapsulated mono-ruthenium complexes within the Q[10] cavity have been investigated to study the formation of supramolecular photocatalyst systems.34 Wallace and coworkers showed that large tris–chelate transition metals (Ru and Ir) can also be accommodated within the Q[10] host molecule.35
Many studies have demonstrated that Q[n] can be used to tune the photophysical properties of luminescent dyes;36 based on this point, we previously applied host–guest chemistry to some large luminescent cyclometalated iridium(III) complexes. Sizable enhancement in the luminescence properties and aqueous solubility of these complexes was obtained upon encapsulation inside Q[10].16,37 Wallace showed that this effect is not universal for all iridium complexes and depends mainly on the electronic structure of the guest.35 Herein, we extend our previous work to consider more polypyridyl iridium(III) complexes with different electronic properties. These complexes have different substituent groups on the bpy ligand ([Ir(ppy)2(bpy-R)]n+) and are commonly used in sensing applications. In addition, the effect of changing the position of the substituent groups on the binding mode, luminescence and solubility properties of these complexes will be explored. Aldehyde (–CHO) and carboxylic acid (–COOH) moieties are used as substituent (R) groups in this study (Fig. 1) because of their benefits as building blocks in the synthesis of sensors.38,39 Furthermore, the ionizable nature of the carboxylic acid group enables pH-dependent behavior, which is valuable for designing pH-sensitive sensors and systems that respond to biological environments.
1H NMR spectra were recorded using a Bruker AVANCE-III 400 MHz NanoBay FT-NMR spectrometer at 25 °C, and tetramethylsilane (TMS) was used as a reference. Luminescence spectra were recorded using a Jasco spectrofluorometer (FP-6500). This spectrofluorometer was also used to study the temperature dependence of the emission spectra, and a circulating water bath (Jeio Tech RW-1025G) was used for controlling the temperature. Samples were excited at 350 nm using quartz cuvettes (1 cm path length). Luminescence lifetimes were determined using a spectrofluorometer from Edinburgh Analytical Instruments FL-900S (Germany) equipped with time-correlated single photon counting. Sample decay was fitted using exponential tail fit analysis. The value of the χ2 was used to assess the quality of the fitted result (which should be close to 1.00 for a successful fit). UV-visible spectra were recorded on a Cary 100 UV-Visible spectrophotometer. Acetate buffer (0.05 M, pH = 4.7) was used to calculate the absorption coefficients (ε) of the iridium complexes.
Quantum yields were measured in acetate buffer at an excitation wavelength of 350 nm. The optically dilute method with single-point measurements was used for this purpose.42 A deaerated [Ru(bpy)3]2+ aqueous solution was used as a reference (λex = 350 nm, ΦR = 0.042). In this method, the quantum yield was found based on eqn (1):
| Φx = ΦR [IX/IR] [AR/Ax] [ηx2/ηR2] | (1) |
Φx and ΦR are the quantum yields of the unknown and reference, respectively; Ix and IR are the integrated emission intensities of our sample and reference, respectively; Ax and AR are the absorbance of our sample and reference at the excitation wavelength, respectively; and ηx and ηR are the refractive indices of the solvents. The refractive index of acetate buffer (solvent) is 1.34.43
:
1 and 1
:
2 complexes, respectively, are given below:
![]() | (2) |
![]() | (3) |
![]() | (4) |
:
1 inclusion complexes (eqn (2)) and by plotting 1/A − A0 vs. 1/[Q[10]]2 for 1
:
2 inclusion complexes (eqn (3)). The slope obtained from the Benesi–Hildebrand plot (using eqn (4)) was used to calculate the binding constant Ka.
As further support for the encapsulation of complex 1 in the cavity of Q[10], complex 1 and Q[10] were combined in D2O; however, it was found that the low solubility of 1 retarded the formation of Q[10]·1. A comparative spectrum of 1 alone was only feasible if 1 was first dissolved in ACN-d3 and then diluted with D2O to a final concentration of 5% ACN-d3/D2O (Fig. 2B). It was also observed that the aldehyde resonance at 10.18 ppm was significantly reduced, indicating partial hydration. Combining 1 and Q[10] in D2O required extended sonication to form a soluble form of Q[10]·1, and the progressive formation was observed by NMR as the free Q[10] doublet resonance at 5.82 ppm decreased and the resonances for the associated guest 1 increased (Fig. 2). The new association product Q[10]·1 appears as a dynamic multiplex with two new upfield doublets occurring for Q[10] at 5.7 and 5.9 ppm, which suggests two association forms in an approximate ratio of 1
:
2, respectively. The broadness of the resonances suggests dynamics of different orientations of the guest, possibly between two forms and/or intermediate exchange. The remaining Q[10] resonances (normally a singlet and doublet) fall near the original chemical shift of free Q[10] as broad peaks at 4.2 and 5.5 ppm. Two pertinent guest proton resonances indicate that the phenyl pyridine ligands are located in the cavity and that the bipyridine ligand is near the portal. This is indicated by an upfield shift to 6.2 ppm for the proton doublet α to the Ir–C bond compared to the original shift of 6.4 ppm38,48 and the downfield shift for the CHO resonance, from 10.2 to 10.3 ppm.
In the case of complex 2, in the pD range of 2–9, the solubility of Q[10]·2 is too low for resonances to be visible by 1H NMR spectroscopy. Adding 2 as a D2O solution or as a solid to a solution of Q[10] in equimolar quantities results in a yellow precipitate with no resonances observable from the supernatant. When the pD of a suspension of the precipitate in D2O is adjusted with the addition of NaOD, resonances begin to appear for the free complex 2 and free Q[10] at pD = 8.5, suggesting that the dicarboxylate ion was released. At lower pD < 3.7, 2 must be associated with Q[10] in the protonated form, and above this value, only one carboxylic group would be deprotonated to form a zwitterion complex, with a corresponding increase in the emission. Increasing the pD > 9 results in complete ionisation, and dissociation begins, consistent with the behaviour of carboxylates from Q[7 or 8], as previously reported.49–57 The 1H NMR spectrum at pD 9 showed the release of 2 in an excess relative to the observable Q[10] with broader resonances of some of the protons of 2 and the downfield doublet of Q[10], indicating intermediate exchange within the NMR time scale (Fig. 3).
![]() | ||
| Fig. 3 1H NMR spectra (D2O, 500 MHz) of (A) complex 2 at pD 10 and (B) suspension of Q[10]·2 in D2O; pD adjusted with NaOD (0.2 M) to 9. | ||
The luminescent nature of the yellow precipitate obtained in the NMR study was determined in the solid form (λem = 587 nm, see the SI). For comparison, equimolar quantities of complex 2 and Q[10] were ground together to a fine homogeneous powder with λem = 608 nm. The 2-fold increase in the intensity and the blue shift of 21 nm indicate that the Q[10]·2 complex is formed almost quantitatively as the yellow precipitate from aqueous solutions (Fig. S12).
| λabs (nm) | λem (nm) | Lifetime (ns) | Quantum yield | Enhancement factor | Ka (M−1) | |
|---|---|---|---|---|---|---|
| 1 | 257 max, 309 sh, 381 sh | 575 | 22 | 0.014 ± 0.002 | — | — |
| Q[8]·1 | — | 552 | 279 | — | 6 times | |
| Q[10]·1 | 260 max, 312 sh, 388 sh | 540 | 499 | 0.142 ± 0.016 | 40 times | 1.02 × 106 |
| 93 | ||||||
| 2 | 254 max, 324 sh, 381 sh | 578 | 13 | 0.023 ± 0.004 | — | — |
| Q[8]·2 | — | 557 | 27 | — | 2 times | — |
| Q[10]·2 | 261 max, 325 sh, 388 sh | 553 | 309 | 0.060 ± 0.002 | 20 times | 2.31 × 105 |
| 98 |
The absorption spectra of 1 and 2 before and after adding excess Q[10] are shown in Fig. 4. Both complexes display high-energy maxima at 257 and 254 nm, with molar absorptivity (ε) of 4.0 × 104 and 5.0 × 104 M−1cm−1, respectively. These bands can be assigned to the overlapping 1LC (π → π*) transitions of ppy and R-bpy ligands.58 Fairly intense shoulders appear at 309 and 324 nm, respectively, which are also attributed to the 1LC (possibly ppy) transition.59 The lower-energy bands (381 and 382 nm for complex 1 and complex 2, respectively) are assigned to 1LLCT and 1MLCT (from Ir dπ–π*bpy), respectively.38,39,59
![]() | ||
| Fig. 4 Absorption spectra of free 1 and the Q[10]·1 inclusion complex (A) and free 2 and the Q[10]·2 inclusion complex (B) measured in a buffer solution (pH 4.7) at ambient temperature. | ||
Adding Q[10] to the buffer solutions of both complexes (1 and 2) has almost similar effects. A minor red shift is noticed with a decrease in the intensity of most absorption bands, especially the high-energy bands (λmax = 257 → 260 nm for 1, and λmax = 254 → 261 nm for 2). The bathochromic shift can be related to the alteration in the polarity of the medium (high polar in the buffer to less polar in the host cavity). On the contrary, researchers reported that the absorptivity coefficients of some organic chromophores decrease after encapsulation within Q[n] because the cavity has lower polarizability.60 The larger shifts of the high-energy bands (1LC), which probably correspond to the ppy ligand,59 compared to the lowest bands (1MLCT and 1LLCT), may be related to the accommodation mode of these complexes inside Q[10] (see the DFT study). Assuming that the ppy ligand sits inside the cavity, it is significantly affected by adding Q[10] compared to the bpy ligand, which is located on the portal. This effect is consistent with that observed in our previous study for some iridium complexes16 and another previous study for polypyridyl ruthenium complexes.34 TD-DFT calculations were performed to compute the UV-visible spectra of 1 and 2 and their complexes with Q[10]. Despite the slight changes in HOMO–LUMO energy gaps upon the formation of host–guest complexes, the calculated UV-visible spectra nicely reproduce the observed bathochromic shift in λmax upon complexation (see Fig. S8 and Tables S1–S3). The isosurface plots of the HOMO and LUMO wave functions of free 1 and 2 and their Q[10] complexes are given in the SI (Fig. S9 and S10).
In the luminescence study, both 1 and 2 showed broad and weak emission profiles in aqueous buffer solutions. Upon adding Q[10], the emission spectra of Ir-complexes showed a blue shift with a clear enhancement in the emission intensities. The enhancement factors are ∼40 times for complex 1 and ∼20 times for complex 2 (Fig. 5). This suggests the formation of the encapsulated species (Q[10]·1 and Q[10]·2 complexes), which also confirms that the size of the host (portal diameter = ∼10.3 Å) is suitable to encapsulate the iridium complexes (∼9.6 Å) within its cavity, providing protection to the guest from quenching factors such as solvents and photodegradation.
![]() | ||
| Fig. 5 Luminescence spectra of free 1, Q[8]·1 and Q[10]·1 complexes (A) and free 2, Q[8]·2 and Q[10]·2 (B), measured in a buffer solution (pH 4.7) at room temperature. | ||
The quantum yield and lifetime of both complexes are also enhanced after adding Q[10] (Table 1). A biexponential model was required to fit the decay profile obtained from the lifetime measurements, which reveals that the encapsulated systems of both complexes have two different emitting species. For Q[10]·1, the long-lived species has a lifetime of 499 ns, while the short-lived species has a lifetime of 93 ns; while in the case of Q[10]·2, the lifetime values are 309 and 98 ns for the long-lived and short-lived components, respectively. The short-lived component has a longer lifetime than the free guests (22 ns for complex 1 and 13 ns for complex 2), suggesting that this system is not a simple 1
:
1 host–guest species but has multiple modes of supramolecular interactions (see the binding study below). This behavior is similar to that noticed in our previous study on other iridium complexes.16,61
The enhancement in the luminescence properties of both complexes upon the addition of Q[10] (Table 1) can be attributed to the restricted motion of the iridium guests within the Q[10] cavity, which likely alters the balance between radiative and non-radiative decay pathways by suppressing non-radiative processes. A similar effect was observed on the emission properties of some free iridium complexes upon cooling to 77 K (rigid environment).62 The blue shift that emission bands display upon adding Q[10] to both complexes can be related to the low polarity of the Q[10] cavity, compared to the aqueous medium. This observation was verified by examining the effect of the solvent polarity on the emission band of unbound 1 and 2. In both cases, the emission peaks shift to higher energy wavelengths (blue shift) upon moving from a highly polar medium (buffer) to a less polar one (CH3CN) (Fig. S5). This blue shift is mainly related to the destabilization of the 3MLCT excited state in less polar media and will locate close to the 3LC(ppy) state but not above it, which explains the structureless emission profiles of both iridium complexes in the less polar solvent. This result implies that the lowest excited state character is not strongly affected by the solvent polarity. However, the previously studied iridium complexes that have substituent groups on ppy ligands showed structured emission profiles upon encapsulation within Q[10], which suggest that the destabilization of 3MLCT is much stronger in those cases, and the energy of the 3MLCT(bpy) state exceeds the energy of the 3LC state; thus, the emission bands have mainly a 3LC character.16
The sensitivity of the luminescence phenomenon of these iridium complexes (1 and 2) was also demonstrated by their supramolecular complexation and Q[8], which has a smaller size (portal diameter = 7.3 Å) compared to Q[10]. Exclusion binding is presumed between Q[8] and iridium complexes (9.6 Å diameter). Adding Q[8] to 1 induces a 23 nm blue shift with a 6-fold enhancement in the emission intensity of 1, while a 21 nm blue shift with an enhancement of 2-fold is observed in the case of 2. The lifetimes of 1 and 2 are also enhanced upon adding Q[8], and a single exponential is used in the fitting of the emission lifetime, indicating the formation of one emitting species. The structures of the emission bands of both complexes do not change upon adding Q[8], suggesting that the emission bands of host–guest portal binding complexes have the 3MLCT character in both complexes. The blue shift and enhancement that have been displayed in both cases upon portal binding with Q[8] (Fig. 5) can be related to the slight destabilization of the lowest excited state (3MLCT), which is obtained by reducing collisions and the partial protection from the solvent. However, a quenching effect is observed for Q[10] portal binding with 3MLCT emitters, based on the behavior observed for a Ru(II) polypyridyl system.33
:
1 and 1
:
2), which is in agreement with the non-exponential emission decays noted in the presence of excess Q[10] (based on lifetime measurements). This result is similar to that obtained for previously studied iridium complexes.61 The modified Benesi–Hildebrand equation was used to calculate the “apparent” binding constant (Ka) of both complexes with the Q[10] host molecule, assuming a 1
:
1 host–guest stoichiometric ratio. The slope and intercept of the double reciprocal plot of 1/(I − I0) vs. 1/[Q[10]] were used to get the Ka values. For a higher order of complexation complexes, this fitting would be nonlinear. The obtained Ka values are 1.02 × 106 and 2.31 × 105 M−1 for complexes 1 and 2, respectively (Fig. 7).
![]() | ||
| Fig. 6 Luminescence titration spectra of 1 (5.9 × 10−6 M) (A) and 2 (7.6 × 10−7 M) (B) in the presence of different concentrations of Q[10], in a buffer solution (pH 4.7) at ambient temperature. | ||
The binding constants of iridium complexes with Q[10] were also estimated using the UV-visible spectroscopic technique. Absorption titration experiments were carried out for each complex with Q[10] in a buffer solution. A decrease in the absorbance values of each iridium complex with a small red shift resulted after Q[10] addition (Fig. S6). The modified Benesi–Hildebrand equation was also used to determine the binding constants Ka of both complexes with Q[10], assuming a 1
:
1 host–guest stoichiometric ratio. The estimated binding constants are found to be 2.6 × 106 M−1 for complex 1 and 6.3 × 105 M−1 for complex 2 (Fig. S7), which are in good agreement with the values obtained by luminescence titrations. The lower Ka value of complex 2 compared to complex 1 may be related to the repulsion force between the carboxylate moiety COO− and the portal carbonyl groups of the host molecule49,51 (see the pH study below).
:
1 host–guest) and shifts the equilibrium to 1
:
1 or 1
:
2 portal-bound forms or the free species (Fig. 9). In the last status, a clear change in the emission peaks is unlikely to be seen, as these emitting species are very weak compared to the strong 1
:
1 encapsulated species. The emission profile can be quenched by increasing the temperature as a result of the activation of non-radiative pathways (collisions and vibrations). Applying the same study to both free complex 1 and complex 2 gave an insignificant change in the emission intensity. In general, emitting species that possess long lifetimes are usually more sensitive to the temperature than those that possess short lifetimes. Thus, the impact of the temperature on the emission profile is probably a result of its influence on the binding equilibrium and non-radiative pathways together.
![]() | ||
| Fig. 8 Luminescence spectra of Q[10]·1 (A) and Q[10]·2 (B) in a buffer solution (pH 4.7) at different temperature (15–55 °C). | ||
![]() | ||
| Fig. 9 Schematic illustrating the effect of temperature on the binding equilibrium between emitting species. | ||
![]() | ||
| Fig. 10 Emission intensity plot for (A) unbound 2 at 584 nm and (B) encapsulated species Q[10]·2 at 553 nm as a function of pH at ambient temperature. | ||
The optimized structures of the Q[10]-complexes are shown in Fig. 12. The substituted iridium cyclometalated complexes show similar binding modes, in which they all fit well inside the cavity of Q[10]. As expected, the hydrophobic part is deeply encapsulated inside the inner cavity, while the hydrophilic groups (–CHO, –COOH, and –COO−) are excluded and more exposed to water. In the case of the protonated carboxylated iridium complex (Ir–COOH), the complex is further stabilized by hydrogen bonding with the carbonyl portal of Q[10] with a hydrogen bond distance of 1.7 Å. The calculated binding energy reveals higher affinity for Q[10]·2 (R = –CO2H), followed by Q[10]·2 (R = –CHO). It should be noted that the weak binding affinity of the dicarboxylated complex 2 is in accordance with its dissociation from the cavity, as seen in 1H NMR experiments (Fig. 3).
| This journal is © The Royal Society of Chemistry 2025 |