Triboluminescence and Metal Phosphor for Organic Light-Emitting

Nov 21, 2016 - We report the utilization of both pyrid-2-yl-imidazolylidene and dianionic bipz chelates as constituents in syntheses of a new series o...
1 downloads 15 Views 1MB Size
Subscriber access provided by UNIV OF WESTERN ONTARIO

Article

Triboluminescence and Metal Phosphor for Organic Lightemitting Diodes: Functional Pt(II) Complexes with both 2-Pyridylimidazol-2-ylidene and Bipyrazolate Chelates Che Wei Hsu, Kiet Tuong Ly, Wei-Kai Lee, Chung-Chih Wu, Lai-Chin Wu, JeyJau Lee, Tzu-Chieh Lin, Shih-Hung Liu, Pi-Tai Chou, Gene-Hsiang Lee, and Yun Chi ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.6b12707 • Publication Date (Web): 21 Nov 2016 Downloaded from http://pubs.acs.org on November 26, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Triboluminescence and Metal Phosphor for Organic Light-emitting Diodes: Functional Pt(II) Complexes with both 2-Pyridylimidazol-2-ylidene and Bipyrazolate Chelates

Che-Wei Hsu,a Kiet Tuong Ly,a Wei-Kai Lee,b Chung-Chih Wu,b,* Lai-Chin Wu,c Jey-Jau Leec Tzu-Chieh Lin,d Shih-Hung Liu,d Pi-Tai Choud,* Gene-Hsiang Lee,d and Yun Chi,a,*

a

Department of Chemistry, National Tsing Hua University, Hsinchu 30013, Taiwan;

E-mail: [email protected] b

Graduate Institute of Electronics Engineering and Department of Electrical

Engineering,

National

Taiwan

University,

Taipei

10617,

Taiwan;

E-mail:

[email protected] c

National Synchrotron Radiation Research Center, Hsinchu 30076, Taiwan

d

Department of Chemistry, National Taiwan University, Taipei 10617, Taiwan; E-mail:

[email protected]

Keywords: Platinum, organic light emitting diode, triboluminescence, mechanoluminescence, carbene, imidazolium, N-donor, pyrazolate

Abstract We report the utilization of both pyrid-2-yl-imidazolylidene and dianionic bipz chelates as constituents in syntheses of a new series of charge-neutral Pt(II) complexes 1‒4, among which complex 4 revealed remarkable triboluminescence, i.e. generation of photoemission upon grinding or cracking of the solid sample. The triboluminescence is found to be sensitive to the subtle changes of the associated substituents of pyrid-2-yl-imidazolylidene chelate, as verified by the disappearance of the triboluminescence for complexes 1‒3. Alternatively, the well-ordered solid packing of 3, as indicated by the grazing incidence X-ray scattering experiment, serves as an ideal emitter for the fabrication of highly efficient OLEDs, rendering high ‒1‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 35

external quantum efficienciy (25.9%) and luminesce efficiency (90 cd⋅A‒1) at the practical brightness of 100 cd∙m‒2. The rather low roll-off in efficiency (24.4%, 85 cd⋅A‒1 at high brightness of 1000 cd∙m‒2) is attributed to the short excited-state lifetime of 3 (∼800 ns) in the solid state, which in turn is associated with the MMLCT transition character. _____________________________________________________________________

Introduction There has been an extensive interest in developing transition metal-based materials for the organic light emitting diodes (OLEDs).1-6 These transition metal complexes are distinct from the organic fluorophores due to the high efficiency and relatively long emission lifetime; the latter is a key characteristic of the so-called phosphorescence

derived

from

the

triplet

excited

states.

The

bright

phosphorescence at room temperature is attributed to the heavy atom effect that induces strong spin-orbit coupling, facilitating the fast intersystem crossing between the singlet and triplet states.7-10 As a result, a large number of emissive Ru(II), Os(II), Ir(III) and Pt(II) metal complexes have been synthesized and applied to the fabrication of OLEDs and other optoelectronic devices.11-13 It has been well established that their fundamental properties are strongly influenced by the constituent chelates; particularly the donor ability, π-conjugation and steric encumbrance have proved to be crucial as part of the complex design strategy. Moreover, the choice of ligand can also permit fine-tuning of the emission color, as well as improving solubility, stability and luminance efficiency, all of which are exceedingly important for the material processing.2,

13

As for the charge-neutral

bidentate ligands, to date, 2,2’-bipyridine (bpy)14-16 and dicarbene17-19 are the two most employed designs in assembly of square-planar Pt(II) metal complexes (cf. Chart 1).

‒2‒ ACS Paragon Plus Environment

Page 3 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Chart 1: Chelating ligands with varying electronic character.

Notably, the charge-neutral Pt(II) complexes bearing either bpy or dicarbene chelate will be produced only if the complexes incorporated two anionic ligands such as halides,20-21 cyanides22-23 or acetylides.24-26 However, the progress in preparation of respective Pt(II) complexes using dianionic chelates has been hampered with limited success due to the lack of suitable designs. Only recently, the dianionic chelates

such

as

2,2’-biphenyl,27

benzene-1,2-dithiolate,28

5,5'-di(trifluoro-

methyl)-3,3’-bi-pyrazolate (bipz),29 5,5'-di(trifluoromethyl)-3,3’bi-triazolate (bitz),29 5,5'-(1-methylethylidene)bis(3-trifluoromethyl-1H-pyrazolate) 3-trifluoromethyl5(4-(trifluoromethylphenyl)-1H-pyrazolate

(mepz),30 (phpz)31

and were

successfully utilized in the syntheses of d8-Pt(II) metal complexes and the Ru(II), Os(II) and Ir(III) based octahedral complexes with a d6-electronic configuration. In these metal complexes, computational analyses suggested that the occupied frontier orbitals primarily contain contributions from both central metal atom and dianionic chelate, whereas the unoccupied frontier orbitals reside on the neutral chelates.27 Accordingly, the dominant transitions are best described as metal‒ligand-to-ligand charge transfers (MLLCT),5 which is distinctive from the typical transition character of combined ligand-centered (LC) ππ* and metal-to-ligand charge transfer (MLCT) transitions in typical transition-metal complexes.

‒3‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 35

In an aim to extend the synthetic scope of functional Pt(II) metal complexes, our attention were drawing to the pro-chelates, namely: 3-alkyl-1-(pyrid-2-yl)1H-imidazolylium, which can be converted to the neutral chelates bearing both pyridyl and imidazolylidene donors, displaying a subtle balance between the electronic and steric properties. In fact, relevant Ru(II), Ir(III) and Pt(II) metal complexes comprising similar carbene chelate have been extensively studied for catalysis applications32-34 and in studying luminescent properties;35-38 the latter may show high potential in organic optoelectronics due to the much destabilized π*-orbitals. In this study, we report the utilization of both pyrid-2-yl-imidazolylidene and dianionic bipz chelates as constituents in syntheses of the corresponding Pt(II) complexes, among which one has showed novel triboluminescence, i.e. generation of photoemission upon gentle grinding or cracking of the solid sample.39-42 Furthermore, upon changing the associated substituents, the triboluminescence vanished; alternatively they serve as the ideal emitters for the fabrication of highly efficient OLEDs.

Results and Discussion Synthesis

and

structural

characterization.

Four

3-alkyl-1-(pyrid-2-yl)-1H-imidazolylium derivatives, i.e. (mpyim)HPF6, R’ = H, R” = Me; (ipyim)HPF6, R’ = H, R” = iPr; (mtpyim)HPF6, R’ = tBu, R” = Me; and (itpyim)HPF6, R’ = t

Bu, R” = iPr, were prepared and isolated as hexafluorophosphate salts according to

the literature method.43-44 The 5,5'-di(trifluoromethyl)-3,3'-bipyrazole chelate (bipzH2) was

alternatively

prepared

using

Claisen

condensation

employing

ethyl

trifluoroacetate and 2,3-butanedione, followed by heating with hydrazine in refluxing ethanol solution.45 The parent imidazolylium derivatives, i.e. (mpyim)HPF6 and (ipyim)HPF6 are relatively insoluble in nonpolar solvents, such that their 1H NMR ‒4‒ ACS Paragon Plus Environment

Page 5 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

spectra were recorded in d6-acetone, in contrast to the t-butyl substituted derivatives (mtpyim)HPF6 and (itpyim)HPF6, for which the NMR spectra are instead recorded in CDCl3 solution. All these pre-chelates exhibited a signal in the regime δ 9.88 ∼ 9.35 assigned to the acidic imidazolylium C-H fragment, which disappeared upon conversion to the imidazolylidene chelate after formation of the demanded Pt(II) metal complexes.46 Pt(II) complexes with formula [Pt(N^C)(bipz)], N^C = pyridyl imidazolylidene chelates such as mpyim, ipyim, mtpyim and itpyim, were typically obtained by treatment of imidazolium derivative with Pt(DMSO)2Cl2 in presence of NaHCO3 at 100oC for 12 h. Without isolation, the dianionic chelate (bipz)H2 were added and heated at 100 °C for another 6 h to afford the Pt(II) metal complexes, with typical synthetic yield spanning 52 ‒ 68%. This synthetic approach is different from that adopted for the analogous Pt(II)-bisacetylene derivatives [(pyim)Pt(CCAr)2] Ar = aryl group, which were alternatively obtained by replacement of the 1,5-cyclooctadiene (COD) chelate in [(COD)Pt(CCAr)2] with pyridyl imidazolylidene Ag(I) iodide in CH2Cl2 solution at RT.47

These Pt(II) complexes 1−4 were initially characterized using

1

H NMR

spectroscopies. The key data were the doublet observed at δ 10.67 (1), 10.75 (2), 10.52 (3) and 10.82 (4), which are assigned to the ortho-CH group of the pyridyl fragment and are much downfield compared to the same signal of the free imidazolylium derivatives (δ 8.66, 8.62, 8.41 and 8.41). Moreover, the N-isopropyl derivatives 2 and 4 also exhibited a notable downfield septet at δ 6.87 (2), and 7.13 (4), which are assigned to the methine group of isopropyl substituent, whereas the same fragment of the uncomplexed imidazolylium ligands appeared at δ 4.77 and ‒5‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

4.92, respectively. These downfield shifted

1

Page 6 of 35

H NMR signals are due to the

inter-ligand C-H∙∙∙N hydrogen bonding coupled to the bipz chelate, which can be utilized as a proof to the formation of Pt(II) metal complexes 1−4. The single crystal X-ray diffraction study on 4 was executed; in fact it is the only complex that can grow the suitable single crystals. The molecules crystallize in a trigonal space group P3121, and the lack of a center of symmetry is demonstrated by the dimeric molecular motif along the c axis. The only crystallographically imposed symmetry is a threefold translational screw axis. The perspective view of single Pt(II) unit, packing in the crystal lattices, selected bond lengths and angles are shown in Figures 1a and 1b and caption. As can be seen, all Pt(II) complexes adopt the distorted square planar arrangement due to the geometrical constraint imposed by both neutral pyridyl imidazolylidene itpyim and dianionic bipz chelates. The metal to itpyim chelate distances in 4, cf. Pt-N(1) = 2.032(6) Å and Pt-C(6) = 1.988(8) Å, are notably shorter than those of Pt(II) metal complex bearing both pyridyl imidazolylidene chelate and two cis-arranged methyl substituents (i.e. Pt-N = 2.128(5) Å and Pt-C = 2.006(5) Å), showing the diminished trans-influence of the bipz chelate versus the alkyl groups.34 Moreover, the inter-ligand H-bond observed in the 1H NMR studies is also evidenced by the shortened non-bonding contacts of N(7)∙∙∙H(13) = 2.102(6) and N(4)∙∙∙H(1) = 2.294(5) Å between the nitrogen atom of bipz chelate and the pyridyl and isopropyl fragment of the pyridyl imidazolylidene chelate. The packing diagram also revealed a reduced Pt∙∙∙Pt contact of 3.480(4) Å within the Pt2 dimer and significantly lengthened Pt∙∙∙Pt contact of 4.321(4) Å between the Pt2 dimer units. The Pt∙∙∙Pt∙∙∙Pt angle is calculated to be 154.4ᵒ, showing a tilted packing arrangement. Moreover, within each dimer unit, the Pt(II) fragments are packed in the pyim-to-pyim and bipz-to-bipz fashion, and are rotated by ∼12.6ᵒ along the Pt∙∙∙Pt axis for reducing possible steric encumbrance between the peripheral bulky substituents. On the other hand, the Pt2 dimer units are stacked in similar

‒6‒ ACS Paragon Plus Environment

Page 7 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

pyim-to-pyim and bipz-to-bipz fashion, but with a much greater degree of rotation along the Pt∙∙∙Pt vector by ∼45ᵒ. Photophysical properties. The pertinent absorption spectra of Pt(II) complexes 1−4 are shown in Figure 2, while numerical data are listed in Table 1. The shoulder shown in the longer wavelength region is assigned to the metal-to-ligand charge transfer (MLCT) transition, whereas the intense absorptions at ∼340 nm in CH2Cl2 solution are due to the intra-ligand π-π* transitions from the dianionic bipz chelate to the neutral N^C chelate. Albeit variation of the extinction coefficient, all observed peak profiles are very close to each other, supporting the existence of similar structure for these Pt(II) complexes in solution state. As for the emission spectra, these Pt(II) complexes were all non-emissive in degassed CH2Cl2 solution (vide infra), but showed intense emission for powders that were obtiend from vacuum sublimation (using a temperature-gradient tube furnace). Their higher emission quantum yield (Q.Y.), on the one hand, is caused by the rigid environment, which is expected to reduce the external quenching as well as to have a more destabilized MC dd excited state due to the hindered stretching motion of metal-ligand bonding that destabilized the dσ*-orbitals. On the other hand, Pt(II) complex 1, bearing the N-methyl substituted imidazolylidene, showed a broadened and structureless profile with peak max. located at 510 nm. The exceedingly large Stokes shift (cf. absorption in CH2Cl2, see Figure 2) and high emission quantum yield Q.Y. = 91% with short phosphorescence lifetime τobs = 1077 ns are in good agreement with the metal‒metal-to-ligand charge transfer (MMLCT) character observed in many square planar Pt(II) complexes that exhibited extensive solid-state aggregation.48-50 For the respective N-isopropyl substituted Pt(II) complex 2, the increased steric encumbrance increases the spatial separation between the Pt(II) metal complexes and reduces the MMLCT character, giving a rather broad emission profile with multiple peak wavelengths at 455, 480, 524 nm, which probably originate from the mixture of monomeric and aggregated species. The concomitant increase of average

‒7‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 35

emission lifetime (τobs = 3074 ns) and reduction of Q.Y. (53%) are also consistent with the reduction of MMLCT character. Similarly, the corresponding mtpyim Pt(II) complex 3 showed the largest red-shifted emission with λmax = 552 nm, Q.Y. = 95 % and τobs = 804 ns. Though lacking the X-ray structural data, it is reasonable to assume that complex 3 adopts similar crystal structure with that of 4, for which the Pt2 dimer units are stacked in the pyim-to-pyim and bipz-to-bipz arrangement. Accordingly the balance between van der Waal attraction and steric interaction for both alkyl substituents becomes crucial. For example, the t-butyl substituent on the pyridyl fragment (cf. complexes 3 and 4) would have improved, rather than prohibited, the inter-molecular packing interaction by rotation and sliding of the adjacent Pt(II) metal complex against one another, so that the unfavorable steric interaction between the pair of Pt2 dimers can be sequestered. As for R’ = tBu and R” = Me in complex 3 it is believed that the void space between two Pt2 dimers is large enough to accommodate tBu and Me groups such that the Pt∙∙∙Pt stacking in solid state is shortened, resulting in the higher proportion of MMLCT character and hence lower emission energy gap. Changing R” = methyl group (mtpyim) in 3 to bulky isopropyl substituent (itpyim) in 4 may increase the steric hindrance. As a result, the spatial separation between the Pt∙∙∙Pt distance increases with corresponding reduction of MMLCT, which accounts for a blue-shift of MMLCT emission maximized at 501 nm for 4 (cf. 552 nm in 3). This phenomenon is virtually the same as the reduction in MMLCT character for isopropyl-to-methyl substitution detected in the parent Pt(II) metal complexes 1 and 2. Nevertheless, 4 exhibits the highest emission Q.Y. of 99 % and comparable lifetime of τobs = 1033 ns as the neat powder. We then employed a computational time-dependent density functional theory (TD-DFT) approach to gain in-depth insight into these Pt(II) complexes at the single molecular level. The calculated singlet and triplet absorption characteristics for monomeric Pt(II) complexes 1‒4 are displayed in Table 2, Figure 3, Tables S1−S4 and Figures S1−S4 (see Supporting Information). The calculated energy gaps for the S0 → ‒8‒ ACS Paragon Plus Environment

Page 9 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

S1 optical transition (in terms of wavelength) are 421.5 nm (1), 420.3 nm (2), 408.8 nm (3), 408.1 nm (4), and for the S0 → T1 optical transition are 446.2 nm (1), 445.2 nm (2), 435.5 nm (3), 434.5 nm (4). In reality, due to the strong spin-orbit coupling induced by Pt(II) metal atom the lowest lying states in the singlet and triplet manifolds, i.e. S1 and T1, should be strongly mixed. On this basis, the calculated energy gaps are in agreement with the threshold/onset of their absorption spectra around 400-430 nm in CH2Cl2 (see Figure 2). For the S0 → S1 and S0 → T1 optical transition, their main assignments are all HOMO → LUMO processes, in which the electron density distribution of HOMO for Pt(II) complexes 1− −4 are mainly localized at bipz fragments and partially at Pt(II) atoms. On the other hand, the electron density distribution of LUMO for Pt(II) complexes 1− −4 are all mainly localized at mpyim (for 1), ipyim (for 2), mtpyim (for 3) and itpyim (for 4) fragments, respectively, with few percentages at Pt(II) atoms (see Figure 3). The computational results thus suggest that HOMO primarily contains contributions from both dianionic chelate and central Pt(II) atom, whereas LUMO reside on the neutral chelates. Accordingly, the lowest lying transitions are commonly described as metal‒ligand-to-ligand charge transfers (MLLCT).5 We then moved one more step to probe the self-assembly of the titled Pt(II) complexes in an attempt to mimic the observed emission in the solid state. In this approach, we mainly focused on complex 4 due to its well resolved X-ray crystal structure (vide supra), from which the packing diagram revealed a Pt∙∙∙Pt contact of 3.480(4) Å and was truncated as a Pt2 dimer to perform the computation. As a result, the energy gaps of the S0 → S1 and S0 → T1 optical transition for Pt(II) complex 4 dimer are calculated to be 463.5 and 482.4 nm, respectively (Table 2 and S5), in which the S0 → T1 optical transition closely matches the observed peak wavelength of the complex 4 emission in solid (see Figure 2). The assignments of the S0 → S1 and S0 → T1 optical transition for dimer of 4 are both solely HOMO → LUMO process. The electron density distribution of HOMO and LUMO for the dimer of 4 are largely localized at bipz and itpyim fragments, respectively (Figure 3 and S5). ‒9‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 35

However, interestingly, as for the truncated Pt2 dimer the electron density distributions have non-negligible contribution from both Pt atoms in HOMO and LUMO, which are located at nonbonding 5dπ and 5dσ* orbitals, respectively. Similar MC dd excited state contributed to the HOMO → LUMO process was also noticed in its monomeric form (vide supra). It is believed that the repulsive dd potential energy surface, in part, accounts for the non-emissive character of the titled monomeric Pt(II) complexes in solution. On this basis, the calculated lowest lying MC dd state in the Pt2 dimer cannot rationalize the high emission QY of 4 in the solid. In other words, the truncated Pt2 dimer is not sufficient to represent the origin of emission in solid. It is notable that our calculation was based on the Pt∙∙∙Pt contact of 3.480(4) Å to simulate the corresponding photophysical properties, for which the employed Pt∙∙∙Pt distance is longer than the known distance with substantial Pt∙∙∙Pt interaction (i.e. ∼3.3 Å)51-52 and the GIXS result of vacuum deposited thin film of 3 (vide infra). Further optimization is unfortunately prohibited due to the limited computation capacity. Triboluminescence. Triboluminescence is the phenomenon where the fracture of material gives production of the spontaneous emission.39,

53

It has been well

established that dipolar structures and noncentrosymmetric arrangements of crystals are the two crucial criteria for achieving the triboluminescence.40-42, 54 In the present system, the Pt(II) complexes 1‒4 are assembled using both the charge neutral and dianionic chelates, which fulfill the requirement of having dipolar structures.55-57 Furthermore, in contrast to the Pt(II) complexes 1‒3, which show no triboluminescence, grinding the powder of 4, for which the molecules crystallized in a noncentrosymmetric space group P3121, with a spatula or glass slide produces prominent bluish green flashes that are visible to naked eyes under the condition of indoor lighting at RT (see video in SI), confirming the second criterion. This triboluminescence is considered as an electric polarization phenomenon, with charge separation being the primary event and the recombination of separated charges during fracture resulting in the spontaneous emission. Hence, breaking the crystals ‒ 10 ‒ ACS Paragon Plus Environment

Page 11 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

of 4 by mechanical stresses electrifies the cracked crystal surfaces, which then electronically excites the molecules to afford the triboluminescence.58 This observation suggests that the existence of both t-butyl and i-propyl groups of chelates among the titled Pt(II) complexes are crucial for the formation of noncentrosymmetric crystals and hence the generation of triboluminescence. Nevertheless, what controlled the formation of noncentrosymmetric arrangements in the crystalline state is still pending resolution. For example, the crystal packing is similar between 3 and 4 whereas triboluminescence was not observed in 3. To

gain

further

insight

into

the

triboluminescent

property,

both

photoluminescence and triboluminescence spectra of solid Pt(II) complex 4 are shown in Figure 4. Comparing the photoluminescence spectra, which reveals typical broad MMLCT emission (vide supra) maximized at ∼506 nm, the triboluminescence shows a similar spectral feature but is slightly red shifted to 509 nm by ∼3 nm. This spectral shift is authentic because an identical detecting system incorporating a charge coupled detector (CCD) was used throughout the measurement. For further comparison, photoluminescence of the sample after being ground is also shown in Figure 4. Clearly, the slight red shift of the emission was also observed. After dissolution of the ground sample in e.g. CH2Cl2 and recovering the solid sample, its photoluminescence regains the original spectrum (prior to grinding). The results demonstrate that upon imposing the mechanic force the surface of origin crystal structure of 4, to certain extents, has been altered, giving rise to the triboluminescence that is slightly different from the photoluminescence of the unperturbed one. In an aim to verify the viewpoint of structural difference before and after mechanic perturbation, Raman spectroscopy was then carried out to probe the possible surface structure transformation. In comparison to the Raman spectrum of the original complex 4, the results shown in Figure 5 clearly reveal that the ground sample causes the changes of several vibrational peaks in terms of frequencies or relative intensity. Careful analyses indicate that almost all vibration peaks were ‒ 11 ‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 35

shifted by ∼2-4 cm−1 to the lower frequency. Especially, the low frequency Raman peak of 151 cm−1 for 4 (Figure 5(a)), which is assigned to the Pt-Pt stretching vibration of A1 symmetry,59 shifts to lower frequency of 149 cm−1 (Figure 5(b)). Also changed were the vibrational peaks ascribable to C=C and C=N stretching around 1600∼1650 cm-1, which were shifted from 1623 cm-1 and 1637 cm-1 to 1619 cm-1 and 1633 cm-1, respectively. These variations lead us to conclude that during grinding, the mechanical stress energy weakens both Pt∙∙∙Pt and π-π interactions at the crystal surfaces, in a way similar to that exerted by photoexcitation. The loosened molecule then regains their ground state architecture and releases energy that induces the observed triboluminescence.

Electrochemistry. Data of cyclic voltammetry were compiled in Figure S6 of electronic supporting information. As can be seen, all Pt(II) complexes exhibit irreversible oxidation peak potentials in a narrow range of 1.12 ∼ 1.17 V, which are consistent with the metal-centered oxidation. Therefore, the substituents on the pyridyl imidazolylidene can only impose minor influence to the oxidation peak potential. On the other hand, the reduction peak potentials occur at a slightly larger range between −2.19 V and −2.34 V, among which both of the 4-t-butylpyridine derivatives 3 and 4 (−2.30 V and −2.34 V) exhibit a notable negative shift versus that of the simple pyridine derivatives 1 and 2 (−2.23 V and −2.19 V). This difference may be rationalized in terms of the electron donating character of t-butyl substituent that destabilized the π*-orbital of the pyridyl imidazolylidene chelate. Electroluminescence. Among complexes 1‒4, complex 3 retains a high emission QY of 96% when vacuum-deposited as a pure/non-doped thin film, with emission properties (i.e., emission spectrum and PLQY) mimicking those of the powder and is promising for use as the pure/non-doped emitting layer for OLEDs.60 Complex 3 was thus subjected to electroluminescent (EL) studies using architecture: ITO/MoO3 (1 nm)/TAPC (65 nm)/mCP (8 nm)/3 (pure/non-doped, 30 nm)/3TPYMB (50 nm)/LiF (1 nm)/Al

(120

nm).

TAPC,

mCP

and

3TPYMB

stand

for ‒ 12 ‒

ACS Paragon Plus Environment

Page 13 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

1,1-bis[(di-4-tolylamino)phenyl]cyclohexane,

N,N-dicarbazolyl-3,5-benzene

and

tris-[3-(3-pyridyl)mesityl]borane, respectively, and serve either as the hole- or electron-transport layers.61-62 Vacuum-deposited thin film of 3 constituted the emitting layer (EML). Representative EL characteristics of OLEDs using the non-doped emitting layer of 3 are shown in Figure 6(a)‒6(c). Its EL spectrum (yellow emission) is similar to its PL spectra in both vacuum-deposited thin films and in powders. The slight difference between the thin-film PL spectrum and the EL spectrum is mainly due to the weak microcavity effect in the device environment. The device exhibits a turn-on voltage of ∼ 2.8−3 V and an operation voltage ∼ 4 V for a brightness of 100 cd∙m‒2. Most importantly, the non-doped device of 3 shows very high maximal EL efficiencies of (26%, 91 cd∙A‒1, 64.4 lm∙W‒1). In addition, the device retains high external quantum efficienciy (EQE) and luminesce efficiency of (25.9%, 90 cd⋅A‒1) at the practical brightness of 100 cd∙m-2 and (24.4%, 85 cd⋅A‒1) even at the high brightness of 1000 cd∙m‒2, respectively. Such a low efficiency roll-off may be attributed to the short excited-state lifetimes of 3 (∼1000 ns in the vacuum-deposited thin film as shown in Figure 7, and ∼800 ns in the solid-state powder as discussed in the previous section), which in turn is associated with the MMLCT transition character of 3. To our knowledge, the recorded EL efficiencies are among the highest of non-doped phosphorescent OLEDs (using pure/non-doped transition metal complex emitting layers) ever documented in literature.63-64 Although >30% EQEs had been reported for red-emitting non-doped phosphorescent OLEDs (with emission wavelength > 600 nm) utilizing MMLCT based Pt(II) complexes,63-64 this work represents a first demonstration that similarly high EQEs can be achieved for shorter emission wavelengths (i.e., < 600 nm) based on the MMLCT transition mechanism of Pt(II) complexes. The results thus shed light to further development of high-efficiency non-doped phosphorescent OLEDs and MMLCT based OLEDs spanning different visible colors.

‒ 13 ‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 35

The formation of ground-state aggregates in vacuum-deposited thin film of 3 is verified by UV-Vis spectroscopy and GIXS (grazing incidence X-ray scattering) analyses.65-66 Figure 8(a) exhibited the UV-Vis absorption spectra of 3 recorded in both CH2Cl2 solution and as vacuum-deposited thin film; the latter clearly shows the emergence of a broad peak profile at ∼400 nm associated with ground-state aggregates. The GIXS data shown in Figure 8(b) also exhibited broadened diffraction peaks along the out-of-plane orientation (Qz), and weak diffractions along the in-plane orientation (i.e., along Qxy). Peaks around Qxy = 1.9 Å‒1 of the in-plane orientation can be attributed to the self-assembly/stacking of 3 into the columnar stacks in the vacuum-deposited thin film.65 In addition, the intermolecular stacking signal in the in-plane direction (i.e., along Qxy) suggests a preferential columnar alignment parallel to the substrate surface.65 The stacking distance is calculated to be 3.33 Å, which is similar to the Pt∙∙∙Pt separation detected by single-crystal XRD experiment on 4 (3.48 Å), and allowed the formation of MMLCT transitions.51-52 Although the detected signal for intermolecular stacking of Pt complex 3 in GIXS is not strong, it is however clearly visible. The absorption spectrum of the vacuum deposited thin film 3 (as shown in Figure 8(a)), when compared to absorption spectrum of 3 in solution, also firmly supports formation of notable ground-state aggregation. The weak GIXS signal for 3 is due to the poorer crystallinity exerted by vacuum deposition. Its intensity in thin film depends on many factors, such as the sizes of crystallites, the ratios and the orientation distribution of aggregates, and the excitation beam intensity, etc. The weakened stacking signal in GIXS analyses and its relatively broad angular distribution may indicate the formation of small (nanoscale) aggregates and somewhat scattered orientations of columnar aggregates. Meanwhile, the broadened diffraction peak along the out-of-plane axis (i.e., along Qz) and the smaller Qz = 0.38 Å‒1 represents the inter-columnar assembly and corresponds to an inter-columnar d-spacing of 22.2 Å‒1.65-66 The broadened and ring-like GIXS pattern, however, indicate a somewhat random distribution of the inter-columnar packings.

‒ 14 ‒ ACS Paragon Plus Environment

Page 15 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

The excitation spectrum of 3 in vacuum-deposited thin films is similar to the corresponding absorption spectra, indicating that the emission of vacuum deposited 3 films originates from the identical ground electronic state, i.e., the ground-state aggregation. The transient PL measurement on vacuum-deposited thin films of 3 (Figure 7) reveals mono-exponential decay characteristics with an excited state lifetime of ∼1000 ns, similar to that in the solid-state powder of 3 as discussed in the previous section). These photophysical properties again support the MMLCT transition character of 3 from the ground-state aggregation in vacuum deposited thin films. Emitting dipole orientations in films of 3 were also characterized by the angleand polarization-dependent photoluminescent experiment.67-69 The preferential horizontal (in-plane) transition dipoles were attained in vacuum-deposited films of 3, with the intermolecular stacking being parallel to the substrate and the MMLCT transition along the stacking axis.70 The measured p-polarized PL intensities of 3 as a function of the emission angles are shown in Figure 9(a). The measured data are compared to results simulated with different horizontal dipole ratios Θ// (the percentage of horizontal dipoles among all emitting dipoles) to extract Θ// in the emitting layer (cf. Θ// = 100% for fully horizontally arranged orientation and Θ// = 67% for the isotropic arranged transition dipole). The measured data match well with those of Θ// = 74% shown in Figure 9(b), clearly revealing its favorable horizontal (in-plane) transition orientation. Recent reports have revealed the importance of having preferential transition dipoles along the in-plane (horizontal) orientation for higher optical out-coupling,67-69 since vertical emitting dipoles in OLEDs generally contribute little to the extracted emission and, hence, OLED with emitters showing preferential horizontal dipole ratio (Θ// > 67%) are desirable for improving OLED optical out-coupling and external quantum efficiencies.67-69 The high external quantum efficiency of 3-based OLED thus may result from both the higher PLQY and the preferential horizontal transition dipoles.

‒ 15 ‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 35

As a comparison, although (crystalline) powder of 4 (obtained from vacuum sublimation) also show a very high PLQY, yet its vacuum-deposited non-doped thin film gives a much reduced PLQY of only 5-6%. As a consequence, the OLED fabricated using vacuum-deposited thin layer of 4 exhibit much inferior EL efficiencies (3.4%, 10.7 cd∙A-1, 6.5 lm∙W-1) (see ESI) vs. the OLED fabricated using vacuum deposited thin film of 3. Furthermore, in sharp contrast to the spectral behavior of 3, neither the UV-Vis spectrum nor the GIXS analyses of 4 have shown the aggregate formation and stacked structure in the vacuum-deposited thin film (see Figure S9 in ESI), while the transient PL exhibited the multi-exponential decay kinetics (see Figure 7). The latter is attributed to the weakened ground-state aggregation, giving various trapping sites that exhibit slightly different excited-state relaxation dynamics. Hence, this result explains the lowered PLQY of vacuum-deposited thin film vs. that of sublimed sample of 4.

Conclusion In summary a new series of charge-neutral Pt(II) metal complexes 1‒4 bearing both pyrid-2-yl-imidazolylidene and dianionic bipz chelates have been strategically designed and synthesized. In serendipity, complex 4 revealed remarkable triboluminescence which is tentatively rationalized by the unique arrangement for the isopropyl and tert-butyl substituents on the pyrid-2-yl-imidazolylidene chelate. The triboluminescence is found to be sensitive to subtle changes of the associated substituents and disappears in 1‒3 upon replacing either isopropyl or tert-butyl substituents by methyl or hydrogen atom, respectively. For complex 3, the well-ordered solid packing, supported by the grazing incidence X-ray scattering experiment, and thus the aligned dipole serves as an excellent emitter for the fabrication of highly efficient OLEDs, achieving superb external quantum efficienciy (25.9%,) and luminesce efficiency (90 cd⋅A‒1) at the practical brightness of 100 cd∙m‒2. The rather low roll-off in efficiency is attributed to the short excited-state lifetime of 3 (∼ 800 ns) in the solid state, which in turn is associated with the MMLCT transition ‒ 16 ‒ ACS Paragon Plus Environment

Page 17 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

character. The results broaden the horizon for further development of high-efficiency non-doped MMLCT phosphorescence based OLEDs.

Experimental Section. General Procedures. All reactions were performed under nitrogen. Solvents were distilled from appropriate drying agents prior to use. Commercially available reagents

were

used

without

further

purification.

The

corresponding

3-alkyl-1-(pyrid-2-yl)-1H-imidazolium salts were synthesized according to literature procedures, alkyl = methyl and iPr.43-44 5,5'-di(trifluoromethyl)-2H,2'H-3,3'-bipyrazole (bipzH2) was prepared from condensation of 2,3-butanedione and ethyl trifluoroacetate, followed by treatment with hydrazine.45 All reactions were monitored by TLC with pre-coated silica gel plates (Merck, 0.20 mm with fluorescent indicator UV254). Flash column chromatography was carried out using silica gel obtained from Merck (230-400 mesh). Mass spectra were obtained on a JEOL SX-102A instrument operating in electron impact (EI) or fast atom bombardment (FAB) mode. 1H and 19F NMR spectra were recorded on a Bruker-400 or INOVA-500 instrument. Elemental analysis was carried out with a Heraeus CHN-O Rapid Elementary Analyzer. Preparation of 3-alkyl-1-(pyrid-2-yl)-1H-imidazolium PF6− salt. To a typical reaction, a DMSO solution of corresponding 2-bromopyridine, imidazole, K2CO3, CuI and L-proline was stirred at 60oC for 24 h; after then, excess of ethyl acetate was added. The solution was washed with deionized water, dried over Na2SO4 and concentrated by rotary evaporation. Flash column chromatography afforded the corresponding 2-pyridylimidazole. It was then treated with methyliodide in acetonitrile at 65oC for 12 h. After cooled to RT, the resulting mixture was filtered, and a saturated KPF6 water solution was added to the filtrate, giving white precipitate as the desired product upon concentration. Data of (mpyim)HPF6, yield: 91 %. 1H NMR (400 MHz, d6-acetone, 298K): δ 9.88 (s, 1H), 8.66 (dd, J = 4.0, 1.2 Hz, 1H), 8.47 (t, J = 2.0 Hz, 1H), 8.21 (dt, J = 7.6, 2.0 Hz, ‒ 17 ‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 35

1H), 8.03 (d, J = 7.6 Hz, 1H), 7.97 (t, J = 2.0 Hz, 1H), 7.66 (ddd, J = 7.6, 4.0, 1.2 Hz, 1H), 4.23 (s, 3H). 19F NMR (376 MHz, d6-acetone, 298K): δ −71.64 (d, JFP = 717 Hz, 6F). Data of (ipyim)HPF6, yield: 54 %. 1H NMR (400 MHz, d6-acetone, 298K): δ 9.99 (s, 1H), 8.62 (d, J = 4.8 Hz, 1H), 8.48 (t, J = 1.6 Hz, 1H), 8.17 (dt, J = 7.6, 1.6 Hz, 1H), 8.10 (t, J = 1.6 Hz, 1H), 7.99 (d, J = 7.6 Hz, 1H), 7.61 (dd, J = 7.6, 4.8 Hz, 1H), 4.77 (septet, J = 6.8 Hz, 1H), 1.53 (d, J = 6.8 Hz, 6H). 19F NMR (376 MHz, d6-acetone, 298K): δ −70.19 (d, JFP = 711 Hz, 6F). Data of (mtpyim)HPF6, yield: 64 %. 1H NMR (400 MHz, CDCl3, 298K): δ 9.35 (s, 1H), 8.41 (d, J = 5.2 Hz, 1H), 8.16 ∼ 8.14 (m, 1H), 7.78 (s, 1H), 7.44 ∼ 7.42 (m, 2H), 4.09 (s, 3H), 1.36 (s, 9H). 19F NMR (376 MHz, CDCl3, 298K): δ −71.97 (d, JFP = 710 Hz, 6F). Data of (itpyim)HPF6, yield: 62 %. 1H NMR (400 MHz, CDCl3, 298K): δ 9.39 (s, 1H), 8.41 (d, J = 5.2 Hz, 1H), 8.21 (s, 1H), 7.83 (s, 1H), 7.51 (s, 1H), 7.44 (d, J = 5.2 Hz, 1H), 4.92 (septet, J = 6.8 Hz, 1H), 1.64 (d, J = 6.8 Hz, 6H), 1.38 (s, 9H). 19F NMR (376 MHz, CDCl3, 298K): δ −71.9 (d, JFP = 714 Hz, 6F). Preparation of Pt(II) complexes. A mixture of Pt(DMSO)2Cl2 (703 mg, 1.7 mmol), (mpyim)HPF6 (560 mg, 1.7 mmol) and NaHCO3 (210 mg, 2.5 mmol) in 10 mL of anhydrous DMSO was heated at 100oC for 12 h. After cooled to RT, bipzH2 (520 mg, 1.7 mmol) was added and the mixture was stirred at 100oC for 6 h. After then, the reaction was quenched by addition of excess of water to induce the immediate precipitation. It was filtered and washed with water and ethyl ether in sequence, and sublimed in vacuo to afford light green colored solid of [Pt(mpyim)(bipz)] (1, 540 mg, 68 %). Other Pt(II) analogues, i.e. colorless [Pt(ipyim)(bipz)] (2, 62%), yellow [Pt(mtpyim)(bipz)] (3, 68%) and light green [Pt(itpyim)(bipz)] (4, 52%), were prepared in an analogous manner. Spectral data of [Pt(mpyim)(bipz)] (1): MS (FAB, 195Pt): m/z 622.7 [M+]; 1H NMR (400 MHz, d6-DMSO, 298K): δ 10.67 (br, 1H), 8.35 (br, 2H), 8.08 (br, 1H), 7.64 (br, 1H), 7.57 (br, 1H), 6.65 (br, 1H), 6.59 (br, 1H) , 4.45 (s, 3H). 19F NMR (376 MHz, d6-DMSO,

‒ 18 ‒ ACS Paragon Plus Environment

Page 19 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

298K): δ −58.87 (s, 3F), −59.34 (s, 3F). Anal. Calcd. for C17H11F6N7Pt: C, 32.81; H, 1.78; N, 15.75. Found: C, 32.59; H, 1.62; N, 15.58. Spectral data of [Pt(ipyim)(bipz)] (2): MS (FAB, 195Pt): m/z 650.4 [M+]; 1H NMR (400 MHz, d6-DMSO, 298K): δ 10.75 (d, J = 4.8 Hz, 1H), 8.48 (d, J = 2.4 Hz, 1H), 8.37 (t, J = 6.8 Hz, 1H), 8.14 (d, J = 8.0 Hz, 1H), 7.97 (d, J = 2.4 Hz, 1H), 7.69 (t, J = 6.8 Hz, 1H), 6.87 (septet, J = 6.6 Hz, 1H), 6.72 (s, 1H), 6.66 (s, 1H), 1.44 (d, J = 6.6 Hz, 6H). 19F NMR (376 MHz, d6-DMSO, 298K): δ −58.90 (s, 3F), −59.54 (s, 3F). Anal. Calcd. for C19H15F6N7Pt: C, 35.08; H, 2.32; N, 15.07. Found: C, 34.85; H, 2.26; N, 15.19. Spectral data of [Pt(mtpyim)(bipz)] (3): MS (FAB, 195Pt): m/z 678.6 [M+]; 1H NMR (400 MHz, d6-acetone, 298K): δ 10.52 (d, J = 6.0 Hz, 1H), 8.47 (d, J = 2.0 Hz, 1H), 8.02 (s, 1H), 7.65 (d, J = 6.0 Hz, 1H), 7.56 (d, J = 2.0 Hz, 1H), 6.63 (s, 1H), 6.56 (s, 1H) , 4.41 (s, 3H), 1.37 (s, 9H). 19F NMR (376 MHz, d6-acetone, 298K): δ −58.83 (s, 3F), −59.31 (s, 3F). Anal. Calcd. for C21H19F6N7Pt: C, 37.17; H, 2.82; N, 14.45. Found: C, 37.49; H, 2.78; N, 14.67. Spectral data of [Pt(itpyim)(bipz)] (4): MS (FAB, 195Pt): m/z 706.8 [M+]; 1H NMR (400 MHz, d6-acetone, 298K): δ 10.82 (d, J = 6.4 Hz, 1H), 8.30 (d, J = 2 Hz, 1H), 8.00 (s, 1H), 7.73 (d, J = 2 Hz, 1H), 7.59 (d, J = 6.4 Hz, 1H), 7.13 (septet, J = 6.6 Hz, 1H), 6.60 (s, 1H), 6.54 (s, 1H), 1.51 (d, J = 6.6 Hz, 6H), 1.46 (s, 9H). 19F NMR (376 MHz, d6-acetone, 298K): δ −60.66 (s, 3F), −61.23 (s, 3F). Anal. Calcd. for C23H23F6N7Pt: C, 39.10; H, 3.28; N, 13.88. Found: C, 39.22; H, 3.44; N, 13.92. Selected crystal data of 4: C23H23F6N7Pt; M = 706.57; trigonal; space group = P3121; a = 14.0948(2) Å, b = 14.0948(2) Å, c = 21.1510(3) Å; α = β = 90o, γ = 120o, V = 3638.98(9) Å3; Z = 6; ρcalcd = 1.935 Mg∙m−3; F(000) = 2052; µ = 5.857 mm−1; crystal size = 0.40 × 0.20 × 0.18 mm3; λ(Mo-Kα) = 0.71073 Å; T = 150(2) K; 16228 reflections collected, 5508 independent reflections (Rint = 0.0339), restraints / parameters = 19 / 349, GOF = 1.160, final R1[I > 2σ(I)] = 0.0296 and wR2(all data) = 0.0817; largest diff. peak and hole = 1.270 and -1.372 e∙Å-3. Single Crystal X-Ray Diffraction Studies: Single crystal X-ray diffraction study was measured with a Bruker SMART Apex CCD diffractometer using (Mo-Kα) ‒ 19 ‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 35

radiation (λ = 0.71073 Å). The data collection was executed using the SMART program. Cell refinement and data reduction were performed with the SAINT program. An empirical absorption was applied based on the symmetry-equivalent reflections and the SADABS program. The structures were solved using the SHELXS-97 program and refined using the SHELXL-97 program by full-matrix least squares on F2 values. The structural analysis and molecular graphics were obtained using the SHELXTL program on a personal computer.

Supporting information. Details of experimental procedures and data. CIF data of Pt(II) complex 4 (with deposition number: CCDC 1499101), video that showed the triboluminescence, results of the TD-DFT calculation of all Pt(II) complexes, and OLED device characteristics.

Acknowledgments. This work was supported by the Ministry of Science and Technology of Taiwan (MOST) under the grant numbers: 104-2119-M-007-001, 105-3113-E-002-008 and 104-2221-E-002-152-MY3.

‒ 20 ‒ ACS Paragon Plus Environment

Page 21 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

(a)

(b)

Figure 1. (a) Structural drawing of [Pt(itpyim)(bipz)] (4) with thermal ellipsoids shown at 50% probability level. Selected bond distances: Pt-C(6) = 1.988(8), Pt-N(6) = 2.019(5), Pt-N(1) = 2.032(6), Pt-N(5) = 2.052(6), N(7)∙∙∙H(13) = 2.102(6) and N(4)∙∙∙H(1) = 2.294(5) Å, and bond angles: N(1)-Pt-C(6) = 79.0(3), N(1)-Pt-N(6) = 174.3(3), N(5)-Pt-C(6) = 171.4(2) and N(5)-Pt-N(6) = 77.8(2)°. (b) Molecular stacking diagram with Pt∙∙∙Pt contacts of 3.480(4) and 4.3214(4) Å and Pt∙∙∙Pt∙∙∙Pt angle of 154.4ᵒ.

‒ 21 ‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1.2

1 2 3 4

3

-1

-1

Extinction Coefficient (x10 M cm )

20

15

1.0 0.8 0.6

10

0.4 5

0.2 0

300

400

500

600

Normalized Intensity (a.u.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 35

0.0 700

Wavelength (nm)

Figure 2. UV-Vis absorption spectra in CH2Cl2 and solid state emission spectra of Pt(II) complexes 1‒4 measured at RT.

1 monomer

2 monomer

3 monomer

4 monomer

4 dimer

Pt: 8.88%

Pt: 8.81%

Pt: 9.33%

Pt: 9.19%

Pt: 11.4%

Pt: 4.11%

Pt: 4.12%

Pt: 4.25%

Pt: 4.31%

Pt: 3%

L

H

Figure 3. Frontier molecular orbitals (HOMO [H] and LUMO [L]) involved in the lowest-lying transition for monomeric Pt(II) complexes 1–4 and dimer of 4 in CH2Cl2. The electron density distribution of Pt(II) atom(s) in each molecular orbital is also showed.

‒ 22 ‒ ACS Paragon Plus Environment

Page 23 of 35

a b c

Intensity (a.u.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

425

450

475

500

525

550

575

600

625

wavelength (nm) Figure 4. Emission spectra of Pt(II) complex 4 at RT. (a) Photoluminescence excited by 375 nm laser (emission maximum 506 nm), (b) triboluminescence (emission maximum 509 nm), and (c) photoluminescence of grinded complex 4 excited by 375 nm (emission maximum 508 nm).

‒ 23 ‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

a b

100

Relative Intensity

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 35

150

800

1400

200

850

1425

2750

250

900

1450

2800

300

950

1475

2850

350

400

450

500

550

600

650

700

750

1000 1050 1100 1150 1200 1250 1300 1350 1400

1500

2900

1525

2950

1550

1575

3000

1600

3050

1625

3100

1650

3150

1675

3200

-1

Raman shift (cm )

Figure 5. Raman spectra of Pt(II) complex 4 excited by 532 nm laser. (a) Before and (b) after grinding.

‒ 24 ‒ ACS Paragon Plus Environment

Page 25 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 6. (a) EL spectrum together with PL spectrum of the vacuum-deposited thin film of 3, (b) current-voltage-luminance (I-V-L) characteristics, and (c) efficiency characteristics of the OLED using Pt(II) complex 3 as the pure/non-doped emitting layer.

‒ 25 ‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1

10

Normalized Intensity

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 35

3 4

0

10

-1

10

-2

10

0

2

4

6

8

10

Time (µs) Figure 7. The transient PL decay curves of vacuum deposited films of 3 and 4. The decay curve of 3 can be fitted by a mono-exponential function with a lifetime of 1050 ns. The curve of 4 is more complicated and cannot be fitted by a mono-exponential decay kinetics.

‒ 26 ‒ ACS Paragon Plus Environment

Page 27 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 8. (a) Absorption spectra of 3 in the dilute CH2Cl2 solution and in the vacuum-deposited thin film. (b) GIXS pattern of the vacuum-deposited non-doped thin film of 3.

‒ 27 ‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 35

Figure 9. (a) Measured (symbols) p-polarized PL intensity (around PL peak wavelength) of the vacuum-deposited pure film of 3 as a function of the emission angle and wavelength. The measured data are compared with simulated ones with different horizontal dipole ratios Θ// (e.g., Θ// = 100% for fully horizontal dipoles and Θ//= 67% for the isotropic dipole orientation) to extract the horizontal emitting dipole ratios of the emitting layer. The measured data match well with those calculated with Θ// = 74% shown in (b).

‒ 28 ‒ ACS Paragon Plus Environment

Page 29 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Table 1. Selected photophysical properties of Pt(II) complexes 1‒4. abs λmax [nm][a] (ε x 103 M-1cm-1)

[b]

Φ[b]

τ[b]

kr

knr

[%]

[ns]

[s-1]

[s-1]

em λmax [nm]

1

344 (11.7)

510

91

1077

8.4 x 105

8.4 x 104

2

343 (7.5)

455, 480, 524

53

3074

1.7 x 105

1.5 x 105

3

340 (13.6)

552

95

804

1.2 x 106

6.2 x 104

4

340 (13.2)

501

99

1033

9.6 x 105

9.7 x 103

[a] UV-Vis spectra were recorded in CH2Cl2 solution. [b] PL data measured in solid state as neat powders at RT. Solid-state powders of all complexes were collected from temperature-gradient vacuum sublimation. [c] kr and knr were calculated according to the equations, kr = Φ /τobs and knr = (1/τobs) − kr.

Table 2. The calculated wavelengths, transition probabilities and charge character of the lowest optical transitions S1 and T1 (with mainly assignments) for monomeric Pt(II) complexes 1–4 and dimer of 4 in CH2Cl2.

1 2 3 4 4x2

state

λcal (nm)

f

assignment

MLCT

T1

446.2

0

HOMO → LUMO (86%)

-4.92%

S1

421.5

0.0002

HOMO → LUMO (97%)

-4.63%

T1

445.2

0

HOMO → LUMO (86%)

-4.98%

S1

420.3

0.0259

HOMO → LUMO (97%)

-4.55%

T1

435.5

0

HOMO → LUMO (84%)

-5.28%

S1

408.8

0.0004

HOMO → LUMO (96%)

-4.88%

T1

434.5

0

HOMO → LUMO (84%)

-5.12%

S1

408.1

0.0005

HOMO → LUMO (96%)

-4.68%

T1

482.4

0

HOMO → LUMO (86%)

-7.35%

S1

463.5

0.0001

HOMO → LUMO (95%)

-7.98%

‒ 29 ‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 35

References 1. Wong, K. M.-C.; Yam, V. W.-W. Self-Assembly of Luminescent Alkynylplatinum(ii) Terpyridyl Complexes: Modulation of Photophysical Properties through Aggregation Behavior. Acc. Chem. Res. 2011, 44, 424-434. 2. Visbal, R.; Gimeno, M. C. N-Heterocyclic Carbene Metal Complexes: Photoluminescence and Applications. Chem. Soc. Rev. 2014, 43, 3551-3574. 3. Chi, Y.; Tong, B.; Chou, P.-T. Metal Complexes with Pyridyl Azolates: Design, Preparation and Applications. Coord. Chem. Rev. 2014, 281, 1-25. 4. Yam, V. W.-W.; Au, V. K.-M.; Leung, S. Y.-L. Light-Emitting Self-Assembled Materials Based on D8 and D10 Transition Metal Complexes. Chem. Rev. 2015, 115, 7589-7728. 5. Li, K.; Ming Tong, G. S.; Wan, Q.; Cheng, G.; Tong, W.-Y.; Ang, W.-H.; Kwong, W.-L.; Che, C.-M. Highly Phosphorescent Platinum(ii) Emitters: Photophysics, Materials and Biological Applications. Chem. Sci. 2016, 7, 1653-1673. 6. Lu, C.-W.; Wang, Y.; Chi, Y. Metal Complexes with Azolate-Functionalized Multidentate Ligands: Tactical Designs and Optoelectronic Applications. Chem. Eur. J. 2016, DOI: 10.1002/chem.201601216. 7. Yersin, H.; Rausch, A. F.; Czerwieniec, R.; Hofbeck, T.; Fischer, T. The Triplet State of Organo-Transition Metal Compounds. Triplet Harvesting and Singlet Harvesting for Efficient Oleds. Coord. Chem. Rev. 2011, 255, 2622-2652. 8. Younker, J. M.; Dobbs, K. D. Correlating Experimental Photophysical Properties of Iridium(iii) Complexes to Spin–Orbit Coupled TDDFT Predictions. J. Phys. Chem. C 2013, 117, 25714-25723. 9. Koseki, S.; Kagita, Y.; Matsumoto, S.; Asada, T.; Yagi, S.; Nakazumi, H.; Matsushita, T. Theoretical Analyses on Phosphorescent Processes in Pt(thpy)2 and Its Derivatives. J. Phys. Chem. C 2014, 118, 15412-15421. 10. Wu, Y.; Shan, G.-G.; Li, H.-B.; Wu, S.-X.; Ren, X.-Y.; Geng, Y.; Su, Z.-M. Theoretical Study and Design of Multifunctional Phosphorescent Platinum(ii) Complexes Containing Triarylboron Moieties for Efficient Oled Emitters. Phys. Chem. Chem. Phys. 2015, 17, 2438-2446. 11. Chou, P.-T.; Chi, Y. Osmium and Ruthenium-Based Phosphorescent Materials: Design, Photophysics and Utilization in OLED Fabrication. Eur. J. Inorg. Chem. 2006, 3319-3332. 12. You, Y.; Nam, W. Photofunctional Triplet Excited States of Cyclometalated Ir(iii) Complexes: Beyond Electroluminescence. Chem. Soc. Rev. 2012, 41, 7061-7084. 13. Xiang, H.; Cheng, J.; Ma, X.; Zhou, X.; Chruma, J. J. Near-Infrared Phosphorescence: Materials and Applications. Chem. Soc. Rev. 2013, 42, 6128-6185. 14. Ladouceur, S.; Fortin, D.; Zysman-Colman, E. Role of Substitution on the Photophysical Properties of 5,5'-Diaryl-2,2'-bipyridine (bpy*) in [Ir(ppy)2(bpy*)]PF6 Complexes: A Combined Experimental and Theoretical Study. Inorg. Chem. 2010, 49, 5625-5641. 15. Werrett, M. V.; Muzzioli, S.; Wright, P. J.; Palazzi, A.; Raiteri, P.; Zacchini, S.; Massi, M.; Stagni, S. Proton-Induced Reversible Modulation of the Luminescent Output of Rhenium(i), Iridium(iii), and Ruthenium(ii) Tetrazolate Complexes. Inorg. Chem. 2014, 53, 229-243. 16. Abbott, G.; Brooks, R.; Rosenberg, E.; Terwilliger, M.; Ross, J. B. A.; Ichire, O. O. L. Surface-Bound Ruthenium Diimine Organometallic Complexes: Excited-State Properties. Organometallics 2014, 33, 2467-2478. 17. Unger, Y.; Zeller, A.; Taige, M. A.; Strassner, T. Near-UV Phosphorescent Emitters: N-Heterocyclic Platinum(ii) Tetracarbene Complexes. Dalton Trans. 2009, 4786. 18. Uesugi, H.; Tsukuda, T.; Takao, K.; Tsubomura, T. Highly Emissive Platinum(ii) Complexes Bearing Carbene and Cyclometalated Ligands. Dalton Trans. 2013, 42, 7396-7403. 19. Zhang, Y.; Clavadetscher, J.; Bachmann, M.; Blacque, O.; Venkatesan, K. Tuning the Luminescent Properties of Pt(ii) Acetylide Complexes through Varying the ‒ 30 ‒ ACS Paragon Plus Environment

Page 31 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Electronic Properties of N-Heterocyclic Carbene Ligands. Inorg. Chem. 2014, 53, 756-771. 20. Ahrens, S.; Strassner, T. Detour-Free Synthesis of Platinum-Bis-NHC Chloride Complexes, Their Structure and Catalytic Activity in the Ch Activation of Methane. Inorg. Chim. Acta 2006, 359, 4789-4796. 21. Wu, Y.; Wu, S.-X.; Li, H.-B.; Geng, Y.; Su, Z.-M. Forward Molecular Design for Highly Efficient OLED Emitters: A Theoretical Analysis of Photophysical Properties of Platinum(ii) Complexes with N-Heterocyclic Carbene Ligands. Dalton Trans. 2011, 40, 4480-4488. 22. Che, C. M.; He, L. Y.; Poon, C. K.; Mak, T. C. W. Solid-State Emission of Dicyanoplatinum(ii) and Palladium(ii) Complexes of Substituted 2,2'-Bipyridines and 1,10-Phenanthroline and X-Ray Crystal Structures of Isomorphous M(bpy)(cn)2 (bpy = 2,2'-Bipyridine; M = Pt, Pd). Inorg. Chem. 1989, 28, 3081-3083. 23. Kato, M.; Kosuge, C.; Morii, K.; Ahn, J. S.; Kitagawa, H.; Mitani, T.; Matsushita, M.; Kato, T.; Yano, S.; Kimura, M. Luminescence Properties and Crystal Structures of Dicyano(Diimine)Platinum(ii) Complexes Controlled by Pt∙∙∙Pt and π Interactions. Inorg. Chem. 1999, 38, 1638-1641. 24. Goeb, S.; Rachford, A. A.; Castellano, F. N. Solvent-Induced Configuration Mixing and Triplet Excited State Inversion Exemplified in a Pt(ii) Complex. Chem. Commun. 2008, 814-816. 25. Guo, H.; Muro-Small, M. L.; Ji, S.; Zhao, J.; Castellano, F. N. Naphthalimide Phosphorescence Finally Exposed in a Platinum(ii) Diimine Complex. Inorg. Chem. 2010, 49, 6802-6804. 26. Zhang, Y.; Garg, J. A.; Michelin, C.; Fox, T.; Blacque, O.; Venkatesan, K. Synthesis and Luminescent Properties of Cis Bis-N-Heterocyclic Carbene Platinum(ii) Bis-Arylacetylide Complexes. Inorg. Chem. 2011, 50, 1220-1228. 27. Rillema, D. P.; Cruz, A. J.; Moore, C.; Siam, K.; Jehan, A.; Base, D.; Nguyen, T.; Huang, W. Electronic and Photophysical Properties of Platinum(ii) Biphenyl Complexes Containing 2,2'-Bipyridine and 1,10-Phenanthroline Ligands. Inorg. Chem. 2013, 52, 596-607. 28. Nguyen, V. H.; Chew, H. Q.; Su, B.; Yip, J. H. K. Synthesis and Spectroscopy of Anionic Cyclometalated Iridium(iii)-Dithiolate and Sulfinates - Effect of Sulfur Dioxygenation on Electronic Structure and Luminescence. Inorg. Chem. 2014, 53, 9739-9750. 29. Chi, Y.; Wu, K.-L.; Wei, T.-C. Ruthenium and Osmium Complexes That Bear Functional Azolate Chelates for Dye-Sensitized Solar Cells. Chem. Asian J. 2015, 10, 1098-1115. 30. Hsu, C.-W.; Zhao, Y.; Yeh, H.-H.; Lu, C.-W.; Fan, C.; Hu, Y.; Robertson, N.; Lee, G.-H.; Sun, X.; Chi, Y. Efficient Pt(ii) Emitters Assembled from Neutral Bipyridine and Dianionic Bipyrazolate: Designs, Photophysical Characterizations and Fabrication of Non-Doped Oleds. J. Mater. Chem. C 2015, 3, 10837-10847. 31. Liao, J.-L.; Chi, Y.; Wang, J.-Y.; Chen, Z.-N.; Tsai, Z.-H.; Hung, W.-Y.; Tseng, M.-R.; Lee, G.-H. Pt(ii) Phosphors Featuring Both Dicarbene and Functional Biazolate Chelates: Synthesis, Luminescent Properties, and Applications in Organic Light-Emitting Diodes. Inorg. Chem. 2016, 55, 6394-6404. 32. Specht, Z. G.; Grotjahn, D. B.; Moore, C. E.; Rheingold, A. L. Effects of Hindrance in N-Pyridyl Imidazolylidenes Coordinated to Iridium on Structure and Catalysis. Organometallics 2013, 32, 6400-6409. 33. Riener, K.; Bitzer, M. J.; Pöthig, A.; Raba, A.; Cokoja, M.; Herrmann, W. A.; Kühn, F. E. On the Concept of Hemilability: Insights into a Donor-Functionalized Iridium(i) NHC Motif and Its Impact on Reactivity. Inorg. Chem. 2014, 53, 12767-12777. 34. Tsoureas, N.; Danopoulos, A. A. Pyridyland Diphenylphosphinoethyl-Functionalised N-Heterocyclic Carbene Platinum Methyl Complexes. J. Organomet. Chem. 2015, 775, 178-187. 35. Barbante, G. J.; Francis, P. S.; Hogan, C. F.; Kheradmand, P. R.; Wilson, D. J. D.; Barnard, P. J. Electrochemiluminescent Ruthenium(ii) N-Heterocyclic Carbene Complexes: A Combined Experimental and Theoretical Study. Inorg. Chem. 2013, 52, 7448-7459. ‒ 31 ‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 35

36. Monti, F.; Kessler, F.; Delgado, M.; Frey, J.; Bazzanini, F.; Accorsi, G.; Armaroli, N.; Bolink, H. J.; Ortí, E.; Scopelliti, R.; Nazeeruddin, M. K.; Baranoff, E. Charged Bis-Cyclometalated Iridium(iii) Complexes with Carbene-Based Ancillary Ligands. Inorg. Chem. 2013, 52, 10292-10305. 37. Zhang, F.; Ma, D.; Duan, L.; Qiao, J.; Dong, G.; Wang, L.; Qiu, Y. Synthesis, Characterization, and Photophysical and Electroluminescent Properties of Blue-Emitting Cationic Iridium(iii) Complexes Bearing Nonconjugated Ligands. Inorg. Chem. 2014, 53, 6596-6606. 38. Barbante, G. J.; Doeven, E. H.; Francis, P. S.; Stringer, B. D.; Hogan, C. F.; Kheradmand, P. R.; Wilson, D. J. D.; Barnard, P. J. Iridium(iii) N-Heterocyclic Carbene Complexes: An Experimental and Theoretical Study of Structural, Spectroscopic, Electrochemical and Electrogenerated Chemiluminescence Properties. Dalton Trans. 2015, 44, 8564-8576. 39. Zink, J. I. Triboluminescence. Acc. Chem. Res. 1978, 11, 289-295. 40. Kuchison, A. M.; Wolf, M. O.; Patrick, B. O. Conjugated Ligand-Based Tribochromic Luminescence. Chem. Commun. 2009, 7387-7389. 41. Marchetti, F.; Di Nicola, C.; Pettinari, R.; Timokhin, I.; Pettinari, C. Synthesis of a Photoluminescent and Triboluminescent Copper(i) Compound: An Experiment for an Advanced Inorganic Chemistry Laboratory. J. Chem. Ed. 2012, 89, 652-655. 42. Nakayama, H.; Nishida, J.-i.; Takada, N.; Sato, H.; Yamashita, Y. Crystal Structures and Triboluminescence Based on Trifluoromethyl and Pentafluorosulfanyl Substituted Asymmetric N-Phenyl Imide Compounds. Chem. Mater. 2012, 24, 671-676. 43. Peters, M.; Breinbauer, R. A Simple Synthesis of Functionalized 3-Methyl-1-Pyridinyl-1h-Imidazolium Salts as Bidentate N-Heterocyclic-Carbene Precursors and Their Application in Ir-Catalyzed Arene Borylation. Tetra. Lett. 2010, 51, 6622-6625. 44. Agarwal, J.; Shaw, T. W.; Stanton, C. J.; Majetich, G. F.; Bocarsly, A. B.; Schaefer, H. F. NHC-Containing Manganese(i) Electrocatalysts for the Two-Electron Reduction of Co2. Angew. Chem. Int. Ed. 2014, 53, 5152-5155. 45. Yeh, H.-H.; Ho, S.-T.; Chi, Y.; Clifford, J. N.; Palomares, E.; Liu, S.-H.; Chou, P.-T. Ru(ii) Sensitizers Bearing Dianionic Biazolate Ancillaries: Ligand Synergy for High Performance Dye Sensitized Solar Cells. J. Mater. Chem. A 2013, 1, 7681-7689. 46. Chen, J. C. C.; Lin, I. J. B. Palladium Complexes Containing a Hemilabile Pyridylcarbene Ligand. Organometallics 2000, 19, 5113-5121. 47. Bachmann, M.; Suter, D.; Blacque, O.; Venkatesan, K. Tunable and Efficient White Light Phosphorescent Emission Based on Single Component N-Heterocyclic Carbene Platinum(ii) Complexes. Inorg. Chem. 2016, 55, 4733-4745. 48. Nishiuchi, Y.; Takayama, A.; Suzuki, T.; Shinozaki, K. A Polymorphic Platinum(ii) Complex: Yellow, Red, and Green Polymorphs and X-Ray Crystallography of [Pt(fdpb)Cl] [Hfdpb = 1,3-bis(5-trifluoromethyl-2-pyridyl)benzene]. Eur. J. Inorg. Chem. 2011, 1815. 49. Cuerva, C.; Campo, J. A.; Ovejero, P.; Torres, M. R.; Oliveira, E.; Santos, S. M.; Lodeiro, C.; Cano, M. Columnar Discotic Pt(ii) Metallomesogens as Luminescence Multifunctional Materials with Chemo and Thermosensor Abilities. J. Mater. Chem. C 2014, 2, 9167-9181. 50. Yoshida, M.; Yashiro, N.; Shitama, H.; Kobayashi, A.; Kato, M. A Redox-Active Dinuclear Platinum Complex Exhibiting Multicolored Electrochromism and Luminescence. Chem. Eur. J. 2016, 22, 491-495. 51. Miskowski, V. M.; Houlding, V. H. Electronic Spectra and Photophysics of Platinum(ii) Complexes with a-Diimine Ligands Solid-State Effects. 2. Metal-Metal Interaction in Double Salts and Linear Chains. Inorg. Chem. 1991, 30, 4446-4452. 52. Connick, W. B.; Henling, L. M.; Marsh, R. E.; Gray, H. B. Emission Spectroscopic Properties of the Red Form of Dichloro(2,2'-bipyridine)platinum(ii). Role of Intermolecular Stacking Interactions. Inorg. Chem. 1996, 35, 6261-6265. 53. Xu, S.; Liu, T.; Mu, Y.; Wang, Y.-F.; Chi, Z.; Lo, C.-C.; Liu, S.; Zhang, Y.; Lien, A.; Xu, J. An Organic Molecule with Asymmetric Structure Exhibiting Aggregation-Induced Emission, Delayed Fluorescence, and Mechanoluminescence. Angew. Chem. Int. Ed. 2015, 54, 874-878. ‒ 32 ‒ ACS Paragon Plus Environment

Page 33 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

54. Lee, Y.-A.; Eisenberg, R. Luminescence Tribochromism and Bright Emission in Gold(I) Thiouracilate Complexes. J. Am. Chem. Soc. 2003, 125, 7778-7779. 55. Biju, S.; Gopakumar, N.; Bünzli, J. C. G.; Scopelliti, R.; Kim, H. K.; Reddy, M. L. P. Brilliant Photoluminescence and Triboluminescence from Ternary Complexes of Dyiii and Tb(iii) with 3-Phenyl-4-propanoyl-5-isoxazolonate and a Bidentate Phosphine Oxide Coligand. Inorg. Chem. 2013, 52, 8750-8758. 56. Balsamy, S.; Natarajan, P.; Vedalakshmi, R.; Muralidharan, S. Triboluminescence and Vapor-Induced Phase Transitions in the Solids of Methyltriphenylphosphonium Tetrahalomanganate(ii) Complexes. Inorg. Chem. 2014, 53, 6054-6059. 57. Chen, J.; Zhang, Q.; Zheng, F.-K.; Liu, Z.-F.; Wang, S.-H.; Wu, A. Q.; Guo, G.-C. Intense Photo- and Tribo-Luminescence of Three Tetrahedral Manganese(ii) Dihalides with Chelating Bidentate Phosphine Oxide Ligand. Dalton Trans. 2015, 44, 3289-3294. 58. Fontenot, R. S.; Bhat, K. N.; Owens, C. A.; Hollerman, W. A.; Aggarwal, M. D. Effects of Added Dibutyl Phosphate on the Luminescent Properties of Europium Tetrakis Dibenzoylmethide Triethylammonium. J. Lumin. 2015, 158, 428-434. 59. Xu, Q.; Heaton, B. T.; Jacob, C.; Mogi, K.; Ichihashi, Y.; Souma, Y.; Kanamori, K.; Eguchi, T. Hexacarbonyldiplatinum(i). Synthesis, Spectroscopy, and Density Functional Calculation of the First Homoleptic, Dinuclear Platinum(I) Carbonyl Cation, [{Pt(CO)3}2]2+, Formed in Concentrated Sulfuric Acid. J. Am. Chem. Soc. 2000, 122, 6862-6870. 60. Chang, S.-Y.; Kavitha, J.; Li, S.-W.; Hsu, C.-S.; Chi, Y.; Yeh, Y.-S.; Chou, P.-T.; Lee, G.-H.; Carty, A. J.; Tao, Y.-T.; Chien, C.-H. Platinum(ii) Complexes with Pyridyl Azolate-Based Chelates:  Synthesis, Structural Characterization, and Tuning of Photoand Electrophosphorescence. Inorg. Chem. 2006, 45, 137-146. 61. Lin, M.-S.; Yang, S.-J.; Chang, H.-W.; Huang, Y.-H.; Tsai, Y.-T.; Wu, C.-C.; Chou, S.-H.; Mondal, E.; Wong, K.-T. Incorporation of a Cn Group into Mcp: A New Bipolar Host Material for Highly Efficient Blue and White Electrophosphorescent Devices. J. Mater. Chem. 2012, 22, 16114-16120. 62. Tsai, W.-L.; Huang, M.-H.; Lee, W.-K.; Hsu, Y.-J.; Pan, K.-C.; Huang, Y.-H.; Ting, H.-C.; Sarma, M.; Ho, Y.-Y.; Hu, H.-C.; Chen, C.-C.; Lee, M.-T.; Wong, K.-T.; Wu, C.-C. A Versatile Thermally Activated Delayed Fluorescence Emitter for Both Highly Efficient Doped and Non-Doped Organic Light Emitting Devices. Chem. Commun. 2015, 51, 13662-13665. 63. Wang, Q.; Oswald, I. W. H.; Yang, X.; Zhou, G.; Jia, H.; Qiao, Q.; Chen, Y.; Hoshikawa-Halbert, J.; Gnade, B. E. A Non-Doped Phosphorescent Organic Light-Emitting Device with above 31% External Quantum Efficiency. Adv. Mater. 2014, 26, 8107-8113. 64. Kim, K.-H.; Liao, J.-L.; Lee, S. W.; Sim, B.; Moon, C.-K.; Lee, G.-H.; Kim, H. J.; Chi, Y.; Kim, J.-J. Crystal Organic Light-Emitting Diodes with Perfectly Oriented Non-Doped Pt-Based Emitting Layer. Adv. Mater. 2016, 28, 2526-2532. 65. Tsai, Y.-T.; Chen, C.-Y.; Chen, L.-Y.; Liu, S.-H.; Wu, C.-C.; Chi, Y.; Chen, S. H.; Hsu, H.-F.; Lee, J.-J. Analyzing Nanostructures in Mesogenic Host–Guest Systems for Polarized Phosphorescence. Org. Electron. 2014, 15, 311-321. 66. Lin, Y.-H.; Tsai, Y.-T.; Wu, C.-C.; Tsai, C.-H.; Chiang, C.-H.; Hsu, H.-F.; Lee, J.-J.; Cheng, C.-Y. Comparative Study of Spectral and Morphological Properties of Blends of P3ht with Pcbm and Icba. Org. Electron. 2012, 13, 2333-2341. 67. Kuei, C.-Y.; Tsai, W.-L.; Tong, B.; Jiao, M.; Lee, W.-K.; Chi, Y.; Wu, C.-C.; Liu, S.-H.; Lee, G.-H.; Chou, P.-T. Bis-Tridentate Ir(iii) Complexes with Nearly Unitary Rgb Phosphorescence and Organic Light-Emitting Diodes with External Quantum Efficiency Exceeding 31%. Adv. Mater. 2016, 28, 2795-2800. 68. Jiao, M.; Lu, C.-Y.; Lee, W.-K.; Chen, C.-Y.; Wu, C.-C. Simple Planar Indium-Tin-Oxide-Free Organic Light-Emitting Devices with Nearly 39% External Quantum Efficiency. Adv. Opt. Mater. 2016, 4, 365-370. 69. Lin, T.-A.; Chatterjee, T.; Tsai, W.-L.; Lee, W.-K.; Wu, M.-J.; Jiao, M.; Pan, K.-C.; Yi, C.-L.; Chung, C.-L.; Wong, K.-T.; Wu, C.-C. Sky-Blue Organic Light Emitting Diode with 37% External Quantum Efficiency Using Thermally Activated Delayed Fluorescence from Spiroacridine-Triazine Hybrid. Adv. Mater. 2016, 28, 6976-6983. ‒ 33 ‒ ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 35

70. Liu, S.-H.; Lin, M.-S.; Chen, L.-Y.; Hong, Y.-H.; Tsai, C.-H.; Wu, C.-C.; Poloek, A.; Chi, Y.; Chen, C.-A.; Chen, S. H.; Hsu, H.-F. Polarized Phosphorescent Organic Light-Emitting Devices Adopting Mesogenic Host–Guest Systems. Org. Electron. 2011, 12, 15-21.

‒ 34 ‒ ACS Paragon Plus Environment

Page 35 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

TOC Illustration:

‒ 35 ‒ ACS Paragon Plus Environment