Dual-phosphorescent Iridium(III) Complexes Extending Oxygen

1 hour ago - ... Organic Pigments, Fluorescent Brighteners, and Photographic Sensitizers · Industrial Carbohydrates · Industrial Organic Chemicals, Le...
3 downloads 0 Views 1MB Size
Subscriber access provided by Kaohsiung Medical University

Article

Dual-phosphorescent Iridium(III) Complexes Extending Oxygen Sensing from Hypoxia to Hyperoxia Kenneth Yin Zhang, Pengli Gao, Guanglan Sun, Taiwei Zhang, Xiangling Li, Shujuan Liu, Qiang Zhao, Kenneth Kam-Wing Lo, and Wei Huang J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.8b02492 • Publication Date (Web): 06 Jun 2018 Downloaded from http://pubs.acs.org on June 6, 2018

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 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 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.

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 9 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

Journal of the American Chemical Society

Dual-phosphorescent Iridium(III) Complexes Extending Oxygen Sensing from Hypoxia to Hyperoxia Kenneth Yin Zhang,† Pengli Gao,† Guanglan Sun,† Taiwei Zhang,† Xiangling Li,† Shujuan Liu,† Qiang Zhao,*,† Kenneth Kam-Wing Lo,*,‡ and Wei Huang*,†,§ †

Key Laboratory for Organic Electronics and Information Displays & Jiangsu Key Laboratory for Biosensors, Institute of Advanced Materials (IAM), Jiangsu National Synergetic Innovation Center for Advanced Materials (SICAM), Nanjing University of Posts & Telecommunications, 9 Wenyuan Road, Nanjing 210023, P. R. China ‡

Department of Chemistry, City University of Hong Kong, Tat Chee Avenue, Hong Kong, P. R. China

§

Key Laboratory of Flexible Electronics and Institute of Advanced Materials, Jiangsu National Synergetic Innovation Center for Advanced Materials, Nanjing Tech University, 30 South Puzhu Road, Nanjing 211816, P. R. China Supporting Information Placeholder

ABSTRACT: Hypoxia and hyperoxia, referring to states of biological tissues in which oxygen supply is in-sufficient and too much, respectively, are often pathological conditions. Many luminescent oxygen probes have been developed for imaging intracellular and in vivo hypoxia but their sensitivity toward hyperoxia becomes very low. Here we report a series of iridium(III) complexes in which limited internal conversion between two excited-states results in dual phosphorescence from two different excited-states upon excitation at a single wavelength. Structural manipulation of the complexes allows rational tuning of the dual-phosphorescence properties and the spectral profile response of the complexes toward oxygen. By manipulating the efficiency of internal conversion between the two emissive states, we obtained a complex exhibiting naked-eye distinguishable green, orange, and red emission in aqueous buffer solution under an atmosphere of N2, air, and O2, respectively. This complex is used for intracellular and in vivo oxygen sensing not only in hypoxia region, but also in normoxia and hyperoxia intervals. To the best of our knowledge, this is the first example of using a molecular probe for simultaneous bioimaging of hypoxia and hyperoxia.

INTRODUCTION Molecular oxygen is important and essential to sustaining life. It is consumed during cellular respiration and takes part in producing adenosine triphosphate (ATP) during oxidative phosphorylation in mitochondria.1,2 The oxygen contents in healthy organs, tissues, and cells are kept in a certain interval. Hypoxia and hyperoxia, referring to states in which oxygen supply is in-sufficient and too much, respectively, are often pathological conditions. The former is one of the most important features of many diseases including solid tumors,3-5 inflammatory diseases,6,7 and cardiac ischemia,7,8 whereas the latter can cause an increased level of reactive oxygen/nitrogen species (RONS)9-11 and oxygen toxicity syndrome,12,13 leading to damage to lungs,14-16 eyes,17-20 and the central nervous system.21 Therefore, the accurate detection of oxygen contents in biological environments is of great importance to early diagnosis of these diseases. Many efforts have been made to develop luminescent probes for intracellular oxygen.22,23 Current achievement mainly includes two families of probes that are fluores-

cent organic dyes24-27 and phosphorescent transitionmetal complexes.28-30 Organic dyes are designed to response to specific reductases irreversibly under hypoxic conditions giving rise to fluorescence change, while transition-metal complexes are used for detection and imaging of oxygen contents in cell and in vivo owing to the efficient phosphorescence quenching of the complexes in triplet excited states by oxygen via energy transfer. Both types of probes exhibit fast and sensitive luminescence response toward hypoxia, but the sensing sensitivity toward high oxygen content significantly drops down. It is a great challenge to design probes that show sensitive responses toward both hypoxia and hyperoxia. The development of probes for hyperoxia has not been reported till now. Under hyperoxic conditions, the reductases are inactive so that fluorescent organic oxygen probes do not work.24-27 During the development of phosphorescent oxygen probes, most efforts have been put on increasing the sensing sensitivity, improving the biocompatibility, and red-shifting the phosphorescence wavelength of the probes for in vivo biological applications.29-35 For these turn-off type probes, high sensitivity ensures

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

good sensing performance toward hypoxia, but, in the meantime, indicates reduced sensing ability toward hyperoxia because the phosphorescence becomes weak under normoxia (normal oxygen level) conditions. Increased excitation power is usually required to obtain trustworthy phosphorescence signals under hyperoxia conditions. In this respect, wavelength-ratiometric probes are good candidates for hyperoxia sensing. The intensity ratio of two wavelengths is used to reflect the oxygen content. Thus, the influence from the inherent variability in incident laser power and the uncertainty of intracellular probe concentration is minimized.36 Current design strategy of ratiometric probes relies on incorporation of an additional non-sensitive fluorescent compound into phosphorescent probes as an internal standard, and the phosphorescence/fluorescence ratio is used for oxygen sensing.37-41 Even so, the utilization of probes based on the covalent or noncovalent assembly of phosphorescent and fluorescent constituents for oxygen sensing is still focused on the hypoxia region, while there is no study extending to hyperoxia sensing because of lack of suitable probes. Compared to the bi-component probes, single component small molecular dual-emissive luminophores are easy to synthesize. Their small molecular size also facilitates cellular internalization. Additionally, the exclusion of uncontrollable intramolecular energy/electron transfer will improve the accuracy of sensing. The development of single-component dual-emissive small-molecular luminophores is slow because, as stated by Kasha’s rule, photon emission occurs only from the lowest excited state.42 In principle, dual-emissive probes should fulfill two basic requirements, which are rich excited-state characters and limited electronic communication between the states, to allow possible radiative decays occurring from two different states. Fraser reported luminescent polymers that displayed fluorescence and phosphorescence simultaneously under hypoxia condition at room temperature.43 The phosphorescence was sensitively quenched by O2 rendering ratiometric response toward trace oxygen. Transition-metal complexes display longlived phosphorescence involving both metal- and ligandcentered orbitals, resulting in a diversity of emissive-state characters.44 Phosphorescent iridium(III) polypyridine complexes of a general formula [Ir(N^C)2(N^N)](PF)6 coordinated with two types of ligands exhibit intense phosphorescence from triplet metal/ligand-to-ligand charge-transfer (3CT) or intraligand (3IL) excited states depending on the electronic structures of the ligands.45,46 Incorporation of functional groups baring unconjugated lone pair electrons, such as amino groups, into the structure of the complexes provides occupied lone pair orbitals, giving rise to additional electronic transitions to unoccupied π* orbitals, which may interrupt the internal conversion of the excited complex to the lowest excited states, resulting in dual phosphorescence from two independent states.47-49 However, a direct link between the chemical structures of the iridium(III) complexes and the occurrence of dual phosphorescence has not yet been established, which is essential to further the rational de-

sign of single component dual-emissive small molecular probes that display wavelength-ratiometric response toward specific analytes. In this work, we designed and synthesized a series of iridium(III) complexes containing aminomethyl substituted phenylpyridine ligands, and investigated their dualphosphorescence properties. The large difference in emission wavelengths and lifetimes of the two phosphorescence bands rendered these complexes single-component small-molecule-based ratiometric sensors for molecular oxygen. Structural manipulation of the complexes allowed rational tuning of the internal-conversion efficiency and thus the dual-phosphorescence properties and the spectral profile response of the complexes toward oxygen. Interestingly, one of the piperidinomethyl substituted complexes exhibited naked-eye distinguishable green, orange, and red emission in aqueous buffer solution under an atmosphere of N2, air, and O2, respectively. This complex was further used for simultaneously imaging both hypoxia and hyperoxia in living cells, zebrafish, and mice with distinct spectral responses.

RESULTS AND DISCUSSION We have recently unexpectedly developed a series of iridium(III) complexes, including complex 1c (Figure 1), which exhibit spectrum-resolved dual phosphorescence.46,48 The high energy (HE) and low energy (LE) phosphorescence bands have been assigned to 3IL and 3CT excited states, respectively. Herein we aim to establish a relation between the chemical structures of the iridium(III) complexes and the occurrence of dual phosphorescence, and to develop novel ratiometric probes for both hypoxia and hyperoxia based on the dualphosphorescent iridium(III) complexes. According to Stern-Volmer Equation,50 ‫ ݋ܫ‬τ‫݋‬ = = 1 + ݇‫ ݍ‬τ‫ ݋‬ሾO2 ሿ ‫ܫ‬ τ where Io and I are phosphorescence intensities in the absence and presence of O2, respectively; τo and τ are lifetimes of the emissive excited state in the absence and presence of O2; kq is the quenching rate coefficient, the quenching efficiency (KSV = kqτo, Stern-Volmer constant) is proportional to the phosphorescence lifetime. A longer excited-state lifetime is in favor of the non-radiative quenching. Hence, the different phosphorescence lifetimes of the HE and LE bands of complex 1c allowed ratiometric luminescence response toward O2. The phosphorescence spectra of complex 1c were recorded in aqueous phosphate buffered saline (PBS)/methanol (9:1, v/v) under an atmosphere of N2, air, and O2 and the phosphorescence lifetimes of the HE and LE were measured. To highlight the spectrum-profile response toward O2, the luminescence spectra were normalized by dividing by the number of total photon emitted, viz., integral of the spectrum, and they are shown in Figure 2. Complex 1c exhibited a HE emission band at ca. 520 nm with a LE shoulder at ca. 620 nm under N2. The emission lifetimes at the two wavelengths were determined to be 2.5 µs and 24 ns,

ACS Paragon Plus Environment

Page 2 of 9

N2 Air O2

550

600

650

700

750

500

550

Wavelength / nm

600

650

550

750

500

550

600

650

600

Wavelength / nm

650

700

750

600

600

Air

650

650

700

750

Wavelength / nm

2d

O2 Luminescence Intensity

550

550

2c N2

500

500

Wavelength / nm

2b Normalized Luminescence Intensity

2a

500

700

Wavelength / nm

1d

Luminescence Intensity

500

1c Normalized Luminescence Intensity

1b Normalized Luminescence Intensity

Normalized Luminescence Intensity

1a

Normalized Luminescence 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

Journal of the American Chemical Society

Normalized Luminescence Intensity

Page 3 of 9

500

550

600

650

Wavelength / nm

Wavelength / nm

700

750

500

550

600

650

700

750

Wavelength / nm

Figure 1. Structures of the iridium(III) complexes and their photoluminescence spectra in PBS/CH3OH (9:1, v/v) under N2 (blue), air (green), and O2 (red) atmospheres. Insets are the corresponding photographs of complex 2b upon excitation by a UV lamp.

respectively, indicating that their sensitivities toward oxygen quenching were different from each other by more than 100 times. As expected, the HE emission was significantly quenched in aerated solution, resulting in predominance of a broad LE band. The HE emission lifetime was shortened to 0.45 µs while that of the LE luminescence kept to be 24 ns. Bubbling neat O2 into the solution led to further shortening of the lifetime of HE emission to 0.14 µs, while the LE emission lifetime was slightly shortened to 20 ns. However, the spectral difference was not significant because the HE band was embedded into the LE emission even under ambient air. We aim to tune the dual phosphorescence properties of the complex via structural manipulation to enable target complexes to exhibit remarkable spectral response toward both hyperoxia and hypoxia. To investigate the effect of the amino groups in the cyclometalating ligands on the dual phosphorescence properties and thus the spectral response toward O2, we firstly changed the number of substituents bound to the nitrogen atom of the amino group, synthesizing three complexes 1a, 1b, and 1d, which do not contain an amino group or contain primary and tertiary amino groups, respectively (Figure 1). All these complexes have been characterized by 1H NMR, 13C NMR, matrix-assisted laser desorption ionization time-of-flight (MALDI-TOF) mass spectrometry (MS), electrospray ion-

ization (ESI) high-resolution MS, IR, and UV-Vis absorption spectroscopy (see in the Supporting Information). Normalized luminescence spectra of these complexes in PBS/methanol (9:1, v/v) were recorded under an atmosphere of N2, air, and O2 and shown in Figure 1. The photoluminescence quantum yields were listed in Table S1. Interestingly, only the tertiary-amine-containing complex 1d, similar to complex 1c, displayed dual phosphorescence upon photoexcitation and O2 quenching-induced luminescence spectral response. Both the amine-free and the primary-amine-containing complexes 1a and 1b exhibited intensity quenching of the single broad LE band in the presence of O2 without wavelength shift. Theoretical time-dependent density functional theory (TD-DFT) calculation revealed that the frontier orbitals contributable to the emissive states are quite similar in energy for all the four complexes (Figure 2 and S1). Since both triplet states TIL and TCT originating from HOMO to LUMO+2 and LUMO, respectively, are present with similar energy for all the complexes 1a − 1d, the occurrence of dual phosphorescence depends on the efficiency of internal conversion from TIL to TCT. The most significant difference among these complexes are the occupied orbital localized at the lone pair of the amino groups, which is absent for complex 1a, occurs at a low-energy level (-6.59 eV, HOMO-5) for complex 1b, becomes higher in energy for complex 1c (-6.02 eV, HOMO-1), and moderately

ACS Paragon Plus Environment

Journal of the American Chemical Society 1a

1b

1c

1d

2a

2b

-1

-1.67 eV

L+2

-1.69 eV

L+2

-1.70 eV

L+2

-1.72 eV

L+2

L+3

-1.72 eV

-1.74 eV

L+3

-2

-2.50 eV

-2.51 eV Energy / eV

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 9

-3

-2.51 eV

L

L

H

H

-2.51 eV L

-2.70 eV

-2.69 eV L

L

L

-4

-5

-5.56 eV

-5.55 eV

-5.56 eV

H

-5.46 eV

H

-5.60 eV

-5.49 eV

-5.45 eV -5.48 eV

H

-6.04 eV

-6

H H-1

-5.67 eV

-6.02 eV

-6.59 eV

H-2

-7

H-1 H-1 H-5

H-1

TCT :

T1 H → L (98.6%)

T1 H → L (98.5%)

T1 H → L (98.0%)

T1 H → L (67.5%)

T1 H → L (98.3%)

TIL :

T4 H → L+2 (71.8%)

T3 H → L+2 (50.1%)

T3 H → L+2 (66.3%)

T5 H → L+2 (50.7%)

T4 H → L+3 (72.2%)

T1 H−2 → L (48.5%) T6 H−2 → L+3 (31.9%)

TNLCT:

not applicable

not observed until T10

T4 H−1 → L (31.7%)

T2 H−1 → L (72.2%)

T7 H−2 → L (86.9%)

T2 H−1 → L (99.5%)

Figure 2. Selected TD-DFT-calculated energy levels, isosurface plots and triplet transitions of complexes 1a − 1d, 2a, and 2b.

contributes to the HOMO for complex 1d (Figure 2). The high energy levels of the amine lone pairs of complexes 1c and 1d give rise to additional 3NLCT (n → π*) transitions (Figure 2) which are probably the cause of limited electronic communications between TIL and TCT states. Although complexes 1c and 1d exhibit sensitive and efficient spectral response toward hypoxia, the huge difference in luminescence lifetimes between HE TIL and LE TCT states rendering highly efficient oxygen quenching of the HE TIL emissive state limits the sensitivity of the luminescence spectrum profile toward hyperoxia. To reduce the quenching sensitivity of the HE phosphorescence by oxygen, complexes 2a and 2b (Figure 1) were designed, synthesized, and characterized. We expected that the additional electron-withdrawing amide substituent in the diimine ligand should stabilize the π* orbital of the diimine ligand, hence facilitating internal conversion from the TIL to the TCT states, shortening the excited-state lifetime of the HE phosphorescence, and thus reducing the quenching efficiency. Theoretical TD-DFT calculation confirmed a lower π*-orbital (LUMO) energy of the diimine ligands (about -2.70 eV) for both complexes compared to complexes 1a − 1d (about -2.50 eV, Figure 2). The energy of π orbitals of the cyclometalating ligands is also slightly reduced owing to the inductive effect along the trans N−Ir−C bonds. Interestingly, the secondary-aminesubstituted complex 2a only exhibited a broad LE emission band at 620 nm even under 100% N2 atmosphere upon photoexcitation (Figure 1). The LE phosphorescence was slightly quenched by O2 owing to the short lifetime (τo = 31 ns). In contrast, the tertiary-amine-substituted complex 2b exhibited dual phosphorescence with comparable HE and LE intensities in PBS under ambient air, giving an orange phosphorescence color (Figure 1). The occurrence of dual phosphorescence of complex 2b is probably due to the presence of an 3NLCT (n → π*) (T2, 2.12 eV) state inhibiting the communication between the 3 IL (T6, 2.65 eV) and 3CT (T1, 1.79 eV) states (Figure 2). For complex 2a, the 3NLCT (T7, 2.78 eV) is much lower in energy than both 3IL (T4, 2.67 eV) and 3CT (T1, 1.76 eV)

states, and thus does not disturb their communication. The comparable intensity of the HE and LE luminescence bands of complex 2b under normoxia conditions renders the complex a potential candidate for both hyperoxia and hypoxia sensing. The lifetimes of the HE and LE phosphorescence of complex 2b were 0.36 µs and 36 ns, respectively, under ambient air. Bubbling pure N2 and O2 into the solution tuned the phosphorescence colors sharply to green and red, respectively (Figure 1). The lifetime of HE phosphorescence was correspondingly elongated and shortened while that of the LE one was negligibly altered. B A

π*(L1)

π*(L2)

π*(L2)

L1

π(L1) amine

III

TIL L2

amine TNLCT

dπ(Ir) TCT

C

TIL TNLCT

TCT S0

Figure 3. (A) Structure skeleton of a potential dual phos-

phorescent iridium(III) complex. (B) Energy diagrams showing molecular orbitals involved in IL, NLCT, and CT transitions. (C) Energy diagrams of ground and excited states showing that NLCT state interrupts the internal conversion, resulting in dual phosphorescence from both 3 CT and 3IL states. Taken together, we tentatively conclude that, in the series of iridium(III) complexes we studied, the nonconjugated amino groups give rise to additional electronic transitions from the lone pairs to the π* orbitals of the diimine ligands, leading to new 3NLCT excited states. Once the energy of the new state stands between those of the traditional emissive states (3IL and 3CT), the internal conversion from the HE state to the LE one is inhibited, resulting in occurrence of dual phosphorescence from two independent emissive states (Figure 3). This is important and helpful to furthering the rational design of single

ACS Paragon Plus Environment

Page 5 of 9

component dual-emissive small molecular probes that display wavelength-ratiometric response toward specific analytes. Owing to the naked-eye distinguishable emission colors under different atmospheres, we further record the phosphorescence spectra of complex 2b under different oxygen contents in detail. The normalized luminescence spectra highlighting the spectral response and the lifetimes of HE and LE phosphorescence have been shown in Figure 4. A good and exciting spectral sensitivity was observed in both hypoxia and hyperoxia regions, and has been ascribed to the different quenching efficiencies of the HE and LE phosphorescence by triplet molecular oxygen. Linear Stern-Volmer plots were obtained (Figure 4), indicating that the internal conversion efficiency depends on the chemical structure of the complex only and that it is independent of the oxygen contents. The quenching sensitivities, i.e., the KSV values, of the HE and LE phosphorescence were 0.15 and 0.0031 %−1, respectively (Figure 4). The 48 times gap in the KSV values was in the same order of magnitude as the 44 times difference in phosphorescence lifetimes in the absence of O2. To further confirm that the ratiometric spectral response of complex 2b toward O2 was due to the different quenching efficiency of the two excited emissive states, we synthesized two model complexes 2c and 2d, in which the diimine and cyclometalating ligands were replaced by acetylacetone and 2-(4-methylphenyl)pyridine, respectively (Figure 1). Upon photoexcitation, complexes 2c and 2d, respectively, exhibited single HE and LE phosphorescence at 516 and 608 nm with lifetimes of 0.72 µs and 33.4 ns. The oxygen quenching efficiency of complex 2c was remarkably higher than that of complex 2d (Figure 1). The KSV values for O2 quenching of complexes 2c and 2d were determined to be 0.13 and 0.0042 %−1, respectively, which were similar to the values for the HE and LE phosphorescence of the dual-emissive complex 2b. Hypoxia

0% − 12% O2

Phosphorescence lifetimes of complex 2b / ns O2 / %

HE

LE

O2 / %

HE

LE

0

1523

38.5

21 (air)

362

35.7

Normoxia 15% − 35% O2

2

1162

37.8

30

282

34.9

5

901

37.0

50

179

33.7

Hyperoxia 40% − 100% O2

10

610

36.6

80

117

30.6

15

479

36.0

100

94

29.6

any viability lose in our experimental conditions during the following experiments (Figure S2). Phosphorescence pH titration ensured that the spectral response of complex 2b maintained high sensitivity in physiology pH range (Figure S3). The phosphorescence spectra recorded at different temperatures (10 − 50 °C) showed that increasing temperature reduced the phosphorescence intensity, but did not change the spectral profile much (Figure S4). The stability of the complex in serum solution has also been evaluated. In the presence of fetal bovine serum (FBS), the phosphorescence spectral profile of complex 2b in PBS buffer did not exhibited noticeable change and the sensitive profile response toward both hypoxia and hyperoxia was maintained (Figure S5). HeLa, HepG2, and 3T3 cells incubated with complex 2b (5 µM) for 2 h showed substantial cytoplasmic staining (Figure 5). The phosphorescence signals collected from the green (520 ± 20 nm) and red (620 ± 20 nm) channels were similar in intensity. Interestingly, when the cells were cultured under low (5%) and high (50%) oxygen environments, the green phosphorescence was much brighter or dimmer than the red one, respectively (Figure 5). It is mentionable that in order to highlight the spectral-profile response, the images of cells under different culture atmospheres were taken under most suitable but different excitation laser powers. Thus, we cannot analyze the signals from the green or red channels independently. It is also an important and attractive advantage of dual emissive probes that the spectral-profile or the intensity ratio of two luminescence band is independent of the excitation laser power and probe concentration. Additionally, the phosphorescence spectra of the internalized complex 2b were recorded via a λ-scanning mode of photoluminescence confocal microscopy. As shown in Figure 5, the cytoplasmic complex displayed a broad emission band covering 480 − 680 nm when the three types of cells were cultured under ambient atmosphere. The green and red phosphorescence became predominant when the cells were cultured in hypoxia and hyperoxia conditions. This is the first demonstration that a small molecular probe showed sensitive spectral response toward both intracellular hypoxia and hyperoxia.

15

HepG2

y = 0.15x + 1 R2 = 1.00

5% O2

Air

HeLa 50% O2

5% O2

3T3

Air

50% O2

5% O2

Air

50% O2

10 520 ± 20 nm

τo / τ

20

40 60 80 Oxygen Content / %

620 ± 20 nm

100

Figure 4. Phosphorescence spectral response and phosphorescence lifetimes of complex 2b in PBS/CH3OH (9:1, v/v) toward 0 to 100% O2, and Stern-Volmer plots of the HE (green) and LE (red) phosphorescence toward oxygen quenching.

The highly sensitive phosphorescence spectra of complex 2b toward O2 encouraged us to use it for detection of both hypoxia and hyperoxia in living cells. The 3-(4,5dimethyl-2-thiazolyl)-2,5-diphenyltetrazolium bromide (MTT) assay confirmed that the complex did not cause

500

550

600

Wavelength / nm

650

500

550

600

Wavelength / nm

650

Luminescence Intensity (A. U.)

0 0

y = 0.0031x + 1 R2 = 0.97

Luminescence Intensity (A. U.)

5

Luminescence 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

Journal of the American Chemical Society

500

550

600

650

Wavelength / nm

Figure 5. Laser-scanning luminescence microscopy images and intracellular phosphorescence spectra of HepG2, HeLa, and 3T3 cells incubated with complex 2b (5 µM, 2 h, 37 °C) cultured under 5% (blue), 21% (green), and 5% (red) oxygen conditions. Scale bar: 30 µm.

ACS Paragon Plus Environment

Journal of the American Chemical Society To demonstrate real-time intracellular oxygen sensing, we designed the following experiment. A coverslip was used to cover the complex 2b-loaded HepG2 cells, and 50% O2 was sustainably and slowly bubbled into the culture dish (Figure 6). Thus, the oxygen content became higher and higher in the dish, but became lower and lower under the coverslip because of consumption by cellular respiration and limited gas circulation. Before bubbling oxygen, the cells display green and red phosphorescence with comparable intensities. As bubbling proceeded, the HE phosphorescence of the cells under the coverslip became dominant while the uncovered cells exhibited predominant red phosphorescence (Figure 6). The cells at the edge of the coverslip maintained green and red dualphosphorescence. The phosphorescence spectra of the covered and uncovered cells and those at the edge of the coverslip have also been recorded and shown in Figure 6, demonstrating remarkable spectral response of complex 2b toward intracellular hypoxia and hyperoxia. 620 ± 20 nm

520 ± 20 nm

Luminescence overlaid

lived autofluorescence was completely filtered off when a time-gate of 5 ns was applied between photoexcitation and signal collection (Figure 7). The response of the phosphorescence signals from the head area toward O2 was analyzed. Bubbling 5% and 50% O2 into the culture media of the zebrafish did not cause noticeable effect on the red phosphorescence (620 ± 20 nm), but, respectively, enhanced and quenched the green phosphorescence (520 ± 20 nm). The phosphorescence spectra clearly showed the sensitive profile response toward in vivo hypoxia and hyperoxia (Figure 7). 520 ± 20 nm

620 ± 20 nm

Time-gated Photoluminescence photoluminescence imaging lifetime imaging

Normoxia

Overlaid

Bright field

covered Bubbling 50% O2 to induce in vivo hyperoxia

0 min

uncovered

Hyperoxia

Bubbling 5% O2 to induce in vivo hypoxia

10 min

Hypoxia

500

550

600

650

Wavelength / nm

20 min

Figure 7. Laser-scanning luminescence microscopy images, photoluminescence lifetime image, and time-gated photoluminescence image (gated time = 5 ns) of zebrafish injected with complex 2b (1 µM, 50 pL) at the head area, and laserscanning luminescence microscopy images and in vivo luminescence spectra of the complex 2b-injected zebrafish under normoxia, hyperoxia, and hypoxia conditions.

c

b a

Luminescence Intensity (A. U.)

30 min

c A

a

In ambient condition for 4 h

Solid tumor bearing mouse

D

High

b

520 ± 20 nm

500

550

600

650

Wavelength / nm

B

620 ± 20 nm

520 ± 20 nm

Low

Intensity analysis

E

Complex 2b has further been used for oxygen monitoring in zebrafish. The complex (10 µM, 50 pL) was injected into 4-days old zebrafish at the head. Upon photoexcitation, the whole zebrafish was emissive intensely (Figure 7). Photoluminescence lifetime imaging indicated that the long-lived phosphorescence signals were only localized in the head area while intense short-lived autofluorescence was observed from the belly (Figure 7). Owing to the long-lived phosphorescence of complex 2b, the short-

520 ± 20 nm

C

620 ± 20 nm

After breathing 10% oxygen for 4 h

I520 ± 20 nm I620 ± 20 nm I620 nm/I520 nm

Luminescence Intensity

Figure 6. Laser-scanning luminescence microscopy images of complex 2b loaded HepG2 cells covered at the top righthand corner by a coverslip upon bubbling gas mixture of 50% O2, 45% N2, and 5% CO2 for different times and intracellular phosphorescence spectra of cells in selected areas after bubbling of the gas mixture for 30 min.

620 ± 20 nm

After breathing pure oxygen for 4 h

2.0

1.5

1.0

Intensity ratio

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 6 of 9

0.5

0.0 Ambient condition 520 ± 20 nm

Breathing pure O2

Breathing 10% O2

Solid tumor

Normal tissue

620 ± 20 nm

Figure 8. In vivo images of the complex 2b-injected mouse under ambient condition for 4 h (A), after breathing pure oxygen (B) and 10% oxygen (C) for 4 h. In vivo images of the HepG2 tumor bearing mouse injected with complex 2b into the tumor (circled) and normal tissue (uncircled) (D). Bar chart showing the intensity analysis of signals from images A − D (E). Error bars represent the standard deviations of six independent measurements.

ACS Paragon Plus Environment

Page 7 of 9 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

Journal of the American Chemical Society Furthermore, complex 2b was used to sense hypoxia and hyperoxia in living mice. The complex (50 µM, 50 µL) was intravenously injected into a male nude mouse. Imaging was performed under excitation at 420 nm and the luminescence at HE (520 ± 20 nm) and LE (620 ± 20 nm) was extracted from autofluorescence via spectra unmixing for intensity analysis. The HE and LE intensities of the complex were similar when the mouse was in ambientcondition (Figure 8). The LE/HE ratio kept almost unchanged in at least 4 h. When letting the 2b-injected mouse breathe in a pure oxygen atmosphere for 4 h, the HE intensity became 60% of the LE intensity and the intensity ratio of LE/HE was remarkably increased to about 1.68 (Figure 8), indicative of a high sensitivity of complex 2b toward in vivo hyperoxia. After letting the mouse breathe in a non-lethal low-oxygen atmosphere (10%), the LE/HE ratio became as low as 0.63, demonstrating the sensitivity of complex 2b toward in vivo hypoxia. Since hypoxia is one of the most important characteristics of solid tumor, complex 2b was used to identify the tumor from normal tissue owing to its spectral response toward hypoxia. The complex was intratumorally injected into the intravenous HepG2 tumor of a male nude mouse and intravenously injected into normal tissue of the same mouse. In the normal tissue, the HE and LE intensities of the complex were similar (Figure 8). In the solid tumor, both the HE and LE emission was reduced in intensity due to the deeper injection depth. Interestingly, the HE intensity is 1.62 fold of the LE intensity in the solid tumor (Figure 8), demonstrating the high sensitivity of complex 2b toward in vivo hypoxia. It is noteworthy that errors in the luminescence ratio (blue bars) is obviously smaller than those in luminescence intensity (green and red bars), demonstrating the advantage of ratiometric sensing.

dual-emissive probes allows naked-eye recognizable color change, which is independent of probe concentration, excitation power, or tissue depth. By tuning the sensitivity of the two excited states, we can broaden the response interval toward the specific analyte. For example, we demonstrated the utilization of rationally designed dual phosphorescent complex 2b for intracellular and in vivo oxygen sensing not only in hypoxia region, but also in normoxia and hyperoxia intervals. To the best of our knowledge, this is the first example that a small molecular probe showed sensitive spectral response toward both hypoxia and hyperoxia. The development of dual-emissive complexes with two red or infrared phosphorescence bands for deep hypoxia and hyperoxia sensing is underway. Apart from what has been demonstrated in this work, incorporation of two recognition units into the dual phosphorescent skeleton is possible to allow the two emissive states sensitive toward multiple analytes simultaneously. We anticipate wide application prospects of the small-molecular dual-phosphorescent probes in bioimaging, biosensing and disease diagnosis.

ASSOCIATED CONTENT Supporting Information. Synthesis, characterization, computational details, experimental information, and additional figures. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author *[email protected] *[email protected] *[email protected]

Notes The authors declare no competing financial interest.

CONCLUSION Although the photophysical properties of transition-metal complexes have been extensively studied and their applications in biological sensing and imaging have been widely investigated, the knowledge of dual phosphorescence has not been systematically built up and the application of dual-phosphorescent complexes in biosensing and bioimaging is still rare. In this work, we investigated the dual phosphorescence properties of a series of iridium(III) complexes of general formula [Ir(N^C)2(N^N)](PF)6. The internal conversion from TIL to TCT states can be inhibited by additional non-conjugate amino groups in the cyclometalating ligands with high lone-pair energy levels, but be facilitated by a low-energy π*(N^N) orbital. Simultaneous occurrence of phosphorescence from the two states is possible by rational structure manipulation tuning the energy levels of related molecular orbitals to allow the energy of 3NLCT being between those of 3IL and 3CT. The small-molecular dual-phosphorescent complexes are attractive candidates for development of wavelengthratiometric probes, which display distinct spectrumprofile response towards a variety of analytes, especially disease related factors. Sensitive spectral response of the

ACKNOWLEDGMENT We thank Dr Juan Song, Qi Wu, and Xuecheng Wang for their help and valuable discussions. We thank National Natural Science Foundation of China (61775104, 21501098, 51473078 and 21671108), National Program for Support of Top-Notch Young Professionals, Natural Science Foundation of Jiangsu Province of China (BK20150833), Scientific and Technological Innovation Teams of Colleges and Universities in Jiangsu Province (TJ215006), and Priority Academic Program Development of Jiangsu Higher Education Institutions (YX03001) for financial support.

REFERENCES (1)

Semenza, G. L. Science 2007, 318, 62-64.

(2) Decker, H.; Van Holde, K. E. Oxygen and the Evolution of Life Springer Verlag, Berlin Heidelberg, 2011. (3)

Harris, A. L. Nat. Rev. Cancer 2002, 2, 38-47.

(4) Brahimi-Horn, M. C.; Chiche, J.; Pouysségur, J. J. Mol. Med. 2007, 85, 1301-1307. (5) 410.

Wilson, W. R.; Hay, M. P. Nat. Rev. Cancer 2011, 11, 393-

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

(6) Murdoch, C.; Muthana, M.; Lewis, C. E. J. Immunol. 2005, 175, 6257-6263.

(32) Hutter, L. H.; Muller, B. J.; Koren, K.; Borisov, S. M.; Klimant, I. J. Mater. Chem. C 2014, 2, 7589-7598.

(7) Eltzschig, H. K.; Bratton, D. L.; Colgan, S. P. Nat. Rev. Drug Discov. 2014, 13, 852-869.

(33) Amela-Cortes, M.; Paofai, S.; Cordier, S.; Folliot, H.; Molard, Y. Chem. Commun. 2015, 51, 8177-8180.

(8) (9) 992.

Semenza, G. L. Annu. Rev. Med. 2003, 54, 17-28. Freeman, B. A.; Crapo, J. D. J. Biol. Chem. 1981, 256, 986-

(10) Chowdhury, A. K.; Watkins, T.; Parinandi, N. L.; Saatian, B.; Kleinberg, M. E.; Usatyuk, P. V.; Natarajan, V. J. Biol. Chem. 2005, 280, 20700-20711. (11) Brueckl, C.; Kaestle, S.; Kerem, A.; Habazettl, H.; Krombach, F.; Kuppe, H.; Kuebler, W. M. Am. J. Respir. Cell Mol. Biol. 2006, 34, 453-463.

(34) Liu, J.; Liu, Y.; Bu, W.; Bu, J.; Sun, Y.; Du, J.; Shi, J. J. Am. Chem. Soc. 2014, 136, 9701-9709. (35) Zheng, X.; Wang, X.; Mao, H.; Wu, W.; Liu, B.; Jiang, X. Nat. Commun. 2015, 6, 5834. (36) Zhang, K. Y.; Zhang, J.; Liu, Y.; Liu, S.; Zhang, P.; Zhao, Q.; Tang, Y.; Huang W. Chem. Sci. 2015, 6, 301-307. (37) Yoshihara, T.; Yamaguchi, Y.; Hosaka, M.; Takeuchi, T.; Tobita, S. Angew. Chem. Int. Ed. 2012, 51, 4148-4151.

(12) Koizumi, M.; Frank L.; Massaro, D. Am. Rev. Respir. Dis. 1985, 131, 907-911.

(38) Zhao, Q.; Zhou, X.; Cao, T.; Zhang, K. Y.; Yang, L.; Liu, S.; Liang, H.; Yang, H.; Li, F.; Huang, W. Chem. Sci. 2015, 6, 18251831.

(13) Gerstner, B.; DeSilva, T. M.; Genz, K.; Armstrong, A.; Brehmer, F.; Neve, R. L.; Felderhoff-Mueser, U.; Volpe, J. J.; Rosenberg, P. A. J. Neurosci. 2008, 28, 1236-1245.

(39) Wang, R.-F.; Peng, H.-Q.; Chen, P.-Z.; Niu, L.-Y.; Gao, J.F.; Wu, L.-Z.; Tung, C.-H.; Chen, Y.-Z.; Yang Q.-Z. Adv. Funct. Mater. 2016, 26, 5419-5425.

(14) Carvalho, C. R.; de Paula Pinto Schettino, G.; Maranhao, B.; Bethlem, E. P. Curr. Opin. Pulm. Med. 1998, 4, 300-304.

(40) Wu, C.; Bull, B.; Christensen, K.; McNeill, J. Angew. Chem. Int. Ed. 2009, 48, 2741-2745.

(15) Bhandari, V.; Choo-Wing, R.; Lee, C. G.; Zhu, Z.; Nedrelow, J. H.; Chupp, G. L.; Zhang, X. C.; Matthay, M. A.; Ware, L. B.; Homer, R. J.; Lee, P. J.; Geick, A.; de Fougerolles, A. R.; Elias, J. A. Nat. Med. 2006, 12, 1286-1293.

(41) Xu, R.; Wang, Y.; Duan, X.; Lu, K.; Micheroni, D.; Hu, A.; Lin, W. J. Am. Chem. Soc. 2016, 138, 2158-2161.

(16) Syed, M.; Das, P.; Pawar, A.; Aghai, Z. H.; Kaskinen, A.; Zhuang, Z. W.; Ambalavanan, N.; Pryhuber, G.; Andersson, S.; Bhandari, V. Nat. Commun. 2017, 8, 1173. (17) Yu, D. Y.; Cringle, S. J. Exp. Eye Res. 2005, 80, 745-751. (18) Shen, J. K.; Yang, X. R.; Dong, A. L.; Petters, R. M.; Peng, Y. W.; Wong, F.; Campochiaro, P. A. J. Cell. Physiol. 2005, 203, 457-464. (19) Wang, P.; Liu, X. C.; Yan, H.; Li, M. Y. Mol. Vis. 2009, 15, 2945-2952. (20) Beebe, D. C.; Holekamp, N. M.; Shui, Y.-B. Ophthalmic Res. 2010, 44, 155-165. (21) Reiter, R. J. Prog. Neurobiol. 1998, 56, 359-384. (22) Papkovsky, D. B.; Dmitriev, R. I. Chem. Soc. Rev. 2013, 42, 8700-8732. (23) Wang, X. D.; Wolfbeis, O. S. Chem. Soc. Rev. 2014, 43, 3666-3761.

(42) Kasha, M. Faraday Soc. Discuss. 1950, 9, 14-19. (43) Zhang, G.; Palmer, G. M.; Dewhirst, M. W.; Fraser, C. L. Nature Mater. 2009, 8, 747-751. (44) Zhao, Q.; Li, F.; Huang, C. Chem. Soc. Rev. 2010, 39, 30073030. (45) Dixon, I. M.; Collin, J. P.; Sauvage, J. P.; Flamigni, L.; Encinas, S.; Barigelletti, F. Chem. Soc. Rev. 2000, 29, 385-391. (46) Zhang, K. Y.; Chen, X.; Sun, G.; Zhang, T.; Liu, S.; Zhao, Q.; Huang, W. Adv. Mater. 2016, 28, 7137–7142. (47) Lo, K. K.-W.; Zhang, K. Y.; Leung, S.-K.; Tang, M.-C. Angew. Chem. Int. Ed. 2008, 47, 2213-2216. (48) You, Y.; Han, Y.; Lee, Y.-M.; Park, S. Y.; Nam, W.; Lippard, S. J. J. Am. Chem. Soc. 2011, 133, 11488-11491. (49) Zhang, K. Y.; Liu, H.-W.; Tang, M.-C.; Choi, A. W.-T.; Zhu, N.; Wei, X.-G.; Lau, K.-C.; Lo, K. K.-W. Inorg. Chem. 2015, 54, 6582-6593. (50) Stern, O.; Volmer, M. Phys. Z 1919, 20, 183-188.

(24) Piao, W.; Tsuda, S.; Tanaka, Y.; Maeda, S.; Liu, F.; Takahashi, S.; Kushida, Y.; Komatsu, T.; Ueno, T.; Terai, T.; Nakazawa, T.; Uchiyama, M.; Morokuma, K.; Nagano, T.; Hanaoka, K. Angew. Chem. Int. Ed. 2013, 52, 13028-13032. (25) Knox, H. J.; Hedhli, J.; Kim, T. W.; Khalili, K.; Dobrucki, L. W.; Chan, J. Nat. Commun. 2017, 8, 1794. (26) Piao, W.; Hanaoka, K.; Fujisawa, T.; Takeuchi, S.; Komatsu, T.; Ueno, T.; Terai, T.; Tahara, T.; Nagano, T.; Urano, Y. J. Am. Chem. Soc. 2017, 139, 13713-13719. (27) Liu, J.-N.; Bu, W.; Shi, J. Chem. Rev. 2017, 117, 6160-6224. (28) Xie, Z.; Ma, L.; deKrafft, K. E.; Jin, A.; Lin, W. J. Am. Chem. Soc. 2010, 132, 922-923. (29) Zhang, S.; Hosaka, M.; Yoshihara, T.; Negishi, K.; Iida, Y.; Tobita, S.; Takeuchi, T. Cancer Res. 2010, 70, 4490-4498. (30) Tobita, S.; Yoshihara, T. Curr. Opin. Chem. Biol. 2016, 33, 39-45. (31) Achatz, D. E.; Meier, R. J.; Fischer, L. H.; Wolfbeis, O. S. Angew. Chem. Int. Ed. 2011, 50, 260-263.

ACS Paragon Plus Environment

Page 8 of 9

Page 9 of 9 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

Journal of the American Chemical Society

Table of Contents artwork R1

L1

N2 N

R2N R2N

C

III

Ir C N

Air

O2

R2 N N

L2

R1

ACS Paragon Plus Environment