Graphene Quantum Dots Integrated in Ionophore-Based Fluorescent

Nov 2, 2018 - Turn-On Fluorescence Probe for Nitric Oxide Detection and ... Probe for Imaging of Nitric Oxide during Endoplasmic Reticulum Stress...
7 downloads 0 Views 2MB Size
Subscriber access provided by Kaohsiung Medical University

Article

Graphene Quantum Dots Integrated in Ionophorebased Fluorescent Nanosensors for Na and K +

+

Renjie Wang, Xinfeng Du, Yaotian Wu, Jingying Zhai, and Xiaojiang Xie ACS Sens., Just Accepted Manuscript • DOI: 10.1021/acssensors.8b00918 • Publication Date (Web): 02 Nov 2018 Downloaded from http://pubs.acs.org on November 4, 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 10 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 Sensors

Graphene Quantum Dots Integrated in Ionophore-based Fluorescent Nanosensors for Na+ and K+ Renjie Wang, Xinfeng Du, Yaotian Wu, Jingying Zhai, and Xiaojiang Xie* Department of Chemistry, Southern University of Science and Technology, Shenzhen, 518055, P. R. China Email: [email protected] Abstract. To enrich the recipes of ion-selective nanosensors, graphene quantum dots (GQDs) were integrated into ionophore-based fluorescent nanosensors with exquisite selectivity and high sensitivity for Na+ and K+. The unique property of GQDs gave the nanosensors ultrasmall size (ca. 10 nm), high brightness, good biocompatibility and potential pH sensing possibility. At pH 7.4, the sensors exhibited a detection range from 0.1 mM to 1 M for Na+ and from 3 µM to 1 mM for K+. The nanosensors were successfully applied to blood serum and urine samples. Chemically induced intracellular sodium concentration change in HeLa cells was also qualitatively monitored. Key words: graphene quantum dots, nanosensors, ion-selective, sodium, potassium Fluorescent ion sensors are known to have high sensitivity and

probes. In addition, diffusion limited response time of the

could provide high spatial and temporal resolution of ion

nanosensors is much faster than some ion-selective biosensors

+

+

concentrations. Among various ions, Na and K are the most

based on nucleic acids.11

common alkali metal ions in the human body fluids. The Na+

Previous

concentration in extracellular fluid could reach ca. 440 mM for

pioneered by the groups of Bakker, Hall, Clark, and

+

Michalska.12-18 The common ground of these nanosensors is the

concentration is typically around 400 mM and 139 mM,

underlying classical ion-selective optode principle.19 However,

respectively.1 The appropriate intracellular and extracellular

the nanosensors have been built on various materials which

invertebrates and ca. 145 mM for vertebrates while cytosolic K

+

+

research

on

ionophore-based

nanosensors

was

help maintain the proper

mostly are optically inert. Apart from a few examples, the

membrane potential which is vital to the regulation of a number

intracellular applications of the nanosensors is also very limited.

concentrations of Na

and K

2

of signal transduction pathways. Imbalance of the electrolytes

A paradigm was demonstrated earlier by Clark and co-workers

concentrations is also related to a number of diseases including

where they successfully utilized plasticized poly (vinyl chloride)

congestive heart failure, kidney malfunction, severe dehydration,

sodium selective nanosensor to record sodium dynamics in

bulimia, and cancer.

3

isolated cardiomyocytes.16 More recently, we reported optical +

nanosensors for Li+, Na+, and K+ based on organosilica

and K+ is quite small.4-5 Commercial synthetic probes for Na+

nanoparticles.20 The crosslinked organosilica particles could be

and K+ such as SBFI and PBFI suffer from short excitation

readily taken into HeLa cells and provided high robustness

However, the palette of fluorescent sensors and probes for Na

wavelength, poor selectivity and sensitivity.

6-7

Over the past few

against dilution and cell metabolism compared with the

decades, different synthetic probes have been reported with

self-assembled nanosensors.

improved characteristics, some also compatible with two-photon

In this work, we present for the first time, ion-selective

microscopy and near infrared spectroscopy.

8-9

Nonetheless,

nanosensors based on graphene quantum dots (GQDs)

small molecular probes still suffer from problems such as

containing highly selective ionophores. While extending the

cellular sequestration and fast leakage after loading. In contrast,

variety of ionophore-based nanosensors, the results also lead to

ion-selective optical nanosensors become advantageous because

an optically active and readily functionalized host material. The

molecular probes are well protected by the matrix host material

surface of GQDs was modified with acetylene groups and

of the nanosensors, preventing the probes from involving in

crosslinked using azide-functionalized poly (ethylene oxide)

10

Nanosensors

through Cu+-catalyzed click reaction. The strongly green

therefore, provides high brightness and stability, and could

fluorescent GQDs was easily traceable under fluorescence

remain in cells for longer time compared with small molecular

microscopy and allowed signal interrogation at multiple

direct metabolic and other interactive processes.

ACS Paragon Plus Environment

ACS Sensors 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 10

wavelengths. In recent years, the functionalization of GQDs for

washed 3 times with THF by centrifuging at 7000 rpm for 3 min

chemical sensing has received extensive attention and several

to remove any unreacted molecules. The THF residue after

groups reported graphene-based materials for the detection of

centrifugation was removed using compressed air. The solids

+

2+

3+

- 21-24

Ag , Hg , Fe , and I .

However, GQD-based nanosensors +

for alkali metal ions are very rare. Here, Na -selective and +

K -selective GQD-based nanosensors were prepared by using +

different ionophores. The Na nanosensors exhibited exquisite +

selectivity over K , rendering them an attractive sensing tool for +

were re-suspended in 1 mL deionized water containing 1.2 mg polyoxyethylene bis(azide), 3.2 mg L-ascorbic acid, and 4.2 mg CuSO4·5H2O and shaken at room temperature for 6 h. The products of the reaction were centrifuged at 7000 rpm for 3 min and washed with deionized water for at least 3 times. The

intracellular experiments. Cytosolic Na level change in HeLa

product PEG-GQDs was suspended in 1 mL deionized water

cells

and stored in refrigerator for further use.

induced

by

gramicidin

and

carbonyl

cyanide

3-chlorophenylhydrazone (CCCP) was successfully observed

Nanosensor preparation For Na+-selective nanosensors, 0.5

under fluorescence microscope with the Na+ nansoensors. The

mg Ox R, 1.5 mg TFPB, 2.5 mg NaX, 4.8 mg DOS were

nanosensors were also applied in human urine and blood serum

dissolved in 2.4 mL of methanol to form a homogeneous

samples.

cocktail solution. A volume of 80 µL PEG-GQDs stock solution

EXPERIMENTAL SECTION

and 25 µL of above-mentioned methanol cocktail were mixed and added into 4 mL of Tris-HCl solution (10 mM, pH 7.4) on a

Reagents Graphene quantum dots (GQDs), polyoxyethylene bis(azide) (MW 5000), L-ascorbic acid, copper(II) sulfate pentahydrate, sodium

ionophore

X

(NaX),

potassium

tetrakis-[3,5-bis(trifluoromethyl)-phenyl]

or

borate

sodium (TFPB),

potassium ionophore I (valinomycin), methanol, tetrahydrofuran (THF),

bis(2-ethylhexyl)

sebacate

(DOS),

2-amino-2-(hydroxymethyl)-1,3-propanediol (Tris), citric acid, boric acid and sodium dihydrogen phosphate were purchased from Sigma-Aldrich. Propargyl bromide was purchased from J&K

Scientific

Ltd.

in

China.

synthesized according to literature.25

All

Ox

R

were

solutions

were

prepared by dissolving appropriate salts into deionized water purified by Milli-Q Integral 5. Dulbecco's modified eagle medium (DMEM), heat-inactivated fetal bovine serum (FBS), penicillin-streptomycin solution (100X) were purchased from Corning. Cell Counting Kit-8 (CCK-8) was obtained from MedChem Express in China. Low autofluorescence cell culture medium TransDetect BrightFluore DMEM was purchased from Beijing TransGen Biotech Co., Ltd. Gramicidin was purchased from Shanghai Jingdu Biological Technology Co., Ltd. Carbonyl

cyanide

3-chlorophenylhydrazone

(CCCP)

was

purchased from Alfa Aesar. Urine sample was obtained from a volunteer. Blood serum sample was provided by Dr. Fanxin Zeng from the Department of Clinic Medical Center, Dazhou Central Hospital, China. Modification of GQDs. 1mg GQDs and 100 µL propargyl bromide were added in a flask and ultrasonically dispersed, and then heated to 60 ℃ for 2 h. The resulting precipitates were

vortexer at 1000 rpm. Methanol was removed by blowing compressed air to the surface of the resulting solution for at least 1 h. Similarly, for the preparation of K+-selective nanosensors, a 2.4 mL of methanol cocktail 0.5 mg Ox R, 1.5 mg TFPB, 2.8 mg valinomycin, 4.8 mg DOS were prepared. Instrumentation and Measurements. The morphology and size of the nanosensors were characterized by transmission electron

microscopy

(TEM,

HT-7700,

Hitachi)

operated at an acceleration voltage of 100 kV. To prepare the samples for TEM characterization, a drop of the nanosensor suspension was dropped onto a copper grid with a carbon support film, and dried in air. Infrared spectra were recorded on an FT-IR spectrometer with an ATR accessory (Nicolet iS10, Thermo Scientific) at room temperature. Fluorescence spectra was measured on a fluorescence spectrometer (Fluorolog-3, Horiba Jobin Yvon). Absorption spectra were measured using an ultraviolet−visible (UV-vis) absorption spectrometer (Evolution 220, Thermo Fisher Scientific). The pH response of PEG-GQDs were evaluated in buffer solutions with 2.5 mM of citric acid, boric acid, and NaH2PO4 adjusted to the required pH. The fluorescence response of ion-selective nanosensors with at different ion concentrations was recorded in pH 7.4 Tris-HCl buffer with stepwise addition of stock solutions. The ratio of the fluorescence intensity at 522 nm and 602 nm were used to for sensor calibration curves. Furthermore, K+ concentration in the urine sample was measure after 100x dilution with the abovementioned process by using the K+-selective nanosensors in Tris-HCl solution containing 1 mM sodium background. The K+ level in the urine sample was separately determined using

ACS Paragon Plus Environment

Page 3 of 10 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 Sensors

ion-selective electrodes in potentiometry. The ion-selective

OH

pH Sensitive Region

Legends:

electrode membrane containing valinomycin as the ionophore 26

was prepared according to the literature.

All measurements

Poly (ethylene oxide) + H+

- H+

Graphene Quantum Dots

O-

Ion-Exchanger (TFPB)

were conducted in triplicated.

Chromoionophore (HInd+, Ind)

Cell culturing and cytotoxicity evaluation. Hela cells were

Na+/K+ Selective Ionophore (L)

seeded

in

DMEM

culture

supplemented

with

10%

heat-inactivated FBS and 1% penicillin-streptomycin, and incubated at 37 ℃ with humidified air containing 5% CO2. Cytotoxicity of the nanosensors was assayed using CCK-8 according to the manual. Briefly, after Hela cells were cultured in a 96-well plate (100 µL, 104 cells per well) for 24 h, different amounts of nanosensor stock solution (10, 30, and 50 µL) were added into experiment group wells and co-incubated for 24 h. The wells without nanosensors were set as the control group. 10 µL of CCK-8 reagents was introduced into each well, and the absorbance at 450 nm was recorded after 2 h on a microplate reader (cytation5, BioTek). Confocal fluorescence imaging. For fluorescence imaging, Hela cells were incubated in confocal dishes with the nanosensors for 3 h, washed carefully 3 times with phosphate buffer (10 mM, pH 7.4). Afterwards, the cells were cultured in TransDetect BrightFluore DMEM with low autofluorescence and imaged using confocal laser scanning microscopy (A1R, Nikon). Laserlines at 488 nm and 561 nm were used as the excitation together with the standard FITC and TRITC filter cubes. Chemically induced cytosolic sodium level change. Hela cells loaded with Na+ nanosensors were firstly cultured in TransDetect BrightFluore DMEM medium. Gramicidin and CCCP was added to the medium to reach final concentrations of 20 µM and 5 µM, respectively. NaCl was also added to the culture medium reach 100 mM concentration. The fluorescence intensities at 522 nm and 602 nm were continuously recorded with the plate reader. Images of the cells before and after perturbation were also captured with Cytation 5 using with colored CCD camera and LED excitation at 469±15 nm and 586±15 nm.

HInd+ TFPB + M+ (aq) L (org)

Ind TFPB + H+ (aq) LM+ (org)

Ion-Selective Region

Figure 1. Schematic illustration of the proposed structure of the ionophore-based nanosensor incorporating GQDs and the mechanism of the optical responses for pH and ions. graphene with the periphery rich in hydroxyl groups and possess strong quantum confinement and edge effects. Therefore, they typically exhibit high dispersity in aqueous environments and their photoluminescence spectra could be tailored by the size, shape, and defects.27 The nanosensors contain pH sensitive regions because of the surface hydroxy groups in the GQDs. It is more challenging, however, to obtain the ion-selective region, which required surface modification with polyethylene oxides. Although GQDs contained a large number of sp2 carbons, they did not possess interface with high enough hydrophobicity to directly adsorb the sensing optode components. In our initial experiments, we attempted to directly mix the sensing ingredients with GQDs. However, the resulting suspension did not show any fluorescence response to sample ion concentration change. To integrate the green fluorescent GQDs into ionophore-based nanosensors, the hydroxyl groups of the GQDs were reacted with propargyl bromide to add carbon-carbon triple bonds to the periphery. The reactant and product of the reaction was compared in Fourier transform infrared spectroscopy (FT-IR). As shown in Figure 2a, the asymmetric carbon-carbon stretching at 2125 cm-1, the C-H stretching at 3300 cm-1, 2930 cm-1, and 2844 cm-1 of the propargyl group confirmed the successful modification. Then, polyoxyethylene bis(azide) was used to crosslink the propargylated GQDs in the presence of Cu+.

After

crosslinking,

the

product

PEG-GQDs

was

centrifuged and washed with water. Note that neither the naked GQDs nor the propargylated GQDs could be collected through

RESULTS AND DISCUSSION

centrifugation in the same conditions, providing another

Figure 1 shows a schematic illustration of the proposed structure

evidence for the crosslinking reaction. An increase in the

of the nanosensors where PEGylated GQDs acted as a host for

hydrodynamic size of the PEG-GQDs nanosensors to around

the ion-selective optode components, i.e., ionophores (L),

150 nm due to aggregation was also observed with dynamic

ion-exchanger (TFPB) and the chromoionophore (Hind+, Ind).

light scattering (DLS, see size distribution and correlation curve

The plasticizer DOS, represented in gray, served to solubilize

in Figure S1). As shown in Figure 2b, the aqueous suspension of

the optode components. GQDs are sheets of few-layered

the PEG-GQDs was also slightly more turbid than the free

ACS Paragon Plus Environment

ACS Sensors 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 10

Figure 3. (a) Fluorescence spectra of the Na+ nanosensors with various Na+ concentrations in pH 7.4 Tris-HCl buffer. (b) Selectivity pattern of the Na+ nanosensors. (c) Fluorescence spectra of the K+ nanosensors with various K+ concentrations in Figure 2. (a) FT-IR spectra of free GQDs, the reactant propargyl

pH 7.4 Tris-HCl buffer. (d) Selectivity pattern of the K+

bromide, and the purified product PEG-GQDs. (b) Pictures of

nanosensors.

the aqueous suspension of the free GQDs (1), PEG-QGDs (2),

compounds, leaving some of the hydroxy groups still exposed to

and the nanosensors with Na+ selective components (3). (c)

the aqueous solution (the pH sensitive region in Figure 1).

TEM images of the well separated nanosensors (1) and

The nanosensors were characterized with transmission electron

aggregates (2 and 3) with different magnification.

microscopy (TEM). As shown in Figure 2c, the majority of the

GQDs. The third image of Figure 2b was from a Na+ nanosensor suspension. After the abovementioned modifications, a more amphiphilic interface was obtained for the incorporation of the sensing ingredient. Here, we used the solvent displacement method to prepare the ion-selective nanosensors.28 Briefly, all the sensing components including PEG-GQDs, cation-exchanger (TFPB), chromoionophore, and ionophore were dissolved in methanol and mixed with buffer solutions (see experimental section for detailed preparation). Hydrophobic interactions drive the sensing ingredient into the crosslinked PEG-GQDs as methanol was evaporated, forming the ion-selective region shown in Figure 1. Previously, different surfactants were used to stabilize various types of nanosensors.14,

16

Here, since poly (ethylene

oxide) was already linked to GQDs, no additional surfactant was required. No sedimentation or flocculation was observed for the nanosensor suspension over an observation period of three

resulting Na+ nanosensors produced TEM images as shown in Figure 2c1 with an average diameter of ca. 10 nm, which is slightly bigger than the original GQDs (< 5 nm). Since poly (ethylene oxide) chains could not be seen under TEM, the contrast should come from the few-layered graphene sheets. The result indicates that GQDs in the nanosensors were probably crosslinked between the layers (as shown in Figure 1) due to the stabilizing π-π interaction between the GQD planes. However, in some area, aggregates of several particles as shown in Figure 2c2 and 2c3 were also observed. Comparing to the length of the polyethylene oxide chains, the possibility for the formation of inter-particular crosslinking can be precluded. Images of the K+ nanosensors were also obtained (Figure S2a). The size and the shape of individual particles was very similar. Aggregation may be formed during the TEM sample preparation. TEM image of the starting GQDs with the same sample preparation was shown in Figure S2b for comparison.

weeks. However, the surface of the GQDs was not completely

The Na+-selective nanosensors were prepared with sodium

covered with polyethylene oxide chains and the hydrophobic

ionophore X (NaX, a calix[4]arene derivative), TFPB, and a H+

ACS Paragon Plus Environment

Page 5 of 10 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 Sensors

turn-on chromoionophore, Ox R, (see Figure S3 in the

even above 0.1 M.

supporting information). This chromoionophore has been

Likewise, K+-selective nanosensors were prepared by replacing

25, 28-29

As

the NaX with valinomycin (potassium ionophore I). As shown

+

in Figure 3c and 3d, the nanosensor showed a highly selective

was observed owing to the good overlap between the emission

response to K+. Notice that the relative fluorescence intensity

of GQDs and the excitation of Ox R. As Na+ concentration

difference of the two peaks between Figure 3a and 3c was

increased, Ox R emission (around 602 nm) decreased because of

simply a result of different doping amount of Ox R. The

the deprotonation of OX R while GQD emission showed an

emission intensity of Ox R showed a decrease as K+

increase, which is a result of both the energy transfer and

concentration went from 1 µM to 1 mM. However, only a

secondary filtration effect between Ox R and GQDs.

slight increase in the GQD emission was observed, indicating a

Importantly,

the

less efficient energy transfer between Ox R and GQDs. We infer

chromoionophore. Without additional chromoionophore, the

that the distribution of the sensing components in the

previously used in several ionophore-based sensors.

shown in Figure 3a, a ratiometric fluorescence response to Na

GQDs

themselves

could

not

act

as

+

suspension of GQDs alone showed very poor response to Na or K

+

(Figure S4). For the PEG-GQD particle suspension

nanosensors could be different and changing with environment. For instance, the adsorption of proteins could influence the

incorporating only the chromoionophore Ox R, only emission

intensity of GQD emission.

from GQDs were fluorescent since the deprotonated from of Ox

Moreover, according to previous studies, energy transfer within

R was nonfluorescent (Figure S5). Here, the nanosensors

the nanosensors is every sensitive to intermolecular distance,

functioned on the basis of ion-exchange theory. Therefore,

which makes the distribution of the sensing components quite

changing the doping level and ratio between the sensing

important.30-31

components could indeed alter the shape of the optical response

The freedom in the choice of the sensing components and their

curve (Figure S6). Therefore, the doping amount of the sensing

ratio is one of the advantages of ionophore-based ion-selective

components was kept consistence throughout all experiments.

optodes. Here, the chromoionophore Ox R was firstly chosen

In pH 7.4 Tris-HCl buffered solutions, the nanosensors showed

because of the spectral overlap with GQDs for ratiometric

+

a Na response from 0.1 mM to 1 M and a good selectivity over +

+

2+

K , Li , Mg , and Ca

2+

(Figure 3b). These interference ions

measurements at constant sample pH. On the other hand, we noted a highly sensitive pH response of the modified GQDs in

were evaluated because of their common presence in

the

intracellular media and other body fluids. Concentrations above

chromoionophore in nansensors containing Ox R, TFPB, and

1 M were not evaluated due to solubility. Notably, K+ ions, the

NaX could enable the dual sensing for e.g., pH and Na+. For this

most abundant intracellular metal ion, caused little interference

purpose,

fluorescence

the

mode

energy

(Figure

transfer

S7).

between

Changing

GQDs

and

the

the

chromoionophore should be reduced. Here, a different chromoionophore (Ox B) was assessed..25 As shown in Figure S3, Ox B is structurally similar compared with Ox R, but has much more red-shifted absorption, and thus much less energy transfer with GQDs. When excited at 460 nm (Figure 4a), the GQD emission around 522 nm was indeed pH sensitive with an apparent pKa around 6.0, which makes these nanosensors potentially useful for observing the cellular acidification processes. Upon excitation at 630 nm (Figure 4b), sample Na+ concentration change was successfully reflected on the decrease of the fluorescence emission around 700 nm. A small emission peak from Ox B was also observed, indicating that there may Figure 4. (a) Fluorescence pH response of GQD-based

still be some energy transfer from GQDs to Ox B. To

nanosensors

independently

containing

Ox

B,

TFPB,

+

and

NaX.

(b)

measure

sample

pH,

excitation

of

the

Fluorescence Na response of the GQD-based nanosensors at

chromoionophore should have minimum overlap with the GQD

various Na+ concentration as indicated.

emission, which was not yet fulfilled with Ox R and Ox B

ACS Paragon Plus Environment

ACS Sensors 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 10

(Figure S8).

of components over time.

The short response time of the nanoscale sensors is an attractive

On the other hand, the small size and the excellent selectivity

feature. Ion-selective nanosensors with fast response have been

rendered the nanosensor very attractive for intracellular

demonstrated in several cases and modeling for the response

measurements. Therefore, the potential application of the Na+

time also has been described.14-15, 32-34 The color change upon

nanosensors was attempted in HeLa cells. The nanosensors were

changing the analyte concentration appeared instantaneous to

introduced into the cells through endocytosis by incubating the

the eyes. To obtain quantitative values, we attempted

cells in culture medium containing the nanosensors. The culture

stopped-flow methods. However, the experiment was limited by

medium was washed away before fluorescence microscopic

the mixing time of the samples and the results were not

imaging. Further reducing the size and some surface

accurate.

modification could help make the cell uptake more efficient.

As preliminary applications, the K+ concentration in a human +

Laser scanning confocal fluorescence microscopy was used to

urine sample was successfully measured with the K nanosensor.

study the intracellular distribution of the Na+ nanosensors

The results (20.1±0.2 mM) was found very close to the value

containing Ox R. As shown in Figure 5, the green color was

+

obtained from potentiometric measurements with K -selective

from the GQD emission while the red from Ox R. The

electrodes (20.4±0.1 mM, external calibration method). The

colocalization analysis confirmed that the leakage of GQDs was

calibration curves were shown in Figure S9a,b in the supporting

negligible. In addition, GQDs were known as a biocompatible

information. We also attempted to apply the nanosensors in

material.27 The cytotoxicity was also evaluated with the

blood serum samples. In absorption mode, blood sodium levels

commercial CCK-8 assay to confirm the biocompatibility of the

were successfully measured (152±3 mM) using the maximum

nanosensors. As shown in Figure 5e, the results indicated very

absorbance of Ox R for calibration (Figure S9c,d). The results

little toxicity after 24 hours of incubation with various amount

were also close to the ones from ISEs (149±1 mM). In

of nanosensors.

fluorescence mode, however, the GQD emission was easily

In order to evaluate whether the nanosensors respond to

influenced by the addition of blood serum, causing a decrease in

intracellular Na+ concentration changes, gramicidin (20µM)

intensity. (Figure S10) The nanosensors remained functional

was added into the culture medium together with CCCP (5µM)

after storing in fridge for 30 days. Figure S11 shows that the

while the extracellular Na+ concentration was changed to 100

response curves were quite similar, with a little change in the

mM. This protocol has been used in previous reports to allow

shape of response curve, which might indicate some degradation

the entry of sodium ions through the cell membrane in which gramicidin

forms

equilibration

of

sodium

channels

intracellular

and

sodium

CCCP to

allows

extracellular

concentrations.35 Since the typical cytosolic Na+ concentration for mammalian cells is around 12 mM,1 such experimental perturbation should cause an increase of the cytosolic Na+ concentration, which in turn causes a decrease in the chromoionophore emission intensity of the nanosensors. Images of the nanosensors-loaded cells in Figure 6a were recorded before adding gramicidin and CCCP while the ones in Figure 6b were recorded 20 minutes after adding gramicidin and CCCP to the cells. The fluorescence intensity of Ox R in the red channel Figure 5. Confocal fluorescence microscopy imaging of HeLa

clearly experienced a decrease, which indeed reflected an

cells loaded with the Na+ nanosensors. (a) FITC channel, (b)

increase of intracellular Na+ concentration. In the meantime, the

TRITC channel, (c) overlay. (d) Fluorescence intensity

fluorescence in the green channel showed no notable change,

distribution along the blue line in Part (a), (b), and (c). Scale bar:

indicating no dramatic intracellular pH change after treating the

25 µm. (e) Cytotoxicity evaluation of the Na+ nanosensors with

cells with gramicidin and CCCP. The dynamic fluorescence

the CCK-8 assay.

signal change was also continuously recorded on a plate reader. As shown in Figure 6c, 6d and 6e, the Ox R emission intensity

ACS Paragon Plus Environment

Page 7 of 10 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 Sensors

Figure 6. Bright field and fluorescence microscopy images of the Na+ nanosensor-loaded HeLa cells before (a) and 20 min after the addition of gramicidin (20µM) and CCCP (5µM) to induce intracellular Na+ level change. Scale bar: 50 µm. The time evolution of the fluorescence intensity at 522 nm (c), 602 nm (d), and their ratio (e) upon addition of gramicidin and CCCP. Data was averaged from observations in 5 parallel wells. of emission at 602 nm gradually decreased for about ca. 20%

nanosensors to measure sample pH. The Na+ nanosensors were

and there was no dramatic change in the emission from the

readily introduced into HeLa cells to monitor the cytosolic Na+

GQDs at 522 nm. The control experiment where the cells were

increase induced by treating the cells with gramicidin and CCCP.

monitored without addition of gramicidin and CCCP also

Preliminary application of the K+ nanosensors also allowed the

confirmed that the reduction of the fluorescence intensity at 602

successful determination of potassium level in a human urine

nm was not a consequence of photobleaching. The fluorescence

sample. This ionophore-based sensing principle could in

in the green channel did not increase as much as observed in in

principle be extended to other ionophores and lead to different

vitro calibrations. As observed for the blood serum sample, the

selectivity. The results here lay the foundation for a highly

GQD emission could be sensitive to intracellular proteins as

valuable optical ion sensing platform to diagnostic and

well.

biological applications.

For quantitative concentration analysis, since in vitro

calibrations could not be used directly, a protocol to calibrate

ASSOCIATED CONTENT

the sodium levels inside the cells is being pursued in the lab.

Supporting Information

CONCLUSIONS

Additional information as noted in the text include: chemical

To summarize, GQDs were successfully modified and integrated

structures of the sensing components, supplementary TEM, the

into ionophore-based optical nanosensors for Na+ and K+. The

pH response of PEG-GQDs, spectral overlap information, and

nanosensors exhibited excellent selectivity, small individual size

calibration curves. This material is available free of charge via

(ca. 10 nm in diameter), ratiometric optical response, fast

the Internet at http://pubs.acs.org.

response, and good biocompatibility. The use of GQDs with the

AUTHOR INFORMATION

proper choice of chromoionophore might also enable the

Corresponding Author

ACS Paragon Plus Environment

ACS Sensors 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 10

* Email: [email protected]

Sensing. PNAS 2015, 112, 5903-5908.

ORCID

12.

Xiaojiang Xie: 0000-0003-2629-8362

ion-selective nanospheres with voltage-sensitive dyes. J. Am.

ACKNOWLEDGEMENTS

Chem. Soc. 2014, 136 (47), 16465-8.

The authors thank the National Natural Science Foundation of

13.

China, the Thousand Talents Program of China, and the

to the nanoscale. Anal. Bioanal. Chem. 2015, 407 (14),

Shenzhen Municipal Science and Technology Innovation

3899-910.

Council for financial support.

14.

REFERENCES

Nanospheres: Maximising Response Range and Minimising

1.

Lodish, H.; Berk, A.; Kaiser, C. A.; Krieger, M.; Bretscher,

A.; Ploegh, H.; Amon, A.; Martin, K. C., Molecular Cell Biology. 8 ed.; W. H. Freeman: 2016. 2.

Alberts, B.; Johnson, A. D.; Lewis, J., Molecular Biology

of the Cell. 6 ed.; W. W. Norton & Company: 2014. 3.

Barrett, K.; Brooks, H.; Boitano, S.; Barman, S., Ganong’s

Review of Medical Physiology. 23 ed.; McGraw-Hill Medical: 2012. 4.

Gao, G.; Cao, Y.; Liu, W.; Li, D.; Zhou, W.; Liu, J.,

Fluorescent Sensors for Sodium Ions. Anal. Methods 2017, 9, 5570-5579. 5.

Song, G.; Sun, R.; Du, J.; Chen, M.; Tian, Y., A Highly

Selective, Colorimetric, and Environment-Sensitive Optical Potassium Ion Sensor. Chem. Commun. 2017, 53, 5602-5605. 6.

MINTA, A.; Tsien, R. Y., Fluorescent Indicators for

Cytosolic Sodium. J. Biol. Chem. 1989, 264, 19449-19457. 7.

Meuwis, K.; Boens, N.; Schryver, F. C. D.; Gallay, J.;

Vincent, M., Photophysics of The Fluorescent K+ Indicator PBFI. Biophys. J. 1995, 68, 2469-2473. 8.

Bandara, H. M. D.; Hua, Z.; Zhang, M.; Pauff, S. M.;

Miller,

S.

C.;

Davie,

E.

A.

C.;

Kobertz,

W.

R.,

Palladium-Mediated Synthesis of a Near-Infrared Fluorescent K+ Sensor. J. Org. Chem. 2017, 82, 8199-8205. 9.

Depauw, A.; Dossi, E.; Kumar, N.; Fiorini-Debuisschert,

C.; Huberfeld, G.; Ha-Thi, M.-H.; Rouach, N.; Leray, I., A Highly Selective Potassium Sensor for the Detection of Potassium in Living Tissues. Chem. Eur. J. 2016, 22, 14902-14911. 10.

Si, D.; Epstein, T.; Lee, Y.-E. K.; Kopelman, R.,

Nanoparticle PEBBLE Sensors for Quantitative Nanomolar Imaging of Intracellular Free Calcium Ions. Anal. Chem. 2012, 84, 978-986. 11. Torabi, S.-F.; Wu, P.; McGhee, C. E.; Chen, L.; Hwang, K.; Zheng, N.; Cheng, J.; Lu, Y., In Vitro Selection of a Sodium-Specific DNAzyme and its Application in Intracellular

Xie, X.; Zhai, J.; Bakker, E., Potentiometric response from

Xie, X.; Bakker, E., Ion selective optodes: from the bulk

Ruedas-Rama, M. J.; Hall, E. A. H., K+-Selective

Response Time. Analyst 2006, 131, 1282-1291. 15.

Walters, J. D.; Hall, E. A. H., An optrode particle

geometry to decrease response time. Analyst 2011, 136, 4718-4723. 16.

Dubach, J. M.; Das, S.; Rosenzweig, A.; Clark, H. A.,

Visualizing sodium dynamics in isolated cardiomyocytes using fluorescent nanosensors. PNAS 2009, 106, 16145-16150. 17.

Dubach, J. M.; Harjes, D. I.; Clark, H. A., Ion-Selective

Nano-optodes Incorporating Quantum Dots. J. Am. Chem. Soc 2007, 129, 8418-8419. 18.

Kłucinśka, K.; Stelmach, E.; Kisiel, A.; Maksymiuk, K.;

Michalska, A., Nanoparticles of Fluorescent Conjugated Polymers: Novel Ion- Selective Optodes. Anal. Chem. 2016, 88, 5644-5648. 19.

Mistlberger,

G.;

Crespo,

G.

A.;

Bakker,

E.,

Ionophore-Based Optical Sensors. Annu. Rev. Anal. Chem. 2014, 7, 483-512. 20.

Du, X.; Yang, L.; Hu, W.; Wang, R.; Zhai, J.; Xie, X., A

plasticizer-free miniaturized optical ion sensing platform with ionophores and silicon-based particles. Anal. Chem. 2018, 90 (9), 5818-5824. 21.

Wen, Y.; Xing, F.; He, S.; Song, S.; Wang, L.; Long, Y.; Li,

D.; Fan, C., A graphene-based fluorescent nanoprobe for silver(I) ions detection by using graphene oxide and a silver-specific oligonucleotide. Chem. Commun. 2010, 46, 2596-2598. 22.

He, L.; Li, J.; Xin, J. H., A novel graphene oxide-based

fluorescent nanosensor for selective detection of Fe3+ with a wide linear concentration and its application in logic gate. Biosens. Bioelectron. 2015, 70, 69-73. 23.

Li, M.; Zhou, X.; Ding, W.; Guo, S.; Wu, N., Fluorescent

aptamer-functionalized graphene oxide biosensor for label-free detection of mercury(II). Biosens. Bioelectron. 2013, 41, 889-893. 24.

Dinda, D.; Shaw, B. K.; Saha, S. K., Thymine

Functionalized Graphene Oxide for Fluorescence “Turn-off- on”

ACS Paragon Plus Environment

Page 9 of 10 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 Sensors

Sensing of Hg2+ and I− in Aqueous Medium. ACS Appl. Mater. Interfaces 2015, 7, 14743-14749. 25.

Xie, X.; Crespo, G. A.; Bakker, E., Oxazinoindolines as

fluorescent H+ turn-on chromoionophores for optical and electrochemical ion sensors. Anal Chem 2013, 85 (15), 7434-40. 26.

Qin, Y.; Bakker, E., Evaluation of the Separate

Equilibrium Processes That Dictate the Upper Detection Limit of Neutral Ionophore-Based Potentiometric Sensors. Anal. Chem. 2002, 74, 3134-3141. 27.

Zhu, Y.; Murali, S.; Cai, W.; Li, X.; Suk, J. W.; Potts, J. R.;

Ruoff, R. S., Graphene and graphene oxide: synthesis, properties, and applications. Adv. Mater. 2010, 22, 3906-3924. 28.

Xie,

X.;

Mistlberger,

G.;

Bakker,

E.,

Ultrasmall

fluorescent ion-exchanging nanospheres containing selective ionophores. Anal Chem 2013, 85 (20), 9932-8. 29.

Xie,

X.;

Zhai,

J.;

Crespo,

G.

A.;

Bakker,

E.,

Ionophore-based ion-selective optical nanosensors operating in exhaustive sensing mode. Anal Chem 2014, 86 (17), 8770-5. 30.

Roy, R.; Hohng, S.; Ha, T., A practical guide to

single-molecule FRET. Nat. Methods 2008, 5 (6), 507-516. 31.

Ruckh, T. T.; Skipwith, C. G.; Chang, W.; Senko, A. W.;

Bulovic, V.; Anikeeva, P. O.; Clark, H. A., Ion-Switchable Quantum Dot Förster Resonance Energy Transfer Rates in Ratiometric Potassium Sensors. ACS Nano 2016, 10 (4), 4020-4030. 32.

Du, X.; Xie, X., Non-equilibrium diffusion controlled

ion-selective optical sensor for blood potassium determination. ACS Sens. 2017, 2, 1410-1414. 33.

Park, E. J.; Brasuel, M.; Behrend, C.; Philbert, M. A.;

Kopelman, R., Ratiometric Optical PEBBLE Nanosensors for Real-Time Magnesium Ion Concentrations Inside Viable Cells. Anal. Chem. 2003, 75, 3784-3791. 34.

Kim, M. D.; Dergunov, S. A.; Lindner, E.; Pinkhassik, E.,

Dye-Loaded Porous Nanocapsules Immobilized in a Permeable Polyvinyl Alcohol Matrix: A Versatile Optical Sensor Platform. Anal. Chem. 2012, 84 (6), 2695-2701. 35.

Lo, C.-J.; Leake, M. C.; Berry, R. M., Fluorescence

Measurement of Intracellular Sodium Concentration in Single Escherichia coli Cells. Biophys. J. 2006, 90, 357-365.

ACS Paragon Plus Environment

ACS Sensors 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

For TOC only:

ACS Paragon Plus Environment

Page 10 of 10