Perylene Bisimide Aggregates as Probes for Subnanomolar

Apr 12, 2019 - Department of Biological and Environmental Sciences and Technologies, DISTEBA, University of Salento , Via per Arnesano, I-73100 Lecce ...
1 downloads 0 Views 725KB Size
Subscriber access provided by UNIV OF LOUISIANA

Surfaces, Interfaces, and Applications

Perylene Bisimide Aggregates as Probes for SubNanomolar Discrimination of Aromatic Biogenic Amines Simona Bettini, Zois Syrgiannis, Rosanna Pagano, Luka #or#evi#, Luca Salvatore, Maurizio Prato, Gabriele Giancane, and Ludovico Valli ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.9b04101 • Publication Date (Web): 12 Apr 2019 Downloaded from http://pubs.acs.org on April 13, 2019

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 49 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

Perylene Bisimide Aggregates as Probes for SubNanomolar Discrimination of Aromatic Biogenic Amines Simona Bettini,a,b‡ Zois Syrgiannis,c†‡ Rosanna Paganod Luka Đorđević,c† Luca Salvatore,a Maurizio Prato,c,e,f* Gabriele Giancane,b,g* and Ludovico Valli b,d

[a]

Department of Engineering for Innovation, Campus University Ecotekne, Via per

Monteroni, I-73100 Lecce, Italy

[b]

Consorzio Interuniversitario Nazionale per la Scienza e Tecnologia dei Materiali,

INSTM, Via G. Giusti, 9, I-50121 Firenze.

[c]

Center of Excellence for Nanostructured Materials (CENMAT) and INSTM, unit

of Trieste, Department of Chemical and Pharmaceutical Sciences, University of Trieste, via L. Giorgieri 1, 34127 Trieste, Italy: [email protected]

ACS Paragon Plus Environment

1

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

[d]

Page 2 of 49

Department of Biological and Environmental Sciences and Technologies,

DISTEBA, University of Salento, Via per Arnesano, I-73100 Lecce, Italy

[e]

Basque Fdn Sci, Ikerbasque, Bilbao 48013, Spain.

[f]

Carbon Nanobiotechnology Laboratory CIC biomaGUNE, Paseo de Miramón

182, 20009 Donostia-San Sebastian, Spain;

[g]

Department of Cultural Heritage, Università del Salento, Via D. Birago, 48, I-

73100 Lecce, Italy: [email protected]

ACS Paragon Plus Environment

2

Page 3 of 49 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

ABSTRACT. Perylene Bisimide (PBI) derivatives show peculiar physical chemical features, such as a highly conjugated system, high extinction coefficients and elevated fluorescence quantum yields, making them suitable for the development of optical sensors of compounds of interest. In particular, they are characterized by the tendency to aggregate into  stacked supramolecular structures. In this contribution, the behavior of the PBI derivative N,N’-bis(2-(trimethylammonium)ethylene)perylene bisimide dichloride was investigated both in aqueous solution and on solid support. The electronic communication among PBI aggregates and biogenic amines was exploited in order to discriminate aromatic amines down to sub-nanomolar concentrations by observing PBI fluorescence variations in presence of various amines and at different concentrations. The experimental findings were corroborated by DFT calculations. In particular, phenylethylamine and tyramine were demonstrated to be selectively detected down to 10-10 M concentration. Then, in order to develop a Surface Plasmon Resonance (SPR) device, PBI was deposited onto a SPR support by means of the layer-by-layer method. The PBI was deposited in the aggregated form and was demonstrated to

ACS Paragon Plus Environment

3

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 49

preserve the capability to discriminate, selectively and with an outstanding analytical sensitivity, tyramine in vapor phase and even if mixed with other aromatic amines.

KEYWORDS. Biogenic amines, Perylene Bisimides, Fluorescence spectroscopy,  stacking, Surface Plasmon Resonance, Layer-by-layer technique, DFT calculations.

1. Introduction

Amines are nearly ubiquitous analytes that can be found in a disparate number of areas, which range from biology (e.g. neurotransmitters, amino acids, etc.)

1-3,

diseases

(e.g. diagnosis of certain lung diseases, including cancer, kidney function, etc.) products (e.g. determination freshness of meat, fish, shellfish)

6-8,

4-5,

food

to industrial waste

(including the pharmaceutical, fertilizer, surfactant, and colorant production, to name a few) 9-10. Among amine compounds, biogenic amines (BAs) (e.g. histamine, tyramine, putrescine, cadaverine and phenylethylamine) are a class of compounds widely spread in food,

ACS Paragon Plus Environment

4

Page 5 of 49 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

beverage and human cells. Their presence in food and beverages is due to the decarboxylation of several amino acids. These BAs are potentially harmful to human health as inducing headaches, respiratory distress, heart palpitation, hyper or hypotension, and several allergenic disorders

11-12.

Therefore, the occurrence of BAs

could be harmful to the health of some sensitive consumers 13. Consequently, the detection and sensing of BAs has attracted attention, due to its importance in the environment and human well-being, and several protocols and strategies have been developed. Preferentially employed are detection methods based on chromatography, electrochemistry, chemiluminescence and capillary electrophoresis 14.

However, simpler, less time consuming, more portable or stable approaches have

been developed in order to overcome some of these issues

15-17.

These include

colorimetric sensors designed to recognize amines either by non-covalent interactions or chemical reactions (i.e. chemodosimeters)18-20 . Among the chromophores employed for the sensing of amines, there is a variety of choice, including Zn-porphyrins, coumarins, merocyanines, (poly)thiophenes, diarylethenes and perylene mono- and bisimides, to name a few 21-24.

ACS Paragon Plus Environment

5

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 6 of 49

Perylene-3,4,9,10-tetracarboxylic bisimide (PBI) analogues have been extensively investigated due to their outstanding optoelectronic properties, as well as their stability under thermal and oxidative stress

25-26.

PBI fluorophores have also been employed in

sensing applications, due to their electron accepting ability and high fluorescence quantum yield in the disassembled state

27-29.

PBIs with properly designed imide

substituents were employed as, for example, sensing metals (Hg2+, Pd2+, Cu2+, …)

30-33,

anions 34-35, polycyclic aromatics 25, pH 36-37, and amines 15, 38-42. Concerning the amines sensing, PBIs have been explored mostly for probing their vapors, since a photoinduced electron transfer (PET) can occur due to favorable energy difference between the PBI HOMO and amine HOMO energy levels, resulting in efficient fluorescence quenching

38, 40, 43-45.

Furthermore, some works exploited the ability of the PBIs to self-

assemble due to strong π-π interactions

46

and prepare aggregated one-dimensional

nanostructures: the long range exciton migration of well-organized fibers was found favorable for preparing thin films for sensing 47-49. In the present work we have exploited the strong π-π stacking interaction between N,N’bis(2-(trimethylammonium)ethylene)perylene bisimide dichloride (PBI, Figure S1)

ACS Paragon Plus Environment

6

Page 7 of 49 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

monomers in aqueous solution. Working under dilute conditions, in water, we explored the effect that biogenic amines (histamine, tyramine, tryptamine, phenylethylamine and putrescine) have on the photophysical properties of PBI aggregates. Further, the PBI derivative herein investigated, even without any ad hoc peripheral substituent, was demonstrated to be a very efficient active solid layer for BAs vapor detection. In fact, both the deposition and the transduction method allowed to obtain a sensing device able to detect down to nanomolar analyte concentration. In particular, the Layer-byLayer (LbL) method was exploited to fabricate PBI films since it is well-known as a humid deposition technique able to ensure a high control over the active layer deposition50-51. As transduction method, Surface Plasmon Resonance approach was used, at the best of our knowledge, for the first time to analyze the effect of BAs vapors on PBI supramolecular aggregation in thin film. In fact, BAs are able to modify the optical properties, such as the refractive index and the extinction coefficient, of the PBI film and this has been exploited to develop a sensing prototypal system based on the SPR effect. Such an effect, indeed, is intrinsically high sensitive to small changes in the spectroscopic features of a medium. In this context, the analysis of vapor phase amines

ACS Paragon Plus Environment

7

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 49

could be very useful for the analysis of real samples, for example, for food quality by analyzing the food spoilage avoiding sophisticated sample preparation.

2. Experimental Section

The PBI derivative was synthetized according to the literature.52-53 All the organic compounds (tyramine, tryptamine, histamine, phenylethylamine, putrescine, 1,2-phenylendiamine) were purchased from Sigma-Aldrich and used as received. Visible spectra were measured by a Cary 5000 (Agilent). Steady State Fluorescence spectroscopy was carried out by a Fluorolog (Horiba) spectrophotometer. PBI derivative films deposition was carried out by means of LbL technique. In particular, six layers of PBIs were deposited alternatively to polystyrene sulfonate (PSS), which was used as the negatively charged polyelectrolyte. The substrate was subsequently immersed in a beaker containing the suspensions in water of PBI (4 × 10-5 M, pH 8.5) and PSS at a speed of 8 mm min-1; after 3 min wait, the substrate was withdrawn (at the same speed) and, after other 3 min delay, washed with MilliQ grade

ACS Paragon Plus Environment

8

Page 9 of 49 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

water. The substrate dipping was monitored and controlled by means of a LangmuirBlodgett equipment KSV 5000 System 3 LB apparatus. The LbL transfer process was monitored by means of openQCM device (Novatech S.r.l.). 10 MHz AT-cut quartz crystal was used as resonance microbalance with a frequency standard deviation of 0.4 Hz. The frequency variation measurements after each LbL dipping were repeated on three different quartz crystal microbalance and averaged. Concerning the SPR measurements, the PBI film was transferred on SPR metal/glass substrate (Corning 7059, with a refractive index of 1.723 at 632.8 nm, gold thickness of 44 nm) by means of the LbL technique. The deposited film was used to record the Reflectance (%) vs Angle Of Incidence (AOI) curve and to measure the SPR angle at the starting point (AOI0). Then, the film was removed from the SPR cell and exposed to amine vapors at different concentrations (ppm) for 30 min and, afterwards, was placed again in the SPR cell for the measurement of the angle shift (AOIi). The SPR angle shift was calculated as the difference between AOI0 and AOIi. After monitoring the amine binding, the recovery of the derivative film was carried out upon exposing the active

ACS Paragon Plus Environment

9

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 49

layer to acid vapors for 30 min under ambient condition; then, the complete recovery of the initial angle was verified before carrying out further investigations. The SPR measurements were performed using a Nanofilm apparatus54-55 by varying the incident light angle from 54° to 58° at a fixed wavelength of 633 nm with a step of 0.05°, before and after exposing the film to amine vapors. A home-made gas control system was used to change the concentration of the investigated BAs.56 The system is made by a mass flow controller connected with a two mass flowmeters that are connected to an outside line that supplies N2 gas. One channel directly connects the N2 flow to the test chamber, another one channel drives the N2 gas in a vial that contains the analyte in liquid phase allowing to modulate the gas concentrations at a fixed total flow in the test chamber. FT-IR Spectrum One (PerkinElmer) in ATR mode was used to characterize the PBI derivative and the different interactions with the different amines. Cast films of PBI, BAs and PBI/BAs mixtures were obtained depositing a drop of the aqueous solution directly on the ATR prism. The PBI film deposited on the SPR support and its interaction with

ACS Paragon Plus Environment

10

Page 11 of 49 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

phenylethylamine and tyramine were also characterized by means of FT-IR spectroscopy by using an AmplifIR accessory distributed by SensIR Technologies57.

3. Results and Discussion

3.1 Spectroscopic Characterization of PBI in aqueous medium

The dicationic PBI derivative, in aqueous solutions, tends to aggregate due to the strong π-π interactions58. In dilute solutions (c = 1 × 10-6 M), the perylene bisimide tends to form dimers, while at higher concentrations the formation of cylindrical nanocrystals is observed, through assembly along the π-π stacking direction. As a preliminary characterization, PBI aqueous solutions at different concentrations were analyzed by means of UV-Vis absorption and the obtained spectra are reported in Figure 1a. Two main absorption peaks are evidenced in all the studied concentrations in the investigated range (400-650 nm) and are located at about 537 and 500 nm, in accordance to previously reported absorbance of PBI derivatives

25.

In particular, these

ACS Paragon Plus Environment

11

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 49

two contributions can be assigned to the 0→0 and the 0→1 vibronic bands of S0→S1 transitions, respectively

58.

Then, two concentrations were deeply investigated, in

particular 4 × 10-5 M and 2 × 10-6 M. These two concentrations were chosen in order to confirm the dependence of the aggregation state on PBI concentration, and in particular the 4 × 10-5 M solution was chosen as the upper absorbance value threshold (around 1) and the 2 × 10-6 M solution as the lower (around 0.05). These two spectra are again reported in Figure 1b, in order to better show the Vis light absorption features (black and red lines correspond to higher and lower concentrations, respectively). The lower concentration solution spectrum is shown multiplied by 20. More in detail, the two aforementioned bands are well resolved in both cases and in the case of the higher concentration, a red shift of about 5 nm is recorded in the case of the 0→0 vibronic band, highlighting a stronger aggregation 58. Moreover, the ratio between the intensities of the two bands, A0→0 /A0→1,59-60 can be calculated and was found to predict the aggregation state of PBI molecules. Such a ratio has to be around 1.6 for the monomeric form of PBI and equal or lower than 0.7 upon aggregation. The A0→0 /A0→1 value calculated for the higher concentration (black

ACS Paragon Plus Environment

12

Page 13 of 49 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

spectrum, Figure 1b), 0.59, confirms the presence of nanotubular aggregates, whilst such an aggregation results drastically reduced for the 2 × 10-6 M solution (red spectrum, Figure 1b), since the bands intensity ratio is 1.39, indicating a smaller aggregation state. Perylene bisimides, substituted with unbranched alkyl chains in the imide position, typically self-assembles into parallel stacked H-type geometry thanks to  interactions.61-63 The presence of such a supramolecular structure is related to particular optical and electronic features that can be exploited for different applications.

ACS Paragon Plus Environment

13

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 49

Figure 1. a) Absorption spectra in the 400-650 nm range of PBI aqueous solutions at different concentrations (from 4 x 10-5 M to 2 x 10-6 M); b) Absorption spectra in the 400650 nm range of PBI aqueous solutions at the two different investigated concentration: 4 × 10-5 M, black line, and 2 × 10-6 M, red line; c) Fluorescence spectra of PBI aqueous solutions at the two different investigated concentration: 4 × 10-5 M, black line, and 2 × 10-6 M, red line (exc = 500 nm).

The different aggregation states were further confirmed by analyzing the fluorescence emissions of both solutions, upon 500 nm excitation (Figure 1c). The presence of the water medium at the investigated concentration of 4 × 10-5 M clearly induces the formation of aggregates and a fluorescence self-quenching (black spectrum, Figure 1c). Furthermore, the emission intensity of the more diluted solution, tested in the same conditions, results higher (red line, Figure 1c).

3.2 DFT calculations

ACS Paragon Plus Environment

14

Page 15 of 49 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

Computational modelling based on the dimer approach was performed, which can be regarded as the smallest aggregate and has been successfully used to understand optical spectra of perylene-based aggregates26,

64-66.

We optimized the dimer (and a

monomer for reference) using the dispersion-corrected density functional theory (DFT) using

Grimme’s

PBEh-3c

approach

67

(Figure

2ab).

The

dimer

represents

perpendicularly (α = 90°) stacked two PBI molecules, distanced 0.33 nm, with a torsion angle β = 33°. We then employed time-dependent DFT (using the long-range corrected ωB97X functional and 6-31G** basis set68) to describe the charge-transfer excitations and obtain corresponding absorption spectra, as depicted in Figure 2c. Compared to the PBI monomer (λmax = 552 nm), the dimer showed a strong H-band transition (at 528 nm), as well as a weaker J-band transition centred at higher wavelengths (at 621 nm). While the former, blue-shifted strong, absorption band is related to H-type aggregates, the latter, partially allowed red-shifted, optical transition is rising from the rotational displacement between the two dye molecules, as rationalized by exciton theory

69-70.

Given the simplicity of the dimer approach, the absorption profiles resemble well the experimental data and confirm the formation of H-stack aggregates.

ACS Paragon Plus Environment

15

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 49

Figure 2. Results from DFT (a,b) and TD-DFT (b) calculations. ab) Top and side view of the optimized perylene bisimide H-stacked dimer (PBEh-3c): c) TD-DFT results (ωB97X) for the monomer (red line) and the dimer (blue line).

We then proceeded to DFT calculations (at the same level of theory performed for the PBIs) on the amines presented in this work (vide infra) and plotted the energies of the frontier orbitals, in order to elucidate if electron transfer could occur from the amines to the PBI aggregates (Figure S2). For all the amines it was found that the highest occupied molecular orbitals were sufficiently high in energy to donate electrons to the PBI dimer. This is true also for protonated amines (Figure S2), except putrescine, as expected for their aqueous solutions, pointing to the importance of the aromatic rings as

ACS Paragon Plus Environment

16

Page 17 of 49 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

electron donors, which could eventually result in the quenching of the PBI fluorescence. Therefore, we proceeded with carrying out the appropriate experiments.

3.3 Study of the Interaction among PBI and Biogenic Amines

In order to study the interaction between PBI and BAs, the compound was dissolved in MilliQ grade water at a concentration of 4 × 10-5 M (pH ~ 8) and mixed with the different BAs. The mixture was characterized by means of steady-state fluorescence spectroscopy. In particular, five BAs were tested: putrescin (as aliphatic BA), tyramine and phenylethylamine (as aromatic BAs), histamine and tryptamine (as heterocyclic BAs). The chemical structures of all the investigated BAs are reported in Figure 3.

ACS Paragon Plus Environment

17

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 49

Figure 3. The chemical structures of the investigated BAs: a) putrescine; b) phenylethylamine, c) tryptamine, d) histamine, e) tyramine.

The emission intensity at 550 nm and 590 nm (exc = 500 nm) of the PBI investigated aqueous solution was tested in the presence of different concentrations of BAs, in a range comprised between 10-4 M and 10-10 M. The results show that the effect of amines on PBI can be divided into two groups: (i) the first, made by tyramine, histamine and tryptamine, has been demonstrated to be able to quench PBI fluorescence, and (ii) the second one, including putrescine and phenylethylamine, able to enhance the fluorescence emission. In detail, PBI has been proved to be very sensitive to aromatic amines. In fact, Figure 4 reports the fluorescence spectra of PBI in presence of different concentrations of phenylethylamine (a) and of tyramine (b).

ACS Paragon Plus Environment

18

Page 19 of 49 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 4. Fluorescence spectra of a) PBI 4 × 10-5 M aqueous solution (blue line) and in presence of different concentrations of phenylethylamine and b) PBI 4 × 10-5 M aqueous solution (blue line) and in presence of different concentrations of tyramine.

ACS Paragon Plus Environment

19

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 49

In the first case (Figure 4a), an enhancement of PBI fluorescence has been recorded down to 10-10 M of phenylethylamine, with a sub-nanomolar detection of this compound in aqueous medium. Besides, as it can be observed from Figure 4b, tyramine induces a PBI fluorescence quenching even down to 10-10 M amine concentration, demonstrating again a very interesting sub-nanomolar detection limit.

The fluorescence spectra of PBI in presence of putrescine (a), histamine (b) and tryptamine (c), in a concentration range comprised between 10-4 and 10-7 M, are reported in Figure S3. In fact, the detection limit for these BAs was higher by three magnitude orders (10-7 M) when compared to the aromatic ones (10-10 M), suggesting an important role played by the aromatic ring in the sensing mechanism, that could be hypothesized to affect the  stacked aggregates. In fact, putrescine has been demonstrated to be able to enhance PBI fluorescence up to 10-7 M (Figure S3a), while heterocyclic amines bring about a quenching of PBI fluorescence down to 10-7 M.

In order to elucidate the sensing mechanisms, FT-IR spectroscopy was carried out for cast films, obtained directly on the ATR plate, of PBI aqueous solution (Figure 5a) and

ACS Paragon Plus Environment

20

Page 21 of 49 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

PBI aqueous solution in presence of phenylethylamine (10-4 M), tyramine (10-4 M) and putrescine (10-4 M) (Figure 5b). The FT-IR spectrum of PBI, reported in the frequency range comprised between 1850 and 750 cm-1, is dominated by the typical ring modes of the aromatic compounds (1600-1350 cm-1), CH in-plane bending modes bands (in the 1200-1000 cm-1 range) and CH out-of-plane bending vibrations between 1000 and 700 cm-1. Moreover, the imide C=O stretching bands are well-defined in the region comprised between 1730 and 1655 cm-1 and the imide C-N stretching modes is located at about 1180 cm-1. Finally, the C-N stretching mode of the tertiary amine of the tetraalkyl ammonium peripheral substituents is present at 1265 cm-1 71.

ACS Paragon Plus Environment

21

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 22 of 49

Figure 5. a) FTIR spectrum of a cast film of PBI directly on the ATR plate; the principal peaks are highlighted and discussed in the main text; b) FTIR spectra of cast films of PBI and PBI in presence of phenylethylamine, tyramine and putrescine.

ACS Paragon Plus Environment

22

Page 23 of 49 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 interaction mechanism can be supposed to be ruled by the electronic communication between the electron-rich amine group of BAs and the electron-poor PBI core. Recently, crown-ether bearing PBI has been found to be a suitable fluorescent probe for amino acids due to charge-transfer (CT) interaction of aromatic moieties with the PBI core, resulting in fluorescent quenching.42 Furthermore, in some cases, formation of hydrogen bonds between ammonium groups (of protonated amines) and imides from PBI can be hypothesized. Therefore, we observe changes in the FT-IR spectra obtained right after putrescine, phenylethylamine and tyramine interacted with PBI compound, in the attempt to understand the types of interactions that ultimately result in the discrimination of the BAs. In fact, in the case of phenylethylamine, the binding affects both C=O modes of the imides, underlining the key role played by the imide group in the interaction. Moreover, all the ring modes are strongly affected, suggesting a further robust  communication. Instead, in the case of putrescine, we can observe changes only in the C=O modes of the imides since such a diamine is aliphatic and the absence of the aromatic ring did not allow the further interaction. The lack of the  interaction leads to a higher detection

ACS Paragon Plus Environment

23

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 24 of 49

limit in the case of putrescine (10-7 M), if compared with phenylethylamine detection limit (10-10 M). When interaction between PBI and tyramine occurs, although such amine is aromatic, the perylene bisimide IR features are little affected (as in the case of the phenylethylamine), underlining the presence of a completely different interaction mechanism: tyramine aromatic ring cloud is partially occupied due to the presence of the p-OH substituents; the presence of a hydroxy, electron inductive, group might be responsible for a higher energy HOMO level of the aromatic ring, therefore leading to electron transfer to the PBIs, finally resulting in fluorescence quenching of the former. Nevertheless, the system remains characterized by sub-nanomolar detection limit of 1010

M, even in this case. Besides charge transfer, this result can be explained by

hypothesizing  interaction of tyramine with the PBI core as well, inducing a stronger aggregation. This is true also for heterocycle-bearing amines, in which the highest occupied molecular orbitals are sufficiently high in energy to cause a photo-induced electron transfer to the HOMO level of the PBIs, as supported by DFT calculations. All the detected changes in the PBI spectra are due to the complex formation and not only to the presence in the mixture of the amines, since the IR spectra of the bare amines

ACS Paragon Plus Environment

24

Page 25 of 49 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

have been recorded (Figure S4) and turn out to be different from the relative complex spectrum. An aggregation/disaggregation sensing mechanism, depending on both the BAs type (aromatic or not, for instance) and the correspondent HOMO level could be rationalized. In the case of phenylethylamine, by means of  interaction and the communication among ammonium and the imide groups, the amine molecules could be intercalated inside the nanotubular aggregates of perylene partially disaggregating the PBI stacked aggregates, inducing an increase of the concentration of the monomeric form in solution. Putrescine showed the same effect, although with lower sensitivity imputable to the lack of the aromatic ring. Indeed, due to the HOMO level of very near similarity to the one of the PBI, a recombination of the electron could takes place. This also can lead to less packed PBIs, ultimately resulting in an enhancement of the fluorescence. Otherwise, in the case of the other aromatic or heterocyclic BAs, this cannot take place due to higher energy difference between the HOMOs. In fact, tyramine, due to its chemical structure and HOMO level position, has been hypothesized to transfer an

ACS Paragon Plus Environment

25

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 26 of 49

electron to the HOMO level of PBI and to induce further aggregation, causing the quenching of the fluorescent emission.

In order to confirm the hypothesis that the presence of phenylethylamine partially disassembles the aggregates of PBI in water solution, a UV-Vis spectrum of an aqueous mixture containing PBI (c = 4 × 10-5 M) and phenylethylamine (10-4 M) was recorded and the A0→0 /A0→1 was calculated, obtaining a value equal to 0.73. The solution of only PBI, at the same concentration, was characterized by a A0→0 /A0→1 value of 0.59 and, considering that a value below 0.7 is imputable to aggregated structures72, the increase of this value is in agreement with the hypothesis that such an amine induces a separation of the  stacked domains. Similarly, the A0→0 /A0→1 has been calculated for the UV-Vis spectrum of a mixture of PBI (c = 4 × 10-5 M) and tyramine (104

M). In this case, the value resulted further decreased to 0.58, underlining that tyramine

does not reduce the aggregation, but probably upgrades stacking among PBI aggregates.

ACS Paragon Plus Environment

26

Page 27 of 49 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

3.4 LbL PBI deposition process

In order to transfer the PBI from aqueous solution (4x10-5 M) onto solid supports, the LbL approach was chosen. This approach ensured the alternate transfer of the hydrophilic compound bearing positive charges with layers of polystyrene sulfonate (PSS, used as negatively charged polyelectrolyte). The transfer process was monitored by means of Quartz Crystal Microbalance (QCM) technique. According to the Sauerbrey equation73, the frequency variation (f) of the QCM is proportional to the mass deposited on the resonator during the transfer process74. In Figure S5a the frequency change has been plotted as a function of the number of LbL runs. The variation in the frequency values during the formation of the PSS-PBI assembly was not linearly dependent on the LbL layers’ number and f decreases during the runs repetitions. This alteration from the linear trend would suggest that PSS is not able to ensure a complete cover of the PBI layer that is partially removed during the washing step75. This tendency reaches an asymptote after 6 LbL runs, after then very small frequency variations are recorded. For this reason, the

ACS Paragon Plus Environment

27

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 49

number of LbL runs used to assembly the active layer was fixed at 6, in order to maximize the number of active sites for the further investigations. The deposition success was further confirmed by means of UV-Vis spectroscopy analysis. Figure S5b reports the spectrum of 6 layers of PBI deposited onto a glass substrate in the 400-550 nm region. The features of this spectrum clearly evidence that the PBI is deposited in the aggregated form.

3.5 SPR-based Amine Vapors Detection

The sensing mechanism dominated by a disaggregation event in the case of phenylethylamine and a charge transfer in the case of tyramine was also evaluated for PBI immobilized onto a solid support, for the development of SPR-based detection system for different amine vapor concentrations. First of all, the sensing performance of the device was tested by exposure to saturated amine vapors. The variation of the SPR angle was monitored at 632 nm upon amine interaction with the film as detailed in the experimental section. Figure 6a shows the

ACS Paragon Plus Environment

28

Page 29 of 49 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

Reflectance (%) variation recorded for the PBI film (black line) and after the interaction with phenylethylamine (2.5 ppm, blue line) and tyramine (0.02 ppm, green line). Depending on the different interaction mechanism, the shift of the SPR angle induced by tyramine was in the opposite direction if compared with the shift induced by phenylethylamine. In fact, the detection of small molecules by means of SPR transduction method can be mainly obtained as a consequence of the changes of the chemical and the physical features of the active layer. For example, oxidation processes promoted by the analyte allowed to detect the presence of uric acid in sub-micromolar concentrations76 as well as the active layer swelling and de-aggregation phenomena were used for small molecules detection.77-78

ACS Paragon Plus Environment

29

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 49

Figure 6. a) SPR curves obtained for a six layers film of PBI (black curve), after phenylethylamine binding (blue line) and tyramine binding (green line); b) FTIR spectra of six layers of PBI derivative obtained onto SPR substrate (in the middle), after binding with phenylethylamine (at the top) and with tyramine (at the bottom).

This peculiar behavior can be rationalized considering the relation between the intensity of SPR wave, photon incident angle (), medium refractive index (n1), gold thin layer refractive index (ng) and deposited thin film refractive index (n2) 79:

𝜃 = 𝑠𝑖𝑛

(

―1

)

1 𝑛22𝑛2𝑔 𝑛1 𝑛22 + 𝑛2𝑔

As it can be deduced, a shift towards higher incident angles implies the refractive index of the thin film after the interaction with the analyte is higher than n2 of the bare PBI film. According to Würthner and co-workers, PBI H-aggregates show lower extinction coefficient then the monomeric form

25.

So, SPR experimental evidences

suggest that phenylethylamine molecules induce a partial disaggregation of PBI stacks,

ACS Paragon Plus Environment

30

Page 31 of 49 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

even in solid film

21,

according to the rationale proposed by means of the

absorption/emission measurements. A different mechanism can be proposed accounting for the relevant shift of the plasmonic peak towards lower angle values obtained when PBI film is interacting with tyramine. Probably charge transfer from tyramine towards PBI induces a reduction of the active layer with a consequent shift of  towards lower values80-81. As a further characterization, the temperature-dependence of the obtained SPR curve for the PBI film was tested and Figure S6 reports the SPR curves obtained at different temperatures, i.e. 4 °C - 10 °C - 20 °C - 30 °C. A good thermal stability has been demonstrated. FT-IR spectra were recorded in order to better investigate the interaction between BAs vapors and PBI active layer. IR bands of the film upon exposure to saturated vapor of phenylethylamine and tyramine are shown in Figure 6b, top and down respectively. The ratio of the intensity of the two imide C=O bands changes as in the case of solution only for phenylethylamine, further corroborating the proposed sensing mechanism. The PBI film was, then, systematically exposed to increasing vapor amine concentrations (ppm) only in the case of tyramine and phenylethylamine, since the other

ACS Paragon Plus Environment

31

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 49

investigated biogenic amines (not aromatic) did not induce detectable SPR angle variation. The obtained results are reported in Figure 7.

ACS Paragon Plus Environment

32

Page 33 of 49 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 7. a) SPR angle shift recorded for different concentration of tyramine (black spots) and phenylethylamine (blue spots) in vapor phase. All measurements are shown as the average of three different experiments; b) SPR angle shift vs amine concentration calibration curves obtained in the investigated static range of response for tyramine (black) and phenylethylamine (blue) in vapor phase.

The PBI-based SPR sensor was tested in a concentration range comprised between 0.002 and 2.5 ppm in the case of tyramine (black dots in Figure 7a) and between 0.01 and 2.5 ppm in the case of phenylethylamine (blue dots in Figure 7a). In this extended concentration range, in both cases, the SPR angle shift was demonstrated to increase until reaching a plateau phase, so a dynamic response was recorded

82.

If the behavior,

from a qualitative point of view, is comparable for the two investigated amines, the SPR detection system was demonstrated to have higher analytical sensitivity and a lower limit of detection (LOD) for tyramine vapors. In fact, the experimental LOD was found to be 0.002 ppm (ca. 10 nM) and 0.01 ppm (ca. 90 nM) for tyramine and phenyltethylamine, respectively. This result was further confirmed by analyzing the

ACS Paragon Plus Environment

33

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 49

concentration static range of response for the amines: 0.002-0.05 ppm for tyramine and 0.01-0.5 ppm for phenylethylamine. In this range, linear fits have been obtained with very good correlation coefficients (>0.99); the curves are reported in Figure 7b. By discussing the different obtained slope values, about 12°/ppm and about 0.8°/ppm for tyramine and phenylethylamine, respectively, the greater sensitivity for tyramine could be assessed. Probably, the interaction mechanism based on a charge transfer towards the HOMO level of PBI resulted easier detectable by using this transduction method if compared to the phenylethylamine binding mechanism. The SPR system was, indeed, used for the detection of vapor of 1,2-phenylendiamine (0.05 ppm), as potential aromatic interfering agent, and for a vapor mixture of 1,2phenylendiamine, tyramine and phenylethylamine at 0.05 ppm. The concentration of 0.05 ppm was chosen in order to work in the static range for both amines. 1,2phenylendiamine induced a SPR angle shift of 0.09°, a value considerably lower than the angle shift induced at 0.05 ppm by the other amines (0.76° and 0.35° for tyramine and phenylethylamine). We could suppose that, nonetheless the presence of two NH2 groups, a cooperative effect should be excluded, and, more probably, the interaction

ACS Paragon Plus Environment

34

Page 35 of 49 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

results “sterically demanding”. The higher molecule rigidity, in this case, could reduce the freedom degrees as well as the interaction among analytes and PBI aggregates. Finally, very interestingly, the vapor mixture exposure induced a SPR shift of 0.8°, a value totally comparable to the one got in the case of 0.05 ppm tyramine vapor exposure. This evidence proposes the PBI-based device as a selective tool for tyramine sensing, otherwise the opposite SPR response in presence of tyramine or phenylethylamine allows to discriminate the contribution of the two different BAs. As a further characterizations, the effect of humidity on the sensor sensibility towards tyramine was evaluated. The analyte concentration was kept at 1 ppm and SPR angle shift was recorded at relative humidity of 20%, 40%, 60% and 80%. The relative humidity did not significantly influenced the device response (Figure S7a). Again, same procedure was repeated for a tyramine concentration of 0.03 ppm, that corresponds to the linear range of response, and as observed for higher concentration the humidity effect on the device response can be considered negligible (Figure S7b). The high analytical sensitivity and the binding mechanism of the SPR system, in the case of tyramine, probably could make it selective for this amine, avoiding the detection

ACS Paragon Plus Environment

35

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 36 of 49

of the other aromatic compounds, proposing this system for the engineering of a future SPR sensor for tyramine in vapor phase based on the PBI derivative LbL film.

4. Conclusions

A PBI derivative has been used to detect BAs in aqueous media by exploiting its physicochemical

characteristics,

including

the

capability,

depending

on

the

concentration, to form  stacked tubular aggregates, which can be followed by fluorescence and UV-Vis absorption spectroscopy. In fact, the monomer has a higher quantum yield if compared to the aggregates

25.

The interaction mechanism between

PBI in a partially aggregated aqueous solution and different kinds of BAs (aliphatic, heterocyclic or aromatic) were investigated by means of spectroscopic techniques and confirmed by DFT calculations. Putrescine and phenylethylamine were shown to be able to enhance the PBI fluorescence, and histamine, tryptamine and tyramine to quench its fluorescence. In details, phenylethylamine and tyramine can be detected

ACS Paragon Plus Environment

36

Page 37 of 49 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

down to 10-10 M. FTIR spectroscopy has evidenced that phenylethylamine binding mechanism involves the whole PBI molecule starting from the electron-poor core; clear changes in the C=O stretching modes and aromatic core modes are exhibited. Instead, tyramine, notwithstanding the comparable limit of detection, is characterized by a different interaction mechanism, since C=O imide groups modes are the most affected upon interaction. The two behaviors could be rationalized considering both BAs  delocalization grade and HOMO position, as reported in the literature for similar systems

43-44.

Phenylethylamine has been supposed to partially disaggregate the 

aggregates, reintroducing PBI fluorescence, even by reason of the HOMO level of very close-by similarity to the one of the PBI. Tyramine, instead, due to the presence of a pOH as a substituent on the benzene ring and HOMO level position, could be hypothesized to transfer an electron to the PBI HOMO level, inducing further aggregation. Such results were also confirmed by Surface Plasmon Resonance experiments. Six alternate layers of PBI derivative and PSS were deposited by means of the layer-by-layer technique onto glass and SPR solid substrates. The PBI was deposited in the aggregated form and has been demonstrated to preserve the ability to

ACS Paragon Plus Environment

37

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 38 of 49

interact and discriminate the aromatic amines even in vapor phase. In fact, after the exposure to phenylethylamine and tyramine saturated vapors, a SPR angle shift was recorded in both cases in two opposite directions. The shift towards lower angles, due to tyramine, could be ascribable to an electron transfer to PBI compound

80-81;

on the

other hand, the shift towards higher angles, upon interaction with phenylethylamine, could be put down to an increment of the concentration of PBI monomeric form, characterized by a higher refractive index. The possibility to deposit a water-soluble derivative and to use it for the detection of analytes even in vapor phase represents an important issue for the development of real sensors for quality food assessment without samples pre-preparation. In this regard, the developed system was exposed to vapors of 1,2-phenylendiamine, as potential aromatic interfering compound, and to a mixture of 1,2-phenylendiamine,

tyramine

and

phenylethylamine.

The

SPR

system

was

demonstrated to have a high sensitivity to tyramine, which is the unique amine detected even in presence of the interfering compounds.

ASSOCIATED CONTENT

ACS Paragon Plus Environment

38

Page 39 of 49 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

Supporting Information. The file reports the chemical structures of the utilized PBI derivative; the fluorescence spectra of PBI aqueous solution in presence of different concentration of putrescine, histamine and tryptamine, FTIR spectra of putrescine, phenylethylamine and tyramine; the optimized DFT geometries; the UV-Vis and QCM characterization of the deposited film; SPR curves recorded for the 6 LbL runs PBI films at different temperatures and the effect of the relative humidity on the sensor response injecting 1 ppm and 0.03 ppm of tyramine.

AUTHOR INFORMATION

Corresponding Authors *Maurizio Prato [email protected] *Gabriele Giancane [email protected]

Author Contributions

ACS Paragon Plus Environment

39

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 40 of 49

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. ‡These authors contributed equally.

Present Address

† Simpson

Querrey Institute, Northwestern University, 303 E. Superior, Chicago, IL, USA

ACKNOWLEDGMENT This research was supported by “FutureInResearch” APQ Ricerca Regione Puglia.

REFERENCES

1.

Mallick, S.; Chandra, F.; Koner, A. L., A ratiometric fluorescent probe for

detection of biogenic primary amines with nanomolar sensitivity. Analyst 2016, 141 (3), 827-831. 2.

Bachrach, U.; Wang, Y. C.; Tabib, A., Polyamines: new cues in cellular signal

transduction. News in physiological sciences : an international journal of physiology

produced jointly by the International Union of Physiological Sciences and the American Physiological Society 2001, 16, 106-9. 3.

Childs, A. C.; Mehta, D. J.; Gerner, E. W., Polyamine-dependent gene

expression. Cellular and molecular life sciences : CMLS 2003, 60 (7), 1394-406.

ACS Paragon Plus Environment

40

Page 41 of 49 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

4.

Agostinelli, E.; Tempera, G.; Molinari, A.; Salvi, M.; Battaglia, V.; Toninello, A.;

Arancia, G., The physiological role of biogenic amines redox reactions in mitochondria. New perspectives in cancer therapy. Amino acids 2007, 33 (2), 175-87. 5.

Stepien, M.; Duarte-Salles, T.; Fedirko, V.; Floegel, A.; Barupal, D. K.; Rinaldi,

S.; Achaintre, D.; Assi, N.; Tjonneland, A.; Overvad, K.; Bastide, N.; Boutron-Ruault, M. C.; Severi, G.; Kuhn, T.; Kaaks, R.; Aleksandrova, K.; Boeing, H.; Trichopoulou, A.; Bamia, C.; Lagiou, P.; Saieva, C.; Agnoli, C.; Panico, S.; Tumino, R.; Naccarati, A.; Bueno-de-Mesquita, H. B.; Peeters, P. H.; Weiderpass, E.; Quiros, J. R.; Agudo, A.; Sanchez, M. J.; Dorronsoro, M.; Gavrila, D.; Barricarte, A.; Ohlsson, B.; Sjoberg, K.; Werner, M.; Sund, M.; Wareham, N.; Khaw, K. T.; Travis, R. C.; Schmidt, J. A.; Gunter, M.; Cross, A.; Vineis, P.; Romieu, I.; Scalbert, A.; Jenab, M., Alteration of amino acid and biogenic amine metabolism in hepatobiliary cancers: Findings from a prospective cohort study. International journal of cancer 2016, 138 (2), 348-60. 6.

Shalaby, A. R., Significance of biogenic amines to food safety and human health.

Food Res Int 1996, 29 (7), 675-690. 7.

Biji, K. B.; Ravishankar, C. N.; Venkateswarlu, R.; Mohan, C. O.; Gopal, T. K. S.,

Biogenic amines in seafood: a review. J Food Sci Tech Mys 2016, 53 (5), 2210-2218. 8.

Doeun, D.; Davaatseren, M.; Chung, M. S., Biogenic amines in foods. Food Sci

Biotechnol 2017, 26 (6), 1463-1474. 9.

Muz, M.; Dann, J. P.; Jager, F.; Brack, W.; Krauss, M., Identification of Mutagenic

Aromatic Amines in River Samples with Industrial Wastewater Impact. Environ Sci

Technol 2017, 51 (8), 4681-4688. 10.

Jurado-Sanchez, B.; Ballesteros, E.; Gallego, M., Occurrence of aromatic amines

and N-nitrosamines in the different steps of a drinking water treatment plant. Water Res 2012, 46 (14), 4543-4555. 11.

Santos, M. H. S., Biogenic amines: Their importance in foods. Int J Food

Microbiol 1996, 29 (2-3), 213-231. 12.

Smit, A. Y.; du Toit, W. J.; du Toit, M., Biogenic Amines in Wine: Understanding

the Headache. S Afr J Enol Vitic 2008, 29 (2), 109-127. 13.

Guo, Y. Y.; Yang, Y. P.; Peng, Q.; Han, Y., Biogenic amines in wine: a review. Int

J Food Sci Tech 2015, 50 (7), 1523-1532.

ACS Paragon Plus Environment

41

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

14.

Page 42 of 49

Costa, M. P.; Balthazar, C. F.; Rodrigues, B. L.; Lazaro, C. A.; Silva, A. C. O.;

Cruz, A. G.; Conte, C. A., Determination of biogenic amines by high-performance liquid chromatography (HPLC-DAD) in probiotic cow's and goat's fermented milks and acceptance. Food Sci Nutr 2015, 3 (3), 172-178. 15.

Sriramulu, D.; Valiyaveettil, S., Perylene derivatives as a fluorescent probe for

sensing of amines in solution. Dyes Pigments 2016, 134, 306-314. 16.

Buccolieri, A.; Bettini, S.; Salvatore, L.; Baldassarre, F.; Ciccarella, G.; Giancane,

G., Sub-nanomolar detection of biogenic amines by SERS effect induced by hairy Janus silver nanoparticles. Sensor Actuat B-Chem 2018, 267, 265-271. 17.

Nelson, T. L.; O'Sullivan, C.; Greene, N. T.; Maynor, M. S.; Lavigne, J. J., Cross-

Reactive Conjugated Polymers:  Analyte-Specific Aggregative Response for Structurally Similar Diamines. J. Am. Chem. Soc. 2006, 128 (17), 5640-5641. 18.

A., R. N.; Avijit, S.; C., J. M.; B., P. J.; S., S. K., Molecular Recognition and

Discrimination of Amines with a Colorimetric Array. Angew. Chem. Int. Ed. 2005, 44 (29), 4528-4532. 19.

Zucchero, A. J.; McGrier, P. L.; Bunz, U. H. F., Cross-Conjugated Cruciform

Fluorophores. Acc. Chem. Res. 2010, 43 (3), 397-408. 20.

Askim, J. R.; Mahmoudi, M.; Suslick, K. S., Optical sensor arrays for chemical

sensing: the optoelectronic nose. Chem. Soc. Rev. 2013, 42 (22), 8649-8682. 21.

Giancane, G.; Borovkov, V.; Inoue, Y.; Conoci, S.; Valli, L., Syn-anti conformation

switching of a bis-porphyrin derivative at the air-water interface and in the solid state as an effective tool for chemical sensing. Soft Matter 2013, 9 (7), 2302-2307. 22.

Hariharan, P. S.; Pitchaimani, J.; Madhu, V.; Anthony, S. P., Perylene Diimide

Based Fluorescent Dyes for Selective Sensing of Nitroaromatic Compounds: Selective Sensing in Aqueous Medium Across Wide pH Range. J Fluoresc 2016, 26 (2), 395-401. 23.

Rella, R.; Rizzo, A.; Licciulli, A.; Siciliano, P.; Troisi, L.; Valli, L., Tests in

controlled atmosphere on new optical gas sensing layers based on TiO2/metalphthalocyanines hybrid system. Mat Sci Eng C-Bio S 2002, 22 (2), 439-443. 24.

Buccolieri, A.; Manno, D.; Serra, A.; Santino, A.; Hasan, M.; Borovkov, V.;

Giancane, G., Highly sensitive conformational switching of ethane-bridged mono-zinc

ACS Paragon Plus Environment

42

Page 43 of 49 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

bis-porphyrin as an application tool for rapid monitoring of aqueous ammonia and acetone. Sensor Actuat B-Chem 2018, 257, 685-691. 25.

Wurthner, F.; Saha-Moller, C. R.; Fimmel, B.; Ogi, S.; Leowanawat, P.; Schmidt,

D., Perylene Bisimide Dye Assemblies as Archetype Functional Supramolecular Materials. Chem Rev 2016, 116 (3), 962-1052. 26.

Burian, M.; Rigodanza, F.; Demitri, N.; D̵ord̵ević, L.; Marchesan, S.; Steinhartova,

T.; Letofsky-Papst, I.; Khalakhan, I.; Mourad, E.; Freunberger, S. A.; Amenitsch, H.; Prato, M.; Syrgiannis, Z., Inter-Backbone Charge Transfer as Prerequisite for LongRange Conductivity in Perylene Bisimide Hydrogels. ACS Nano 2018, 12 (6), 58005806. 27.

Jones, B. A.; Ahrens, M. J.; Yoon, M. H.; Facchetti, A.; Marks, T. J.; Wasielewski,

M. R., High-mobility air-stable n-type semiconductors with processing versatility: Dicyanoperylene-3,4 : 9,10-bis(dicarboximides). Angew Chem Int Edit 2004, 43 (46), 6363-6366. 28.

Zhao, C. T.; Zhang, Y. X.; Li, R. J.; Li, X. Y.; Jiang, J. Z., Di(alkoxy)- and

di(alkylthio)-substituted

perylene-3,4;9,10-tetracarboxy

diimides

with

tunable

electrochemical and photophysical properties. J Org Chem 2007, 72 (7), 2402-2410. 29.

Dwivedi, A. K.; Pandeeswar, M.; Govindaraju, T., Assembly Modulation of PDI

Derivative as a Supramolecular Fluorescence Switching Probe for Detection of Cationic Surfactant and Metal Ions in Aqueous Media. Acs Appl Mater Inter 2014, 6 (23), 2136921379. 30.

Singh, P.; Mittal, L. S.; Kumar, S.; Bhargava, G.; Kumar, S., Perylene Diimide

Appended with 8-Hydroxyquinoline for Ratiometric Detection of Cu2+ Ions and Metal Displacement Driven "Turn on" Cyanide Sensing. J Fluoresc 2014, 24 (3), 909-915. 31.

Wang, H. X.; Lang, Y. H.; Wang, H. X.; Lou, J. J.; Guo, H. M.; Li, X. Y., Perylene

diimide based 'turn-on' fluorescence sensor for detection of Pd2+ in mixed aqueous media. Tetrahedron 2014, 70 (11), 1997-2002. 32.

Malkondu, S.; Erdemir, S., A novel perylene-bisimide dye as "turn on" fluorescent

sensor for Hg2+ ion found in DMF/H2O. Dyes Pigments 2015, 113, 763-769.

ACS Paragon Plus Environment

43

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

33.

Page 44 of 49

Cheng, H. R.; Qian, Y., Two novel rhodamine-perylenediimide fluorescent

probes: Synthesis, photophysical properties, and cell imaging. Chinese Chem Lett 2016, 27 (6), 879-886. 34.

Barendt, T. A.; Ferreira, L.; Marques, I.; Felix, V.; Beer, P. D., Anion- and

Solvent-Induced Rotary Dynamics and Sensing in a Perylene Diimide [3]Catenane. J

Am Chem Soc 2017, 139 (26), 9026-9037. 35.

Goodson, F. S.; Panda, D. K.; Ray, S.; Mitra, A.; Guha, S.; Saha, S., Tunable

electronic interactions between anions and perylenediimide. Org Biomol Chem 2013, 11 (29), 4797-4803. 36.

Georgiev, N. I.; Sakr, A. R.; Bojinov, V. B., Design and synthesis of novel

fluorescence sensing perylene diimides based on photoinduced electron transfer. Dyes

Pigments 2011, 91 (3), 332-339. 37.

Aigner, D.; Borisov, S. M.; Klimant, I., New fluorescent perylene bisimide

indicators-a platform for broadband pH optodes. Anal Bioanal Chem 2011, 400 (8), 2475-2485. 38.

Liu, Y.; Wang, K. R.; Guo, D. S.; Jiang, B. P., Supramolecular Assembly of

Perylene Bisimide with beta-Cyclodextrin Grafts as a Solid-State Fluorescence Sensor for Vapor Detection. Adv Funct Mater 2009, 19 (14), 2230-2235. 39.

Sriramulu, D.; Turaga, S. P.; Bettiol, A. A.; Valiyaveettil, S., Molecular

Organization Induced Anisotropic Properties of Perylene - Silica Hybrid Nanoparticles.

Sci Rep-Uk 2017, 7. 40.

Huang, Y. W.; Liu, X. T.; Wang, Q.; Fu, J. Y.; Zhao, L. L.; Liu, Z. H.; Jin, D. Z.,

Highly responsive ethylenediamine vapor sensor based on a perylenediimidecamphorsulfonic acid complex via ionic self-assembly. J Mater Chem C 2017, 5 (30), 7644-7651. 41.

Hu, Y. Y.; Ma, X. J.; Zhang, Y. B.; Che, Y. K.; Zhao, J. C., Detection of Amines

with Fluorescent Nanotubes: Applications in the Assessment of Meat Spoilage. Acs

Sensors 2016, 1 (1), 22-25. 42.

Annike, W.; R., S. M. C.; Frank, W., Optical Sensing of Aromatic Amino Acids

and Dipeptides by a Crown‐Ether‐Functionalized Perylene Bisimide Fluorophore. Chem.

Eur. J. 2018, 24 (31), 8009-8016.

ACS Paragon Plus Environment

44

Page 45 of 49 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

43.

Che, Y. K.; Yang, X. M.; Loser, S.; Zang, L., Expedient vapor probing of organic

amines using fluorescent nanofibers fabricated from an n-type organic semiconductor.

Nano Lett 2008, 8 (8), 2219-2223. 44.

Fan, J. Y.; Chang, X. M.; He, M. X.; Shang, C. D.; Wang, G.; Yin, S. W.; Peng, H.

N.; Fang, Y., Functionality-Oriented Derivatization of Naphthalene Diimide: A Molecular Gel Strategy-Based Fluorescent Film for Aniline Vapor Detection. Acs Appl Mater Inter 2016, 8 (28), 18584-18592. 45.

Weissenstein, A.; Saha-Moller, C. R.; Wurthner, F., Optical Sensing of Aromatic

Amino Acids and Dipeptides by a Crown-Ether-Functionalized Perylene Bisimide Fluorophore. Chem-Eur J 2018, 24 (31), 8009-8016. 46.

Fang, H. P.; Shellaiah, M.; Singh, A.; Raju, M. V. R.; Wu, Y. H.; Lin, H. C., Naked

eye and fluorescent detections of Hg2+ ions and Cysteine via J-aggregation and deaggregation of a perylene bisimide derivative. Sensor Actuat B-Chem 2014, 194, 229-237. 47.

Datar, A.; Balakrishnan, K.; Zang, L., One-dimensional self-assembly of a water

soluble perylene diimide molecule by pH triggered hydrogelation. Chem Commun 2013,

49 (61), 6894-6896. 48.

Kalita, A.; Hussain, S.; Malik, A. H.; Subbarao, N. V. V.; Iyer, P. K., Vapor phase

sensing of ammonia at the sub-ppm level using a perylene diimide thin film device. J

Mater Chem C 2015, 3 (41), 10767-10774. 49.

Shang, C. D.; Wang, L.; An, Y. Q.; Chen, P.; Chang, X. M.; Qi, Y. Y.; Kang, R.;

Fang, Y., Langmuir-Blodgett films of perylene bisimide derivatives and fluorescent recognition of diamines. Phys Chem Chem Phys 2017, 19 (35), 23898-23904. 50.

Elosua, C.; Lopez-Torres, D.; Hernaez, M.; Matias, I. R.; Arregui, F. J.,

Comparative study of layer-by-layer deposition techniques for poly(sodium phosphate) and poly(allylamine hydrochloride). Nanoscale Res Lett 2013, 8, 539-548. 51.

S. Bettini, L. V., A. Santino, C. Martinelli, G.M. Farinola, A. Cardone, V. Sgobba,

G. Giancane, Spectroscopic investigations, characterization and chemical sensor application

of

composite

Langmuir–Schäfer

films

of

anthocyanins

and

oligophenylenevinylene derivatives. Dyes and Pigments 2012, 94 (1), 156-162.

ACS Paragon Plus Environment

45

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

52.

Page 46 of 49

Tenori, E.; Colusso, A.; Syrgiannis, Z.; Bonasera, A.; Osella, S.; Ostric, A.;

Lazzaroni, R.; Meneghetti, M.; Prato, M., Perylene Derivatives As Useful SERRS Reporters, Including Multiplexing Analysis. ACS Appl. Mater. Interfaces 2015, 7 (51), 28042-28048. 53.

Rigodanza, F.; Tenori, E.; Bonasera, A.; Syrgiannis, Z.; Prato, M., Fast and

Efficient Microwave-Assisted Synthesis of Perylenebisimides. Eur. J. Org. Chem. 2015,

2015 (23), 5060-5063. 54.

Bettini, S.; Maglie, E.; Pagano, R.; Borovkov, V.; Inoue, Y.; Valli, L.; Giancane,

G., Conformational switching of ethano-bridged Cu,H-2-bis-porphyrin induced by aromatic amines. Beilstein J Nanotech 2015, 6, 2154-2160. 55.

Bettini, S.; Pagano, R.; Borovkov, V.; Giancane, G.; Valli, L., The role of the

central metal ion of ethane-bridged bis-porphyrins in histidine sensing. J Colloid Interf

Sci 2019, 533, 762-770. 56.

Manera, M. G.; Rella, R., Improved gas sensing performances in SPR sensors

by transducers activation. Sensor Actuat B-Chem 2013, 179, 175-186. 57.

Manera, M. G.; Ferreiro-Vila, E.; Cebollada, A.; Garcia-Martin, J. M.; Garcia-

Martin, A.; Giancane, G.; Valli, L.; Rella, R., Ethane-Bridged Zn Porphyrins Dimers in Langmuir-Schafer Thin Films: Spectroscopic, Morphologic, and Magneto-Optical Surface Plasmon Resonance Characterization. J Phys Chem C 2012, 116 (19), 1073410742. 58.

Burian, M.; Rigodanza, F.; Amenitsch, H.; Almasy, L.; Khalakhan, I.; Syrgiannis,

Z.; Prato, M., Structural and optical properties of a perylene bisimide in aqueous media.

Chem Phys Lett 2017, 683, 454-458. 59.

Wang, W.; Han, J. J.; Wang, L.-Q.; Li, L.-S.; Shaw, W. J.; Li, A. D. Q., Dynamic

π−π Stacked Molecular Assemblies Emit from Green to Red Colors. Nano Letters 2003,

3 (4), 455-458. 60.

Ford, W. E., Photochemistry of 3,4,9,10-perylenetetracarboxylic dianhydride

dyes: visible absorption and fluorescence of the di(glycyl)imide derivative monomer and dimer in basic aqueous solutions. Journal of Photochemistry 1987, 37 (1), 189-204.

ACS Paragon Plus Environment

46

Page 47 of 49 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

61.

Chen, S.; Slattum, P.; Wang, C.; Zang, L., Self-Assembly of Perylene Imide

Molecules into 1D Nanostructures: Methods, Morphologies, and Applications. Chem.

Rev. 2015, 115 (21), 11967-11998. 62.

Korevaar, P. A.; Schaefer, C.; de Greef, T. F. A.; Meijer, E. W., Controlling

Chemical Self-Assembly by Solvent-Dependent Dynamics. J. Am. Chem. Soc. 2012,

134 (32), 13482-13491. 63.

Yagai, S.; Seki, T.; Karatsu, T.; Kitamura, A.; Würthner, F., Transformation from

H- to J-Aggregated Perylene Bisimide Dyes by Complexation with Cyanurates. Angew.

Chem. Int. Ed. 2008, 47 (18), 3367-3371. 64.

Zhao, H.-M.; Pfister, J.; Settels, V.; Renz, M.; Kaupp, M.; Dehm, V. C.; Würthner,

F.; Fink, R. F.; Engels, B., Understanding Ground- and Excited-State Properties of Perylene Tetracarboxylic Acid Bisimide Crystals by Means of Quantum Chemical Computations. J. Am. Chem. Soc. 2009, 131 (43), 15660-15668. 65.

Engels, B.; Engel, V., The dimer-approach to characterize opto-electronic

properties of and exciton trapping and diffusion in organic semiconductor aggregates and crystals. Physical Chemistry Chemical Physics 2017, 19 (20), 12604-12619. 66.

Draper, E. R.; Greeves, B. J.; Barrow, M.; Schweins, R.; Zwijnenburg, M. A.;

Adams, D. J., pH-Directed Aggregation to Control Photoconductivity in Self-Assembled Perylene Bisimides. Chem 2017, 2 (5), 716-731. 67.

Grimme, S.; Brandenburg, J. G.; Bannwarth, C.; Hansen, A., Consistent

structures and interactions by density functional theory with small atomic orbital basis sets. J Chem Phys 2015, 143 (5), 054107. 68.

Chai, J. D.; Head-Gordon, M., Systematic optimization of long-range corrected

hybrid density functionals. J Chem Phys 2008, 128 (8), 084106. 69.

M. Kasha, H. R. R., M. Ashraf El-Bayoumi, The exciton model in molecular

spectroscopy. Pure and Applied Chemistry 1965, 11 (3-4), 371-392. 70.

Chen, Z. S., V.; Dehm, V.; Prins, P.; Siebbeles, L. D.; Seibt, J.; Marquetand, P.;

Engel, V.; Wurthner, F. , Photoluminescence and conductivity of self-assembled pi-pi stacks of perylene bisimide dyes. . Chemistry 2007, 13 (2), 436-449. 71.

Silverstein, R. B. C. B., G.; Morril, T.C. , Spectrometric Identification of Organic

Compounds. 1981, Fourth edition, John Wiley & Sons Edition, 95-180.

ACS Paragon Plus Environment

47

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

72.

Page 48 of 49

Würthner, F.; Saha-Möller, C. R.; Fimmel, B.; Ogi, S.; Leowanawat, P.; Schmidt,

D., Perylene Bisimide Dye Assemblies as Archetype Functional Supramolecular Materials. Chem. Rev. 2016, 116 (3), 962-1052. 73.

Giancane, G.; Guascito, M. R.; Malitesta, C.; Mazzotta, E.; Picca, R. A.; Valli, L.,

QCM sensors for aqueous phenols based on active layers constituted by tetrapyrrolic macrocycle Langmuir films. J Porphyr Phthalocya 2009, 13 (11), 1129-1139. 74.

Giancane, G.; Basova, T.; Hassan, A.; Gumus, G.; Gurek, A. G.; Ahsen, V.; Valli,

L., Investigations and application in piezoelectric phenol sensor of Langmuir-Schafer films of a copper phthalocyanine derivative functionalized with bulky substituents. J

Colloid Interf Sci 2012, 377, 176-183. 75.

Alonso, T.; Irigoyen, J.; Iturri, J. J.; Larena, I. L.; Moya, S. E., Study of the

multilayer assembly and complex formation of poly(diallyldimethylammonium chloride) (PDADMAC) and poly(acrylic acid) (PAA) as a function of pH. Soft Matter 2013, 9 (6), 1920-1928. 76.

Miyazaki, C. M.; Shimizu, F. M.; Mejia-Salazar, J. R.; Oliveira, O. N.; Ferreira, M.,

Surface plasmon resonance biosensor for enzymatic detection of small analytes.

Nanotechnology 2017, 28 (14) 145501-145506. 77.

Gestwicki, J. E.; Hsieh, H. V.; Pitner, J. B., Using receptor conformational change

to detect low molecular weight analytes by surface plasmon resonance. Anal Chem 2001, 73 (23), 5732-5737. 78.

Matsui, J.; Akamatsu, K.; Hara, N.; Miyoshi, D.; Nawafune, H.; Tamaki, K.;

Sugimoto, N., SPR sensor chip for detection of small molecules using molecularly imprinted polymer with embedded gold nanoparticles. Anal Chem 2005, 77 (13), 42824285. 79.

Tang, Y. J.; Zeng, X. Q.; Liang, J., Surface Plasmon Resonance: An Introduction

to a Surface Spectroscopy Technique. J Chem Educ 2010, 87 (7), 742-746. 80.

Chen, D. Q.; Mei, Y. H.; Hu, W. H.; Li, C. M., Electrochemically enhanced

antibody immobilization on polydopamine thin film for sensitive surface plasmon resonance immunoassay. Talanta 2018, 182, 470-475. 81.

Baba, A.; Taranekar, P.; Ponnapati, R. R.; Knoll, W.; Advincula, R. C.,

Electrochemical Surface Plasmon Resonance and Waveguide-Enhanced Glucose

ACS Paragon Plus Environment

48

Page 49 of 49 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

Biosensing with N-Alkylaminated Polypyrrole/Glucose Oxidase Multilayers. Acs Appl

Mater Inter 2010, 2 (8), 2347-2354. 82.

Vergara, D.; Bianco, M.; Pagano, R.; Priore, P.; Lunetti, P.; Guerra, F.; Bettini, S.;

Carallo, S.; Zizzari, A.; Pitotti, E.; Giotta, L.; Capobianco, L.; Bucci, C.; Valli, L.; Maffia, M.; Arima, V.; Gaballo, A., An SPR based immunoassay for the sensitive detection of the soluble epithelial marker E-cadherin. Nanomed-Nanotechnol 2018, 14 (7), 19631971.

ACS Paragon Plus Environment

49