Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect
ACS Publications. Most Trusted. Most Cited. Most Read
Investigation of Earth-Abundant Metal Salts for the Inhibition of Asphalt-Derived Volatile Organic Compounds
My Activity
  • Open Access
Article

Investigation of Earth-Abundant Metal Salts for the Inhibition of Asphalt-Derived Volatile Organic Compounds
Click to copy article linkArticle link copied!

  • Harpreet Kaur
    Harpreet Kaur
    Department of Chemistry, Emory University, Atlanta, Georgia 30322, United States
  • Reem Nsouli
    Reem Nsouli
    Department of Chemistry, Emory University, Atlanta, Georgia 30322, United States
    More by Reem Nsouli
  • Gabriella Cerna
    Gabriella Cerna
    School of Molecular Sciences, Arizona State University, 660 S. College Avenue, Tempe, Arizona 85287-3005, United States
  • Saba Shariati
    Saba Shariati
    School of Sustainable Engineering and the Built Environment, Arizona State University, 660 S. College Avenue, Tempe, Arizona 85287-3005, United States
  • Marco Flores
    Marco Flores
    School of Molecular Sciences, Arizona State University, 660 S. College Avenue, Tempe, Arizona 85287-3005, United States
    More by Marco Flores
  • Elham H. Fini*
    Elham H. Fini
    School of Sustainable Engineering and the Built Environment, Arizona State University, 660 S. College Avenue, Tempe, Arizona 85287-3005, United States
    *[email protected] (E.H.F.).
  • Laura K. G. Ackerman-Biegasiewicz*
    Laura K. G. Ackerman-Biegasiewicz
    Department of Chemistry, Emory University, Atlanta, Georgia 30322, United States
    *[email protected] (L.K.G.A.-B.).
Open PDFSupporting Information (1)

ACS Omega

Cite this: ACS Omega 2024, 9, 21, 22941–22951
Click to copy citationCitation copied!
https://doi.org/10.1021/acsomega.4c02095
Published May 13, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

Asphalt is used globally in construction for roads, pavements, and buildings; however, as a fossil-derived material, it is known to generate volatile organic compounds (VOCs) upon exposure to heat and light that can be harmful to human health. Several heterogeneous strategies have been reported for the inhibition of these VOCs; however, the direct use of inexpensive, accessible Earth-Abundant metals has not been extensively explored. In this study, simple metal salts are examined for their coordination capability toward asphalt-derived VOCs. From UV–visible (UV–vis) spectroscopic studies, FeCl3 emerged relative to other metal salts (metal = Mn, Co, Ni, Cu, Zn) as a promising candidate for the adsorption and retention of Lewis basic compounds. Coordination of an example oxygen-containing VOC, benzofuran (Bf), to Fe yielded a paramagnetic semi-octahedral complex Fe(Bf)3Cl3. Evaluation by thermal gravimetric analysis (TGA) coupled to infrared spectroscopy (IR) demonstrated that the complex was stable up to 360 °C. Spectroscopic evaluation demonstrated the stability of the complex upon visible light irradiation and in the presence of a variety of organic pollutants. The potential application of Fe was demonstrated by subjecting biochar to FeCl3 followed by the addition of Bf. It was discovered that this Fe-rich biochar was successful at adsorbing Bf suggesting the possibility of introducing Fe to biochar late-stage in processing to deter asphalt degradation and VOC emissions. An understanding of the binding and stability of Fe salts to VOCs provides insight into how a sustainable infrastructure can be achieved.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2024 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

Volatile organic compounds (VOCs) are organic pollutants that cause significant harm to human health including skin irritation, neurological impairments, respiratory problems, and cardiovascular and reproductive diseases. (1−5) Due to growing industrialization, increasing amounts of VOCs are being released into the environment from construction sites, commercial products, and vehicle emissions. (6) Moreover, these compounds contribute to air pollution and are precursors to secondary organic aerosols. Recently, asphalt has been identified as a nontraditional source of VOCs. Small organic molecules such as alkanes, alkenes, aldehydes, aromatics, and heterocyclic compounds have been reported as major constituents of asphalt emissions during the construction and service life of pavement (Figure 1A). (4) In the United States, more than 2.5 million miles of pavement contribute to VOC emissions. (6) The loss of VOCs from asphalt increases by 300% upon irradiation and by 70% for every 20 °C rise in temperature. (7−9) This can also lead to a reduction in the durability of asphalt, further promoting the rate of deterioration of asphalt-surfaced areas. (10) Studies conducted to evaluate emissions found that asphalt-related VOCs are harmful not only for construction workers but also the public who is frequently exposed to asphalt surfaces. (11−13) Recently, there has been a significant emphasis on identifying simple and cost-effective solutions for limiting these emissions. (14−16)

Figure 1

Figure 1. Significance of inhibiting asphalt-derived VOC emissions. (A) Prevalence of VOC emissions upon exposure of asphalt to heat and light. (B) Limitations of the incorporation of inorganic materials in asphalt as VOC inhibitors. (C) Studying the use of Earth-Abundant metal salts in asphalt binder to deter VOC emission.

The two main strategies proposed to reduce emissions from asphalt-surfaced areas include lowering the construction temperature of hot-mix asphalt and incorporating inhibitors into asphalt that are aimed at trapping and retaining VOCs. (17,18) While the first strategy is effective during the construction phase, it does not reduce emissions that occur during the service life of asphalt. Therefore, alternative strategies that focus on preventing the degradation of asphalt over time have been explored. (16) The two categories of VOC inhibitors are cross-linking agents and adsorbents. However, these inhibitors are often limited in suppressing VOCs when they are practically applied to asphalt. In some cases, the inhibitors even reduce the longevity of asphalt and react with the VOCs, leading to the emission of new uncharacterized byproducts. (19−21) Alternatively, asphalt can be modified using inorganic materials such as metal oxides, salts, and porous substances to reduce VOC emissions without the formation of unintended byproducts. The modification of asphalt is typically achieved heterogeneously with double hydroxides, porous geopolymers, zeolites and other materials that must be carefully formulated prior to use. (22−25) In contrast, there has been little to no investigation into the direct use of Earth-Abundant metal salts for the adsorption of VOCs that can be added late-stage to the asphalt binder (Figure 1B). Therefore, there is a need to examine scalable and economically viable inorganic salts capable of integrating into asphalt as a modifier that could serve as emission reducing reagents.
Recent work by Fini et al. focused on the selective adsorption of VOCs from asphalt using Fe-rich biochar made from a hybrid feedstock of Cyanidioschyzon merolae algae and swine manure (20:80 ratio) with an Fe metal content of 21.8 g/kg. The results indicated that when incorporating Fe-rich biochar into asphalt, up to 76% reduced emissions were observed. (26) This research finding motivated us to explore the binding capability of Earth-Abundant metal salts like Fe toward VOCs. We began the investigation of Bf as recent studies on the emission of asphalt reported 30% of emissions can consist of oxygen-containing compounds. (8) Additionally, Bf has been reported by the United States Agency for Toxic Substances and Disease Registry to be a carcinogenic compound whose oral exposure in the range of 60–240 mg kg–1 per day was shown by the Centers for Disease Control and Prevention to cause cancer in mice. In this study, we have identified that the incorporation of a simple FeCl3 salt into biochar can effectively adsorb Bf (Figure 1C). The interaction between FeCl3 and biochar along with the resulting material’s resilience was examined using TGA-DSC analysis, while the coordination of Bf was explored via UV–vis spectroscopy.

Results and Discussion

Click to copy section linkSection link copied!

I. Investigation of the Binding Activity of Earth-Abundant Metal Salts

To probe the fundamental binding properties of oxygen-containing VOCs, the binding of VOCs to Earth-Abundant metal salts was first investigated. Bf was chosen as a model carcinogenic VOC; it has an aromatic core, Lewis basic oxygen, and a relatively low boiling point. (27) In addition, Bf has been identified as being emitted from asphalt over a range of temperatures, along with other oxygen- and sulfur-containing VOCs. (9) To identify metal salts that could effectively bind Bf, and eventually be incorporated into biochar, initial studies were conducted using UV–vis spectroscopy with a noncoordinating solvent system. For each trial, equimolar amounts of the metal salt and Bf were dissolved in 1 mL of dichloromethane (DCM). The solutions were stirred at 28 °C for 1 h prior to UV–vis analysis (Figure 2A). Spectra were examined for bathochromic shifts observed relative to the absorbance of the independent metal salt, suggesting coordination of Bf to the metal. Interestingly, among the 14 Earth-Abundant metal salts studied, only the spectrum collected using FeCl3 indicated an interaction with Bf. The appearance of a new red-shifted peak at 554 nm was observed in the spectrum of FeCl3 and Bf stirred at room temperature, which could be attributed to an n-π* ligand-to-metal transition (Figure 2B). It was noted that when FeCl3 and Bf were mixed in DCM, a deep purple solution was observed that upon evaporation resulted in a purple-colored complex. This is consistent with known Fe complex formation in which FeCl3 serves as a strong Lewis Acid. (28)

Figure 2

Figure 2. (A) Change in λmax between metal salts with and without Bf. Metal salt (0.05 mmol) and Bf (0.05 mmol) were dissolved in 1 mL of DCM and stirred at 28 °C for 1 h prior to UV–vis analysis. (B) UV–vis spectrum of 200 μM Bf (green), 200 μM FeCl3 (red), and a 1:1 mixture of FeCl3 and Bf (purple). aBathochromic shift was observed relative to the absorbance of the metal salt.

As part of a preliminary assessment of the potential for FeCl3 to be an effective metal for direct incorporation into biochar, the effects of temperature and irradiation on Fe-Bf binding were next examined. In prior reports related to asphalt, VOC emissions were observed to be temperature-dependent: as the temperature rises to 60 °C, the amount of emissions doubles, reaching the highest level when the temperature is over 140 °C. (8) The laying of asphalt is conventionally achieved at 100 °C; therefore it is crucial to investigate the stability of the complex at high temperature. (29,30) Additionally, upon exposure to light, the emission rate of VOCs accelerates. (8) Therefore, it is important to investigate the stability of Fe-Bf interactions under these conditions. After the purple complex was formed in DCM (50 mM) by refluxing the metal salt with Bf at 60 °C (Table S1), it was then exposed to elevated temperatures (up to 100 °C). Similarly, the Fe-Bf complex solution was irradiated at different wavelengths (390, 456, and 525 nm), and the stability of the complex was studied in solution phase. The UV–vis absorption spectra were noted for each of the samples and compared with the untreated Fe-Bf complex (Figure 3).

Figure 3

Figure 3. UV–vis spectrum of (A) 200 μM FeCl3 (pink), 200 μM Bf (green), and a 1:1 mixture of FeCl3 and Bf, stirred for 1 h in DCM in one of four conditions: (i) at 28 °C (purple); (ii) heated at 60 °C (mauve), (iii) 80 °C (dark blue), and (iv) 100 °C (light blue); and (B) 200 μM FeCl3 (pink), 200 μM Bf (green), and a 1:1 mixture of FeCl3 and Bf, stirred for 1 h in DCM under irradiation conditions at: (i) 390 nm (light blue), (ii) 456 nm (dark blue), and (iii) 525 nm (mauve) wavelengths.

The Fe-Bf complex was stable up to 80 °C, and no degradation was observed upon irradiation at 456 and 525 nm wavelength (Figure 3 and Figure S2). The red-shifted peak was observed at 28 °C, 60 °C, and 80 °C; however, upon further increasing the temperature to 100 °C, the complex began to degrade, as indicated by the diminished intensity of the peak at 554 nm. It is notable that upon heating from 28 to 60 °C, the intensity of the absorbance at 554 nm increased significantly, indicating that mild heating could provide more favorable conditions for the formation of the Fe-Bf complex. The complex remained stable upon irradiation at 456 and 525 nm (visible light) as the absorption peak at 554 nm was unchanged after introduction to light. However, upon irradiation of the mixture at 390 nm (near UV), no peak at 554 nm was observed, suggesting degradation. As the majority of sunlight that is not absorbed by the atmosphere is in the visible range, it is significant that the complex still remained intact when exposed to blue and green light.

TGA-FTIR Studies to Support the Binding of Bf to FeCl3

To further support the formation of the Fe-Bf complex, we used TGA-FTIR analysis to monitor the extrusion of Bf from the Fe-Bf complex through evolved gas IR analysis. The FTIR stretch at 3065 cm–1 that corresponds to the Csp2–H bond on Bf was examined over time. Upon coordination of Bf, the complex would be expected to have a delayed decomposition time or prolonged degradation at higher temperatures, in comparison to Bf independently. Figure 4A and 4B show the decomposition times and profile curves of the Bf alone and Fe and Bf mixture. Upon analysis by TGA-FTIR pure Bf is detected by FTIR approximately after 13 min at 82 °C. However, when the Fe and Bf mixture was subjected to similar analysis no changes in the IR stretching band at 3065 cm–1 were observed until 30 min at 363 °C. The observed delay in the appearance of Bf in the FTIR spectra supports the formation of the Fe-Bf complex. Notably, when this complex was subjected to prolonged heating (at 200 °C for 40 min), which would be expected during the asphalt’s production phase, it remained intact (Figure 4A, entry 3).

Figure 4

Figure 4. (A) FeCl3 (16.2 mg, 0.1 mmol) and Bf (10.8 μL, 0.1 mmol) were mixed with dichloromethane (3.0 mL) at room temperature for 1 h. The solution was then concentrated, and a portion of the sample (16 mg) was transferred to a high-temperature Pt pan for TGA-FTIR analysis. Entry 1 displays a control run for the detection of pure Bf by FTIR at 82 °C upon thermal degradation of 10.8 μL Bf dissolved in 200 μL of DCM analyzed using TGA-FTIR. Entry 2 displays FTIR detection of Bf at 363 °C upon thermal degradation of the Fe-Bf complex using TGA at a ramp of 20 °C/min to 1000 °C. Entry 3 displays FTIR detection of Bf at 363 °C upon thermal degradation of Fe-Bf complex using TGA at a ramp of 20 °C/min to 40 °C followed by an isothermal hold at 40 °C for 10 min to ensure solvent evaporation. This was followed by another ramp at 20 °C/min to 200 °C followed by another isothermal hold for 40 min to assess the thermal stability of the Fe-Bf complex at environmental conditions. Finally, a ramp at 20 °C/min up to 1000 °C was applied to investigate overall thermal stability. (B) Gram-Schmidt profile for Bf (green) and a mixture of FeCl3 and Bf samples (purple). The method used included a ramp rate of 20 °C/min to 40 °C, followed by an isotherm hold at 40 °C for 10 min, and ramp rate of 20 °C/min up to 1000 °C.

II. Characterization of the Fe-Bf Complex

Stoichiometry Study of Fe to Bf

From UV–vis analysis, it was evident that FeCl3 was uniquely capable of binding Bf and that the Fe-Bf complex was stable under elevated temperatures and irradiation. To gain insight into the molecular structure of the Fe-Bf complex, the stoichiometry of the complex was investigated using the Job plot and Molar ratio methods. (31) In the Job plot method, the Fe-Bf complex was prepared by using different mole fractions of FeCl3 and Bf while keeping the total molar concentration of the solution constant (Table S3). UV–vis spectra were collected for the samples, and a plot of the mole fraction of Fe (III) versus the absorbance intensity at 554 nm was obtained. The stoichiometry of the complex was established from the maxima of the Job plot curve (Figure 5A). The nonlinear fitting curve displayed a maximum value at 0.25, suggesting a 1:3 ratio of Fe to Bf for the Fe-Bf complex. (31) Subsequently, a Molar ratio experiment was conducted. In this study, the number of moles of Bf was varied, keeping the moles of Fe(III) constant in the solution, and the changes in the absorption intensity at 554 nm were noted. The variation in the absorbance intensity of the absorption peak was plotted with respect to the molar ratio of the complex, and the stoichiometry of the complex formed was deduced from the highest absorbance in the plot (Figure 5B).

Figure 5

Figure 5. (A) Job plot of the Fe-Bf complex prepared using different mole fractions of Fe(III) and Bf where the total molar concentration of the complex is kept constant; and (B) Molar Ratio plot of the Fe-Bf complex prepared keeping the number of moles of Fe(III) (nM) constant and varying the number of moles of Bf (nL). In both plots, the variations in the absorption intensity of the peak at 554 nm were observed, and the stoichiometry of the Fe-Bf complex was deduced from the maximal position in the Job plot and the highest absorbance in the Molar Ratio plot, suggesting a 1:3 stoichiometry for the Fe-Bf complex.

Furthermore, the strength of the interaction between FeCl3 and Bf was calculated using the Benesi–Hildebrand relation (1) and was found to be 0.632 × 105 M–1, indicating excellent binding affinity of the metal salt toward Bf, as shown in Figure S6. (32)
log(AAo)(AfAo)=log[Fe]+logKb
(1)
A0, A, and Af are the absorption values in the absence of ferric ion, at the intermediate level of ferric ion concentration, and at the saturation of the ferric ion, respectively. [Fe] is the concentration of the Fe salt in M, and “Kb” is the binding constant in M–1. The stoichiometry studies suggest that the Fe-Bf complex consists of three Bf molecules coordinated to one Fe atom, which may or may not have displaced chlorine atoms. Therefore, subsequent experiments focused on the analysis of the chemical structure by gaining information about the magnetic properties of the Fe-Bf complex.

Electron Paramagnetic Resonance (EPR) Spectroscopy

To study the magnetic behavior of the Fe-Bf complex, the EPR spectrum was collected at 120 K (Figure 6). The X-band (9.40 GHz) EPR spectrum of the Fe-Bf complex (DCM, T = 120 K) showed signals consistent with the presence of a single high-spin Fe(III) (black line in Figure 6). (33,34) To obtain the EPR parameters, the respective spin Hamiltonian was fitted to the data (red line in Figure 6). The observed spectral features were well fit (see Supporting Information) considering an S = 5/2 sextet state with an isotropic g-value (giso = 1.872, see inset (Figure 6) for all fitting parameters) and axial zero-field interaction (|D| = 5282 MHz and E = 0). Furthermore, the value of |D| approaches the conditions for resonance at a magnetic field value equal to zero, i.e., |D| ≈ ν/2 where ν is the microwave frequency. All these properties indicate an octahedral Fe coordination environment. (34)

Figure 6

Figure 6. Experimental (black line) and simulated (red line) X-band EPR spectra of Fe-Bf at 120 K. The narrow line around 340 mT (marked with an asterisk) belongs to a minor unidentified species. Inset: parameters used to fit the EPR Spectrum of the Fe-Bf complex at 9.40 GHz and T = 120 K. aThe fitting parameters were the isotropic g-value (giso), the zero-field splitting parameters (D and E), and the isotropic line width (ΔB).

Moreover, the best fit of the EPR spectrum was obtained by considering a partially ordered frozen-solution sample. It was considered that most of the complexes are arranged with the planes formed by their equatorial ligands being parallel. Such ordering is probably a consequence of pi stacking among complexes and produces additional broadening of the EPR spectrum (Figure 6). The narrow signal around 340 mT (marked with an asterisk) was not included in the fit, since it belongs to a minor nonidentified species. Coincidentally, its line width (1.2 mT) is like those observed for organic radicals.

Magnetic Moment Measurements

To further support the characterization of the Fe-Bf complex the magnetic susceptibility (χm) of the paramagnetic Fe-Bf complex was evaluated using Guoy’s method. (35) In this experiment, a known amount of the complex was analyzed with [Ni(en)3][S2O3] as a calibrant. The balance reading (R) was used to calculate the gram susceptibility (χg) using the relationship (2) where ‘C’ is the calibration constant calculated using a calibrant whose χg is known. In this formula ‘L’ is the length of the sample in the sample tube in cm, ‘R’ is the balance reading for the sample, “Ro” is the balance reading in the absence of the sample, and ‘m’ is the sample mass in g.
χg=CLRRom×109
(2)
The value of χg for [Ni(en)3][S2O3] is 1.103 × 10–5 emu/g. After calculating the value of ‘C’, the “χg” for the Fe-Bf complex was calculated to be 4.7931 × 10–5 emu/g. Subsequently, the molar susceptibility (χm) was calculated using eq 3 where ‘M’ is the molecular weight of the complex in g/mol.
χm=χg×M
(3)
Since the molecular weight of the complex was unknown, the value of “χm” was obtained from the molecular weight of the possible Fe-Bf structures that could be formed with a 1:3 ratio of FeCl3 to Bf.
χm is in turn related to the effective magnetic moment (μeff) by the relationship (4) and μeff can be used to calculate the number of unpaired electrons (n) present in a paramagnetic complex using (5).
μeff=2.828(χm×T)
(4)
μeff=n(n+2)
(5)
It was deduced from the calculations that the possibility of the Fe atom to exist in low-spin conditions is negligible, which is also supported by the experimental findings using EPR analysis. The value of μeff was calculated for the following possible geometries of the Fe-Bf complex (Figure 7). The molecular weight of the structure is 516.60 g/mol, and the value of χm was calculated to be 0.01451801 emu/mol. From the calculated χm, the μeff was obtained to be 5.88 μB which is close to the experimental value of 5.9 μB. Relationship (5) was then employed to calculate the number of unpaired electrons which was found to be 4.97.

Figure 7

Figure 7. Proposed chemical structures of the high-spin Fe-Bf complex. (A) [Fe(Bf)3Cl2]Cl and (B) Fe(Bf)3Cl3 based on EPR and magnetic susceptibility results.

The spectroscopic analysis indicated a stoichiometry of 1:3 Fe/Bf in the complex. EPR analysis as well as magnetic susceptibility calculations suggested the presence of a high-spin Fe(III) center in the complex structure and supported the plausibility of the complex to exist in either an octahedral or trigonal bipyramidal geometry. In addition, the EPR analysis indicated that it is highly unlikely that the complex would have two metal centers. Therefore, based on the spectroscopic, magnetic susceptibility, and EPR results, the following two geometries of the Fe-Bf complex were proposed in Figure 7.

III. DFT Modeling

To assess the intermolecular interactions between FeCl3 and Bf, four methods of interaction were probed, depicted in eqs 69. The geometries of the species involved in the reactions were optimized by using an implicit continuum solvation model, COSMO. Stabilization energy (ES) for these interactions were calculated using eq 1 (Section I in the Supporting Information). Equation 6 demonstrates the addition of three Bf molecules to FeCl3 while the three chlorine atoms remain attached to the metal center. Equations 79 illustrate the formation of Fe-Bf ions with three, two, and one Bf molecule(s), respectively. Results indicate that separation of the Cl atom from the iron center and formation of the complex ions in a nonaqueous medium (DCM solvent with a dielectric constant of 9.08) was not thermodynamically favorable. These findings were supported by positive ES values of 212.8, 130.1, and 56.6 kcal/mol for eq 7, eq 8, and eq 9, respectively. Moreover, a significant and negative value of Es was detected for the formation of the neutral Fe(Bf)3Cl3 complex (eq 6). Two distinct stable geometries were detected, as shown in Figure 8. In one complex, the iron center interacted with the oxygen atoms of the Bf molecules (Figure 8A). The other complex showed the π-electron cloud of the benzene rings of the Bf molecules interacting with the iron center (Figure 8B).
FeCl3+3BfDCMFe(Bf)3Cl3
(6)
FeCl3+3BfDCM[Fe(Bf)3]3++3Cl
(7)
FeCl3+2BfDCM[Fe(Bf)2Cl]2++2Cl
(8)
[FeCl3+BfDCMFe(Bf)Cl2]++Cl
(9)

Figure 8

Figure 8. Optimized geometries of the Fe(Bf)3Cl3 complexes displaying bond length (Å) of Fe–O bond and Fe–C bond and stabilization energy (ES) of the two optimized geometries in kcal/mol. Atom colors are Fe-violet, Cl-green, O-red, C-gray, and H-white.

IV. Selectivity Study of the Complex

Literature precedent has demonstrated that asphalt can be a source of a wide range of VOC emissions containing anthracene, fluoranthene, fluorene, 1-methyl naphthalene, benzothiophene, dibenzothiophene and other oxygen-, nitrogen-, and sulfur-containing organic molecules. (36) The complexity of the composition as well as the emissions of asphalt fumes increase with age, usage, temperature, and UV exposure. Therefore, to develop a robust system that can efficiently suppress the emission of these hazardous compounds, the selectivity of the material in the presence of different VOCs and different concentrations is crucial. To address this need, we investigated the stability of the Fe-Bf complex upon being subjected to different VOCs that have been reported in the literature. To gain insight into the possible interactions between the competing VOCs and the Fe-Bf complex, the Fe-Bf complex was examined in the presence of different VOCs by using UV–vis spectroscopy. Control samples of FeCl3 with different VOCs were also independently analyzed to rule out the possibility of new observed absorption bands resulting from electronic transitions of the VOC molecule.
First, the Fe-Bf complex was tested in the presence of different oxygen-containing VOCs. We examined the UV–vis spectra of the complex in the presence of 0.3 mmol of 2,3-dihydrobenzofuran (2,3-DHB), 9-hydroxy fluorene, catechol, triethylene glycol (TEG) and salicylic acid in DCM (1.0 mL). Among the studied VOCs, only 2,3-DHB and TEG displayed an interaction with the complex (Figures 9A, S7, and S10C). In the presence of 2,3-DHB, a new distinct absorption peak at 535 nm was observed and the absorption peak at 554 nm diminished. The interaction between the Fe-Bf complex and 2,3-DHB was investigated by slowly subjecting the preformed complex to small aliquots of 2,3-DHB and monitoring the absorption spectrum after each addition (Figure S12). It was inferred from the absorption spectrum that 2,3-DHB could destabilize the Fe-Bf complex, as the intensity of the peak at 554 nm decreased. Alternatively, when the Fe-Bf complex was subjected to TEG, distinct absorption bands at 316 and 365 nm were observed (Figure 9A). Analogously, a titration experiment conducted in the presence of TEG showed that the complex was gradually destabilized with no Fe-Bf complex detected after the addition of 40 mM TEG (Figure 10B). This aligned with our molecular modeling calculations showing that the TEG molecule could substitute for the Bf ligand in the Fe-Bf complex (Figure 12B).

Figure 9

Figure 9. (A) Schematic representation of the binding activity of the complex with Oxygen-containing VOCs evaluated using UV–vis absorption spectroscopy displaying distinct absorption peaks (λmax) summarized in tabular form. Note: FeCl3 (0.1 mmol), Bf (0.3 mmol), and VOC (0.3 mmol) were dissolved in DCM (2 mL). The solutions were stirred at 60 °C for 1 h prior to UV–vis analysis. Note: ‘-’ indicates that no spectroscopic changes were observed. (B) Bar graph displaying variation in the intensity of the absorption peak at 554 nm (characteristic of the Fe-Bf complex) in the presence of different VOCs. The UV–vis absorption peak was observed after subjecting the Fe-Bf complex (50 mM) to 0.3 mmol of VOCs in DCM and diluting it before recording the spectra.

Figure 10

Figure 10. (A) UV–vis absorption spectra of the Fe-Bf complex in the presence of triethylene glycol, along with FeCl3, FeCl3-triethylene glycol, triethylene glycol, and the Fe-Bf complex as controls under the same conditions. (B) UV–vis absorption spectra of the Fe-Bf complex in the presence of increasing concentrations of triethylene glycol (0–40 mM), displaying distinct absorption peaks (316 and 365 nm) along with the disappearance of the absorption peak at 554 nm.

We further investigated the stability of the Fe-Bf complex in the presence of benzothiophene, a sulfur-containing VOC that is emitted from asphalt. The solution containing the complex and benzothiophene was analyzed by using UV–visible spectroscopy to observe any spectroscopic changes. The absorption spectrum displayed no characteristic peak in the UV–vis region, suggesting no interaction between Fe(III) and benzothiophene (Figure S10A). The stability of the complex in the presence of other VOCs such as 1-methyl naphthalene, quinoline, anthracene, fluoranthene, and toluene was also investigated. The complex was subjected to 0.3 mmol of these VOCs and analyzed spectrophotometrically. Among these analyses, quinoline was found to destabilize the Fe-Bf complex, and a new peak at 365 nm was observed, indicating an interaction between FeCl3 and quinoline (Figure 9B and Figure S8). The selectivity results are summarized in Table S4. Interestingly, no new peaks were observed in the presence of anthracene. This could indicate that anthracene disrupts the complexation of the Fe-Bf species without the formation of any new species. The interaction between the Fe-Bf complex and quinoline was further studied by a titration study. This demonstrated that upon increasing the concentration of quinoline, the Fe-Bf complex was gradually destabilized and was not observed after the addition of 20 mM quinoline (Figure 11B). Our computational modeling using density functional theory supported the higher stability of the Fe-quinoline complex, which is explained in greater detail in the DFT section (Figure 12A).

Figure 11

Figure 11. (A) UV–vis absorption spectra of the Fe-Bf complex in the presence of quinoline, along with FeCl3, FeCl3-quinoline, quinoline, and the Fe-Bf complex as controls under the same conditions. (B) UV–vis absorption spectra of the Fe-Bf complex in the presence of increasing concentrations of quinoline (0–20 mM) displayed distinct absorption peaks along with the disappearance of the absorption peak at 554 nm.

V. DFT Modeling for Evaluation of the Stable Fe-VOC Complexes

2,3-DHB, TEG and quinoline were the three VOCs that could displace the Bf ligand in the Fe-Bf complex. To assess the stability of the Fe(Bf)3Cl3 complex in the presence of these VOCs, chemical reactions 10 and 11 were considered. In cases where different VOC orientations were possible, the structure with the maximal stabilization energy was reported. All of these VOC molecules have oxygen or nitrogen heteroatoms in their structures that can form a complex similar to that shown in Figure 12A. To analyze the stability of the new Fe-VOC complexes, their ES values were calculated for the formation of Fe-VOC complexes from an FeCl3 molecule (eq 10) and the ΔES for the formation of Fe-VOC complexes from the Fe(Bf)3Cl3 complex (eq 11).
FeCl3+3VOCDCMFe(VOC)3Cl3
(10)
Fe(Bf)3Cl3+3VOCDCMFe(VOC)3Cl3+3Bf
(11)

Figure 12

Figure 12. Optimized geometries ES and ΔES for the formation of (A) Fe(quinoline)3Cl3, (B) Fe(TEG)3Cl3, and (C) Fe(2,3-DHB)3Cl3. Atom colors are Fe-violet, Cl-green, O-red, N-blue, C-gray, and H-white.

DFT results indicated that these VOCs formed new complexes with FeCl3 that were more stable than the Fe(Bf)3Cl3 complex and consequently could replace the Bf ligands in Fe(Bf)3Cl3. The geometries of these new complexes and their ES and ΔES are presented in Figure 12. The higher stability of these new complexes was reflected in the ΔES values that were −26.6 kcal/mol, −25.6 kcal/mol, and −15.7 kcal/mol for interactions of quinoline, TEG, and 2,3-DHB, respectively. Alongside the ΔES values, the optimized geometries showed shorter distances between the iron center and heteroatoms of these VOCs compared to those of Bf, indicating stronger coordination between iron and the heteroatoms of these ligands.

VI. Direct Application of Fe-Incorporated Biochar to Bf Inhibition

The interaction between Bf and FeCl3 to form a stable Fe(Bf)3Cl3 complex motivated us to apply the metal salt as a late-stage dopant in an asphalt modifier to reduce Bf emissions. Since Fe-rich biochar has precedent in reducing VOC emissions up to 76%, we focused on utilizing algae biochar. Algae biochar was treated with FeCl3 (Section III, Supporting Information) and analyzed spectroscopically. It was observed from the UV–vis spectra that a new shoulder peak appeared at 405 nm with the treated biochar, which was not a characteristic peak of pristine biochar or FeCl3 independently (Figure S13). The treated biochar was also analyzed using FTIR spectroscopy to confirm the incorporation of the Fe salt into biochar, and the spectra displayed a few defined vibrational stretches at 1020–1160 and 1590 cm–1 in addition to the loss of characteristic IR stretching and bending vibrations of FeCl3·6H2O as shown in Figure S14A. Both pieces of data indicated that Fe had been altered by association with the biochar. With evidence for Fe incorporation into biochar, the adsorption capacity of the biochar for Fe(III) was then calculated to be 13.02 mg/g. To obtain the adsorption capacity, a spectroscopic calibration plot of the known concentrations of the metal salt was obtained (Figure S15) and the amount of Fe(III) remaining was calculated in the filtrate after the treated biochar. Next, the thermal stability of the treated biochar was examined to ensure that the biochar does not degrade with the addition of Fe salt. The TGA and DTG curves of the biochar after treatment with FeCl3 displayed a robust material in comparison to the pristine biochar (Figure S14B and Figure S16). After validating the incorporation of Fe(III) into biochar and investigating the stability of the material thermogravimetrically, the treated biochar’s binding capability to Bf was tested via spectroscopic analysis. The modified biochar was solubilized in DCM and subjected to 3 equiv of Bf (with respect to adsorbed Fe(III) concentration calculated using a calibration plot) and stirred at 60 °C for 1 h. The complex formed displayed a sharp absorption band at 554 nm (Figure S13) consistent with the stable Fe(Bf)3Cl3 complex. This result confirmed that Fe-incorporated biochar can be employed to suppress Bf emissions. Additionally, this simple procedure for identifying and incorporating Earth-Abundant metal salts into modifiers could be broadly applicable in sustainable engineering.

Conclusion

Click to copy section linkSection link copied!

Asphalt is a nontraditional source of pollutants that adversely affects air quality and public health. To investigate sustainable and cost-effective asphalt materials that can retain VOCs under environmentally relevant conditions, we examined the use of Earth-Abundant metal salt dopants. It was found that FeCl3 was a promising candidate for binding a selection of VOCs using UV–vis and TGA-FTIR spectroscopy. The formation of an octahedral Fe(Bf)3Cl3 complex when FeCl3 was in the presence of an excess of Bf was supported by stoichiometry studies, DFT, EPR, and magnetic moment analyses. The Fe-Bf complex was determined to be thermally stable up to 360 °C and resistant to degradation upon exposure to visible light. Selectivity studies between distinct VOCs in a homogeneous medium demonstrated that once Bf coordinated to the Fe salt, the complex was stable even in the presence of 1-methyl naphthalene, fluoranthene, anthracene, 9-hydroxy fluorene, catechol, salicylic acid, and benzothiophene. However, nitrogen-containing VOCs such as quinoline and oxygen-containing VOCs such as triethylene glycol and 2,3-dihydrobenzofuran were able to successfully replace Bf and form new Fe coordination complexes. The feasibility of applying this promising dopant to asphalt was demonstrated when FeCl3 was successfully adsorbed on Algal Biochar, which is a known asphalt binder. The Fe-enriched Biochar exhibited thermal stability by TGA-DSC analysis. Additionally, Fe-biochar was shown to bind Bf, exhibiting the same spectroscopic signatures characteristic of Fe(Bf)3Cl3. The present study is promising for the development of sustainable asphalt engineering where asphalt modifiers with the late-stage addition of inexpensive Earth-Abundant metal salts can be used to combat VOC emissions.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.4c02095.

  • This includes the following: general information about reagents, solvents, and instrumentation as well as procedures for experimentation and analysis (UV–vis,TGA-FTIR, TGA-DSC, EPR, and DFT studies (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
    • Elham H. Fini - School of Sustainable Engineering and the Built Environment, Arizona State University, 660 S. College Avenue, Tempe, Arizona 85287-3005, United States Email: [email protected]
    • Laura K. G. Ackerman-Biegasiewicz - Department of Chemistry, Emory University, Atlanta, Georgia 30322, United States Email: [email protected]
  • Authors
    • Harpreet Kaur - Department of Chemistry, Emory University, Atlanta, Georgia 30322, United States
    • Reem Nsouli - Department of Chemistry, Emory University, Atlanta, Georgia 30322, United States
    • Gabriella Cerna - School of Molecular Sciences, Arizona State University, 660 S. College Avenue, Tempe, Arizona 85287-3005, United States
    • Saba Shariati - School of Sustainable Engineering and the Built Environment, Arizona State University, 660 S. College Avenue, Tempe, Arizona 85287-3005, United StatesOrcidhttps://orcid.org/0000-0002-1993-8758
    • Marco Flores - School of Molecular Sciences, Arizona State University, 660 S. College Avenue, Tempe, Arizona 85287-3005, United StatesOrcidhttps://orcid.org/0000-0003-4139-7094
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

The authors acknowledge the contribution and support of the Transportation Research Board Minority Student Fellows Program from the Arizona Department of Transportation. This research is sponsored in-part by the Gordon and Betty Moore Foundation GBMF11403 (10.37807/GBMF11403) and the National Science Foundation (Award 1935723).

References

Click to copy section linkSection link copied!

This article references 36 other publications.

  1. 1
    Lin, S.; Hung, W.; Leng, Z. Air pollutant emissions and acoustic performance of hot mix asphalts. Constr. Build. Mater. 2016, 129, 110,  DOI: 10.1016/j.conbuildmat.2016.11.013
  2. 2
    Montero-Montoya, R.; López-Vargas, R.; Arellano-Aguilar, O. Volatile Organic Compounds in Air: Sources, Distribution, Exposure and Associated Illnesses in Children. Ann. Glob. Health. 2018, 84 (2), 225238,  DOI: 10.29024/aogh.910
  3. 3
    The Government of the Hong Kong Special Administrative Region Indoor Air Quality Management Group Guidance Notes for the Management of Indoor Air Quality in Offices and Public Places. www.iaq.gov.hk/wp-content/uploads/2021/04/gn_officeandpublicplace_eng-2019.pdf (accessed 2019).
  4. 4
    Mousavi, M.; Aldagari, S.; Fini, E. H. Adsorbing Volatile Organic Compounds within Bitumen Improves Colloidal Stability and Air Quality. ACS Sustainable Chem. Eng. 2023, 11, 95819594,  DOI: 10.1021/acssuschemeng.3c00539
  5. 5
    The World Health Organization Agents Classified by the IARC Monographs. monographs.iarc.who.int/agents-classified-by-the-iarc/ (assessed 2022).
  6. 6
    Cui, P.; Schito, G.; Cui, Q. VOC emissions from asphalt pavement and health risks to construction workers. J. Clean. Prod. 2020, 244, 118757118768,  DOI: 10.1016/j.jclepro.2019.118757
  7. 7
    Zhou, B.; Gong, G.; Wang, C. Characteristics and assessment of volatile organic compounds from different asphalt binders in laboratory. Transp. Res. D 2023, 118, 103708103724,  DOI: 10.1016/j.trd.2023.103708
  8. 8
    Khare, P.; Machesky, J.; Soto, R.; He, M.; Presto, A. A.; Gentner, D. R. Asphalt-related emissions are a major missing nontraditional source of secondary organic aerosol precursors. Science 2020, 6,  DOI: 10.1126/sciadv.abb9785 .
  9. 9
    Dhar, P. Scientists are engineering Asphalt that is safer for humans and the environment. ACS Cent. Sci. 2023, 9, 10731075,  DOI: 10.1021/acscentsci.3c00653
  10. 10
    Hung, A. M.; Kazembeyki, M.; Hoover, C. G.; Fini, E. H. Evolution of Morphological and Nanomechanical Properties of Bitumen Thin Films as a Result of Compositional Changes Due to Ultraviolet Radiation. ACS Sustain. Chem. Eng. 2019, 7, 1800518014,  DOI: 10.1021/acssuschemeng.9b04846
  11. 11
    Clark, C. R.; Burnett, D. M.; Parker, C. M.; Arp, E. W.; Swanson, M. S.; Minsavage, G. D.; Kriech, A. J.; Osborn, L. V.; Freeman, J. J.; Barter, R. A.; Newton, P. E.; Beazley, S. L.; Stewart, C. W. Asphalt fume dermal carcinogenicity potential: I. Dermal carcinogenicity evaluation of asphalt (bitumen) fume condensates. Regul. Toxicol. Pharmacol. 2011, 61, 916,  DOI: 10.1016/j.yrtph.2011.04.003
  12. 12
    Freeman, J. J.; Schreiner, C. A.; Beazley, S.; Burnett, D. M.; Clark, C. R.; Mahagaokar, S.; Parker, C. M.; Stewart, C. W.; Swanson, M. S.; Arp, E. W. Asphalt fume dermal carcinogenicity potential: II. Initiation-promotion assay of type III built-up roofing asphalt. Regul. Toxicol. Pharmacol. 2011, 61, 1722,  DOI: 10.1016/j.yrtph.2011.05.008
  13. 13
    Grahn, K.; Gustavsson, P.; Andersson, T.; Lindén, A.; Hemmingsson, T.; Selander, J.; Wiebert, P. Occupational exposure to particles and increased risk of developing chronic obstructive pulmonary disease (COPD): A population-based cohort study in Stockholm, Sweden. Environ. Res. 2021, 200, 111739111750,  DOI: 10.1016/j.envres.2021.111739
  14. 14
    Ye, W.; Jiang, W.; Li, P.; Yuan, D.; Shan, J.; Xiao, J. Analysis of mechanism and time-temperature equivalent effects of asphalt binder in short-term aging. Constr. Build. Mater. 2019, 215, 823838,  DOI: 10.1016/j.conbuildmat.2019.04.197
  15. 15
    Zhang, H.; Duan, H.; Zhu, C.; Chen, C.; Luo, H. Mini-Review on the Application of Nanomaterials in Improving Anti-Aging Properties of Asphalt. Energy Fuels 2021, 35, 1101711036,  DOI: 10.1021/acs.energyfuels.1c01035
  16. 16
    Pahlavan, F.; Gholipour, A.; Zhou, T.; Fini, E. H. Cleaner Asphalt Production by Suppressing Emissions Using Phenolic Compounds. ACS Sustainable Chem. Eng. 2023, 11, 27372751,  DOI: 10.1021/acssuschemeng.2c05345
  17. 17
    Cheraghian, G.; Cannone Falchetto, A.; You, Z.; Chen, S.; Kim, Y. S.; Westerhoff, J.; Moon, K. H.; Wistuba, M. P. Warm mix asphalt technology: An up to date review. J. Clean. Prod. 2020, 268, 122128122145,  DOI: 10.1016/j.jclepro.2020.122128
  18. 18
    Wang, M.; Wang, C.; Huang, S.; Yuan, H. Study on asphalt volatile organic compounds emission reduction: A state-of-the-art review. J. Clean. Prod. 2021, 318, 128596128610,  DOI: 10.1016/j.jclepro.2021.128596
  19. 19
    Mousavi, M.; Martis, V.; Fini, E. H. Inherently Functionalized Carbon from Algae to Adsorb Precursors of Secondary Organic Aerosols in Noncombustion Sources. ACS Sustainable Chem. Eng. 2021, 9, 1437514384,  DOI: 10.1021/acssuschemeng.1c03827
  20. 20
    Yang, X.; Shen, A.; Su, Y.; Zhao, W. Effects of Alumina Trihydrate (ATH) and Organic Montmorillonite (OMMT) on Asphalt Fume Emission and Flame Retardancy Properties of SBS Modified Asphalt. Constr. Build. Mater. 2020, 236, 117576117587,  DOI: 10.1016/j.conbuildmat.2019.117576
  21. 21
    Zhang, X.; Xiao, Y.; Long, Y.; Chen, Z.; Cui, P.; Wu, R.; Chang, X. VOCs Reduction in Bitumen Binder with Optimally Designed Ca(OH)2-Incorporated Zeolite. Constr. Build. Mater. 2021, 279, 122485122496,  DOI: 10.1016/j.conbuildmat.2021.122485
  22. 22
    Cui, P.; Wu, S.; Xiao, Y.; Wan, M.; Cui, P. Inhibiting effect of Layered Double Hydroxides on the emissions of volatile organic compounds from bituminous materials. J. Clean. Prod. 2015, 108, 987991,  DOI: 10.1016/j.jclepro.2015.06.115
  23. 23
    Chang, X.; Wan, L.; Long, Y.; Xiao, Y.; Xue, Y. Optimal Zeolite structure design for VOC emission Reduction in Asphalt materials. Const. Build. Mater. 2023, 366, 130227130236,  DOI: 10.1016/j.conbuildmat.2022.130227
  24. 24
    Wu, R.; Xiao, Y.; Zhang, P.; Lin, J.; Cheng, G.; Chen, Z.; Yu, R. Asphalt VOCs Reduction of Zeolite synthesized from solid wastes of Red Mud and Steel Slag. J. Clean. Prod. 2022, 345, 131078131089,  DOI: 10.1016/j.jclepro.2022.131078
  25. 25
    Tang, N.; Yang, K.-K.; Alrefaei, Y.; Dai, J.-G.; Wu, L.-M.; Wang, Q. Reduce VOCs and PM emissions of warm-mix asphalt using geopolymer additives. Constr. Build. Mater. 2020, 244, 118338118350,  DOI: 10.1016/j.conbuildmat.2020.118338
  26. 26
    Mousavi, M.; Aldagari, S.; Crocker, M. S.; Ackerman-Biegasiewicz, L. K. G.; Fini, E. H. Iron-Rich Biochar to Adsorb Volatile Organic Compounds Emitted from Asphalt-Surfaced Areas. ACS Sustainable Chem. Eng. 2023, 11, 28852896,  DOI: 10.1021/acssuschemeng.2c06292
  27. 27
    United States Agency for Toxic Substances and Disease Registry Toxicological Profile for 2,3-Benzofuran, US Department of Health and Human Services, Washington DC 1992. https://wwwn.cdc.gov/TSP/ToxProfiles/ToxProfiles.aspx?id=915&tid=187.
  28. 28
    Pei, Y.; Qin, J.; Wang, J.; Hu, Y. Fe-based metal organic framework derivative with enhanced Lewis acidity and hierarchical pores for excellent adsorption of oxygenated volatile organic compounds. Sci. Total Environ. 2021, 790, 148132148143,  DOI: 10.1016/j.scitotenv.2021.148132
  29. 29
    Li, X.; Zhang, L.; Yang, Z.; Wang, P.; Yan, Y.; Ran, J. Adsorption materials for volatile organic compounds (VOCs) and the key factors for VOCs adsorption process: A review. Sep. Purif. Technol. 2020, 235, 116213116273,  DOI: 10.1016/j.seppur.2019.116213
  30. 30
    Espinoza, J.; Medina, C.; Calabi-Floody, A.; Sánchez-Alonso, E.; Valdés, G.; Quiroz, A. Evaluation of Reductions in Fume Emissions (VOCs and SVOCs) from Warm Mix Asphalt Incorporating Natural Zeolite and Reclaimed Asphalt Pavement for Sustainable Pavements. Sustainability. 2020, 12, 95469563,  DOI: 10.3390/su12229546
  31. 31
    Kaur, M.; Raj, P.; Singh, N.; Kuwar, A.; Kaur, N. Benzimidazole-Based Imine-Linked Copper Complexes in Food Safety: Selective Detection of Cyproheptadine and Thiabendazole. ACS Sustainable Chem. Eng. 2018, 6, 37233732,  DOI: 10.1021/acssuschemeng.7b04084
  32. 32
    Kaur, H.; Kaur, N.; Singh, N. Nitrogen and sulfur co-doped fluorescent carbon dots for the trapping of Hg(II) ions from water. Mater. Adv. 2020, 1, 30093021,  DOI: 10.1039/D0MA00448K
  33. 33
    Priem, A.; van Bentum, P. J. M.; Hagen, W. R.; Reijerse, E. J. Estimation of High-Order Magnetic Spin Interactions of Fe(III) and Gd(III) Ions Doped in α-Alumina Powder with Multifrequency EPR. Appl. Magn. Reson. 2001, 21, 535548,  DOI: 10.1007/BF03162427
  34. 34
    Solano-Peralta, A.; Saucedo-Vázquez, J. P.; Escudero, R.; Höpfl, H.; El-Mkami, H.; Smith, G. M.; Sosa-Torres, M. E. Magnetic and High-Frequency EPR Studies of an Octahedral Fe(III) Compound with Unusual Zero-Field Splitting Parameters. Dalton Trans. 2009, 9, 16681674,  DOI: 10.1039/b814225d
  35. 35
    Nie, L.; Feng, X.; Song, H.; Li, Z.; Yao, S. A New Integrated Method of Magnetic Separation of Isoquinoline Alkaloids from Coptis chinensis based on their Magnetized Derivatives and Key Physical Properties. New J. Chem. 2020, 44, 71057115,  DOI: 10.1039/D0NJ00731E
  36. 36
    Rozewski, E.; Taqi, O.; Fini, E. H.; Lewinski, N. A.; Klein-Seetharaman, J. Systems biology of asphalt pollutants and their human molecular targets. Front. Syst. Biol. 2023, 2 DOI: 10.3389/fsysb.2022.928962 .

Cited By

Click to copy section linkSection link copied!

This article has not yet been cited by other publications.

Open PDF

ACS Omega

Cite this: ACS Omega 2024, 9, 21, 22941–22951
Click to copy citationCitation copied!
https://doi.org/10.1021/acsomega.4c02095
Published May 13, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

693

Altmetric

-

Citations

-
Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Significance of inhibiting asphalt-derived VOC emissions. (A) Prevalence of VOC emissions upon exposure of asphalt to heat and light. (B) Limitations of the incorporation of inorganic materials in asphalt as VOC inhibitors. (C) Studying the use of Earth-Abundant metal salts in asphalt binder to deter VOC emission.

    Figure 2

    Figure 2. (A) Change in λmax between metal salts with and without Bf. Metal salt (0.05 mmol) and Bf (0.05 mmol) were dissolved in 1 mL of DCM and stirred at 28 °C for 1 h prior to UV–vis analysis. (B) UV–vis spectrum of 200 μM Bf (green), 200 μM FeCl3 (red), and a 1:1 mixture of FeCl3 and Bf (purple). aBathochromic shift was observed relative to the absorbance of the metal salt.

    Figure 3

    Figure 3. UV–vis spectrum of (A) 200 μM FeCl3 (pink), 200 μM Bf (green), and a 1:1 mixture of FeCl3 and Bf, stirred for 1 h in DCM in one of four conditions: (i) at 28 °C (purple); (ii) heated at 60 °C (mauve), (iii) 80 °C (dark blue), and (iv) 100 °C (light blue); and (B) 200 μM FeCl3 (pink), 200 μM Bf (green), and a 1:1 mixture of FeCl3 and Bf, stirred for 1 h in DCM under irradiation conditions at: (i) 390 nm (light blue), (ii) 456 nm (dark blue), and (iii) 525 nm (mauve) wavelengths.

    Figure 4

    Figure 4. (A) FeCl3 (16.2 mg, 0.1 mmol) and Bf (10.8 μL, 0.1 mmol) were mixed with dichloromethane (3.0 mL) at room temperature for 1 h. The solution was then concentrated, and a portion of the sample (16 mg) was transferred to a high-temperature Pt pan for TGA-FTIR analysis. Entry 1 displays a control run for the detection of pure Bf by FTIR at 82 °C upon thermal degradation of 10.8 μL Bf dissolved in 200 μL of DCM analyzed using TGA-FTIR. Entry 2 displays FTIR detection of Bf at 363 °C upon thermal degradation of the Fe-Bf complex using TGA at a ramp of 20 °C/min to 1000 °C. Entry 3 displays FTIR detection of Bf at 363 °C upon thermal degradation of Fe-Bf complex using TGA at a ramp of 20 °C/min to 40 °C followed by an isothermal hold at 40 °C for 10 min to ensure solvent evaporation. This was followed by another ramp at 20 °C/min to 200 °C followed by another isothermal hold for 40 min to assess the thermal stability of the Fe-Bf complex at environmental conditions. Finally, a ramp at 20 °C/min up to 1000 °C was applied to investigate overall thermal stability. (B) Gram-Schmidt profile for Bf (green) and a mixture of FeCl3 and Bf samples (purple). The method used included a ramp rate of 20 °C/min to 40 °C, followed by an isotherm hold at 40 °C for 10 min, and ramp rate of 20 °C/min up to 1000 °C.

    Figure 5

    Figure 5. (A) Job plot of the Fe-Bf complex prepared using different mole fractions of Fe(III) and Bf where the total molar concentration of the complex is kept constant; and (B) Molar Ratio plot of the Fe-Bf complex prepared keeping the number of moles of Fe(III) (nM) constant and varying the number of moles of Bf (nL). In both plots, the variations in the absorption intensity of the peak at 554 nm were observed, and the stoichiometry of the Fe-Bf complex was deduced from the maximal position in the Job plot and the highest absorbance in the Molar Ratio plot, suggesting a 1:3 stoichiometry for the Fe-Bf complex.

    Figure 6

    Figure 6. Experimental (black line) and simulated (red line) X-band EPR spectra of Fe-Bf at 120 K. The narrow line around 340 mT (marked with an asterisk) belongs to a minor unidentified species. Inset: parameters used to fit the EPR Spectrum of the Fe-Bf complex at 9.40 GHz and T = 120 K. aThe fitting parameters were the isotropic g-value (giso), the zero-field splitting parameters (D and E), and the isotropic line width (ΔB).

    Figure 7

    Figure 7. Proposed chemical structures of the high-spin Fe-Bf complex. (A) [Fe(Bf)3Cl2]Cl and (B) Fe(Bf)3Cl3 based on EPR and magnetic susceptibility results.

    Figure 8

    Figure 8. Optimized geometries of the Fe(Bf)3Cl3 complexes displaying bond length (Å) of Fe–O bond and Fe–C bond and stabilization energy (ES) of the two optimized geometries in kcal/mol. Atom colors are Fe-violet, Cl-green, O-red, C-gray, and H-white.

    Figure 9

    Figure 9. (A) Schematic representation of the binding activity of the complex with Oxygen-containing VOCs evaluated using UV–vis absorption spectroscopy displaying distinct absorption peaks (λmax) summarized in tabular form. Note: FeCl3 (0.1 mmol), Bf (0.3 mmol), and VOC (0.3 mmol) were dissolved in DCM (2 mL). The solutions were stirred at 60 °C for 1 h prior to UV–vis analysis. Note: ‘-’ indicates that no spectroscopic changes were observed. (B) Bar graph displaying variation in the intensity of the absorption peak at 554 nm (characteristic of the Fe-Bf complex) in the presence of different VOCs. The UV–vis absorption peak was observed after subjecting the Fe-Bf complex (50 mM) to 0.3 mmol of VOCs in DCM and diluting it before recording the spectra.

    Figure 10

    Figure 10. (A) UV–vis absorption spectra of the Fe-Bf complex in the presence of triethylene glycol, along with FeCl3, FeCl3-triethylene glycol, triethylene glycol, and the Fe-Bf complex as controls under the same conditions. (B) UV–vis absorption spectra of the Fe-Bf complex in the presence of increasing concentrations of triethylene glycol (0–40 mM), displaying distinct absorption peaks (316 and 365 nm) along with the disappearance of the absorption peak at 554 nm.

    Figure 11

    Figure 11. (A) UV–vis absorption spectra of the Fe-Bf complex in the presence of quinoline, along with FeCl3, FeCl3-quinoline, quinoline, and the Fe-Bf complex as controls under the same conditions. (B) UV–vis absorption spectra of the Fe-Bf complex in the presence of increasing concentrations of quinoline (0–20 mM) displayed distinct absorption peaks along with the disappearance of the absorption peak at 554 nm.

    Figure 12

    Figure 12. Optimized geometries ES and ΔES for the formation of (A) Fe(quinoline)3Cl3, (B) Fe(TEG)3Cl3, and (C) Fe(2,3-DHB)3Cl3. Atom colors are Fe-violet, Cl-green, O-red, N-blue, C-gray, and H-white.

  • References


    This article references 36 other publications.

    1. 1
      Lin, S.; Hung, W.; Leng, Z. Air pollutant emissions and acoustic performance of hot mix asphalts. Constr. Build. Mater. 2016, 129, 110,  DOI: 10.1016/j.conbuildmat.2016.11.013
    2. 2
      Montero-Montoya, R.; López-Vargas, R.; Arellano-Aguilar, O. Volatile Organic Compounds in Air: Sources, Distribution, Exposure and Associated Illnesses in Children. Ann. Glob. Health. 2018, 84 (2), 225238,  DOI: 10.29024/aogh.910
    3. 3
      The Government of the Hong Kong Special Administrative Region Indoor Air Quality Management Group Guidance Notes for the Management of Indoor Air Quality in Offices and Public Places. www.iaq.gov.hk/wp-content/uploads/2021/04/gn_officeandpublicplace_eng-2019.pdf (accessed 2019).
    4. 4
      Mousavi, M.; Aldagari, S.; Fini, E. H. Adsorbing Volatile Organic Compounds within Bitumen Improves Colloidal Stability and Air Quality. ACS Sustainable Chem. Eng. 2023, 11, 95819594,  DOI: 10.1021/acssuschemeng.3c00539
    5. 5
      The World Health Organization Agents Classified by the IARC Monographs. monographs.iarc.who.int/agents-classified-by-the-iarc/ (assessed 2022).
    6. 6
      Cui, P.; Schito, G.; Cui, Q. VOC emissions from asphalt pavement and health risks to construction workers. J. Clean. Prod. 2020, 244, 118757118768,  DOI: 10.1016/j.jclepro.2019.118757
    7. 7
      Zhou, B.; Gong, G.; Wang, C. Characteristics and assessment of volatile organic compounds from different asphalt binders in laboratory. Transp. Res. D 2023, 118, 103708103724,  DOI: 10.1016/j.trd.2023.103708
    8. 8
      Khare, P.; Machesky, J.; Soto, R.; He, M.; Presto, A. A.; Gentner, D. R. Asphalt-related emissions are a major missing nontraditional source of secondary organic aerosol precursors. Science 2020, 6,  DOI: 10.1126/sciadv.abb9785 .
    9. 9
      Dhar, P. Scientists are engineering Asphalt that is safer for humans and the environment. ACS Cent. Sci. 2023, 9, 10731075,  DOI: 10.1021/acscentsci.3c00653
    10. 10
      Hung, A. M.; Kazembeyki, M.; Hoover, C. G.; Fini, E. H. Evolution of Morphological and Nanomechanical Properties of Bitumen Thin Films as a Result of Compositional Changes Due to Ultraviolet Radiation. ACS Sustain. Chem. Eng. 2019, 7, 1800518014,  DOI: 10.1021/acssuschemeng.9b04846
    11. 11
      Clark, C. R.; Burnett, D. M.; Parker, C. M.; Arp, E. W.; Swanson, M. S.; Minsavage, G. D.; Kriech, A. J.; Osborn, L. V.; Freeman, J. J.; Barter, R. A.; Newton, P. E.; Beazley, S. L.; Stewart, C. W. Asphalt fume dermal carcinogenicity potential: I. Dermal carcinogenicity evaluation of asphalt (bitumen) fume condensates. Regul. Toxicol. Pharmacol. 2011, 61, 916,  DOI: 10.1016/j.yrtph.2011.04.003
    12. 12
      Freeman, J. J.; Schreiner, C. A.; Beazley, S.; Burnett, D. M.; Clark, C. R.; Mahagaokar, S.; Parker, C. M.; Stewart, C. W.; Swanson, M. S.; Arp, E. W. Asphalt fume dermal carcinogenicity potential: II. Initiation-promotion assay of type III built-up roofing asphalt. Regul. Toxicol. Pharmacol. 2011, 61, 1722,  DOI: 10.1016/j.yrtph.2011.05.008
    13. 13
      Grahn, K.; Gustavsson, P.; Andersson, T.; Lindén, A.; Hemmingsson, T.; Selander, J.; Wiebert, P. Occupational exposure to particles and increased risk of developing chronic obstructive pulmonary disease (COPD): A population-based cohort study in Stockholm, Sweden. Environ. Res. 2021, 200, 111739111750,  DOI: 10.1016/j.envres.2021.111739
    14. 14
      Ye, W.; Jiang, W.; Li, P.; Yuan, D.; Shan, J.; Xiao, J. Analysis of mechanism and time-temperature equivalent effects of asphalt binder in short-term aging. Constr. Build. Mater. 2019, 215, 823838,  DOI: 10.1016/j.conbuildmat.2019.04.197
    15. 15
      Zhang, H.; Duan, H.; Zhu, C.; Chen, C.; Luo, H. Mini-Review on the Application of Nanomaterials in Improving Anti-Aging Properties of Asphalt. Energy Fuels 2021, 35, 1101711036,  DOI: 10.1021/acs.energyfuels.1c01035
    16. 16
      Pahlavan, F.; Gholipour, A.; Zhou, T.; Fini, E. H. Cleaner Asphalt Production by Suppressing Emissions Using Phenolic Compounds. ACS Sustainable Chem. Eng. 2023, 11, 27372751,  DOI: 10.1021/acssuschemeng.2c05345
    17. 17
      Cheraghian, G.; Cannone Falchetto, A.; You, Z.; Chen, S.; Kim, Y. S.; Westerhoff, J.; Moon, K. H.; Wistuba, M. P. Warm mix asphalt technology: An up to date review. J. Clean. Prod. 2020, 268, 122128122145,  DOI: 10.1016/j.jclepro.2020.122128
    18. 18
      Wang, M.; Wang, C.; Huang, S.; Yuan, H. Study on asphalt volatile organic compounds emission reduction: A state-of-the-art review. J. Clean. Prod. 2021, 318, 128596128610,  DOI: 10.1016/j.jclepro.2021.128596
    19. 19
      Mousavi, M.; Martis, V.; Fini, E. H. Inherently Functionalized Carbon from Algae to Adsorb Precursors of Secondary Organic Aerosols in Noncombustion Sources. ACS Sustainable Chem. Eng. 2021, 9, 1437514384,  DOI: 10.1021/acssuschemeng.1c03827
    20. 20
      Yang, X.; Shen, A.; Su, Y.; Zhao, W. Effects of Alumina Trihydrate (ATH) and Organic Montmorillonite (OMMT) on Asphalt Fume Emission and Flame Retardancy Properties of SBS Modified Asphalt. Constr. Build. Mater. 2020, 236, 117576117587,  DOI: 10.1016/j.conbuildmat.2019.117576
    21. 21
      Zhang, X.; Xiao, Y.; Long, Y.; Chen, Z.; Cui, P.; Wu, R.; Chang, X. VOCs Reduction in Bitumen Binder with Optimally Designed Ca(OH)2-Incorporated Zeolite. Constr. Build. Mater. 2021, 279, 122485122496,  DOI: 10.1016/j.conbuildmat.2021.122485
    22. 22
      Cui, P.; Wu, S.; Xiao, Y.; Wan, M.; Cui, P. Inhibiting effect of Layered Double Hydroxides on the emissions of volatile organic compounds from bituminous materials. J. Clean. Prod. 2015, 108, 987991,  DOI: 10.1016/j.jclepro.2015.06.115
    23. 23
      Chang, X.; Wan, L.; Long, Y.; Xiao, Y.; Xue, Y. Optimal Zeolite structure design for VOC emission Reduction in Asphalt materials. Const. Build. Mater. 2023, 366, 130227130236,  DOI: 10.1016/j.conbuildmat.2022.130227
    24. 24
      Wu, R.; Xiao, Y.; Zhang, P.; Lin, J.; Cheng, G.; Chen, Z.; Yu, R. Asphalt VOCs Reduction of Zeolite synthesized from solid wastes of Red Mud and Steel Slag. J. Clean. Prod. 2022, 345, 131078131089,  DOI: 10.1016/j.jclepro.2022.131078
    25. 25
      Tang, N.; Yang, K.-K.; Alrefaei, Y.; Dai, J.-G.; Wu, L.-M.; Wang, Q. Reduce VOCs and PM emissions of warm-mix asphalt using geopolymer additives. Constr. Build. Mater. 2020, 244, 118338118350,  DOI: 10.1016/j.conbuildmat.2020.118338
    26. 26
      Mousavi, M.; Aldagari, S.; Crocker, M. S.; Ackerman-Biegasiewicz, L. K. G.; Fini, E. H. Iron-Rich Biochar to Adsorb Volatile Organic Compounds Emitted from Asphalt-Surfaced Areas. ACS Sustainable Chem. Eng. 2023, 11, 28852896,  DOI: 10.1021/acssuschemeng.2c06292
    27. 27
      United States Agency for Toxic Substances and Disease Registry Toxicological Profile for 2,3-Benzofuran, US Department of Health and Human Services, Washington DC 1992. https://wwwn.cdc.gov/TSP/ToxProfiles/ToxProfiles.aspx?id=915&tid=187.
    28. 28
      Pei, Y.; Qin, J.; Wang, J.; Hu, Y. Fe-based metal organic framework derivative with enhanced Lewis acidity and hierarchical pores for excellent adsorption of oxygenated volatile organic compounds. Sci. Total Environ. 2021, 790, 148132148143,  DOI: 10.1016/j.scitotenv.2021.148132
    29. 29
      Li, X.; Zhang, L.; Yang, Z.; Wang, P.; Yan, Y.; Ran, J. Adsorption materials for volatile organic compounds (VOCs) and the key factors for VOCs adsorption process: A review. Sep. Purif. Technol. 2020, 235, 116213116273,  DOI: 10.1016/j.seppur.2019.116213
    30. 30
      Espinoza, J.; Medina, C.; Calabi-Floody, A.; Sánchez-Alonso, E.; Valdés, G.; Quiroz, A. Evaluation of Reductions in Fume Emissions (VOCs and SVOCs) from Warm Mix Asphalt Incorporating Natural Zeolite and Reclaimed Asphalt Pavement for Sustainable Pavements. Sustainability. 2020, 12, 95469563,  DOI: 10.3390/su12229546
    31. 31
      Kaur, M.; Raj, P.; Singh, N.; Kuwar, A.; Kaur, N. Benzimidazole-Based Imine-Linked Copper Complexes in Food Safety: Selective Detection of Cyproheptadine and Thiabendazole. ACS Sustainable Chem. Eng. 2018, 6, 37233732,  DOI: 10.1021/acssuschemeng.7b04084
    32. 32
      Kaur, H.; Kaur, N.; Singh, N. Nitrogen and sulfur co-doped fluorescent carbon dots for the trapping of Hg(II) ions from water. Mater. Adv. 2020, 1, 30093021,  DOI: 10.1039/D0MA00448K
    33. 33
      Priem, A.; van Bentum, P. J. M.; Hagen, W. R.; Reijerse, E. J. Estimation of High-Order Magnetic Spin Interactions of Fe(III) and Gd(III) Ions Doped in α-Alumina Powder with Multifrequency EPR. Appl. Magn. Reson. 2001, 21, 535548,  DOI: 10.1007/BF03162427
    34. 34
      Solano-Peralta, A.; Saucedo-Vázquez, J. P.; Escudero, R.; Höpfl, H.; El-Mkami, H.; Smith, G. M.; Sosa-Torres, M. E. Magnetic and High-Frequency EPR Studies of an Octahedral Fe(III) Compound with Unusual Zero-Field Splitting Parameters. Dalton Trans. 2009, 9, 16681674,  DOI: 10.1039/b814225d
    35. 35
      Nie, L.; Feng, X.; Song, H.; Li, Z.; Yao, S. A New Integrated Method of Magnetic Separation of Isoquinoline Alkaloids from Coptis chinensis based on their Magnetized Derivatives and Key Physical Properties. New J. Chem. 2020, 44, 71057115,  DOI: 10.1039/D0NJ00731E
    36. 36
      Rozewski, E.; Taqi, O.; Fini, E. H.; Lewinski, N. A.; Klein-Seetharaman, J. Systems biology of asphalt pollutants and their human molecular targets. Front. Syst. Biol. 2023, 2 DOI: 10.3389/fsysb.2022.928962 .
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.4c02095.

    • This includes the following: general information about reagents, solvents, and instrumentation as well as procedures for experimentation and analysis (UV–vis,TGA-FTIR, TGA-DSC, EPR, and DFT studies (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.