Skip to main page content
U.S. flag

An official website of the United States government

Dot gov

The .gov means it’s official.
Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

Https

The site is secure.
The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
. 2023 Nov 13;8(47):44942-44954.
doi: 10.1021/acsomega.3c06360. eCollection 2023 Nov 28.

2D Fluorinated Graphene Oxide (FGO)-Polyethyleneimine (PEI) Based 3D Porous Nanoplatform for Effective Removal of Forever Toxic Chemicals, Pharmaceutical Toxins, and Waterborne Pathogens from Environmental Water Samples

Affiliations

2D Fluorinated Graphene Oxide (FGO)-Polyethyleneimine (PEI) Based 3D Porous Nanoplatform for Effective Removal of Forever Toxic Chemicals, Pharmaceutical Toxins, and Waterborne Pathogens from Environmental Water Samples

Avijit Pramanik et al. ACS Omega. .

Abstract

Although water is essential for life, as per the United Nations, around 2 billion people in this world lack access to safely managed drinking water services at home. Herein we report the development of a two-dimensional (2D) fluorinated graphene oxide (FGO) and polyethylenimine (PEI) based three-dimensional (3D) porous nanoplatform for the effective removal of polyfluoroalkyl substances (PFAS), pharmaceutical toxins, and waterborne pathogens from contaminated water. Experimental data show that the FGO-PEI based nanoplatform has an estimated adsorption capacity (qm) of ∼219 mg g-1 for perfluorononanoic acid (PFNA) and can be used for 99% removal of several short- and long-chain PFAS. A comparative PFNA capturing study using different types of nanoplatforms indicates that the qm value is in the order FGO-PEI > FGO > GO-PEI, which indicates that fluorophilic, electrostatic, and hydrophobic interactions play important roles for the removal of PFAS. Reported data show that the FGO-PEI based nanoplatform has a capability for 100% removal of moxifloxacin antibiotics with an estimated qm of ∼299 mg g-1. Furthermore, because the pore size of the nanoplatform is much smaller than the size of pathogens, it has a capability for 100% removal of Salmonella and Escherichia coli from water. Moreover, reported data show around 96% removal of PFAS, pharmaceutical toxins, and pathogens simultaneously from spiked river, lake, and tap water samples using the nanoplatform.

PubMed Disclaimer

Conflict of interest statement

The authors declare no competing financial interest.

Figures

Figure 1
Figure 1
Scheme shows the synthetic path used for the development of the 2D fluorinated graphene oxide and polyethylenimine (PEI) based three-dimensional (3D) porous nanoplatform.
Figure 2
Figure 2
(A) TEM image shows the morphology of 2D graphene oxide. Inserted energy dispersive spectrometry (EDS) mapping shows the presence of C and O in 2D graphene oxide. (B) TEM image shows the morphology of 2D fluorinated graphene oxide. (C) SEM image shows the morphology of the 2D fluorinated graphene oxide and PEI based 3D nanoplatform. Inserted energy dispersive spectrometry (EDS) mapping shows the presence of C, O, N, and F in the 3D nanoplatform. Another inserted image shows the nanoplatform developed using the FGO-PEI based 3D nanoplatform. (D) The XPS spectrum from FGO confirms peaks at 288.5, 533.2, and 688.3 eV, which are due to C, O, and F, respectively. (E) XRD data from FGO show the presence of (002) and (100) reflection peaks. (F) FTIR spectra from the FGO-PEI based 3D nanoplatform shows the presence of −C–F, −C=O, −O–H, amide-I, and amide-A peaks. (G) Raman spectra from the FGO-PEI based nanoplatform shows the presence of D and G peaks. (H) N2 adsorption/desorption isotherm of the FGO-PEI based 3D nanoplatform. (I) Pore size distributions from the FGO-PEI based 3D nanoplatform show that pore size varies from 5 to 100 nm, with the highest pore density being around 30 nm.
Figure 3
Figure 3
(A) Perfluorobutanoic acid (PFBA) removal efficiency from drinking water using the 2D-GO based nanoplatform, PEI-attached 2D-GO (GO-PEI) based nanoplatform, and PEI-attached 2D F-GO (FGO-PEI) based nanoplatform. For this experiment, we used 1000 ng/L of PFBA infected drinking water. (B) Perfluorobutanesulfonic acid (PFBS) removal efficiency from drinking water using the 2D-GO based nanoplatform, GO-PEI based nanoplatform, and FGO-PEI based nanoplatform. For this experiment, we used 1000 ng/L of PFBS infected drinking water. (C) Perfluorohexanesulfonate (PFHxS) removal efficiency from drinking water using the 2D-GO based nanoplatform, GO-PEI based nanoplatform, and FGO-PEI based nanoplatform. For this experiment, we used 1000 ng/L of PFHxS infected drinking water. (D) Variation of perfluorononanoic acid (PFNA) removal efficiency with time for the GO based nanoplatform, PEI-attached 2D-GO (GO-PEI) based nanoplatform, and PEI-attached 2D F-GO (FGO-PEI) based nanoplatform. (E) Plots show the time-dependent removal efficiency for PFBS and PFHxS using the FGO-PEI based nanoplatform. For this experiment, we used 1000 ng/L of PFBS or PFHxS infected water samples. (F) Plot shows the variation of (t/qt) with time for PFNA using the PEI-FGO and GO adsorber separately, where qt is the quantity of PFNA removed per gram of the FGO-PEI or GO nanoplatform. (G) Plot shows the variation of 1/qe with 1/Ce for PFNA using the PEI-FGO and PEI-GO adsorber separately, where qe is the quantity of PFNA adsorbed at equilibrium and Ce is the concentration of PFNA. (H) Plot shows how the PFNA removal efficiency varies with pH when the FGO-PEI based nanoplatform was used. (I) PFNA removal efficiency from tap water, Mississippi river water, lake water, and drinking water using the FGO-PEI based nanoplatform. For this experiment, we used 1000 ng/L of PFBA infected water samples. (J) PFBS removal efficiency from tap water, Mississippi river water, lake water, and drinking water using the FGO-PEI based nanoplatform. For this experiment, we used 1000 ng/L of PFBS infected water samples. (K) Removal efficiency of PFBS, PFBA, PFHxS, and PFNA simultaneously from tap water, Mississippi river water, lake water, and drinking water using the FGO-PEI based nanoplatform. For this experiment, we used 250 ng/L of PFBS, 250 ng/L ng/L of PFBA, 250 ng/L of PFHxS, and 250 ng/L of PFNA infected water samples. (L) Plot shows how the removal efficiency for PFBS, PFBA, PFHxS, and PFNA together varies with the number of cycles of filtration when we used the FGO-PEI based nanoplatform. For this experiment, we used 250 ng/L of PFBS, 250 ng/L of PFBA, 250 ng/L of PFHxS, and 250 ng/L of PFNA infected water samples.
Figure 4
Figure 4
(A) Surface enhanced Raman spectra (SERS) from water samples of moxifloxacin antibiotics (1000 ng/L) before filtration. SERS from water samples after filtration using the FGO-PEI based nanoplatform. (B) SERS from water samples of tetracycline antibiotics (1000 ng/L) before filtration. SERS from water samples after filtration using the FGO-PEI based nanoplatform. (C) Plot shows how the SERS intensity at 1620 cm–1 for the C=C stretch from moxifloxacin antibiotics varies with the concentration (ng/L). I0 is the SERS intensity at 1620 cm–1 when the concentration is 1000 ng/L. If is the SERS intensity at 1620 cm–1 when the concentration varies from 1000 ng/L to 5 pg/L. (D) Plot shows how the SERS intensity at 1230 cm–1 for amide-III band from tetracycline antibiotics varies with the concentration (ng/L). I0is the SERS intensity at 1230 cm–1 when the concentration is 1000 ng/L. If is the SERS intensity at 1230 cm–1 when the concentration varies from 1000 ng/L to 5 pg/L. (E) Tetracycline and moxifloxacin antibiotic removal efficiency from drinking water using theFGO-PEI based nanoplatform. For this experiment, we used 1000 ng/L of antibiotic infected drinking water. (F) Tetracycline and moxifloxacin antibiotic removal efficiency from drinking water using the GO-PEI based nanoplatform. For this experiment, we used 1000 ng/L of each antibiotic separately. (G) Tetracycline and moxifloxacin antibiotic removal efficiency from drinking water using the FGO based nanoplatform. For this experiment, we used 1000 ng/L of antibiotic infected drinking water. (H) Variation of moxifloxacin antibiotic removal efficiency with time for the FGO and FGO-PEI based nanoplatform. (I) Plot shows the variation of 1/qe with 1/Ce for moxifloxacin antibiotics using the PEI-FGO adsorber, where qe is the quantity of moxifloxacin antibiotics absorbed at equilibrium and Ce is the concentration of moxifloxacin antibiotics. (J) Moxifloxacin antibiotic removal efficiency from tap water, Mississippi river water, lake water, and drinking water using the FGO-PEI based nanoplatform. For this experiment, we used 1000 ng/L of moxifloxacin antibiotic infected water samples. (K) Tetracycline antibiotic removal efficiency from tap water, Mississippi river water, lake water, and drinking water using the FGO-PEI based nanoplatform. For this experiment, we used 1000 ng/L of tetracycline antibiotic infected water samples. (L) Removal efficiency of PFNA, PFBS, tetracycline, and moxifloxacin antibiotics simultaneously from tap water, Mississippi river water, lake water, and drinking water using the FGO-PEI based nanoplatform. For this experiment, we used 250 ng/L of perfluorobutanesulfonic acid (PFBS), 250 ng/L of PFNA, 250 ng/L of tetracycline, and 250 ng/L of moxifloxacin antibiotic infected water samples.
Figure 5
Figure 5
(A) SEM image shows that Salmonella waterborne pathogens are captured by the FGO-PEI based nanoplatform during filtration. (B) TEM image shows that Escherichia coli waterborne pathogens are captured by the FGO-PEI based nanoplatform during filtration. (C) Colony counting data show that Escherichia coli is present in water samples before filtration. (D) Colony counting data show that no Escherichia coli is present in water samples after filtration, which indicates that 100% bacteria is captured during filtration. (E) Salmonella and Escherichia coli waterborne pathogen removal efficiency from drinking water using the FGO-PEI based nanoplatform. For this experiment, we used 103 CFU/mL of each bacterium. (F) Salmonella and Escherichia coli waterborne pathogen removal efficiency from drinking water using the GO-PEI based nanoplatform. For this experiment, we used 103 CFU/mL of each bacterium. G) Salmonella and Escherichia coli waterborne pathogen removal efficiency from drinking water using the FGO based nanoplatform. For this experiment, we used 103 CFU/mL of each bacterium. (H) Escherichia coli waterborne pathogen removal efficiency from tap water, Mississippi river water, lake water, and drinking water using the FGO-PEI based nanoplatform. For this experiment, we used 103 CFU/mL of bacteria infected water samples. (I) Salmonella waterborne pathogen removal efficiency from tap water, Mississippi river water, lake water, and drinking water using the FGO-PEI based nanoplatform. For this experiment, we used 103 CFU/mL of bacteria infected water samples. (J) Simultaneously Salmonella and Escherichia coli waterborne pathogen removal efficiency from tap water, Mississippi river water, lake water, and drinking water using the FGO-PEI based nanoplatform. For this experiment, we used 500 CFU/mL of bacteria infected water samples. (K). Removal efficiency of PFBS, PFBA, tetracycline, and moxifloxacin antibiotics and Salmonella and Escherichia coli waterborne pathogens simultaneously from tap water, Mississippi river water, lake water, and drinking water using the FGO-PEI based nanoplatform. For this experiment, we used water samples infected with 333 ng/L of PFBS, 333 ng/L of PFBA, 333 ng/L of moxifloxacin antibiotics, and 500 CFU/mL of each bacterium. (L) Plot shows how the removal efficiency PFBS, PFBA, tetracycline, and moxifloxacin antibiotics varies with the number of cycles of filtration when we have used the FGO-PEI based nanoplatform in the presence of Salmonella and Escherichia coli waterborne pathogens. For this experiment, we used water samples infected with 333 ng/L of PFBS, 333 ng/L of PFBA, 333 ng/L of moxifloxacin antibiotics, and 500 CFU/mL of each bacterium.

Similar articles

References

    1. 2.1 billion people lack safe drinking water at home, more than twice as many lack safe sanitation https://www.unicef.cn/en/press-releases/21-billion-people-lack-safe-drin...2017.
    1. United States Environmental Protection Agency . Drinking Water Health Advisories for PFAS: Fact Sheet for Public Water Systems (PFOA, PFOS, GenX Chemicals and PFBS). https://www.epa.gov/system/files/documents/2022-06/drinking-water-ha-pfa...2022.
    1. Smalling K. L.; Romanok K. M.; Bradley P. M.; Morriss M. C.; Gray J. L.; Kanagy L. K.; Gordon S. E.; Williams B. M.; Breitmeyer S. E.; Jones D. K.; DeCicco L. A.; Eagles-Smith C. A.; Wagner T. Per- and Polyfluoroalkyl Substances (PFAS) in U.S. Tapwater: Comparison of Public-Supply and Underserved Private-Well Exposures and Associated Health Implications. Environ. Int. 2023, 178, 10803310.1016/j.envint.2023.108033. - DOI - PubMed
    1. Turner S. W. D.; Rice J. S.; Nelson K. D.; Vernon C. R.; McManamay R.; Dickson K.; Marston L. Comparison of Potential Drinking Water Source Contamination across One Hundred U.S. Cities. Nat. Commun. 2021, 12, 7254.10.1038/s41467-021-27509-9. - DOI - PMC - PubMed
    1. Evich M. G.; Davis M. J. B.; McCord J. P.; Acrey B.; Awkerman J. A.; Knappe D. R. U.; Lindstrom A. B.; Speth T. F.; Tebes-Stevens C.; Strynar M. J.; Wang Z.; Weber E. J.; Henderson W. M.; Washington J. W. Per- and polyfluoroalkyl substances in the environment. Science 2022, 375, eabg906510.1126/science.abg9065. - DOI - PMC - PubMed