[go: up one dir, main page]

 
 

Carbon Nanostructures as Promising Future Materials: 2nd Edition

A special issue of Nanomaterials (ISSN 2079-4991). This special issue belongs to the section "2D and Carbon Nanomaterials".

Deadline for manuscript submissions: 30 September 2024 | Viewed by 15413

Special Issue Editors


E-Mail Website
Guest Editor
“Cristofor Simionescu” Faculty of Chemical Engineering and Environment Protection, “Gheorghe Asachi” Technical University, Iasi, Romania
Interests: polysaccharide modification; bioactive polymers; biomaterials; hydrogels; interpenetrated networks; micro- and nanoparticles (spheres and capsules); hybrid and functionalized nanoparticles for drug targeting; drug delivery; polymer–drug conjugates
Special Issues, Collections and Topics in MDPI journals

Special Issue Information

Dear Colleagues,

We are pleased to invite you to submit an article to our Special Issue entitled “Carbon Nanostructures as Promising Future Materials”. Carbon is an element well-known for its allotropic states, which are determined by various structures found in diamond, graphite, graphene, etc., that have various uses. The last four decades have marked a relaunch of carbon-based materials, beginning with the discovery of new nanostructures such as fullerenes (1985, with Nobel Prize for Robert Curl, Harold Kroto, and Richard Smalley in 1996), carbon nanotubes (1991), graphenes (Nobel Prize for Andre Geim in 2004 and Konstantin Novoselov in 2010), and carbon dots. The preparation of carbon nanostructures can be achieved through several strategies, two of which stand out as the most important: pyrolysis of organic precursors under an inert atmosphere, which is applicable to large-scale production but offers limited control over the carbon nanostructure; and physical/chemical vapor deposition techniques, which offer atomic-scale precision in controlling the nanostructure but require complex equipment. Carbon nanostructures have found a wide range of applications, such as in electron transport and nanoscale electronics, advanced fillers, adsorbents, active materials in energy accumulating systems (batteries), hydrogen storage systems, supercapacitors, additives for polymers, ceramics, metals and metal alloys, glasses, textiles and composite materials, filtering media, catalysts or supports for catalysts, delivery of moisture and essential elements for plants growth, theranostic platform (drug immobilization, transport, and delivery, medical imaging, etc.)

This Special Issue aims to present the latest research regarding the preparation, characterization, and application of carbon nanostructures, and intends to serve as a platform for debating and disseminating new results in this very versatile and practical research domain.

For this Special Issue, original research articles and reviews are welcome. Research areas may include (but are not limited to) carbon nanostructures and nanocomposites, energy storage, medical applications, and carbon dots.

We look forward to receiving your contributions.

Prof. Dr. Marcel Popa
Prof. Dr. Leonard Ionut Atanase
Guest Editors

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. Nanomaterials is an international peer-reviewed open access semimonthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 2900 CHF (Swiss Francs). Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Keywords

  • carbon nanostructures
  • graphene
  • fullerene
  • nanotube
  • carbon dots
  • carbon nanocomposites
  • support for catalysts
  • medical applications
  • energy storage
  • fuel storage (including hydrogen)

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Published Papers (8 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Research

Jump to: Review

25 pages, 3593 KiB  
Article
Simulations of Infrared Reflectivity and Transmission Phonon Spectra for Undoped and Doped GeC/Si (001)
by Devki N. Talwar and Jason T. Haraldsen
Nanomaterials 2024, 14(17), 1439; https://doi.org/10.3390/nano14171439 - 3 Sep 2024
Viewed by 459
Abstract
Exploring the phonon characteristics of novel group-IV binary XC (X = Si, Ge, Sn) carbides and their polymorphs has recently gained considerable scientific/technological interest as promising alternatives to Si for high-temperature, high-power, optoelectronic, gas-sensing, and photovoltaic applications. Historically, the effects of phonons on [...] Read more.
Exploring the phonon characteristics of novel group-IV binary XC (X = Si, Ge, Sn) carbides and their polymorphs has recently gained considerable scientific/technological interest as promising alternatives to Si for high-temperature, high-power, optoelectronic, gas-sensing, and photovoltaic applications. Historically, the effects of phonons on materials were considered to be a hindrance. However, modern research has confirmed that the coupling of phonons in solids initiates excitations, causing several impacts on their thermal, dielectric, and electronic properties. These studies have motivated many scientists to design low-dimensional heterostructures and investigate their lattice dynamical properties. Proper simulation/characterization of phonons in XC materials and ultrathin epilayers has been challenging. Achieving the high crystalline quality of heteroepitaxial multilayer films on different substrates with flat surfaces, intra-wafer, and wafer-to-wafer uniformity is not only inspiring but crucial for their use as functional components to boost the performance of different nano-optoelectronic devices. Despite many efforts in growing strained zinc-blende (zb) GeC/Si (001) epifilms, no IR measurements exist to monitor the effects of surface roughness on spectral interference fringes. Here, we emphasize the importance of infrared reflectivity Rω  and transmission Tω spectroscopy at near normal θi = 0 and oblique θi ≠ 0 incidence (Berreman effect) for comprehending the phonon characteristics of both undoped and doped GeC/Si (001) epilayers. Methodical simulations of Rω and Tω revealing atypical fringe contrasts in ultrathin GeC/Si are linked to the conducting transition layer and/or surface roughness. This research provided strong perspectives that the Berreman effect can complement Raman scattering spectroscopy for allowing the identification of longitudinal optical ωLO phonons, transverse optical ωTO phonons, and LO-phonon–plasmon coupled ωLPP+  modes, respectively. Full article
(This article belongs to the Special Issue Carbon Nanostructures as Promising Future Materials: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Sketch of a three-phase ideal model (‘air/epifilm/substrate’) with dielectric functions 1 air <math display="inline"><semantics> <mrow> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (air), 2 <math display="inline"><semantics> <mrow> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mn>2</mn> </msub> <mo>=</mo> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mrow> <mi>tf</mi> </mrow> </msub> </mrow> </semantics></math> (zb GeC thin film), and 3 <math display="inline"><semantics> <mrow> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mn>3</mn> </msub> <mo>=</mo> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mi mathvariant="normal">s</mi> </msub> </mrow> </semantics></math> (Si substrate) for studying the reflectance/transmission spectra of thin zb GeC/Si (001) films grown on a substrate. The modified model with the dielectric functions 1 air <math display="inline"><semantics> <mrow> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (air), 2<math display="inline"><semantics> <mrow> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="true">∼</mo> </mover> </mrow> <mn>2</mn> </msub> <mo>=</mo> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mrow> <mi>tf</mi> </mrow> </msub> </mrow> </semantics></math> (thin film) transition layer 2′ <math display="inline"><semantics> <mrow> <mrow> <msup> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> <mo>′</mo> </msup> </mrow> <mo>=</mo> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mrow> <mi>tl</mi> </mrow> </msub> </mrow> </semantics></math>, and 3 <math display="inline"><semantics> <mrow> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mn>3</mn> </msub> <mo>=</mo> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mi mathvariant="normal">s</mi> </msub> </mrow> </semantics></math> (substrate). Scattering factors χ and χ<sub>2</sub> due to roughness between GeC/air and GeC//TL surface (see Equations (8b,c)) are also included for studying the reflectivity and transmission spectra of thin films grown on a substrate [see text].</p>
Full article ">Figure 2
<p>Calculated reflectance spectra at near-normal incidence for semi-infinite n-type zb GeC. The blue and red lines reflect the spectra for undoped η = 0 and n-doped with η = 0.5 E+19 cm<sup>−3</sup>, respectively. The positions of <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">ω</mi> <mrow> <mi>TO</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">ω</mi> <mrow> <mi>LO</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>modes of zb GeC are also marked (see text).</p>
Full article ">Figure 3
<p>(<b>a</b>) Calculated infrared reflectance spectra at near-normal incidence θ<sub>i</sub> ≈ 0 for the GeC/Si (001) epilayers of different film thicknesses. The results include bulk zb GeC as well as 8 μm, 6 μm, 4 μm, 2 μm, 1 μm, 0.5 μm, and 0.05 μm thick films. (<b>b</b>) Reflectivity spectra of 4 μm thick GeC/Si (001) epifilm, with blue- and red-colored lines indicating undoped η = 0 and n-doped η = 0.5 E+19 cm<sup>−3</sup>, respectively. (<b>c</b>) Polarization-dependent reflectivity of 0.5 μm thick GeC/Si (001) epifilm at oblique incident angle θ<sub>i</sub> = 45°, where blue- and red-colored lines indicate s- and p-polarization spectra. The positions of <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">ω</mi> <mrow> <mi>TO</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">ω</mi> <mrow> <mi>LO</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>modes of GeC are also marked (see text).</p>
Full article ">Figure 4
<p>(<b>a</b>) Calculated infrared transmission spectra at near-normal incidence θ<sub>i</sub> ≈ 0 for the GeC/Si (001) epilayers of different film thicknesses. The results include 8 μm, 6 μm, 4 μm, 2 μm, 1 μm, 0.5 μm, and 0.05 μm thick films. (<b>b</b>) Transmission spectra of 4 μm thick GeC/Si (001) epifilm, with blue- and red-colored lines indicating undoped η = 0 and n-doped η = 0.5 E+19 cm<sup>−3</sup>, respectively. (<b>c</b>) Polarization-dependent transmission spectra of 0.5 μm thick GeC/Si (001) epifilm at oblique incident angle θ<sub>i</sub> = 45°, where blue- and red-colored lines indicate s- and p-polarization spectra. The positions of <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">ω</mi> <mrow> <mi>TO</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">ω</mi> <mrow> <mi>LO</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>modes of GeC are also marked (see text).</p>
Full article ">Figure 5
<p>Calculated LO-phonon–plasmon coupled <math display="inline"><semantics> <mrow> <msubsup> <mi mathvariant="sans-serif">ω</mi> <mrow> <mrow> <mi>LPP</mi> <mo> </mo> </mrow> </mrow> <mo>±</mo> </msubsup> </mrow> </semantics></math>mode frequencies in n-type GeC as a function of free carrier concentration η. The values of ω<sub>LO</sub>, ω<sub>TO</sub> modes (dotted lines) of GeC are indicated by sky blue arrows. Variation in ω<sub>P</sub> (sky blue line) with η is also displayed (see text).</p>
Full article ">Figure 6
<p>(<b>a</b>) Calculated plasma frequency <span class="html-italic">ω</span><sub>P</sub> in cm<sup>−1</sup> versus charge carrier concentration <span class="html-italic">η</span> (cm<sup>−3</sup>) in n-type GeC. (<b>b</b>) Calculated low field mobility μ in (cm<sup>2</sup>/Vs) (left) and plasmon coupling coefficient γ in cm<sup>−1</sup> versus charge carrier concentration <span class="html-italic">η</span> (cm<sup>−3</sup>) in n-type GeC (see text).</p>
Full article ">Figure 7
<p>Calculated infrared spectrum of 5.0 μm thick n-type GeC/Si(100) epifilm: (<b>a</b>) Reflectivity spectra as a function of frequency (cm<sup>−1</sup>) for a fixed value of γ = 150 cm<sup>−1</sup> while changing ω<sub>P</sub> from 300, 500, 700, 900, and 1200 cm<sup>−1</sup>. (<b>b</b>) Same key as for (<b>a</b>) but for the transmission spectra of 5.0 μm thick n-type GeC/Si(100) epifilm. (<b>c</b>) Reflectivity spectra as a function of frequency (cm<sup>−1</sup>) for a fixed value of ω<sub>P</sub> = 1000 cm<sup>−1</sup> while changing γ from 100, 200, 300, 400, and 500 cm<sup>−1</sup>. (<b>d</b>) Same key as for (<b>c</b>) but for the transmission spectra of 5.0 μm thick n-type GeC/Si(100) epifilm (see text).</p>
Full article ">Figure 8
<p>(<b>a</b>) Calculated infrared reflectivity spectra at oblique incidence (θ<sub>i</sub> = 45°) for n-type GeC/Si (001) 1.0 μm thick film in the s- and p-polarization (different colors). The charge carrier concentration η increased from 6.2 E+18 cm<sup>−3</sup> → 1.1 E+19 cm<sup>−3</sup> → 1.7 E+19 cm<sup>−3</sup> → 2.5 E+19 cm<sup>−3</sup>, respectively. The calculated shifts of <math display="inline"><semantics> <mrow> <msubsup> <mi mathvariant="sans-serif">ω</mi> <mrow> <mrow> <mi>LPP</mi> <mo> </mo> </mrow> </mrow> <mo>+</mo> </msubsup> </mrow> </semantics></math> modes in the p-polarization spectra of GeC/Si are shown by the magenta-colored vertical arrows (see text). (<b>b</b>) Same key as for (<b>a</b>) but for the simulated transmission spectra of 1.0 μm thick epifilm with different charge carrier concentrations η.</p>
Full article ">Figure 9
<p>(<b>a</b>) Calculated reflectance at near-normal incidence for a 4 μm thick GeC/Si (100) epifilm (<math display="inline"><semantics> <mi mathvariant="sans-serif">η</mi> </semantics></math>~1.01 E+17 cm<sup>−3</sup>) with different air/film surface roughnesses δ (≡0.05 μm, 0.10 μm, and 0.15 μm). (<b>b</b>) Same key as for (<b>a</b>) but for different film/substrate interface roughnesses δ<sub>2</sub> (≡0.10 μm, 0.15 μm, 0.20 μm, and 0.25 μm) (see text).</p>
Full article ">Figure 10
<p>(<b>a</b>) Calculated reflectance spectra at near normal incidence (θ<sub>i</sub> ≈ 0) for a 4 μm thick GeC/Si (100) epifilm (<math display="inline"><semantics> <mi mathvariant="sans-serif">η</mi> </semantics></math>~1.01 E+17 cm<sup>−3</sup>) for a fixed value of transition layer thickness<math display="inline"><semantics> <mrow> <msub> <mrow> <mrow> <mo> </mo> <mi mathvariant="normal">d</mi> </mrow> </mrow> <mn>2</mn> </msub> <mo stretchy="false">(</mo> <mo>≡</mo> <mo> </mo> </mrow> </semantics></math>0.05 μm) and varying air/film surface roughness δ (≡ 0.10 μm, 0.15 μm, 0.20 μm, and 0.25 μm). (<b>b</b>) Same key as for (<b>a</b>) with a fixed value of transition layer thickness<math display="inline"><semantics> <mrow> <msub> <mrow> <mrow> <mo> </mo> <mi mathvariant="normal">d</mi> </mrow> </mrow> <mn>2</mn> </msub> <mo stretchy="false">(</mo> <mo>≡</mo> <mo> </mo> </mrow> </semantics></math>0.05 μm) and varying film/substrate interface roughness δ<sub>2</sub> (≡ 0.10 μm, 0.15 μm, 0.20 μm, and 0.25 μm) (see text).</p>
Full article ">
11 pages, 3998 KiB  
Article
Flexible Mechanical Sensors Fabricated with Graphene Oxide-Coated Commercial Silk
by Hyun-Seok Jang, Ki Hoon Lee and Byung Hoon Kim
Nanomaterials 2024, 14(12), 1000; https://doi.org/10.3390/nano14121000 - 8 Jun 2024
Viewed by 825
Abstract
Many studies on flexible strain and pressure sensors have been reported due to growing interest in wearable devices for healthcare purposes. Here, we present flexible pressure and strain (motion) sensors prepared with only graphene oxide (GO) and commercial silk fabrics and yarns. The [...] Read more.
Many studies on flexible strain and pressure sensors have been reported due to growing interest in wearable devices for healthcare purposes. Here, we present flexible pressure and strain (motion) sensors prepared with only graphene oxide (GO) and commercial silk fabrics and yarns. The pressure sensors were fabricated by simply dipping the silk fabric into GO solution followed by applying a thermal treatment at 400 °C to obtain reduced GO (rGO). The pressure sensors were made from rGO-coated fabrics, which were stacked in three, five, and seven layers. A super-sensitivity of 2.58 × 103 kPa−1 at low pressure was observed in the seven-layer pressure sensor. The strain sensors were obtained from rGO-coated twisted silk yarns whose gauge factor was 0.307. Although this value is small or comparable to the values for other sensors, it is appropriate for motion sensing. The results of this study show a cost-effective and simple method for the fabrication of pressure and motion sensors with commercial silk and GO. Full article
(This article belongs to the Special Issue Carbon Nanostructures as Promising Future Materials: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) AFM image and (<b>b</b>) the sizes of the GO particles. Optical images of (<b>c</b>) GOS and (<b>d</b>) rGOS fabrics. SEM images of (<b>e</b>) GOS and (<b>f</b>) rGOS fabrics. The inset in (<b>e</b>) shows the Raman D and G peaks of GO, GOS, and rGOS. The inset in (<b>f</b>) shows that GO samples were well coated onto the surface of the silk.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD patterns of the GO, GOS, and rGOS. XPS C1s peaks of (<b>b</b>) GO, (<b>c</b>) GOS, and (<b>d</b>) rGOS. A decrease in the number of oxygen functional groups was observed in the rGOS.</p>
Full article ">Figure 3
<p>Side−views of (<b>a</b>) a single rGOS fabric and (<b>b</b>) a 7−layer rGOS fabric. The sensitivities of the rGOS pressure sensors fabricated with (<b>c</b>) 3−layer, (<b>d</b>) 5−layer, and (<b>e</b>) 7−layer fabrics. (<b>f</b>) The operating mechanism of the pressure sensor. The yellow circles are charge carriers. (<b>g</b>) A schematic illustration of the textile−based pressure sensor module and (<b>h</b>) an optical image of the two−by−two pressure sensor module.</p>
Full article ">Figure 4
<p>Real-time pressure sensing of each sensor in the module depicted using the Labview program.</p>
Full article ">Figure 5
<p>(<b>a</b>) Repeatability of the 7−layer pressure sensor during 1000 cycles and near (<b>b</b>) the first cycles and (<b>c</b>) 1000 cycles.</p>
Full article ">Figure 6
<p>(<b>a</b>) Optical image of the twisted rGOS yarn. (<b>b</b>) The variation in force during the twisting of the yarn. (<b>c</b>) SEM image of the twisted rGOS yarn.</p>
Full article ">Figure 7
<p>(<b>a</b>) Optical images of the application of strain to the twisted rGOS yarn (white boxes). The yarn was maximally stretched by 162.5%. (<b>b</b>) The strain-dependent <span class="html-italic">R</span> and (<b>c</b>) Δ<span class="html-italic">R</span>/<span class="html-italic">R</span><sub>0</sub> (variation); here, Δ<span class="html-italic">R</span> = <span class="html-italic">R</span> − <span class="html-italic">R</span><sub>0</sub>, and R<sub>0</sub> is the initial R (when the strain is zero). The negative value means that the <span class="html-italic">R</span> decreased as the strain increased. (<b>d</b>) Schematic illustration of the decrease in <span class="html-italic">R</span> due to strain. Yellow arrows are the electrically conductive paths.</p>
Full article ">Figure 8
<p>(<b>a</b>) Test of the motion sensor using a human finger and (<b>b</b>) its result. The current increases as the finger bends. (<b>c</b>) An artificial hand fabricated using a 3D printer. (<b>d</b>) Cyclability of the yarn during 5000 cycles and near (<b>e</b>) 2500 cycles and (<b>f</b>) 5000 cycles.</p>
Full article ">
0 pages, 4537 KiB  
Article
Photoluminescence of Argan-Waste-Derived Carbon Nanodots Embedded in Polymer Matrices
by Corneliu S. Stan, Noumane Elouakassi, Cristina Albu, Conchi O. Ania, Adina Coroaba, Laura E. Ursu, Marcel Popa, Hamid Kaddami and Abdemaji Almaggoussi
Nanomaterials 2024, 14(1), 83; https://doi.org/10.3390/nano14010083 - 27 Dec 2023
Cited by 1 | Viewed by 1274
Abstract
In this work, photoluminescent (PL) carbon nano dots (CNDs) prepared from argan waste were embedded in highly optical transparent poly(styrene-co-acrylonitrile) (PSA) and cyclo-olefin copolymer (COC) matrices, which were further processed into thin films. In the first step, the luminescent CNDs were prepared through [...] Read more.
In this work, photoluminescent (PL) carbon nano dots (CNDs) prepared from argan waste were embedded in highly optical transparent poly(styrene-co-acrylonitrile) (PSA) and cyclo-olefin copolymer (COC) matrices, which were further processed into thin films. In the first step, the luminescent CNDs were prepared through thermal processing of fine-groundargan waste, followed, in the second step, by direct dispersion in the polymer solutions, obtained by solving PSA and COC in selected solvents. These two polymer matrices were selected due to their high optical transparency, resilience to various environmental factors, and ability to be processed as quality thin films. The structural configuration of the CNDs was investigated through EDX, XPS, and FTIR, while DLS, HR-SEM, and STEM were used for their morphology investigation. The luminescence of the prepared CNDs and resulted polymer nanocomposites was thoroughly investigated through steady-state, absolute PLQY, and lifetime fluorescence. The quality of the resulted CND–polymer nanocomposite thin films was evaluated through AFM. The prepared highly luminescent thin films with a PL conversion efficiency of 30% are intended to be applied as outer photonic conversion layers on solar PV cells for increasing their conversion efficiency through valorization of the UV component of the solar radiation. Full article
(This article belongs to the Special Issue Carbon Nanostructures as Promising Future Materials: 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Prepared argan-derived-CNDs dispersed in water and a thick film of CND/COC nanocomposite under UV-A excitation.</p>
Full article ">Figure 2
<p>Record FT−IR spectra of the (<b>a</b>) argan cake waste and (<b>b</b>) derived CNDs.</p>
Full article ">Figure 3
<p>Average hydrodynamic diameter distribution of CND particles dispersed in water and chloroform, obtained by dynamic light scattering.</p>
Full article ">Figure 4
<p>HR-SEM micrographs of the CNDs dispersed in (<b>a</b>) water and (<b>b</b>) chloroform.</p>
Full article ">Figure 5
<p>TEM micrographs of the CNDs recorded at (<b>a</b>) 230 k and (<b>b</b>) 405 k magnifications.</p>
Full article ">Figure 6
<p>AFM images of the CND/polymer nanocomposites: (<b>a</b>) CND/COC and (<b>b</b>) CND/PSA.</p>
Full article ">Figure 7
<p>Steady-state PL emission of the argan-prepared CNDs dispersed in (<b>a</b>) water, (<b>b</b>) chloroform, and (<b>c</b>) tetrahydrofuran.</p>
Full article ">Figure 8
<p>Steady-state PL emission of the (<b>a</b>) COC/CND and (<b>b</b>) PSA/CND nanocomposites.</p>
Full article ">Figure 9
<p>PL emission chromaticity parameters (CIE1931) of the (<b>a</b>) solvent-dispersed CNDs and the (<b>b</b>) CND/polymer nanocomposites.</p>
Full article ">Figure 10
<p>PL lifetime decay recorded for the CLF-dispersed CNDs and COC–CND nanocomposites.</p>
Full article ">Figure 11
<p>PV cell provided with the COC–CND nanocomposite outer thin layer: (<b>a</b>) under normal illumination conditions and (<b>b</b>) under UV-A excitation.</p>
Full article ">
17 pages, 1311 KiB  
Article
Density Functional Theory for Buckyballs within Symmetrized Icosahedral Basis
by Chung-Yuan Ren, Raj Kumar Paudel and Yia-Chung Chang
Nanomaterials 2023, 13(13), 1912; https://doi.org/10.3390/nano13131912 - 23 Jun 2023
Cited by 1 | Viewed by 1718
Abstract
We have developed a highly efficient computation method based on density functional theory (DFT) within a set of fully symmetrized basis functions for the C60 buckyball, which possesses the icosahedral (Ih) point-group symmetry with 120 symmetry operations. We demonstrate [...] Read more.
We have developed a highly efficient computation method based on density functional theory (DFT) within a set of fully symmetrized basis functions for the C60 buckyball, which possesses the icosahedral (Ih) point-group symmetry with 120 symmetry operations. We demonstrate that our approach is much more efficient than the conventional approach based on three-dimensional plane waves. When applied to the calculation of optical transitions, our method is more than one order of magnitude faster than the existing DFT package with a conventional plane-wave basis. This makes it very convenient for modeling optical and transport properties of quantum devices related to buckyball crystals. The method introduced here can be easily extended to other fullerene-like materials. Full article
(This article belongs to the Special Issue Carbon Nanostructures as Promising Future Materials: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of buckyball [<a href="#B25-nanomaterials-13-01912" class="html-bibr">25</a>].</p>
Full article ">Figure 2
<p>Imaginary part of dielectric response function, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> <mfenced separators="|"> <mrow> <mo>ħ</mo> <mi>ω</mi> </mrow> </mfenced> </mrow> </semantics></math> of the C<sub>60</sub> buckyball calculated by the current method. The broadening parameter (<span class="html-italic">Γ</span>) used is 0.005 eV.</p>
Full article ">
17 pages, 5327 KiB  
Article
Nanoporous Hollow Carbon Spheres Derived from Fullerene Assembly as Electrode Materials for High-Performance Supercapacitors
by Lok Kumar Shrestha, Zexuan Wei, Gokulnath Subramaniam, Rekha Goswami Shrestha, Ravi Singh, Marappan Sathish, Renzhi Ma, Jonathan P. Hill, Junji Nakamura and Katsuhiko Ariga
Nanomaterials 2023, 13(5), 946; https://doi.org/10.3390/nano13050946 - 5 Mar 2023
Cited by 6 | Viewed by 4763
Abstract
The energy storage performances of supercapacitors are expected to be enhanced by the use of nanostructured hierarchically micro/mesoporous hollow carbon materials based on their ultra-high specific surface areas and rapid diffusion of electrolyte ions through the interconnected channels of their mesoporous structures. In [...] Read more.
The energy storage performances of supercapacitors are expected to be enhanced by the use of nanostructured hierarchically micro/mesoporous hollow carbon materials based on their ultra-high specific surface areas and rapid diffusion of electrolyte ions through the interconnected channels of their mesoporous structures. In this work, we report the electrochemical supercapacitance properties of hollow carbon spheres prepared by high-temperature carbonization of self-assembled fullerene-ethylenediamine hollow spheres (FE-HS). FE-HS, having an average external diameter of 290 nm, an internal diameter of 65 nm, and a wall thickness of 225 nm, were prepared by using the dynamic liquid-liquid interfacial precipitation (DLLIP) method at ambient conditions of temperature and pressure. High temperature carbonization (at 700, 900, and 1100 °C) of the FE-HS yielded nanoporous (micro/mesoporous) hollow carbon spheres with large surface areas (612 to 1616 m2 g−1) and large pore volumes (0.925 to 1.346 cm3 g−1) dependent on the temperature applied. The sample obtained by carbonization of FE-HS at 900 °C (FE-HS_900) displayed optimum surface area and exhibited remarkable electrochemical electrical double-layer capacitance properties in aq. 1 M sulfuric acid due to its well-developed porosity, interconnected pore structure, and large surface area. For a three-electrode cell setup, a specific capacitance of 293 F g−1 at a 1 A g−1 current density, which is approximately 4 times greater than the specific capacitance of the starting material, FE-HS. The symmetric supercapacitor cell was assembled using FE-HS_900 and attained 164 F g−1 at 1 A g−1 with sustained 50% capacitance at 10 A g−1 accompanied by 96% cycle life and 98% coulombic efficiency after 10,000 consecutive charge/discharge cycles. The results demonstrate the excellent potential of these fullerene assemblies in the fabrication of nanoporous carbon materials with the extensive surface areas required for high-performance energy storage supercapacitor applications. Full article
(This article belongs to the Special Issue Carbon Nanostructures as Promising Future Materials: 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Electron microscopy observations of the hollow carbon spheres obtained by high temperature carbonization of FE-HS at 900 °C. (<b>a</b>,<b>b</b>) SEM images; (<b>c</b>) TEM image; (<b>d</b>) HR-TEM image; and (<b>e</b>,<b>f</b>) histograms of the outer diameter (<b>e</b>) and inner diameter (<b>f</b>) the FE-HS_900.</p>
Full article ">Figure 2
<p>(<b>a</b>) pXRD patterns of pristine C<sub>60</sub> (pC<sub>60</sub>), FE-HS, FE-HS_700, FE-HS_900; and the corresponding Raman scattering spectra (<b>b</b>), and FTIR spectra (<b>c</b>).</p>
Full article ">Figure 3
<p>X-ray photoelectron spectra of the materials. (<b>a</b>) XPS survey spectra of pC<sub>60</sub>, FE-HS, FE-HS_700, FE-HS_900, and FE-HS_1100; (<b>b</b>) XPS C 1s core level spectra with the deconvoluted peaks; (<b>c</b>) XPS O 1s spectra with the deconvoluted peaks; and (<b>d</b>) XPS N 1s spectra with the deconvoluted peaks.</p>
Full article ">Figure 4
<p>(<b>a</b>) Nitrogen adsorption isotherms FE-HS, FE-HS_700, FE-HS_900, and FE-HS_1100; (<b>b</b>) corresponding pore size distribution profile obtained by the density functional theory (DFT) method; and (<b>c</b>) pore size distribution and the Barrett–Joyner–Halenda (BJH) model.</p>
Full article ">Figure 5
<p>(<b>a</b>) CV curves of FE-HS, FE-HS_700, FE-HS_900, and FE-HS_1100 recorded at a scan rate of 5 mV s<sup>−1</sup> in the three-electrode cell setup; (<b>b</b>) CV curves vs. scan rates for FE-HS_700; (<b>c</b>) CV curves vs. scan rates for FE-HS_900; (<b>d</b>) CV curves vs. scan rates for FE-HS_1100.</p>
Full article ">Figure 6
<p>(<b>a</b>) GCD profiles of FE-HS, FE-HS_700, FE-HS_900, and FE-HS_1100 measured at a current density of 1 A g<sup>−1</sup>; GCD curves at different current densities (1 to 20 A g<sup>−1</sup>) for (<b>b</b>) FE-HS; (<b>c</b>) FE-HS_700; (<b>d</b>) FE-HS_900; and (<b>e</b>) FE-HS_1100; and (<b>f</b>) calculated specific capacitance from the GCD curves against current density for all the samples.</p>
Full article ">Figure 7
<p>Electrochemical performance of the assembled symmetric supercapacitors. (<b>a</b>) Comparison of the CV curves recorded at a fixed scan rate of 5 mV s<sup>−1</sup> for FE-HS_900 and FE-HS_1100; (<b>b</b>) CV curves vs. scan rate for FE-HS_900; (<b>c</b>) GCD curves of FE-HS_900 and FE-HS_1100 at 1 A g<sup>−1</sup>; GCD profiles vs. current density for (<b>d</b>) FE-HS_900; and (<b>e</b>) FE-HS_1100; (<b>f</b>) calculated specific capacitance vs. current density; (<b>g</b>) cycle performance with coulombic efficiency; (<b>h</b>) Nyquist plots for FE-HS_900 and FE-HS_1100; and (<b>i</b>) Ragone plot of the FE-HS_900 and FE-HS_1100 supercapacitors.</p>
Full article ">Scheme 1
<p>Preparation of self-assembled fullerene-ethylenediamine hollow spheres (FE-HS) and their direct conversion to hierarchically porous carbon spheres. (<b>a</b>) Schematic of the DLLIP method; (<b>b</b>) transmission electron microscopy (TEM) image of FE-HS with a representative model in the inset; (<b>c</b>) TEM image of the resulting hollow carbon spheres with a model structure in the inset; and (<b>d</b>) high-resolution TEM (HR-TEM) image of the hollow carbon spheres showing abundant micro/mesopores.</p>
Full article ">
11 pages, 2902 KiB  
Communication
Reversible Hydrogen Storage Media by g-CN Monolayer Decorated with NLi4: A First-Principles Study
by Xihao Chen, Wenjie Hou, Fuqiang Zhai, Jiang Cheng, Shuang Yuan, Yihan Li, Ning Wang, Liang Zhang and Jie Ren
Nanomaterials 2023, 13(4), 647; https://doi.org/10.3390/nano13040647 - 7 Feb 2023
Cited by 7 | Viewed by 1690
Abstract
A two-dimensional graphene-like carbon nitride (g-CN) monolayer decorated with the superatomic cluster NLi4 was studied for reversible hydrogen storage by first-principles calculations. Molecular dynamics simulations show that the g-CN monolayer has good thermal stability at room temperature. The NLi4 is firmly [...] Read more.
A two-dimensional graphene-like carbon nitride (g-CN) monolayer decorated with the superatomic cluster NLi4 was studied for reversible hydrogen storage by first-principles calculations. Molecular dynamics simulations show that the g-CN monolayer has good thermal stability at room temperature. The NLi4 is firmly anchored on the g-CN monolayer with a binding energy of −6.35 eV. Electronic charges are transferred from the Li atoms of NLi4 to the g-CN monolayer, mainly due to the hybridization of Li(2s), C(2p), and N(2p) orbitals. Consequently, a spatial local electrostatic field is formed around NLi4, leading to polarization of the adsorbed hydrogen molecules and further enhancing the electrostatic interactions between the Li atoms and hydrogen. Each NLi4 can adsorb nine hydrogen molecules with average adsorption energies between −0.152 eV/H2 and −0.237 eV/H2. This range is within the reversible hydrogen storage energy window. Moreover, the highest achieved gravimetric capacity is up to 9.2 wt%, which is superior to the 5.5 wt% target set by the U.S. Department of Energy. This study shows that g-CN monolayers decorated with NLi4 are a good candidate for reversible hydrogen storage. Full article
(This article belongs to the Special Issue Carbon Nanostructures as Promising Future Materials: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Top view and (<b>b</b>) side view of the optimized structure of g-CN monolayer. Red and black balls represent N and C atoms, respectively.</p>
Full article ">Figure 2
<p>The PDOS of pure g-CN monolayer. The PDOS is shown for carbon and nitrogen atoms. Energy calculation was conducted with reference to Fermi energy level.</p>
Full article ">Figure 3
<p>The first-principles MD study about g-CN monolayer. Temperature (<b>red</b>) and energy (<b>blue</b>) under room temperature (300 k) against time. The time step is 0.5 fs and the total testing is 4 ps, including 8000 steps.</p>
Full article ">Figure 4
<p>The top view (<b>a</b>) and the side view (<b>b</b>) of the optimized structure of NLi<sub>4</sub> decorated on the g-CN monolayer. Red, black, and blue balls are the symbols for N, C, and Li atoms, respectively.</p>
Full article ">Figure 5
<p>The PDOS of N, C, and Li atoms of NLi<sub>4</sub> decorated on the g-CN monolayer. Energy calculation was conducted with reference to Fermi energy level.</p>
Full article ">Figure 6
<p>The top view (<b>a</b>) and the side view (<b>b</b>) of the charge density difference for the NLi<sub>4</sub> on the g-CN monolayer. Bule and yellow region mean charge loss and gain. The isosurface of charge density is set to 0.0012 (number of charge/Bohr<sup>3</sup>).</p>
Full article ">Figure 7
<p>(<b>a</b>–<b>f</b>) The top and side views of processing of multiple H<sub>2</sub> absorption on optimized NLi<sub>4</sub> decorated g-CN monolayer in the increasing order.</p>
Full article ">Figure 8
<p>(<b>a</b>) Top view and (<b>b</b>) side view of charge density difference for the adsorbed H<sub>2</sub> molecules on NLi<sub>4</sub>-decorated g-CN monolayer, where blue and yellow are charge loss and charge gain region, respectively. The isosurface is set to 0.005 (number of charge/Bohr<sup>3</sup>).</p>
Full article ">

Review

Jump to: Research

24 pages, 2408 KiB  
Review
Carbon Dots in Photodynamic/Photothermal Antimicrobial Therapy
by Siqi Wang, Colin P. McCoy, Peifeng Li, Yining Li, Yinghan Zhao, Gavin P. Andrews, Matthew P. Wylie and Yi Ge
Nanomaterials 2024, 14(15), 1250; https://doi.org/10.3390/nano14151250 - 25 Jul 2024
Viewed by 1100
Abstract
Antimicrobial resistance (AMR) presents an escalating global challenge as conventional antibiotic treatments become less effective. In response, photodynamic therapy (PDT) and photothermal therapy (PTT) have emerged as promising alternatives. While rooted in ancient practices, these methods have evolved with modern innovations, particularly through [...] Read more.
Antimicrobial resistance (AMR) presents an escalating global challenge as conventional antibiotic treatments become less effective. In response, photodynamic therapy (PDT) and photothermal therapy (PTT) have emerged as promising alternatives. While rooted in ancient practices, these methods have evolved with modern innovations, particularly through the integration of lasers, refining their efficacy. PDT harnesses photosensitizers to generate reactive oxygen species (ROS), which are detrimental to microbial cells, whereas PTT relies on heat to induce cellular damage. The key to their effectiveness lies in the utilization of photosensitizers, especially when integrated into nano- or micron-scale supports, which amplify ROS production and enhance antimicrobial activity. Over the last decade, carbon dots (CDs) have emerged as a highly promising nanomaterial, attracting increasing attention owing to their distinctive properties and versatile applications, including PDT and PTT. They can not only function as photosensitizers, but also synergistically combine with other photosensitizers to enhance overall efficacy. This review explores the recent advancements in CDs, underscoring their significance and potential in reshaping advanced antimicrobial therapeutics. Full article
(This article belongs to the Special Issue Carbon Nanostructures as Promising Future Materials: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of three types of mechanisms in photodynamic therapy (PDT) including a series of potential energy transfer processes that occur in photosensitizers after light excitation (Jablonski diagram), along with possible photodynamic therapy mechanisms (with (Type I and Type II) or without oxygen (Type III)) that may take place during this process. The Jablonski diagram illustrates the electronic states of a molecule, including the ground state (S<sub>0</sub>), singlet excited states (S<sub>1</sub>, S<sub>2</sub>, S<sub>n</sub>), and triplet states (T<sub>1</sub>). Upon absorption of a photon, electrons are excited from S<sub>0</sub> to higher singlet states (S<sub>1</sub>, S<sub>2</sub>, Sn). Fluorescence occurs when electrons return from an excited singlet state (S<sub>1</sub>) to the ground state (S<sub>0</sub>). Intersystem crossing (ISC) is a non-radiative transition from a singlet state (S<sub>1</sub>) to a triplet state (T<sub>1</sub>). Phosphorescence involves the emission of a photon when electrons return from a triplet state (T<sub>1</sub>) to the ground state (S<sub>0</sub>). Internal conversion (IC) represents non-radiative transitions between singlet states (e.g., S<sub>n</sub> to S<sub>1</sub>) [<a href="#B40-nanomaterials-14-01250" class="html-bibr">40</a>,<a href="#B41-nanomaterials-14-01250" class="html-bibr">41</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic representation of the synthesis of citric acid-derived carbon dots (C-DOTS) and in vivo and in vitro antibacterial photodynamic therapeutic studies. Photoexcited C-DOTS reduced the number of bacterial colonies (log CFU/mL) based on the light dose 450 nm, 40 mW/cm<sup>2</sup> delivered [<a href="#B135-nanomaterials-14-01250" class="html-bibr">135</a>]. (<b>b</b>) Synthesis of red-carbon dots (R-CDs) by solvothermal method and its dual effect of intrinsically antibacterial and photodynamic antibacterial effect against MRAB and biofilms [<a href="#B140-nanomaterials-14-01250" class="html-bibr">140</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) Photodynamic efficiency of amine-functionalized CDs (CDs-NH<sub>2</sub>) for the inactivation of <span class="html-italic">E. coli</span> upon irradiation at 0.3 W for 10 min and 20 min [<a href="#B143-nanomaterials-14-01250" class="html-bibr">143</a>]. (<b>b</b>) Influence of the CDs-NH<sub>2</sub> and CDs-AMP concentration on the treatment efficiency of <span class="html-italic">E. coli</span> without (solid lines) and with (dash lines) visible light illumination (20 min, 0.3 W). The error bars represent the standard deviation of three independent experiments [<a href="#B143-nanomaterials-14-01250" class="html-bibr">143</a>]. (<b>c</b>) Scheme for synthesis of BSA-coated CDs (BSA-CD) for visible-light-induced ROS generation and simultaneous release of ciprofloxacin for antibacterial activity [<a href="#B144-nanomaterials-14-01250" class="html-bibr">144</a>]. (<b>d</b>) Percentage cell survival of <span class="html-italic">S. aureus</span> in the presence of BSA-CD loaded with ciprofloxacin. Percentage cell survival of <span class="html-italic">E. coli</span> in the presence of BSA-CD loaded with ciprofloxacin [<a href="#B144-nanomaterials-14-01250" class="html-bibr">144</a>]. Data were considered statistically significant and highly significant when <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.001, respectively (**, 0.05 &lt; <span class="html-italic">p</span> &lt; 0.01; ***, 0.01 &lt; <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>(<b>a</b>) Synthetic routes of ZnPc-CQDs for PDT/PTT antibacterial effects [<a href="#B146-nanomaterials-14-01250" class="html-bibr">146</a>]. (<b>b</b>) Bacterial viability after treatment with different concentrations of ZnPc-CQDs with and without irradiation [<a href="#B146-nanomaterials-14-01250" class="html-bibr">146</a>]. (<b>c</b>) CDs-MR synthesis scheme with its cell penetration and antimicrobial photoactivation [<a href="#B147-nanomaterials-14-01250" class="html-bibr">147</a>]. (<b>d</b>) Disk-diffusion (Kirby–Bauer) tests of the antimicrobial activity of CD-MR (120 μg) on MHA plates against <span class="html-italic">C. albicans</span>, <span class="html-italic">C. neoformans</span>, and <span class="html-italic">S. aureus</span> [<a href="#B147-nanomaterials-14-01250" class="html-bibr">147</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) Fabrication of PAN-CQDs NFs and the photodynamic inactivation of bacteria upon visible light illumination [<a href="#B157-nanomaterials-14-01250" class="html-bibr">157</a>]. (<b>b</b>) Photodynamic inactivation studies employing PAN-CQDs NFs against Gram-negative bacteria <span class="html-italic">E. coli</span> and <span class="html-italic">P. aeruginosa</span> as well as Gram-positive bacteria <span class="html-italic">S. aureus</span> and <span class="html-italic">B. subtilis</span>. Displayed is the survival rate for the PAN-CQDs-2.5% NFs dark control (dark grey bar) and illuminated PAN NFs light control (red bar) conditions, blue and green bars represent the PAN-CQDs NFs (0.6 and 2.5%) against bacteria, respectively. Studies were performed with a 60 min dark pre-incubation followed by 90 min illumination [<a href="#B157-nanomaterials-14-01250" class="html-bibr">157</a>]. (<b>c</b>) Design of CS/nHA/CDs scaffolds for enhancing BMSC adhesion and differentiation, promoting vascularized new bone formation, tumor ablation, and bacterial eradication by PTT [<a href="#B158-nanomaterials-14-01250" class="html-bibr">158</a>]. (<b>d</b>) Number of clinically relevant <span class="html-italic">S. aureus</span> (top) and <span class="html-italic">E. coli</span> (bottom) bacterial colonies after bacteria from the harvested samples were cultured for 24 h after 1-week treatments in vivo. Each value is the mean ± standard deviation; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 [<a href="#B158-nanomaterials-14-01250" class="html-bibr">158</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) Synthesis of the CDs/Cur nanocomposite and bactericidal activities of CDs/Cur upon dual wavelength (405 + 808 nm) illumination [<a href="#B36-nanomaterials-14-01250" class="html-bibr">36</a>]. (<b>b</b>) Agar plate photographs of <span class="html-italic">E. coli</span> treated with Cur or CDs/Cur, respectively, were taken under (non) light irradiation conditions. The corresponding dependence of <span class="html-italic">E. coli</span> survival fraction on the concentration of Cur and CDs/Cur was measured, respectively, under non-light, 405 nm light, and 405 + 808 nm light. Values are means ± standard deviation (SD) (n = 3) [<a href="#B36-nanomaterials-14-01250" class="html-bibr">36</a>]. (<b>c</b>) Design of hybrid hydrogel derived from carbon dots (CDs), protoporphyrin IX (PpIX), and DNA [<a href="#B37-nanomaterials-14-01250" class="html-bibr">37</a>]. (<b>d</b>) Plate count assay for <span class="html-italic">S. aureus.</span> cell survival with CD-DNA-PpIX hydrogel (PpIX:1 mM) exposed to UV light [<a href="#B37-nanomaterials-14-01250" class="html-bibr">37</a>]. Each value is the mean ± standard deviation; *** <span class="html-italic">p</span> &lt; 0.001, ns: non-significant.</p>
Full article ">
25 pages, 18302 KiB  
Review
Cytotoxicity of Carbon Nanotubes, Graphene, Fullerenes, and Dots
by Marianna V. Kharlamova and Christian Kramberger
Nanomaterials 2023, 13(9), 1458; https://doi.org/10.3390/nano13091458 - 25 Apr 2023
Cited by 7 | Viewed by 2224
Abstract
The cytotoxicity of carbon nanomaterials is a very important issue for microorganisms, animals, and humans. Here, we discuss the issues of cytotoxicity of carbon nanomaterials, carbon nanotubes, graphene, fullerene, and dots. Cytotoxicity issues, such as cell viability and drug release, are considered. The [...] Read more.
The cytotoxicity of carbon nanomaterials is a very important issue for microorganisms, animals, and humans. Here, we discuss the issues of cytotoxicity of carbon nanomaterials, carbon nanotubes, graphene, fullerene, and dots. Cytotoxicity issues, such as cell viability and drug release, are considered. The main part of the review is dedicated to important cell viability issues. They are presented for A549 human melanoma, E. coli, osteosarcoma, U2-OS, SAOS-2, MG63, U87, and U118 cell lines. Then, important drug release issues are discussed. Bioimaging results are shown here to illustrate the use of carbon derivatives as markers in any type of imaging used in vivo/in vitro. Finally, perspectives of the field are presented. The important issue is single-cell viability. It can allow a correlation of the functionality of organelles of single cells with the development of cancer. Such organelles are mitochondria, nuclei, vacuoles, and reticulum. It allows for finding biochemical evidence of cancer prevention in single cells. The development of investigation methods for single-cell level detection of viability stimulates the cytotoxicity investigative field. The development of single-cell microscopy is needed to improve the resolution and accuracy of investigations. The importance of cytotoxicity is drug release. It is important to control the amount of drug that is released. This is performed with pH, temperature, and electric stimulation. Further development of drug loading and bioimaging is important to decrease the cytotoxicity of carbon nanomaterials. We hope that this review is useful for researchers from all disciplines across the world. Full article
(This article belongs to the Special Issue Carbon Nanostructures as Promising Future Materials: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) (<b>a1</b>) The SEM images of non−treated bacterial cell. (<b>a2</b>) Bacterial cells treated with GONPs. (<b>a3</b>) Bacteria cells treated with GONPs–PEG. (<b>a4</b>) Bacterial cells exposed with GONPs–PEG–N. sativa. Left panels correspond to <span class="html-italic">E. coli</span>, and right panels belong to <span class="html-italic">S. aureus</span>. Copyright 2021 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open−access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B56-nanomaterials-13-01458" class="html-bibr">56</a>]. (<b>b</b>) Low, and high−magnification HRTEM images of MWCNT/PPGP<sub>s</sub> (<b>b1</b>,<b>b2</b>) and MWCNT/PPGP<sub>c</sub> (<b>b3</b>,<b>b4</b>), and SWCNT/PPGP<sub>s</sub> (<b>b5</b>,<b>b6</b>) and SWCNT/PPGPc (<b>b7</b>,<b>b8</b>), accordingly. Copyright 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open−access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B41-nanomaterials-13-01458" class="html-bibr">41</a>]. (<b>c</b>) (1) Ultraviolet (UV)−visible absorption spectra of synthetic oxazolidinone antibiotic linezolid (LNZ), bovine serum albumin carbon dots (BCDs), and LNZ–BCDs nanocomposite. (Inset) Photos of BCDs. Copyright 2023 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B67-nanomaterials-13-01458" class="html-bibr">67</a>]. (<b>d</b>) Absorbance spectra (<b>d1</b>), and fluorescence spectra (<b>d2</b>) of functionalized with poly(ethylene imine) (PEI) reduced GO, and PEI−rGO/DOX. Copyright 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B50-nanomaterials-13-01458" class="html-bibr">50</a>]. (<b>e</b>) The Raman spectra of raw and functionalized magnetic multi−walled carbon nanotubes. Copyright 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B43-nanomaterials-13-01458" class="html-bibr">43</a>]. (<b>f</b>) FT−IR spectra of the purified reduced carbon nanotube nanocomposite Nano 6 (<b>f1</b>) and the same sample at the end of hydrocortisone release (<b>f2</b>). Copyright 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open−access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B44-nanomaterials-13-01458" class="html-bibr">44</a>]. (<b>g</b>) Z−Average sizes and zeta−potential values of (<b>g1</b>) PK<sub>5</sub>E<sub>5</sub> (PEI−rGO), (<b>g2</b>) PK<sub>5</sub>E<sub>7</sub> (PEI−rGO), (<b>g3</b>) PK<sub>5</sub>E<sub>9</sub> (PEI−rGO), and (<b>g4</b>) PK<sub>5</sub>E<sub>13</sub> (PEI−rGO) with various weight ratios (PKE: PEI−rGO). Copyright 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open−access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B50-nanomaterials-13-01458" class="html-bibr">50</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) The schematics of drug delivery system from SWCNTs, self−assembled ribbon−like structures (SRLS), Congo Red (CR) with drug doxorubicin. Copyright 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open−access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B42-nanomaterials-13-01458" class="html-bibr">42</a>]. (<b>b</b>) The structure of SWCNT−CR-DOX systems at neutral pH values: (<b>b1</b>) SWCNT (10,0), 20 CR (red) and 10 DOX (blue), (<b>b2</b>) SWCNT (30,0), 20 CR and 10 DOX, (<b>b3</b>) SWCNT (30,0), 40 CR and 10 DOX. Inset: SEM image. Copyright 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open−access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B42-nanomaterials-13-01458" class="html-bibr">42</a>]. (<b>c</b>) The structure of SWNT−CR−DOX systems at medium pH values (5.0 &lt; pH &lt; 7.4). (<b>c1</b>) SWCNT (10,0), 20 CR (red), 10 DOX (blue), (<b>c2</b>) SWCNT (30,0), 20 CR, and 10 DOX, (<b>c3</b>) SWCNT (30,0), 40 CR and 10 DOX. Copyright 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open−access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B42-nanomaterials-13-01458" class="html-bibr">42</a>]. (<b>d</b>) The schematics of the dendro [<a href="#B60-nanomaterials-13-01458" class="html-bibr">60</a>] fullerene. Copyright 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open−access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B63-nanomaterials-13-01458" class="html-bibr">63</a>]. (<b>e</b>) Theoretical studies of functionalized drug−loaded fullerenes in water. Copyright 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open−access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B63-nanomaterials-13-01458" class="html-bibr">63</a>]. (<b>f</b>) Theoretical studies of functionalized drug−loaded fullerenes in organic solvent. Copyright 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open−access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B63-nanomaterials-13-01458" class="html-bibr">63</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) (<b>a1</b>) Control; (<b>a2</b>) CNT-1; (<b>a3</b>) CNT-2 treated red algae P. purpureum. Copyright 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B7-nanomaterials-13-01458" class="html-bibr">7</a>]. (<b>b</b>) Images of gills (upper), and digestive glands (lower) in Mytilus galloprovincialis influenced by different materials stained with hematoxylin. Scale bar = 50 µm. CTL: control, RSW: seawater after remediation (at 21 °C). Copyright 2021 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B17-nanomaterials-13-01458" class="html-bibr">17</a>]. (<b>c</b>) Hematoxylin–eosin (HE) staining of lung sections exposed to MWCNTs. Copyright 2012 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B1-nanomaterials-13-01458" class="html-bibr">1</a>]. (<b>d</b>) Animals exposed to CNTs and euthanized at 0 week post-exposure as well as 2 weeks post-exposure. Copyright 2014 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B2-nanomaterials-13-01458" class="html-bibr">2</a>]. (<b>e</b>) (<b>e1</b>,<b>e2</b>) show the representative light image and NIRF image of control sample of an intestinal cross-section of fish. (<b>e3</b>,<b>e4</b>) show representative light image and NIRF image of SWCNT fed fish. Scale bar = 150 μm. Copyright 2015 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B5-nanomaterials-13-01458" class="html-bibr">5</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) The viability of U2-OS (<b>a1</b>), MG63 (<b>a2</b>), SAOS-2 cell lines (<b>a3</b>) and images of the cells cultured for 72 h in the presence of GO (<b>a4</b>). Scale bars, 100 μm. Copyright 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B60-nanomaterials-13-01458" class="html-bibr">60</a>]. (<b>b</b>) The viability of U87 (<b>b1</b>), U118 (<b>b2</b>), and images of the cells cultured for 72 h in the presence of material (<b>b3</b>). Scale bars, 100 μm. Copyright 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B60-nanomaterials-13-01458" class="html-bibr">60</a>]. (<b>c</b>) The comparison of the PELI values of different graphene samples, such as untreated graphene oxide (<b>c1</b>), UV treated graphene oxide (<b>c2</b>), thermally reduced graphene oxide (<b>c3</b>), 45 min sonicated graphene oxide (<b>c4</b>), 2.5 h sonicated graphene oxide (<b>c5</b>). The enrichment in toxicity is marked with *. Copyright 2021 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B19-nanomaterials-13-01458" class="html-bibr">19</a>]. (<b>d</b>) The cytotoxicity of graphene nanoparticles in human keratinocytes cells (HaCaT) for exposure period of 24 h and concentrations from 0 (control) to 10 µg/mL measured by the (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium) (MTT) assay. Copyright 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B21-nanomaterials-13-01458" class="html-bibr">21</a>]. * <span class="html-italic">p</span>-value &lt; 0.05, ** <span class="html-italic">p</span>-value &lt; 0.01,*** <span class="html-italic">p</span>-value &lt; 0.001, **** <span class="html-italic">p</span>-value &lt; 0.0001, α, β, Φ: the mean ± standard error of the mean, n.s: not significant.</p>
Full article ">Figure 5
<p>(<b>a</b>) The proliferative activity of graphene nanoparticles in the HaCaT cell was investigated at 0.005 and 0.01 μg/mL for 72 (<b>a1</b>) or 96 h (<b>a2</b>). The enrichment in toxicity is marked with *. Copyright 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B21-nanomaterials-13-01458" class="html-bibr">21</a>]. (<b>b</b>) Exposure of embryos (<span class="html-italic">n</span> = 32) to reduced graphene oxide revealed that for the 2 μm × 2 μm reduced graphene oxide, there was significant mortality for concentrations starting 10.7 μg/mL, whereas 400 nm × 400 nm reduced graphene oxide did not lead to significant mortality at tested concentrations. MO24 = 24 post-fertilization (hpf) mortality; MORT = 120 hpf mortality. Copyright 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B22-nanomaterials-13-01458" class="html-bibr">22</a>]. (<b>c</b>) The waste water quality parameters, biological oxygen demand (BOD), chemical oxygen demand (COD), and total organic carbon (TOC) before and after photocatalytic treatment with aerogel photocatalytic membrane based on graphene oxide (GO) with Cr–Mn-doped TiO<sub>2</sub>. Copyright 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B26-nanomaterials-13-01458" class="html-bibr">26</a>]. (<b>d</b>) The dependence of efficiency of Cr–Mn-doped TiO<sub>2</sub>@GO aerogel photocatalytic membranes on number of cycles. Copyright 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B26-nanomaterials-13-01458" class="html-bibr">26</a>]. (<b>e</b>) (<b>e1</b>,<b>e2</b>) Cell viability of TM3, and TM4 cells treated with graphene oxide samples with the average size of 20 nm (GO-20), and 100 nm (GO-100). (<b>e3</b>,<b>e4</b>) TM3, and TM4 cell morphology, respectively, under a light microscope, compared with control sample (Con), and silver nanoparticles (AgNPs) (scale bar 200 µm). The enrichment in toxicity is marked with *. Copyright 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B155-nanomaterials-13-01458" class="html-bibr">155</a>]. (<b>f</b>) TEM images of cellular internalization of reduced graphene oxide (rGO) and graphene oxide (GO) in Caco-2 cells. (<b>f1</b>,<b>f4</b>) untereated control sample. (<b>f2</b>,<b>f3</b>) Caco-2 treated with rGO after 24 h and 48 h, respectively. (<b>f5</b>,<b>f6</b>) Caco-2 treated with GO after 24 h and 48 h, respectively (scale bar: 2 μm). Signs are h—heterophagosome; N—nucleus; <span class="html-italic">n</span>—nucleolus; *—mitochondria; ▴—endoplasmic reticulum; ì—dense bodies; apoptotic bodies (circle) and graphene materials (black arrows). Copyright 2022 The Authors. Published by Elsevier B.V. This is an open-access article under the CC BY license [<a href="#B156-nanomaterials-13-01458" class="html-bibr">156</a>]. (<b>g</b>) (<b>g1</b>) Cell viability for different concentrations of samples carbon dots, carbon dots-gel, DS-carbon dots, DS-carbon dots-Gel, DS-Gel for 24 h. (<b>g2</b>) Histopathological microscopy of ocular tissues. Copyright 2021 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B64-nanomaterials-13-01458" class="html-bibr">64</a>]. * <span class="html-italic">p</span>-value &lt; 0.05, ** <span class="html-italic">p</span>-value &lt; 0.01,*** <span class="html-italic">p</span>-value &lt; 0.001, **** <span class="html-italic">p</span>-value &lt; 0.0001, n.s: not significant.</p>
Full article ">Figure 6
<p>(<b>a</b>) Lemna minor morphology after influence of pristine graphene and graphene oxide. (<b>1</b>) Control group. (<b>a1</b>–<b>a6,a13,a14</b>,<b>a16,a17</b>) after influence of pristine graphene. (<b>a7</b>–<b>a12</b>,<b>a15</b>,<b>a18</b>,<b>a19</b>) after graphene oxide influence. Copyright 2021 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B30-nanomaterials-13-01458" class="html-bibr">30</a>]. (<b>b</b>) Brightfield imaging of 6 hpf embryos for different post-exposure periods, in comparison with exposures to ultrapure water controls. Copyright 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B22-nanomaterials-13-01458" class="html-bibr">22</a>]. (<b>c</b>) The images of a whole-body single-photon emission computed tomography/computed tomography (SPECT/CT) imaging, and quantitative γ-counting of tissue biodistribution of <sup>153</sup>Sm@SWCNTs and <sup>153</sup>Sm@MWCNTs. Reprinted with permission from [<a href="#B76-nanomaterials-13-01458" class="html-bibr">76</a>]. Copyright 2020 American Chemical Society. (<b>d</b>) Targeting ability of <sup>64</sup>Cu-NOTA-GO-TRC105 towards 4T1 tumor-bearing mice (NOTA = a chelating agent of <sup>64</sup>Cu). (<b>d1</b>) The post injected PET images. (<b>d2</b>) <sup>64</sup>Cu-NOTA-GO alone TRC105. (<b>d3</b>) Preinjected blocking dose of TRC105. Reprinted with permission from Ref. [<a href="#B94-nanomaterials-13-01458" class="html-bibr">94</a>]. Copyright 2012 American Chemical Society.</p>
Full article ">Figure 7
<p>(<b>a</b>) The cell viability (MTT assay) for different concentrations of unloaded DOX, MWCNTs, nanocomposites, loaded DOX against A549 human melanoma cells after 48 h of treatment. Copyright 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B41-nanomaterials-13-01458" class="html-bibr">41</a>]. (<b>b</b>) (<b>b1</b>) TEM images of treated cells. (<b>b2</b>) AlamarBlue cell viability assay 24 h after treatment with GO–Rg3–DOX. (<b>b3</b>) SEM of treated cells. (<b>b4</b>) Reactive oxygen species (ROS). (<b>b5</b>) Schematics of GO–Rg3–DOX internalization. The enrichment in toxicity is marked with *. Copyright 2023 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B27-nanomaterials-13-01458" class="html-bibr">27</a>]. (<b>c</b>) Rg3, DOX, graphene oxide cytotoxicity in human breast cancer MDA-MB-231 cells. AlamarBlue assay 24 h after GO (<b>c1</b>), GO–Rg3 (<b>c2</b>), GO–Rg3–DOX (<b>c3</b>) treatment is shown together with ROS production (on the right side). The enrichment in toxicity is marked with *. Copyright 2023 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B27-nanomaterials-13-01458" class="html-bibr">27</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) The drug release from the functionalized carbon nanotubes at different pH assessed by fluorescence measurements under excitation at 259 nm. Copyright 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B43-nanomaterials-13-01458" class="html-bibr">43</a>]. (<b>b</b>) Sequential versus simultaneous addition of CR and DOX to SWNT. Percent of DOX added to the sample found in complexes. Congo Red (CR), anticancer drug doxorubicin. Copyright 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B40-nanomaterials-13-01458" class="html-bibr">40</a>]. (<b>c</b>) DOX release from SWCNT-CR-DOX complex at pH 5, and pH 7.4. Copyright 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B40-nanomaterials-13-01458" class="html-bibr">40</a>]. (<b>d</b>) The dependence of DOX release on time PK<sub>5</sub>E<sub>7</sub> (PEI-rGO/DOX) at different pH. Copyright 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B50-nanomaterials-13-01458" class="html-bibr">50</a>]. (<b>e</b>) Cumulative release of Quercetin from the microfiber scaffolds without electric stimulus (<b>e1</b>), and under 10 Hz (<b>e2</b>), 50 Hz (<b>e3</b>). Copyright 2021 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B52-nanomaterials-13-01458" class="html-bibr">52</a>]. (<b>f</b>) In vitro release profiles. Copyright 2021 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B64-nanomaterials-13-01458" class="html-bibr">64</a>]. (<b>g</b>) The comparison of in vitro release profile of the LNZ, LNZ–BCDs nanocomposite. Copyright 2023 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license [<a href="#B67-nanomaterials-13-01458" class="html-bibr">67</a>].</p>
Full article ">
Back to TopTop