Next Article in Journal
Functionalization of Carbon Nanotubes Surface by Aryl Groups: A Review
Previous Article in Journal
Temperature-Dependent Carrier Transport in GaN Nanowire Wrap-Gate Transistor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Inductively Coupled Nonthermal Plasma Synthesis of Size-Controlled γ-Al2O3 Nanocrystals

1
Department of Mechanical Engineering, University of Minnesota, 111 Church Street SE, Minneapolis, MN 55455, USA
2
Chemical Engineering and Materials Science Department, University of Minnesota, Minneapolis, MN 55455, USA
3
Ames National Laboratory, United States Department of Energy, Department of Chemistry, Iowa State University, Ames, IA 50011, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2023, 13(10), 1627; https://doi.org/10.3390/nano13101627
Submission received: 8 April 2023 / Revised: 8 May 2023 / Accepted: 10 May 2023 / Published: 12 May 2023

Abstract

:
Gamma alumina (γ-Al2O3) is widely used as a catalyst and catalytic support due to its high specific surface area and porosity. However, synthesis of γ-Al2O3 nanocrystals is often a complicated process requiring high temperatures or additional post-synthetic steps. Here, we report a single-step synthesis of size-controlled and monodisperse, facetted γ-Al2O3 nanocrystals in an inductively coupled nonthermal plasma reactor using trimethylaluminum and oxygen as precursors. Under optimized conditions, we observed phase-pure, cuboctahedral γ-Al2O3 nanocrystals with defined surface facets. Nuclear magnetic resonance studies revealed that nanocrystal surfaces are populated with AlO6, AlO5 and AlO4 units with clusters of hydroxyl groups. Nanocrystal size tuning was achieved by varying the total reactor pressure yielding particles as small as 3.5 nm, below the predicted thermodynamic stability limit for γ-Al2O3.

1. Introduction

Alumina, Al2O3, is extensively used for a wide range of applications because of its superior chemical, mechanical, and thermal properties [1,2]. Among the crystalline Al2O3 polymorphs, gamma alumina (γ-Al2O3) has attracted significant attention as a catalyst and a catalytic support [3,4,5,6] nanomaterial in heterogeneous catalysis and petroleum refining processes [7,8,9,10] due to its high specific surface area [3,4,5], high porosity [3], and excellent thermal stability [4,9].
Typically, gibbsite (γ-Al(OH)3) and boehmite (γ-AlOOH) are used as precursors to synthesize nanostructured γ-Al2O3 particles [11,12]. Various methods have been developed to synthesize γ-Al2O3 nanocrystals including precipitation [3,13,14,15,16,17,18], sol-gel synthesis [4,19,20,21,22], combustion [23], hydrolysis [24,25], and high-pressure compaction [26]. It was shown that the final structure, morphology, and properties of γ-Al2O3 nanocrystals largely depend on the synthetic routes and experimental parameters. However, most of these methods require high-reaction temperatures, long reaction times (~35 h–6 days), and additional calcination steps.
The detailed structure of γ-Al2O3 still remains unclear owing to the structural complexity and the notorious presence of mixed alumina polymorphs during synthesis [27,28,29]. Many properties and surface characteristics of γ-Al2O3 are still actively debated [29,30]. Nanostructured γ-Al2O3 often results in cube octahedral or octahedral morphology demonstrating distinct surface facets [7,31,32]. These surface facets directly influence the anchoring ability, dispersion, and surface interaction of metal catalyst atoms and their catalytic behaviour [7,31]. Reported synthetic methods that utilize high-temperature conditions often produce polydisperse nanocrystal aggregates with non-reproducible facet orientations, which hamper their catalytic activity.
Here, we report a rapid, single-step, low temperature synthesis of phase-pure facetted γ-Al2O3 nanocrystals using a flow-through nonthermal inductively coupled plasma (ICP) reactor. The nonthermal plasma approach has emerged as a competitive technique in producing a variety of nanoparticles with high-purity and narrow size distributions [33,34,35]. One of the major advantages of nanocrystals synthesis in plasma is the unipolar electrical charging of particles that reduces or eliminates the agglomeration of nanoparticles [33,34]. Exploiting this unique feature, we synthesized γ-Al2O3 nanocrystals with defined surface planes using a low-pressure nonthermal ICP reactor utilizing trimethylaluminum and oxygen as precursors. Cendejas et al., recently studied the crystallization of Al2O3 nanoparticles in a nonthermal capacitively coupled plasma (CCP) reactor and achieved the γ-Al2O3 phase, but their study did not demonstrate control of the nanocrystal size and lacked detailed microstructural characterization [36]. The plasma density of an ICP used in this study is usually more than one order of magnitude higher than that of a CCP [37]. We demonstrate, here, that this causes effective crystallization of the γ-Al2O3 phase [38,39]. We also performed detailed microstructural characterizations of the synthesized γ-Al2O3 nanocrystals, which revealed cuboctahedra morphology with (111) stepped facets. Furthermore, size tuning of γ-Al2O3 nanocrystals was achieved here by varying the total reactor pressure.

2. Methods

2.1. γ-Al2O3 Nanocrystal Synthesis

γ-Al2O3 nanocrystals were synthesized using a flow-through, low-pressure, nonthermal ICP reactor, Figure 1, which is similar to a previously published report [40]. A simplified schematic of this system is shown in Figure 1. Briefly, a low-pressure discharge was generated by the application of 120 W radio frequency (13.56 MHz) power to an induction copper coil with three turns, approximately 2.5 cm in length and 2.5 cm in diameter, wrapped around a quartz tube with an outer diameter of 2.5 cm. The RF power was generated by a 13.56 MHz RF power supply (AG 0313, T&C Power Conversion, Rochester, NY, USA) and applied through an impedance matching network (Model HFT1500, Vectronics, Starkville, MS, USA). Trimethylaluminum (TMA, 97%, Sigma Aldrich, St. Louis, MO, USA) vapor, carried by an argon flow, entered through the top inlet mixed with a diluted oxygen/argon flow that entered from the side arm tube before the plasma discharge. In a typical recipe, the TMA volumetric flow rate was about 0.2 standard cubic centimetres per minute (sccm) carried by 6 sccm of argon flowing through the TMA bubbler, in which pressure was stable around 350 Torr. The oxygen volumetric flow rate and the associated argon volumetric flow rate were 2.5 sccm and 60 sccm, respectively. γ-Al2O3 nanocrystals were collected on silicon wafer substrates by inertial impaction [41] using an orifice with a 0.25 × 8 mm rectangular opening. The total reactor pressure was maintained at approximately 3.8 Torr for the typical recipe. The collection rate of γ-Al2O3 nanocrystals was 18 mg/h. Deposited nanocrystals were stored under ambient conditions. Size tuning experiments were carried out by varying the orifice diameter, leading to a change in reactor pressure. While the above-described reaction conditions were kept the same for 0.25 mm and 0.5 mm orifice diameters, 15 sccm of Ar flow with TMA and 5 sccm of O2 were used for an orifice of 1 mm diameter. The 0.5 mm and 1 mm orifice diameters resulted in total reactor pressures of 2.5 and 1.3 Torr, respectively.

2.2. X-ray Diffraction (XRD)

XRD was performed using a Bruker D8 Discover 2D X-ray diffractometer (Bruker, Billerica, MA, USA) equipped with a Co Kα (λ = 1.79 Å) radiation point source in Bragg–Brentano configuration. The diffraction patterns were collected at 25 °C with a step size of 0.01° per step and a retention time of 5 s per increment. Nanocrystals were directly deposited onto Si wafers for XRD analysis. The XRD patterns were converted to Cu source (λ = 1.54 Å) for data analysis. Data analysis was performed using the Material Data Incorporated Jade 8.0 software package.

2.3. Fourier-Transform Infrared (FT-IR) Spectroscopy

FT-IR spectroscopy was performed on a Bruker Alpha spectrometer (Bruker, Billerica, MA, USA) using the attenuated total reflection (ATR) module in a nitrogen filled glovebox. The as synthesized nanocrystals were dissolved in methanol, drop cast onto the ATR crystal, and 20 scans were taken for each measurement at 2 cm−1 resolution.

2.4. X-ray Photoelectron Spectroscopy (XPS)

XPS was performed with a PHI Versa Probe III XPS and UPS system (Physical Electronics, Chanhassen, MN, USA). Samples were directly deposited onto Si wafers for XPS analysis. The binding energy of C 1 s at 284.6 eV was used as a reference. A 55 eV bandpass energy was used to collect high-resolution scans. Peaks were fitted using PHI’s Multipak software v9. XPS survey scans were taken at a bandpass energy of 280 eV. The atomic percentages were calculated using Multipak software v9.

2.5. Transmission Electron Microscopy (TEM)

High-angle annular dark-field scanning transmission microscopy (HAADF-STEM) images were collected using an aberration-corrected FEI Titan G2 60-300 (FEI, Hillsboro, OR, USA) operated at 300 kV with a 25 mrad semi-convergence angle. Transmission electron microscopy (TEM) images were collected using a Thermo Scientific Talos F200X (Thermo Scientific, Waltham, MA, USA) operated at 200 kV. Samples were directly deposited onto lacy/thin double carbon-coated TEM grids. Nanocrystal dimensions were measured using ImageJ. A minimum of 300 particles were counted for every size distribution and the nanoparticle dimensions were reported with geometric standard deviations.

2.6. Solid-State Nuclear Magnetic Resonance (NMR) Spectroscopy

Solid-State NMR spectroscopy was performed at the National High Magnetic Field Laboratory in Tallahassee, FL, USA, on a 19.6 T (ν0(1H) = 833 MHz) magnet equipped with a Bruker NEO console (Bruker, Billerica, MA, USA) and a 3.2 mm Low-E HX magic-angle spinning (MAS) NMR probe. The magnet field strength was referenced to the 17O chemical shift of tap water [δiso(17O) = 2.825 ppm]. 1H and 27Al shifts were indirectly referenced to 1% TMS in CDCl3 and aqueous Al(NO3)3 solution, respectively, using the previously published relative NMR frequencies [42]. The as synthesized γ-Al2O3 nanocrystals were packed into the appropriate sized NMR rotor in a N2 filled glove-box. All NMR spectra were processed in Bruker Topspin 4.0.7. The 1H π/2 and π pulse lengths were 5 and 10 μs in duration, respectively, corresponding to a 50 kHz radio frequency. 27Al central-transition (CT) selective π/2 and π pulse lengths were 5 and 10 μs in duration, respectively, corresponding to a 16.7 kHz RF field and 50 kHz CT nutation frequency. 1H and 27Al longitudinal relaxation constants (T1) were measured using a saturation recovery experiment. A quantitative single-pulse 27Al NMR spectrum was recorded with a 30 s recycle delay [ca. 10–30 × T1(27Al)] and a 10° tip-angle excitation pulse. A 2D 27Al → 1H Dipolar-mediated Refocused Insensitive Nuclei Enhanced by a Polarization Transfer (D-RINEPT) 2D correlation NMR spectrum was recorded with previously described NMR pulse sequences [43,44], and the symmetry-based SR 4 1 2 heteronuclear dipolar-recoupling sequence [45] was applied to the 1H spins. A 2D 1H dipolar double quantum-single quantum (DQ-SQ) homonuclear correlation NMR spectrum was recorded with the previously described back-to-back (BABA) NMR pulse sequence [46,47]. 2D 27Al dipolar DQ-SQ homonuclear correlation NMR spectra were recorded with BR 2 2 1   homonuclear dipolar recoupling and the previously described pulse sequences [48]. A CT (central transition) selective π pulse was applied during the indirect dimension (t1) evolution period to ensure only DQ coherence between two 27Al spins in the CT spin states were observed [49]. Rotor-assisted population transfer (RAPT) was applied to ±400 kHz off-resonance before all 27Al → 1H D-RINEPT and 27Al DQ-SQ NMR experiments to enhance the 27Al CT NMR signals [50,51]. Small phase incremental alternation with 64 step (SPINAL-64) [52] heteronuclear decoupling with a 50 kHz 1H RF field was applied during the detection of 27Al NMR signals.

3. Results and Discussion

Freestanding γ-Al2O3 nanocrystals were synthesized using the flow-through nonthermal plasma process discussed above. Even though nucleophilic molecular oxygen is highly reactive towards the electron deficient TMA under ambient conditions, it is found to be rather inert or unreactive with many organometallic precursors including TMA at low-pressure and low-temperature conditions [53,54]. Thus, the pre-mixing of TMA with molecular oxygen before the plasma region did not lead to a reaction of the gas before entering the plasma. The detailed chemistry occurring in the plasma region is poorly understood, but according to the previous reports, oxidation of TMA will proceed through initial decomposition [53], which is followed by the reaction with atomic oxygen. In the plasma region, TMA is expected to rapidly decompose to produce reactive intermediates such as metal atoms (Al) released by electron impact or partially decomposed fragments generated by radical abstraction reactions. These reactive intermediates will react with atomic oxygen to form fully oxidized γ-Al2O3 nanocrystals. Nucleation and growth of these nanocrystals are believed to follow the typical nanoparticle growth in plasmas, where the growth proceeds through the nucleation of clusters that rapidly coagulate to form nanoparticles [55]. Crystallization is induced by the heat generated by energetic surface reactions such as electron-ion recombination and surface chemical reactions [36,39]. Due to the higher plasma density of the ICP, and thus more intensive nanoparticle heating, the nominal power reported here (120 W) is around 2.5 times lower than the required nominal power (300 W) [36] for full crystalline nanoparticles using a CCP.
As synthesized γ-Al2O3 nanocrystals were structurally characterized through the means of XRD, TEM, and FT-IR, XPS, and solid-state NMR spectroscopies. These materials often adopt a defect spinel structure with oxygen atoms forming a face centered cubic structure, and Al cations occupying the interstitial tetrahedral and octahedral sites [12,30,56,57]. Figure 2a shows the theoretical and experimental powder-XRD patterns of γ-Al2O3 nanocrystals. The experimental pattern exhibits six reflections at 32.0°, 37.9°, 39.9°, 46.2°, 61.2°, and 66.8°, which correspond to the standard γ-Al2O3 crystal planes of 220, 331, 222, 400, 511, and 440, respectively (JCPDS No. 29-1486).
FT-IR and XPS were performed to evaluate the surface composition of the plasma synthesized γ-Al2O3 nanocrystals. As shown in Figure 2b, the absorption band ranging from 430–890 cm−1 can be attributed to the characteristic γ-Al2O3 stretching modes of four-fold (AlO4) and six-fold coordinated (AlO6) Al sites [58,59,60]. While peaks at ~890 cm−1 and the shoulder at ~760 cm−1 can be assigned to the Al-O stretching modes of AlO4, the shoulder at ~620 cm−1 relates to the Al-O stretching modes of AlO6. The broad absorption band at ~3570 cm−1 corresponds to the surface-bound and free –OH groups. In Figure 2c, XPS survey of these nanocrystals revealed three main peaks corresponding to the Al 2 p, C 1 s, and O 1s at 74, 285, and 530 eV lines, respectively. The atomic percentage of C was around 12%, which may originate from the methyl groups in TMA or partly due to the contamination in air as samples were briefly exposed to the air during transfer. A marginal C contamination is commonly observed when TMA is used due to the strong Al-C bonds [61,62]. The high-resolution XPS spectrum of Al showed a single peak at 74.2 eV, which corresponds to Al-O bonding in Al2O3 (Figure 2d) [63,64]. The O 1s peak at 530.5 eV can be deconvoluted to two individual peaks, indicating –OH and Al-O surface species (Figure 2e) [63,64,65,66]. The observed XPS features are consistent with the observed absorption peaks in the FT-IR spectrum, as well as the solid-state NMR spectra discussed below.
We further investigated the local structure of the plasma synthesized γ-Al2O3 nanocrystals via high-field (19.6 T) 1H and 27Al magic-angle spinning (MAS) solid-state NMR spectroscopy. A two-dimensional (2D) 27Al → 1H dipolar-refocused insensitive nuclei enhanced by the polarization transfer (D-RINEPT) NMR spectrum reveals a broad 1H NMR signal centered at ca. 3 ppm, correlating to three 27Al NMR signals centered at ca. 12, 35 and 70 ppm, which are assigned to AlO6, AlO5 and AlO4 species, respectively (Figure 3a). The short duration of dipolar recoupling (τrec = 448 μs) ensures that the observed Al sites are likely on the surface of the nanocrystals and capped with hydroxyl groups. A 1H dipolar double-quantum-single-quantum (DQ-SQ) homonuclear correlation NMR spectrum reveals that the hydroxyl groups are clustered (i.e., spatially proximate to each other) on the surface of the nanocrystal, consistent with the broad stretching band observed in the FT-IR spectrum (Figure S1). Integration of a quantitative 10° single-pulse (SP) 27Al NMR spectrum reveals that the population of AlO6, AlO5 and AlO4 is ca. 65, 3 and 32%, respectively (Figure 3b). The much larger population of AlO4 observed in the SP 27Al NMR spectrum, as compared to the surface-selective D-RINEPT NMR spectrum, suggests that AlO6 predominantly terminates the surface of the nanocrystal (Figure 3b). Lastly, we recorded 2D 27Al dipolar DQ-SQ NMR spectra to probe the AlOx-AlOx (x = 4–6) linkages with the nanocrystals (Figure S2). The 2D 27Al DQ-SQ NMR spectra reveal intense AlO6-AlO6 and AlO6-AlO4 homonuclear correlations, suggesting that the majority of the nanocrystal contains AlO6 linked to either another AlO6 or AlO4, consistent with prior NMR experiments on γ-Al2O3 [48]. There were additional weak AlO4-AlO4 correlations, suggesting that this linkage is relatively rare. No correlations involving AlO5 were observed due to the low population of the AlO5 site. We note that the observed 27Al homonuclear correlations of the synthesized γ-Al2O3 nanocrystals are near identical to that of commercially available γ-Al2O3 (Figure S2).
HAADF-STEM images were analysed to examine the morphology and the size distribution of plasma synthesized γ-Al2O3 nanocrystals. Figure 4 shows representative HAADF-STEM images and the size histogram of γ-Al2O3 nanocrystals. Overall, the sample consisted of well-dispersed, facetted nanocrystals with an average crystal diameter of 12.0 nm with a geometric standard deviation of 1.45. The primary morphology of these particles appeared to be cuboctahedral [32], exposing (111), (110), and (001) facets. High-resolution images revealed that the (110) surface was not atomically flat but consisted of stepped facets of alternating (111) surfaces, as shown in Figure 4b.
This observation of surface reconstruction of the (110) surface is consistent with the DFT calculations by Pinto et al., suggesting that the (110) facet was thermodynamically favored to reconstruct into (111) facets [67]. Furthermore, enhanced surface contrast was observed in (111) terminating facets (Figure 4a and Figure 5a,b), which corroborates previous literature reports of facetted γ-Al2O3 nanocrystals [7,31,32]. The enhanced surface contrast found in cuboctahedral γ-Al2O3 nanocrystals is hypothesized [31] to occur due to the excess Al3+ cations on (111) surface planes. The presence of excess Al3+ cations at the nanoparticle’s surface is a modification from the bulk structure, which is found to be critical in improving the dispersion and the thermal stability of fine metal particle catalysts in γ-Al2O3 catalyst supports [31].
To evaluate the impact of total reactor pressure on the size of γ-Al2O3 nanocrystals, we varied the reactor pressure by tuning the orifice diameter at constant volumetric flow rates. The orifice width was tuned to 0.25 mm, 0.5 mm, and 1 mm, resulting in 3.8 Torr, 2.5 Torr, and 1.3 Torr total reactor pressures, respectively. The size of γ-Al2O3 nanocrystals decreased with decreasing reactor pressure, producing 12.0 nm particles at 3.8 Torr, 8.5 nm particles at 2.5 Torr, and 4.8 nm particles at 1.3 Torr with geometric standard deviations of 1.45, 1.42 and 1.25, respectively (Figure 5a–d). The morphology of γ-Al2O3 nanocrystals remained similar under different pressure conditions exhibiting their typical facetted nature. In many studies, it has been found that the average nanoparticle size correlates nearly linearly with the residence time of particles in the plasma [68]. A slower gas velocity yields longer residence time, thus increasing the size of particles leaving the plasma. Here, for essentially the same total of volumetric flow rate, a wider orifice reduces the pressure in the reactor while increasing the gas velocity. Therefore, an increasing orifice size results in a decreasing residence time and a reduction of the average particle sizes of γ-Al2O3 nanocrystals. Assuming 2 cm for the plasma length, the gas residence times are estimated as 33 ms, 22 ms, and 10 ms with the orifice width of 0.25 mm, 0.5 mm, and 1 mm, respectively. Here, the average nanoparticle diameters are found to increase nearly linearly with the estimated gas residence times. Recently, we found the size of the particles leaving a flow-through nonthermal capacitively coupled plasma reactor is mainly determined by the balance of gas drag forces and electrostatic forces acting on the particles [69]. However, it is unclear whether trapping plays a role in the inductively coupled plasma reactor.
XRD patterns confirmed the crystalline nature of γ-Al2O3 nanoparticles synthesized under all three pressure conditions (Figure 5e). At the nanoscale, the thermodynamics of the growth process drives the crystalline structure of Al2O3 nanocrystals. It was reported that Al2O3 nanoparticles larger than ~20 nm should adopt an α structure, while Al2O3 nanoparticles smaller than ~6.5 nm will be amorphous [70]. Hence, γ-Al2O3 nanocrystals were predicted to be thermodynamically stable in the range ~20–6.5 nm [70]. Our approach demonstrated that the size tuning of γ-Al2O3 is feasible in a range of 12–5 nm with some nanocrystals as small as 3.5 nm, as shown in Figure 5c, which is below the thermodynamically predicted size limit. It will be interesting to explore whether the nonthermal plasma environment, where nanoparticles are charged while they grow, more generally allows for the synthesis of material phases outside of thermodynamically predicted size limits.

4. Conclusions

We demonstrated the synthesis of γ-Al2O3 nanocrystals using an inductively coupled nonthermal plasma. While XRD patterns confirmed phase-pure crystalline γ-Al2O3 nanoparticles, TEM images revealed cuboctahedra morphology with (111) stepped facets. The total pressure of the reactor was varied by tuning the orifice diameter at a constant gas flow to yield γ-Al2O3 particles ranging between 5–12 nm. We observed γ-Al2O3 nanocrystals as small as 3.5 nm, which is below the size at which thermodynamics would predict amorphous alumina to be the most stable phase.
Overall, this study demonstrates the single-step synthesis of size-tunable, facetted γ-Al2O3 nanocrystals without additional post-synthetic calcination or annealing steps. These particles, specifically the γ-Al2O3 as small as 3.5 nm, could be used as heterogeneous catalysts and catalytic supports. Their specific surface areas along with catalytic performance need to be probed in future work.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13101627/s1, Figure S1: 2D 1H dipolar DQ-SQ homonuclear correlation NMR spectrum recorded with a 17.857 kHz MAS frequency and 112 μs (i.e., two rotor cycles) of total homonuclear dipolar recoupling; Figure S2: 2D 27Al dipolar DQ-SQ homonuclear correlation NMR spectra of plasma synthesized gamma alumina nanocrystals and commercially available gamma alumina.

Author Contributions

Z.X., H.P.A., R.W.D., A.R. and U.R.K. conceived and designed experiments and analysis. Z.X., Y.-J.J. and H.P.A. performed nanocrystal synthesis and characterization; J.T.H. performed HAADF-STEM and analyzed the results; R.W.D. and A.R. performed Solid-State NMR and analyzed the results. All authors participated in the discussion and analysis of the data. H.P.A. and Z.X. drafted the manuscript using contributions from all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Army Research Office under MURI Grant W911NF-18-1-0240. Parts of this work were carried out in the Characterization Facility, University of Minnesota, which receives partial support from the US National Science Foundation through the MRSEC (Award Number DMR-2011401) and the NNCI (Award Number ECCS-2025124) programs. Solid-State NMR Spectroscopy (R.D. and A.R.) was supported work was supported by the National Science Foundation under Grant No. CBET-1916809. 19.6 T NMR experiments were performed at the National High Magnetic Field Laboratory. The National High Magnetic Field Laboratory is supported by the National Science Foundation through NSF/DMR-1644779 and the State of Florida.

Data Availability Statement

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

Acknowledgments

The authors are grateful to Ivan Hung for assistance with 27Al NMR experiments performed at the National High Magnetic Field Laboratory.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wefers, K. Alumina Chemicals: Science and Technology Handbook; The American Ceramic Society: Westerville, OH, USA, 1990. [Google Scholar]
  2. Ziva, A.Z.; Suryana, Y.K.; Kurniadianti, Y.S.; Nandiyanto, A.B.D.; Kurniawan, T. Recent Progress on the Production of Aluminum Oxide (Al2O3) Nanoparticles: A Review. Mech. Eng. Soc. Ind. 2021, 1, 54–77. [Google Scholar] [CrossRef]
  3. Potdar, H.S.; Jun, K.-W.; Bae, J.W.; Kim, S.-M.; Lee, Y.-J. Synthesis of nano-sized porous γ-alumina powder via a precipitation/digestion route. Appl. Catal. A Gen. 2007, 321, 109–116. [Google Scholar] [CrossRef]
  4. Hosseini, Z.; Taghizadeh, M.; Yaripour, F. Synthesis of nanocrystalline γ-Al2O3 by sol-gel and precipitation methods for methanol dehydration to dimethyl ether. J. Nat. Gas Chem. 2011, 20, 128–134. [Google Scholar] [CrossRef]
  5. Baghalha, M.; Mohammadi, M.; Ghorbanpour, A. Coke deposition mechanism on the pores of a commercial Pt–Re/γ-Al2O3 naphtha reforming catalyst. Fuel Process. Technol. 2010, 91, 714–722. [Google Scholar] [CrossRef]
  6. Sánchez, M.; Navas, M.; Ruggera, J.F.; Casella, M.L.; Aracil, J.; Martinez, M. Biodiesel production optimization using γAl2O3 based catalysts. Energy 2014, 73, 661–669. [Google Scholar] [CrossRef]
  7. Rozita, Y.; Brydson, R.; Comyn, T.P.; Scott, A.J.; Hammond, C.; Brown, A.; Chauruka, S.; Hassanpour, A.; Young, N.P.; Kirkland, A.I.; et al. A Study of Commercial Nanoparticulate γ-Al2O3Catalyst Supports. Chemcatchem 2013, 5, 2695–2706. [Google Scholar] [CrossRef]
  8. Ishaq, K.; Saka, A.A.; Kamardeen, A.O.; Abdulrahman, A.; Adekunle, I.K.; Afolabi, A.S. Application of γ alumina as catalyst support for the synthesis of CNTs in a CVD reactor. Adv. Nat. Sci. Nanosci. Nanotechnol. 2018, 9, 035012. [Google Scholar] [CrossRef]
  9. Trueba, M.; Trasatti, S.P. γ-Alumina as a Support for Catalysts: A Review of Fundamental Aspects. Eur. J. Inorg. Chem. 2005, 2005, 3393–3403. [Google Scholar] [CrossRef]
  10. Bose, S.; Das, C. Preparation, characterization, and activity of γ-alumina-supported molybdenum/cobalt catalyst for the removal of elemental sulfur. Appl. Catal. A Gen. 2016, 512, 15–26. [Google Scholar] [CrossRef]
  11. Zhang, X.; Huestis, P.L.; Pearce, C.I.; Hu, J.Z.; Page, K.; Anovitz, L.M.; Aleksandrov, A.B.; Prange, M.P.; Kerisit, S.; Bowden, M.E.; et al. Boehmite and Gibbsite Nanoplates for the Synthesis of Advanced Alumina Products. ACS Appl. Nano Mater. 2018, 1, 7115–7128. [Google Scholar] [CrossRef]
  12. Paglia, G.; Buckley, C.E.; Rohl, A.L.; Hart, R.D.; Winter, K.; Studer, A.J.; Hunter, B.A.; Hanna, J.V. Boehmite Derived γ-Alumina System. 1. Structural Evolution with Temperature, with the Identification and Structural Determination of a New Transition Phase, γ‘-Alumina. Chem. Mater. 2004, 16, 220–236. [Google Scholar] [CrossRef]
  13. Wang, S.; Li, X.; Wang, S.; Li, Y.; Zhai, Y. Synthesis of γ-alumina via precipitation in ethanol. Mater. Lett. 2008, 62, 3552–3554. [Google Scholar] [CrossRef]
  14. Jbara, A.S.; Othaman, Z.; Ati, A.A.; Saeed, M.A. Characterization of γ-Al2O3 nanopowders synthesized by Co-precipitation method. Mater. Chem. Phys. 2017, 188, 24–29. [Google Scholar] [CrossRef]
  15. Wang, J.; Zhao, D.; Zhou, G.; Zhang, C.; Zhang, P.; Hou, X. Synthesis of nano-sized γ-Al2O3 with controllable size by simple homogeneous precipitation method. Mater. Lett. 2020, 279, 128476. [Google Scholar] [CrossRef]
  16. Yi, J.-H.; Sun, Y.-Y.; Gao, J.-F.; Xu, C.-Y. Synthesis of crystalline γ-Al2O3 with high purity. Trans. Nonferrous Met. Soc. China 2009, 19, 1237–1242. [Google Scholar] [CrossRef]
  17. Huang, B.; Bartholomew, C.H.; Woodfield, B.F. Facile structure-controlled synthesis of mesoporous γ-alumina: Effects of alcohols in precursor formation and calcination. Microporous Mesoporous Mater. 2013, 177, 37–46. [Google Scholar] [CrossRef]
  18. Ali, S.; Abbas, Y.; Zuhra, Z.; Butler, I.S. Synthesis of γ-alumina (Al2O3) nanoparticles and their potential for use as an adsorbent in the removal of methylene blue dye from industrial wastewater. Nanoscale Adv. 2019, 1, 213–218. [Google Scholar] [CrossRef]
  19. Mohamad, S.N.S.; Mahmed, N.; Che Halin, D.S.; Abdul Razak, K.; Norizan, M.N.; Mohamad, I.S. Synthesis of alumina nanoparticles by sol-gel method and their applications in the removal of copper ions (Cu2+) from the solution. In Proceedings of the IOP Conference Series: Materials Science and. Engineering, Wuhan, China, 10–12 October 2019; Volume 701. [Google Scholar] [CrossRef]
  20. Dubey, S.; Singh, A.; Nim, B.; Singh, I.B. Optimization of molar concentration of AlCl3 salt in the sol–gel synthesis of nanoparticles of gamma alumina and their application in the removal of fluoride of water. J. Sol-Gel Sci. Technol. 2017, 82, 468–477. [Google Scholar] [CrossRef]
  21. Kim, S.-M.; Lee, Y.-J.; Jun, K.-W.; Park, J.-Y.; Potdar, H.S. Synthesis of thermo-stable high surface area alumina powder from sol–gel derived boehmite. Mater. Chem. Phys. 2007, 104, 56–61. [Google Scholar] [CrossRef]
  22. Yuan, Q.; Yin, A.-X.; Luo, C.; Sun, L.-D.; Zhang, Y.-W.; Duan, W.-T.; Liu, H.-C.; Yan, C.-H. Facile Synthesis for Ordered Mesoporous γ-Aluminas with High Thermal Stability. J. Am. Chem. Soc. 2008, 130, 3465–3472. [Google Scholar] [CrossRef]
  23. Afruz, F.B.; Tafreshi, M.J. Synthesis of γ-Al2O3 Nano Particles by Different Combustion Modes Using Ammonium Carbonate. Indian J. Pure Appl. Phys. 2014, 52, 385–387. [Google Scholar]
  24. Wang, Y.; Wang, J.; Shen, M.; Wang, W. Synthesis and properties of thermostable γ-alumina prepared by hydrolysis of phosphide aluminum. J. Alloys Compd. 2009, 467, 405–412. [Google Scholar] [CrossRef]
  25. Ramesh, S.; Sominska, E.; Cina, B.; Chaim, R.; Gedanken, A. Nanocrystalline -Alumina Synthesized by Sonohydrolysis of Alkoxide Precursor in the Presence of Organic Acids: Structure and Morphological Properties. J. Am. Ceram. Soc. 2000, 83, 89–94. [Google Scholar] [CrossRef]
  26. Costa, T.M.H.; Gallas, M.R.; Benvenutti, E.V.; Da Jornada, J.A.H. Study of Nanocrystalline γ-Al2O3 Produced by High-Pressure Compaction. J. Phys. Chem. B 1999, 103, 4278–4284. [Google Scholar] [CrossRef]
  27. Paglia, G.; Buckley, C.E.; Rohl, A.L.; Hunter, B.A.; Hart, R.D.; Hanna, J.V.; Byrne, L.T. Tetragonal structure model for boehmite-derived γ-alumina. Phys. Rev. B Condens. Matter Mater Phys. 2003, 68, 144110. [Google Scholar] [CrossRef]
  28. Peintinger, M.F.; Kratz, M.J.; Bredow, T. Quantum-chemical study of stable, meta-stable and high-pressure alumina polymorphs and aluminum hydroxides. J. Mater. Chem. A Mater. 2014, 2, 13143–13158. [Google Scholar] [CrossRef]
  29. Acikgoz, M.; Harrell, J.; Pavanello, M. Seeking a Structure–Function Relationship for γ-Al2O3 Surfaces. J. Phys. Chem. C 2018, 122, 25314–25330. [Google Scholar] [CrossRef]
  30. Ayoola, H.O.; House, S.D.; Bonifacio, C.S.; Kisslinger, K.; Saidi, W.A.; Yang, J.C. Evaluating the accuracy of common γ-Al2O3 structure models by selected area electron diffraction from high-quality crystalline γ-Al2O3. Acta Mater. 2020, 182, 257–266. [Google Scholar] [CrossRef]
  31. Jefferson, D.A. The surface activity of ultrafine particles. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2000, 358, 2683–2692. [Google Scholar] [CrossRef]
  32. Rozita, Y.; Brydson, R.; Scott, A.J. An investigation of commercial gamma-Al2O3 nanoparticles. J. Phys. Conf. Ser. 2010, 241, 012096. [Google Scholar] [CrossRef]
  33. Kortshagen, U.R.; Sankaran, R.M.; Pereira, R.N.; Girshick, S.L.; Wu, J.J.; Aydil, E.S. Nonthermal Plasma Synthesis of Nanocrystals: Fundamental Principles, Materials, and Applications. Chem. Rev. 2016, 116, 11061–11127. [Google Scholar] [CrossRef] [PubMed]
  34. Kortshagen, U. Nonthermal Plasma Synthesis of Nanocrystals: Fundamentals, Applications, and Future Research Needs. Plasma Chem. Plasma Process. 2016, 36, 73–84. [Google Scholar] [CrossRef]
  35. Mangolini, L.; Thimsen, E.; Kortshagen, U. High-Yield Plasma Synthesis of Luminescent Silicon Nanocrystals. Nano Lett. 2005, 5, 655–659. [Google Scholar] [CrossRef] [PubMed]
  36. Cendejas, A.J.; Sun, H.; Hayes, S.E.; Kortshagen, U.; Thimsen, E. Predicting plasma conditions necessary for synthesis of γ-Al2O3 nanocrystals. Nanoscale 2021, 13, 11387–11395. [Google Scholar] [CrossRef]
  37. El-Fayoumi, I.M.; Jones, I.R.; Turner, M.M. Hysteresis in the E- to H-mode transition in a planar coil, inductively coupled rf argon discharge. J. Phys. D Appl. Phys. 1998, 31, 3082–3094. [Google Scholar] [CrossRef]
  38. Lopez, T.; Mangolini, L. On the nucleation and crystallization of nanoparticles in continuous-flow nonthermal plasma reactors. J. Vac. Sci. Technol. B 2014, 32, 061802. [Google Scholar] [CrossRef]
  39. Kramer, N.J.; Anthony, R.J.; Mamunuru, M.; Aydil, E.S.; Kortshagen, U.R. Plasma-induced crystallization of silicon nanoparticles. J. Phys. D Appl. Phys. 2014, 47, 75202. [Google Scholar] [CrossRef]
  40. Li, Z.; Wray, P.R.; Su, M.P.; Tu, Q.; Andaraarachchi, H.P.; Jeong, Y.J.; Atwater, H.A.; Kortshagen, U.R. Aluminum Oxide Nanoparticle Films Deposited from a Nonthermal Plasma: Synthesis, Characterization, and Crystallization. ACS Omega 2020, 5, 24754–24761. [Google Scholar] [CrossRef]
  41. Holman, Z.C.; Kortshagen, U.R. A flexible method for depositing dense nanocrystal thin films: Impaction of germanium nanocrystals. Nanotechnology 2010, 21, 335302. [Google Scholar] [CrossRef]
  42. Harris, R.K.; Becker, E.D.; Cabral de Menezes, S.M.; Goodfellow, R.; Granger, P. NMR nomenclature. Nuclear spin properties and conventions for chemical shifts(IUPAC Recommendations 2001). Pure Appl. Chem. 2001, 73, 1795–1818. [Google Scholar] [CrossRef]
  43. Trébosc, J.; Hu, B.; Amoureux, J.P.; Gan, Z. Through-space R3-HETCOR experiments between spin-1/2 and half-integer quadrupolar nuclei in solid-state NMR. J. Magn. Reson. 2007, 186, 220–227. [Google Scholar] [CrossRef] [PubMed]
  44. Venkatesh, A.; Hanrahan, M.P.; Rossini, A.J. Proton detection of MAS solid-state NMR spectra of half-integer quadrupolar nuclei. Solid State Nucl. Magn. Reson. 2017, 84, 171–181. [Google Scholar] [CrossRef] [PubMed]
  45. Brinkmann, A.; Kentgens, A.P.M. Proton-Selective17O−1H Distance Measurements in Fast Magic-Angle-Spinning Solid-State NMR Spectroscopy for the Determination of Hydrogen Bond Lengths. J. Am. Chem. Soc. 2006, 128, 14758–14759. [Google Scholar] [CrossRef] [PubMed]
  46. Schnell, I.; Lupulescu, A.; Hafner, S.; Demco, D.E.; Spiess, H.W. Resolution Enhancement in Multiple-Quantum MAS NMR Spectroscopy. J. Magn. Reson. 1998, 133, 61–69. [Google Scholar] [CrossRef] [PubMed]
  47. Feike, M.; Demco, D.E.; Graf, R.; Gottwald, J.; Hafner, S.; Spiess, H. Broadband Multiple-Quantum NMR Spectroscopy. J. Magn. Reson. A 1996, 122, 214–221. [Google Scholar] [CrossRef]
  48. Wang, Q.; Hu, B.; Lafon, O.; Trébosc, J.; Deng, F.; Amoureux, J. Double-quantum homonuclear NMR correlation spectroscopy of quadrupolar nuclei subjected to magic-angle spinning and high magnetic field. J. Magn. Reson. 2009, 200, 251–260. [Google Scholar] [CrossRef]
  49. Mali, G.; Fink, G.; Taulelle, F. Double-quantum homonuclear correlation magic angle sample spinning nuclear magnetic resonance spectroscopy of dipolar-coupled quadrupolar nuclei. J. Chem. Phys. 2004, 120, 2835. [Google Scholar] [CrossRef]
  50. Kwak, H.-T.; Prasad, S.; Clark, T.; Grandinetti, P.J. Enhancing sensitivity of quadrupolar nuclei in solid-state NMR with multiple rotor assisted population transfers. Solid State Nucl. Magn. Reson. 2003, 24, 71–77. [Google Scholar] [CrossRef]
  51. Yao, Z.; Kwak, H.-T.; Sakellariou, D.; Emsley, L.; Grandinetti, P.J. Sensitivity enhancement of the central transition NMR signal of quadrupolar nuclei under magic-angle spinning. Chem. Phys. Lett. 2000, 327, 85–90. [Google Scholar] [CrossRef]
  52. Fung, B.M.; Khitrin, A.K.; Ermolaev, K. An Improved Broadband Decoupling Sequence for Liquid Crystals and Solids. J. Magn. Reson. 2000, 142, 97–101. [Google Scholar] [CrossRef]
  53. Szymanski, S.F.; Seman, M.T.; Wolden, C.A. Plasma and gas-phase characterization of a pulsed plasma-enhanced chemical vapor deposition system engineered for self-limiting growth of aluminum oxide thin films. Surf. Coat. Technol. 2007, 201, 8991–8997. [Google Scholar] [CrossRef]
  54. Nguyen, H.M.T.; Tang, H.-Y.; Huang, W.-F.; Lin, M.C. Mechanisms for reactions of trimethylaluminum with molecular oxygen and water. Comput. Theor. Chem. 2014, 1035, 39–43. [Google Scholar] [CrossRef]
  55. Belenguer, P.; Blondeau, J.P.; Boufendi, L.; Toogood, M.; Plain, A.; Bouchoule, A.; Laure, C.; Boeuf, J.P. Numerical and experimental diagnostics of rf discharges in pure and dusty argon. Phys. Rev. A 1992, 46, 7923–7933. [Google Scholar] [CrossRef] [PubMed]
  56. Lee, M.-H.; Cheng, C.-F.; Heine, V.; Klinowski, J. Distribution of tetrahedral and octahedral A1 sites in gamma alumina. Chem. Phys. Lett. 1997, 265, 673–676. [Google Scholar] [CrossRef]
  57. Zhou, R.-S.; Snyder, R.L. Structures and transformation mechanisms of the η, γ and θ transition aluminas. Acta Crystallogr. Sect. B 1991, 47, 617–630. [Google Scholar] [CrossRef]
  58. Bradley, S.M.; Hanna, J.V. 27Al and 23Na MAS NMR and Powder X-ray Diffraction Studies of Sodium Aluminate Speciation and the Mechanistics of Aluminum Hydroxide Precipitation upon Acid Hydrolysis. J. Am. Chem. Soc. 1994, 116, 7771–7783. [Google Scholar] [CrossRef]
  59. Boumaza, A.; Favaro, L.; Lédion, J.; Sattonnay, G.; Brubach, J.B.; Berthet, P.; Huntz, A.M.; Roy, P.; Tétot, R. Transition alumina phases induced by heat treatment of boehmite: An X-ray diffraction and infrared spectroscopy study. J. Solid State Chem. 2009, 182, 1171–1176. [Google Scholar] [CrossRef]
  60. Zagrajczuk, B.; Dziadek, M.; Olejniczak, Z.; Sulikowski, B.; Cholewa-Kowalska, K.; Laczka, M. Structural investigation of gel-derived materials from the SiO2Al2O3 system. J. Mol. Struct. 2018, 1167, 23–32. [Google Scholar] [CrossRef]
  61. Kuech, T.F.; Veuhoff, E.; Kuan, T.S.; Deline, V.; Potemski, R. The influence of growth chemistry on the MOVPE growth of GaAs and AlxGa1−xAs layers and heterostructures. J. Cryst. Growth 1986, 77, 257–271. [Google Scholar] [CrossRef]
  62. Kobayashi, N.; Makimoto, T. Reduced Carbon Contamination in OMVPE Grown GaAs and AlGaAs. Jpn. J. Appl. Phys. 1985, 24, L824. [Google Scholar] [CrossRef]
  63. Paparazzo, E. XPS analysis of iron aluminum oxide systems. Appl. Surf. Sci. 1986, 25, 1–12. [Google Scholar] [CrossRef]
  64. van den Brand, J.; Snijders, P.C.; Sloof, W.G.; Terryn, H.; De Wit, J.H.W. Acid−Base Characterization of Aluminum Oxide Surfaces with XPS. J. Phys. Chem. B 2004, 108, 6017–6024. [Google Scholar] [CrossRef]
  65. McCafferty, E.; Wightman, J.P. Determination of the Concentration of Surface Hydroxyl Groups on Metal Oxide Films by a Quantitative XPS Method. Surf. Interface Anal. 1998, 26, 549–564. [Google Scholar] [CrossRef]
  66. van den Brand, J.; Sloof, W.G.; Terryn, H.; De Wit, J.H.W. Correlation between hydroxyl fraction and O/Al atomic ratio as determined from XPS spectra of aluminium oxide layers. Surf. Interface Anal. 2004, 36, 81–88. [Google Scholar] [CrossRef]
  67. Pinto, H.P.; Nieminen, R.M.; Elliott, S.D. Ab Initio study of γ−Al2O3 surfaces. Phys. Rev. B 2004, 70, 125402. [Google Scholar] [CrossRef]
  68. Gresback, R.; Holman, Z.; Kortshagen, U. Nonthermal plasma synthesis of size-controlled, monodisperse, freestanding germanium nanocrystals. Appl. Phys. Lett. 2007, 91, 093119. [Google Scholar] [CrossRef]
  69. Xiong, Z.; Lanham, S.; Husmann, E.; Nelson, G.; Eslamisaray, M.A.; Polito, J.; Liu, Y.; Goree, J.; Thimsen, E.; Kushner, M.J.; et al. Particle trapping, size-filtering, and focusing in the nonthermal plasma synthesis of sub-10 nanometer particles. J. Phys. D Appl. Phys. 2022, 55, 235202. [Google Scholar] [CrossRef]
  70. Tavakoli, A.H.; Maram, P.S.; Widgeon, S.J.; Rufner, J.; van Benthem, K.; Ushakov, S.; Sen, S.; Navrotsky, A. Amorphous Alumina Nanoparticles: Structure, Surface Energy, and Thermodynamic Phase Stability. J. Phys. Chem. C 2013, 117, 17123–17130. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the nonthermal ICP setup for the synthesis of γ–Al2O3 nanocrystals.
Figure 1. Schematic diagram of the nonthermal ICP setup for the synthesis of γ–Al2O3 nanocrystals.
Nanomaterials 13 01627 g001
Figure 2. (a) XRD patterns of experimental (black) and reference (red) γ-Al2O3 nanocrystals (PDF number-29-1486). (b) FT-IR spectrum, (c) XPS survey scan, and (d,e) high-resolution Al 2p and O 1s spectra, respectively. Purple lines indicate background spectra.
Figure 2. (a) XRD patterns of experimental (black) and reference (red) γ-Al2O3 nanocrystals (PDF number-29-1486). (b) FT-IR spectrum, (c) XPS survey scan, and (d,e) high-resolution Al 2p and O 1s spectra, respectively. Purple lines indicate background spectra.
Nanomaterials 13 01627 g002
Figure 3. (a) 2D 27Al → 1H D-RINEPT NMR spectrum of the synthesized γ–Al2O3 nanocrystals recorded at B0 = 19.6 T with a 17.857 kHz MAS frequency and 448 μs of total SR412 dipolar recoupling applied to the 1H spins. The direct excitation 1H NMR spectrum is shown above the 2D 1H projection. (b) Comparison of a quantitative (upper) 10° tip-angle single-pulse (SP) (*) 27Al NMR spectrum with that of the (lower) 2D 27Al → 1H D-RINEPT 27Al projection.
Figure 3. (a) 2D 27Al → 1H D-RINEPT NMR spectrum of the synthesized γ–Al2O3 nanocrystals recorded at B0 = 19.6 T with a 17.857 kHz MAS frequency and 448 μs of total SR412 dipolar recoupling applied to the 1H spins. The direct excitation 1H NMR spectrum is shown above the 2D 1H projection. (b) Comparison of a quantitative (upper) 10° tip-angle single-pulse (SP) (*) 27Al NMR spectrum with that of the (lower) 2D 27Al → 1H D-RINEPT 27Al projection.
Nanomaterials 13 01627 g003
Figure 4. HAADF-STEM analysis of γ-alumina NCs. (a) Average-sized (~11 nm) (110)-oriented NC showing (111) and (001) facets. The outer atomic layer of (111) facets exhibit enhanced contrast due to excess Al3+ cations. (b) Larger (~16 nm) (110)-oriented NC showing (111) and (001) facets. The (110) facets have a stepped structure, exposing alternating (111) facets (red arrows). (c) Representative image of a collection of NCs. (d) Size distribution of 800 NCs. The mean μg and geometric standard deviations σg were estimated by fitting the histogram with a log-normal distribution.
Figure 4. HAADF-STEM analysis of γ-alumina NCs. (a) Average-sized (~11 nm) (110)-oriented NC showing (111) and (001) facets. The outer atomic layer of (111) facets exhibit enhanced contrast due to excess Al3+ cations. (b) Larger (~16 nm) (110)-oriented NC showing (111) and (001) facets. The (110) facets have a stepped structure, exposing alternating (111) facets (red arrows). (c) Representative image of a collection of NCs. (d) Size distribution of 800 NCs. The mean μg and geometric standard deviations σg were estimated by fitting the histogram with a log-normal distribution.
Nanomaterials 13 01627 g004
Figure 5. TEM images of γ-Al2O3 nanocrystals (a) with 0.25 mm orifice width at 3.8 Torr pressure, (b) with 0.5 mm orifice width at 2.5 Torr pressure, (c) with 1 mm orifice width at 1.3 Torr pressure (red arrow indicates the shown particle dimension), (d) respective particle size distributions, and (e) XRD patterns.
Figure 5. TEM images of γ-Al2O3 nanocrystals (a) with 0.25 mm orifice width at 3.8 Torr pressure, (b) with 0.5 mm orifice width at 2.5 Torr pressure, (c) with 1 mm orifice width at 1.3 Torr pressure (red arrow indicates the shown particle dimension), (d) respective particle size distributions, and (e) XRD patterns.
Nanomaterials 13 01627 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xiong, Z.; Andaraarachchi, H.P.; Held, J.T.; Dorn, R.W.; Jeong, Y.-J.; Rossini, A.; Kortshagen, U.R. Inductively Coupled Nonthermal Plasma Synthesis of Size-Controlled γ-Al2O3 Nanocrystals. Nanomaterials 2023, 13, 1627. https://doi.org/10.3390/nano13101627

AMA Style

Xiong Z, Andaraarachchi HP, Held JT, Dorn RW, Jeong Y-J, Rossini A, Kortshagen UR. Inductively Coupled Nonthermal Plasma Synthesis of Size-Controlled γ-Al2O3 Nanocrystals. Nanomaterials. 2023; 13(10):1627. https://doi.org/10.3390/nano13101627

Chicago/Turabian Style

Xiong, Zichang, Himashi P. Andaraarachchi, Jacob T. Held, Rick W. Dorn, Yong-Jin Jeong, Aaron Rossini, and Uwe R. Kortshagen. 2023. "Inductively Coupled Nonthermal Plasma Synthesis of Size-Controlled γ-Al2O3 Nanocrystals" Nanomaterials 13, no. 10: 1627. https://doi.org/10.3390/nano13101627

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop