Xiao Cheng Zeng
University of Nebraska–Lincoln
Network
Latest external collaboration on country level. Dive into details by clicking on the dots.
Publication
Featured researches published by Xiao Cheng Zeng.
Nature | 2001
Kenichiro Koga; G.T. Gao; Hideki Tanaka; Xiao Cheng Zeng
Following their discovery, carbon nanotubes have attracted interest not only for their unusual electrical and mechanical properties, but also because their hollow interior can serve as a nanometre-sized capillary, mould or template in material fabrication. The ability to encapsulate a material in a nanotube also offers new possibilities for investigating dimensionally confined phase transitions. Particularly intriguing is the conjecture that matter within the narrow confines of a carbon nanotube might exhibit a solid–liquid critical point beyond which the distinction between solid and liquid phases disappears. This unusual feature, which cannot occur in bulk material, would allow for the direct and continuous transformation of liquid matter into a solid. Here we report simulations of the behaviour of water encapsulated in carbon nanotubes that suggest the existence of a variety of new ice phases not seen in bulk ice, and of a solid–liquid critical point. Using carbon nanotubes with diameters ranging from 1.1 nm to 1.4 nm and applied axial pressures of 50 MPa to 500 MPa, we find that water can exhibit a first-order freezing transition to hexagonal and heptagonal ice nanotubes, and a continuous phase transformation into solid-like square or pentagonal ice nanotubes.
Journal of Physical Chemistry Letters | 2014
Jun Dai; Xiao Cheng Zeng
Phosphorene, a monolayer of black phosphorus, is promising for nanoelectronic applications not only because it is a natural p-type semiconductor but also because it possesses a layer-number-dependent direct bandgap (in the range of 0.3 to 1.5 eV). On basis of the density functional theory calculations, we investigate electronic properties of the bilayer phosphorene with different stacking orders. We find that the direct bandgap of the bilayers can vary from 0.78 to 1.04 eV with three different stacking orders. In addition, a vertical electric field can further reduce the bandgap to 0.56 eV (at the field strength 0.5 V/Å). More importantly, we find that when a monolayer of MoS2 is superimposed with the p-type AA- or AB-stacked bilayer phosphorene, the combined trilayer can be an effective solar-cell material with type-II heterojunction alignment. The power conversion efficiency is predicted to be ∼18 or 16% with AA- or AB-stacked bilayer phosphorene, higher than reported efficiencies of the state-of-the-art trilayer graphene/transition metal dichalcogenide solar cells.
Advanced Materials | 2015
Rui Dong; Yanjun Fang; Jungseok Chae; Jun Dai; Zhengguo Xiao; Qingfeng Dong; Yongbo Yuan; Andrea Centrone; Xiao Cheng Zeng; Jinsong Huang
Dr. R. Dong, Dr. Y. Fang, Z. Xiao, Dr. Q. Dong, Dr. Y. Yuan, Prof. X. C. Zeng, Prof. J. Huang Department of Mechanical and Materials Engineering University of Nebraska-Lincoln Lincoln , NE 68588 , USA E-mail: [email protected] Dr. J. Chae, Dr. A. Centrone Center for Nanoscale Science and Technology National Institute of Standards and Technology 100 Bureau Drive , Gaithersburg , MD 20899 , USA Dr. J. Chae Maryland Nanocenter University of Maryland College Park , MD 20742 , USA Dr. J. Dai, Prof. X. C. Zeng Department of Chemistry University of Nebraska-Lincoln Lincoln , NE 68588 , USA
ACS Nano | 2012
Xiaojun Wu; Jun Dai; Yu Zhao; Zhiwen Zhuo; Jinlong Yang; Xiao Cheng Zeng
Boron, a nearest-neighbor of carbon, is possibly the second element that can possess free-standing flat monolayer structures, evidenced by recent successful synthesis of single-walled and multiwalled boron nanotubes (MWBNTs). From an extensive structural search using the first-principles particle-swarm optimization (PSO) global algorithm, two boron monolayers (α(1)- and β(1)-sheet) are predicted to be the most stable α- and β-types of boron sheets, respectively. Both boron sheets possess greater cohesive energies than the state-of-the-art two-dimensional boron structures (by more than 60 meV/atom based on density functional theory calculation using PBE0 hybrid functional), that is, the α-sheet previously predicted by Tang and Ismail-Beigi and the g(1/8)- and g(2/15)-sheets (both belonging to the β-type) recently reported by Yakobson and co-workers. Moreover, the PBE0 calculation predicts that the α-sheet is a semiconductor, while the α(1)-, β(1)-, g(1/8)-, and g(2/15)-sheets are all metals. When two α(1) monolayers are stacked on top each other, the bilayer α(1)-sheet remains flat with an optimal interlayer distance of ~3.62 Å, which is close to the measured interlayer distance (~3.2 Å) in MWBNTs.
Physical Chemistry Chemical Physics | 2012
Peng Lu; Xiaojun Wu; Wanlin Guo; Xiao Cheng Zeng
We investigate the strain-dependent electronic and magnetic properties of two-dimensional (2D) monolayer and bilayer MoS(2), as well as 1D MoS(2) nanoribbons and nanotubes using first-principles calculations. For 2D monolayer MoS(2) subjected to isotropic or uniaxial tensile strain, the direct band gap of MoS(2) changes to an indirect gap that decreases monotonically with increasing strain; while under the compressive strain, the original direct band gap is enlarged first, followed by gap reduction when the strain is beyond -2%. The effect of isotropic strain is even stronger than that of uniaxial strain. For bilayer MoS(2) subjected to isotropic tensile strain, its indirect gap reduces monotonically to zero at strain about 6%; while under the isotropic compressive strain, its indirect gap increases first and then reduces and turns into direct gap when the strain is beyond -4%. For strained 1D metallic zigzag MoS(2) nanoribbons, the net magnetic moment increases slightly with axial strain from about -5% to 5%, but drops to zero when the compressive strain is beyond -5% or increases with a power law beyond 5%. For 1D armchair MoS(2) nanotubes, tensile or compressive axial strain reduces or enlarges the band gap linearly, and the gap can be fully closed for nanotubes with relatively small diameter or under large tensile strain. For zigzag MoS(2) nanotubes, the strain effect becomes nonlinear and the tensile strain can reduce the band gap, whereas compressive strain can initially enlarge the band gap and then decrease it. The strain induced change in projected orbitals energy of Mo and the coupling between the Mo atom d orbital and the S atom p orbital are analyzed to explain the strong strain effect on the band gap and magnetic properties.
Proceedings of the National Academy of Sciences of the United States of America | 2009
Takahiro Koishi; Kenji Yasuoka; Shigenori Fujikawa; Toshikazu Ebisuzaki; Xiao Cheng Zeng
Water droplets on rugged hydrophobic surfaces typically exhibit one of the following two states: (i) the Wenzel state [Wenzel RN (1936) Ind Eng Chem 28:988–994] in which water droplets are in full contact with the rugged surface (referred as the wetted contact) or (ii) the Cassie state [Cassie, ABD, Baxter S (1944) Trans Faraday Soc 40:546–551] in which water droplets are in contact with peaks of the rugged surface as well as the “air pockets” trapped between surface grooves (the composite contact). Here, we show large-scale molecular dynamics simulation of transition between Wenzel state and Cassie state of water droplets on a periodic nanopillared hydrophobic surface. Physical conditions that can strongly affect the transition include the height of nanopillars, the spacing between pillars, the intrinsic contact angle, and the impinging velocity of water nanodroplet (“raining” simulation). There exists a critical pillar height beyond which water droplets on the pillared surface can be either in the Wenzel state or in the Cassie state, depending on their initial location. The free-energy barrier separating the Wenzel and Cassie state was computed on the basis of a statistical-mechanics method and kinetic raining simulation. The barrier ranges from a few tenths of kBT0 (where kB is the Boltzmann constant, and T0 is the ambient temperature) for a rugged surface at the critical pillar height to ≈8 kBT0 for the surface with pillar height greater than the length scale of water droplets. For a highly rugged surface, the barrier from the Wenzel-to-Cassie state is much higher than from Cassie-to-Wenzel state. Hence, once a droplet is trapped deeply inside the grooves, it would be much harder to relocate on top of high pillars.
Journal of Chemical Physics | 1991
Xiao Cheng Zeng; David W. Oxtoby
We have applied a nonclassical density functional theory of nucleation to the gas–liquid and liquid–gas transitions of a Lennard‐Jones fluid. For the liquid‐to‐gas transition (cavitation) deviations from classical theory are extremely large: 15 orders of magnitude in rates. For the gas‐to‐liquid transition (condensation) the deviations are smaller in magnitude but still systematic. Our nonclassical theory agrees with classical theory in its prediction of the dependence of nucleation rates on supersaturation, but it differs in its prediction of temperature dependence. Good agreement is found between our theory and experiments on condensation nucleation of nonane.
Nature | 2000
Kenichiro Koga; Hideki Tanaka; Xiao Cheng Zeng
Supercooled water and amorphous ice have a rich metastable phase behaviour. In addition to transitions between high- and low-density amorphous solids, and between high- and low-density liquids, a fragile-to-strong liquid transition has recently been proposed, and supported by evidence from the behaviour of deeply supercooled bilayer water confined in hydrophilic slit pores. Here we report evidence from molecular dynamics simulations for another type of first-order phase transition—a liquid-to-bilayer amorphous transition—above the freezing temperature of bulk water at atmospheric pressure. This transition occurs only when water is confined in a hydrophobic slit pore with a width of less than one nanometre. On cooling, the confined water, which has an imperfect random hydrogen-bonded network, transforms into a bilayer amorphous phase with a perfect network (owing to the formation of various hydrogen-bonded polygons) but no long-range order. The transition shares some characteristics with those observed in tetrahedrally coordinated substances such as liquid silicon, liquid carbon and liquid phosphorus.
Journal of Chemical Physics | 2009
Soohaeng Yoo; Xiao Cheng Zeng; Sotiris S. Xantheas
The melting temperature (T(m)) of ice I(h) was determined from constant enthalpy and pressure (NPH) Born-Oppenheimer molecular dynamics simulations to be 417+/-3 K for the Perdew-Burke-Ernzerhof and 411+/-4 K for the Becke-Lee-Yang-Parr density functionals using a coexisting ice (I(h))-liquid phase at constant pressures of P=2500 and 10,000 bar and a density rho=1 g/cm(3), respectively. This suggests that ambient condition simulations at rho=1 g/cm(3) will rather describe a supercooled state that is overstructured when compared to liquid water.
Proceedings of the National Academy of Sciences of the United States of America | 2006
Jaeil Bai; Jun Wang; Xiao Cheng Zeng
We report six phases of high-density nano-ice predicted to form within carbon nanotubes (CNTs) at high pressure. High-density nano-ice self-assembled within smaller-diameter CNT (17,0) exhibits a double-walled helical structure where the outer wall consists of four double-stranded helixes, which resemble a DNA double helix, and the inner wall is a quadruple-stranded helix. Four other double-walled nano-ices, self-assembled respectively in two larger-diameter CNTs (20,0 and 22,0), display tubular structure. Within CNT (24,0), the confined water can freeze spontaneously into a triple-walled helical nano-ice where the outer wall is an 18-stranded helix and the middle and inner walls are hextuple-stranded helixes.