Journal articles on the topic 'Transition voltage'

To see the other types of publications on this topic, follow the link: Transition voltage.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Transition voltage.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Ohtsuki, Hiroshi, Akihiko Yokoyama, and Yasuji Swkine. "Voltage Stability Transition during Voltage Collapse." IEEJ Transactions on Power and Energy 112, no. 7 (1992): 615–21. http://dx.doi.org/10.1541/ieejpes1990.112.7_615.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Ledwell, Jennifer L., and Richard W. Aldrich. "Mutations in the S4 Region Isolate the Final Voltage-dependent Cooperative Step in Potassium Channel Activation." Journal of General Physiology 113, no. 3 (March 1, 1999): 389–414. http://dx.doi.org/10.1085/jgp.113.3.389.

Full text
Abstract:
Charged residues in the S4 transmembrane segment play a key role in determining the sensitivity of voltage-gated ion channels to changes in voltage across the cell membrane. However, cooperative interactions between subunits also affect the voltage dependence of channel opening, and these interactions can be altered by making substitutions at uncharged residues in the S4 region. We have studied the activation of two mutant Shaker channels that have different S4 amino acid sequences, ILT (V369I, I372L, and S376T) and Shaw S4 (the S4 of Drosophila Shaw substituted into Shaker), and yet have very similar ionic current properties. Both mutations affect cooperativity, making a cooperative transition in the activation pathway rate limiting and shifting it to very positive voltages, but analysis of gating and ionic current recordings reveals that the ILT and Shaw S4 mutant channels have different activation pathways. Analysis of gating currents suggests that the dominant effect of the ILT mutation is to make the final cooperative transition to the open state of the channel rate limiting in an activation pathway that otherwise resembles that of Shaker. The charge movement associated with the final gating transition in ILT activation can be measured as an isolated component of charge movement in the voltage range of channel opening and accounts for 13% (∼1.8 e0) of the total charge moved in the ILT activation pathway. The remainder of the ILT gating charge (87%) moves at negative voltages, where channels do not open, and confirms the presence of Shaker-like conformational changes between closed states in the activation pathway. In contrast to ILT, the activation pathway of Shaw S4 seems to involve a single cooperative charge-moving step between a closed and an open state. We cannot detect any voltage-dependent transitions between closed states for Shaw S4. Restoring basic residues that are missing in Shaw S4 (R1, R2, and K7) rescues charge movement between closed states in the activation pathway, but does not alter the voltage dependence of the rate-limiting transition in activation.
APA, Harvard, Vancouver, ISO, and other styles
3

Zagotta, W. N., T. Hoshi, J. Dittman, and R. W. Aldrich. "Shaker potassium channel gating. II: Transitions in the activation pathway." Journal of General Physiology 103, no. 2 (February 1, 1994): 279–319. http://dx.doi.org/10.1085/jgp.103.2.279.

Full text
Abstract:
Voltage-dependent gating behavior of Shaker potassium channels without N-type inactivation (ShB delta 6-46) expressed in Xenopus oocytes was studied. The voltage dependence of the steady-state open probability indicated that the activation process involves the movement of the equivalent of 12-16 electronic charges across the membrane. The sigmoidal kinetics of the activation process, which is maintained at depolarized voltages up to at least +100 mV indicate the presence of at least five sequential conformational changes before opening. The voltage dependence of the gating charge movement suggested that each elementary transition involves 3.5 electronic charges. The voltage dependence of the forward opening rate, as estimated by the single-channel first latency distribution, the final phase of the macroscopic ionic current activation, the ionic current reactivation and the ON gating current time course, showed movement of the equivalent of 0.3 to 0.5 electronic charges were associated with a large number of the activation transitions. The equivalent charge movement of 1.1 electronic charges was associated with the closing conformational change. The results were generally consistent with models involving a number of independent and identical transitions with a major exception that the first closing transition is slower than expected as indicated by tail current and OFF gating charge measurements.
APA, Harvard, Vancouver, ISO, and other styles
4

Huisman, Everardus H., Constant M. Guédon, Bart J. van Wees, and Sense Jan van der Molen. "Interpretation of Transition Voltage Spectroscopy." Nano Letters 9, no. 11 (November 11, 2009): 3909–13. http://dx.doi.org/10.1021/nl9021094.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Lúcio da Silva Martins, Mário, Jumar Luís Russi, José Renes Pinheiro, and Hélio Leães Hey. "Zero-current Zero-voltage Transition Pwm Converters With Magnetically Coupled Auxiliary Circuit." Eletrônica de Potência 13, no. 4 (November 1, 2008): 201–8. http://dx.doi.org/10.18618/rep.2008.4.201208.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Guichao Hua, Ching-Shan Leu, Yimin Jiang, and F. C. Y. Lee. "Novel zero-voltage-transition PWM converters." IEEE Transactions on Power Electronics 9, no. 2 (March 1994): 213–19. http://dx.doi.org/10.1109/63.286814.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Chan, Y. H., C. C. Lim, K. T. Lau, and S. H. Foo. "Low power critical voltage transition logic." Microelectronics International 23, no. 3 (September 2006): 3–8. http://dx.doi.org/10.1108/13565360610680695.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Richter, C., Z. Wu, and L. Menon. "Pattern Formation in Nanoporous Titania Templates." Journal of Nanoscience and Nanotechnology 7, no. 2 (February 1, 2007): 704–7. http://dx.doi.org/10.1166/jnn.2007.122.

Full text
Abstract:
We have carried out a systematic investigation into the formation of nanoscaled patterns in titania (TiO2) templates under dc anodization of Ti in HF acid. At lower acid concentrations (around 0.5 wt% HF) either pores or tubes form at the surface of anodized titanium foil. The pores or nanotubes are separated from the bottom Ti layer by a thin barrier layer of TiO2. The critical voltage where the transition from pores to tubes occurs has been determined. It is observed that the transition voltage shift towards higher voltages as acid concentration is increased, with pore formation disappearing altogether at high acid concentrations. We have also carried out a systematic investigation into the dependence of pore and tube parameters on the applied dc anodization voltage. Our results indicate that the barrier layer thickness, pore and tube length increase as a function of applied voltage.
APA, Harvard, Vancouver, ISO, and other styles
9

Zagotta, W. N., and R. W. Aldrich. "Voltage-dependent gating of Shaker A-type potassium channels in Drosophila muscle." Journal of General Physiology 95, no. 1 (January 1, 1990): 29–60. http://dx.doi.org/10.1085/jgp.95.1.29.

Full text
Abstract:
The voltage-dependent gating mechanism of A1-type potassium channels coded for by the Shaker locus of Drosophila was studied using macroscopic and single-channel recording techniques on embryonic myotubes in primary culture. From a kinetic analysis of data from single A1 channels, we have concluded that all of the molecular transitions after first opening, including the inactivation transition, are voltage independent and therefore not associated with charge movement through the membrane. In contrast, at least some of the activation transitions leading to first opening are considerably voltage dependent and account for all of the voltage dependence seen in the macroscopic currents. This mechanism is similar in many ways to that of vertebrate neuronal voltage-sensitive sodium channels, and together with the sequence similarities in the S4 region suggests a conserved mechanism for voltage-dependent gating among channels with different selectivities. By testing independent and coupled models for activation and inactivation we have determined that the final opening transition and inactivation are not likely to arise from the independent action of multiple subunits, each with simple gating transitions, but rather come about through their aggregate properties. A partially coupled model accurately reproduces all of the single-channel and macroscopic data. This model will provide a framework on which to organize and understand alterations in gating that occur in Shaker variants and mutants.
APA, Harvard, Vancouver, ISO, and other styles
10

Horrigan, Frank T., and Richard W. Aldrich. "Coupling between Voltage Sensor Activation, Ca2+ Binding and Channel Opening in Large Conductance (BK) Potassium Channels." Journal of General Physiology 120, no. 3 (August 26, 2002): 267–305. http://dx.doi.org/10.1085/jgp.20028605.

Full text
Abstract:
To determine how intracellular Ca2+ and membrane voltage regulate the gating of large conductance Ca2+-activated K+ (BK) channels, we examined the steady-state and kinetic properties of mSlo1 ionic and gating currents in the presence and absence of Ca2+ over a wide range of voltage. The activation of unliganded mSlo1 channels can be accounted for by allosteric coupling between voltage sensor activation and the closed (C) to open (O) conformational change (Horrigan, F.T., and R.W. Aldrich. 1999. J. Gen. Physiol. 114:305–336; Horrigan, F.T., J. Cui, and R.W. Aldrich. 1999. J. Gen. Physiol. 114:277–304). In 0 Ca2+, the steady-state gating charge-voltage (QSS-V) relationship is shallower and shifted to more negative voltages than the conductance-voltage (GK-V) relationship. Calcium alters the relationship between Q-V and G-V, shifting both to more negative voltages such that they almost superimpose in 70 μM Ca2+. This change reflects a differential effect of Ca2+ on voltage sensor activation and channel opening. Ca2+ has only a small effect on the fast component of ON gating current, indicating that Ca2+ binding has little effect on voltage sensor activation when channels are closed. In contrast, open probability measured at very negative voltages (less than −80 mV) increases more than 1,000-fold in 70 μM Ca2+, demonstrating that Ca2+ increases the C-O equilibrium constant under conditions where voltage sensors are not activated. Thus, Ca2+ binding and voltage sensor activation act almost independently, to enhance channel opening. This dual-allosteric mechanism can reproduce the steady-state behavior of mSlo1 over a wide range of conditions, with the assumption that activation of individual Ca2+ sensors or voltage sensors additively affect the energy of the C-O transition and that a weak interaction between Ca2+ sensors and voltage sensors occurs independent of channel opening. By contrast, macroscopic IK kinetics indicate that Ca2+ and voltage dependencies of C-O transition rates are complex, leading us to propose that the C-O conformational change may be described by a complex energy landscape.
APA, Harvard, Vancouver, ISO, and other styles
11

Chen, Shan, Jing Wang, Lei Zhou, Meena S. George, and Steven A. Siegelbaum. "Voltage Sensor Movement and cAMP Binding Allosterically Regulate an Inherently Voltage-independent Closed−Open Transition in HCN Channels." Journal of General Physiology 129, no. 2 (January 29, 2007): 175–88. http://dx.doi.org/10.1085/jgp.200609585.

Full text
Abstract:
The hyperpolarization-activated cyclic nucleotide-modulated cation (HCN) channels are regulated by both membrane voltage and the binding of cyclic nucleotides to a cytoplasmic, C-terminal cyclic nucleotide-binding domain (CNBD). Here we have addressed the mechanism of this dual regulation for HCN2 channels, which activate with slow kinetics that are strongly accelerated by cAMP, and HCN1 channels, which activate with rapid kinetics that are weakly enhanced by cAMP. Surprisingly, we find that the rate of opening of HCN2 approaches a maximal value with extreme hyperpolarization, indicating the presence of a voltage-independent kinetic step in the opening process that becomes rate limiting at very negative potentials. cAMP binding enhances the rate of this voltage-independent opening step. In contrast, the rate of opening of HCN1 is much greater than that of HCN2 and does not saturate with increasing hyperpolarization over the voltage range examined. Domain-swapping chimeras between HCN1 and HCN2 reveal that the S4–S6 transmembrane region largely determines the limiting rate in opening kinetics at negative voltages. Measurements of HCN2 tail current kinetics also reveal a voltage-independent closing step that becomes rate limiting at positive voltages; the rate of this closing step is decreased by cAMP. These results are consistent with a cyclic allosteric model in which a closed–open transition that is inherently voltage independent is subject to dual allosteric regulation by voltage sensor movement and cAMP binding. This mechanism accounts for several properties of HCN channel gating and has potentially important physiological implications.
APA, Harvard, Vancouver, ISO, and other styles
12

Artés, Juan M., Montserrat López-Martínez, Arnaud Giraudet, Ismael Díez-Pérez, Fausto Sanz, and Pau Gorostiza. "Current–Voltage Characteristics and Transition Voltage Spectroscopy of Individual Redox Proteins." Journal of the American Chemical Society 134, no. 50 (December 4, 2012): 20218–21. http://dx.doi.org/10.1021/ja3080242.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Bui, Duong Minh, Shi-Lin Chen, Keng-Yu Lien, and Jheng-Lun Jiang. "A Generalised Fault Protection Structure Proposed for Uni-grounded Low-Voltage AC Microgrids." International Journal of Emerging Electric Power Systems 17, no. 2 (April 1, 2016): 69–89. http://dx.doi.org/10.1515/ijeeps-2015-0151.

Full text
Abstract:
Abstract This paper presents three main configurations of uni-grounded low-voltage AC microgrids. Transient situations of a uni-grounded low-voltage (LV) AC microgrid (MG) are simulated through various fault tests and operation transition tests between grid-connected and islanded modes. Based on transient simulation results, available fault protection methods are proposed for main and back-up protection of a uni-grounded AC microgrid. In addition, concept of a generalised fault protection structure of uni-grounded LVAC MGs is mentioned in the paper. As a result, main contributions of the paper are: (i) definition of different uni-grounded LVAC MG configurations; (ii) analysing transient responses of a uni-grounded LVAC microgrid through line-to-line faults, line-to-ground faults, three-phase faults and a microgrid operation transition test, (iii) proposing available fault protection methods for uni-grounded microgrids, such as: non-directional or directional overcurrent protection, under/over voltage protection, differential current protection, voltage-restrained overcurrent protection, and other fault protection principles not based on phase currents and voltages (e.g. total harmonic distortion detection of currents and voltages, using sequence components of current and voltage, 3I0 or 3V0 components), and (iv) developing a generalised fault protection structure with six individual protection zones to be suitable for different uni-grounded AC MG configurations.
APA, Harvard, Vancouver, ISO, and other styles
14

Nguyen, Thanh Hai, Tan Luong Van, Asif Nawaz, and Ammar Natsheh. "Feedback Linearization-Based Control Strategy for Interlinking Inverters of Hybrid AC/DC Microgrids with Seamless Operation Mode Transition." Energies 14, no. 18 (September 7, 2021): 5613. http://dx.doi.org/10.3390/en14185613.

Full text
Abstract:
This study proposes an advanced control scheme for the interlinking inverters of the hybrid AC/DC microgrids, which facilitates a seamless transition between grid-connected and stand-alone/islanding modes for the microgrid. Due to a nonlinear relationship between the terminal voltages of the voltage-source inverter (VSI) interfacing through inductor–capacitor (LC) filters with the grid voltages and currents, a feedback linearization technique (FLT) is employed to control the interlinking VSI under both grid-connected and islanding operations. The FLT-based current controllers are applied in the grid-connected mode, in which they adjust the power exchange between the DC and AC subgrids and mitigate the distortion of the grid currents produced by nonlinear loads. Under the stand-alone operation, the AC bus voltages are directly regulated by the FLT-voltage controllers of the interlinking VSI. In order to reduce the inrush currents and voltage overshot at the instant of mode switching, the FLT-based controllers are performed all the time regardless of the operating modes, where the voltage references for the VSI are not changed abruptly. The control performance of the VSI is highly satisfactory with low-transient overshoot values of the voltages and currents and low total harmonic distortion (THD) values of the grid currents and AC bus voltages are about 3.5% and 2.7%, respectively, under the nonlinear load condition. The validity of the new control strategy is verified by the simulation work, which investigates the operation of a hybrid AC/DC microgrid.
APA, Harvard, Vancouver, ISO, and other styles
15

Abdel-Fattah, E., and Omar F. Farag. "Alpha to gamma mode transition in hydrogen capacitive radio-frequency discharge." Canadian Journal of Physics 91, no. 12 (December 2013): 1062–67. http://dx.doi.org/10.1139/cjp-2013-0144.

Full text
Abstract:
Electron energy distribution functions (EEDFs) were measured with increasing discharge voltages in hydrogen capacitively coupled plasmas by means of radio-frequency compensated Langmuir probe. The results are compared with EEDF in argon plasmas. It was found that, in the hydrogen capacitive discharge, abnormally low-energy electrons became highly populated and the EEDF evolved to a non-Maxwellian distribution as the discharge voltage was increased. This voltage dependence of the EEDF in the hydrogen is contrary to argon capacitively coupled plasma, where at high discharge voltage, low-energy electrons are significantly thermalized due to γ heating and the EEDF evolves to the Maxwellian distribution. The highly populated low-energy electrons at high gas pressure, which was not observed in capacitively coupled argon plasma, show that the γ heating mechanism is somehow inefficient in terms of the molecular gas in capacitive discharges. It appears that this inefficient γ heating seems to be attributed to an efficient vibrational excitation in hydrogen capacitive plasma.
APA, Harvard, Vancouver, ISO, and other styles
16

Horrigan, Frank T., Jianmin Cui, and Richard W. Aldrich. "Allosteric Voltage Gating of Potassium Channels I." Journal of General Physiology 114, no. 2 (August 1, 1999): 277–304. http://dx.doi.org/10.1085/jgp.114.2.277.

Full text
Abstract:
Activation of large conductance Ca2+-activated K+ channels is controlled by both cytoplasmic Ca2+ and membrane potential. To study the mechanism of voltage-dependent gating, we examined mSlo Ca2+-activated K+ currents in excised macropatches from Xenopus oocytes in the virtual absence of Ca2+ (<1 nM). In response to a voltage step, IK activates with an exponential time course, following a brief delay. The delay suggests that rapid transitions precede channel opening. The later exponential time course suggests that activation also involves a slower rate-limiting step. However, the time constant of IK relaxation [τ(IK)] exhibits a complex voltage dependence that is inconsistent with models that contain a single rate limiting step. τ(IK) increases weakly with voltage from −500 to −20 mV, with an equivalent charge (z) of only 0.14 e, and displays a stronger voltage dependence from +30 to +140 mV (z = 0.49 e), which then decreases from +180 to +240 mV (z = −0.29 e). Similarly, the steady state GK–V relationship exhibits a maximum voltage dependence (z = 2 e) from 0 to +100 mV, and is weakly voltage dependent (z ≅ 0.4 e) at more negative voltages, where Po = 10−5–10−6. These results can be understood in terms of a gating scheme where a central transition between a closed and an open conformation is allosterically regulated by the state of four independent and identical voltage sensors. In the absence of Ca2+, this allosteric mechanism results in a gating scheme with five closed (C) and five open (O) states, where the majority of the channel's voltage dependence results from rapid C–C and O–O transitions, whereas the C–O transitions are rate limiting and weakly voltage dependent. These conclusions not only provide a framework for interpreting studies of large conductance Ca2+-activated K+ channel voltage gating, but also have important implications for understanding the mechanism of Ca2+ sensitivity.
APA, Harvard, Vancouver, ISO, and other styles
17

Smith-Maxwell, Catherine J., Jennifer L. Ledwell, and Richard W. Aldrich. "Uncharged S4 Residues and Cooperativity in Voltage-dependent Potassium Channel Activation." Journal of General Physiology 111, no. 3 (March 1, 1998): 421–39. http://dx.doi.org/10.1085/jgp.111.3.421.

Full text
Abstract:
Substitution of the S4 of Shaw into Shaker alters cooperativity in channel activation by slowing a cooperative transition late in the activation pathway. To determine the amino acids responsible for the functional changes in Shaw S4, we created several mutants by substituting amino acids from Shaw S4 into Shaker. The S4 amino acid sequences of Shaker and Shaw S4 differ at 11 positions. Simultaneous substitution of just three noncharged residues from Shaw S4 into Shaker (V369I, I372L, S376T; ILT) reproduces the kinetic and voltage-dependent properties of Shaw S4 channel activation. These substitutions cause very small changes in the structural and chemical properties of the amino acid side chains. In contrast, substituting the positively charged basic residues in the S4 of Shaker with neutral or negative residues from the S4 of Shaw S4 does not reproduce the shallow voltage dependence or other properties of Shaw S4 opening. Macroscopic ionic currents for ILT could be fit by modifying a single set of transitions in a model for Shaker channel gating (Zagotta, W.N., T. Hoshi, and R.W. Aldrich. 1994. J. Gen. Physiol. 103:321–362). Changing the rate and voltage dependence of a final cooperative step in activation successfully reproduces the kinetic, steady state, and voltage-dependent properties of ILT ionic currents. Consistent with the model, ILT gating currents activate at negative voltages where the channel does not open and, at more positive voltages, they precede the ionic currents, confirming the existence of voltage-dependent transitions between closed states in the activation pathway. Of the three substitutions in ILT, the I372L substitution is primarily responsible for the changes in cooperativity and voltage dependence. These results suggest that noncharged residues in the S4 play a crucial role in Shaker potassium channel gating and that small steric changes in these residues can lead to large changes in cooperativity within the channel protein.
APA, Harvard, Vancouver, ISO, and other styles
18

Smith-Maxwell, Catherine J., Jennifer L. Ledwell, and Richard W. Aldrich. "Role of the S4 in Cooperativity of Voltage-dependent Potassium Channel Activation." Journal of General Physiology 111, no. 3 (March 1, 1998): 399–420. http://dx.doi.org/10.1085/jgp.111.3.399.

Full text
Abstract:
Charged residues in the S4 transmembrane segment of voltage-gated cation channels play a key role in opening channels in response to changes in voltage across the cell membrane. However, the molecular mechanism of channel activation is not well understood. To learn more about the role of the S4 in channel gating, we constructed chimeras in which S4 segments from several divergent potassium channels, Shab, Shal, Shaw, and Kv3.2, were inserted into a Shaker potassium channel background. These S4 donor channels have distinctly different voltage-dependent gating properties and S4 amino acid sequences. None of the S4 chimeras have the gating behavior of their respective S4 donor channels. The conductance–voltage relations of all S4 chimeras are shifted to more positive voltages and the slopes are decreased. There is no consistent correlation between the nominal charge content of the S4 and the slope of the conductance–voltage relation, suggesting that the mutations introduced by the S4 chimeras may alter cooperative interactions in the gating process. We compared the gating behavior of the Shaw S4 chimera with its parent channels, Shaker and Shaw, in detail. The Shaw S4 substitution alters activation gating profoundly without introducing obvious changes in other channel functions. Analysis of the voltage-dependent gating kinetics suggests that the dominant effect of the Shaw S4 substitution is to alter a single cooperative transition late in the activation pathway, making it rate limiting. This interpretation is supported further by studies of channels assembled from tandem heterodimer constructs with both Shaker and Shaw S4 subunits. Activation gating in the heterodimer channels can be predicted from the properties of the homotetrameric channels only if it is assumed that the mutations alter a cooperative transition in the activation pathway rather than independent transitions.
APA, Harvard, Vancouver, ISO, and other styles
19

Czyz, Piotr, Panteleimon Papamanolis, Francesc Trunas Bruguera, Thomas Guillod, Florian Krismer, Vladan Lazarevic, Jonas Huber, and Johann W. Kolar. "Load-Independent Voltage Balancing of Multi-Level Flying Capacitor Converters in Quasi-2-Level Operation." Electronics 10, no. 19 (October 2, 2021): 2414. http://dx.doi.org/10.3390/electronics10192414.

Full text
Abstract:
Quasi-2-level (Q2L) operation of multi-level bridge-legs, especially of flying-capacitor converters (FCC), is an interesting option for realizing single-cell power conversion in applications whose system voltages exceed the ratings of available power semiconductors. To ensure equal voltage sharing among a Q2L-FCC’s switches, the voltages of a Q2L-FCC’s minimized flying capacitors (FCs) must always be balanced. Thus, we propose a concept for load-independent FC voltage balancing: For non-zero load current, we use a model predictive control (MPC) approach to identify the commutation sequence of the individual switches within a Q2L transition that minimizes the FC or cell voltage errors. In case of zero load current, we employ a novel MPC-based approach using cell multiple switching (CMS), i.e., the insertion of additional zero-current commutations within a Q2L transition, to exchange charge between the FCs via the charging currents of the switches’ parasitic capacitances. Experiments with a 5-level FCC half-bridge demonstrator confirm the validity of the derived models and verify the performance of the proposed load-independent balancing concept.
APA, Harvard, Vancouver, ISO, and other styles
20

Guo, Shaoyin, Gang Zhou, and Nongjian Tao. "Single Molecule Conductance, Thermopower, and Transition Voltage." Nano Letters 13, no. 9 (August 6, 2013): 4326–32. http://dx.doi.org/10.1021/nl4021073.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Jung-Goo Cho, Ju-Won Baek, Geun-Hie Rim, and Iouri Kang. "Novel zero-voltage-transition PWM multiphase converters." IEEE Transactions on Power Electronics 13, no. 1 (January 1998): 152–59. http://dx.doi.org/10.1109/63.654970.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Lu, Xinying, and Fang Yuan. "Breakdown voltage and transition zone of concrete." Cement and Concrete Research 30, no. 4 (April 2000): 643–44. http://dx.doi.org/10.1016/s0008-8846(00)00207-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Tseng, C. J., and C. L. Chen. "Comparisons of zero-voltage-transition Cúk converters." IEE Proceedings - Electric Power Applications 146, no. 4 (1999): 433. http://dx.doi.org/10.1049/ip-epa:19990222.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Moo, Chin Sien, Yu Jen Chen, Hung Liang Cheng, and Yao Ching Hsieh. "Twin-Buck Converter With Zero-Voltage Transition." IEEE Transactions on Industrial Electronics 58, no. 6 (June 2011): 2366–71. http://dx.doi.org/10.1109/tie.2010.2069072.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Bennett, Neil, Gengzhao Xu, Louisa J. Esdaile, Harry L. Anderson, J. Emyr Macdonald, and Martin Elliott. "Transition Voltage Spectroscopy of Porphyrin Molecular Wires." Small 6, no. 22 (October 20, 2010): 2604–11. http://dx.doi.org/10.1002/smll.201001046.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

Jayawardana, Chamithri, and Brett L. Lucht. "Difluorophosphoric Acid Generation and Crossover Reactions in LiNixCoyMnzO2 Cathodes." ECS Meeting Abstracts MA2022-01, no. 2 (July 7, 2022): 263. http://dx.doi.org/10.1149/ma2022-012263mtgabs.

Full text
Abstract:
Layered LiNixCoyMnzO2 (NCM) materials are one of the most widely utilized cathode materials for Lithium-ion batteries (LIB) with their high theoretical capacity and relatively high voltage window. Increasing the Nickel content and the cycling voltage can increase the capacity of these cells at a lower cost. But these high Nickel cathodes cycled to a higher upper cut off voltage have poor cycling stability. The cause of this rapid capacity decrease has been attributed to decomposition of the anode solid electrolyte interphase (SEI) from transition metal deposition on the anode surface and impedance growth on the cathode surface. Recent studies have also attributed this to acidic species generated from oxidative decomposition of the electrolyte and their crossover reactions degrading the SEI, particularly the Difluorophosphoric acid. In this study the role of transition metal dissolution and Difluorophosphoric acid generation has been investigated through various quantitative analysis techniques including, ICP-MS, XPS, quantitative solution NMR, and EIS. Full cells were constructed using graphite as the negative electrode and NCM cathodes with different Nickel contents as the positive electrode. These cells contain the same base electrolyte formulation (1.2M LiPF6 and EC: EMC, 3:7) and they were cycled at different temperatures and voltages to investigate the correlation between acid and transition metal concentration, cell voltage and capacity fade.
APA, Harvard, Vancouver, ISO, and other styles
27

Qu, Ke-qing, Guo-cheng Chen, Cheng-bo Sun, Jie Li, and Guo-xiang Wu. "A zero-voltage transition for three-phase double-PWM voltage-fed converter." Journal of Shanghai University (English Edition) 11, no. 5 (October 2007): 496–501. http://dx.doi.org/10.1007/s11741-007-0511-1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Delemotte, Lucie, Marina A. Kasimova, Michael L. Klein, Mounir Tarek, and Vincenzo Carnevale. "Free-energy landscape of ion-channel voltage-sensor–domain activation." Proceedings of the National Academy of Sciences 112, no. 1 (December 22, 2014): 124–29. http://dx.doi.org/10.1073/pnas.1416959112.

Full text
Abstract:
Voltage sensor domains (VSDs) are membrane-bound protein modules that confer voltage sensitivity to membrane proteins. VSDs sense changes in the transmembrane voltage and convert the electrical signal into a conformational change called activation. Activation involves a reorganization of the membrane protein charges that is detected experimentally as transient currents. These so-called gating currents have been investigated extensively within the theoretical framework of so-called discrete-state Markov models (DMMs), whereby activation is conceptualized as a series of transitions across a discrete set of states. Historically, the interpretation of DMM transition rates in terms of transition state theory has been instrumental in shaping our view of the activation process, whose free-energy profile is currently envisioned as composed of a few local minima separated by steep barriers. Here we use atomistic level modeling and well-tempered metadynamics to calculate the configurational free energy along a single transition from first principles. We show that this transition is intrinsically multidimensional and described by a rough free-energy landscape. Remarkably, a coarse-grained description of the system, based on the use of the gating charge as reaction coordinate, reveals a smooth profile with a single barrier, consistent with phenomenological models. Our results bridge the gap between microscopic and macroscopic descriptions of activation dynamics and show that choosing the gating charge as reaction coordinate masks the topological complexity of the network of microstates participating in the transition. Importantly, full characterization of the latter is a prerequisite to rationalize modulation of this process by lipids, toxins, drugs, and genetic mutations.
APA, Harvard, Vancouver, ISO, and other styles
29

Teale, C., J. Sherman, and J. Kitching. "Degenerate two-photon Rydberg atom voltage reference." AVS Quantum Science 4, no. 2 (June 2022): 024403. http://dx.doi.org/10.1116/5.0090892.

Full text
Abstract:
We implement a DC voltage reference by measuring Stark shifts of cesium Rydberg atoms in a vapor cell. Cesium atoms are excited from the ground state to the 15s state via a degenerate two-photon transition that provides a narrow, Doppler free line. The 15s state experiences a scalar, quadratic stark shift, which is used to measure the voltage across a parallel plate capacitor integrated into the vapor cell. We demonstrate a sensitivity of 82 mV/[Formula: see text] at a bias voltage of 100 V. The device could be adapted for even larger voltages by increasing the plate spacing or using a lower energy state.
APA, Harvard, Vancouver, ISO, and other styles
30

Walls, Brian, Oisín Murtagh, Sergey I. Bozhko, Andrei Ionov, Andrey A. Mazilkin, Daragh Mullarkey, Ainur Zhussupbekova, et al. "VOx Phase Mixture of Reduced Single Crystalline V2O5: VO2 Resistive Switching." Materials 15, no. 21 (October 31, 2022): 7652. http://dx.doi.org/10.3390/ma15217652.

Full text
Abstract:
The strongly correlated electron material, vanadium dioxide (VO2), has seen considerable attention and research application in metal-oxide electronics due to its metal-to-insulator transition close to room temperature. Vacuum annealing a V2O5(010) single crystal results in Wadsley phases (VnO2n+1, n > 1) and VO2. The resistance changes by a factor of 20 at 342 K, corresponding to the metal-to-insulator phase transition of VO2. Macroscopic voltage-current measurements with a probe separation on the millimetre scale result in Joule heating-induced resistive switching at extremely low voltages of under a volt. This can reduce the hysteresis and facilitate low temperature operation of VO2 devices, of potential benefit for switching speed and device stability. This is correlated to the low resistance of the system at temperatures below the transition. High-resolution transmission electron microscopy measurements reveal a complex structural relationship between V2O5, VO2 and V6O13 crystallites. Percolation paths incorporating both VO2 and metallic V6O13 are revealed, which can reduce the resistance below the transition and result in exceptionally low voltage resistive switching.
APA, Harvard, Vancouver, ISO, and other styles
31

Lei, Tongfei, Saleem Riaz, Noor Zanib, Munira Batool, Feng Pan, and Shaoguo Zhang. "Performance Analysis of Grid-Connected Distributed Generation System Integrating a Hybrid Wind-PV Farm Using UPQC." Complexity 2022 (March 18, 2022): 1–14. http://dx.doi.org/10.1155/2022/4572145.

Full text
Abstract:
This work presents a distributed generation system (DG) that combines system of a wind turbine (WT) and photovoltaic (PV) using a unified power quality conditioner (UPQC). Along with providing active power (AP) to the utility grid, Wind-PV-UPQC improves PQ indicators, for example, voltage drops/surges, harmonics of grid voltages, and PF. Since Wind-PV-UPQC depends on dual compensation scheme, the parallel converter works as a sinusoidal voltage source, while the series converter works as a sinusoidal current source. In this way, a smooth transition from grid operation to island operation and vice versa can be achieved without load voltage transitions. In addition, in order to overcome the problems through abrupt solar radiation or wind speed variations, a faster power balance is achieved between the wind turbines, the PV array, and the grid, as FFCL pursue the production of the current references of series converter. Consequently, the dynamic reactions of the converter currents and the voltage of dc bus are enhanced. A comprehensive analysis of flow of the AP through the converters is done to ensure a proper understanding of how Wind-PV-UPQC works. Finally, the simulation results are shown to estimate the dynamic and static performance of Wind-PV-UPQC in conjunction with the power distribution system.
APA, Harvard, Vancouver, ISO, and other styles
32

Akhlaghi, Baharak, Morteza Esteki, and Hosein Farzanehfard. "Family of zero voltage transition interleaved converters with low voltage and current stress." IET Power Electronics 11, no. 12 (September 19, 2018): 1886–93. http://dx.doi.org/10.1049/iet-pel.2017.0656.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Gurunathan, R., and A. K. S. Bhat. "A zero-voltage transition boost converter using a zero-voltage switching auxiliary circuit." IEEE Transactions on Power Electronics 17, no. 5 (September 2002): 658–68. http://dx.doi.org/10.1109/tpel.2002.802184.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Chernov, Pavel, Aleksandr Ponomarev, and Aleksei N. Lachinov. "Research of Changes of Value of Potential Barrier at the Metal/Polymer Interface during Phase Transitions in Metal." Materials Science Forum 845 (March 2016): 61–64. http://dx.doi.org/10.4028/www.scientific.net/msf.845.61.

Full text
Abstract:
The article is devoted to the study of influence of the first-order phase transition in a metal (In) on changes in the metal/polymer potential barrier. The metal that undergoes the phase transition is not in direct contact with the polymer film. Analysis of the current-voltage characteristics allows to calculate the changes in the magnitude of the metal/polymer potential barrier at the phase transitions in In.
APA, Harvard, Vancouver, ISO, and other styles
35

An, Kun, Xuechen Jin, Jiang Meng, Xiao Li, and Yifeng Ren. "Frequency Invariability of (Pb,La)(Zr,Ti)O3 Antiferroelectric Thick-Film Micro-Cantilevers." Sensors 18, no. 5 (May 13, 2018): 1542. http://dx.doi.org/10.3390/s18051542.

Full text
Abstract:
Micro-electromechanical systems comprising antiferroelectric layers can offer both actuation and transduction to integrated technologies. Micro-cantilevers based on the (Pb0.97La0.02)(Zr0.95Ti0.05)O3 (PLZT) antiferroelectric thick film are fabricated by the micro-nano manufacturing process, to utilize the effect of phase transition induced strain and sharp phase switch of antiferroelectric materials. When micro-cantilevers made of antiferroelectric thick films were driven by sweep voltages, there were two resonant peaks corresponding to the natural frequency shift from 27.8 to 27.0 kHz, before and after phase transition. This is the compensation principle for the PLZT micro-cantilever to tune the natural frequency by the amplitude modulation of driving voltage, rather than of frequency modulation. Considering the natural frequency shift about 0.8 kHz and the frequency tuning ability about 156 Hz/V before the phase transition, this can compensate the frequency shift caused by increasing temperature by tuning only the amplitude of driving voltage, when the ultrasonic micro-transducer made of antiferroelectric thick films works for such a long period. Therefore, antiferroelectric thick films with hetero-structures incorporated into PLZT micro-cantilevers not only require a lower driving voltage (no more than 40 V) than rival bulk piezoelectric ceramics, but also exhibit better performance of frequency invariability, based on the amplitude modulation.
APA, Harvard, Vancouver, ISO, and other styles
36

Peiravan, Zahra, Majid Delshad, and Mohammad Reza Amini. "A new zero voltage transition interleaved flyback converter." International Journal of Power Electronics and Drive Systems (IJPEDS) 13, no. 2 (June 1, 2022): 1026. http://dx.doi.org/10.11591/ijpeds.v13.i2.pp1026-1036.

Full text
Abstract:
The paper introduced a new zero voltage transition (ZVT) interleaved flyback converter which has two similar flyback converters. Two flyback converters are in parallel connection and auxiliary circuit in this converter provides ZVT condition for all of the main switches and also provides zero current switching and zero voltage zero current switching (ZVZCS) conditions for the auxiliary switch. Also, ZCS conditions are created for diodes turning off, so reverse recovery problem is solved. The auxiliary circuit in the suggested converter is modular, and by adding parallel branches to the flyback circuit, this circuit can provide soft switching conditions for all switches without significantly change. A complete analysis of the converter is provided and its operating intervals are explained. A 180 W laboratory prototype of the converter is made to approve the theoretical calculations. The experimental results show 7.7% increase in efficiency.
APA, Harvard, Vancouver, ISO, and other styles
37

Sun, Xiaoning, Zhaoming Qu, Jianghang Yuan, Erwei Cheng, Pingping Wang, and Qingguo Wang. "Voltage-induced phase transition of VO2@SiO2 nanoparticles." Ceramics International 47, no. 20 (October 2021): 29011–22. http://dx.doi.org/10.1016/j.ceramint.2021.07.063.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Yao-Ching Hsieh, Te-Chin Hsueh, and Hau-Chen Yen. "An Interleaved Boost Converter With Zero-Voltage Transition." IEEE Transactions on Power Electronics 24, no. 4 (April 2009): 973–78. http://dx.doi.org/10.1109/tpel.2008.2010397.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Adib, E., and H. Farzanehfard. "Zero-Voltage-Transition PWM Converters With Synchronous Rectifier." IEEE Transactions on Power Electronics 25, no. 1 (January 2010): 105–10. http://dx.doi.org/10.1109/tpel.2009.2024153.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Adib, E., and H. Farzanehfard. "Family of isolated zero-voltage transition PWM converters." IET Power Electronics 1, no. 1 (2008): 144. http://dx.doi.org/10.1049/iet-pel:20070125.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Dusmez, Serkan, Alireza Khaligh, and Amin Hasanzadeh. "A Zero-Voltage-Transition Bidirectional DC/DC Converter." IEEE Transactions on Industrial Electronics 62, no. 5 (May 2015): 3152–62. http://dx.doi.org/10.1109/tie.2015.2404825.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Chien-Ming Wang. "Novel zero-Voltage-transition PWM DC-DC converters." IEEE Transactions on Industrial Electronics 53, no. 1 (February 2006): 254–62. http://dx.doi.org/10.1109/tie.2005.862253.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Bâldea, Ioan. "Transition voltage spectroscopy: Artefacts of the Simmons approach." Journal of Physics and Chemistry of Solids 73, no. 9 (September 2012): 1151–53. http://dx.doi.org/10.1016/j.jpcs.2012.05.006.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Bâldea, Ioan. "Revealing molecular orbital gating by transition voltage spectroscopy." Chemical Physics 377, no. 1-3 (November 2010): 15–20. http://dx.doi.org/10.1016/j.chemphys.2010.08.009.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Shin, J. U., and T. Kim. "Technique for Transition Energy-Aware Dynamic Voltage Assignment." IEEE Transactions on Circuits and Systems II: Express Briefs 53, no. 9 (September 2006): 956–60. http://dx.doi.org/10.1109/tcsii.2006.881808.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Gooth, Johannes, Robert Zierold, Jan G. Gluschke, Tim Boehnert, Stefan Edinger, Sven Barth, and Kornelius Nielsch. "Gate voltage induced phase transition in magnetite nanowires." Applied Physics Letters 102, no. 7 (February 18, 2013): 073112. http://dx.doi.org/10.1063/1.4793529.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Hwu, Kuo-Ing, Jenn-Jong Shieh, and Wen-Zhuang Jiang. "Three-level boost converter with zero voltage transition." Journal of Engineering 2017, no. 7 (July 1, 2017): 354–61. http://dx.doi.org/10.1049/joe.2017.0149.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Martins, M. L., J. L. Russi, and H. L. Hey. "Zero-voltage transition PWM converters: a classification methodology." IEE Proceedings - Electric Power Applications 152, no. 2 (2005): 323. http://dx.doi.org/10.1049/ip-epa:20041230.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Laitko, Ulrike, Peter F. Juranka, and Catherine E. Morris. "Membrane Stretch Slows the Concerted Step prior to Opening in a Kv Channel." Journal of General Physiology 127, no. 6 (May 30, 2006): 687–701. http://dx.doi.org/10.1085/jgp.200509394.

Full text
Abstract:
In the simplest model of channel mechanosensitivity, expanded states are favored by stretch. We showed previously that stretch accelerates voltage-dependent activation and slow inactivation in a Kv channel, but whether these transitions involve expansions is unknown. Thus, while voltage-gated channels are mechanosensitive, it is not clear whether the simplest model applies. For Kv pore opening steps, however, there is excellent evidence for concerted expansion motions. To ask how these motions respond to stretch, therefore, we have used a Kv1 mutant, Shaker ILT, in which the step immediately prior to opening is rate limiting for voltage-dependent current. Macroscopic currents were measured in oocyte patches before, during, and after stretch. Invariably, and directly counter to prediction for expansion-derived free energy, ILT current activation (which is limited by the concerted step prior to pore opening) slowed with stretch and the g(V) curve reversibly right shifted. In WTIR (wild type, inactivation removed), the g(V) (which reflects independent voltage sensor motions) is left shifted. Stretch-induced slowing of ILT activation was fully accounted for by a decreased basic forward rate, with no change of gating charge. We suggest that for the highly cooperative motions of ILT activation, stretch-induced disordering of the lipid channel interface may yield an entropy increase that dominates over any stretch facilitation of expanded states. Since tail current τ(V) reports on the opposite (closing) motions, ILT and WTIR τ(V)tail were determined, but the stretch responses were too complex to shed much light. Shaw is the Kv3 whose voltage sensor, introduced into Shaker, forms the chimera that ILT mimics. Since Shaw2 F335A activation was reportedly a first-order concerted transition, we thought its activation might, like ILT's, slow with stretch. However, Shaw2 F335A activation proved to be sigmoid shaped, so its rate-limiting transition was not a concerted pore-opening transition. Moreover, stretch, via an unidentified non–rate-limiting transition, augmented steady-state current in Shaw2 F335A. Since putative area expansion and compaction during ILT pore opening and closing were not the energetically consequential determinants of stretch modulation, models incorporating fine details of bilayer structural forces will probably be needed to explain how, for Kv channels, bilayer stretch slows some transitions while accelerating others.
APA, Harvard, Vancouver, ISO, and other styles
50

Börjesson, Sara I., and Fredrik Elinder. "An electrostatic potassium channel opener targeting the final voltage sensor transition." Journal of General Physiology 137, no. 6 (May 30, 2011): 563–77. http://dx.doi.org/10.1085/jgp.201110599.

Full text
Abstract:
Free polyunsaturated fatty acids (PUFAs) modulate the voltage dependence of voltage-gated ion channels. As an important consequence thereof, PUFAs can suppress epileptic seizures and cardiac arrhythmia. However, molecular details for the interaction between PUFA and ion channels are not well understood. In this study, we have localized the site of action for PUFAs on the voltage-gated Shaker K channel by introducing positive charges on the channel surface, which potentiated the PUFA effect. Furthermore, we found that PUFA mainly affects the final voltage sensor movement, which is closely linked to channel opening, and that specific charges at the extracellular end of the voltage sensor are critical for the PUFA effect. Because different voltage-gated K channels have different charge profiles, this implies channel-specific PUFA effects. The identified site and the pharmacological mechanism will potentially be very useful in future drug design of small-molecule compounds specifically targeting neuronal and cardiac excitability.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography