To see the other types of publications on this topic, follow the link: Space size.

Journal articles on the topic 'Space size'

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 'Space size.'

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

Chukova, S., and V. E. Gauzelman. "Size Distortions: Space-Time Interaction." Perception 26, no. 1_suppl (1997): 114. http://dx.doi.org/10.1068/v970102.

Full text
Abstract:
We used a modified method of constant stimuli to measure spatial interval discrimination thresholds. Horizontal intervals were indicated by a pair of dark vertical lines on a bright background. In each experimental session, thresholds were measured for seven reference stimuli, presented in random order. Reference stimulus separations varied from 9.52 to 16.66 min−1 in increments of 1.95 min−1. The interstimulus interval (ISI) was varied (50, 200, 500, and 1000 ms) between experimental sessions. Stimulus duration was constant at 500 ms. For all ISI durations, the point of subjective equality (PSE) for small spatial separation references was less than physical equality, the PSE for larger separations was greater, and the PSE was close to physical equality for reference stimuli in the centre of the range. This result is consistent with the modular model [V D Glezer, 1995 Vision and Mind (Mahwah, NJ: Lawrence Erlbaum)]. However, the magnitude of the PSE shifts was affected by the ISI duration: at 50 and 1000 ms, the small spatial intervals were more underestimated and the large ones were more overestimated than at 200 or 500 ms. The discriminability thresholds based on the slopes of the psychometric functions varied inversely with the ISI duration, but at the ISI of 1000 ms increased again. These findings demonstrate that in the sequential mode of presentation the temporal separation can be as important as the spatial separation distribution in determining the PSE. This suggests that these size distortions result more from memory processing than from spatial processing.
APA, Harvard, Vancouver, ISO, and other styles
2

Ben-Sasson, Eli. "Size-Space Tradeoffs for Resolution." SIAM Journal on Computing 38, no. 6 (2009): 2511–25. http://dx.doi.org/10.1137/080723880.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Gnad, Daniel, and Jorg Hoffmann. "On the Relation between Star-Topology Decoupling and Petri Net Unfolding." Proceedings of the International Conference on Automated Planning and Scheduling 29 (May 25, 2021): 172–80. http://dx.doi.org/10.1609/icaps.v29i1.3473.

Full text
Abstract:
Petri net unfolding expands concurrent sub-threads of a transition system separately. In AI Planning, star-topology decoupling (STD) finds a partitioning of state variables into components whose dependencies take a star shape, and expands leafcomponent state spaces separately. Thus both techniques rely on the separate expansion of state-space composites. How do they relate? We show that, provided compatible search orderings, STD state space size dominates that of unfolding if every component contains a single state variable, and unfolding dominates STD in the absence of prevail conditions (nondeleted action preconditions). In all other cases, exponential state space size advantages are possible on either side. Thus the sources of exponential advantages of STD are exactly a) state space size in the presence of prevail conditions (our results), and b) decidability of reachability in time linear in state space size vs. NP-hard for unfolding (known results).
APA, Harvard, Vancouver, ISO, and other styles
4

Neal, Valerie. "Space policy and the size of the space shuttle fleet." Space Policy 20, no. 3 (2004): 157–69. http://dx.doi.org/10.1016/j.spacepol.2004.06.001.

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

Zimmer, Michael F. "Finite-size scaling in space-time." Physical Review E 57, no. 5 (1998): 5190–95. http://dx.doi.org/10.1103/physreve.57.5190.

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

HASEGAWA, Go, Yoshiharu TSUKAMOTO, and Masahiro TANAKA. "ROOM ARRANGEMENT COMPARING SIZE OF SPACE." Journal of Architecture and Planning (Transactions of AIJ) 79, no. 699 (2014): 1257–64. http://dx.doi.org/10.3130/aija.79.1257.

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

González-Val, Rafael. "US city-size distribution and space." Spatial Economic Analysis 14, no. 3 (2019): 283–300. http://dx.doi.org/10.1080/17421772.2019.1572917.

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

Benzi, R., and R. Petronzio. "Finite-Size Real-Space Renormalization Group." Europhysics Letters (EPL) 9, no. 1 (1989): 17–22. http://dx.doi.org/10.1209/0295-5075/9/1/004.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Bauer, Reinhard, Tobias Columbus, Ignaz Rutter, and Dorothea Wagner. "Search-space size in contraction hierarchies." Theoretical Computer Science 645 (September 2016): 112–27. http://dx.doi.org/10.1016/j.tcs.2016.07.003.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Petronzio, Roberto. "Finite size real space renormalisation group." Nuclear Physics B - Proceedings Supplements 9 (June 1989): 533–35. http://dx.doi.org/10.1016/0920-5632(89)90157-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Knight, C. A., and J. M. Beaulieu. "Genome Size Scaling through Phenotype Space." Annals of Botany 101, no. 6 (2008): 759–66. http://dx.doi.org/10.1093/aob/mcm321.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Hattab, Hawete. "On quasi-orbital space." Applied General Topology 18, no. 1 (2017): 53. http://dx.doi.org/10.4995/agt.2017.4676.

Full text
Abstract:
<p align="left">Let <em><span style="font-family: CMMI8; font-size: xx-small;"><em><span style="font-family: CMMI8; font-size: xx-small;">G </span></em></span></em><span style="font-family: CMR8; font-size: xx-small;"><span style="font-family: CMR8; font-size: xx-small;">be a subgroup of the group Homeo(</span></span><em><span style="font-family: CMMI8; font-size: xx-small;"><em><span style="font-family: CMMI8; font-size: xx-small;">E</span></em></span></em><span style="font-family: CMR8; font-size: xx-small;">) of homeomorphisms </span>of a Hausdorff topological space <em><span style="font-family: CMMI8; font-size: xx-small;"><em><span style="font-family: CMMI8; font-size: xx-small;">E</span></em></span></em><span style="font-family: CMR8; font-size: xx-small;"><span style="font-family: CMR8; font-size: xx-small;">. The class of an orbit </span></span><em><span style="font-family: CMMI8; font-size: xx-small;"><em><span style="font-family: CMMI8; font-size: xx-small;">O </span></em></span></em><span style="font-family: CMR8; font-size: xx-small;"><span style="font-family: CMR8; font-size: xx-small;">of </span></span><em><span style="font-family: CMMI8; font-size: xx-small;"><em><span style="font-family: CMMI8; font-size: xx-small;">G </span></em></span></em><span style="font-family: CMR8; font-size: xx-small;">is the union of </span>all orbits having the same closure as <em><span style="font-family: CMMI8; font-size: xx-small;"><em><span style="font-family: CMMI8; font-size: xx-small;">O</span></em></span></em><span style="font-family: CMR8; font-size: xx-small;"><span style="font-family: CMR8; font-size: xx-small;">. We denote by </span></span><em><span style="font-family: CMMI8; font-size: xx-small;"><em><span style="font-family: CMMI8; font-size: xx-small;">E=</span></em></span></em><span style="font-family: CMEX8; font-size: xx-small;"><span style="font-family: CMEX8; font-size: xx-small;">e</span></span><em><span style="font-family: CMMI8; font-size: xx-small;"><em><span style="font-family: CMMI8; font-size: xx-small;">G </span></em></span></em><span style="font-family: CMR8; font-size: xx-small;">the space of classes </span>of orbits called quasi-orbit space. A space <em><span style="font-family: CMMI8; font-size: xx-small;"><em><span style="font-family: CMMI8; font-size: xx-small;">X </span></em></span></em><span style="font-family: CMR8; font-size: xx-small;">is called a quasi-orbital space if </span>it is homeomorphic to <em><span style="font-family: CMMI8; font-size: xx-small;"><em><span style="font-family: CMMI8; font-size: xx-small;">E=</span></em></span></em><span style="font-family: CMEX8; font-size: xx-small;"><span style="font-family: CMEX8; font-size: xx-small;">e</span></span><em><span style="font-family: CMMI8; font-size: xx-small;"><em><span style="font-family: CMMI8; font-size: xx-small;">G </span></em></span></em><span style="font-family: CMR8; font-size: xx-small;"><span style="font-family: CMR8; font-size: xx-small;">where </span></span><em><span style="font-family: CMMI8; font-size: xx-small;"><em><span style="font-family: CMMI8; font-size: xx-small;">E </span></em></span></em><span style="font-family: CMR8; font-size: xx-small;">is a compact Hausdorff space. In this </span>paper, we show that every in nite second countable quasi-compact <em><span style="font-family: CMMI8; font-size: xx-small;"><em><span style="font-family: CMMI8; font-size: xx-small;">T</span></em></span></em><span style="font-family: CMR6; font-size: xx-small;"><span style="font-family: CMR6; font-size: xx-small;">0</span></span><span style="font-family: CMR8; font-size: xx-small;">-space </span>is the quotient of a quasi-orbital space.</p><p align="left"> </p><p align="left"> </p>
APA, Harvard, Vancouver, ISO, and other styles
13

Bernal-González, Luis. "The algebraic size of the family of injective operators." Open Mathematics 15, no. 1 (2017): 13–20. http://dx.doi.org/10.1515/math-2017-0005.

Full text
Abstract:
Abstract In this paper, a criterion for the existence of large linear algebras consisting, except for zero, of one-to-one operators on an infinite dimensional Banach space is provided. As a consequence, it is shown that every separable infinite dimensional Banach space supports a commutative infinitely generated free linear algebra of operators all of whose nonzero members are one-to-one. In certain cases, the assertion holds for nonseparable Banach spaces.
APA, Harvard, Vancouver, ISO, and other styles
14

Bebko, Adam O., and Nikolaus F. Troje. "The size of objects in visual space compared to pictorial space." Journal of Vision 19, no. 10 (2019): 16. http://dx.doi.org/10.1167/19.10.16.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Voyiatzaki, Eleni, Spyros Papadakis, Eleni Rossiou, Nikos Avouris, Konstantinos Paparrizos, and Thanasis Hadzilacos. "One Size does not fit all." Proceedings of the International Conference on Networked Learning 6 (May 5, 2008): 847–48. http://dx.doi.org/10.54337/nlc.v6.9350.

Full text
Abstract:
This paper describes the work on the selection and integration of different pedagogical methods and tools in order to address specific needs of a given authentic educational setting. Asynchronous communication with LMS (COMPUS), synchronous communication within virtual classroom implemented by CENTRA and collaborative networked activities supported by SYNERGO are combined to support f2f traditional teaching of Algorithms to first year students of University of Macedonia, Thessaloniki, Greece. The integration of these activities brought up the need to define conceptual “spaces”. Students participated actively and moved successfully, through ‘spaces’: the “virtual preparatory space” where all participants shared material, and prepared “class” work (like a library), the “ virtual classroom” for blended learning activities (like a class), the “private group space” where students in dyads or triads solved problems (like a desk), a “private room” for the teacher-student private discussion when feedback and additional support was needed (like teacher’s desk), “small private rooms” for small groups meta-cognitive activities (like a meeting table). Various communication channels have been integrated to support the participants as video, audio, chat, synchronous drawing tool, etc. Participants, instructors and students, had successfully switched among “spaces” and tools. Students (85%) reproduced the whole activity during the preparation for their exams. Generally, students have improved their performance and avoided typical mistakes discussed during the activity, in their exams.
APA, Harvard, Vancouver, ISO, and other styles
16

Willerton, Simon. "Spread: A Measure of the Size of Metric Spaces." International Journal of Computational Geometry & Applications 25, no. 03 (2015): 207–25. http://dx.doi.org/10.1142/s0218195915500120.

Full text
Abstract:
Motivated by Leinster-Cobbold measures of biodiversity, the notion of the spread of a finite metric space is introduced. This is related to Leinster’s magnitude of a metric space. Spread is generalized to infinite metric spaces equipped with a measure and is calculated for spheres and straight lines. For Riemannian manifolds the spread is related to the volume and total scalar curvature. A notion of scale-dependent dimension is introduced and seen for approximations to certain fractals to be numerically close to the Minkowski dimension of the original fractals.
APA, Harvard, Vancouver, ISO, and other styles
17

Kasim, Priawanto ,., Bobby ,. Polii, and Zetly ,. Tamod. "OPTIMASI TUTUPAN LAHAN RUANG TERBUKA HIJAU PUBLIK KOTA KOTAMOBAGU." AGRI-SOSIOEKONOMI 13, no. 3A (2017): 9. http://dx.doi.org/10.35791/agrsosek.13.3a.2017.17950.

Full text
Abstract:
This study was conducted to (1) identify the public green opened space based on covering area in Kotamobagu city and (2) analyze the optimization of development for the public green opened space in Kotamobagu city. Method used in this study was the mix method. The mix method was applicated to study and analyze the index of covering area based on size of existing of the public green opened space. This existing public green opened space would define direction of continouos public green opened space development in Kotamobagu city. Result of the study showed that (1) the size of public green opened space in Kotamobagu city was 1.059 ha with the covering area index of 16 percents. The size of public green opened space in Kotamobagu city consists of several types of public green opened spaces, which were not optimally managed such as those at Pobundayan city parks, Gelora Ambang city parks and Bonawang city forest. (2) Development of public green opened space in Kotamobagu city was focused on minimum size target of 20 percents from entired size of Kotamobagu areas as the indicator of green city standard. The potential locations for size development were including the areas of plantation, city forest and river common border of 50 m.
APA, Harvard, Vancouver, ISO, and other styles
18

Shelah, Saharon, and Stevo Todorcevic. "A Note on Small Baire Spaces." Canadian Journal of Mathematics 38, no. 3 (1986): 659–65. http://dx.doi.org/10.4153/cjm-1986-033-8.

Full text
Abstract:
A Baire space is a topological space which satisfies the Baire Category Theorem, i.e., in which the intersection of countably many dense open sets is dense. In this note we shall be interested in the size of Baire spaces, so to avoid trivialities we shall consider only non-atomic spaces, that is, spaces X whose regular open algebras ro(X) are non-atomic. All natural examples of Baire spaces, such as complete metric spaces or compact spaces, seem to have sizes at least 2ℵ0. So a natural question, asked first by W. Fleissner and K. Kunen [5], is whether there exists a Baire space of the minimal possible size ℵ1. The purpose of this note is to show that such a space need not exist by proving the following result.
APA, Harvard, Vancouver, ISO, and other styles
19

Gilmore, Dehn. "Life-Size." Victorian Literature and Culture 51, no. 3 (2023): 451–54. http://dx.doi.org/10.1017/s1060150323000141.

Full text
Abstract:
This entry posits the life-size as a form of “technology” for collapsing time and space and closes by suggesting some ways in which literary authors were inspired by the proliferation of life-size forms in culture.
APA, Harvard, Vancouver, ISO, and other styles
20

Arend, Harald, and H. Hermann Koelle. "Size and economics of big space freighters." Journal of Spacecraft and Rockets 26, no. 4 (1989): 240–44. http://dx.doi.org/10.2514/3.26061.

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

Spencer, Robin. "The Size and Shape of "Idea Space"." International Journal of Innovation Science 4, no. 2 (2012): 71–76. http://dx.doi.org/10.1260/1757-2223.4.2.71.

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

HASEGAWA, Go, and Yoshiharu TSUKAMOTO. "COMPOSITION OF STORIES COMPARING SIZE OF SPACE." Journal of Architecture and Planning (Transactions of AIJ) 80, no. 709 (2015): 737–43. http://dx.doi.org/10.3130/aija.80.737.

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

Bruda, S. D., and S. G. Akl. "Size Matters: Logarithmic Space is Real Time." International Journal of Computers and Applications 29, no. 4 (2007): 327–36. http://dx.doi.org/10.1080/1206212x.2007.11441863.

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

Luo, G., T. Garaas, M. Pomplun, and E. Peli. "Does saccadic space compression mean size shrinking?" Journal of Vision 9, no. 8 (2010): 443. http://dx.doi.org/10.1167/9.8.443.

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

Cornille, Henri. "Half-Space Large Size Discrete Velocity Models." Annales Henri Poincaré 4, S2 (2003): 889–903. http://dx.doi.org/10.1007/s00023-003-0969-z.

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

Ayvazyan, Sergey, and Mikhail Afanasyev. "The Size of Innovation Space as a Factor of Innovation Activity in Regions." MONTENEGRIN JOURNAL OF ECONOMICS 12, no. 2 (2016): 7–27. http://dx.doi.org/10.14254/1800-5845.2016/12-2/11.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Mills, L. Scott, and Frederick F. Knowlton. "Coyote space use in relation to prey abundance." Canadian Journal of Zoology 69, no. 6 (1991): 1516–21. http://dx.doi.org/10.1139/z91-212.

Full text
Abstract:
Food abundance is an important factor determining space use in many species, but its effect on carnivore home range and territory size has rarely been investigated. We explored the relationship between food abundance for the coyote (Canis latrans) and space use in two study areas in the northern Great Basin, where the primary prey, the black-tailed jackrabbit (Lepus californicus), fluctuates dramatically in abundance. At one site, home ranges and territories were significantly larger during a time of prey-scarcity than when prey was abundant. Coyotes on the second site had similar-size home ranges and territories at low and high prey abundance, but a higher proportion and probably a higher number of individuals were transients during the prey-scarcity period. We propose mortality rates of coyotes as an important factor mediating adjustments in space use to food abundance, and suggest two mechanisms by which mortality might interact with food abundance. Higher mortality rates may simply permit more rapid adjustment of home range size to changing food conditions. Alternatively, higher mortality may selectively eliminate transients, thus reducing the impact of intruders in limiting the size of the remaining territories.
APA, Harvard, Vancouver, ISO, and other styles
28

Majeed, Sultan J., and Marcus Hutter. "Exact Reduction of Huge Action Spaces in General Reinforcement Learning." Proceedings of the AAAI Conference on Artificial Intelligence 35, no. 10 (2021): 8874–83. http://dx.doi.org/10.1609/aaai.v35i10.17074.

Full text
Abstract:
The reinforcement learning (RL) framework formalizes the notion of learning with interactions. Many real-world problems have large state-spaces and/or action-spaces such as in Go, StarCraft, protein folding, and robotics or are non-Markovian, which cause significant challenges to RL algorithms. In this work we address the large action-space problem by sequentializing actions, which can reduce the action-space size significantly, even down to two actions at the expense of an increased planning horizon. We provide explicit and exact constructions and equivalence proofs for all quantities of interest for arbitrary history-based processes. In the case of MDPs, this could help RL algorithms that bootstrap. In this work we show how action-binarization in the non-MDP case can significantly improve Extreme State Aggregation (ESA) bounds. ESA allows casting any (non-MDP, non-ergodic, history-based) RL problem into a fixed-sized non-Markovian state-space with the help of a surrogate Markovian process. On the upside, ESA enjoys similar optimality guarantees as Markovian models do. But a downside is that the size of the aggregated state-space becomes exponential in the size of the action-space. In this work, we patch this issue by binarizing the action-space. We provide an upper bound on the number of states of this binarized ESA that is logarithmic in the original action-space size, a double-exponential improvement.
APA, Harvard, Vancouver, ISO, and other styles
29

Gregory, Jim S., and J. S. Griffith. "Winter concealment by subyearling rainbow trout: space size selection and reduced concealment under surface ice and in turbid water conditions." Canadian Journal of Zoology 74, no. 3 (1996): 451–55. http://dx.doi.org/10.1139/z96-052.

Full text
Abstract:
The proportion of rainbow trout (Oncorhynchus mykiss) concealing themselves in simulated interstitial spaces was examined in the presence of surface ice, in turbid water, and in clear water. Tests were conducted in enclosures in a small Idaho stream with structures that provided five rectangular spaces varying in width and height, one circular space, and one triangular space. Space use was assessed each morning by trapping test fish inside the structures. Significantly more fish concealed themselves under clear water conditions than under either surface ice or turbid water conditions. Spaces narrower than the width of a test fish with extended pectoral fins and spaces taller than the height of a test fish with dorsal fin extended were used less than would be expected if space use was random. The frequency with which two or more fish occurred together in the same space was similar to that expected if fish occurred together at random. Fish rarely returned to the same space on consecutive nights.
APA, Harvard, Vancouver, ISO, and other styles
30

Fernandes, Luciana Q. P., Rhita C. Almeida, Barbara N. G. de Andrade, Felipe de Assis R. Carvalho, Marco Antonio de O. Almeida, and Flavia R. G. Artese. "Tooth size discrepancy: Is the E space similar to the leeway space?" Journal of the World Federation of Orthodontists 2, no. 2 (2013): e49-e51. http://dx.doi.org/10.1016/j.ejwf.2013.03.001.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Lelis, Levi, Roni Stern, and Nathan Sturtevant. "Estimating Search Tree Size with Duplicate Detection." Proceedings of the International Symposium on Combinatorial Search 5, no. 1 (2021): 114–22. http://dx.doi.org/10.1609/socs.v5i1.18328.

Full text
Abstract:
In this paper we introduce Stratified Sampling with Duplicate Detection (SSDD), an algorithm for estimating the number of state expansions performed by heuristic search algorithms seeking solutions in state spaces represented by undirected graphs. SSDD is general and can be applied to estimate other state-space properties. We test SSDD on two tasks: (i) prediction of the number of A* expansions in a given f-layer when using a consistent heuristic function, and (ii) prediction of the state-space radius. SSDD has the asymptotic guarantee of producing perfect estimates in both tasks. Our empirical results show that in task (i) SSDD produces good estimates in all four domains tested, being in most cases orders of magnitude more accurate than a competing scheme, and in task (ii) SSDD quickly produces accurate estimates of the radii of the 4x4 Sliding-Tile Puzzle and the 3x3x3 Rubik's Cube.
APA, Harvard, Vancouver, ISO, and other styles
32

Leith, Scott A., and Anne E. Wilson. "When size justifies: intergroup attitudes and subjective size judgments of “sacred space”." Journal of Experimental Social Psychology 54 (September 2014): 122–30. http://dx.doi.org/10.1016/j.jesp.2014.05.003.

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

Siklóssy, Laurent, and Eduard Tulp. "The space reduction method: a method to reduce the size of search spaces." Information Processing Letters 38, no. 4 (1991): 187–92. http://dx.doi.org/10.1016/0020-0190(91)90098-3.

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

Han, Shuyan, Dexuan Song, Feng Shi, Hu Du, Yuhao Zhang, and Mingjun Yang. "Assessing Neighbourhood Preference: An Evaluation of Environmental Features within Small-Scale Open Spaces." Land 13, no. 4 (2024): 531. http://dx.doi.org/10.3390/land13040531.

Full text
Abstract:
Well-designed urban public spaces often attract residents and play a critical role in improving people’s wellbeing. Many studies have examined the importance of one or a few environmental features in urban public spaces, such as the size of the space, greenery coverage, seating arrangements, recreational facilities, etc. However, there is a lack of systematic understanding regarding (1) which environmental features have a significant impact on the usage of urban public spaces and (2) how these features influence people’s environmental preferences. To answer these questions, this investigation adopts a two-fold analytical structure: (1) first, an expert inquiry was conducted to evaluate the environmental features, and the analytic hierarchy process (AHP) was applied to determine the weight of each influencing factor; then, (2) on-site measurements were conducted across 104 spaces, accompanied by structured interviews with users of the spaces, based on which a decision tree analysis was employed to elucidate the decision-making processes of residents regarding their outdoor activities. The main findings of this investigation are as follows: (1) the site size, internal pedestrian flow, sky view factor, green-vision rate, and seat–circumference ratio are primary indicators affecting outdoor space usage, which are used in the objective evaluation index; (2) advantage value intervals for the sky view factor, green-vision rate, and seat–circumference ratio variables were calculated, and these three factors were found to significantly outweigh site size and internal pedestrian flow in terms of their effect on spatial preference. The interaction between the green-vision rate and seat–circumference ratio can affect the environmental preferences of residents: spaces with more seats exhibit lower requirements for greenery, while spaces with fewer seats should prioritise trees and greenery. Based on this study, an index based on influencing factors is proposed, enabling a better understanding of the environmental features affecting the usage of space. This study also provides valuable insights for future neighbourhood design through investigating the environmental preferences of residents, as well as the importance of various spatial features and their associated advantage value intervals.
APA, Harvard, Vancouver, ISO, and other styles
35

Song, H. Q., and T. T. S. Kuo. "Model-space-size dependence of nuclear matter model-space Brueckner-Hartree-Fock calculations." Physical Review C 43, no. 6 (1991): 2883–86. http://dx.doi.org/10.1103/physrevc.43.2883.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Schiermeier, Quirin. "Space probe set to size up polar ice." Nature 464, no. 7289 (2010): 658. http://dx.doi.org/10.1038/464658a.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Gorea, A., S. Belkoura, and J. A. Solomon. "Summary statistics for size over space and time." Journal of Vision 14, no. 9 (2014): 22. http://dx.doi.org/10.1167/14.9.22.

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

Baker, Rose D. "Testing for space-time clusters of unknown size." Journal of Applied Statistics 23, no. 5 (1996): 543–54. http://dx.doi.org/10.1080/02664769624080.

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

Lande, B., and W. Mitzner. "Assessment of air space size and surface area." Journal of Applied Physiology 108, no. 3 (2010): 760. http://dx.doi.org/10.1152/japplphysiol.00006.2010.

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

Kondyurin, Alexey. "Large-size space laboratory for biological orbit experiments." Advances in Space Research 28, no. 4 (2001): 665–71. http://dx.doi.org/10.1016/s0273-1177(01)00376-3.

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

Murphy, Samuel William, Carlos Roberto de Souza Filho, and Clive Oppenheimer. "Monitoring volcanic thermal anomalies from space: Size matters." Journal of Volcanology and Geothermal Research 203, no. 1-2 (2011): 48–61. http://dx.doi.org/10.1016/j.jvolgeores.2011.04.008.

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

Jeschke, Hartwig. "Chip size estimation for SOC design space exploration." Journal of Systems Architecture 53, no. 10 (2007): 764–76. http://dx.doi.org/10.1016/j.sysarc.2007.01.012.

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

de Berg, M. "Linear Size Binary Space Partitions for Uncluttered Scenes." Algorithmica 28, no. 3 (2000): 353–66. http://dx.doi.org/10.1007/s004530010047.

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

Weirich, Melanie, and Adrian P. Simpson. "Acoustic vowel space size and perceived speech tempo." Journal of the Acoustical Society of America 133, no. 5 (2013): 3571. http://dx.doi.org/10.1121/1.4806540.

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

Regnier, Eva D., and Steven M. Shechter. "State-space size considerations for disease-progression models." Statistics in Medicine 32, no. 22 (2013): 3862–80. http://dx.doi.org/10.1002/sim.5808.

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

Chambers, Jeanne C. "Patterns of growth and reproduction in a perennial tundra forb (Geum rossii): effects of clone area and neighborhood." Canadian Journal of Botany 69, no. 9 (1991): 1977–83. http://dx.doi.org/10.1139/b91-248.

Full text
Abstract:
The relationships of clone area and neighborhood to ramet size, reproductive effort, and spatial distribution within Geum rossii clones were studied in an alpine ecosystem on the Beartooth Plateau, Montana. Clones growing on an early seral site in relative isolation were compared to clones on a late seral site within dense, heterogeneous neighborhoods. Individual clones of G. rossii required a minimum clone area of about 200 cm2 before maximum ramet size and reproductive effort were achieved. Mean ramet size and reproductive effort were fairly constant among clones larger than 200 cm2 on both the early and later seral sites. Within clones the size and reproductive effort of ramets were positively related. Pattern analysis revealed that ramets became more widely and irregularly spaced as clone area increased on the early seral site. This may have been a geometric function of an increase in the space required as clones aged and became larger. On the late seral site, clones were characterized by ramets that were widely and erratically spaced, that had low leaf numbers and mass, and that had low reproductive effort. For clones of comparable area on the early seral site, ramets were more closely and uniformly spaced, and leaf number, mass, and reproduction per ramet were higher. Conservative patterns of growth and reproduction make G. rossii well suited to dominate in dense, heterogeneous neighborhoods of late seral sites and to colonize mineral soils of early seral sites. Similar to other clonal species, site characteristics and the type of neighborhood determine the trade-off between the physical occupation of space and the allocation to ramet growth and reproduction in G. rossii. Key words: Geum rossii, alpine, reproductive effort, growth, clone area, pattern analysis, succession, neighborhood.
APA, Harvard, Vancouver, ISO, and other styles
47

Spoolder, H. A. M., S. A. Edwards, and S. Corning. "Effects of group size and feeder space allowance on welfare in finishing pigs." Animal Science 69, no. 3 (1999): 481–89. http://dx.doi.org/10.1017/s135772980005133x.

Full text
Abstract:
AbstractCompared with small groups, housing in large groups offers the pig more total available space, resulting potentially in an increased degree of control over its (micro) environment. For the producer, large groups require fewer pen divisions and offer more possibilities for the sharing of resources such as feeders and drinkers. However, whilst large groups may offer benefits to higher ranking animals in the group, there may be serious disadvantages for those further down the social hierarchy, who also need to compete for access to resources. This study investigated the interactive effects on welfare of food availability (one single space hopper per 20 or per 10 pigs) and group size (20, 40 or 80 pigs per pen), at constant stocking density (0·55 m2per pig) in part-slatted pens. Groups provided with two feeding spaces per 20 pigs were less active than groups with one feeding space per 20 pigs. The number of aggressive interactions per pig at the food trough was not affected by group size but decreased with number of feeder spaces per 20 pigs. The number of skin lesions increased with group size. Average daily gain in the first half of the finishing period was negatively influenced by group size and positively by number of feeding spaces. No effect on weight gain was found subsequently. Within-group variation in growth was not affected by group size or number of feeder spaces. No differences between treatments were found in the number of pigs removed for health reasons. Interactive effects of the two treatments were found on some behaviours but not on any of the performance variables measured. It is concluded that, from a welfare point of view, the number of pigs per feeder space should be lower than 20, although performance levels appear acceptable at 20 pigs per feeder. Further research will have to identify whether the effects of group size on general aggression is common to all finishing pig systems, or whether the presence of straw can serve as a mitigating factor.
APA, Harvard, Vancouver, ISO, and other styles
48

LIZZI, FEDELE. "STRINGS, NONCOMMUTATIVE GEOMETRY AND THE SIZE OF THE TARGET SPACE." International Journal of Modern Physics A 14, no. 28 (1999): 4501–17. http://dx.doi.org/10.1142/s0217751x99002116.

Full text
Abstract:
We describe how the presence of the antisymmetric tensor (torsion) on the world sheet action of string theory renders the size of the target space a gauge noninvariant quantity. This generalizes the R ↔ 1/R symmetry in which momenta and windings are exchanged, to the whole O(d,d,ℤ). The crucial point is that, with a transformation, it is possible always to have all of the lowest eigenvalues of the Hamiltonian to be momentum modes. We interpret this in the framework of noncommutative geometry, in which algebras take the place of point spaces, and of the spectral action principle for which the eigenvalues of the Dirac operator are the fundamental objects, out of which the theory is constructed. A quantum observer, in the presence of many low energy eigenvalues of the Dirac operator (and hence of the Hamiltonian) will always interpreted the target space of the string theory as effectively uncompactified.
APA, Harvard, Vancouver, ISO, and other styles
49

Appleby, MC. "What causes crowding? Effects of space, facilities and group size on behaviour, with particular reference to furnished cages for hens." Animal Welfare 13, no. 3 (2004): 313–20. http://dx.doi.org/10.1017/s0962728600028426.

Full text
Abstract:
AbstractTheoretical models are presented of the effects of space, facilities and group size on the behaviour of chickens at high stocking densities, with relevance for all animals. The appropriateness of each model is supported by published data, although such data are scant for some important variables. Freedom of movement is analysed by taking the area of a hen as 475 cm2 and finding the number of free bird spaces left at different space allowances. This provides support for current recommendations of a maximum of seven laying hens per m2 on deep litter, but suggests that a maximum for broilers of 34 kg/m2 unacceptably restricts freedom of movement. In cages, freedom of movement increases with space allowance per hen, and, for a given space allowance, with cage and group size. Nesting behaviour is analysed for synchrony, which decreases with group size. Perching and feeding are often synchronous and the space needed for these is determined by body width. Recommendations are derived for hens in furnished cages. The main part of the cage should be as large as possible; an absolute minimum of 600 cm2 per bird is suggested, but 675 cm2 per bird is probably the minimum practical. Perch and feeder space should be provided at 14 cm or more per bird, with a possible derogation for light hybrids to 12 cm. The number of nest spaces needed varies with the number of birds, with nest spaces being 300 cm2 each. These recommendations sum to a minimum of 800 cm2 per bird for groups of eight or more, 850 cm2 for groups of four to seven, and 900 cm2 for groups of three or fewer, plus litter area. Crowding is primarily caused by limited space allowance, but for a given space allowance it is worse in small enclosures and groups.
APA, Harvard, Vancouver, ISO, and other styles
50

Wang, Penglong, Yanyan Ma, Xueyan Zhao, Bao Wang, Jianghao Wang, and Feng Gao. "Regional Differences and Influential Factors of Open Public Space in Chinese Cities Based on Big Earth Data." Sustainability 12, no. 6 (2020): 2514. http://dx.doi.org/10.3390/su12062514.

Full text
Abstract:
Urban open public spaces that provide multiple services for residents are essential for improving life quality and urban ecosystem function and promoting healthy development, the safety of human settlements and the sustainable development of urban cities. Based on Sustainable Development Goal 11.7 of the United Nations (UN) 2030 Agenda, this study combines the big earth data with the Theil index, a coefficient of variation and Exploratory Spatial Data Analysis (ESDA) to analyze the regional differences and spatial distribution of urban open public space in 2015 for China, and uses the geographical detector to identify key factors that affect the distribution of open public spaces. The results show that (1) open public space scales in provincial-level cities have an ‘East–Central–West’ low-lying land pattern in spatial distribution, where the eastern region has a relatively larger open public space scale. (2) In the prefecture-level cities, the open public space scale increases with an increase in city size and economic development level, and the differences in urban open public space reduce with an increase in city size and increase with a decrease in the economic development level. (3) Factors including economic development level, residents’ living standards, the urbanization level and the population size have sound explanatory powers in varying degrees on the scale of open public spaces; interactions between these factors have improved the explanatory power of the scale of urban open public space.
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!