Sample records for xfy-1 vertically rising

  1. Flight Tests of a 0.13-Scale Model of the Convair XFY-1 Vertically Rising Airplane with the Lower Vertical Tail Removed, TED No.DE 368

    NASA Technical Reports Server (NTRS)

    Lovell, Powell M., Jr.

    1954-01-01

    An experimental investigation has been conducted to determine the dynamic stability and control characteristics in hovering and transition flight of a 0.13-scale flying model of the Convair XFY-1 vertically rising airplane with the lower vertical tail removed. The purpose of the tests was to obtain a general indication of the behavior of a vertically rising airplane of the same general type as the XFY-1 but without a lower vertical tail in order to simplify power-off belly landings in an emergency. The model was flown satisfactorily in hovering flight and in the transition from hovering to normal unstalled forward flight (angle of attack approximately 30deg). From an angle of attack of about 30 down to the lowest angle of attack covered in the flight tests (approximately 15deg) the model became progressively more difficult to control. These control difficulties were attributed partly to a lightly damped Dutch roll oscillation and partly to the fact that the control deflections required for hovering and transition flight were too great for smooth flight at high speeds. In the low-angle-of-attack range not covered in the flight tests, force tests have indicated very low static directional stability which would probably result in poor flight characteristics. It appears, therefore, that the attainment of satisfactory directional stability, at angles of attack less than 10deg, rather than in the hovering and transition ranges of flight is the critical factor in the design of the vertical tail for such a configuration.

  2. Vertical Descent and Landing Tests of a 0.13-Scale Model of the Convair XFY-1 Vertically Rising Airplane in Still Air, TED No. NACA DE 368

    NASA Technical Reports Server (NTRS)

    Smith, Charlee C., Jr.; Lovell, Powell M., Jr.

    1954-01-01

    An investigation is being conducted to determine the dynamic stability and control characteristics of a 0.13-scale flying model of Convair XFY-1 vertically rising airplane. This paper presents the results of flight and force tests to determine the stability and control characteristics of the model in vertical descent and landings in still air. The tests indicated that landings, including vertical descent from altitudes representing up to 400 feet for the full-scale airplane and at rates of descent up to 15 or 20 feet per second (full scale), can be performed satisfactorily. Sustained vertical descent in still air probably will be more difficult to perform because of large random trim changes that become greater as the descent velocity is increased. A slight steady head wind or cross wind might be sufficient to eliminate the random trim changes.

  3. Wind-Tunnel Investigation at Low Speed of the Pitching Stability Derivatives of a 1/9-Scale Powered Model of the Convair XFY-1 Vertically Rising Airplane, TED No. NACA DE 373

    NASA Technical Reports Server (NTRS)

    Queijo, M. J.; Wolhart, Walter D.; Fletcher, H. S.

    1953-01-01

    An experimental investigation has been conducted in the Langley stability tunnel at low speed to determine the pitching stability derivatives of a 1/9-scale powered model of the Convair XFY-1 vertically rising airplane. Effects of thrust coefficient, control deflections, and propeller blade angle were investigated. The tests were made through an angle-of-attack range from about -4deg to 29deg, and the thrust coefficient range was from 0 to 0.7. In order to expedite distribution of these data, no analysis of the data has been prepared for this paper.

  4. Wind-Tunnel Investigation at Low Speed of the Rolling Stability Derivatives of a 1/9-Scale Powered Model of the Convair XFY-1 Vertically Rising Airplane, TED No. NACA DE 373

    NASA Technical Reports Server (NTRS)

    Queijo, M. J.; Wolhart, Walter D.; Fletcher, H. S.

    1953-01-01

    An experimental investigation has been conducted in the Langley stability tunnel at low speed to determine the rolling stability derivatives of a 1/9-scale powered model of the Convair XFY-1 vertically rising airplane. Effects of thrust coefficient were investigated for the complete model and for certain components of the model. Effects of control deflections and of propeller blade angle were investigated for the complete model. Most of the tests were made through an angle-of-attack range from about -4deg to 29deg, and the thrust coefficient range was from 0 to 0.7. In order to expedite distribution of these data, no analysis of the data has been prepared for this paper.

  5. Wind-Tunnel Investigation at Low Speed of the Yawing Stability Derivatives of a 1/9-Scale Powered Model of the Convair XFY-1 Vertically Rising Airplane, TED No. NACA DE 373

    NASA Technical Reports Server (NTRS)

    Queijo, M. J.; Wolhart, w. D.; Fletcher, H. S.

    1953-01-01

    An experimental investigation has been conducted in the Langley stability tunnel at low speed to deter+nine the yawing stability derivatives of a 1/9-scale powered model of the Convair XFY-1 vertically rising airplane. Effects of thrust coefficient were investigated for the complete model and for certain components of the model. Effects of control deflections and of propeller blade angle were investigated for the complete model. Most of the tests were made through an angle-of-attack range from about -4deg to 29deg, and the thrust coefficient range was from 0 to 0.7. In order to expedite distribution of these data, no analysis of the data has been prepared for this.

  6. Flight Investigation of the Stability and Control Characteristics of a 0.13-Scale Model of the Convair XFY-1 Vertically Rising Airplane During Constant-Altitude Transitions, TED No. NACA DE 368

    NASA Technical Reports Server (NTRS)

    Lovell, Powell M., Jr.; Kibry, Robert H.; Smith, Charles C., Jr.

    1953-01-01

    An investigation is being conducted to determine the dynamic stability and control characteristics of a 0.13-scale flying model of the Convair XFY-1 vertically rising airplane. This paper presents the results of flight tests to determine the stability and control characteristics of the model during constant-altitude slow transitions from hovering to normal unstalled forward flight. The tests indicated that the airplane can be flown through the transition range fairly easily although some difficulty will probably encountered in controlling the yawing motions at angles of attack between about 60 and 40. An increase in the size of the vertical tail will not materially improve the controllability of the yawing motions in this range of angle of attack but the use of a yaw damper will make the yawing motions easy to control throughout the entire transitional flight range. The tests also indicated that the airplane can probably be flown sideways satisfactorily at speeds up to approximately 33 knots (full scale) with the normal control system and up to approximately 37 knots (full scale) with both elevons and rudders rigged to move differentially for roll control. At sideways speeds above these values, the airplane will have a strong tendency to diverge uncontrollably in roll.

  7. Take-Off and Landing Characteristics of a 0.13-Scale Model of the Convair XFY-1 Vertically Rising Airplane in Steady Winds, TED No. NACA DE 368

    NASA Technical Reports Server (NTRS)

    Schade, Robert O.; Smith, Charles C., Jr.; Lovell, P. M., Jr.

    1954-01-01

    An experimental investigation has been conducted to determine the stability and control characteristics of a 0.13-scale free-flight model of the Convair XFY-1 airplane during take-offs and landings in steady winds. The tests indicated that take-offs in headwinds up to at least 20 knots (full scale) will be fairly easy to perform although the airplane may be blown downstream as much as 3 spans before a trim condition can be established. The distance that the airplane will be blown down-stream can be reduced by restraining the upwind landing gear until the instant of take-off. The tests also indicated that spot landings in headwinds up to at least 30 knots (full scale) and in crosswinds up to at least 20 knots (full scale) can be accomplished with reasonable accuracy although, during the landing approach, there will probably be an undesirable nosing-up tendency caused by ground effect and by the change in angle of attack resulting from vertical descent. Some form of arresting gear will probably be required to prevent the airplane from rolling downwind or tipping over after contact. This rolling and tipping can be prevented by a snubbing line attached to the tip of the upwind' wing or tail or by an arresting gear consisting of a wire mesh on the ground and hooks on the landing gear to engage the mesh.

  8. Flight Tests of a 0.13-Scale Model of the Convair XFY-1 Vertically Rising Airplane in a Setup Simulating that Proposed for Captive-Flight Tests in a Hangar, TED No. NACA DE 368

    NASA Technical Reports Server (NTRS)

    Lovell, Powell M., Jr.

    1953-01-01

    An experimental investigation has been conducted to determine the dynamic stability and control characteristics of a 0.13-scale free-flight model of the Convair XFY-1 airplane in test setups representing the setup proposed for use in the first flight tests of the full-scale airplane in the Moffett Field airship hangar. The investigation was conducted in two parts: first, tests with the model flying freely in an enclosure simulating the hangar, and second, tests with the model partially restrained by an overhead line attached to the propeller spinner and ground lines attached to the wing and tail tips. The results of the tests indicated that the airplane can be flown without difficulty in the Moffett Field airship hangar if it does not approach too close to the hangar walls. If it does approach too close to the walls, the recirculation of the propeller slipstream might cause sudden trim changes which would make smooth flight difficult for the pilot to accomplish. It appeared that the tethering system proposed by Convair could provide generally satisfactory restraint of large-amplitude motions caused by control failure or pilot error without interfering with normal flying or causing any serious instability or violent jerking motions as the tethering lines restrained the model.

  9. Flight Determination of the Longitudinal Stability Characteristics of a 0.133-Scale Rocket-Powered Model of the Consolidated Vultee XFY-1 Airplane without Propellers at Mach Numbers from 0.73 to 1.19, TED No. NACA DE 369

    NASA Technical Reports Server (NTRS)

    Hastings, Earl E., Jr.; Mitcham, Grady L.

    1954-01-01

    A flight test has been conducted to determine the longitudinal stability and control,characteristics of a 0.133-scale model of the Consolidated Vultee XFY-1 airplane without propellers for the Mach number range between 0.73 and 1.19.

  10. Measurement of the Static Stability and Control and the Damping Derivatives of a 0.13-Scale Model of the Convair XFY-1 Airplane, TED No. NACA DE 368

    NASA Technical Reports Server (NTRS)

    Johnson, Joseph L.

    1954-01-01

    An investigation has been conducted to determine the static stability and control and damping in roll and yaw of a 0.13-scale model of the Convair XFY-1 airplane with propellers off from 0 deg to 90 deg angle of attack. The tests showed that a slightly unstable pitch-up tendency occurred simultaneously with a break in the normal-force curve in the angle-of-attack range from about 27 deg to 36 deg. The top vertical tail contributed positive values of static directional stability and effective dihedral up to an angle of attack of about 35 deg. The bottom tail contributed positive values of static directional stability but negative values of effective dihedral throughout the angle-of-attack range. Effectiveness of the control surfaces decreased to very low values at the high angles of attack, The model had positive damping in yaw and damping in roll about the body axes over the angle-of-attack range but the damping in yaw decreased to about zero at 90 deg angle of attack.

  11. Consideration of vertical uncertainty in elevation-based sea-level rise assessments: Mobile Bay, Alabama case study

    USGS Publications Warehouse

    Gesch, Dean B.

    2013-01-01

    The accuracy with which coastal topography has been mapped directly affects the reliability and usefulness of elevationbased sea-level rise vulnerability assessments. Recent research has shown that the qualities of the elevation data must be well understood to properly model potential impacts. The cumulative vertical uncertainty has contributions from elevation data error, water level data uncertainties, and vertical datum and transformation uncertainties. The concepts of minimum sealevel rise increment and minimum planning timeline, important parameters for an elevation-based sea-level rise assessment, are used in recognition of the inherent vertical uncertainty of the underlying data. These concepts were applied to conduct a sea-level rise vulnerability assessment of the Mobile Bay, Alabama, region based on high-quality lidar-derived elevation data. The results that detail the area and associated resources (land cover, population, and infrastructure) vulnerable to a 1.18-m sea-level rise by the year 2100 are reported as a range of values (at the 95% confidence level) to account for the vertical uncertainty in the base data. Examination of the tabulated statistics about land cover, population, and infrastructure in the minimum and maximum vulnerable areas shows that these resources are not uniformly distributed throughout the overall vulnerable zone. The methods demonstrated in the Mobile Bay analysis provide an example of how to consider and properly account for vertical uncertainty in elevation-based sea-level rise vulnerability assessments, and the advantages of doing so.

  12. Investigation of Spinning and Tumbling Characteristics of a 1/20-Scale Model of the Consolidated Vultee XFY-1 Airplane in the Free-Spinning Tunnel, TED No. NACA DE 370

    NASA Technical Reports Server (NTRS)

    Lee, Henry A.

    1952-01-01

    An investigation has been conducted in the Langley 20-foot free-spinning tunnel on a l/20-scale model of the Consolidated Vultee XFY-1 airplane with a windmilling propeller simulated to determine the effects of control setting and movements upon the erect spin and recovery characteristics for a range of airplane-loading conditions. The effects on the model's spin-recovery characteristics of removing the lower vertical tail, removing the gun pods, and fixing the rudders at neutral were also investigated briefly. The investigation included determination of the size parachute required for emergency recovery from demonstration spins. The tumbling tendencies of the model were also investigated. Brief static force tests were made to determine the aerodynamic characteristics in pitch at high angles of attack. The investigation indicated that the spin and recovery characteristics of the airplane with propeller windmilling will be satisfactory for all loading conditions if recovery is attempted by full rudder reversal accompanied by simultaneous movement of the stick laterally to full with the spin (stick right in a right spin) and longitudinally to neutral. Inverted spins should be satisfactorily terminated by fully reversing the rudder followed immediately by moving the stick laterally towards the forward rudder pedal and longitudinally to neutral. Removal of the gun pods or fixing the rudders at neutral will not adversely affect the airplane's spin-recovery characteristics, but removal of the lower vertical tail will result in unsatisfactory spin-recovery characteristics. The model-test results showed that a 13.3-foot wing-tip conventional parachute (drag coefficient approximately 0.7) should be effective as an emergency spin-recovery device during demonstration spins of the airplane. It was indicated that the airplane should not tumble and that no unusual longitudinal-trim characteristics should be obtained for the center-of-gravity positions investigated.

  13. Vertical cities - the new form of high-rise construction evolution

    NASA Astrophysics Data System (ADS)

    Akristiniy, Vera A.; Boriskina, Yulia I.

    2018-03-01

    The article considers the basic principles of the vertical cities formation for the creation of a comfortable urban environment in conditions of rapid population growth and limited territories. As urban growth increases, there is a need for new concepts and approaches to urban space planning through the massive introduction of high-rise construction. The authors analyzed and systematized the list of high-tech solutions for arrangement the space of vertical cities, which are an integral part of the creation of the methodology for forming a high-rise buildings. Their concept differs in scale, presence of the big areas of public spaces, tendencies to self-sufficiency and sustainability, opportunity to offer the new unique comfortable environment to the population living in them.

  14. Organizing vertical layout environments: a forward-looking development strategy for high-rise building projects

    NASA Astrophysics Data System (ADS)

    Magay, A. A.; Bulgakova, E. A.; Zabelina, S. A.

    2018-03-01

    The article highlights issues surrounding development of high rise buildings. With the rapid increase of the global population there has been a trend for people to migrate into megacities and has caused the expansion of big city territories. This trend, coupled with the desire for a comfortable living environment, has resulted in numerous problems plaguing the megacity. This article proposes that a viable solution to the problems facing megacities is to create vertical layout environments. Potential options for creating vertical layout environments are set out below including the construction of buildings with atriums. Further, the article puts forth suggested spatial organization of the environment as well as optimal landscaping of high-rise buildings and constructions for the creation of vertical layout environments. Finally, the persuasive reasons for the adoption of vertical layout environments is that it will decrease the amount of developed urban areas, decrease traffic and increase environmental sustainability.

  15. The Impact of Sea Level Rise on Geodetic Vertical Datum of Peninsular Malaysia

    NASA Astrophysics Data System (ADS)

    Din, A. H. M.; Abazu, I. C.; Pa'suya, M. F.; Omar, K. M.; Hamid, A. I. A.

    2016-09-01

    Sea level rise is rapidly turning into major issues among our community and all levels of the government are working to develop responses to ensure these matters are given the uttermost attention in all facets of planning. It is more interesting to understand and investigate the present day sea level variation due its potential impact, particularly on our national geodetic vertical datum. To determine present day sea level variation, it is vital to consider both in-situ tide gauge and remote sensing measurements. This study presents an effort to quantify the sea level rise rate and magnitude over Peninsular Malaysia using tide gauge and multi-mission satellite altimeter. The time periods taken for both techniques are 32 years (from 1984 to 2015) for tidal data and 23 years (from 1993 to 2015) for altimetry data. Subsequently, the impact of sea level rise on Peninsular Malaysia Geodetic Vertical Datum (PMGVD) is evaluated in this study. the difference between MSL computed from 10 years (1984 - 1993) and 32 years (1984 - 2015) tidal data at Port Kelang showed that the increment of sea level is about 27mm. The computed magnitude showed an estimate of the long-term effect a change in MSL has on the geodetic vertical datum of Port Kelang tide gauge station. This will help give a new insight on the establishment of national geodetic vertical datum based on mean sea level data. Besides, this information can be used for a wide variety of climatic applications to study environmental issues related to flood and global warming in Malaysia.

  16. Vertical rise velocity of equatorial plasma bubbles estimated from Equatorial Atmosphere Radar (EAR) observations and HIRB model simulations

    NASA Astrophysics Data System (ADS)

    Tulasi Ram, S.; Ajith, K. K.; Yokoyama, T.; Yamamoto, M.; Niranjan, K.

    2017-06-01

    The vertical rise velocity (Vr) and maximum altitude (Hm) of equatorial plasma bubbles (EPBs) were estimated using the two-dimensional fan sector maps of 47 MHz Equatorial Atmosphere Radar (EAR), Kototabang, during May 2010 to April 2013. A total of 86 EPBs were observed out of which 68 were postsunset EPBs and remaining 18 EPBs were observed around midnight hours. The vertical rise velocities of the EPBs observed around the midnight hours are significantly smaller ( 26-128 m/s) compared to those observed in postsunset hours ( 45-265 m/s). Further, the vertical growth of the EPBs around midnight hours ceases at relatively lower altitudes, whereas the majority of EPBs at postsunset hours found to have grown beyond the maximum detectable altitude of the EAR. The three-dimensional numerical high-resolution bubble (HIRB) model with varying background conditions are employed to investigate the possible factors that control the vertical rise velocity and maximum attainable altitudes of EPBs. The estimated rise velocities from EAR observations at both postsunset and midnight hours are, in general, consistent with the nonlinear evolution of EPBs from the HIRB model. The smaller vertical rise velocities (Vr) and lower maximum altitudes (Hm) of EPBs during midnight hours are discussed in terms of weak polarization electric fields within the bubble due to weaker background electric fields and reduced background ion density levels.Plain Language SummaryEquatorial plasma bubbles are plasma density irregularities in the ionosphere. The radio waves passing through these irregular density structures undergo severe degradation/scintillation that could cause severe disruption of satellite-based communication and augmentation systems such as GPS navigation. These bubbles develop at geomagnetic equator, grow <span class="hlt">vertically</span>, and elongate along the field lines to latitudes away from the equator. The knowledge on bubble <span class="hlt">rise</span> velocities and their</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFD.F7006M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFD.F7006M"><span><span class="hlt">Rising</span> dynamics of a bubble confined in <span class="hlt">vertical</span> cells with rectangular cross-sections</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Murano, Mayuko; Okumura, Ko</p> <p>2017-11-01</p> <p>Recently, the drag friction acting on a fluid drop in confined space has been actively studied. Here, we investigate the <span class="hlt">rising</span> velocity of a bubble in a <span class="hlt">vertical</span> cell with a rectangular cross-section, both theoretically and experimentally, in which understanding of the drag force acting on the <span class="hlt">rising</span> bubble is crucial. Although the drag force in such confined space could involve several regimes, we study a special case in which the bubble is long and the aspect-ratio of the rectangular cross-section of the cell is high. As a result, we found new scaling law for the <span class="hlt">rising</span> velocity and the drag force, and confirmed the laws experimentally. Crossover to the <span class="hlt">rising</span> dynamics in a Hele-Shaw cell will be also discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20090023602','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20090023602"><span>Flight Test of the Lateral Stability of a 0.133-Scale Model of the Convair <span class="hlt">XFY</span>-<span class="hlt">1</span> Airplane with Windmilling Propellers at Mach Numbers from 0.70 to <span class="hlt">1</span>.12 (TED No. NACA DE 369)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hollinger, James A.; Mitcham, Grady L.</p> <p>1955-01-01</p> <p>A flight test of a rocket-propelled model of the Convair <span class="hlt">XFY</span>-<span class="hlt">1</span> airplane was conducted to determine the lateral stability and control characteristics, The 0.133-scale model had windmilling propellers for this test, which covered a Mach number range of O.70 to <span class="hlt">1</span>.12. The center of gravity was located at 13.9 percent of the mean aerodynamic chord. The methods of analysis included both a solution by vector diagrams and simple one- and two-degree-of-freedom methods. The model was both statically and dynamically stable throughout the speed range of the testa The roll damping was good, and the slope of the side-force curve varied little with speed. The rudder was effective throughout the test speed range, although it was reduced to about 43 percent of its subsonic value at supersonic speeds.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFD.F7008M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFD.F7008M"><span>Dynamics of the liquid film around elongated bubbles <span class="hlt">rising</span> in <span class="hlt">vertical</span> capillaries</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Magnini, Mirco; Khodaparast, Sepideh; Matar, Omar K.; Stone, Howard A.; Thome, John R.</p> <p>2017-11-01</p> <p>We performed a theoretical, numerical and experimental study on elongated bubbles <span class="hlt">rising</span> in <span class="hlt">vertical</span> tubes in co-current liquid flows. The flow conditions were characterized by capillary, Reynolds and Bond numbers within the range of Ca = 0.005 - 0.<span class="hlt">1</span> , Re = <span class="hlt">1</span> - 2000 and Bo = 0 - 20 . Direct numerical simulations of the two-phase flows are run with a self-improved version of OpenFOAM, implementing a coupled Level Set and Volume of Fluid method. A theoretical model based on an extension of the traditional Bretherton theory, accounting for inertia and the gravity force, is developed to obtain predictions of the profiles of the front and rear menisci of the bubble, liquid film thickness and bubble velocity. Different from the traditional theory for bubbles <span class="hlt">rising</span> in a stagnant liquid, the gravity force impacts the flow already when Bo < 4 . Gravity effects speed up the bubble compared to the Bo = 0 case, making the liquid film thicker and reducing the amplitude of the undulation on the surface of the bubble near its tail. Gravity effects are more apparent in the visco-capillary regime, i.e. when the Reynolds number is below <span class="hlt">1</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA12A..06Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA12A..06Y"><span><span class="hlt">Vertical</span> <span class="hlt">Rise</span> Velocity of Equatorial Plasma Bubbles Estimated from Equatorial Atmosphere Radar Observations and High-Resolution Bubble Model Simulations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yokoyama, T.; Ajith, K. K.; Yamamoto, M.; Niranjan, K.</p> <p>2017-12-01</p> <p>Equatorial plasma bubble (EPB) is a well-known phenomenon in the equatorial ionospheric F region. As it causes severe scintillation in the amplitude and phase of radio signals, it is important to understand and forecast the occurrence of EPBs from a space weather point of view. The development of EPBs is presently believed as an evolution of the generalized Rayleigh-Taylor instability. We have already developed a 3D high-resolution bubble (HIRB) model with a grid spacing of as small as <span class="hlt">1</span> km and presented nonlinear growth of EPBs which shows very turbulent internal structures such as bifurcation and pinching. As EPBs have field-aligned structures, the latitude range that is affected by EPBs depends on the apex altitude of EPBs over the dip equator. However, it was not easy to observe the apex altitude and <span class="hlt">vertical</span> <span class="hlt">rise</span> velocity of EPBs. Equatorial Atmosphere Radar (EAR) in Indonesia is capable of steering radar beams quickly so that the growth phase of EPBs can be captured clearly. The <span class="hlt">vertical</span> <span class="hlt">rise</span> velocities of the EPBs observed around the midnight hours are significantly smaller compared to those observed in postsunset hours. Further, the <span class="hlt">vertical</span> growth of the EPBs around midnight hours ceases at relatively lower altitudes, whereas the majority of EPBs at postsunset hours found to have grown beyond the maximum detectable altitude of the EAR. The HIRB model with varying background conditions are employed to investigate the possible factors that control the <span class="hlt">vertical</span> <span class="hlt">rise</span> velocity and maximum attainable altitudes of EPBs. The estimated <span class="hlt">rise</span> velocities from EAR observations at both postsunset and midnight hours are, in general, consistent with the nonlinear evolution of EPBs from the HIRB model.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li class="active"><span>1</span></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_1 --> <div id="page_2" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_1");'>1</a></li> <li class="active"><span>2</span></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="21"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JGRC..123.1196M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JGRC..123.1196M"><span>GPS <span class="hlt">Vertical</span> Land Motion Corrections to Sea-Level <span class="hlt">Rise</span> Estimates in the Pacific Northwest</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Montillet, J.-P.; Melbourne, T. I.; Szeliga, W. M.</p> <p>2018-02-01</p> <p>We construct coastal Pacific Northwest profiles of <span class="hlt">vertical</span> land motion (VLM) known to bias long-term tide-gauge measurements of sea-level <span class="hlt">rise</span> (SLR) and use them to estimate absolute sea-level <span class="hlt">rise</span> with respect to Earth's center of mass. Multidecade GPS measurements at 47 coastal stations along the Cascadia subduction zone show VLM varies regionally but smoothly along the Pacific coast and inland Puget Sound with rates ranging from + 4.9 to -<span class="hlt">1</span>.2 mm/yr. Puget Sound VLM is characterized by uniform subsidence at relatively slow rates of -0.<span class="hlt">1</span> to -0.3 mm/yr. Uplift rates of 4.5 mm/yr persist along the western Olympic Peninsula of northwestern Washington State and decrease southward becoming nearly 0 mm/yr south of central coastal Washington through Cape Blanco, Oregon. South of Cape Blanco, uplift increases to <span class="hlt">1</span>-2 mm/yr, peaks at 4 mm/yr near Crescent City, California, and returns to zero at Cape Mendocino, California. Using various stochastic noise models, we estimate long-term (˜50 -100 yr) relative sea-level <span class="hlt">rise</span> rates at 18 coastal Cascadia tide gauges and correct them for VLM. Uncorrected SLR rates are scattered, ranging between -2 mm/yr and + 5 mm/yr with mean 0.52 ± <span class="hlt">1</span>.59 mm/yr, whereas correcting for VLM increases the mean value to <span class="hlt">1</span>.99 mm/yr and reduces the uncertainty to ± <span class="hlt">1</span>.18 mm/yr, commensurate with, but approximately 17% higher than, twentieth century global mean.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29689415','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29689415"><span>Prediction of the rate of the <span class="hlt">rise</span> of an air bubble in nanofluids in a <span class="hlt">vertical</span> tube.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cho, Heon Ki; Nikolov, Alex D; Wasan, Darsh T</p> <p>2018-04-19</p> <p>Our recent experiments have demonstrated that when a bubble <span class="hlt">rises</span> through a nanofluid (a liquid containing dispersed nanoparticles) in a <span class="hlt">vertical</span> tube, a nanofluidic film with several particle layers is formed between the gas bubble and the glass tube wall, which significantly changes the bubble velocity due to the nanoparticle layering phenomenon in the film. We calculated the structural nanofilm viscosity as a function of the number of particle layers confined in it and found that the film viscosity increases rather steeply when the film contains only one or two particle layers. The nanofilm viscosity was found to be several times higher than the bulk viscosity of the fluid. Consequently, the Bretherton equation cannot accurately predict the rate of the <span class="hlt">rise</span> of a slow-moving long bubble in a <span class="hlt">vertical</span> tube in a nanofluid because it is valid only for very thick films and uses the bulk viscosity of the fluid. However, in this brief note, we demonstrate that the Bretherton equation can indeed be used for predicting the rate of the <span class="hlt">rise</span> of a long single bubble through a <span class="hlt">vertical</span> tube filled with a nanofluid by simply replacing the bulk viscosity with the proper structural nanofilm viscosity of the fluid. Copyright © 2018. Published by Elsevier Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011PNAS..10813019B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011PNAS..10813019B"><span>Comparing the role of absolute sea-level <span class="hlt">rise</span> and <span class="hlt">vertical</span> tectonic motions in coastal flooding, Torres Islands (Vanuatu)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ballu, Valérie; Bouin, Marie-Noëlle; Siméoni, Patricia; Crawford, Wayne C.; Calmant, Stephane; Boré, Jean-Michel; Kanas, Tony; Pelletier, Bernard</p> <p>2011-08-01</p> <p>Since the late 1990s, <span class="hlt">rising</span> sea levels around the Torres Islands (north Vanuatu, southwest Pacific) have caused strong local and international concern. In 2002-2004, a village was displaced due to increasing sea incursions, and in 2005 a United Nations Environment Programme press release referred to the displaced village as perhaps the world's first climate change "refugees." We show here that <span class="hlt">vertical</span> motions of the Torres Islands themselves dominate the apparent sea-level <span class="hlt">rise</span> observed on the islands. From 1997 to 2009, the absolute sea level rose by 150 + /-20 mm. But GPS data reveal that the islands subsided by 117 + /-30 mm over the same time period, almost doubling the apparent gradual sea-level <span class="hlt">rise</span>. Moreover, large earthquakes that occurred just before and after this period caused several hundreds of mm of sudden <span class="hlt">vertical</span> motion, generating larger apparent sea-level changes than those observed during the entire intervening period. Our results show that <span class="hlt">vertical</span> ground motions must be accounted for when evaluating sea-level change hazards in active tectonic regions. These data are needed to help communities and governments understand environmental changes and make the best decisions for their future.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21795605','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21795605"><span>Comparing the role of absolute sea-level <span class="hlt">rise</span> and <span class="hlt">vertical</span> tectonic motions in coastal flooding, Torres Islands (Vanuatu).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ballu, Valérie; Bouin, Marie-Noëlle; Siméoni, Patricia; Crawford, Wayne C; Calmant, Stephane; Boré, Jean-Michel; Kanas, Tony; Pelletier, Bernard</p> <p>2011-08-09</p> <p>Since the late 1990s, <span class="hlt">rising</span> sea levels around the Torres Islands (north Vanuatu, southwest Pacific) have caused strong local and international concern. In 2002-2004, a village was displaced due to increasing sea incursions, and in 2005 a United Nations Environment Programme press release referred to the displaced village as perhaps the world's first climate change "refugees." We show here that <span class="hlt">vertical</span> motions of the Torres Islands themselves dominate the apparent sea-level <span class="hlt">rise</span> observed on the islands. From 1997 to 2009, the absolute sea level rose by 150 + /-20 mm. But GPS data reveal that the islands subsided by 117 + /-30 mm over the same time period, almost doubling the apparent gradual sea-level <span class="hlt">rise</span>. Moreover, large earthquakes that occurred just before and after this period caused several hundreds of mm of sudden <span class="hlt">vertical</span> motion, generating larger apparent sea-level changes than those observed during the entire intervening period. Our results show that <span class="hlt">vertical</span> ground motions must be accounted for when evaluating sea-level change hazards in active tectonic regions. These data are needed to help communities and governments understand environmental changes and make the best decisions for their future.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.G11D..02P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.G11D..02P"><span><span class="hlt">Vertical</span> land motion controls regional sea level <span class="hlt">rise</span> patterns on the United States east coast since 1900</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Piecuch, C. G.; Huybers, P. J.; Hay, C.; Mitrovica, J. X.; Little, C. M.; Ponte, R. M.; Tingley, M.</p> <p>2017-12-01</p> <p>Understanding observed spatial variations in centennial relative sea level trends on the United States east coast has important scientific and societal applications. Past studies based on models and proxies variously suggest roles for crustal displacement, ocean dynamics, and melting of the Greenland ice sheet. Here we perform joint Bayesian inference on regional relative sea level, <span class="hlt">vertical</span> land motion, and absolute sea level fields based on tide gauge records and GPS data. Posterior solutions show that regional <span class="hlt">vertical</span> land motion explains most (80% median estimate) of the spatial variance in the large-scale relative sea level trend field on the east coast over 1900-2016. The posterior estimate for coastal absolute sea level <span class="hlt">rise</span> is remarkably spatially uniform compared to previous studies, with a spatial average of <span class="hlt">1</span>.4-2.3 mm/yr (95% credible interval). Results corroborate glacial isostatic adjustment models and reveal that meaningful long-period, large-scale <span class="hlt">vertical</span> velocity signals can be extracted from short GPS records.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27858504','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27858504"><span><span class="hlt">Rise</span> and Shock: Optimal Defibrillator Placement in a High-<span class="hlt">rise</span> Building.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chan, Timothy C Y</p> <p>2017-01-01</p> <p>Out-of-hospital cardiac arrests (OHCA) in high-<span class="hlt">rise</span> buildings experience lower survival and longer delays until paramedic arrival. Use of publicly accessible automated external defibrillators (AED) can improve survival, but "<span class="hlt">vertical</span>" placement has not been studied. We aim to determine whether elevator-based or lobby-based AED placement results in shorter <span class="hlt">vertical</span> distance travelled ("response distance") to OHCAs in a high-<span class="hlt">rise</span> building. We developed a model of a single-elevator, n-floor high-<span class="hlt">rise</span> building. We calculated and compared the average distance from AED to floor of arrest for the two AED locations. We modeled OHCA occurrences using floor-specific Poisson processes, the risk of OHCA on the ground floor (λ <span class="hlt">1</span> ) and the risk on any above-ground floor (λ). The elevator was modeled with an override function enabling direct travel to the target floor. The elevator location upon override was modeled as a discrete uniform random variable. Calculations used the laws of probability. Elevator-based AED placement had shorter average response distance if the number of floors (n) in the building exceeded three quarters of the ratio of ground-floor OHCA risk to above-ground floor risk (λ <span class="hlt">1</span> /λ) plus one half (n ≥ 3λ <span class="hlt">1</span> /4λ + 0.5). Otherwise, a lobby-based AED had shorter average response distance. If OHCA risk on each floor was equal, an elevator-based AED had shorter average response distance. Elevator-based AEDs travel less <span class="hlt">vertical</span> distance to OHCAs in tall buildings or those with uniform <span class="hlt">vertical</span> risk, while lobby-based AEDs travel less <span class="hlt">vertical</span> distance in buildings with substantial lobby, underground, and nearby street-level traffic and OHCA risk.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3156165','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3156165"><span>Comparing the role of absolute sea-level <span class="hlt">rise</span> and <span class="hlt">vertical</span> tectonic motions in coastal flooding, Torres Islands (Vanuatu)</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ballu, Valérie; Bouin, Marie-Noëlle; Siméoni, Patricia; Crawford, Wayne C.; Calmant, Stephane; Boré, Jean-Michel; Kanas, Tony; Pelletier, Bernard</p> <p>2011-01-01</p> <p>Since the late 1990s, <span class="hlt">rising</span> sea levels around the Torres Islands (north Vanuatu, southwest Pacific) have caused strong local and international concern. In 2002–2004, a village was displaced due to increasing sea incursions, and in 2005 a United Nations Environment Programme press release referred to the displaced village as perhaps the world’s first climate change “refugees.” We show here that <span class="hlt">vertical</span> motions of the Torres Islands themselves dominate the apparent sea-level <span class="hlt">rise</span> observed on the islands. From 1997 to 2009, the absolute sea level rose by 150 + /-20 mm. But GPS data reveal that the islands subsided by 117 + /-30 mm over the same time period, almost doubling the apparent gradual sea-level <span class="hlt">rise</span>. Moreover, large earthquakes that occurred just before and after this period caused several hundreds of mm of sudden <span class="hlt">vertical</span> motion, generating larger apparent sea-level changes than those observed during the entire intervening period. Our results show that <span class="hlt">vertical</span> ground motions must be accounted for when evaluating sea-level change hazards in active tectonic regions. These data are needed to help communities and governments understand environmental changes and make the best decisions for their future. PMID:21795605</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E3SWC..3301010L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E3SWC..3301010L"><span>Gardening as vector of a humanization of high-<span class="hlt">rise</span> building</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lekareva, Nina; Zaslavskaya, Anna</p> <p>2018-03-01</p> <p>Article is devoted to issue of integration of <span class="hlt">vertical</span> gardening into structure of high-<span class="hlt">rise</span> building in the conditions of the constrained town-planning situation. On the basis of the analysis of the existing experience of design and building of "biopositive" high-<span class="hlt">rise</span> building ecological, town-planning, social and constructive advantages of the organization of gardens on roofs and <span class="hlt">vertical</span> gardens are considered [<span class="hlt">1</span>]. As the main mechanism of increase in investment appeal of high-<span class="hlt">rise</span> building the principle of a humanization due to gardening of high-<span class="hlt">rise</span> building taking into account requirements of ecology, energy efficiency of buildings and improvement of quality of construction with minimization of expenses and maximizing comfort moves forward. The National Standards of Green construction designed to adapt the international requirements of architecture and construction of the energy efficient, eco-friendly and comfortable building or a complex to local conditions are considered [2,3].</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20090023311','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20090023311"><span>Longitudinal Stability Characteristics of the Consolidated Vultee <span class="hlt">XFY</span>-<span class="hlt">1</span> Airplane with Windmilling Propellers as Obtained from Flight of 0.133-Scale Rocket-Propelled Model at Mach Numbers from 0.70 to <span class="hlt">1</span>.13</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hastings, Earl C.; Mitcham, Grady L.</p> <p>1954-01-01</p> <p>A flight test has been conducted to determine the longitudinal stability and control characteristics of a 0.133-scale model of the Consolidated Vultee <span class="hlt">XFY</span>-<span class="hlt">1</span> airplane with windmilling propellers for the Mach number range between 0.70 and <span class="hlt">1</span>.13. The variation of lift-curve slope C(sub L(sub alpha) with Mach number was gradual with a maximum value of 0.074 occurring at a Mach number of 0.97. Propellers had little effect upon the values of lift-curve slope or the linearity of lift coefficient with angle of attack. At lift coefficients between approximately 0.25 and 0.45 with an elevon angle of approximately -l0 deg, there was a region of neutral longitudinal stability at Mach numbers below 0.93 introduced by the addition of windmilling propellers. Below a lift coefficient of 0.10 and above a lift coefficient of 0.45, the model was longitudinally stable throughout the Mach number range of the test. There was a forward shift in the aerodynamic center of about 3-percent mean aerodynamic chord introduced by the addition of propellers. The aerodynamic center as determined at low lift moved gradually from a value of 28.5-percent mean aerodynamic chord at a Mach number of 0.75 to a value of 47-percent mean aerodynamic chord at a Mach number of <span class="hlt">1</span>.10. There was an abrupt decrease in pitch damping between Mach numbers of 0.88 and 0.99 followed by a rapid increase in damping to a Mach number of <span class="hlt">1</span>.06. The propellers had little effect upon the pitch damping characteristics . The transonic trim change was a large pitching-down tendency with and without windmilling propellers. The elevons were effective pitch controls throughout the speed range; however, their effectiveness was reduced about 50 percent at supersonic speeds. The propellers had no appreciable effect upon the control effectiveness.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1431478','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1431478"><span>Historical Cavern Floor <span class="hlt">Rise</span> for All SPR Sites</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Moriarty, Dylan Michael</p> <p>2016-09-01</p> <p>The Strategic Petroleum Reserve (SPR) contains the largest supply is the largest stockpile of government-owned emergency crude oil in the world. The oil is stored in multiple salt caverns spread over four sites in Louisiana and Texas. Cavern infrastructure near the bottom of the cavern can be damaged from <span class="hlt">vertical</span> floor movement. This report presents a comprehensive history of floor movements in each cavern. Most of the cavern floor <span class="hlt">rise</span> rates ranged from 0.5-3.5 ft/yr, however, there were several caverns with much higher <span class="hlt">rise</span> rates. BH103, BM106, and BH105 had the three highest <span class="hlt">rise</span> rates. Information from this report willmore » be used to better predict future <span class="hlt">vertical</span> floor movements and optimally place cavern infrastructure. The reasons for floor <span class="hlt">rise</span> are not entirely understood and should be investigated.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22150642','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22150642"><span>The effect of <span class="hlt">rising</span> vs. falling glucose level on amperometric glucose sensor lag and accuracy in Type <span class="hlt">1</span> diabetes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ward, W K; Engle, J M; Branigan, D; El Youssef, J; Massoud, R G; Castle, J R</p> <p>2012-08-01</p> <p>Because declining glucose levels should be detected quickly in persons with Type <span class="hlt">1</span> diabetes, a lag between blood glucose and subcutaneous sensor glucose can be problematic. It is unclear whether the magnitude of sensor lag is lower during falling glucose than during <span class="hlt">rising</span> glucose. Initially, we analysed 95 data segments during which glucose changed and during which very frequent reference blood glucose monitoring was performed. However, to minimize confounding effects of noise and calibration error, we excluded data segments in which there was substantial sensor error. After these exclusions, and combination of data from duplicate sensors, there were 72 analysable data segments (36 for <span class="hlt">rising</span> glucose, 36 for falling). We measured lag in two ways: (<span class="hlt">1</span>) the time delay at the <span class="hlt">vertical</span> mid-point of the glucose change (regression delay); and (2) determination of the optimal time shift required to minimize the difference between glucose sensor signals and blood glucose values drawn concurrently. Using the regression delay method, the mean sensor lag for <span class="hlt">rising</span> vs. falling glucose segments was 8.9 min (95%CI 6.<span class="hlt">1</span>-11.6) vs. <span class="hlt">1</span>.5 min (95%CI -2.6 to 5.5, P<0.005). Using the time shift optimization method, results were similar, with a lag that was higher for <span class="hlt">rising</span> than for falling segments [8.3 (95%CI 5.8-10.7) vs. <span class="hlt">1</span>.5 min (95% CI -2.2 to 5.2), P<0.001]. Commensurate with the lag results, sensor accuracy was greater during falling than during <span class="hlt">rising</span> glucose segments. In Type <span class="hlt">1</span> diabetes, when noise and calibration error are minimized to reduce effects that confound delay measurement, subcutaneous glucose sensors demonstrate a shorter lag duration and greater accuracy when glucose is falling than when <span class="hlt">rising</span>. © 2011 The Authors. Diabetic Medicine © 2011 Diabetes UK.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17385599','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17385599"><span>Stable plume <span class="hlt">rise</span> in a shear layer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Overcamp, Thomas J</p> <p>2007-03-01</p> <p>Solutions are given for plume <span class="hlt">rise</span> assuming a power-law wind speed profile in a stably stratified layer for point and finite sources with initial <span class="hlt">vertical</span> momentum and buoyancy. For a constant wind speed, these solutions simplify to the conventional plume <span class="hlt">rise</span> equations in a stable atmosphere. In a shear layer, the point of maximum <span class="hlt">rise</span> occurs further downwind and is slightly lower compared with the plume <span class="hlt">rise</span> with a constant wind speed equal to the wind speed at the top of the stack. If the predictions with shear are compared with predictions for an equivalent average wind speed over the depth of the plume, the plume <span class="hlt">rise</span> with shear is higher than plume <span class="hlt">rise</span> with an equivalent average wind speed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Heel&id=EJ961912','ERIC'); return false;" href="https://eric.ed.gov/?q=Heel&id=EJ961912"><span>Reliability and Validity of the Standing Heel-<span class="hlt">Rise</span> Test</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Yocum, Allison; McCoy, Sarah Westcott; Bjornson, Kristie F.; Mullens, Pamela; Burton, Gay Naganuma</p> <p>2010-01-01</p> <p>A standardized protocol for a pediatric heel-<span class="hlt">rise</span> test was developed and reliability and validity are reported. Fifty-seven children developing typically (CDT) and 34 children with plantar flexion weakness performed three tests: unilateral heel <span class="hlt">rise</span>, <span class="hlt">vertical</span> jump, and force measurement using handheld dynamometry. Intraclass correlation…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5082943','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5082943"><span>Modeled Tradeoffs between Developed Land Protection and Tidal Habitat Maintenance during <span class="hlt">Rising</span> Sea Levels</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Cadol, Daniel; Elmore, Andrew J.; Guinn, Steven M.; Engelhardt, Katharina A. M.; Sanders, Geoffrey</p> <p>2016-01-01</p> <p>Tidal habitats host a diversity of species and provide hydrological services such as shoreline protection and nutrient attenuation. Accretion of sediment and biomass enables tidal marshes and swamps to grow <span class="hlt">vertically</span>, providing a degree of resilience to <span class="hlt">rising</span> sea levels. Even if accelerating sea level <span class="hlt">rise</span> overcomes this <span class="hlt">vertical</span> resilience, tidal habitats have the potential to migrate inland as they continue to occupy land that falls within the new tide range elevations. The existence of developed land inland of tidal habitats, however, may prevent this migration as efforts are often made to dyke and protect developments. To test the importance of inland migration to maintaining tidal habitat abundance under a range of potential rates of sea level <span class="hlt">rise</span>, we developed a spatially explicit elevation tracking and habitat switching model, dubbed the Marsh Accretion and Inundation Model (MAIM), which incorporates elevation-dependent net land surface elevation gain functions. We applied the model to the metropolitan Washington, DC region, finding that the abundance of small National Park Service units and other public open space along the tidal Potomac River system provides a refuge to which tidal habitats may retreat to maintain total habitat area even under moderate sea level <span class="hlt">rise</span> scenarios (0.7 m and <span class="hlt">1.1</span> m <span class="hlt">rise</span> by 2100). Under a severe sea level <span class="hlt">rise</span> scenario associated with ice sheet collapse (<span class="hlt">1</span>.7 m by 2100) habitat area is maintained only if no development is protected from <span class="hlt">rising</span> water. If all existing development is protected, then 5%, 10%, and 40% of the total tidal habitat area is lost by 2100 for the three sea level <span class="hlt">rise</span> scenarios tested. PMID:27788209</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27788209','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27788209"><span>Modeled Tradeoffs between Developed Land Protection and Tidal Habitat Maintenance during <span class="hlt">Rising</span> Sea Levels.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cadol, Daniel; Elmore, Andrew J; Guinn, Steven M; Engelhardt, Katharina A M; Sanders, Geoffrey</p> <p>2016-01-01</p> <p>Tidal habitats host a diversity of species and provide hydrological services such as shoreline protection and nutrient attenuation. Accretion of sediment and biomass enables tidal marshes and swamps to grow <span class="hlt">vertically</span>, providing a degree of resilience to <span class="hlt">rising</span> sea levels. Even if accelerating sea level <span class="hlt">rise</span> overcomes this <span class="hlt">vertical</span> resilience, tidal habitats have the potential to migrate inland as they continue to occupy land that falls within the new tide range elevations. The existence of developed land inland of tidal habitats, however, may prevent this migration as efforts are often made to dyke and protect developments. To test the importance of inland migration to maintaining tidal habitat abundance under a range of potential rates of sea level <span class="hlt">rise</span>, we developed a spatially explicit elevation tracking and habitat switching model, dubbed the Marsh Accretion and Inundation Model (MAIM), which incorporates elevation-dependent net land surface elevation gain functions. We applied the model to the metropolitan Washington, DC region, finding that the abundance of small National Park Service units and other public open space along the tidal Potomac River system provides a refuge to which tidal habitats may retreat to maintain total habitat area even under moderate sea level <span class="hlt">rise</span> scenarios (0.7 m and <span class="hlt">1.1</span> m <span class="hlt">rise</span> by 2100). Under a severe sea level <span class="hlt">rise</span> scenario associated with ice sheet collapse (<span class="hlt">1</span>.7 m by 2100) habitat area is maintained only if no development is protected from <span class="hlt">rising</span> water. If all existing development is protected, then 5%, 10%, and 40% of the total tidal habitat area is lost by 2100 for the three sea level <span class="hlt">rise</span> scenarios tested.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1926b0013D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1926b0013D"><span>Some special values of <span class="hlt">vertices</span> of trees on the suborbital graphs</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Deǧer, A. H.; Akbaba, Ü.</p> <p>2018-01-01</p> <p>In the present study, the action of a congruence subgroup of S L(2, Z) on ℚ ^ is examined. From this action and its properties, <span class="hlt">vertices</span> of paths of minimal length on the suborbital graph Fu,N give <span class="hlt">rise</span> to some special sequence values, that are alternate sequences such as identity, Fibonacci and Lucas sequences. These types of <span class="hlt">vertices</span> also give <span class="hlt">rise</span> to special continued fractions, hence from recurrence relations for continued fractions, values of these <span class="hlt">vertices</span> and values of special sequences were associated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70098419','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70098419"><span>How mangrove forests adjust to <span class="hlt">rising</span> sea level</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Krauss, Ken W.; McKee, Karen L.; Lovelock, Catherine E.; Cahoon, Donald R.; Saintilan, Neil; Reef, Ruth; Chen, Luzhen</p> <p>2014-01-01</p> <p>Mangroves are among the most well described and widely studied wetland communities in the world. The greatest threats to mangrove persistence are deforestation and other anthropogenic disturbances that can compromise habitat stability and resilience to sea-level <span class="hlt">rise</span>. To persist, mangrove ecosystems must adjust to <span class="hlt">rising</span> sea level by building <span class="hlt">vertically</span> or become submerged. Mangroves may directly or indirectly influence soil accretion processes through the production and accumulation of organic matter, as well as the trapping and retention of mineral sediment. In this review, we provide a general overview of research on mangrove elevation dynamics, emphasizing the role of the vegetation in maintaining soil surface elevations (i.e. position of the soil surface in the <span class="hlt">vertical</span> plane). We summarize the primary ways in which mangroves may influence sediment accretion and <span class="hlt">vertical</span> land development, for example, through root contributions to soil volume and upward expansion of the soil surface. We also examine how hydrological, geomorphological and climatic processes may interact with plant processes to influence mangrove capacity to keep pace with <span class="hlt">rising</span> sea level. We draw on a variety of studies to describe the important, and often under-appreciated, role that plants play in shaping the trajectory of an ecosystem undergoing change.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMNH31B0212H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMNH31B0212H"><span>Estimating <span class="hlt">Vertical</span> Land Motion in the Chesapeake Bay</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Houttuijn Bloemendaal, L.; Hensel, P.</p> <p>2017-12-01</p> <p>This study aimed to provide a modern measurement of subsidence in the Chesapeake Bay region and establish a methodology for measuring <span class="hlt">vertical</span> land motion using static GPS, a cheaper alternative to InSAR or classical leveling. <span class="hlt">Vertical</span> land motion in this area is of particular concern because tide gages are showing up to 5 mm/yr of local, relative sea level <span class="hlt">rise</span>. While a component of this rate is the actual eustatic sea level <span class="hlt">rise</span> itself, part of the trend may also be <span class="hlt">vertical</span> land motion, in which subsidence exacerbates the effects of actual changes in sea level. Parts of this region are already experiencing an increase in the frequency and magnitude of near-shore coastal flooding, but the last comprehensive study of <span class="hlt">vertical</span> land motion in this area was conducted by NOAA in 1974 (Holdahl & Morrison) using repeat leveled lines. More recent measures of <span class="hlt">vertical</span> land motion can help inform efforts on resilience to sea level <span class="hlt">rise</span>, such as in the Hampton Roads area. This study used measured GPS-derived <span class="hlt">vertical</span> heights in conjunction with legacy GPS data to calculate rates of <span class="hlt">vertical</span> motion at several points in time for a selection of benchmarks scattered throughout the region. Seventeen marks in the stable Piedmont area and in the areas suspected of subsidence in the Coastal Plain were selected for the analysis. Results indicate a significant difference between the rates of <span class="hlt">vertical</span> motion in the Piedmont and Coastal Plain, with a mean rate of -4.10 mm/yr in the Coastal Plain and 0.15 mm/yr in the Piedmont. The rates indicate particularly severe subsidence at the southern Delmarva Peninsula coast and the Hampton-Roads area, with a mean rate of -6.57 mm/yr in that region. By knowing local rates of subsidence as opposed to sea level change itself, coastal managers may make better informed decisions regarding natural resource use, such as deciding whether or not to reduce subsurface fluid withdrawals or to consider injecting treated water back into the aquifer to slow</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JIEIA..98..237O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JIEIA..98..237O"><span>Disposal of Kitchen Waste from High <span class="hlt">Rise</span> Apartment</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ori, Kirki; Bharti, Ajay; Kumar, Sunil</p> <p>2017-09-01</p> <p>The high <span class="hlt">rise</span> building has numbers of floor and rooms having variety of users or tenants for residential purposes. The huge quantities of heterogenous mixtures of domestic food waste are generated from every floor of the high <span class="hlt">rise</span> residential buildings. Disposal of wet and biodegradable domestic kitchen waste from high <span class="hlt">rise</span> buildings are more expensive in regards of collection and <span class="hlt">vertical</span> transportation. This work is intended to address the technique to dispose of the wet organic food waste from the high <span class="hlt">rise</span> buildings or multistory building at generation point with the advantage of gravity and vermicomposting technique. This innovative effort for collection and disposal of wet organic solid waste from high <span class="hlt">rise</span> apartment is more economical and hygienic in comparison with present system of disposal.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009EGUGA..11.6006A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009EGUGA..11.6006A"><span>Sea level <span class="hlt">rise</span> within the west of Arabian Gulf using tide gauge and continuous GPS measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ayhan, M. E.; Alothman, A.</p> <p>2009-04-01</p> <p>Arabian Gulf is connected to Indian Ocean and located in the south-west of the Zagros Trust Belt. To investigate sea level variations within the west of Arabian Gulf, monthly means of sea level at 13 tide gauges along the coast of Saudi Arabia and Bahrain, available in the database of the Permanent Service for Mean Sea Level (PSMSL), are studied. We analyzed individually the monthly means at each station, and estimated secular sea level rate by a robust linear trend fitting. We computed the average relative sea level <span class="hlt">rise</span> rate of <span class="hlt">1</span>.96 ± 0.21 mm/yr within the west of Arabian Gulf based on 4 stations spanning longer than 19 years. <span class="hlt">Vertical</span> land motions are included into the relative sea level measurements at the tide gauges. Therefore sea level rates at the stations are corrected for <span class="hlt">vertical</span> land motions using the ICE-5G v<span class="hlt">1</span>.2 VM4 Glacial Isostatic Adjustment (GIA) model then we found the average sea level <span class="hlt">rise</span> rate of 2.27 mm/yr. Bahrain International GPS Service (IGS) GPS station, which is close to the Mina Sulman tide gauge station in Bahrain, is the only continuous GPS station accessible in the region. The weekly GPS time series of <span class="hlt">vertical</span> component at Bahrain IGS-GPS station referring to the ITRF97 from 1999.2 to 2008.6 are downloaded from http://www-gps.mit.edu/~tah/. We fitted a linear trend with an annual signal and one break to the GPS <span class="hlt">vertical</span> time series and found a <span class="hlt">vertical</span> land motion rate of 0.48 ± 0.11 mm/yr. Assuming the <span class="hlt">vertical</span> rate at Bahrain IGS-GPS station represents the <span class="hlt">vertical</span> rate at each of the other tide gauge stations studied here in the region, we computed average sea level <span class="hlt">rise</span> rate of 2.44 ± 0.21 mm/yr within the west of Arabian Gulf.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_1");'>1</a></li> <li class="active"><span>2</span></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_2 --> <div id="page_3" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_1");'>1</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li class="active"><span>3</span></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="41"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5478375','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5478375"><span>Mapping the dynamics of cortical neuroplasticity of skilled motor learning using micro X-ray fluorescence and histofluorescence imaging of zinc in the rat</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Alaverdashvili, Mariam; Paterson, Phyllis G.</p> <p>2017-01-01</p> <p>Synchrotron-based X-ray fluorescence imaging (<span class="hlt">XFI</span>) of zinc (Zn) has been recently implemented to understand the efficiency of various therapeutic interventions targeting post-stroke neuroprotection and neuroplasticity. However, it is uncertain if micro <span class="hlt">XFI</span> can resolve neuroplasticity-induced changes. Thus, we explored if learning-associated behavioral changes would be accompanied by changes in cortical Zn concentration measured by <span class="hlt">XFI</span> in healthy adult rats. Proficiency in a skilled reach-to-eat task during early and late stages of motor learning served as a functional measure of neuroplasticity. c-Fos protein and vesicular Zn expression were employed as indirect neuronal measures of brain plasticity. A total Zn map (20 × 20 × 30 μm3 resolution) generated by micro <span class="hlt">XFI</span> failed to reflect increases in either c-Fos or vesicular Zn in the motor cortex contralateral to the trained forelimb or improved proficiency in the skilled reaching task. Remarkably, vesicular Zn increased in the late stage of motor learning along with a concurrent decrease in the number of c-fos-ip neurons relative to the early stage of motor learning. This inverse dynamics of c-fos and vesicular Zn level as the motor skill advances suggest that a qualitatively different neural population, comprised of fewer active but more efficiently connected neurons, supports a skilled action in the late versus early stage of motor learning. The lack of sensitivity of the <span class="hlt">XFI</span>-generated Zn map to visualize the plasticity-associated changes in vesicular Zn suggests that the Zn level measured by micro <span class="hlt">XFI</span> should not be used as a surrogate marker of neuroplasticity in response to the acquisition of skilled motor actions. Nanoscopic <span class="hlt">XFI</span> could be explored in future as a means of imaging these subtle physiological changes. PMID:27840249</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/56156','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/56156"><span>Optimizing smoke and plume <span class="hlt">rise</span> modeling approaches at local scales</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Derek V. Mallia; Adam K. Kochanski; Shawn P. Urbanski; John C. Lin</p> <p>2018-01-01</p> <p>Heating from wildfires adds buoyancy to the overlying air, often producing plumes that <span class="hlt">vertically</span> distribute fire emissions throughout the atmospheric column over the fire. The height of the <span class="hlt">rising</span> wildfire plume is a complex function of the size of the wildfire, fire heat flux, plume geometry, and atmospheric conditions, which can make simulating plume <span class="hlt">rises</span> difficult...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/39381','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/39381"><span>Validation of smoke plume <span class="hlt">rise</span> models using ground based lidar</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Cyle E. Wold; Shawn Urbanski; Vladimir Kovalev; Alexander Petkov; Wei Min Hao</p> <p>2010-01-01</p> <p>Biomass fires can significantly degrade regional air quality. Plume <span class="hlt">rise</span> height is one of the critical factors determining the impact of fire emissions on air quality. Plume <span class="hlt">rise</span> models are used to prescribe the <span class="hlt">vertical</span> distribution of fire emissions which are critical input for smoke dispersion and air quality models. The poor state of model evaluation is due in...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3509803','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3509803"><span>TRANSMISIÓN <span class="hlt">VERTICAL</span> DE HTLV-<span class="hlt">1</span> EN EL PERÚ</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Villaverde, Jorge Alarcón; Romaní, Franco Romaní; Torres, Silvia Montano; Zunt, Joseph R.</p> <p>2012-01-01</p> <p>La infección por el virus linfotrópico humano de células T tipo <span class="hlt">1</span> (HTLV-<span class="hlt">1</span>) ha sido descrita en muchas áreas del mundo, como en los países del Caribe, Japón, África, Oceanía y en Sudamérica. En la presente revisión definimos la endemicidad del HTLV-<span class="hlt">1</span> en el país, planteando cuatro criterios epidemiológicos. Luego discutimos el tema central de la revisión: la transmisión <span class="hlt">vertical</span> del HTLV-<span class="hlt">1</span>, que en nuestro país sería uno de los principales mecanismos de transmisión. Dentro del desarrollo de este aspecto en particular, presentamos una estimación de la tasa de transmisión <span class="hlt">vertical</span> y los factores de riesgo asociados con la transmisión <span class="hlt">vertical</span> sobre la base de una revisión exhaustiva de estudios nacionales y extranjeros. Con esta revisión pretendemos dar una primera aproximación al estudio de la trasmisión <span class="hlt">vertical</span> de HTLV-<span class="hlt">1</span>, un aspecto poco estudiado en nuestro medio. PMID:21537777</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4632590','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4632590"><span>Keep up or drown: adjustment of western Pacific coral reefs to sea-level <span class="hlt">rise</span> in the 21st century</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>van Woesik, R.; Golbuu, Y.; Roff, G.</p> <p>2015-01-01</p> <p>Since the Mid-Holocene, some 5000 years ago, coral reefs in the Pacific Ocean have been <span class="hlt">vertically</span> constrained by sea level. Contemporary sea-level <span class="hlt">rise</span> is releasing these constraints, providing accommodation space for <span class="hlt">vertical</span> reef expansion. Here, we show that Porites microatolls, from reef-flat environments in Palau (western Pacific Ocean), are ‘keeping up’ with contemporary sea-level <span class="hlt">rise</span>. Measurements of 570 reef-flat Porites microatolls at 10 locations around Palau revealed recent <span class="hlt">vertical</span> skeletal extension (78±13 mm) over the last 6–8 years, which is consistent with the timing of the recent increase in sea level. We modelled whether microatoll growth rates will potentially ‘keep up’ with predicted sea-level <span class="hlt">rise</span> in the near future, based upon average growth, and assuming a decline in growth for every <span class="hlt">1</span>°C increase in temperature. We then compared these estimated extension rates with rates of sea-level <span class="hlt">rise</span> under four Representative Concentration Pathways (RCPs). Our model suggests that under low–mid RCP scenarios, reef-coral growth will keep up with sea-level <span class="hlt">rise</span>, but if greenhouse gas concentrations exceed 670 ppm atmospheric CO2 levels and with +2.2°C sea-surface temperature by 2100 (RCP 6.0 W m−2), our predictions indicate that Porites microatolls will be unable to keep up with projected rates of sea-level <span class="hlt">rise</span> in the twenty-first century. PMID:26587277</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E3SWC..3301008S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E3SWC..3301008S"><span>High-<span class="hlt">rise</span> construction in the Saint Petersburg agglomeration in 1703-1950s</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sementsov, Sergey; Akulova, Nadezhda; Kurakina, Severina</p> <p>2018-03-01</p> <p>Regularities of high-<span class="hlt">rise</span> construction (implemented projects and developments) in Saint Petersburg and the Saint Petersburg agglomeration since the foundation of the city in 1703 till the 1950s are considered. Based on these regularities, a single spatially developed system of <span class="hlt">vertical</span> dominants is formed. High-<span class="hlt">rise</span> construction in the city and its suburbs started in the 1710s and continues up to the present time. In the considered decades (1703-1950s), high-<span class="hlt">rise</span> construction mostly performed urban-planning functions (with <span class="hlt">vertical</span> and symbolic dominants), relying on patterns of the visual perception of man-made landscapes under development. Since the 1710s, the construction of <span class="hlt">vertical</span> dominants (mainly temples, spires of towers, lighthouses, etc.) of five ranks (depending on the altitude range and in relation to the background development) was conducted in territories of the entire agglomeration. These dominants were arranged in landscapes of the city and suburbs with almost mathematically precise accuracy and according to special regulations. Such dominants obtained particular descriptive and silhouette characteristics in accordance with the conditions of spatial perception. In some periods of city development, attempts were made to create monuments (symbolic dominants) of specific height and include those in the spatial system of high-<span class="hlt">rise</span> dominants as significant elements of the city silhouette.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20180001313','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20180001313"><span>Chapter 12: Sea Level <span class="hlt">Rise</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sweet, W. V.; Horton, R.; Kopp, R. E.; LeGrande, A. N.; Romanou, A.</p> <p>2017-01-01</p> <p>Global mean sea level (GMSL) has risen by about 7-8 inches (about 16-21 cm) since 1900, with about 3 of those inches (about 7 cm) occurring since 1993. Human-caused climate change has made a substantial contribution to GMSL <span class="hlt">rise</span> since 1900, contributing to a rate of <span class="hlt">rise</span> that is greater than during any preceding century in at least 2,800 years. Relative to the year 2000, GMSL is very likely to <span class="hlt">rise</span> by 0.3-0.6 feet (9-18 cm) by 2030, 0.5-<span class="hlt">1</span>.2 feet (15-38 cm) by 2050, and <span class="hlt">1</span>.0-4.3 feet (30-130 cm) by 2100. Future pathways have little effect on projected GMSL <span class="hlt">rise</span> in the first half of the century, but significantly affect projections for the second half of the century. Emerging science regarding Antarctic ice sheet stability suggests that, for high emission scenarios, a GMSL <span class="hlt">rise</span> exceeding 8 feet (2.4 m) by 2100 is physically possible, although the probability of such an extreme outcome cannot currently be assessed. Regardless of pathway, it is extremely likely that GMSL <span class="hlt">rise</span> will continue beyond 2100. Relative sea level (RSL) <span class="hlt">rise</span> in this century will vary along U.S. coastlines due, in part, to changes in Earth's gravitational field and rotation from melting of land ice, changes in ocean circulation, and <span class="hlt">vertical</span> land motion (very high confidence). For almost all future GMSL <span class="hlt">rise</span> scenarios, RSL <span class="hlt">rise</span> is likely to be greater than the global average in the U.S. Northeast and the western Gulf of Mexico. In intermediate and low GMSL <span class="hlt">rise</span> scenarios, RSL <span class="hlt">rise</span> is likely to be less than the global average in much of the Pacific Northwest and Alaska. For high GMSL <span class="hlt">rise</span> scenarios, RSL <span class="hlt">rise</span> is likely to be higher than the global average along all U.S. coastlines outside Alaska. Almost all U.S. coastlines experience more than global mean sea level <span class="hlt">rise</span> in response to Antarctic ice loss, and thus would be particularly affected under extreme GMSL <span class="hlt">rise</span> scenarios involving substantial Antarctic mass loss. As sea levels have risen, the number of tidal floods each year that cause minor</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://dx.doi.org/10.1023/A:1009904816246','USGSPUBS'); return false;" href="http://dx.doi.org/10.1023/A:1009904816246"><span><span class="hlt">Vertical</span> accretion and shallow subsidence in a mangrove forest of southwestern Florida, U.S.A</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Cahoon, D.R.; Lynch, J.C.</p> <p>1997-01-01</p> <p>Simultaneous measurements of <span class="hlt">vertical</span> accretion from artificial soil marker horizons and soil elevation change from sedimentation-erosion table (SET) plots were used to evaluate the processes related to soil building in range, basin, and overwash mangrove forests located in a low-energy lagoon which recieves minor inputs of terregenous sediments. <span class="hlt">Vertical</span> accretion measures reflect the contribution of surficial sedimentation (sediment deposition and surface root growth). Measures of elevation change reflect not only the contributions of <span class="hlt">vertical</span> accretion but also those of subsurface processes such as compaction, decomposition and shrink-swell. The two measures were used to calculate amounts of shallow subsidence (accretion minus elevation change) in each mangrove forest. The three forest types represent different accretionary envrionments. The basin forest was located behind a natural berm. Hydroperiod here was controlled primarily by rainfall rather than tidal exchange, although the basin flooded during extreme tidal events. Soil accretion here occurred primarily by autochthonous organic matter inputs, and elevation was controlled by accretion and shrink-swell of the substrate apparently related to cycles of flooding-drying and/or root growth-decomposition. This hydrologically-restricted forest did not experience an accretion or elevation deficit relative to sea-level <span class="hlt">rise</span>. The tidally dominated fringe and overwash island forests accreted through mineral sediment inputs bound in place by plant roots. Filamentous turf algae played an important role in stabilizing loose muds in the fringe forest where erosion was prevalent. Elevation in these high-energy environments was controlled not only by accretion but also by erosion and/or shallow subsidence. The rate of shallow subsidence was consistently 3-4 mm y-<span class="hlt">1</span> in the fringe and overwash island forests but was negligible in the basin forest. Hence, the <span class="hlt">vertical</span> development of mangrove soils was influenced by both</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006SPIE.6230E..17S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006SPIE.6230E..17S"><span>The <span class="hlt">RiSE</span> climbing robot: body and leg design</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Saunders, A.; Goldman, D. I.; Full, R. J.; Buehler, M.</p> <p>2006-05-01</p> <p>The <span class="hlt">RiSE</span> robot is a biologically inspired, six legged climbing robot, designed for general mobility in scansorial (<span class="hlt">vertical</span> walls, horizontal ledges, ground level) environments. It exhibits ground reaction forces that are similar to animal climbers and does not rely on suction, magnets or other surface-dependent specializations to achieve adhesion and shear force. We describe <span class="hlt">RiSE</span>'s body and leg design as well as its electromechanical, communications and computational infrastructure. We review design iterations that enable <span class="hlt">RiSE</span> to climb 90° carpeted, cork covered and (a growing range of) stucco surfaces in the quasi-static regime.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110013482','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110013482"><span><span class="hlt">Vertical</span> Crustal Motion Derived from Satellite Altimetry and Tide Gauges, and Comparisons with DORIS Measurements</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ray, R. D.; Beckley, B. D.; Lemoine, F. G.</p> <p>2010-01-01</p> <p>A somewhat unorthodox method for determining <span class="hlt">vertical</span> crustal motion at a tide-gauge location is to difference the sea level time series with an equivalent time series determined from satellite altimetry, To the extent that both instruments measure an identical ocean signal, the difference will be dominated by <span class="hlt">vertical</span> land motion at the gauge. We revisit this technique by analyzing sea level signals at 28 tide gauges that are colocated with DORIS geodetic stations. Comparisons of altimeter-gauge <span class="hlt">vertical</span> rates with DORIS rates yield a median difference of <span class="hlt">1</span>.8 mm/yr and a weighted root-mean-square difference of2.7 mm/yr. The latter suggests that our uncertainty estimates, which are primarily based on an assumed AR(l) noise process in all time series, underestimates the true errors. Several sources of additional error are discussed, including possible scale errors in the terrestrial reference frame to which altimeter-gauge rates are mostly insensitive, One of our stations, Male, Maldives, which has been the subject of some uninformed arguments about sea-level <span class="hlt">rise</span>, is found to have almost no <span class="hlt">vertical</span> motion, and thus is vulnerable to <span class="hlt">rising</span> sea levels. Published by Elsevier Ltd. on behalf of COSPAR.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70029941','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70029941"><span>Mars reconnaissance orbiter's high resolution imaging science experiment (Hi<span class="hlt">RISE</span>)</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>McEwen, A.S.; Eliason, E.M.; Bergstrom, J.W.; Bridges, N.T.; Hansen, C.J.; Delamere, W.A.; Grant, J. A.; Gulick, V.C.; Herkenhoff, K. E.; Keszthelyi, L.; Kirk, R.L.; Mellon, M.T.; Squyres, S. W.; Thomas, N.; Weitz, C.M.</p> <p>2007-01-01</p> <p>The Hi<span class="hlt">RISE</span> camera features a 0.5 m diameter primary mirror, 12 m effective focal length, and a focal plane system that can acquire images containing up to 28 Gb (gigabits) of data in as little as 6 seconds. Hi<span class="hlt">RISE</span> will provide detailed images (0.25 to <span class="hlt">1</span>.3 m/pixel) covering ???<span class="hlt">1</span>% of the Martian surface during the 2-year Primary Science Phase (PSP) beginning November 2006. Most images will include color data covering 20% of the potential field of view. A top priority is to acquire ???1000 stereo pairs and apply precision geometric corrections to enable topographic measurements to better than 25 cm <span class="hlt">vertical</span> precision. We expect to return more than 12 Tb of Hi<span class="hlt">RISE</span> data during the 2-year PSP, and use pixel binning, conversion from 14 to 8 bit values, and a lossless compression system to increase coverage. Hi<span class="hlt">RISE</span> images are acquired via 14 CCD detectors, each with 2 output channels, and with multiple choices for pixel binning and number of Time Delay and Integration lines. Hi<span class="hlt">RISE</span> will support Mars exploration by locating and characterizing past, present, and future landing sites, unsuccessful landing sites, and past and potentially future rover traverses. We will investigate cratering, volcanism, tectonism, hydrology, sedimentary processes, stratigraphy, aeolian processes, mass wasting, landscape evolution, seasonal processes, climate change, spectrophotometry, glacial and periglacial processes, polar geology, and regolith properties. An Internet Web site (HiWeb) will enable anyone in the world to suggest Hi<span class="hlt">RISE</span> targets on Mars and to easily locate, view, and download Hi<span class="hlt">RISE</span> data products. Copyright 2007 by the American Geophysical Union.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70194542','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70194542"><span>Will fluctuations in salt marsh–mangrove dominance alter vulnerability of a subtropical wetland to sea‐level <span class="hlt">rise</span>?</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Mckee, Karen L.; Vervaeke, William</p> <p>2018-01-01</p> <p>To avoid submergence during sea-level <span class="hlt">rise</span>, coastal wetlands build soil surfaces <span class="hlt">vertically</span> through accumulation of inorganic sediment and organic matter. At climatic boundaries where mangroves are expanding and replacing salt marsh, wetland capacity to respond to sea-level <span class="hlt">rise</span> may change. To compare how well mangroves and salt marshes accommodate sea-level <span class="hlt">rise</span>, we conducted a manipulative field experiment in a subtropical plant community in the subsiding Mississippi River Delta. Experimental plots were established in spatially equivalent positions along creek banks in monospecific stands of Spartina alterniflora (smooth cordgrass) or Avicennia germinans (black mangrove) and in mixed stands containing both species. To examine the effect of disturbance on elevation dynamics, vegetation in half of the plots was subjected to freezing (mangrove) or wrack burial (salt marsh), which caused shoot mortality. <span class="hlt">Vertical</span> soil development was monitored for 6 years with the surface elevation table-marker horizon system. Comparison of land movement with relative sea-level <span class="hlt">rise</span> showed that this plant community was experiencing an elevation deficit (i.e., sea level was <span class="hlt">rising</span> faster than the wetland was building <span class="hlt">vertically</span>) and was relying on elevation capital (i.e., relative position in the tidal frame) to survive. Although Avicennia plots had more elevation capital, suggesting longer survival, than Spartina or mixed plots, vegetation type had no effect on rates of accretion, <span class="hlt">vertical</span> movement in root and sub-root zones, or net elevation change. Thus, these salt marsh and mangrove assemblages were accreting sediment and building <span class="hlt">vertically</span> at equivalent rates. Small-scale disturbance of the plant canopy also had no effect on elevation trajectories—contrary to work in peat-forming wetlands showing elevation responses to changes in plant productivity. The findings indicate that in this deltaic setting with strong physical influences controlling elevation (sediment</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018EJPh...39b5002C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018EJPh...39b5002C"><span>Why does a spinning egg <span class="hlt">rise</span>?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cross, Rod</p> <p>2018-03-01</p> <p>Experimental and theoretical results are presented concerning the <span class="hlt">rise</span> of a spinning egg. It was found that an egg <span class="hlt">rises</span> quickly while it is sliding and then more slowly when it starts rolling. The angular momentum of the egg projected in the XZ plane changed in the same direction as the friction torque, as expected, by rotating away from the <span class="hlt">vertical</span> Z axis. The latter result does not explain the <span class="hlt">rise</span>. However, an even larger effect arises from the Y component of the angular momentum vector. As the egg <span class="hlt">rises</span>, the egg rotates about the Y axis, an effect that is closely analogous to rotation of the egg about the Z axis. Both effects can be described in terms of precession about the respective axes. Steady precession about the Z axis arises from the normal reaction force in the Z direction, while precession about the Y axis arises from the friction force in the Y direction. Precession about the Z axis ceases if the normal reaction force decreases to zero, and precession about the Y axis ceases if the friction force decreases to zero.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ACP....16.9201W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ACP....16.9201W"><span>The importance of plume <span class="hlt">rise</span> on the concentrations and atmospheric impacts of biomass burning aerosol</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Walter, Carolin; Freitas, Saulo R.; Kottmeier, Christoph; Kraut, Isabel; Rieger, Daniel; Vogel, Heike; Vogel, Bernhard</p> <p>2016-07-01</p> <p>We quantified the effects of the plume <span class="hlt">rise</span> of biomass burning aerosol and gases for the forest fires that occurred in Saskatchewan, Canada, in July 2010. For this purpose, simulations with different assumptions regarding the plume <span class="hlt">rise</span> and the <span class="hlt">vertical</span> distribution of the emissions were conducted. Based on comparisons with observations, applying a one-dimensional plume <span class="hlt">rise</span> model to predict the injection layer in combination with a parametrization of the <span class="hlt">vertical</span> distribution of the emissions outperforms approaches in which the plume heights are initially predefined. Approximately 30 % of the fires exceed the height of 2 km with a maximum height of 8.6 km. Using this plume <span class="hlt">rise</span> model, comparisons with satellite images in the visible spectral range show a very good agreement between the simulated and observed spatial distributions of the biomass burning plume. The simulated aerosol optical depth (AOD) with data of an AERONET station is in good agreement with respect to the absolute values and the timing of the maximum. Comparison of the <span class="hlt">vertical</span> distribution of the biomass burning aerosol with CALIPSO (Cloud-Aerosol Lidar and Infrared Pathfinder Satellite Observation) retrievals also showed the best agreement when the plume <span class="hlt">rise</span> model was applied. We found that downwelling surface short-wave radiation below the forest fire plume is reduced by up to 50 % and that the 2 m temperature is decreased by up to 6 K. In addition, we simulated a strong change in atmospheric stability within the biomass burning plume.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013PhFl...25j3302B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013PhFl...25j3302B"><span>On the <span class="hlt">rising</span> motion of a drop in stratified fluids</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bayareh, M.; Doostmohammadi, A.; Dabiri, S.; Ardekani, A. M.</p> <p>2013-10-01</p> <p>The <span class="hlt">rising</span> dynamics of a deformable drop in a linearly stratified fluid is numerically obtained using a finite-volume/front-tracking method. Our results show that the drag coefficient of a spherical drop in a stratified fluid enhances as C_{d,s}/C_{d,h}-<span class="hlt">1</span>˜ Fr_d^{-2.86} for drop Froude numbers in the range of 4 < Frd < 16. The role of the deformability of the drop on the temporal evolution of the motion is investigated along with stratification and inertial effects. We also present the important role of stratification on the transient <span class="hlt">rising</span> motion of the drop. It is shown that a drop can levitate in the presence of a <span class="hlt">vertical</span> density gradient. The drop undergoes a fading oscillatory motion around its neutrally buoyant position except for high viscosity ratio drops where the oscillation occurs around a density level lighter than the neutral buoyancy level. In addition, a detailed characterization of the flow signature of a <span class="hlt">rising</span> drop in a linearly stratified fluid including the buoyancy induced vortices and the resultant buoyant jet is presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010cosp...38..283A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010cosp...38..283A"><span>Dtection of Sea Level <span class="hlt">Rise</span> within the Arabian Gulf Using Space Based GNSS Measurements and Insitu Tide Gauge data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Alothman, Abdulaziz; Ayhan, Mehmet</p> <p></p> <p>In the 21st century, sea level <span class="hlt">rise</span> is expected to be about 30 cm or even more (up to 60 cm). Saudi Arabia has very long coasts of about 3400 km and hundreds of islands. Therefore, sea level monitoring may be important in particular along coastal low lands on Red Sea and Arabian Gulf coasts. Arabian Gulf is connected to Indian Ocean and lying along a parallel course in the south-west of the Zagros Trust Belt. We expect <span class="hlt">vertical</span> land motion within the area due to both tectonic structures of the Arabian Peninsula and oil production activities. Global Navigation Satellite System (GNSS) Continues observations were used to estimate the <span class="hlt">vertical</span> crustal motion. Bahrain International GPS Service (IGS-GPS) station is the only continuous GPS station accessible in the region, and it is close to the Mina Sulman tide gauge station in Bahrain. The weekly GPS time series of <span class="hlt">vertical</span> component at Bahrain IGS-GPS station referring to the ITRF97 from 1999.2 to 2008.6 are used in the computation. We fitted a linear trend with an annual signal and a break to the GPS <span class="hlt">vertical</span> time series and found a <span class="hlt">vertical</span> land motion rate of 0.46 0.11 mm/yr. To investigate sea level variation within the west of Arabian Gulf, monthly means of sea level at 13 tide gauges along the coast of Saudi Arabia and Bahrain, available in the database of the Permanent Service for Mean Sea Level (PSMSL), are studied. We analyzed separately the monthly mean sea level measurements at each station, and estimated secular sea level rate by a robust linear trend fitting. We computed the average relative sea level <span class="hlt">rise</span> rate of <span class="hlt">1</span>.96 0.21 mm/yr within the west of Arabian Gulf based on 4 stations spanning longer than 19 years. Sea level rates at the stations are first corrected for <span class="hlt">vertical</span> land motion contamination using the ICE-5G v<span class="hlt">1</span>.2 VM4 Glacial Isostatic Adjustment (GIA) model, and the average sea level rate is found 2.27 0.21 mm/yr. Assuming the <span class="hlt">vertical</span> rate at Bahrain IGS-GPS station represents the <span class="hlt">vertical</span> rate</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29738545','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29738545"><span>Keeping up with sea-level <span class="hlt">rise</span>: Carbonate production rates in Palau and Yap, western Pacific Ocean.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>van Woesik, Robert; Cacciapaglia, Christopher William</p> <p>2018-01-01</p> <p>Coral reefs protect islands from tropical storm waves and provide goods and services for millions of islanders worldwide. Yet it is unknown how coral reefs in general, and carbonate production in particular, will respond to sea-level <span class="hlt">rise</span> and thermal stress associated with climate change. This study compared the reef-building capacity of different shallow-water habitats at twenty-four sites on each of two islands, Palau and Yap, in the western Pacific Ocean. We were particularly interested in estimating the inverse problem of calculating the value of live coral cover at which net carbonate production becomes negative, and whether that value varied across habitats. Net carbonate production varied among habitats, averaging 10.2 kg CaCO3 m-2 y-<span class="hlt">1</span> for outer reefs, 12.7 kg CaCO3 m-2 y-<span class="hlt">1</span> for patch reefs, and 7.2 kg CaCO3 m-2 y-<span class="hlt">1</span> for inner reefs. The value of live coral cover at which net carbonate production became negative varied across habitats, with highest values on inner reefs. These results suggest that some inner reefs tend to produce less carbonate, and therefore need higher coral cover to produce enough carbonate to keep up with sea-level <span class="hlt">rise</span> than outer and patch reefs. These results also suggest that inner reefs are more vulnerable to sea-level <span class="hlt">rise</span> than other habitats, which stresses the need for effective land-use practices as the climate continues to change. Averaging across all reef habitats, the rate of carbonate production was 9.7 kg CaCO3 m-2 y-<span class="hlt">1</span>, or approximately 7.9 mm y-<span class="hlt">1</span> of potential <span class="hlt">vertical</span> accretion. Such rates of <span class="hlt">vertical</span> accretion are higher than projected averages of sea-level <span class="hlt">rise</span> for the representative concentration pathway (RCP) climate-change scenarios 2.6, 4.5, and 6, but lower than for the RCP scenario 8.5.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.G23B0477S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.G23B0477S"><span>A Unified Geodetic <span class="hlt">Vertical</span> Velocity Field (UGVVF), Version <span class="hlt">1</span>.0</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schmalzle, G.; Wdowinski, S.</p> <p>2014-12-01</p> <p>Tectonic motion, volcanic inflation or deflation, as well as oil, gas and water pumping can induce <span class="hlt">vertical</span> motion. In southern California these signals are inter-mingled. In tectonics, properly identifying regions that are contaminated by other signals can be important when estimating fault slip rates. Until recently <span class="hlt">vertical</span> deformation rates determined by high precision Global Positioning Systems (GPS) had large uncertainties compared to horizontal components and were rarely used to constrain tectonic models of fault motion. However, many continuously occupied GPS stations have been operating for ten or more years, often delivering uncertainties of ~<span class="hlt">1</span> mm/yr or less, providing better constraints for tectonic modeling. Various processing centers produced GPS time series and estimated <span class="hlt">vertical</span> velocity fields, each with their own set of processing techniques and assumptions. We compare <span class="hlt">vertical</span> velocity solutions estimated by seven data processing groups as well as two combined solutions (Figure <span class="hlt">1</span>). These groups include: Central Washington University (CWU) and New Mexico Institute of Technology (NMT), and their combined solution provided by the Plate Boundary Observatory (PBO) through the UNAVCO website. Also compared are the Jet Propulsion Laboratory (JPL) and Scripps Orbit and Permanent Array Center (SOPAC) and their combined solution provided as part of the NASA MEaSUREs project. Smaller velocity fields included are from Amos et al., 2014, processed at the Nevada Geodetic Laboratory, Shen et al., 2011, processed by UCLA and called the Crustal Motion Map 4.0 (CMM4) dataset, and a new velocity field provided by the University of Miami (UM). Our analysis includes estimating and correcting for systematic <span class="hlt">vertical</span> velocity and uncertainty differences between groups. Our final product is a unified velocity field that contains the median values of the adjusted velocity fields and their uncertainties. This product will be periodically updated when new velocity fields</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1812573P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1812573P"><span>Contribution of <span class="hlt">vertical</span> land motions to coastal sea level variations: a global synthesis of multisatellite altimetry, tide gauge and GPS measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pfeffer, Julia; Allemand, Pascal</p> <p>2016-04-01</p> <p>Coastal sea level variations result from a complex mix of climatic, oceanic and geodynamical processes driven by natural and anthropogenic constraints. Combining data from multiple sources is one solution to identify particular processes and progress towards a better understanding of the sea level variations and the assessment of their impacts at coast. Here, we present a global database merging multisatellite altimetry with tide gauges and Global Positioning System (GPS) measurements. <span class="hlt">Vertical</span> land motions and sea level variations are estimated simultaneously for a network of 886 ground stations with median errors lower than <span class="hlt">1</span> mm/yr. The contribution of <span class="hlt">vertical</span> land motions to relative sea level variations is explored to better understand the natural hazards associated with sea level <span class="hlt">rise</span> in coastal areas. Worldwide, <span class="hlt">vertical</span> land motions dominate 30 % of observed coastal trends. The role of the crust is highly heterogeneous: it can amplify, restrict or counter the effects of climate-induced sea level change. A set of 182 potential vulnerable localities are identified by large coastal subsidence which increases by several times the effects of sea level <span class="hlt">rise</span>. Though regional behaviours exist, principally caused by GIA (Glacial Isostatic Adjustment), the local variability in <span class="hlt">vertical</span> land motion prevails. An accurate determination of the <span class="hlt">vertical</span> motions observed at the coast is fundamental to understand the local processes which contribute to sea level <span class="hlt">rise</span>, to appraise its impacts on coastal populations and make future predictions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.G53A..04H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.G53A..04H"><span>GPS Imaging of Global <span class="hlt">Vertical</span> Land Motion for Sea Level Studies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hammond, W. C.; Blewitt, G.; Hamlington, B. D.</p> <p>2015-12-01</p> <p>Coastal <span class="hlt">vertical</span> land motion contributes to the signal of local relative sea level change. Moreover, understanding global sea level change requires understanding local sea level <span class="hlt">rise</span> at many locations around Earth. It is therefore essential to understand the regional secular <span class="hlt">vertical</span> land motion attributable to mantle flow, tectonic deformation, glacial isostatic adjustment, postseismic viscoelastic relaxation, groundwater basin subsidence, elastic rebound from groundwater unloading or other processes that can change the geocentric height of tide gauges anchored to the land. These changes can affect inferences of global sea level <span class="hlt">rise</span> and should be taken into account for global projections. We present new results of GPS imaging of <span class="hlt">vertical</span> land motion across most of Earth's continents including its ice-free coastlines around North and South America, Europe, Australia, Japan, parts of Africa and Indonesia. These images are based on data from many independent open access globally distributed continuously recording GPS networks including over 13,500 stations. The data are processed in our system to obtain solutions aligned to the International Terrestrial Reference Frame (ITRF08). To generate images of <span class="hlt">vertical</span> rate we apply the Median Interannual Difference Adjusted for Skewness (MIDAS) algorithm to the <span class="hlt">vertical</span> times series to obtain robust non-parametric estimates with realistic uncertainties. We estimate the <span class="hlt">vertical</span> land motion at the location of 1420 tide gauges locations using Delaunay-based geographic interpolation with an empirically derived distance weighting function and median spatial filtering. The resulting image is insensitive to outliers and steps in the GPS time series, omits short wavelength features attributable to unstable stations or unrepresentative rates, and emphasizes long-wavelength mantle-driven <span class="hlt">vertical</span> rates.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_1");'>1</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li class="active"><span>3</span></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_3 --> <div id="page_4" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="61"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.T33C2945F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.T33C2945F"><span>Record Of Both Tectonic Related <span class="hlt">Vertical</span> Motions and Global Sea Level <span class="hlt">Rise</span> by Marine Terraces along an Active Arc Volcano. Example of Basse-Terre, Lesser Antilles (French West-Indies).</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fabre, M.; Moysan, M.; Graindorge, D.; Jean-Frederic, L.; Philippon, M. M.; Marcaillou, B.; Léticée, J. L.</p> <p>2015-12-01</p> <p>Volcano-tectonic history of the Caribbean plate provides direct insight onto the dynamic of the North American Plate westward subduction. Basse-Terre Island is a volcanic chain that belongs to the Lesser Antilles active volcanic arc with a southward decreasing age of volcanism from 3 Ma to present day.We investigate records of <span class="hlt">vertical</span> motion along Basse-Terre through a morphostructural analysis of the Pleistocene-Holocene shallow-water carbonate platforms and associated terraces that surround Basse-Terre Island. This study is based on new high-resolution bathymetric and dense seismic data acquired during the GEOTREF oceanographic survey (2015, February). Our bathymetric and topographic Digital Terrain Model together with the "Litto3D" Lidar data (IGN/SHOM) images the island topography and the platform bathymetry to a depth of 200m with horizontal and <span class="hlt">vertical</span> resolutions of 5m and ~cm respectively. This detailed study highlights the morphostructure of terraces built during the last transgression in order to identify and quantify their <span class="hlt">vertical</span> motions. We analyze inherited morphology and structures of the forearc that affect the platform to discuss effects of the regional tectonics context. A particular emphasis is put on the influence of the NW-SE arc parallel transtensive Montserrat-Bouillante fault system onto the platform geometry. At last, the distribution of Basse-Terre terraces is compared with terraces distribution around other Lesser Antilles island and the Bahamas stable margin platform. We aim at discriminating the influence of the Pleistocene global sea-level <span class="hlt">rise</span> from the one of tectonic <span class="hlt">vertical</span> deformations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA048752','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA048752"><span>Morphological Studies of <span class="hlt">Rising</span> Equatorial Spread F Bubbles</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1977-11-01</p> <p>depletions. In the present paper , we wish to discuss equatorial Spread F bubble shapes and <span class="hlt">vertical</span> <span class="hlt">rise</span> rates within the context of the collisional...simulation results are needed to ascertain which model fits best. All of the models described in this paper , based on collisional Rayleigh-Taylor type...Analysis of Barium Clouds - Semi-Annual Technical Report, RADC-TR-72-103, Vol. I, Avco Everett Reserach Laboratory, Everett, Mass., January 1972</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012EGUGA..14.9121S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012EGUGA..14.9121S"><span>A High School Project Seminar on Sea Level <span class="hlt">Rise</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Seitz, M.; Bosch, W.</p> <p>2012-04-01</p> <p>In Bavaria the curriculum of the upper grade of high school includes a so called project seminar, running over one and a half year. The aims of the seminar are to let the pupils learn to work on a specific topic, to organize themselves in a team, to improve their soft skills and become familiar with the working life. The topic of the project seminar, jointly organized by the Bertold-Brecht-Gymnasium in Munich and the Deutsche Geodätische Forschungsinstitut (DGFI) was on the "Global sea level <span class="hlt">rise</span>". A team of 13 pupils computed the mean sea level <span class="hlt">rise</span> by using on the one hand altimetry data of TOPEX, Jason-<span class="hlt">1</span> and Jason2 and on the other hand data of globally distributed tide gauges, corrected for <span class="hlt">vertical</span> crustal movements derived from GPS products. The results of the two independent approaches were compared with each other and discussed considering also statements and discussions found in press, TV, and the web. Finally, a presentation was prepared and presented at school.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014NatCC...4..493R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014NatCC...4..493R"><span>Oyster reefs can outpace sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rodriguez, Antonio B.; Fodrie, F. Joel; Ridge, Justin T.; Lindquist, Niels L.; Theuerkauf, Ethan J.; Coleman, Sara E.; Grabowski, Jonathan H.; Brodeur, Michelle C.; Gittman, Rachel K.; Keller, Danielle A.; Kenworthy, Matthew D.</p> <p>2014-06-01</p> <p>In the high-salinity seaward portions of estuaries, oysters seek refuge from predation, competition and disease in intertidal areas, but this sanctuary will be lost if <span class="hlt">vertical</span> reef accretion cannot keep pace with sea-level <span class="hlt">rise</span> (SLR). Oyster-reef abundance has already declined ~85% globally over the past 100 years, mainly from over harvesting, making any additional losses due to SLR cause for concern. Before any assessment of reef response to accelerated SLR can be made, direct measures of reef growth are necessary. Here, we present direct measurements of intertidal oyster-reef growth from cores and terrestrial lidar-derived digital elevation models. On the basis of our measurements collected within a mid-Atlantic estuary over a 15-year period, we developed a globally testable empirical model of intertidal oyster-reef accretion. We show that previous estimates of <span class="hlt">vertical</span> reef growth, based on radiocarbon dates and bathymetric maps, may be greater than one order of magnitude too slow. The intertidal reefs we studied should be able to keep up with any future accelerated rate of SLR (ref. ) and may even benefit from the additional subaqueous space allowing extended <span class="hlt">vertical</span> accretion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26442712','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26442712"><span>Maximizing oyster-reef growth supports green infrastructure with accelerating sea-level <span class="hlt">rise</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ridge, Justin T; Rodriguez, Antonio B; Joel Fodrie, F; Lindquist, Niels L; Brodeur, Michelle C; Coleman, Sara E; Grabowski, Jonathan H; Theuerkauf, Ethan J</p> <p>2015-10-07</p> <p>Within intertidal communities, aerial exposure (emergence during the tidal cycle) generates strong <span class="hlt">vertical</span> zonation patterns with distinct growth boundaries regulated by physiological and external stressors. Forecasted accelerations in sea-level <span class="hlt">rise</span> (SLR) will shift the position of these critical boundaries in ways we cannot yet fully predict, but landward migration will be impaired by coastal development, amplifying the importance of foundation species' ability to maintain their position relative to <span class="hlt">rising</span> sea levels via <span class="hlt">vertical</span> growth. Here we show the effects of emergence on <span class="hlt">vertical</span> oyster-reef growth by determining the conditions at which intertidal reefs thrive and the sharp boundaries where reefs fail, which shift with changes in sea level. We found that oyster reef growth is unimodal relative to emergence, with greatest growth rates occurring between 20-40% exposure, and zero-growth boundaries at 10% and 55% exposures. Notably, along the lower growth boundary (10%), increased rates of SLR would outpace reef accretion, thereby reducing the depth range of substrate suitable for reef maintenance and formation, and exacerbating habitat loss along developed shorelines. Our results identify where, within intertidal areas, constructed or natural oyster reefs will persist and function best as green infrastructure to enhance coastal resiliency under conditions of accelerating SLR.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4595829','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4595829"><span>Maximizing oyster-reef growth supports green infrastructure with accelerating sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ridge, Justin T.; Rodriguez, Antonio B.; Joel Fodrie, F.; Lindquist, Niels L.; Brodeur, Michelle C.; Coleman, Sara E.; Grabowski, Jonathan H.; Theuerkauf, Ethan J.</p> <p>2015-01-01</p> <p>Within intertidal communities, aerial exposure (emergence during the tidal cycle) generates strong <span class="hlt">vertical</span> zonation patterns with distinct growth boundaries regulated by physiological and external stressors. Forecasted accelerations in sea-level <span class="hlt">rise</span> (SLR) will shift the position of these critical boundaries in ways we cannot yet fully predict, but landward migration will be impaired by coastal development, amplifying the importance of foundation species’ ability to maintain their position relative to <span class="hlt">rising</span> sea levels via <span class="hlt">vertical</span> growth. Here we show the effects of emergence on <span class="hlt">vertical</span> oyster-reef growth by determining the conditions at which intertidal reefs thrive and the sharp boundaries where reefs fail, which shift with changes in sea level. We found that oyster reef growth is unimodal relative to emergence, with greatest growth rates occurring between 20–40% exposure, and zero-growth boundaries at 10% and 55% exposures. Notably, along the lower growth boundary (10%), increased rates of SLR would outpace reef accretion, thereby reducing the depth range of substrate suitable for reef maintenance and formation, and exacerbating habitat loss along developed shorelines. Our results identify where, within intertidal areas, constructed or natural oyster reefs will persist and function best as green infrastructure to enhance coastal resiliency under conditions of accelerating SLR. PMID:26442712</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.C53B0574L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.C53B0574L"><span>Ice Shelf-Ocean Interactions Near Ice <span class="hlt">Rises</span> and Ice Rumples</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lange, M. A.; Rückamp, M.; Kleiner, T.</p> <p>2013-12-01</p> <p>, focusing on the floating ice parts of the Brunt and Riiser-Larsen ice shelves. The major response of the ice is observed instantaneously and is caused by the time independent nature of the Stokes equations and the used Glen-type rheology. The influence of ice temperatures and therefore the time-dependent effect on the flow-rate are small, given a 100 year time frame and applying a fixed-geometry setting.. A particularly important result of the current project lies in the fact that we have numerically simulated the three-dimensional stress fields in an ice shelf. Common numerical models that utilize a <span class="hlt">vertically</span> integrated Shallow Shelf Approximation (SSA-models), do not provide that information. Due to the detailed horizontal resolution of <span class="hlt">1</span>km in our models, we were able to also model the observed heavily fractured areas in the vicinity of McDonald Ice <span class="hlt">Rise</span>, a region that is characterized by simulated tensile stresses reaching maximum <span class="hlt">vertical</span> extension in the ice column.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/pa1551.photos.141531p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/pa1551.photos.141531p/"><span><span class="hlt">1</span>. GENERAL VIEW. OVERHANG, PAINTED RED, HAS <span class="hlt">VERTICAL</span> SIDING AND ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p><span class="hlt">1</span>. GENERAL VIEW. OVERHANG, PAINTED RED, HAS <span class="hlt">VERTICAL</span> SIDING AND FADED PAINTINGS OF FARM ANIMALS: COW, DONKEYS AND HORSE. - De Turck House, Barn, State Route 662 vicinity, Oley Township, Oley, Berks County, PA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017E%26PSL.473...24S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017E%26PSL.473...24S"><span>Uncertainty of the 20th century sea-level <span class="hlt">rise</span> due to <span class="hlt">vertical</span> land motion errors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Santamaría-Gómez, Alvaro; Gravelle, Médéric; Dangendorf, Sönke; Marcos, Marta; Spada, Giorgio; Wöppelmann, Guy</p> <p>2017-09-01</p> <p>Assessing the <span class="hlt">vertical</span> land motion (VLM) at tide gauges (TG) is crucial to understanding global and regional mean sea-level changes (SLC) over the last century. However, estimating VLM with accuracy better than a few tenths of a millimeter per year is not a trivial undertaking and many factors, including the reference frame uncertainty, must be considered. Using a novel reconstruction approach and updated geodetic VLM corrections, we found the terrestrial reference frame and the estimated VLM uncertainty may contribute to the global SLC rate error by ± 0.2 mmyr-<span class="hlt">1</span>. In addition, a spurious global SLC acceleration may be introduced up to ± 4.8 ×10-3 mmyr-2. Regional SLC rate and acceleration errors may be inflated by a factor 3 compared to the global. The difference of VLM from two independent Glacio-Isostatic Adjustment models introduces global SLC rate and acceleration biases at the level of ± 0.<span class="hlt">1</span> mmyr-<span class="hlt">1</span> and 2.8 ×10-3 mmyr-2, increasing up to 0.5 mm yr-<span class="hlt">1</span> and 9 ×10-3 mmyr-2 for the regional SLC. Errors in VLM corrections need to be budgeted when considering past and future SLC scenarios.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.476.5629K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.476.5629K"><span>Magnetic activity in the Galactic Centre region - fast downflows along <span class="hlt">rising</span> magnetic loops</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kakiuchi, Kensuke; Suzuki, Takeru K.; Fukui, Yasuo; Torii, Kazufumi; Enokiya, Rei; Machida, Mami; Matsumoto, Ryoji</p> <p>2018-06-01</p> <p>We studied roles of the magnetic field on the gas dynamics in the Galactic bulge by a three-dimensional global magnetohydrodynamical simulation data, particularly focusing on <span class="hlt">vertical</span> flows that are ubiquitously excited by magnetic activity. In local regions where the magnetic field is stronger, it is frequently seen that fast downflows slide along inclined magnetic field lines that are associated with buoyantly <span class="hlt">rising</span> magnetic loops. The <span class="hlt">vertical</span> velocity of these downflows reaches ˜100 km s-<span class="hlt">1</span> near the footpoint of the loops by the gravitational acceleration towards the Galactic plane. The two footpoints of <span class="hlt">rising</span> magnetic loops are generally located at different radial locations and the field lines are deformed by the differential rotation. The angular momentum is transported along the field lines, and the radial force balance breaks down. As a result, a fast downflow is often observed only at the one footpoint located at the inner radial position. The fast downflow compresses the gas to form a dense region near the footpoint, which will be important in star formation afterwards. Furthermore, the horizontal components of the velocity are also fast near the footpoint because the downflow is accelerated along the magnetic sliding slope. As a result, the high-velocity flow creates various characteristic features in a simulated position-velocity diagram, depending on the viewing angle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMGC23F1296S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMGC23F1296S"><span>Measuring Sea Level <span class="hlt">Rise</span>-Induced Shoreline Changes and Inundation in Real Time</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shilling, F.; Waetjen, D.; Grijalva, E.</p> <p>2016-12-01</p> <p>We describe a method to monitor shoreline inundation and changes in response to sea level <span class="hlt">rise</span> (SLR) using a network of time-lapse cameras. We found for coastal tidal marshes that this method was sensitive to <span class="hlt">vertical</span> changes in sea level of <<span class="hlt">1</span> cm, roughly equivalent to <span class="hlt">1</span>-2 years of sea level <span class="hlt">rise</span> under the A<span class="hlt">1</span> scenario. SLR of >20 cm has occurred in the San Francisco Bay and other US coastal areas and is likely to <span class="hlt">rise</span> by another 30-45 cm by mid-century, which will flood and erode many coastal ecosystems, highways, and urban areas. This rapid degree of <span class="hlt">rise</span> means that it is imperative to co-plan for natural and built systems. Many public facilities are adjacent to shoreline ecosystems, which both protect infrastructure from wave and tide energy and are home to regulated species and habitats. Accurate and timely information about the actual extent of SLR impacts to shorelines will be critical during built-system adaptation. Currently, satellite-sourced imagery cannot provide the spatial or temporal resolution necessary to investigate fine-scale shoreline changes, leaving a gap between predictive models and knowing how, where and when these changes are occurring. The method described is feasible for near-term (<span class="hlt">1</span> to 10 years) to long-term application and can be used for measuring fine-resolution shoreline changes (<<span class="hlt">1</span> m2) in response to SLR and associated wave action inundation of marshes and infrastructure. We demonstrate the method with networks of cameras in 2 coastal states (CA and GA), using web-informatics and services to organize photographs that could be combined with related external data (e.g., gauged water levels) to create an information mashup. This information could be used to validate models predicting shoreline inundation and loss, inform SLR-adaptation planning, and to visualize SLR impacts to the public.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.6334Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.6334Z"><span>Modelling the long-term <span class="hlt">vertical</span> dynamics of salt marshes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zoccarato, Claudia; Teatini, Pietro</p> <p>2017-04-01</p> <p>Salt marshes are vulnerable environments hosting complex interactions between physical and biological processes with a strong influence on the dynamics of the marsh evolution. The estimation and prediction of the elevation of a salt-marsh platform is crucial to forecast the marsh growth or regression under different scenarios considering, for example, the potential climate changes. The long-term <span class="hlt">vertical</span> dynamics of a salt marsh is predicted with the aid of an original finite-element (FE) numerical model accounting for the marsh accretion and compaction and for the variation rates of the relative sea level <span class="hlt">rise</span>, i.e., land subsidence of the marsh basement and eustatic <span class="hlt">rise</span> of the sea level. The accretion term considers the <span class="hlt">vertical</span> sedimentation of organic and inorganic material over the marsh surface, whereas the compaction reflects the progressive consolidation of the porous medium under the increasing load of the overlying younger deposits. The modelling approach is based on a 2D groundwater flow simulator, which provides the pressure evolution within a compacting/accreting <span class="hlt">vertical</span> cross-section of the marsh assuming that the groundwater flow obeys the relative Darcy's law, coupled to a <span class="hlt">1</span>D <span class="hlt">vertical</span> geomechanical module following Terzaghi's principle of effective intergranular stress. Soil porosity, permeability, and compressibility may vary with the effective intergranular stress according to empirically based relationships. The model also takes into account the geometric non-linearity arising from the consideration of large solid grain movements by using a Lagrangian approach with an adaptive FE mesh. The element geometry changes in time to follow the deposit consolidation and the element number increases in time to follow the sedimentation of new material. The numerical model is tested on different realistic configurations considering the influence of (i) the spatial distribution of the sedimentation rate in relation to the distance from the marsh margin, (ii</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ut0700.photos.364253p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ut0700.photos.364253p/"><span>Trestle #<span class="hlt">1</span>, detail of bolts on northeast abutment lower <span class="hlt">vertical</span> ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>Trestle #<span class="hlt">1</span>, detail of bolts on northeast abutment lower <span class="hlt">vertical</span> support timbers. View to north - Promontory Route Railroad Trestles, S.P. Trestle 779.91, One mile southwest of junction of State Highway 83 and Blue Creek, Corinne, Box Elder County, UT</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1377826','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1377826"><span>Tidal Marshes across a Chesapeake Bay Subestuary Are Not Keeping up with Sea-Level <span class="hlt">Rise</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Beckett, Leah H.; Baldwin, Andrew H.; Kearney, Michael S.</p> <p></p> <p>Sea-level <span class="hlt">rise</span> is a major factor in wetland loss worldwide, and inmuch of Chesapeake Bay (USA) the rate of sea-level <span class="hlt">rise</span> is higher than the current global rate of 3.2 mmyr -<span class="hlt">1</span> due to regional subsidence.Marshes along estuarine salinity gradients differ in vegetation composition, productivity, decomposition pathways, and sediment dynamics, andmay exhibit different responses to sea-level <span class="hlt">rise</span>. Coastal marshes persist by building <span class="hlt">vertically</span> at rates at or exceeding regional sea-level <span class="hlt">rise</span>. In one of the first studies to examine elevation dynamics across an estuarine salinity gradient, we installed 15 surface elevation tables (SET) and accretion marker-horizon plots (MH) in tidalmore » freshwater, oligohaline, and brackish marshes across a Chesapeake Bay subestuary. Over the course of four years, wetlands across the subestuary decreased <span class="hlt">1</span>.8 ± 2.7 mmyr -<span class="hlt">1</span> in elevation on average, at least 5 mmyr -<span class="hlt">1</span> below that needed to keep pace with global sea-level <span class="hlt">rise</span>. Elevation change rates did not significantly differ among themarshes studied, and ranged from-9.8 ± 6.9 to 4.5 ± 4.3 mmyr -<span class="hlt">1</span>. Surface accretion of depositedmineral and organic matter was uniformly high across the estuary (~9–15 mmyr -<span class="hlt">1</span>), indicating that elevation loss was not due to lack of accretionary input. Position in the estuary and associated salinity regime were not related to elevation change or surface matter accretion. In conclusion, previous studies have focused on surface elevation change inmarshes of uniformsalinity (e.g., salt marshes); however, our findings highlight the need for elevation studies inmarshes of all salinity regimes and different geomorphic positions, and warn that brackish, oligohaline, and freshwater tidal wetlands may be at similarly high risk of submergence in some estuaries.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4965100','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4965100"><span>Tidal Marshes across a Chesapeake Bay Subestuary Are Not Keeping up with Sea-Level <span class="hlt">Rise</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Beckett, Leah H.; Baldwin, Andrew H.; Kearney, Michael S.</p> <p>2016-01-01</p> <p>Sea-level <span class="hlt">rise</span> is a major factor in wetland loss worldwide, and in much of Chesapeake Bay (USA) the rate of sea-level <span class="hlt">rise</span> is higher than the current global rate of 3.2 mm yr-<span class="hlt">1</span> due to regional subsidence. Marshes along estuarine salinity gradients differ in vegetation composition, productivity, decomposition pathways, and sediment dynamics, and may exhibit different responses to sea-level <span class="hlt">rise</span>. Coastal marshes persist by building <span class="hlt">vertically</span> at rates at or exceeding regional sea-level <span class="hlt">rise</span>. In one of the first studies to examine elevation dynamics across an estuarine salinity gradient, we installed 15 surface elevation tables (SET) and accretion marker-horizon plots (MH) in tidal freshwater, oligohaline, and brackish marshes across a Chesapeake Bay subestuary. Over the course of four years, wetlands across the subestuary decreased <span class="hlt">1</span>.8 ± 2.7 mm yr-<span class="hlt">1</span> in elevation on average, at least 5 mm yr-<span class="hlt">1</span> below that needed to keep pace with global sea-level <span class="hlt">rise</span>. Elevation change rates did not significantly differ among the marshes studied, and ranged from -9.8 ± 6.9 to 4.5 ± 4.3 mm yr-<span class="hlt">1</span>. Surface accretion of deposited mineral and organic matter was uniformly high across the estuary (~9–15 mm yr-<span class="hlt">1</span>), indicating that elevation loss was not due to lack of accretionary input. Position in the estuary and associated salinity regime were not related to elevation change or surface matter accretion. Previous studies have focused on surface elevation change in marshes of uniform salinity (e.g., salt marshes); however, our findings highlight the need for elevation studies in marshes of all salinity regimes and different geomorphic positions, and warn that brackish, oligohaline, and freshwater tidal wetlands may be at similarly high risk of submergence in some estuaries. PMID:27467784</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1377826-tidal-marshes-across-chesapeake-bay-subestuary-keeping-up-sea-level-rise','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1377826-tidal-marshes-across-chesapeake-bay-subestuary-keeping-up-sea-level-rise"><span>Tidal Marshes across a Chesapeake Bay Subestuary Are Not Keeping up with Sea-Level <span class="hlt">Rise</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Beckett, Leah H.; Baldwin, Andrew H.; Kearney, Michael S.; ...</p> <p>2016-07-28</p> <p>Sea-level <span class="hlt">rise</span> is a major factor in wetland loss worldwide, and inmuch of Chesapeake Bay (USA) the rate of sea-level <span class="hlt">rise</span> is higher than the current global rate of 3.2 mmyr -<span class="hlt">1</span> due to regional subsidence.Marshes along estuarine salinity gradients differ in vegetation composition, productivity, decomposition pathways, and sediment dynamics, andmay exhibit different responses to sea-level <span class="hlt">rise</span>. Coastal marshes persist by building <span class="hlt">vertically</span> at rates at or exceeding regional sea-level <span class="hlt">rise</span>. In one of the first studies to examine elevation dynamics across an estuarine salinity gradient, we installed 15 surface elevation tables (SET) and accretion marker-horizon plots (MH) in tidalmore » freshwater, oligohaline, and brackish marshes across a Chesapeake Bay subestuary. Over the course of four years, wetlands across the subestuary decreased <span class="hlt">1</span>.8 ± 2.7 mmyr -<span class="hlt">1</span> in elevation on average, at least 5 mmyr -<span class="hlt">1</span> below that needed to keep pace with global sea-level <span class="hlt">rise</span>. Elevation change rates did not significantly differ among themarshes studied, and ranged from-9.8 ± 6.9 to 4.5 ± 4.3 mmyr -<span class="hlt">1</span>. Surface accretion of depositedmineral and organic matter was uniformly high across the estuary (~9–15 mmyr -<span class="hlt">1</span>), indicating that elevation loss was not due to lack of accretionary input. Position in the estuary and associated salinity regime were not related to elevation change or surface matter accretion. In conclusion, previous studies have focused on surface elevation change inmarshes of uniformsalinity (e.g., salt marshes); however, our findings highlight the need for elevation studies inmarshes of all salinity regimes and different geomorphic positions, and warn that brackish, oligohaline, and freshwater tidal wetlands may be at similarly high risk of submergence in some estuaries.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19893839','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19893839"><span><span class="hlt">Vertical</span> integration - Reducing the load on GP teachers.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Anderson, Katrina; Thomson, Jennifer</p> <p>2009-11-01</p> <p>With the increased medical student numbers in Australia there is an expectation that general practice will train students, junior doctors and registrars, and the teaching burden for busy general practitioners will <span class="hlt">rise</span>. We discuss the model of <span class="hlt">vertical</span> integration of general practice education set up at the Australian National University Medical School in the Australian Capital Territory and southeast New South Wales. This model of <span class="hlt">vertical</span> integration is unique. It could be adapted in a range of vocational settings and spans medical student, prevocational doctor, registrar and international medical graduate teaching. A key aim of these strategies is to reduce the load on the clinical GP teacher as sustaining their contribution is crucial to the future of training in general practice.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMOS23B1406S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMOS23B1406S"><span>Quantifying and Projecting Relative Sea-Level <span class="hlt">Rise</span> in The Deltaic Regions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shum, C. K.; Chung-Yen, K.; Calmant, S.; Yang, T. Y.; Guo, Q.; Jia, Y.; Ballu, V.; Guo, J.; Karptychev, M.; Krien, Y.; Kusche, J.; Tseng, K. H.; Wan, J.; Uebbing, B.</p> <p>2017-12-01</p> <p>Half of the world's population lives within 200 km of coastlines. Accelerated sea-level <span class="hlt">rise</span>, compounded by effects of population growth, severe land subsidence due to fluvial sediment compaction/load, and anthropogenic oil and natural gas and ground water extraction, tectonic motion, and the increasing threat of more intense and more frequent cyclone-driven storm surges, have exacerbated the vulnerability of many of world's deltaic regions, including the Bangladesh and the Mississippi River Deltas. At present, understanding and quantifying the natural and anthropogenic processes governing these solid Earth <span class="hlt">vertical</span> motion processes remain elusive to enable addressing coastal vulnerability due to current and future projection of relative sea-level <span class="hlt">rise</span> for deltaic regions at the regional scales. Bangladesh, a low-lying and one of the most densely populated countries in the world located at the Bay of Bengal, is prone to transboundary monsoonal flooding, and is believed to be aggravated by more frequent and intensified cyclones resulting from anthropogenic climate change. The Mississippi River Deltaic region has been severely subsiding due primarily to fluvial sediment compaction and load during the last 10 centuries, oil/gas and groundwater extractions, and commercial developments, making it vulnerable to sea-level <span class="hlt">rise</span> hazards. Here we present results of global geocentric sea-level <span class="hlt">rise</span>, 1950-2016, separating <span class="hlt">vertical</span> land motion at global tide gauge datum, by integrating tide gauge and radar altimeter records in a novel sea-level reconstruction scheme, focusing on the Mississippi River and the Bangladesh Deltas. We then integrate the resulting sea level estimates with historic imageries, GPS and InSAR data, as well as sediment isostatic and load model predicted present-day land subsidence, to constrain the 3D land motion to study the impacts of various scenarios of future relative sea level projections on the Bangladesh Delta to the end of the 21st Century and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.5474A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.5474A"><span>An Update of Sea Level <span class="hlt">Rise</span> in the northwestern part of the Arabian Gulf</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Alothman, Abdulaziz; Bos, Machiel; Fernandes, Rui</p> <p>2017-04-01</p> <p>Relative sea level variations in the northwestern part of the Arabian Gulf have been estimated in the past using no more than 10 to 15 years of observations. In Alothman et al. (2014), we have almost doubled the period to 28.7 years by examining all available tide gauge data in the area and constructing a mean gauge time-series from seven coastal tide gauges. We found for the period 1979-2007 a relative sea level <span class="hlt">rise</span> of about 2mm/yr, which correspond to an absolute sea level <span class="hlt">rise</span> of about <span class="hlt">1</span>.5mm/yr based on the <span class="hlt">vertical</span> displacement of GNSS stations in the region. By taking into account the temporal correlations we concluded that previous published results underestimate the true sea level rate error in this area by a factor of 5-10. In this work, we discuss and update the methodology and results from Alothman et al. (2014), particularly by checking and extending the GNSS solutions. Since 3 of the 6 GPS stations used only started observing in the end of 2011, the longer time series have now significantly lower uncertainties in the estimated <span class="hlt">vertical</span> rate. In addition, we compare our results with GRACE derived ocean bottom pressure time series which are a good proxy of the changes in water mass in this area over time.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/pa2685.photos.142474p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/pa2685.photos.142474p/"><span>4. VIEW OF <span class="hlt">VERTICAL</span> BORING MACHINE. (Bullard) <span class="hlt">Vertical</span> turning lathe ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>4. VIEW OF <span class="hlt">VERTICAL</span> BORING MACHINE. (Bullard) <span class="hlt">Vertical</span> turning lathe (VTL). Machining the fixture for GE Turboshroud. G.S. O'Brien, operator. - Juniata Shops, Machine Shop No. <span class="hlt">1</span>, East of Fourth Avenue at Third Street, Altoona, Blair County, PA</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_4 --> <div id="page_5" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li class="active"><span>5</span></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="81"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70185454','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70185454"><span>A pressure-packer system for conducting <span class="hlt">rising</span> head tests in water table wells</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Levy, Benjamin S.; Pannell, Lawrence J.; Dadoly, John P.</p> <p>1993-01-01</p> <p>The pressure system developed for fully-saturated well screens has been modified for conducting <span class="hlt">rising</span> head tests in water table wells installed in highly permeable aquifers. The pressure system consists of a compressed air source and <span class="hlt">1</span> inch diameter PVC piping with a packer attached at the end. The pressure system was evaluated in a series of <span class="hlt">rising</span> head tests conducted in a well at a Superfund site in New England. The well was tested with slugs and with the pressure system. Within each technique, estimates of hydraulic conductivity showed no difference. Comparison of hydraulic conductivity estimates between techniques (slug test vs. pressure test) showed differences due to stratigraphy. The interval tested using slug tests crossed two stratigraphic units; the pressure system tested only one of these units. We conclude that the pressure system may be used to characterize the <span class="hlt">vertical</span> hydraulic conductivity distribution in a series of successive tests by changing the packer position and the screened interval tested.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFM.P32B..05H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.P32B..05H"><span>Hi<span class="hlt">RISE</span> Observations of the Polar Regions of Mars</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Herkenhoff, K. E.; Byrne, S.; Fishbaugh, K.; Russell, P.; Fortezzo, C.; McEwen, A.</p> <p>2008-12-01</p> <p>Digital elevation models (DEMs) derived from MRO Hi<span class="hlt">RISE</span> stereo images allow meter-scale topographic measurements in the north polar layered deposits (NPLD) and distinction of slope vs. albedo effects on apparent brightness of individual layers. Hi<span class="hlt">RISE</span> images do not show thin layers at the limit of resolution. Rather, fine layering, if it exists, appears to have been obscured by a more dust-rich mantling deposit which shows signs of eolian erosion and slumping. Stratigraphic sequences within the NPLD appear to be repeated within exposures observed by Hi<span class="hlt">RISE</span>, indicative of a record of periodic climate changes. Granular flows sourced from within the dark, basal unit are suggestive of, but do not require, the presence of water during their formation. Active mass wasting of frost and dust has been observed on steep NPLD scarps in early spring, similar to dry, loose snow avalanches on terrestrial slopes. Bright and dark streaks are seen to evolve during the northern summer, evidence for active eolian redistribution of frost and perhaps dark (non- volatile) material. Relatively dark reddish patches observed within the north polar residual cap during the summer indicate that the cap is very thin (<<span class="hlt">1</span> m) or more transparent in places. Hi<span class="hlt">RISE</span> images of exposures of the south polar layered deposits (SPLD) show rectilinear fractures that are continuous across several layers and whose orientation is not affected by the topography of the exposure, suggesting that they were formed before erosion of the SPLD. They appear to extend laterally and <span class="hlt">vertically</span> through the SPLD, like a joint set. While NPLD tectonism appears limited to isolated grabens, several faults have been observed by Hi<span class="hlt">RISE</span> in the SPLD, showing structural details including reverse fault splays that merge into bedding planes and possible evidence for thrust duplication. The faults may be the result of basal sliding (decollements) ramping into thrust faults near the margin of the SPLD.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70175253','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70175253"><span>Barriers to and opportunities for landward migration of coastal wetlands with sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Enwright, Nicholas M.; Griffith, Kereen T.; Osland, Michael J.</p> <p>2016-01-01</p> <p>In the 21st century, accelerated sea-level <span class="hlt">rise</span> and continued coastal development are expected to greatly alter coastal landscapes across the globe. Historically, many coastal ecosystems have responded to sea-level fluctuations via horizontal and <span class="hlt">vertical</span> movement on the landscape. However, anthropogenic activities, including urbanization and the construction of flood-prevention infrastructure, can produce barriers that impede ecosystem migration. Here we show where tidal saline wetlands have the potential to migrate landward along the northern Gulf of Mexico coast, one of the most sea-level <span class="hlt">rise</span> sensitive and wetland-rich regions of the world. Our findings can be used to identify migration corridors and develop sea-level <span class="hlt">rise</span> adaptation strategies to help ensure the continued availability of wetland-associated ecosystem goods and services.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70159462','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70159462"><span>The vulnerability of Indo-Pacific mangrove forests to sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Lovelock, Catherine E.; Cahoon, Donald R.; Friess, Daniel A.; Guntenspergen, Glenn R.; Krauss, Ken W.; Reef, Ruth; Rogers, Kerrylee; Saunders, Megan L.; Sidik, Frida; Swales, Andrew; Saintilan, Neil; Thuyen, Le Xuan; Triet, Tran</p> <p>2015-01-01</p> <p>Sea-level <span class="hlt">rise</span> can threaten the long-term sustainability of coastal communities and valuable ecosystems such as coral reefs, salt marshes and mangroves. Mangrove forests have the capacity to keep pace with sea-level <span class="hlt">rise</span> and to avoid inundation through <span class="hlt">vertical</span> accretion of sediments, which allows them to maintain wetland soil elevations suitable for plant growth. The Indo-Pacific region holds most of the world’s mangrove forests, but sediment delivery in this region is declining, owing to anthropogenic activities such as damming of rivers. This decline is of particular concern because the Indo-Pacific region is expected to have variable, but high, rates of future sea-level <span class="hlt">rise</span>. Here we analyse recent trends in mangrove surface elevation changes across the Indo-Pacific region using data from a network of surface elevation table instruments. We find that sediment availability can enable mangrove forests to maintain rates of soil-surface elevation gain that match or exceed that of sea-level <span class="hlt">rise</span>, but for 69 per cent of our study sites the current rate of sea-level <span class="hlt">rise</span> exceeded the soil surface elevation gain. We also present a model based on our field data, which suggests that mangrove forests at sites with low tidal range and low sediment supply could be submerged as early as 2070.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26466567','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26466567"><span>The vulnerability of Indo-Pacific mangrove forests to sea-level <span class="hlt">rise</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lovelock, Catherine E; Cahoon, Donald R; Friess, Daniel A; Guntenspergen, Glenn R; Krauss, Ken W; Reef, Ruth; Rogers, Kerrylee; Saunders, Megan L; Sidik, Frida; Swales, Andrew; Saintilan, Neil; Thuyen, Le Xuan; Triet, Tran</p> <p>2015-10-22</p> <p>Sea-level <span class="hlt">rise</span> can threaten the long-term sustainability of coastal communities and valuable ecosystems such as coral reefs, salt marshes and mangroves. Mangrove forests have the capacity to keep pace with sea-level <span class="hlt">rise</span> and to avoid inundation through <span class="hlt">vertical</span> accretion of sediments, which allows them to maintain wetland soil elevations suitable for plant growth. The Indo-Pacific region holds most of the world's mangrove forests, but sediment delivery in this region is declining, owing to anthropogenic activities such as damming of rivers. This decline is of particular concern because the Indo-Pacific region is expected to have variable, but high, rates of future sea-level <span class="hlt">rise</span>. Here we analyse recent trends in mangrove surface elevation changes across the Indo-Pacific region using data from a network of surface elevation table instruments. We find that sediment availability can enable mangrove forests to maintain rates of soil-surface elevation gain that match or exceed that of sea-level <span class="hlt">rise</span>, but for 69 per cent of our study sites the current rate of sea-level <span class="hlt">rise</span> exceeded the soil surface elevation gain. We also present a model based on our field data, which suggests that mangrove forests at sites with low tidal range and low sediment supply could be submerged as early as 2070.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19750006925','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19750006925"><span>Liquid jet pumped by <span class="hlt">rising</span> gas bubbles</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hussain, N. A.; Siegel, R.</p> <p>1975-01-01</p> <p>A two-phase mathematical model is proposed for calculating the induced turbulent <span class="hlt">vertical</span> liquid flow. Bubbles provide a large buoyancy force and the associated drag on the liquid moves the liquid upward. The liquid pumped upward consists of the bubble wakes and the liquid brought into the jet region by turbulent entrainment. The expansion of the gas bubbles as they <span class="hlt">rise</span> through the liquid is taken into account. The continuity and momentum equations are solved numerically for an axisymmetric air jet submerged in water. Water pumping rates are obtained as a function of air flow rate and depth of submergence. Comparisons are made with limited experimental information in the literature.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3619298','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3619298"><span>Critical width of tidal flats triggers marsh collapse in the absence of sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Mariotti, Giulio; Fagherazzi, Sergio</p> <p>2013-01-01</p> <p>High rates of wave-induced erosion along salt marsh boundaries challenge the idea that marsh survival is dictated by the competition between <span class="hlt">vertical</span> sediment accretion and relative sea-level <span class="hlt">rise</span>. Because waves pounding marshes are often locally generated in enclosed basins, the depth and width of surrounding tidal flats have a pivoting control on marsh erosion. Here, we show the existence of a threshold width for tidal flats bordering salt marshes. Once this threshold is exceeded, irreversible marsh erosion takes place even in the absence of sea-level <span class="hlt">rise</span>. This catastrophic collapse occurs because of the positive feedbacks among tidal flat widening by wave-induced marsh erosion, tidal flat deepening driven by wave bed shear stress, and local wind wave generation. The threshold width is determined by analyzing the 50-y evolution of 54 marsh basins along the US Atlantic Coast. The presence of a critical basin width is predicted by a dynamic model that accounts for both horizontal marsh migration and <span class="hlt">vertical</span> adjustment of marshes and tidal flats. Variability in sediment supply, rather than in relative sea-level <span class="hlt">rise</span> or wind regime, explains the different critical width, and hence erosion vulnerability, found at different sites. We conclude that sediment starvation of coastlines produced by river dredging and damming is a major anthropogenic driver of marsh loss at the study sites and generates effects at least comparable to the accelerating sea-level <span class="hlt">rise</span> due to global warming. PMID:23513219</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23513219','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23513219"><span>Critical width of tidal flats triggers marsh collapse in the absence of sea-level <span class="hlt">rise</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mariotti, Giulio; Fagherazzi, Sergio</p> <p>2013-04-02</p> <p>High rates of wave-induced erosion along salt marsh boundaries challenge the idea that marsh survival is dictated by the competition between <span class="hlt">vertical</span> sediment accretion and relative sea-level <span class="hlt">rise</span>. Because waves pounding marshes are often locally generated in enclosed basins, the depth and width of surrounding tidal flats have a pivoting control on marsh erosion. Here, we show the existence of a threshold width for tidal flats bordering salt marshes. Once this threshold is exceeded, irreversible marsh erosion takes place even in the absence of sea-level <span class="hlt">rise</span>. This catastrophic collapse occurs because of the positive feedbacks among tidal flat widening by wave-induced marsh erosion, tidal flat deepening driven by wave bed shear stress, and local wind wave generation. The threshold width is determined by analyzing the 50-y evolution of 54 marsh basins along the US Atlantic Coast. The presence of a critical basin width is predicted by a dynamic model that accounts for both horizontal marsh migration and <span class="hlt">vertical</span> adjustment of marshes and tidal flats. Variability in sediment supply, rather than in relative sea-level <span class="hlt">rise</span> or wind regime, explains the different critical width, and hence erosion vulnerability, found at different sites. We conclude that sediment starvation of coastlines produced by river dredging and damming is a major anthropogenic driver of marsh loss at the study sites and generates effects at least comparable to the accelerating sea-level <span class="hlt">rise</span> due to global warming.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AtmRe.186..107W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AtmRe.186..107W"><span>Effects of cloud condensate <span class="hlt">vertical</span> alignment on radiative transfer calculations in deep convective regions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Xiaocong</p> <p>2017-04-01</p> <p>Effects of cloud condensate <span class="hlt">vertical</span> alignment on radiative transfer process were investigated using cloud resolving model explicit simulations, which provide a surrogate for subgrid cloud geometry. Diagnostic results showed that the decorrelation length Lcw varies in the <span class="hlt">vertical</span> dimension, with larger Lcw occurring in convective clouds and smaller Lcw in cirrus clouds. A new parameterization of Lcw is proposed that takes into account such varying features and gives <span class="hlt">rise</span> to improvements in simulations of cloud radiative forcing (CRF) and radiative heating, i.e., the peak of bias is respectively reduced by 8 W m- 2 for SWCF and 2 W m- 2 for LWCF in comparison with Lcw = <span class="hlt">1</span> km. The role of Lcw in modulating CRFs is twofold. On the one hand, larger Lcw tends to increase the standard deviation of optical depth στ, as dense and tenuous parts of the clouds would be increasingly aligned in the <span class="hlt">vertical</span> dimension, thereby broadening the probability distribution. On the other hand, larger στ causes a decrease in the solar albedo and thermal emissivity, as implied in their convex functions on τ. As a result, increasing (decreasing) Lcwleads to decreased (increased) CRFs, as revealed by comparisons among Lcw = 0, Lcw = <span class="hlt">1</span> km andLcw = ∞. It also affects the <span class="hlt">vertical</span> structure of radiative flux and thus influences the radiative heating. A better representation of στ in the <span class="hlt">vertical</span> dimension yields an improved simulation of radiative heating. Although the importance of <span class="hlt">vertical</span> alignment of cloud condensate is found to be less than that of cloud cover in regards to their impacts on CRFs, it still has enough of an effect on modulating the cloud radiative transfer process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29051488','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29051488"><span>Coralgal reef morphology records punctuated sea-level <span class="hlt">rise</span> during the last deglaciation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Khanna, Pankaj; Droxler, André W; Nittrouer, Jeffrey A; Tunnell, John W; Shirley, Thomas C</p> <p>2017-10-19</p> <p>Coralgal reefs preserve the signatures of sea-level fluctuations over Earth's history, in particular since the Last Glacial Maximum 20,000 years ago, and are used in this study to indicate that punctuated sea-level <span class="hlt">rise</span> events are more common than previously observed during the last deglaciation. Recognizing the nature of past sea-level <span class="hlt">rises</span> (i.e., gradual or stepwise) during deglaciation is critical for informing models that predict future <span class="hlt">vertical</span> behavior of global oceans. Here we present high-resolution bathymetric and seismic sonar data sets of 10 morphologically similar drowned reefs that grew during the last deglaciation and spread 120 km apart along the south Texas shelf edge. Herein, six commonly observed terrace levels are interpreted to be generated by several punctuated sea-level <span class="hlt">rise</span> events forcing the reefs to shrink and backstep through time. These systematic and common terraces are interpreted to record punctuated sea-level <span class="hlt">rise</span> events over timescales of decades to centuries during the last deglaciation, previously recognized only during the late Holocene.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMOS21B..07D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMOS21B..07D"><span><span class="hlt">Vertical</span> land motion along the coast of Louisiana: Integrating satellite altimetry, tide gauge and GPS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dixon, T. H.; A Karegar, M.; Uebbing, B.; Kusche, J.; Fenoglio-Marc, L.</p> <p>2017-12-01</p> <p>Coastal Louisiana is experiencing the highest rate of relative sea-level <span class="hlt">rise</span> in North America due to the combination of sea-level <span class="hlt">rise</span> and subsidence of the deltaic plain. The land subsidence in this region is studied using various techniques, with continuous GPS site providing high temporal resolution. Here, we use high resolution tide-gauge data and advanced processing of satellite altimetry to derive <span class="hlt">vertical</span> displacements time series at NOAA tide-gauge stations along the coast (Figure <span class="hlt">1</span>). We apply state-of-the-art retracking techniques to process raw altimetry data, allowing high accuracy on range measurements close to the coast. Data from Jason-<span class="hlt">1</span>, -2 and -3, Envisat, Saral and Cryosat-2 are used, corrected for solid Earth tide, pole tide and tidal ocean loading, using background models consistent with the GPS processing technique. We reprocess the available GPS data using precise point positioning and estimate the rate uncertainty accounting for correlated noise. The displacement time series are derived by directly subtracting tide-gauge data from the altimetry sea-level anomaly data. The quality of the derived displacement rates is evaluated in Grand Isle, Amerada Pass and Shell Beach where GPS data are available adjacent to the tide gauges. We use this technique to infer <span class="hlt">vertical</span> displacement at tide gauges in New Orleans (New Canal Station) and Port Fourchon and Southwest Pass along the coastline.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24144771','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24144771"><span>Head and pelvic movement symmetry in horses during circular motion and in <span class="hlt">rising</span> trot.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Robartes, Helen; Fairhurst, Harriet; Pfau, Thilo</p> <p>2013-12-01</p> <p>Lameness examinations in horses often include lungeing and ridden exercise. To incorporate these exercises into the evidence-based decision making process aided by quantitative sensor based gait analysis, guideline values for movement asymmetry are needed. In this study, movement symmetry (MS) was quantified in horses during unridden and ridden trot on the straight and on the circle. Systematic changes in MS were expected as a result of the 'asymmetrical loading' caused by circular movement, the <span class="hlt">rising</span> trot and the combination of the two. Out of 23 horses (age 4-20 years, height 13.3-17.2 hands), 13 presented within normal limits for head movement and 22 for pelvic movement. Inertial measurement units assessed MS of <span class="hlt">vertical</span> head and sacral movement during trot in-hand, on the lunge and in <span class="hlt">rising</span> trot (straight, left/right circle). Changes in MS between straight line trot and ridden exercise on the circle were more pronounced for the head than for the sacrum. The highest amount of asymmetry was observed during <span class="hlt">rising</span> trot on the circle (symmetry index of the head: <span class="hlt">1</span>.23 for the left rein, 0.83 for the right rein; symmetry index of the sacrum 0.84 for the left rein, <span class="hlt">1</span>.15 for the right rein). Change in MS was significant between exercise conditions except for the difference between head displacement maxima. Horses had greatest asymmetry during <span class="hlt">rising</span> trot on the circle, with MS values of comparable magnitude to mild lameness. Copyright © 2013 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015GMD.....8...69R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015GMD.....8...69R"><span>Modelling turbulent <span class="hlt">vertical</span> mixing sensitivity using a <span class="hlt">1</span>-D version of NEMO</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Reffray, G.; Bourdalle-Badie, R.; Calone, C.</p> <p>2015-01-01</p> <p>Through two numerical experiments, a <span class="hlt">1</span>-D <span class="hlt">vertical</span> model called NEMO<span class="hlt">1</span>D was used to investigate physical and numerical turbulent-mixing behaviour. The results show that all the turbulent closures tested (k+l from Blanke and Delecluse, 1993, and two equation models: generic length scale closures from Umlauf and Burchard, 2003) are able to correctly reproduce the classical test of Kato and Phillips (1969) under favourable numerical conditions while some solutions may diverge depending on the degradation of the spatial and time discretization. The performances of turbulence models were then compared with data measured over a <span class="hlt">1</span>-year period (mid-2010 to mid-2011) at the PAPA station, located in the North Pacific Ocean. The modelled temperature and salinity were in good agreement with the observations, with a maximum temperature error between -2 and 2 °C during the stratified period (June to October). However, the results also depend on the numerical conditions. The <span class="hlt">vertical</span> RMSE varied, for different turbulent closures, from 0.<span class="hlt">1</span> to 0.3 °C during the stratified period and from 0.03 to 0.15 °C during the homogeneous period. This <span class="hlt">1</span>-D configuration at the PAPA station (called PAPA<span class="hlt">1</span>D) is now available in NEMO as a reference configuration including the input files and atmospheric forcing set described in this paper. Thus, all the results described can be recovered by downloading and launching PAPA<span class="hlt">1</span>D. The configuration is described on the NEMO site (<a href="http://www.nemo-ocean.eu/Using-NEMO/Configurations/C<span class="hlt">1</span>D_PAPA">http://www.nemo-ocean.eu/Using-NEMO/Configurations/C<span class="hlt">1</span>D_PAPA</a>). This package is a good starting point for further investigation of <span class="hlt">vertical</span> processes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014GMDD....7.5249R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014GMDD....7.5249R"><span>Modelling turbulent <span class="hlt">vertical</span> mixing sensitivity using a <span class="hlt">1</span>-D version of NEMO</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Reffray, G.; Bourdalle-Badie, R.; Calone, C.</p> <p>2014-08-01</p> <p>Through two numerical experiments, a <span class="hlt">1</span>-D <span class="hlt">vertical</span> model called NEMO<span class="hlt">1</span>D was used to investigate physical and numerical turbulent-mixing behaviour. The results show that all the turbulent closures tested (k + l from Blanke and Delecluse, 1993 and two equation models: Generic Lengh Scale closures from Umlauf and Burchard, 2003) are able to correctly reproduce the classical test of Kato and Phillips (1969) under favourable numerical conditions while some solutions may diverge depending on the degradation of the spatial and time discretization. The performances of turbulence models were then compared with data measured over a one-year period (mid-2010 to mid-2011) at the PAPA station, located in the North Pacific Ocean. The modelled temperature and salinity were in good agreement with the observations, with a maximum temperature error between -2 and 2 °C during the stratified period (June to October). However the results also depend on the numerical conditions. The <span class="hlt">vertical</span> RMSE varied, for different turbulent closures, from 0.<span class="hlt">1</span> to 0.3 °C during the stratified period and from 0.03 to 0.15 °C during the homogeneous period. This <span class="hlt">1</span>-D configuration at the PAPA station (called PAPA<span class="hlt">1</span>D) is now available in NEMO as a reference configuration including the input files and atmospheric forcing set described in this paper. Thus, all the results described can be recovered by downloading and launching PAPA<span class="hlt">1</span>D. The configuration is described on the NEMO site (<a href="http://www.nemo-ocean.eu/Using-NEMO/Configurations/C<span class="hlt">1</span>D_PAPA">http://www.nemo-ocean.eu/Using-NEMO/Configurations/C<span class="hlt">1</span>D_PAPA</a>). This package is a good starting point for further investigation of <span class="hlt">vertical</span> processes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29148355','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29148355"><span><span class="hlt">Vertical</span> Integration of Hospitals and Physicians: Economic Theory and Empirical Evidence on Spending and Quality.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Post, Brady; Buchmueller, Tom; Ryan, Andrew M</p> <p>2017-08-01</p> <p>Hospital-physician <span class="hlt">vertical</span> integration is on the <span class="hlt">rise</span>. While increased efficiencies may be possible, emerging research raises concerns about anticompetitive behavior, spending increases, and uncertain effects on quality. In this review, we bring together several of the key theories of <span class="hlt">vertical</span> integration that exist in the neoclassical and institutional economics literatures and apply these theories to the hospital-physician relationship. We also conduct a literature review of the effects of <span class="hlt">vertical</span> integration on prices, spending, and quality in the growing body of evidence ( n = 15) to evaluate which of these frameworks have the strongest empirical support. We find some support for <span class="hlt">vertical</span> foreclosure as a framework for explaining the observed results. We suggest a conceptual model and identify directions for future research. Based on our analysis, we conclude that <span class="hlt">vertical</span> integration poses a threat to the affordability of health services and merits special attention from policymakers and antitrust authorities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MS%26E..350a2016S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MS%26E..350a2016S"><span>Effect of Capillary Tube’s Shape on Capillary <span class="hlt">Rising</span> Regime for Viscos Fluids</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Soroush, F.; Moosavi, A.</p> <p>2018-05-01</p> <p>When properties of the displacing fluid are considered, the <span class="hlt">rising</span> profile of the penetrating fluid in a capillary tube deviates from its classical Lucas-Washburn profile. Also, shape of capillary tube can affect the <span class="hlt">rising</span> profile in different aspects. In this article, effect of capillary tube’s shape on the <span class="hlt">vertical</span> capillary motion in presence of gravity is investigated by considering the properties of the displacing fluid. According to the fact that the differential equation of the capillary <span class="hlt">rising</span> for a non-simple wall type is very difficult to solve analytically, a finite element simulation model is used for this study. After validation of the simulation model with an experiment that has been done with a simple capillary tube, shape of the capillary tube’s wall is changed in order to understand its effects on the capillary <span class="hlt">rising</span> and different motion regimes that may appear according to different geometries. The main focus of this article is on the sinusoidal wall shapes and comparing them with a simple wall.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16537141','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16537141"><span>On the rate and causes of twentieth century sea-level <span class="hlt">rise</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Miller, Laury; Douglas, Bruce C</p> <p>2006-04-15</p> <p>Both the rate and causes of twentieth century global sea-level <span class="hlt">rise</span> (GSLR) have been controversial. Estimates from tide-gauges range from less than one, to more than two millimetre yr(-<span class="hlt">1</span>). In contrast, values based on the processes mostly responsible for GSLR-mass increase (from mountain glaciers and the great high latitude ice masses) and volume increase (expansion due to ocean warming)-fall below this range. Either the gauge estimates are too high, or one (or both) of the component estimates is too low. Gauge estimates of GSLR have been in dispute for several decades because of <span class="hlt">vertical</span> land movements, especially due to glacial isostatic adjustment (GIA). More recently, the possibility has been raised that coastal tide-gauges measure exaggerated rates of sea-level <span class="hlt">rise</span> because of localized ocean warming. Presented here are two approaches to a resolution of these problems. The first is morphological, based on the limiting values of observed trends of twentieth century relative sea-level <span class="hlt">rise</span> as a function of distance from the centres of the ice loads at last glacial maximum. This observational approach, which does not depend on a geophysical model of GIA, supports values of GSLR near 2 mm yr(-<span class="hlt">1</span>). The second approach involves an analysis of long records of tide-gauge and hydrographic (in situ temperature and salinity) observations in the Pacific and Atlantic Oceans. It was found that sea-level trends from tide-gauges, which reflect both mass and volume change, are 2-3 times higher than rates based on hydrographic data which reveal only volume change. These results support those studies that put the twentieth century rate near 2 mm yr(-<span class="hlt">1</span>), thereby indicating that mass increase plays a much larger role than ocean warming in twentieth century GSLR.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28853277','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28853277"><span>Ambipolar Graphene-Quantum Dot Hybrid <span class="hlt">Vertical</span> Photodetector with a Graphene Electrode.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Che, Yongli; Zhang, Yating; Cao, Xiaolong; Zhang, Haiting; Song, Xiaoxian; Cao, Mingxuan; Yu, Yu; Dai, Haitao; Yang, Junbo; Zhang, Guizhong; Yao, Jianquan</p> <p>2017-09-20</p> <p>A strategy to fabricate an ambipolar near-infrared <span class="hlt">vertical</span> photodetector (VPD) by sandwiching a photoactive material as a channel film between the bottom graphene and top metal electrodes was developed. The channel length in the <span class="hlt">vertical</span> architecture was determined by the channel layer thickness, which can provide an ultrashort channel length without the need for a high-precision manufacturing process. The performance of VPDs with two types of semiconductor layers, a graphene-PbS quantum dot hybrid (GQDH) and PbS quantum dots (QDs), was measured. The GQDH VPD showed better photoelectric properties than the QD VPD because of the high mobility of graphene doped in the channel. The GQDH VPD exhibited excellent photoresponse properties with a responsivity of <span class="hlt">1</span>.6 × 10 4 A/W in the p-type regime and a fast response speed with a <span class="hlt">rise</span> time of 8 ms. The simple manufacture and the promising photoresponse of the GQDH VPDs reveal that an easy and effective way to fabricate high-performance ambipolar photodetectors was developed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20030066757&hterms=DENNIS+ALAN&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D10%26Ntt%3DDENNIS%252C%2BALAN','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20030066757&hterms=DENNIS+ALAN&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D10%26Ntt%3DDENNIS%252C%2BALAN"><span>MRO High Resolution Imaging Science Experiment (Hi<span class="hlt">RISE</span>): Instrument Development</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Delamere, Alan; Becker, Ira; Bergstrom, Jim; Burkepile, Jon; Day, Joe; Dorn, David; Gallagher, Dennis; Hamp, Charlie; Lasco, Jeffrey; Meiers, Bill</p> <p>2003-01-01</p> <p>The primary functional requirement of the Hi<span class="hlt">RISE</span> imager is to allow identification of both predicted and unknown features on the surface of Mars to a much finer resolution and contrast than previously possible. This results in a camera with a very wide swath width, 6km at 300km altitude, and a high signal to noise ratio, >100:<span class="hlt">1</span>. Generation of terrain maps, 30 cm <span class="hlt">vertical</span> resolution, from stereo images requires very accurate geometric calibration. The project limitations of mass, cost and schedule make the development challenging. In addition, the spacecraft stability must not be a major limitation to image quality. The nominal orbit for the science phase of the mission is a 3pm orbit of 255 by 320 km with periapsis locked to the south pole. The track velocity is approximately 3,400 m/s.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120015715','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120015715"><span>Space-based Observational Constraints for <span class="hlt">1</span>-D Plume <span class="hlt">Rise</span> Models</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Martin, Maria Val; Kahn, Ralph A.; Logan, Jennifer A.; Paguam, Ronan; Wooster, Martin; Ichoku, Charles</p> <p>2012-01-01</p> <p>We use a space-based plume height climatology derived from observations made by the Multi-angle Imaging SpectroRadiometer (MISR) instrument aboard the NASA Terra satellite to evaluate the ability of a plume-<span class="hlt">rise</span> model currently embedded in several atmospheric chemical transport models (CTMs) to produce accurate smoke injection heights. We initialize the plume-<span class="hlt">rise</span> model with assimilated meteorological fields from the NASA Goddard Earth Observing System and estimated fuel moisture content at the location and time of the MISR measurements. Fire properties that drive the plume-<span class="hlt">rise</span> model are difficult to estimate and we test the model with four estimates for active fire area and four for total heat flux, obtained using empirical data and Moderate Resolution Imaging Spectroradiometer (MODIS) re radiative power (FRP) thermal anomalies available for each MISR plume. We show that the model is not able to reproduce the plume heights observed by MISR over the range of conditions studied (maximum r2 obtained in all configurations is 0.3). The model also fails to determine which plumes are in the free troposphere (according to MISR), key information needed for atmospheric models to simulate properly smoke dispersion. We conclude that embedding a plume-<span class="hlt">rise</span> model using currently available re constraints in large-scale atmospheric studies remains a difficult proposition. However, we demonstrate the degree to which the fire dynamical heat flux (related to active fire area and sensible heat flux), and atmospheric stability structure influence plume <span class="hlt">rise</span>, although other factors less well constrained (e.g., entrainment) may also be significant. Using atmospheric stability conditions, MODIS FRP, and MISR plume heights, we offer some constraints on the main physical factors that drive smoke plume <span class="hlt">rise</span>. We find that smoke plumes reaching high altitudes are characterized by higher FRP and weaker atmospheric stability conditions than those at low altitude, which tend to remain confined</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li class="active"><span>5</span></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_5 --> <div id="page_6" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="101"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA11757&hterms=opal&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dopal','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA11757&hterms=opal&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dopal"><span>Rover's Wheel Churns Up Bright Martian Soil (<span class="hlt">Vertical</span>)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2009-01-01</p> <p><p/> NASA's Mars Exploration Rover Spirit acquired this mosaic on the mission's <span class="hlt">1</span>,202nd Martian day, or sol (May 21, 2007), while investigating the area east of the elevated plateau known as 'Home Plate' in the 'Columbia Hills.' The mosaic shows an area of disturbed soil, nicknamed 'Gertrude Weise' by scientists, made by Spirit's stuck right front wheel. <p/> The trench exposed a patch of nearly pure silica, with the composition of opal. It could have come from either a hot-spring environment or an environment called a fumarole, in which acidic, volcanic steam <span class="hlt">rises</span> through cracks. Either way, its formation involved water, and on Earth, both of these types of settings teem with microbial life. <p/> The image is presented here as a <span class="hlt">vertical</span> projection, as if looking straight down, and in false color, which brings out subtle color differences.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E3SWC..3302076K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E3SWC..3302076K"><span>Modelling of energy consumption at construction of high-<span class="hlt">rise</span> buildings</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Korol, Elena; Korol, Oleg</p> <p>2018-03-01</p> <p>High-<span class="hlt">rise</span> building structures in the course of its erection suppose primary use of methods provided for erection, concrete and external finishing works. Erection works do not differ significantly from usual ones: traditional equipment, accessories and techniques are used which are based on erection of structures in project position using a crane. Structures to be assembled in building frame include steel columns and beams, wall panels, form elements of columns, walls and floor structures. We can note heightened attention to operational control for quality of erection, but it is attributable to all works in the course of high-<span class="hlt">rise</span> construction. During high-<span class="hlt">rise</span> erection by means of cast in-situ reinforced concrete all formworks to be used do not have any special differences except systems specially designed for high-<span class="hlt">rise</span> erection using sliding formwork or <span class="hlt">vertical</span> traveling forms. In these systems special attention is paid to safety of elevated works. Working methods of placement and curing of concrete and structures as a whole remain traditional - the requirements for controlling such operations become toughened. The most evident differences in high-<span class="hlt">rise</span> erection with regard to equipment, machinery and accessories used are in means provided for load transportation and safety of works at heights. Particularity of internal finishing works which are also obligatory during construction of skyscrapers allows not considering them in as technological differences from usual construction as far as the «height» of its execution is limited by height of particular floor and determined by price and building class.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E%26ES..126a2013P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E%26ES..126a2013P"><span>Multi-layer planting as a strategy of greening the transitional space in high-<span class="hlt">rise</span> buildings: A review</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Prihatmanti, Rani; Taib, Nooriati</p> <p>2018-03-01</p> <p>The issues regarding the rapid development in the urban have resulted in the increasing number of infrastructure built, including the high-<span class="hlt">rise</span> buildings to accommodate the urban dwellers. Lack of greeneries due to the land limitation in the urban area has increased the surface radiation as well as the air temperature that leads to the Urban Heat Island (UHI) phenomena. Where urban land is limited, growing plants <span class="hlt">vertically</span> could be a solution. Plants, which are widely known as one of the sustainability elements in the built environment could be integrated in building as a part of urban faming by growing edible plant species. This is also to address the food security issue in the urban as well as high-density cities. Since space is limited, the function of transitional space could be optimized for the green space. This paper explores the strategy of greening transitional space in the high-<span class="hlt">rise</span> setting. To give a maximum impact in a limited space, multi-layer planting concept could be introduced. This concept is believed that multiple layers of plants could modify the microclimate, as well as the radiation to the building, compare to single layer plant. In addition to that, the method selected also determines the efficacy of the <span class="hlt">vertical</span> greeneries. However, there are many other limitations related to the multi-layer planting method if installed in a transitional space that needs to be further studied. Despite its limitations, the application of <span class="hlt">vertical</span> greeneries with multi-layer planting concept could be a promising solution for greening the limited space as well as improving the thermal comfort in the high-<span class="hlt">rise</span> building.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.B51N..05A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.B51N..05A"><span>Spatial Variability of Salt Marsh <span class="hlt">Vertical</span> Accretion and Carbon Burial Rates along the Gulf of Mexico at Local and Regional Scales</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Arriola, J.; Cable, J. E.</p> <p>2017-12-01</p> <p>Many studies quantifying salt marsh <span class="hlt">vertical</span> accretion and carbon burial have been conducted along the Gulf of Mexico over the past several decades. These results are often used in conjunction with sea level <span class="hlt">rise</span> estimates to evaluate the long term storage, and potential release, of carbon as salt marshes are overtaken by <span class="hlt">rising</span> waters. However, results from these studies are not always comparable because of diverse sampling and analytical methods, which may skew regional averages. In addition, salt marsh <span class="hlt">vertical</span> accretion and carbon burial rates can be highly variable on local scales depending on sampling locations within the marsh, e.g. levee vs marsh plain, and methods to determine carbon quantity, such as utilizing linear relationships between % organic matter and % carbon from other studies. Anthropogenic impacts on accretion and carbon burial may also influence interpretation of results. Utilizing consistent methods for local and regional marsh research will improve the accuracy of accretion and burial rates which is fundamental to our ability to predict responses to climate change. Our study examined sediment cores extracted from 6 salt marshes - 5 marshes along Texas to Florida coasts and <span class="hlt">1</span> marsh on the Florida Atlantic coast. These marshes were selected for minimal human influence and consistent sampling and analytical methodologies were employed to compare <span class="hlt">vertical</span> accretion and carbon burial variability on local and regional scales. Total organic carbon (TOC) and total nitrogen were determined via direct measurement and accretion rates were calculated based on 210Pb via 210Po alpha spectrometry. The lowest TOC inventory was found at Mission-Aransas, TX (18.57 g OC), whereas the highest was found at Apalachicola, FL (35.05 g OC). Anahuac, TX, was found to have the highest modern <span class="hlt">vertical</span> accretion rates of all 6 sites, whereas Guana Tolomato-Matanzas, FL, has the lowest. This research yields regional carbon burial estimates for the Gulf of Mexico using</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.G43B1052B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.G43B1052B"><span>Relative and Geocentric Sea Level <span class="hlt">Rise</span> Along the U.S. West Coast</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Burgette, R. J.; Watson, C. S.</p> <p>2015-12-01</p> <p>The rate of sea level change relative to the land along the West Coast of the U.S. varies over a range of +5 to -2 mm/yr, as observed across the set of long-running tide gauges. We analyze tide gauge data in a network approach that accounts for temporal and spatial correlations in the time series of water levels observed at the stations. This analysis yields a set of rate estimates and realistic uncertainties that are minimally affected by varying durations of observations. The analysis has the greatest impact for tide gauges with short records, as the adjusted rate uncertainties for 2 to 3 decade duration tide gauges approach those estimated from unadjusted century-scale time series. We explore the sources of the wide range of observed relative sea level rates through comparison with: <span class="hlt">1</span>) estimated <span class="hlt">vertical</span> deformation rates derived from repeated leveling and GPS, 2) relative sea level change predicted from models of glacial isostatic adjustment, and 3) geocentric sea level rates estimated from satellite altimetry and century-scale reconstructions. Tectonic deformation is the dominant signal in the relative sea level rates along the Cascadia portion of the coast, and is consistent with along-strike variation in locking behavior on the plate interface. Rates of <span class="hlt">vertical</span> motion are lower along the transform portion of the plate boundary and include anthropogenic effects, but there are significant tectonic signals, particularly in the western Transverse Ranges of California where the crust is shortening across reverse faults. Preliminary analysis of different strategies of estimating the magnitude of geocentric sea level <span class="hlt">rise</span> suggest significant discrepancies between approaches. We will examine the implications of these discrepancies for understanding the process of regional geocentric sea level <span class="hlt">rise</span> in the northeastern Pacific Ocean, and associated projected impacts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20180001857','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20180001857"><span>Global and Regional Sea Level <span class="hlt">Rise</span> Scenarios for the United States</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sweet, William V.; Kopp, Robert E.; Weaver, Christopher P.; Obeysekera, Jayantha; Horton, Radley M.; Thieler, E. Robert; Zervas, Chris</p> <p>2017-01-01</p> <p>The Sea Level <span class="hlt">Rise</span> and Coastal Flood Hazard Scenarios and Tools Interagency Task Force, jointly convened by the U.S. Global Change Research Program (USGCRP) and the National Ocean Council (NOC), began its work in August 2015. The Task Force has focused its efforts on three primary tasks: <span class="hlt">1</span>) updating scenarios of global mean sea level (GMSL) <span class="hlt">rise</span>, 2) integrating the global scenarios with regional factors contributing to sea level change for the entire U.S. coastline, and 3) incorporating these regionally appropriate scenarios within coastal risk management tools and capabilities deployed by individual agencies in support of the needs of specific stakeholder groups and user communities. This technical report focuses on the first two of these tasks and reports on the production of gridded relative sea level (RSL, which includes both ocean-level change and <span class="hlt">vertical</span> land motion) projections for the United States associated with an updated set of GMSL scenarios. In addition to supporting the longer-term Task Force effort, this new product will be an important input into the USGCRP Sustained Assessment process and upcoming Fourth National Climate Assessment (NCA4) due in 2018. This report also serves as a key technical input into the in-progress USGCRP Climate Science Special Report (CSSR).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApSS..422..388W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApSS..422..388W"><span>Copper <span class="hlt">vertical</span> micro dendrite fin arrays and their superior boiling heat transfer capability</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Ya-Qiao; Lyu, Shu-Shen; Luo, Jia-Li; Luo, Zhi-Yong; Fu, Yuan-Xiang; Heng, Yi; Zhang, Jian-Hui; Mo, Dong-Chuan</p> <p>2017-11-01</p> <p>Micro pin fin arrays have been widely used in electronic cooling, micro reactors, catalyst support, and wettability modification and so on, and a facile way to produce better micro pin fin arrays is demanded. Herein, a simple electrochemical method has been developed to fabricate copper <span class="hlt">vertical</span> micro dendrite fin arrays (Cu-VMDFA) with controllable shapes, number density and height. High copper sulphate concentration is one key point to make the dendrite stand <span class="hlt">vertically</span>. Besides, the applied current should <span class="hlt">rise</span> at an appropriate rate to ensure the copper dendrite can grow <span class="hlt">vertically</span> on its own. The Cu-VMDFA can significantly enhance the heat transfer coefficient by approximately twice compared to the plain copper surface. The Cu-VMDFA may be widely used in boiling heat transfer areas such as nuclear power plants, electronic cooling, heat exchangers, and so on.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED017119.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED017119.pdf"><span>HIGH <span class="hlt">RISE</span> OR LOW <span class="hlt">RISE</span>. A STUDY OF DECISION FACTORS IN RESIDENCE HALLS PLANNING.</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Educational Facilities Labs., Inc., New York, NY.</p> <p></p> <p>THE PURPOSE OF THIS REPORT IS TO SERVE COLLEGE OFFICIALS, HOUSING ADMINISTRATORS, PLANNING GROUPS AND ARCHITECTS BY FOCUSING ON THE DECISION FACTORS WHICH RELATE TO HIGH-<span class="hlt">RISE</span> AND LOW-<span class="hlt">RISE</span> STUDENT HOUSING. DECISION FACTORS INCLUDE--(<span class="hlt">1</span>) LAND USE IMPLICATIONS, (2) SITE REQUIREMENTS--BUILDING CODES, SUB-SOIL CONSIDERATIONS, NATURAL TERRAIN,…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29504758','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29504758"><span>Crystallization Behavior of Poly(ethylene oxide) in <span class="hlt">Vertically</span> Aligned Carbon Nanotube Array.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sheng, Jiadong; Zhou, Shenglin; Yang, Zhaohui; Zhang, Xiaohua</p> <p>2018-03-27</p> <p>We investigate the effect of the presence of <span class="hlt">vertically</span> aligned multiwalled carbon nanotubes (CNTs) on the orientation of poly(ethylene oxide) (PEO) lamellae and PEO crystallinity. The high alignment of carbon nanotubes acting as templates probably governs the orientation of PEO lamellae. This templating effect might result in the lamella planes of PEO crystals oriented along a direction parallel to the long axis of the nanotubes. The presence of aligned carbon nanotubes also gives <span class="hlt">rise</span> to the decreases in PEO crystallinity, crystallization temperature, and melting temperature due to the perturbation of carbon nanotubes to the crystallization of PEO. These effects have significant implications for controlling the orientation of PEO lamellae and decreasing the crystallinity of PEO and thickness of PEO lamellae, which have significant impacts on ion transport in PEO/CNT composite and the capacitive performance of PEO/CNT composite. Both the decreased PEO crystallinity and the orientation of PEO lamellae along the long axes of <span class="hlt">vertically</span> aligned CNTs give <span class="hlt">rise</span> to the decrease in the charge transfer resistance, which is associated with the improvements in the ion transport and capacitive performance of PEO/CNT composite.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EurSS..46....1K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EurSS..46....1K"><span>Complex of solonetzes and <span class="hlt">vertic</span> chestnut soils in the manych-gudilo depression</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kovda, I. V.; Morgun, E. P.; Il'ina, L. P.</p> <p>2013-01-01</p> <p>Morphological, physicochemical, and isotopic properties of a two-member soil complex developed under dry steppe have been studied in the central part of the Manych Depression. The soils are formed on chocolate-colored clayey sediments, and have pronounced microrelief and the complex vegetation pattern. A specific feature of the studied soil complex is the inverse position of its components: <span class="hlt">vertic</span> chestnut soil occupies the microhigh, while solonetz is in the microlow. The formation of such complexes is explained by the biological factor, i.e., by the destruction of the solonetzic horizon under the impact of vegetation and earth-burrowing animals with further transformation under steppe plants and dealkalinization of the soil in the microhighs. The manifestation of <span class="hlt">vertic</span> features and shrink-swell process in soils of the complex developing in dry steppe are compared with those in the <span class="hlt">vertic</span> soils of the Central Pre-Caucasus formed under more humid environment. It is supposed that slickensides in the investigated <span class="hlt">vertic</span> chestnut soil are relict feature inherited from the former wetter stage of the soil development and are subjected to a gradual degradation at present. In the modern period, <span class="hlt">vertic</span> processes are weak and cannot be distinctly diagnosed. However, their activation may take place upon an increase of precipitation or the <span class="hlt">rise</span> in the groundwater level.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70035507','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70035507"><span>Analysis of lidar elevation data for improved identification and delineation of lands vulnerable to sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Gesch, Dean B.</p> <p>2009-01-01</p> <p>The importance of sea-level <span class="hlt">rise</span> in shaping coastal landscapes is well recognized within the earth science community, but as with many natural hazards, communicating the risks associated with sea-level <span class="hlt">rise</span> remains a challenge. Topography is a key parameter that influences many of the processes involved in coastal change, and thus, up-to-date, high-resolution, high-accuracy elevation data are required to model the coastal environment. Maps of areas subject to potential inundation have great utility to planners and managers concerned with the effects of sea-level <span class="hlt">rise</span>. However, most of the maps produced to date are simplistic representations derived from older, coarse elevation data. In the last several years, vast amounts of high quality elevation data derived from lidar have become available. Because of their high <span class="hlt">vertical</span> accuracy and spatial resolution, these lidar data are an excellent source of up-to-date information from which to improve identification and delineation of vulnerable lands. Four elevation datasets of varying resolution and accuracy were processed to demonstrate that the improved quality of lidar data leads to more precise delineation of coastal lands vulnerable to inundation. A key component of the comparison was to calculate and account for the <span class="hlt">vertical</span> uncertainty of the elevation datasets. This comparison shows that lidar allows for a much more detailed delineation of the potential inundation zone when compared to other types of elevation models. It also shows how the certainty of the delineation of lands vulnerable to a given sea-level <span class="hlt">rise</span> scenario is much improved when derived from higher resolution lidar data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AtmEn..54..603S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AtmEn..54..603S"><span>Cloud <span class="hlt">rise</span> model for radiological dispersal devices events</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sharon, Avi; Halevy, Itzhak; Sattinger, Daniel; Yaar, Ilan</p> <p>2012-07-01</p> <p>As a part of the preparedness and response to possible radiological terror events, it is important to model the evolution of the radioactive cloud immediately after its formation, as a function of time, explosive quantity and local meteorological conditions. One of the major outputs of a cloud <span class="hlt">rise</span> models is the evaluation of cloud top height, which is an essential input for most of the succeeding atmospheric dispersion models. This parameter strongly affects the radiological consequences of the event. Most of the cloud <span class="hlt">rise</span> models used today, have been developed according to experiments were large quantities of explosives were used, within the range of hundreds of kilograms of TNT. The majority of these models, however, fail to address Radiological Dispersion Devices (RDD) events, which are typically characterized by smaller amounts of TNT. In this paper, a new, semi-empirical model that describes the <span class="hlt">vertical</span> evolution of the cloud up to its effective height as a function of time, explosive quantity, atmospheric stability and horizontal wind speed, is presented. The database for this model is taken from five sets of experiments done in Israel during 2006-2009 under the "Green Field" (GF) project, using 0.25-100 kg of TNT.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFMGC21A0506X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMGC21A0506X"><span>Assessing Sea Level <span class="hlt">Rise</span> Impacts on the Surficial Aquifer in the Kennedy Space Center Region</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xiao, H.; Wang, D.; Hagen, S. C.; Medeiros, S. C.; Warnock, A. M.; Hall, C. R.</p> <p>2014-12-01</p> <p>Global sea level <span class="hlt">rise</span> in the past century due to climate change has been seen at an average rate of approximately <span class="hlt">1</span>.7-2.2 mm per year, with an increasing rate over the next century. The increasing SLR rate poses a severe threat to the low-lying land surface and the shallow groundwater system in the Kennedy Space Center in Florida, resulting in saltwater intrusion and groundwater induced flooding. A three-dimensional groundwater flow and salinity transport model is implemented to investigate and evaluate the extent of floods due to <span class="hlt">rising</span> water table as well as saltwater intrusion. The SEAWAT model is chosen to solve the variable-density groundwater flow and salinity transport governing equations and simulate the regional-scale spatial and temporal evolution of groundwater level and chloride concentration. The horizontal resolution of the model is 50 m, and the <span class="hlt">vertical</span> domain includes both the Surficial Aquifer and the Floridan Aquifer. The numerical model is calibrated based on the observed hydraulic head and chloride concentration. The potential impacts of sea level <span class="hlt">rise</span> on saltwater intrusion and groundwater induced flooding are assessed under various sea level <span class="hlt">rise</span> scenarios. Based on the simulation results, the potential landward movement of saltwater and freshwater fringe is projected. The existing water supply wells are examined overlaid with the projected salinity distribution map. The projected Surficial Aquifer water tables are overlaid with data of high resolution land surface elevation, land use and land cover, and infrastructure to assess the potential impacts of sea level <span class="hlt">rise</span>. This study provides useful tools for decision making on ecosystem management, water supply planning, and facility management.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-KSC-314D-0449_018.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-KSC-314D-0449_018.html"><span>EFT-<span class="hlt">1</span> Delta IV Heavy lift to <span class="hlt">vertical</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2014-10-01</p> <p>This close-up view shows the United Launch Alliance Delta IV Heavy rocket for Exploration Flight Test-<span class="hlt">1</span> being raised into the <span class="hlt">vertical</span> position at the pad at Space Launch Complex 37 at Cape Canaveral Air Force Station in Florida. The Delta IV Heavy is being readied to launch Orion on its first flight test. During its first flight test, Orion will travel farther into space than any human spacecraft has gone in more than 40 years. The data gathered during the flight will influence design decisions, validate existing computer models and innovative new approaches to space systems development, as well as reduce overall mission risks and costs for later Orion flights. Liftoff of Orion on the first flight test is planned for December 2014.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-KSC-314D-0449_027.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-KSC-314D-0449_027.html"><span>EFT-<span class="hlt">1</span> Delta IV Heavy lift to <span class="hlt">vertical</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2014-10-01</p> <p>The United Launch Alliance Delta IV Heavy rocket for Exploration Flight Test-<span class="hlt">1</span> is lifted to the <span class="hlt">vertical</span> position in the mobile service tower on the pad at Space Launch Complex 37 at Cape Canaveral Air Force Station in Florida. The Delta IV Heavy is being readied to launch Orion on its first flight test. During its first flight test, Orion will travel farther into space than any human spacecraft has gone in more than 40 years. The data gathered during the flight will influence design decisions, validate existing computer models and innovative new approaches to space systems development, as well as reduce overall mission risks and costs for later Orion flights. Liftoff of Orion on the first flight test is planned for December 2014.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-KSC-314D-0449_001.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-KSC-314D-0449_001.html"><span>EFT-<span class="hlt">1</span> Delta IV Heavy lift to <span class="hlt">vertical</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2014-10-01</p> <p>The United Launch Alliance Delta IV Heavy rocket for Exploration Flight Test-<span class="hlt">1</span> is being lifted to the <span class="hlt">vertical</span> position at the pad at Space Launch Complex 37 at Cape Canaveral Air Force Station in Florida. The Delta IV Heavy is being readied to launch Orion on its first flight test. During its first flight test, Orion will travel farther into space than any human spacecraft has gone in more than 40 years. The data gathered during the flight will influence design decisions, validate existing computer models and innovative new approaches to space systems development, as well as reduce overall mission risks and costs for later Orion flights. Liftoff of Orion on the first flight test is planned for December 2014.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29759973','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29759973"><span>Enhanced Glucose Control Following <span class="hlt">Vertical</span> Sleeve Gastrectomy Does Not Require a β-Cell Glucagon-Like Peptide <span class="hlt">1</span> Receptor.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Douros, Jonathan D; Lewis, Alfor G; Smith, Eric P; Niu, JingJing; Capozzi, Megan; Wittmann, April; Campbell, Jonathan; Tong, Jenny; Wagner, Constance; Mahbod, Parinaz; Seeley, Randy; D'Alessio, David A</p> <p>2018-05-14</p> <p>Bariatric surgeries, including <span class="hlt">vertical</span> sleeve gastrectomy (VSG), resolve diabetes in 40-50% of patients. Studies examining the molecular mechanisms underlying this effect have centered on the role of the insulinotropic glucagon-like peptide <span class="hlt">1</span> (GLP-<span class="hlt">1</span>), in great part because of the ∼10-fold <span class="hlt">rise</span> in its circulating levels after surgery. However, there is currently debate over the role of direct β-cell signaling by GLP-<span class="hlt">1</span> to mediate improved glucose tolerance following surgery. In order to assess the importance of β-cell GLP-<span class="hlt">1</span> receptor (GLP-<span class="hlt">1</span>R) for improving glucose control after VSG, a mouse model of this procedure was developed and combined with a genetically modified mouse line allowing an inducible, β-cell specific Glp<span class="hlt">1</span>r knockdown ( Glp<span class="hlt">1</span>r β-cell-ko ). Mice with VSG lost ∼20% of body weight over 30 days compared to sham-operated controls and had a ∼60% improvement in glucose tolerance. Isolated islets from VSG mice had significantly greater insulin responses to glucose than controls. Glp<span class="hlt">1</span>r knockdown in β-cells caused glucose intolerance in diet-induced obese mice compared to obese controls, but VSG improved glycemic profiles to similar levels during oral and intraperitoneal glucose challenges in Glp<span class="hlt">1</span>r βcell-ko and Glp<span class="hlt">1</span>r WT mice. Therefore, while the β-cell GLP-<span class="hlt">1</span>R seems to be important for maintaining glucose tolerance in obese mice, in these experiments it is dispensable for the improvement in glucose tolerance after VSG. Moreover, the metabolic physiology activated by VSG can overcome the deficits in glucose regulation caused by lack of β-cell GLP-<span class="hlt">1</span> signaling in obesity. © 2018 by the American Diabetes Association.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017TCry...11.1327R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017TCry...11.1327R"><span>Brief communication: The global signature of post-1900 land ice wastage on <span class="hlt">vertical</span> land motion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Riva, Riccardo E. M.; Frederikse, Thomas; King, Matt A.; Marzeion, Ben; van den Broeke, Michiel R.</p> <p>2017-06-01</p> <p>Melting glaciers, ice caps and ice sheets have made an important contribution to sea-level <span class="hlt">rise</span> through the last century. Self-attraction and loading effects driven by shrinking ice masses cause a spatially varying redistribution of ocean waters that affects reconstructions of past sea level from sparse observations. We model the solid-earth response to ice mass changes and find significant <span class="hlt">vertical</span> deformation signals over large continental areas. We show how deformation rates have been strongly varying through the last century, which implies that they should be properly modelled before interpreting and extrapolating recent observations of <span class="hlt">vertical</span> land motion and sea-level change.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/392450-application-unsteady-state-model-predicting-vertical-temperature-distribution-existing-atrium','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/392450-application-unsteady-state-model-predicting-vertical-temperature-distribution-existing-atrium"><span>Application of an unsteady-state model for predicting <span class="hlt">vertical</span> temperature distribution to an existing atrium</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Takemasa, Yuichi; Togari, Satoshi; Arai, Yoshinobu</p> <p>1996-11-01</p> <p><span class="hlt">Vertical</span> temperature differences tend to be great in a large indoor space such as an atrium, and it is important to predict variations of <span class="hlt">vertical</span> temperature distribution in the early stage of the design. The authors previously developed and reported on a new simplified unsteady-state calculation model for predicting <span class="hlt">vertical</span> temperature distribution in a large space. In this paper, this model is applied to predicting the <span class="hlt">vertical</span> temperature distribution in an existing low-<span class="hlt">rise</span> atrium that has a skylight and is affected by transmitted solar radiation. Detailed calculation procedures that use the model are presented with all the boundary conditions, andmore » analytical simulations are carried out for the cooling condition. Calculated values are compared with measured results. The results of the comparison demonstrate that the calculation model can be applied to the design of a large space. The effects of occupied-zone cooling are also discussed and compared with those of all-zone cooling.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21187763','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21187763"><span><span class="hlt">Vertical</span> position of the orbits in nonsyndromic plagiocephaly in childhood and its relation to <span class="hlt">vertical</span> strabismus.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Eveleens, Jordi R J; Mathijssen, Irene M; Lequin, Maarten H; Polling, Jan-Roelof; Looman, Caspar W N; Simonsz, Huibert J</p> <p>2011-01-01</p> <p>To determine the existence of a correlation between the <span class="hlt">vertical</span> angle of strabismus and the <span class="hlt">vertical</span> angle between the orbital axes in nonsyndromic plagiocephaly in childhood. Patients were included when diagnosed with plagiocephaly. Orthoptic measurements showed a <span class="hlt">vertical</span> strabismus and three-dimensional computed tomographic (CT) imaging of the skull was available. Patients were excluded if plagiocephaly was part of a syndrome or if any surgical intervention had taken place before our measurements. Three-dimensional CT imaging was used to calculate the <span class="hlt">vertical</span> angle between the orbital axes in 3 reference planes (VAO) perpendicular to a line of reference through the lower borders of the maxilla (VAOmax), both auditory canals (VAOaud), and the lower points of the external occipital protuberances (VAOocc). Fourteen patients were included (mean age, 14 mo). Three-dimensional CT measurements showed a mean (SD) VAOmax of <span class="hlt">1</span>.70 (2.31) degrees, VAOaud of -<span class="hlt">1</span>.54 (<span class="hlt">1</span>.46) degrees, and VAOocc of -2.06 (4.29) degrees (a negative value indicates that the eye on the affected side was situated lower in the head). The mean <span class="hlt">vertical</span> angle of strabismus was -2.39 (4.69) degrees in gaze toward the affected side, 3.66 (3.77) degrees in gaze ahead, and 8.14 (5.63) degrees in gaze toward the nonaffected side. The Pearson test showed no significant correlations. The clinical observation that <span class="hlt">vertical</span> strabismus in adult plagiocephaly is correlated with the <span class="hlt">vertical</span> angle of the orbital axes could not be confirmed in young children.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_6 --> <div id="page_7" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="121"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=nucleation&id=EJ823750','ERIC'); return false;" href="https://eric.ed.gov/?q=nucleation&id=EJ823750"><span>On Capillary <span class="hlt">Rise</span> and Nucleation</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Prasad, R.</p> <p>2008-01-01</p> <p>A comparison of capillary <span class="hlt">rise</span> and nucleation is presented. It is shown that both phenomena result from a balance between two competing energy factors: a volume energy and a surface energy. Such a comparison may help to introduce nucleation with a topic familiar to the students, capillary <span class="hlt">rise</span>. (Contains <span class="hlt">1</span> table and 3 figures.)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018HMT....54..353L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018HMT....54..353L"><span>Recognition and measurement gas-liquid two-phase flow in a <span class="hlt">vertical</span> concentric annulus at high pressures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Hao; Sun, Baojiang; Guo, Yanli; Gao, Yonghai; Zhao, Xinxin</p> <p>2018-02-01</p> <p>The air-water flow characteristics under pressure in the range of <span class="hlt">1</span>-6 MPa in a <span class="hlt">vertical</span> annulus were evaluated in this report. Time-resolved bubble <span class="hlt">rising</span> velocity and void fraction were also measured using an electrical void fraction meter. The results showed that the pressure has remarkable effect on the density, bubble size and <span class="hlt">rise</span> velocity of the gas. Four flow patterns (bubble, cap-bubble, cap-slug, and churn) were also observed instead of Taylor bubble at high pressure. Additionally, the transition process from bubble to cap-bubble was investigated at atmospheric and high pressures, respectively. The results revealed that the flow regime transition criteria for atmospheric pressure do not work at high pressure, hence a new flow regime transition model for annular flow channel geometry was developed to predict the flow regime transition, which thereafter exhibited high accuracy at high pressure condition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010EGUGA..12.7656J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010EGUGA..12.7656J"><span>Sea-level Fingerprinting, <span class="hlt">Vertical</span> Crustal Motion from GIA, and Projections of Relative Sea-level Change in the Canadian Arctic</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>James, Thomas; Simon, Karen; Forbes, Donald; Dyke, Arthur; Mazzotti, Stephane</p> <p>2010-05-01</p> <p>We present projections of relative sea-level <span class="hlt">rise</span> in the 21st century for communities in the Canadian Arctic. First, for selected communities, we determine the sea-level fingerprinting response from Antarctica, Greenland, and mountain glaciers and ice caps. Then, for various published projections of global sea-level change in the 21st century, we determine the local amount of "absolute" sea-level change. We next determine the <span class="hlt">vertical</span> land motion arising from glacial isostatic adjustment (GIA) and incorporate this into the estimates of absolute sea-level change to obtain projections of relative sea-level change. The sea-level fingerprinting effect is especially important in the Canadian Arctic owing to proximity to Arctic ice caps and especially to the Greenland ice sheet. Its effect is to reduce the range of projected relative sea-level change compared to the range of global sea-level projections. <span class="hlt">Vertical</span> crustal motion is assessed through empirically derived regional isobases, the Earth's predicted response to ice-sheet loading and unloading by the ICE-5G ice sheet reconstruction, and Global Positioning System <span class="hlt">vertical</span> velocities. Owing to the large rates of crustal uplift from glacial isostatic adjustment across a large region of central Arctic Canada, many communities are projected to experience relative sea-level fall despite projections of global sea-level <span class="hlt">rise</span>. Where uplift rates are smaller, such as eastern Baffin Island and the western Canadian Arctic, sea-level is projected to <span class="hlt">rise</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26774785','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26774785"><span>Moonlight Drives Ocean-Scale Mass <span class="hlt">Vertical</span> Migration of Zooplankton during the Arctic Winter.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Last, Kim S; Hobbs, Laura; Berge, Jørgen; Brierley, Andrew S; Cottier, Finlo</p> <p>2016-01-25</p> <p>In extreme high-latitude marine environments that are without solar illumination in winter, light-mediated patterns of biological migration have historically been considered non-existent [<span class="hlt">1</span>]. However, diel <span class="hlt">vertical</span> migration (DVM) of zooplankton has been shown to occur even during the darkest part of the polar night, when illumination levels are exceptionally low [2, 3]. This paradox is, as yet, unexplained. Here, we present evidence of an unexpected uniform behavior across the entire Arctic, in fjord, shelf, slope and open sea, where <span class="hlt">vertical</span> migrations of zooplankton are driven by lunar illumination. A shift from solar-day (24-hr period) to lunar-day (24.8-hr period) <span class="hlt">vertical</span> migration takes place in winter when the moon <span class="hlt">rises</span> above the horizon. Further, mass sinking of zooplankton from the surface waters and accumulation at a depth of ∼50 m occurs every 29.5 days in winter, coincident with the periods of full moon. Moonlight may enable predation of zooplankton by carnivorous zooplankters, fish, and birds now known to feed during the polar night [4]. Although primary production is almost nil at this time, lunar <span class="hlt">vertical</span> migration (LVM) may facilitate monthly pulses of carbon remineralization, as they occur continuously in illuminated mesopelagic systems [5], due to community respiration of carnivorous and detritivorous zooplankton. The extent of LVM during the winter suggests that the behavior is highly conserved and adaptive and therefore needs to be considered as "baseline" zooplankton activity in a changing Arctic ocean [6-9]. VIDEO ABSTRACT. Copyright © 2016 The Authors. Published by Elsevier Ltd.. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28783473','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28783473"><span>Heel-<span class="hlt">Rise</span> Height Deficit <span class="hlt">1</span> Year After Achilles Tendon Rupture Relates to Changes in Ankle Biomechanics 6 Years After Injury.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Brorsson, Annelie; Willy, Richard W; Tranberg, Roy; Grävare Silbernagel, Karin</p> <p>2017-11-01</p> <p>It is unknown whether the height of a heel-<span class="hlt">rise</span> performed in the single-leg standing heel-<span class="hlt">rise</span> test <span class="hlt">1</span> year after an Achilles tendon rupture (ATR) correlates with ankle biomechanics during walking, jogging, and jumping in the long-term. To explore the differences in ankle biomechanics, tendon length, calf muscle recovery, and patient-reported outcomes at a mean of 6 years after ATR between 2 groups that, at <span class="hlt">1</span>-year follow-up, had less than 15% versus greater than 30% differences in heel-<span class="hlt">rise</span> height. Cohort study; Level of evidence, 3. Seventeen patients with less than 15% (<15% group) and 17 patients with greater than 30% (>30% group) side-to-side difference in heel-<span class="hlt">rise</span> height at <span class="hlt">1</span> year after ATR were evaluated at a mean (SD) 6.<span class="hlt">1</span> (2.0) years after their ATR. Ankle kinematics and kinetics were sampled via standard motion capture procedures during walking, jogging, and jumping. Patient-reported outcome was evaluated with Achilles tendon Total Rupture Score (ATRS), Physical Activity Scale (PAS), and Foot and Ankle Outcome Score (FAOS). Tendon length was evaluated by ultrasonography. The Limb Symmetry Index (LSI = [Injured Side ÷ Healthy Side] × 100) was calculated for side differences. The >30% group had significantly more deficits in ankle kinetics during all activities compared with patients in the <15% group at a mean of 6 years after ATR (LSI, 70%-149% and 84%-106%, respectively; P = .010-.024). The >30% group, compared with the <15% group, also had significantly lower values in heel-<span class="hlt">rise</span> height (LSI, 72% and 95%, respectively; P < .001) and heel-<span class="hlt">rise</span> work (LSI, 58% and 91%, respectively; P < .001) and significantly larger side-to-side difference in tendon length (114% and 106%, respectively; P = .012). Achilles tendon length correlated with ankle kinematic variables ( r = 0.38-0.44; P = .015-.027) whereas heel-<span class="hlt">rise</span> work correlated with kinetic variables ( r = -0.57 to 0.56; P = .001-.047). LSI tendon length correlated negatively with LSI heel-<span class="hlt">rise</span> height ( r</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ISPAr42W7..543S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ISPAr42W7..543S"><span>Elevation-Based Sea-Level <span class="hlt">Rise</span> Vulnerability Assessment of Mindanao, Philippines: are Freely-Available 30-M Dems Good Enough?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Santillan, J. R.; Makinano-Santillan, M.</p> <p>2017-09-01</p> <p>We assessed the <span class="hlt">vertical</span> accuracies and uncertainties of three freely-available global DEMs as inputs to elevation-based sea-level <span class="hlt">rise</span> vulnerability assessment of Mindanao, Philippines - an area where above average SLR of 14.7 mm/year was recently found. These DEMs are the Shuttle Radar Topography Mission (SRTM) DEM, ASTER Global DEM (GDEM Version 2), and ALOS World 3D-30 (AW3D30). Using 2,076 ground control points, we computed each DEM's <span class="hlt">vertical</span> accuracies and uncertainties, and from these we determined the smallest increment of sea-level <span class="hlt">rise</span> (SLRImin) that should be considered when using the DEMs for SLR impact assessment, as well as the Minimum Planning Timeline (TLmin) for an elevation-based SLR assessment. Results of <span class="hlt">vertical</span> accuracy assessment revealed Root Mean Square Errors of 9.80 m for ASTER GDEM V2, 5.16 m for SRTM DEM, and 4.32 m for AW3D30. <span class="hlt">Vertical</span> uncertainties in terms of the Linear Error at 95 % Confidence (LE95) were found to be as follows: 19.21 m for ASTER GDEM V2, 10.12 m for SRTM DEM, and 8.47 m for AW3D30. From these, we found that ASTER GDEM2 is suitable to model SLR increments of at least 38.41 m and it will take 2,613 years for the cumulative water level increase of 14.7 mm/year to reach the minimum SLR increment afforded by this DEM. For the SRTM DEM, SLRImin and TLmin were computed as 20.24 m and <span class="hlt">1</span>,377 years, respectively. For the AW3D30, SLRImin and TLmin were computed as 16.92 m and <span class="hlt">1</span>,151 years, respectively. These results suggest that the readily available global DEMs' suitability for mapping coastal inundations due to SLR in our study area is limited by their low <span class="hlt">vertical</span> accuracies and high uncertainties. All the three DEMs do not have the necessary accuracy and minimum uncertainties that will make them suitable for mapping inundations of Mindanao at smaller increments of SLR (e.g., SLR ≤ 5 m). Hence, users who apply any of these DEMs</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-KSC-314D-0449_031.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-KSC-314D-0449_031.html"><span>EFT-<span class="hlt">1</span> Delta IV Heavy lift to <span class="hlt">vertical</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2014-10-01</p> <p>The United Launch Alliance Delta IV Heavy rocket for Exploration Flight Test-<span class="hlt">1</span> is being lifted to the <span class="hlt">vertical</span> position in the mobile service tower on the pad at the pad at Space Launch Complex 37 at Cape Canaveral Air Force Station in Florida. The Delta IV Heavy is being readied to launch Orion on its first flight test. During its first flight test, Orion will travel farther into space than any human spacecraft has gone in more than 40 years. The data gathered during the flight will influence design decisions, validate existing computer models and innovative new approaches to space systems development, as well as reduce overall mission risks and costs for later Orion flights. Liftoff of Orion on the first flight test is planned for December 2014.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-KSC-314D-0449_108.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-KSC-314D-0449_108.html"><span>EFT-<span class="hlt">1</span> Delta IV Heavy lift to <span class="hlt">vertical</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2014-10-01</p> <p>United Launch Alliance, or ULA, workers monitor the progress as the ULA Delta IV Heavy rocket for Exploration Flight Test-<span class="hlt">1</span> is lifted to the <span class="hlt">vertical</span> position in the mobile service tower on the pad at Space Launch Complex 37 at Cape Canaveral Air Force Station in Florida. The Delta IV Heavy is being readied to launch Orion on its first flight test. During its first flight test, Orion will travel farther into space than any human spacecraft has gone in more than 40 years. The data gathered during the flight will influence design decisions, validate existing computer models and innovative new approaches to space systems development, as well as reduce overall mission risks and costs for later Orion flights. Liftoff of Orion on the first flight test is planned for December 2014.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70197720','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70197720"><span>U.S. Pacific coastal wetland resilience and vulnerability to sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Thorne, Karen M.; MacDonald, Glen M.; Guntenspergen, Glenn R.; Ambrose, Richard F.; Buffington, Kevin J.; Dugger, Bruce D.; Freeman, Chase; Janousek, Christopher; Brown, Lauren N.; Rosencranz, Jordan A.; Homquist, James; Smol, John P.; Hargan, Kathryn; Takekawa, John Y.</p> <p>2018-01-01</p> <p>We used a first-of-its-kind comprehensive scenario approach to evaluate both the <span class="hlt">vertical</span> and horizontal response of tidal wetlands to projected changes in the rate of sea-level <span class="hlt">rise</span> (SLR) across 14 estuaries along the Pacific coast of the continental United States. Throughout the U.S. Pacific region, we found that tidal wetlands are highly vulnerable to end-of-century submergence, with resulting extensive loss of habitat. Using higher-range SLR scenarios, all high and middle marsh habitats were lost, with 83% of current tidal wetlands transitioning to unvegetated habitats by 2110. The wetland area lost was greater in California and Oregon (100%) but still severe in Washington, with 68% submerged by the end of the century. The only wetland habitat remaining at the end of the century was low marsh under higher-range SLR rates. Tidal wetland loss was also likely under more conservative SLR scenarios, including loss of 95% of high marsh and 60% of middle marsh habitats by the end of the century. Horizontal migration of most wetlands was constrained by coastal development or steep topography, with just two wetland sites having sufficient upland space for migration and the possibility for nearly <span class="hlt">1:1</span> replacement, making SLR threats particularly high in this region and generally undocumented. With low <span class="hlt">vertical</span> accretion rates and little upland migration space, Pacific coast tidal wetlands are at imminent risk of submergence with projected rates of rapid SLR.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5834000','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5834000"><span>U.S. Pacific coastal wetland resilience and vulnerability to sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Thorne, Karen; MacDonald, Glen; Guntenspergen, Glenn; Ambrose, Richard; Buffington, Kevin; Dugger, Bruce; Freeman, Chase; Janousek, Christopher; Brown, Lauren; Rosencranz, Jordan; Holmquist, James; Smol, John; Hargan, Kathryn; Takekawa, John</p> <p>2018-01-01</p> <p>We used a first-of-its-kind comprehensive scenario approach to evaluate both the <span class="hlt">vertical</span> and horizontal response of tidal wetlands to projected changes in the rate of sea-level <span class="hlt">rise</span> (SLR) across 14 estuaries along the Pacific coast of the continental United States. Throughout the U.S. Pacific region, we found that tidal wetlands are highly vulnerable to end-of-century submergence, with resulting extensive loss of habitat. Using higher-range SLR scenarios, all high and middle marsh habitats were lost, with 83% of current tidal wetlands transitioning to unvegetated habitats by 2110. The wetland area lost was greater in California and Oregon (100%) but still severe in Washington, with 68% submerged by the end of the century. The only wetland habitat remaining at the end of the century was low marsh under higher-range SLR rates. Tidal wetland loss was also likely under more conservative SLR scenarios, including loss of 95% of high marsh and 60% of middle marsh habitats by the end of the century. Horizontal migration of most wetlands was constrained by coastal development or steep topography, with just two wetland sites having sufficient upland space for migration and the possibility for nearly <span class="hlt">1:1</span> replacement, making SLR threats particularly high in this region and generally undocumented. With low <span class="hlt">vertical</span> accretion rates and little upland migration space, Pacific coast tidal wetlands are at imminent risk of submergence with projected rates of rapid SLR. PMID:29507876</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhPl...25e2107A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhPl...25e2107A"><span><span class="hlt">Vertical</span> sizes of <span class="hlt">1</span>-D and 2-D electrostatic solitons with nonextensive and trapped electrons in the upper ionosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ali Shan, Shaukat; Saleem, Hamid</p> <p>2018-05-01</p> <p>The <span class="hlt">vertical</span> sizes of one-dimensional (<span class="hlt">1</span>-D) and two dimensional (2-D) electrostatic solitons are estimated in the oxygen-hydrogen (O - H) and pure oxygen plasmas of the upper ionosphere taking into account the effects of non-extensive and trapped electrons. The field-aligned flow of oxygen ions is also considered. It is found that both electron trapping and non-extensivity play a constructive role in the formation of <span class="hlt">1</span>-D and 2-D solitary structures. The <span class="hlt">vertical</span> size of the solitons is not known through observations, but here it is pointed out that the <span class="hlt">vertical</span> size of these structures should be of the order of a few meters at the altitude of 800 km in the <span class="hlt">1</span>-D case. On the other hand, in the 2-D case, the <span class="hlt">vertical</span> size is much larger than the horizontal size and it turns out to be of the order of a few kilometers, while the width is about a few hundred meters in agreement with the observations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018VSD....56..529B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018VSD....56..529B"><span>Distributed support modelling for <span class="hlt">vertical</span> track dynamic analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Blanco, B.; Alonso, A.; Kari, L.; Gil-Negrete, N.; Giménez, J. G.</p> <p>2018-04-01</p> <p>The finite length nature of rail-pad supports is characterised by a Timoshenko beam element formulation over an elastic foundation, giving <span class="hlt">rise</span> to the distributed support element. The new element is integrated into a <span class="hlt">vertical</span> track model, which is solved in frequency and time domain. The developed formulation is obtained by solving the governing equations of a Timoshenko beam for this particular case. The interaction between sleeper and rail via the elastic connection is considered in an analytical, compact and efficient way. The modelling technique results in realistic amplitudes of the 'pinned-pinned' vibration mode and, additionally, it leads to a smooth evolution of the contact force temporal response and to reduced amplitudes of the rail <span class="hlt">vertical</span> oscillation, as compared to the results from concentrated support models. Simulations are performed for both parametric and sinusoidal roughness excitation. The model of support proposed here is compared with a previous finite length model developed by other authors, coming to the conclusion that the proposed model gives accurate results at a reduced computational cost.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.G21A0862A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.G21A0862A"><span>Using GNSS for Assessment Recent Sea Level <span class="hlt">Rise</span> in the Northwestern Part of the Arabian Gulf</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Alothman, A. O.; Bos, M. S.; Fernandes, R.</p> <p>2017-12-01</p> <p>Due to the global warming acting recently (in the 21st century) on the planet Earth, an associated sea level <span class="hlt">rise</span> is predicted to reach up to 30 cm to 60 cm in some regions. Sea level monitoring is important for the Kingdom of Saudi Arabia, since it is surrounded by very long cost of about 3400 km in length and hundreds of isolated islands. The eastern coast line of KSA, in the Arabian Gulf, needs some monitoring in the long term, due to low land nature of the region. Also, the ongoing oil withdrawal activities in the area, may affect the regional sea level <span class="hlt">rise</span>. In addition to these two facts, the tectonic structure of the Arabian Peninsula is one factor. The Regional Relative sea level in the eastern cost of Saudi Arabia has been estimated in the past using tide gauge data of more than 28 years using the <span class="hlt">vertical</span> displacement of permanent Global Navigation Satellite System GNSS stations having time span of only about 3 years. In this paper, we discuss and update the methodology and results from Alothman et al. (2014), particularly by checking and extending the GNSS solutions. Since 3 of the 6 GPS stations used only started observing in the end of 2011, the longer time series have now significantly lower uncertainties in the estimated <span class="hlt">vertical</span> rate. Longer time span of GNSS observations were included and 500 synthetic time series were estimated and seasonal signals were analysed. it is concluded that the varying seasonal signal present in the GNSS time series causes an underestimation of 0.<span class="hlt">1</span> mm/yr for short time series of 3 years. In addition to the implications of using short time series to estimate the <span class="hlt">vertical</span> land motion, we found that if the varying seasonal signals are present in the data, the problem is aggravated. This finding can be useful for other studies analyzing short GNSS time series.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11252043','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11252043"><span>Strategic interaction among hospitals and nursing facilities: the efficiency effects of payment systems and <span class="hlt">vertical</span> integration.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Banks, D; Parker, E; Wendel, J</p> <p>2001-03-01</p> <p><span class="hlt">Rising</span> post-acute care expenditures for Medicare transfer patients and increasing <span class="hlt">vertical</span> integration between hospitals and nursing facilities raise questions about the links between payment system structure, the incentive for <span class="hlt">vertical</span> integration and the impact on efficiency. In the United States, policy-makers are responding to these concerns by initiating prospective payments to nursing facilities, and are exploring the bundling of payments to hospitals. This paper develops a static profit-maximization model of the strategic interaction between the transferring hospital and a receiving nursing facility. This model suggests that the post-1984 system of prospective payment for hospital care, coupled with nursing facility payments that reimburse for services performed, induces inefficient under-provision of hospital services and encourages <span class="hlt">vertical</span> integration. It further indicates that the extension of prospective payment to nursing facilities will not eliminate the incentive to <span class="hlt">vertically</span> integrate, and will not result in efficient production unless such integration takes place. Bundling prospective payments for hospitals and nursing facilities will neither remove the incentive for <span class="hlt">vertical</span> integration nor induce production efficiency without such <span class="hlt">vertical</span> integration. However, bundled payment will induce efficient production, with or without <span class="hlt">vertical</span> integration, if nursing facilities are reimbursed for services performed. Copyright 2001 John Wiley & Sons, Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013GeoJI.194..719B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013GeoJI.194..719B"><span>Characterizing and minimizing the effects of noise in tide gauge time series: relative and geocentric sea level <span class="hlt">rise</span> around Australia</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Burgette, Reed J.; Watson, Christopher S.; Church, John A.; White, Neil J.; Tregoning, Paul; Coleman, Richard</p> <p>2013-08-01</p> <p>.8 mm yr-<span class="hlt">1</span>, respectively. While the temporal pattern of the rate estimates is consistent with acceleration in sea level <span class="hlt">rise</span>, it may not be significant, as the uncertainties for the shorter analysis periods may not capture the full range of temporal variation. Analysis of the available continuous GPS records that have been collected within 80 km of Australian tide gauges suggests that rates of <span class="hlt">vertical</span> crustal motion are generally low, with the majority of sites showing motion statistically insignificant from zero. A notable exception is the significant component of <span class="hlt">vertical</span> land motion that contributes to the rapid rate of relative sea level change (>4 mm yr-<span class="hlt">1</span>) at the Hillarys site in the Perth area. This corresponds to crustal subsidence that we estimate in our GPS analysis at a rate of -3.<span class="hlt">1</span> ± 0.7 mm yr-<span class="hlt">1</span>, and appears linked to groundwater withdrawal. Uncertainties on the rates of <span class="hlt">vertical</span> displacement at GPS sites collected over a decade are similar to what we measure in several decades of tide gauge data. Our results motivate continued observations of relative sea level using tide gauges, maintained with high-accuracy terrestrial and continuous co-located satellite-based surveying.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70197429','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70197429"><span>Coastal wetland adaptation to sea level <span class="hlt">rise</span>: Quantifying potential for landward migration and coastal squeeze</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Borchert, Sinéad M.; Osland, Michael J.; Enwright, Nicholas M.; Griffith, Kereen</p> <p>2018-01-01</p> <p>Coastal wetland ecosystems are expected to migrate landwards in response to <span class="hlt">rising</span> seas. However, due to differences in topography and coastal urbanization, estuaries vary in their ability to accommodate migration. Low‐lying urban areas can constrain migration and lead to wetland loss (i.e. coastal squeeze), especially where existing wetlands cannot keep pace with <span class="hlt">rising</span> seas via <span class="hlt">vertical</span> adjustments. In many estuaries, there is a pressing need to identify landward migration corridors and better quantify the potential for landward migration and coastal squeeze.We quantified and compared the area available for landward migration of tidal saline wetlands and the area where urban development is expected to prevent migration for 39 estuaries along the wetland‐rich USA Gulf of Mexico coast. We did so under three sea level <span class="hlt">rise</span> scenarios (0.5, <span class="hlt">1</span>.0, and <span class="hlt">1</span>.5 m by 2100).Within the region, the potential for wetland migration is highest within certain estuaries in Louisiana and southern Florida (e.g. Atchafalaya/Vermilion Bays, Mermentau River, Barataria Bay, and the North and South Ten Thousand Islands estuaries).The potential for coastal squeeze is highest in estuaries containing major metropolitan areas that extend into low‐lying lands. The Charlotte Harbor, Tampa Bay, and Crystal‐Pithlachascotee estuaries (Florida) have the highest amounts of urban land expected to constrain wetland migration. Urban barriers to migration are also high in the Galveston Bay (Texas) and Atchafalaya/Vermilion Bays (Louisiana) estuaries.Synthesis and applications. Coastal wetlands provide many ecosystem services that benefit human health and well‐being, including shoreline protection and fish and wildlife habitat. As the rate of sea level <span class="hlt">rise</span> accelerates in response to climate change, coastal wetland resources could be lost in areas that lack space for landward migration. Migration corridors are particularly important in highly urbanized estuaries where, due to low‐lying coastal</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.A23C0259H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.A23C0259H"><span><span class="hlt">Vertical</span> velocity-CCN correlations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hudson, J. G.; Noble, S. R.</p> <p>2013-12-01</p> <p>The realization that smaller cloud droplets evaporate more readily (Xue and Feingold 2006; Jiang et al. 2002) gives <span class="hlt">rise</span> to an anti-indirect aerosol effect (IAE); less cloudiness with pollution. The greater latent heat exchange of the greater evaporation in more polluted clouds adds TKE and buoyancy gradients that can enhance <span class="hlt">vertical</span> velocity (W), mixing and entrainment (Zhao and Austin 2005). Stronger W can increase horizontal motions, which can further enhance droplet evaporation, which further enhances latent heat exchange and <span class="hlt">vertical</span> motions, thus, positive feedback. This could also include latent heat released during condensation (Lee and Feingold 2010), which is more rapid for the greater surface areas of the smaller more numerous droplets. These theories imply a positive relationship between within-cloud W variations; i.e., standard deviation of W (σw) and CCN concentration (NCCN) rather than W and NCCN. This implies greater turbulence in polluted clouds, which could possibly counteract the reduction of cloudiness of anti-IAE. During two stratus cloud projects, 50 cloud penetrations in 9 MASE flights and 34 cloud penetrations in 13 POST flights, within-cloud σw-NCCN showed correlation coefficients (R) of 0.50 and 0.39. Panel a shows similar within-cloud σw-NCCN R in all altitude bands for 17 RICO flights in small cumulus clouds. R for W-NCCN showed similar values but only at low altitudes. Out-of-cloud σw-NCCN showed similar high values except at the highest altitudes. Within-cloud σw showed higher R than within-cloud W with droplet concentrations (Nc), especially at higher altitudes. Panel b for 13 ICE-T cumulus cloud flights in the same location as RICO but during the opposite season, however, showed σw and W uncorrelated with NCCN at all altitudes; and W and σw correlated with Nc only at the highest altitudes. On the other hand, out-of-cloud σw was correlated with NCCN at all altitudes with R similar to the corresponding R of the other projects</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015GMDD....8.3079G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015GMDD....8.3079G"><span>OESbathy version <span class="hlt">1</span>.0: a method for reconstructing ocean bathymetry with realistic continental shelf-slope-<span class="hlt">rise</span> structures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Goswami, A.; Olson, P. L.; Hinnov, L. A.; Gnanadesikan, A.</p> <p>2015-04-01</p> <p>We present a method for reconstructing global ocean bathymetry that uses a plate cooling model for the oceanic lithosphere, the age distribution of the oceanic crust, global oceanic sediment thicknesses, plus shelf-slope-<span class="hlt">rise</span> structures calibrated at modern active and passive continental margins. Our motivation is to reconstruct realistic ocean bathymetry based on parameterized relationships of present-day variables that can be applied to global oceans in the geologic past, and to isolate locations where anomalous processes such as mantle convection may affect bathymetry. Parameters of the plate cooling model are combined with ocean crustal age to calculate depth-to-basement. To the depth-to-basement we add an isostatically adjusted, multicomponent sediment layer, constrained by sediment thickness in the modern oceans and marginal seas. A continental shelf-slope-<span class="hlt">rise</span> structure completes the bathymetry reconstruction, extending from the ocean crust to the coastlines. Shelf-slope-<span class="hlt">rise</span> structures at active and passive margins are parameterized using modern ocean bathymetry at locations where a complete history of seafloor spreading is preserved. This includes the coastal regions of the North, South, and Central Atlantic Ocean, the Southern Ocean between Australia and Antarctica, and the Pacific Ocean off the west coast of South America. The final products are global maps at 0.<span class="hlt">1</span>° × 0.<span class="hlt">1</span>° resolution of depth-to-basement, ocean bathymetry with an isostatically adjusted, multicomponent sediment layer, and ocean bathymetry with reconstructed continental shelf-slope-<span class="hlt">rise</span> structures. Our reconstructed bathymetry agrees with the measured ETOPO<span class="hlt">1</span> bathymetry at most passive margins, including the east coast of North America, north coast of the Arabian Sea, and northeast and southeast coasts of South America. There is disagreement at margins with anomalous continental shelf-slope-<span class="hlt">rise</span> structures, such as around the Arctic Ocean, the Falkland Islands, and Indonesia.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015GMD.....8.2735G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015GMD.....8.2735G"><span>OESbathy version <span class="hlt">1</span>.0: a method for reconstructing ocean bathymetry with generalized continental shelf-slope-<span class="hlt">rise</span> structures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Goswami, A.; Olson, P. L.; Hinnov, L. A.; Gnanadesikan, A.</p> <p>2015-09-01</p> <p>We present a method for reconstructing global ocean bathymetry that combines a standard plate cooling model for the oceanic lithosphere based on the age of the oceanic crust, global oceanic sediment thicknesses, plus generalized shelf-slope-<span class="hlt">rise</span> structures calibrated at modern active and passive continental margins. Our motivation is to develop a methodology for reconstructing ocean bathymetry in the geologic past that includes heterogeneous continental margins in addition to abyssal ocean floor. First, the plate cooling model is applied to maps of ocean crustal age to calculate depth to basement. To the depth to basement we add an isostatically adjusted, multicomponent sediment layer constrained by sediment thickness in the modern oceans and marginal seas. A three-parameter continental shelf-slope-<span class="hlt">rise</span> structure completes the bathymetry reconstruction, extending from the ocean crust to the coastlines. Parameters of the shelf-slope-<span class="hlt">rise</span> structures at active and passive margins are determined from modern ocean bathymetry at locations where a complete history of seafloor spreading is preserved. This includes the coastal regions of the North, South, and central Atlantic, the Southern Ocean between Australia and Antarctica, and the Pacific Ocean off the west coast of South America. The final products are global maps at 0.<span class="hlt">1</span>° × 0.<span class="hlt">1</span>° resolution of depth to basement, ocean bathymetry with an isostatically adjusted multicomponent sediment layer, and ocean bathymetry with reconstructed continental shelf-slope-<span class="hlt">rise</span> structures. Our reconstructed bathymetry agrees with the measured ETOPO<span class="hlt">1</span> bathymetry at most passive margins, including the east coast of North America, north coast of the Arabian Sea, and northeast and southeast coasts of South America. There is disagreement at margins with anomalous continental shelf-slope-<span class="hlt">rise</span> structures, such as around the Arctic Ocean, the Falkland Islands, and Indonesia.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21844609','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21844609"><span>Relative net <span class="hlt">vertical</span> impulse determines jumping performance.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kirby, Tyler J; McBride, Jeffrey M; Haines, Tracie L; Dayne, Andrea M</p> <p>2011-08-01</p> <p>The purpose of this investigation was to determine the relationship between relative net <span class="hlt">vertical</span> impulse and jump height in a countermovement jump and static jump performed to varying squat depths. Ten college-aged males with 2 years of jumping experience participated in this investigation (age: 23.3 ± <span class="hlt">1</span>.5 years; height: 176.7 ± 4.5 cm; body mass: 84.4 ± 10.<span class="hlt">1</span> kg). Subjects performed a series of static jumps and countermovement jumps in a randomized fashion to a depth of 0.15, 0.30, 0.45, 0.60, and 0.75 m and a self-selected depth (static jump depth = 0.38 ± 0.08 m, countermovement jump depth = 0.49 ± 0.06 m). During the concentric phase of each jump, peak force, peak velocity, peak power, jump height, and net <span class="hlt">vertical</span> impulse were recorded and analyzed. Net <span class="hlt">vertical</span> impulse was divided by body mass to produce relative net <span class="hlt">vertical</span> impulse. Increasing squat depth corresponded to a decrease in peak force and an increase in jump height and relative net <span class="hlt">vertical</span> impulse for both static jump and countermovement jump. Across all depths, relative net <span class="hlt">vertical</span> impulse was statistically significantly correlated to jump height in the static jump (r = .9337, p < .0001, power = <span class="hlt">1</span>.000) and countermovement jump (r = .925, p < .0001, power = <span class="hlt">1</span>.000). Across all depths, peak force was negatively correlated to jump height in the static jump (r = -0.3947, p = .0018, power = 0.8831) and countermovement jump (r = -0.4080, p = .0012, power = 0.9050). These results indicate that relative net <span class="hlt">vertical</span> impulse can be used to assess <span class="hlt">vertical</span> jump performance, regardless of initial squat depth, and that peak force may not be the best measure to assess <span class="hlt">vertical</span> jump performance.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_7 --> <div id="page_8" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="141"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3282823','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3282823"><span><span class="hlt">Rising</span> Incidence of Type <span class="hlt">1</span> Diabetes Is Associated With Altered Immunophenotype at Diagnosis</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Long, Anna E.; Gillespie, Kathleen M.; Rokni, Saba; Bingley, Polly J.; Williams, Alistair J.K.</p> <p>2012-01-01</p> <p>The incidence of type <span class="hlt">1</span> diabetes has increased rapidly over recent decades, particularly in young children. We aimed to determine whether this <span class="hlt">rise</span> was associated with changes in patterns of humoral islet autoimmunity at diagnosis. Autoantibodies to insulin (IAA), GAD (GADA), islet antigen-2 (IA-2A), and zinc transporter 8 (ZnT8A) were measured by radioimmunoassay in sera collected from children and young adults with newly diagnosed type <span class="hlt">1</span> diabetes between 1985 and 2002. The influence of date of diagnosis on prevalence and level of autoantibodies was investigated by logistic regression with adjustment for age and HLA class II genetic risk. Prevalence of IA-2A and ZnT8A increased significantly over the period studied, and this was mirrored by raised levels of IA-2A, ZnT8A, and IA-2β autoantibodies (IA-2βA). IAA and GADA prevalence and levels did not change. Increases in IA-2A, ZnT8A, and IA-2βA at diagnosis during a period of <span class="hlt">rising</span> incidence suggest that the process leading to type <span class="hlt">1</span> diabetes is now characterized by a more intense humoral autoimmune response. Understanding how changes in environment or lifestyle alter the humoral autoimmune response to islet antigens should help explain why the incidence of type <span class="hlt">1</span> diabetes is increasing and may suggest new strategies for preventing disease. PMID:22315309</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMEP23C..06C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMEP23C..06C"><span>Modeling storm and sea level <span class="hlt">rise</span> impacts on marsh transgression</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Carr, J. A.; Guntenspergen, G. R.; Kirwan, M. L.</p> <p>2016-12-01</p> <p>Coastal salt marsh systems provide critical ecosystem services, including key habitat and coastal protection. Both lateral extent, and <span class="hlt">vertical</span> stability of salt marshes to sea level <span class="hlt">rise</span> have been shown to be functions of both biotic, and abiotic drivers and feedbacks. As a result, the ecogeomorphic evolution of the system can exhibit strong non-linearities, discontinuities and thresholds. We developed a two-dimensional transect model to explore controls on marsh lateral extent, <span class="hlt">vertical</span> stability and the potential for marsh transgression inland and upland. Salt marsh and upland regions in the model are discretized in <span class="hlt">1</span> m increments with inundation frequency determined by the elevation of the individual cells, organogenic soil formation and mineral deposition rates, and the history of stochastic water levels. The transect extends from an idealized back barrier bay across the salt marsh platform and into the upland forest and is forced with auto and cross correlated synthetic stochastic wind speed, wind direction and water levels. The model incorporates key feedbacks between fetch, wave growth and subsequent lateral erosion rates and sediment supply to the marsh platform. Deposition of mineral sediment from the bay and/or internal ponds onto the marsh platform cells is dependent both on the inundation frequency and distance from a marsh edge. For each element along the transect, a Markov chain successional model was implemented that considers six distinct states, grass/saltmarsh, seedling, sapling, tree, dead standing tree, and bare. A non-static transition probability matrix, dependent on both inundation of the element and the prior vegetation state, was used in order to allow for feedbacks, both positive and negative, among different vegetation states and environmental drivers. The model was used to examine the qualitative behavior of the coupled systems under varied rates of sea level <span class="hlt">rise</span>, external sediment supply, wind and storm statistics, tidal range, upland</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA210505','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA210505"><span>Consumer Electronics Testing to Fast-<span class="hlt">Rise</span> EMP (Electromagnetic Pulse) (VEMPS (<span class="hlt">Vertical</span> Electromagnetic Pulse Simulator) 2 Development)</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1989-06-01</p> <p>Fast-<span class="hlt">Rise</span> EMP ( VEMPS 11 Development) 12. PERSONAL AUTHOR(S) Vincent J. Ellis 13a. TYPE OF REPORT 13b. TIME COVERED 14. DATE OF REPORT (Year Mot.Dy...SUBJECT TERMS (Continue on reverse if necessary andJ identity by block nwitib") 09EL 03U SBGR EMP, VEMPS 11, consumer electronics, FEMPS, EMP simulation... VEMPS 11), Because of the unique petrr! ance characteristics of VEMPS 11 and the technological changes In consumer electronics over the past 10 years</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28894195','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28894195"><span>Nuisance Flooding and Relative Sea-Level <span class="hlt">Rise</span>: the Importance of Present-Day Land Motion.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Karegar, Makan A; Dixon, Timothy H; Malservisi, Rocco; Kusche, Jürgen; Engelhart, Simon E</p> <p>2017-09-11</p> <p>Sea-level <span class="hlt">rise</span> is beginning to cause increased inundation of many low-lying coastal areas. While most of Earth's coastal areas are at risk, areas that will be affected first are characterized by several additional factors. These include regional oceanographic and meteorological effects and/or land subsidence that cause relative sea level to <span class="hlt">rise</span> faster than the global average. For catastrophic coastal flooding, when wind-driven storm surge inundates large areas, the relative contribution of sea-level <span class="hlt">rise</span> to the frequency of these events is difficult to evaluate. For small scale "nuisance flooding," often associated with high tides, recent increases in frequency are more clearly linked to sea-level <span class="hlt">rise</span> and global warming. While both types of flooding are likely to increase in the future, only nuisance flooding is an early indicator of areas that will eventually experience increased catastrophic flooding and land loss. Here we assess the frequency and location of nuisance flooding along the eastern seaboard of North America. We show that <span class="hlt">vertical</span> land motion induced by recent anthropogenic activity and glacial isostatic adjustment are contributing factors for increased nuisance flooding. Our results have implications for flood susceptibility, forecasting and mitigation, including management of groundwater extraction from coastal aquifers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://pubs.usgs.gov/of/2014/1252/pdf/ofr2014-1252.pdf','USGSPUBS'); return false;" href="http://pubs.usgs.gov/of/2014/1252/pdf/ofr2014-1252.pdf"><span>Evaluating coastal landscape response to sea-level <span class="hlt">rise</span> in the northeastern United States: approach and methods</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Lentz, Erika E.; Stippa, Sawyer R.; Thieler, E. Robert; Plant, Nathaniel G.; Gesch, Dean B.; Horton, Radley M.</p> <p>2014-02-13</p> <p>The U.S. Geological Survey is examining effects of future sea-level <span class="hlt">rise</span> on the coastal landscape from Maine to Virginia by producing spatially explicit, probabilistic predictions using sea-level projections, <span class="hlt">vertical</span> land movement rates (due to isostacy), elevation data, and land-cover data. Sea-level-<span class="hlt">rise</span> scenarios used as model inputs are generated by using multiple sources of information, including Coupled Model Intercomparison Project Phase 5 models following representative concentration pathways 4.5 and 8.5 in the Intergovernmental Panel on Climate Change Fifth Assessment Report. A Bayesian network is used to develop a predictive coastal response model that integrates the sea-level, elevation, and land-cover data with assigned probabilities that account for interactions with coastal geomorphology as well as the corresponding ecological and societal systems it supports. The effects of sea-level <span class="hlt">rise</span> are presented as (<span class="hlt">1</span>) level of landscape submergence and (2) coastal response type characterized as either static (that is, inundation) or dynamic (that is, landform or landscape change). Results are produced at a spatial scale of 30 meters for four decades (the 2020s, 2030s, 2050s, and 2080s). The probabilistic predictions can be applied to landscape management decisions based on sea-level-<span class="hlt">rise</span> effects as well as on assessments of the prediction uncertainty and need for improved data or fundamental understanding. This report describes the methods used to produce predictions, including information on input datasets; the modeling approach; model outputs; data-quality-control procedures; and information on how to access the data and metadata online.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20160004969','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20160004969"><span>Evaluating Coastal Landscape Response to Sea-Level <span class="hlt">Rise</span> in the Northeastern United States - Approach and Methods</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lentz, Erika E.; Stippa, Sawyer R.; Thieler, E. Robert; Plant, Nathaniel G.; Gesch, Dean B.; Horton, Radley M.</p> <p>2015-01-01</p> <p>The U.S. Geological Survey is examining effects of future sea-level <span class="hlt">rise</span> on the coastal landscape from Maine to Virginia by producing spatially explicit, probabilistic predictions using sea-level projections, <span class="hlt">vertical</span> land movement rates (due to isostacy), elevation data, and land-cover data. Sea-level-<span class="hlt">rise</span> scenarios used as model inputs are generated by using multiple sources of information, including Coupled Model Intercomparison Project Phase 5 models following representative concentration pathways 4.5 and 8.5 in the Intergovernmental Panel on Climate Change Fifth Assessment Report. A Bayesian network is used to develop a predictive coastal response model that integrates the sea-level, elevation, and land-cover data with assigned probabilities that account for interactions with coastal geomorphology as well as the corresponding ecological and societal systems it supports. The effects of sea-level <span class="hlt">rise</span> are presented as (<span class="hlt">1</span>) level of landscape submergence and (2) coastal response type characterized as either static (that is, inundation) or dynamic (that is, landform or landscape change). Results are produced at a spatial scale of 30 meters for four decades (the 2020s, 2030s, 2050s, and 2080s). The probabilistic predictions can be applied to landscape management decisions based on sea-level-<span class="hlt">rise</span> effects as well as on assessments of the prediction uncertainty and need for improved data or fundamental understanding. This report describes the methods used to produce predictions, including information on input datasets; the modeling approach; model outputs; data-quality-control procedures; and information on how to access the data and metadata online.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PrOce.151...49S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PrOce.151...49S"><span>The diel <span class="hlt">vertical</span> migration patterns and individual swimming behavior of overwintering sprat Sprattus sprattus</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Solberg, Ingrid; Kaartvedt, Stein</p> <p>2017-02-01</p> <p>We addressed the behavioral patterns and DVM dynamics of sprat overwintering in a Norwegian fjord (150 m) with increasing hypoxia by depth. An upward-facing echosounder deployed at the bottom and cabled to shore provided 4 months of continuous acoustic data. This enabled detailed studies of individual behavior, specifically allowing assessment of individual <span class="hlt">vertical</span> migrations at dusk and dawn in relation to light, analysis of so-called <span class="hlt">rise</span>-and-sink swimming, and investigation of the sprat' swimming activity and behavior in severely hypoxic waters. Field campaigns supplemented the acoustic studies. The acoustic records showed that the main habitat for sprat was the upper ∼65 m where oxygen concentrations were ⩾0.7 mL O2 L-<span class="hlt">1</span>. The sprat schooled at ∼50 m during daytime and initiated an upward migration about <span class="hlt">1</span> h prior to sunset. While some sprat migrated to surface waters, other individuals interrupted the ascent when at ∼20-30 m, and returned to deeper waters ∼20-50 min after sunset. Sprat at depth was on average larger, yet individuals made excursions to- and from upper layers. Sprat were swimming in a "<span class="hlt">rise</span> and sink" pattern at depth, likely related to negative buoyancy. Short-term dives into waters with less than 0.45 mL O2 L-<span class="hlt">1</span> were interpreted as feeding forays for abundant overwintering Calanus spp. The deep group of sprat initiated a dawn ascent less than <span class="hlt">1</span> h before sunrise, ending at 20-30 m where they formed schools. They subsequently returned to deeper waters about ∼20 min prior to sunrise. Measurements of surface light intensities indicated that the sprat experienced lower light levels in upper waters at dawn than at dusk. The <span class="hlt">vertical</span> swimming speed varied significantly between the behavioral tasks. The mixed DVM patterns and dynamic nocturnal behavior of sprat persisted throughout winter, likely shaped by individual strategies involving optimized feeding and predator avoidance, as well as relating to temperature, hypoxia and negative</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5384722','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5384722"><span>Survival of HIV-<span class="hlt">1</span> <span class="hlt">vertically</span> infected children</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Davies, Mary-Ann; Gibb, Diana; Turkova, Anna</p> <p>2017-01-01</p> <p>Purpose of review It is 20 years since the start of the combination antiretroviral therapy (cART) era and >10 years since cART scale-up began in resource-limited settings. We examined survival of <span class="hlt">vertically</span> HIV-infected infants and children in the cART era. Recent findings Good survival has been achieved on cART in all settings with up to ten-fold mortality reductions compared to before cART availability. Although mortality risk remains high in the first few months after cART initiation in young children with severe disease, it drops rapidly thereafter even for those who started with advanced disease, and longer term mortality risk is low. However, suboptimal retention on cART in routine programs threatens good survival outcomes and even on treatment children continue to experience high comorbidity risk; infections remain the major cause of death. Interventions to address infection risk include co-trimoxazole prophylaxis, isoniazid preventive therapy, routine childhood and influenza immunization and improving maternal survival. Summary Pediatric survival has improved substantially with cART and HIV-infected children are aging into adulthood. It is important to ensure access to diagnosis and early cART, good program retention and optimal co-morbidity prophylaxis and treatment to achieve the best possible long-term survival and health outcomes for <span class="hlt">vertically</span> infected children. PMID:27716730</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhPl...25e6106P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhPl...25e6106P"><span>Modelling of NSTX hot <span class="hlt">vertical</span> displacement events using M 3 D -C <span class="hlt">1</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pfefferlé, D.; Ferraro, N.; Jardin, S. C.; Krebs, I.; Bhattacharjee, A.</p> <p>2018-05-01</p> <p>The main results of an intense <span class="hlt">vertical</span> displacement event (VDE) modelling activity using the implicit 3D extended MHD code M3D-C<span class="hlt">1</span> are presented. A pair of nonlinear 3D simulations are performed using realistic transport coefficients based on the reconstruction of a so-called NSTX frozen VDE where the feedback control was purposely switched off to trigger a <span class="hlt">vertical</span> instability. The <span class="hlt">vertical</span> drift phase is solved assuming axisymmetry until the plasma contacts the first wall, at which point the intricate evolution of the plasma, decaying to large extent in force-balance with induced halo/wall currents, is carefully resolved via 3D nonlinear simulations. The faster 2D nonlinear runs allow to assess the sensitivity of the simulations to parameter changes. In the limit of perfectly conducting wall, the expected linear relation between <span class="hlt">vertical</span> growth rate and wall resistivity is recovered. For intermediate wall resistivities, the halo region contributes to slowing the plasma down, and the characteristic VDE time depends on the choice of halo temperature. The evolution of the current quench and the onset of 3D halo/eddy currents are diagnosed in detail. The 3D simulations highlight a rich structure of toroidal modes, penetrating inwards from edge to core and cascading from high-n to low-n mode numbers. The break-up of flux-surfaces results in a progressive stochastisation of field-lines precipitating the thermalisation of the plasma with the wall. The plasma current then decays rapidly, inducing large currents in the halo region and the wall. Analysis of normal currents flowing in and out of the divertor plate reveals rich time-varying patterns.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007AGUFM.T43B1333C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AGUFM.T43B1333C"><span>Holocene <span class="hlt">vertical</span> tectonic movements of the Taipei Basin, northern Taiwan and its implications</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, B.; Hsieh, M.; Lai, T.; Liew, P.</p> <p>2007-12-01</p> <p>Many geological data of the Taipei Basin, although, have been published by various studies in past decades, however, <span class="hlt">vertical</span> tectonic movement rate of the Basin was not well understood so far. This study, therefore, used radiocarbon dates, obtained from fifteen boreholes in the Basin, to calculate the Holocene <span class="hlt">vertical</span> tectonic movement rate. In addition to the derived tectonic movement rate, this study also discussed the causes of the tectonic patterns of the Taipei Basin. The Taipei Basin, located in the northern Taiwan, was a half graben subsided and extended along the western boundary, the Shangiao Normal Fault, of the Basin. The Holocene <span class="hlt">vertical</span> tectonic movement rate of the Basin were calculated based on 94 radiocarbon dates in fifteen boreholes, the elevations of the radiocarbon dating samples, and the eustatic sea-level curve of the past 15 ka. The results showed the rate in the western part of the Basin, was -2.2 -- -0.9 mm/yr (negative value indicates subsiding, and positive value indicates uplifting). In the central part of the Basin, the rate was ca. -<span class="hlt">1</span> -- <span class="hlt">1</span> mm/yr while in the eastern part of the Basin, the rate was 0.<span class="hlt">1</span> -- <span class="hlt">1</span>.6 mm/yr. Along the Shiangiao Fault, the rate of the hanging-wall was ca. -<span class="hlt">1</span>.6 -- -0.4 mm/yr and the rate of the footwall was ca. 0 mm/yr. According to the results of this study, the present territory of the Taipei Basin was not actually consistent with the tectonic subsiding region. The <span class="hlt">vertical</span> tectonic movement pattern demonstrated subsidence in the western part and uplift in the eastern part of the Taipei Basin. The subsidence of the western part was controlled by the extension of the Shangiao Faul. The uplift of the eastern part might be ascribed to the roll-over of the Fault. Another possibility is that the uplift of the east was controlled by the same behavior as the Western Foothills.Consequently, the deposition of the eastern part of the Basin, wass mainly related to the accommodations due to sea-level <span class="hlt">rise</span> but not</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4142119','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4142119"><span>Learning to Read <span class="hlt">Vertical</span> Text in Peripheral Vision</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Subramanian, Ahalya; Legge, Gordon E.; Wagoner, Gunther Harrison; Yu, Deyue</p> <p>2014-01-01</p> <p>Purpose English–language text is almost always written horizontally. Text can be formatted to run <span class="hlt">vertically</span>, but this is seldom used. Several studies have found that horizontal text can be read faster than <span class="hlt">vertical</span> text in the central visual field. No studies have investigated the peripheral visual field. Studies have also concluded that training can improve reading speed in the peripheral visual field for horizontal text. We aimed to establish whether the horizontal <span class="hlt">vertical</span> differences are maintained and if training can improve <span class="hlt">vertical</span> reading in the peripheral visual field. Methods Eight normally sighted young adults participated in the first study. Rapid Serial Visual Presentation (RSVP) reading speed was measured for horizontal and <span class="hlt">vertical</span> text in the central visual field and at 10° eccentricity in the upper or lower (horizontal text), and right or left (<span class="hlt">vertical</span> text) visual fields. Twenty-one normally sighted young adults split equally between 2 training and <span class="hlt">1</span> control group participated in the second study. Training consisted of RSVP reading either using <span class="hlt">vertical</span> text in the left visual field or horizontal text in the inferior visual field. Subjects trained daily over 4 days. Pre and post horizontal and <span class="hlt">vertical</span> RSVP reading speeds were carried out for all groups. For the training groups these measurements were repeated <span class="hlt">1</span> week and <span class="hlt">1</span> month post training. Results Prior to training, RSVP reading speeds were faster for horizontal text in the central and peripheral visual fields when compared to <span class="hlt">vertical</span> text. Training <span class="hlt">vertical</span> reading improved <span class="hlt">vertical</span> reading speeds by an average factor of 2.8. There was partial transfer of training to the opposite (right) hemifield. The training effects were retained for up to a month. Conclusions RSVP training can improve RSVP <span class="hlt">vertical</span> text reading in peripheral vision. These findings may have implications for patients with macular degeneration or hemianopic field loss. PMID:25062130</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1815715M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1815715M"><span><span class="hlt">Vertical</span> profiles of BC direct radiative effect over Italy: high <span class="hlt">vertical</span> resolution data and atmospheric feedbacks</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Močnik, Griša; Ferrero, Luca; Castelli, Mariapina; Ferrini, Barbara S.; Moscatelli, Marco; Grazia Perrone, Maria; Sangiorgi, Giorgia; Rovelli, Grazia; D'Angelo, Luca; Moroni, Beatrice; Scardazza, Francesco; Bolzacchini, Ezio; Petitta, Marcello; Cappelletti, David</p> <p>2016-04-01</p> <p>Black carbon (BC), and its <span class="hlt">vertical</span> distribution, affects the climate. Global measurements of BC <span class="hlt">vertical</span> profiles are lacking to support climate change research. To fill this gap, a campaign was conducted over three Italian basin valleys, Terni Valley (Appennines), Po Valley and Passiria Valley (Alps), to characterize the impact of BC on the radiative budget under similar orographic conditions. 120 <span class="hlt">vertical</span> profiles were measured in winter 2010. The BC <span class="hlt">vertical</span> profiles, together with aerosol size distribution, aerosol chemistry and meteorological parameters, have been determined using a tethered balloon-based platform equipped with: a micro-Aethalometer AE51 (Magee Scientific), a <span class="hlt">1</span>.107 Grimm OPC (0.25-32 μm, 31 size classes), a cascade impactor (Siuotas SKC), and a meteorological station (LSI-Lastem). The aerosol chemical composition was determined from collected PM2.5 samples. The aerosol absorption along the <span class="hlt">vertical</span> profiles was measured and optical properties calculated using the Mie theory applied to the aerosol size distribution. The aerosol optical properties were validated with AERONET data and then used as inputs to the radiative transfer model libRadtran. <span class="hlt">Vertical</span> profiles of the aerosol direct radiative effect, the related atmospheric absorption and the heating rate were calculated. <span class="hlt">Vertical</span> profile measurements revealed some common behaviors over the studied basin valleys. From below the mixing height to above it, a marked concentration drop was found for both BC (from -48.4±5.3% up to -69.<span class="hlt">1</span>±5.5%) and aerosol number concentration (from -23.9±4.3% up to -46.5±7.3%). These features reflected on the optical properties of the aerosol. Absorption and scattering coefficients decreased from below the MH to above it (babs from -47.6±2.5% up to -71.3±3.0% and bsca from -23.5±0.8% up to -61.2±3.<span class="hlt">1</span>%, respectively). Consequently, the Single Scattering Albedo increased above the MH (from +4.9±2.2% to +7.4±<span class="hlt">1</span>.0%). The highest aerosol absorption was</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3274193','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3274193"><span><span class="hlt">Vertical</span> Guidance Performance Analysis of the L<span class="hlt">1</span>–L5 Dual-Frequency GPS/WAAS User Avionics Sensor</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Jan, Shau-Shiun</p> <p>2010-01-01</p> <p>This paper investigates the potential <span class="hlt">vertical</span> guidance performance of global positioning system (GPS)/wide area augmentation system (WAAS) user avionics sensor when the modernized GPS and Galileo are available. This paper will first investigate the airborne receiver code noise and multipath (CNMP) confidence (σair). The σair will be the dominant factor in the availability analysis of an L<span class="hlt">1</span>–L5 dual-frequency GPS/WAAS user avionics sensor. This paper uses the MATLAB Algorithm Availability Simulation Tool (MAAST) to determine the required values for the σair, so that an L<span class="hlt">1</span>–L5 dual-frequency GPS/WAAS user avionics sensor can meet the <span class="hlt">vertical</span> guidance requirements of APproach with <span class="hlt">Vertical</span> guidance (APV) II and CATegory (CAT) I over conterminous United States (CONUS). A modified MAAST that includes the Galileo satellite constellation is used to determine under what user configurations WAAS could be an APV II system or a CAT I system over CONUS. Furthermore, this paper examines the combinations of possible improvements in signal models and the addition of Galileo to determine if GPS/WAAS user avionics sensor could achieve 10 m <span class="hlt">Vertical</span> Alert Limit (VAL) within the service volume. Finally, this paper presents the future <span class="hlt">vertical</span> guidance performance of GPS user avionics sensor for the United States’ WAAS, Japanese MTSAT-based satellite augmentation system (MSAS) and European geostationary navigation overlay service (EGNOS). PMID:22319263</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24843934','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24843934"><span><span class="hlt">Vertical</span> transmission of Key West dengue-<span class="hlt">1</span> virus by Aedes aegypti and Aedes albopictus (Diptera: Culicidae) mosquitoes from Florida.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Buckner, Eva A; Alto, Barry W; Lounibos, L Philip</p> <p>2013-11-01</p> <p>Following the 2009 and 2010 dengue-<span class="hlt">1</span> (DENV-<span class="hlt">1</span>) outbreaks in Key West, FL, we used Florida Aedes aegypti (L.) mosquitoes and DENV-<span class="hlt">1</span> isolated from Key West in 2010 to test the hypothesis that if the 2009 and 2010 DENV-<span class="hlt">1</span> genome sequences are similar, then <span class="hlt">vertical</span> transmission of DENV-<span class="hlt">1</span> from infected Ae. aegypti female mosquitoes to their eggs could have served as an interepidemic reservoir between outbreaks. We also investigated the ability of Florida Aedes albopictus (Skuse) mosquitoes to <span class="hlt">vertically</span> transmit DENV-<span class="hlt">1</span>. In addition, we determined the rates of infection and dissemination of these Florida mosquito species for DENV-<span class="hlt">1</span> and the effect of DENV-<span class="hlt">1</span> infection on oviposition success and number of mosquito eggs laid by females. <span class="hlt">Vertical</span> transmission of DENV-<span class="hlt">1</span> was documented, with rates of 11.11% (2 out of 18) for Ae. albopictus and 8.33% (3 out of 36) for Ae. aegypti. Approximately 93% (111 out of 119) of Ae. aegypti that fed on DENV-<span class="hlt">1</span> in blood became infected, and 80% (89 out of 111) of infections were disseminated. Similarly, 93% of Ae. albopictus became infected (53 out of 57), and 85% (45 out of 53) of infections were disseminated. No significant differences were detected in numbers of eggs laid by either species after imbibing DENV-<span class="hlt">1</span> in blood, suggesting little cost of infection on number of eggs laid. Our results demonstrate that Florida Ae. aegypti and Ae. albopictus mosquitoes are competent vectors for DENV-<span class="hlt">1</span>, whose maintenance between the 2009 and 2010 Key West outbreaks may have been facilitated by <span class="hlt">vertical</span> transmission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26418656','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26418656"><span>Oviposition and <span class="hlt">vertical</span> dispersal of Aedes mosquitoes in multiple storey buildings in Colombo district, Sri Lanka.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jayathilake, T A Hasini D G; Wickramasinghe, Mervyn B; de Silva, B G D Nissanka K</p> <p>2015-09-01</p> <p>The Colombo City in Sri Lanka is experiencing tremendous development and construction of multiple storey buildings and high <span class="hlt">rise</span> apartments. The change in housing types and microhabitats might have altered the flight and breeding behaviour of Aedes mosquito population. This study was carried out to determine the <span class="hlt">vertical</span> dispersal and abundance of Aedes mosquitoes in multiple storey buildings in the Colombo district, with respect to abiotic factors such as rainfall, humidity and wind speed. Hence, this study is of paramount importance, particularly for planning and implementation of control measures against Aedes mosquitoes. An ovitrap based study was carried out at four selected multiple storey buildings in four residential areas located in Colombo, Sri Lanka, from August to December 2013. Results were analyzed using four indices; ovitrap index, mean number of larvae, mean number of eggs and mean number of larvae per ovipaddle. The results implied that Aedes mosquitoes could be found in different elevations from ground floor to the highest floor (130 ft). There was a significant difference between height and ovitrap index (p<0.05), and height and mean number of larvae per recovered ovipaddle (p<0.05). The highest index value for mean number of eggs was observed as 3.492 ± 0.655 at the 6th floor (60 ft high from ground level). At the same height (60 ft height) other indices (ovitrap index, mean number of larvae and mean number of larvae per ovipaddle) also displayed higher values, i.e. 13.19 ± 2.98%, <span class="hlt">1</span>.366 ± 0.527, and 2.070 ± 0.421%, respectively. Abiotic factors such as wind speed, coastal nature, etc. displayed a significant effect to the <span class="hlt">vertical</span> dispersal of Aedes mosquitoes (p<0.05). The study suggested that Aedes mosquitoes are able to breed at any level of the buildings and not restricted by their height. The indices (mean number of larvae, mean number of eggs) representing the <span class="hlt">vertical</span> dispersal with respect to abundance seemed to be statistically non</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AAS...23221102K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AAS...23221102K"><span>Sun Radio Interferometer Space Experiment (Sun<span class="hlt">RISE</span>)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kasper, Justin C.; SunRISE Team</p> <p>2018-06-01</p> <p>The Sun Radio Interferometer Space Experiment (Sun<span class="hlt">RISE</span>) is a NASA Heliophysics Explorer Mission of Opportunity currently in Phase A. Sun<span class="hlt">RISE</span> is a constellation of spacecraft flying in a 10-km diameter formation and operating as the first imaging radio interferometer in space. The purpose of Sun<span class="hlt">RISE</span> is to reveal critical aspects of solar energetic particle (SEP) acceleration at coronal mass ejections (CMEs) and transport into space by making the first spatially resolved observations of coherent Type II and III radio bursts produced by electrons accelerated at CMEs or released from flares. Sun<span class="hlt">RISE</span> will focus on solar Decametric-Hectometric (DH, 0.<span class="hlt">1</span> < f < 15 MHz) radio bursts that always are detected from space before major SEP events, but cannot be seen on Earth due to ionospheric absorption. This talk will describe Sun<span class="hlt">RISE</span> objectives and implementation. Presented on behalf of the entire Sun<span class="hlt">RISE</span> team.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NatSD...580044W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NatSD...580044W"><span>A Mediterranean coastal database for assessing the impacts of sea-level <span class="hlt">rise</span> and associated hazards</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wolff, Claudia; Vafeidis, Athanasios T.; Muis, Sanne; Lincke, Daniel; Satta, Alessio; Lionello, Piero; Jimenez, Jose A.; Conte, Dario; Hinkel, Jochen</p> <p>2018-03-01</p> <p>We have developed a new coastal database for the Mediterranean basin that is intended for coastal impact and adaptation assessment to sea-level <span class="hlt">rise</span> and associated hazards on a regional scale. The data structure of the database relies on a linear representation of the coast with associated spatial assessment units. Using information on coastal morphology, human settlements and administrative boundaries, we have divided the Mediterranean coast into 13 900 coastal assessment units. To these units we have spatially attributed 160 parameters on the characteristics of the natural and socio-economic subsystems, such as extreme sea levels, <span class="hlt">vertical</span> land movement and number of people exposed to sea-level <span class="hlt">rise</span> and extreme sea levels. The database contains information on current conditions and on plausible future changes that are essential drivers for future impacts, such as sea-level <span class="hlt">rise</span> rates and socio-economic development. Besides its intended use in risk and impact assessment, we anticipate that the Mediterranean Coastal Database (MCD) constitutes a useful source of information for a wide range of coastal applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5870338','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5870338"><span>A Mediterranean coastal database for assessing the impacts of sea-level <span class="hlt">rise</span> and associated hazards</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wolff, Claudia; Vafeidis, Athanasios T.; Muis, Sanne; Lincke, Daniel; Satta, Alessio; Lionello, Piero; Jimenez, Jose A.; Conte, Dario; Hinkel, Jochen</p> <p>2018-01-01</p> <p>We have developed a new coastal database for the Mediterranean basin that is intended for coastal impact and adaptation assessment to sea-level <span class="hlt">rise</span> and associated hazards on a regional scale. The data structure of the database relies on a linear representation of the coast with associated spatial assessment units. Using information on coastal morphology, human settlements and administrative boundaries, we have divided the Mediterranean coast into 13 900 coastal assessment units. To these units we have spatially attributed 160 parameters on the characteristics of the natural and socio-economic subsystems, such as extreme sea levels, <span class="hlt">vertical</span> land movement and number of people exposed to sea-level <span class="hlt">rise</span> and extreme sea levels. The database contains information on current conditions and on plausible future changes that are essential drivers for future impacts, such as sea-level <span class="hlt">rise</span> rates and socio-economic development. Besides its intended use in risk and impact assessment, we anticipate that the Mediterranean Coastal Database (MCD) constitutes a useful source of information for a wide range of coastal applications. PMID:29583140</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22904355','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22904355"><span>Processing <span class="hlt">vertical</span> size disparities in distinct depth planes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Duke, Philip A; Howard, Ian P</p> <p>2012-08-17</p> <p>A textured surface appears slanted about a <span class="hlt">vertical</span> axis when the image in one eye is horizontally enlarged relative to the image in the other eye. The surface appears slanted in the opposite direction when the same image is <span class="hlt">vertically</span> enlarged. Two superimposed textured surfaces with different horizontal size disparities appear as two surfaces that differ in slant. Superimposed textured surfaces with equal and opposite <span class="hlt">vertical</span> size disparities appear as a single frontal surface. The <span class="hlt">vertical</span> disparities are averaged. We investigated whether <span class="hlt">vertical</span> size disparities are averaged across two superimposed textured surfaces in different depth planes or whether they induce distinct slants in the two depth planes. In Experiment <span class="hlt">1</span>, two superimposed textured surfaces with different <span class="hlt">vertical</span> size disparities were presented in two depth planes defined by horizontal disparity. The surfaces induced distinct slants when the horizontal disparity was more than ±5 arcmin. Thus, <span class="hlt">vertical</span> size disparities are not averaged over surfaces with different horizontal disparities. In Experiment 2 we confirmed that <span class="hlt">vertical</span> size disparities are processed in surfaces away from the horopter, so the results of Experiment <span class="hlt">1</span> cannot be explained by the processing of <span class="hlt">vertical</span> size disparities in a fixated surface only. Together, these results show that <span class="hlt">vertical</span> size disparities are processed separately in distinct depth planes. The results also suggest that <span class="hlt">vertical</span> size disparities are not used to register slant globally by their effect on the registration of binocular direction of gaze.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014WRR....50.3994C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014WRR....50.3994C"><span>The <span class="hlt">vertical</span> variability of hyporheic fluxes inferred from riverbed temperature data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cranswick, Roger H.; Cook, Peter G.; Shanafield, Margaret; Lamontagne, Sebastien</p> <p>2014-05-01</p> <p>We present detailed profiles of <span class="hlt">vertical</span> water flux from the surface to <span class="hlt">1</span>.2 m beneath the Haughton River in the tropical northeast of Australia. A <span class="hlt">1</span>-D numerical model is used to estimate <span class="hlt">vertical</span> flux based on raw temperature time series observations from within downwelling, upwelling, neutral, and convergent sections of the hyporheic zone. A Monte Carlo analysis is used to derive error bounds for the fluxes based on temperature measurement error and uncertainty in effective thermal diffusivity. <span class="hlt">Vertical</span> fluxes ranged from 5.7 m d-<span class="hlt">1</span> (downward) to -0.2 m d-<span class="hlt">1</span> (upward) with the lowest relative errors for values between 0.3 and 6 m d-<span class="hlt">1</span>. Our <span class="hlt">1</span>-D approach provides a useful alternative to <span class="hlt">1</span>-D analytical and other solutions because it does not incorporate errors associated with simplified boundary conditions or assumptions of purely <span class="hlt">vertical</span> flow, hydraulic parameter values, or hydraulic conditions. To validate the ability of this <span class="hlt">1</span>-D approach to represent the <span class="hlt">vertical</span> fluxes of 2-D flow fields, we compare our model with two simple 2-D flow fields using a commercial numerical model. These comparisons showed that: (<span class="hlt">1</span>) the <span class="hlt">1</span>-D <span class="hlt">vertical</span> flux was equivalent to the mean <span class="hlt">vertical</span> component of flux irrespective of a changing horizontal flux; and (2) the subsurface temperature data inherently has a "spatial footprint" when the <span class="hlt">vertical</span> flux profiles vary spatially. Thus, the mean <span class="hlt">vertical</span> flux within a 2-D flow field can be estimated accurately without requiring the flow to be purely <span class="hlt">vertical</span>. The temperature-derived <span class="hlt">1</span>-D <span class="hlt">vertical</span> flux represents the integrated <span class="hlt">vertical</span> component of flux along the flow path intersecting the observation point. This article was corrected on 6 JUN 2014. See the end of the full text for details.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_8 --> <div id="page_9" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="161"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-VAFB-20160706-RKB-MH-01-0001-Delta_II_Interstage_Hoist_Horizontal_Vertical_on_truck_+Bldg_836_VAFB-3132117.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-VAFB-20160706-RKB-MH-01-0001-Delta_II_Interstage_Hoist_Horizontal_Vertical_on_truck_+Bldg_836_VAFB-3132117.html"><span>JPSS-<span class="hlt">1</span> Delta II Interstage Hoisted from Horizontal and Rotated to <span class="hlt">Vertical</span> for Transport</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2016-07-06</p> <p>The interstage section of the United Launch Alliance Delta II rocket that will launch the Joint Polar Satellite System-<span class="hlt">1</span> (JPSS-<span class="hlt">1</span>) is hoisted to <span class="hlt">vertical</span> in Building 836 on Vandenberg Air Force Base in California. JPSS, a next-generation environmental satellite system, is a collaborative program between the National Oceanic and Atmospheric Administration (NOAA) and NASA. Launch is targeted for March 27, 2017. To learn more about JPSS-<span class="hlt">1</span>, visit www.jpss.noaa.gov.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.A51B0022C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.A51B0022C"><span>Convection Fingerprints on the <span class="hlt">Vertical</span> Profiles of Q<span class="hlt">1</span> and Q2</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chang, C.; Lin, H.; Chou, C.</p> <p>2013-12-01</p> <p>Different types of tropical convection left their fingerprints on <span class="hlt">vertical</span> structures of apparent heat source (Q<span class="hlt">1</span>) and apparent moisture sink (Q2). Profile of deep convection on condensation heating and drying has been well-documented, yet direct assessment of shallow convection remains to be explored. Shallow convection prevails over subtropical ocean, where large-scale subsidence is primarily balanced by radiative cooling and moistening due to surface evaporation instead of moist convection. In this study a united framework is designed to investigate the <span class="hlt">vertical</span> structures of tropical marine convections in three reanalysis data, including ERA-Interim, MERRA, and CFSR. It starts by sorting and binning data from the lightest to the heaviest rain. Then the differences between two neighboring bins are used to examine the direct effects for precipitation change, in light of the fact that non-convective processes would change slowly from bin to bin. It is shown that all three reanalyses reveal the shallow convective processes in light rain bins, featured by re-evaporating and detraining at the top of boundary layer and lower free troposphere. For heavy rain bins, three reanalyses mainly differ in their numbers and altitudes of heating and drying peaks, implying no universal agreement has been reached on partitioning of cloud populations. Coherent variations in temperature, moisture, and <span class="hlt">vertical</span> motion are also discussed. This approach permits a systematical survey and comparison of tropical convection in GCM-type models, and preliminary studies of three reanalyses suggest certain degree of inconsistency in simulated convective feedback to large-scale heat and moisture budgets.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21141661','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21141661"><span>Contemporary sea level <span class="hlt">rise</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cazenave, Anny; Llovel, William</p> <p>2010-01-01</p> <p>Measuring sea level change and understanding its causes has considerably improved in the recent years, essentially because new in situ and remote sensing observations have become available. Here we report on most recent results on contemporary sea level <span class="hlt">rise</span>. We first present sea level observations from tide gauges over the twentieth century and from satellite altimetry since the early 1990s. We next discuss the most recent progress made in quantifying the processes causing sea level change on timescales ranging from years to decades, i.e., thermal expansion of the oceans, land ice mass loss, and land water-storage change. We show that for the 1993-2007 time span, the sum of climate-related contributions (2.85 +/- 0.35 mm year(-<span class="hlt">1</span>)) is only slightly less than altimetry-based sea level <span class="hlt">rise</span> (3.3 +/- 0.4 mm year(-<span class="hlt">1</span>)): approximately 30% of the observed rate of <span class="hlt">rise</span> is due to ocean thermal expansion and approximately 55% results from land ice melt. Recent acceleration in glacier melting and ice mass loss from the ice sheets increases the latter contribution up to 80% for the past five years. We also review the main causes of regional variability in sea level trends: The dominant contribution results from nonuniform changes in ocean thermal expansion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20050234670','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20050234670"><span>The Tropical Convective Spectrum. Part <span class="hlt">1</span>; Archetypal <span class="hlt">Vertical</span> Structures</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Boccippio, Dennis J.; Petersen, Walter A.; Cecil, Daniel J.</p> <p>2005-01-01</p> <p>A taxonomy of tropical convective and stratiform <span class="hlt">vertical</span> structures is constructed through cluster analysis of 3 yr of Tropical Rainfall Measuring Mission (TRMM) "warm-season" (surface temperature greater than 10 C) precipitation radar (PR) <span class="hlt">vertical</span> profiles, their surface rainfall, and associated radar-based classifiers (convective/ stratiform and brightband existence). Twenty-five archetypal profile types are identified, including nine convective types, eight stratiform types, two mixed types, and six anvil/fragment types (nonprecipitating anvils and sheared deep convective profiles). These profile types are then hierarchically clustered into 10 similar families, which can be further combined, providing an objective and physical reduction of the highly multivariate PR data space that retains <span class="hlt">vertical</span> structure information. The taxonomy allows for description of any storm or local convective spectrum by the profile types or families. The analysis provides a quasi-independent corroboration of the TRMM 2A23 convective/ stratiform classification. The global frequency of occurrence and contribution to rainfall for the profile types are presented, demonstrating primary rainfall contribution by midlevel glaciated convection (27%) and similar depth decaying/stratiform stages (28%-31%). Profiles of these types exhibit similar 37- and 85-GHz passive microwave brightness temperatures but differ greatly in their frequency of occurrence and mean rain rates, underscoring the importance to passive microwave rain retrieval of convective/stratiform discrimination by other means, such as polarization or texture techniques, or incorporation of lightning observations. Close correspondence is found between deep convective profile frequency and annualized lightning production, and pixel-level lightning occurrence likelihood directly tracks the estimated mean ice water path within profile types.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPhCS.995a2100K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPhCS.995a2100K"><span>Identification of Natural Frequency of Low <span class="hlt">Rise</span> Building on Soft Ground Profile using Ambient Vibration Method</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kamarudin, A. F.; Zainal Abidin, M. H.; Mokhatar, S. N.; Daud, M. E.; Ibrahim, A.; Ibrahim, Z.; Noh, M. S. Md</p> <p>2018-04-01</p> <p>Natural frequency is the rate at which a body to vibrate or oscillate. Application of ambient vibration (AV) excitation is widely used nowadays as the input motion for building predominant frequency, fo, and ground fundamental frequency, Fo, prediction due to simple, fast, non-destructive, simple handling operation and reliable result. However, it must be emphasized and caution to isolate these frequencies (fo and Fo) from spurious frequencies of site-structure effects especially to low <span class="hlt">rise</span> building on soft ground deposit. In this study, identification of fo and Fo by using AV measurements were performed on ground and 4-storey primary school reinforced concrete (RC) building at Sekolah Kebangsaan (SK) Sg. Tongkang, Rengit, Johor using <span class="hlt">1</span> Hz of tri-axial seismometer sensor. Overlapping spectra between Fourier Amplitude Spectra (FAS) from and Horizontal to <span class="hlt">Vertical</span> Spectra Ratio (HVSR) were used to distinguish respective frequencies of building and ground natural frequencies. Three dominant frequencies were identified from the FAS curves at <span class="hlt">1</span>.91 Hz, <span class="hlt">1</span>.98 Hz and 2.79 Hz in longitudinal (East West-EW), transverse (North South-NS) and <span class="hlt">vertical</span> (UD) directions. It is expected the building has deformed in translational mode based on the first peak frequency by respective NS and EW components of FAS spectrum. <span class="hlt">Vertical</span> frequency identified from the horizontal spectrums, might induces to the potential of rocking effect experienced by the school building. Meanwhile, single peak HVSR spectrum at low ground fundamental frequency concentrated at 0.93 Hz indicates to the existence deep contrast of soft deposit. Strong interaction between ground and building at similar frequency (0.93 Hz) observed from the FAS curves on the highest floor has shown the building to behave as a dependent unit against ground response as one rigid mass.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4444649','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4444649"><span>Association of affect with <span class="hlt">vertical</span> position in L<span class="hlt">1</span> but not in L2 in unbalanced bilinguals</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Li, Degao; Liu, Haitao; Ma, Bosen</p> <p>2015-01-01</p> <p>After judging the valence of the positive (e.g., happy) and the negative words (e.g., sad), the participants' response to the letter (q or p) was faster and slower, respectively, when the letter appeared at the upper end than at the lower end of the screen in Meier and Robinson's (2004) second experiment. To compare this metaphorical association of affect with <span class="hlt">vertical</span> position in Chinese-English bilinguals' first language (L<span class="hlt">1</span>) and second language (L2) (language), we conducted four experiments in an affective priming task. The targets were one set of positive or negative words (valence), which were shown <span class="hlt">vertically</span> above or below the center of the screen (position). The primes, presented at the center of the screen, were affective words that were semantically related to the targets, affective words that were not semantically related to the targets, affective icon-pictures, and neutral strings in Experiment <span class="hlt">1</span>–4, respectively. In judging the targets' valence, the participants showed different patterns of interactions between language, valence, and position in reaction times across the experiments. We concluded that metaphorical association between affect and <span class="hlt">vertical</span> position works in L<span class="hlt">1</span> but not in L2 for unbalanced bilinguals. PMID:26074847</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004JCrGr.270..179Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004JCrGr.270..179Z"><span><span class="hlt">Vertical</span> Bridgman growth of Hg <span class="hlt">1</span>-xMn xTe with variational withdrawal rate</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhi, Gu; Wan-Qi, Jie; Guo-Qiang, Li; Long, Zhang</p> <p>2004-09-01</p> <p>Based on the solute redistribution models, <span class="hlt">Vertical</span> Bridgman growth of Hg<span class="hlt">1</span>-xMnxTe with variational withdrawal rate is studied. Both theoretical analysis and experimental results show that the axial composition uniformity is improved and the crystal growth rate is also increased at the optimized variational method of withdrawal rate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940027929','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940027929"><span>THE <span class="hlt">VERTICAL</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Albert, Stephen L.; Spencer, Jeffrey B.</p> <p>1994-01-01</p> <p>'THE <span class="hlt">VERTICAL</span>' computer keyboard is designed to address critical factors which contribute to Repetitive Motion Injuries (RMI) (including Carpal Tunnel Syndrome) in association with computer keyboard usage. This keyboard splits the standard QWERTY design into two halves and positions each half 90 degrees from the desk. In order to access a computer correctly. 'THE <span class="hlt">VERTICAL</span>' requires users to position their bodies in optimal alignment with the keyboard. The orthopaedically neutral forearm position (with hands palms-in and thumbs-up) reduces nerve compression in the forearm. The <span class="hlt">vertically</span> arranged keypad halves ameliorate onset occurrence of keyboard-associated RMI. By utilizing visually-reference mirrored mylar surfaces adjustable to the user's eye, the user is able to readily reference any key indicia (reversed) just as they would on a conventional keyboard. Transverse adjustability substantially reduces cumulative musculoskeletal discomfort in the shoulders. 'THE <span class="hlt">VERTICAL</span>' eliminates the need for an exterior mouse by offering a convenient finger-accessible curser control while the hands remain in the <span class="hlt">vertically</span> neutral position. The potential commercial application for 'THE <span class="hlt">VERTICAL</span>' is enormous since the product can effect every person who uses a computer anywhere in the world. Employers and their insurance carriers are spending hundreds of millions of dollars per year as a result of RMI. This keyboard will reduce the risk.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DFDH10009S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DFDH10009S"><span>Heat and momentum transport scalings in <span class="hlt">vertical</span> convection</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shishkina, Olga</p> <p>2016-11-01</p> <p>For <span class="hlt">vertical</span> convection, where a fluid is confined between two differently heated isothermal <span class="hlt">vertical</span> walls, we investigate the heat and momentum transport, which are measured, respectively, by the Nusselt number Nu and the Reynolds number Re . For laminar <span class="hlt">vertical</span> convection we derive analytically the dependence of Re and Nu on the Rayleigh number Ra and the Prandtl number Pr from our boundary layer equations and find two different scaling regimes: Nu Pr <span class="hlt">1</span> / 4 Ra <span class="hlt">1</span> / 4 , Re Pr - <span class="hlt">1</span> / 2 Ra <span class="hlt">1</span> / 2 for Pr << <span class="hlt">1</span> and Nu Pr0 Ra <span class="hlt">1</span> / 4 , Re Pr-<span class="hlt">1</span> Ra <span class="hlt">1</span> / 2 for Pr >> <span class="hlt">1</span> . Direct numerical simulations for Ra from 105 to 1010 and Pr from 0.01 to 30 are in excellent ageement with our theoretical findings and show that the transition between the regimes takes place for Pr around 0.<span class="hlt">1</span>. We summarize the results from and present new theoretical and numerical results for transitional and turbulent <span class="hlt">vertical</span> convection. The work is supported by the Deutsche Forschungsgemeinschaft (DFG) under the Grant Sh 405/4 - Heisenberg fellowship.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22902397','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22902397"><span>Polymorphisms in DC-SIGN and L-SIGN genes are associated with HIV-<span class="hlt">1</span> <span class="hlt">vertical</span> transmission in a Northeastern Brazilian population.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>da Silva, Ronaldo Celerino; Segat, Ludovica; Zanin, Valentina; Arraes, Luiz Claudio; Crovella, Sergio</p> <p>2012-11-01</p> <p>DC-SIGN and L-SIGN are receptors expressed on specialized macrophages in decidua, (Hofbauer and placental capillary endothelial cells), known to interact with several pathogens, including HIV-<span class="hlt">1</span>. To disclose the possible involvement of these molecules in the susceptibility to HIV <span class="hlt">vertical</span> transmission, we analyzed DC-SIGN and L-SIGN gene single nucleotide polymorphisms (SNPs) in 192 HIV-<span class="hlt">1</span> positive children and 58 HIV-<span class="hlt">1</span> negative children all born to HIV-<span class="hlt">1</span> positive mothers, as well as 96 healthy uninfected children not exposed to HIV-<span class="hlt">1</span>, all from Northeast Brazil. The frequency of three SNPs in the DC-SIGN promoter (-139G>A, -201G>T and -336A>G) were significantly different when comparing HIV positive children with HIV-<span class="hlt">1</span> exposed uninfected children, indicating an association with susceptibility to HIV-<span class="hlt">1</span> <span class="hlt">vertical</span> transmission. This genetic association suggests that DC-SIGN molecule may play a role in susceptibility to HIV-<span class="hlt">1</span> infection through <span class="hlt">vertical</span> transmission. Copyright © 2012 American Society for Histocompatibility and Immunogenetics. Published by Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1215471','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1215471"><span>Direct Observation of Ultralow <span class="hlt">Vertical</span> Emittance using a <span class="hlt">Vertical</span> Undulator</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Wootton, Kent</p> <p>2015-09-17</p> <p>In recent work, the first quantitative measurements of electron beam <span class="hlt">vertical</span> emittance using a <span class="hlt">vertical</span> undulator were presented, with particular emphasis given to ultralow <span class="hlt">vertical</span> emittances [K. P. Wootton, et al., Phys. Rev. ST Accel. Beams, 17, 112802 (2014)]. Using this apparatus, a geometric <span class="hlt">vertical</span> emittance of 0.9 ± 0.3 pm rad has been observed. A critical analysis is given of measurement approaches that were attempted, with particular emphasis on systematic and statistical uncertainties. The method used is explained, compared to other techniques and the applicability of these results to other scenarios discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5127314','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5127314"><span>Coastal sea level <span class="hlt">rise</span> with warming above 2 °C</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Jevrejeva, Svetlana; Jackson, Luke P.; Riva, Riccardo E. M.; Grinsted, Aslak; Moore, John C.</p> <p>2016-01-01</p> <p>Two degrees of global warming above the preindustrial level is widely suggested as an appropriate threshold beyond which climate change risks become unacceptably high. This “2 °C” threshold is likely to be reached between 2040 and 2050 for both Representative Concentration Pathway (RCP) 8.5 and 4.5. Resulting sea level <span class="hlt">rises</span> will not be globally uniform, due to ocean dynamical processes and changes in gravity associated with water mass redistribution. Here we provide probabilistic sea level <span class="hlt">rise</span> projections for the global coastline with warming above the 2 °C goal. By 2040, with a 2 °C warming under the RCP8.5 scenario, more than 90% of coastal areas will experience sea level <span class="hlt">rise</span> exceeding the global estimate of 0.2 m, with up to 0.4 m expected along the Atlantic coast of North America and Norway. With a 5 °C <span class="hlt">rise</span> by 2100, sea level will <span class="hlt">rise</span> rapidly, reaching 0.9 m (median), and 80% of the coastline will exceed the global sea level <span class="hlt">rise</span> at the 95th percentile upper limit of <span class="hlt">1</span>.8 m. Under RCP8.5, by 2100, New York may expect <span class="hlt">rises</span> of <span class="hlt">1</span>.09 m, Guangzhou may expect <span class="hlt">rises</span> of 0.91 m, and Lagos may expect <span class="hlt">rises</span> of 0.90 m, with the 95th percentile upper limit of 2.24 m, <span class="hlt">1</span>.93 m, and <span class="hlt">1</span>.92 m, respectively. The coastal communities of rapidly expanding cities in the developing world, and vulnerable tropical coastal ecosystems, will have a very limited time after midcentury to adapt to sea level <span class="hlt">rises</span> unprecedented since the dawn of the Bronze Age. PMID:27821743</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27821743','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27821743"><span>Coastal sea level <span class="hlt">rise</span> with warming above 2 °C.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jevrejeva, Svetlana; Jackson, Luke P; Riva, Riccardo E M; Grinsted, Aslak; Moore, John C</p> <p>2016-11-22</p> <p>Two degrees of global warming above the preindustrial level is widely suggested as an appropriate threshold beyond which climate change risks become unacceptably high. This "2 °C" threshold is likely to be reached between 2040 and 2050 for both Representative Concentration Pathway (RCP) 8.5 and 4.5. Resulting sea level <span class="hlt">rises</span> will not be globally uniform, due to ocean dynamical processes and changes in gravity associated with water mass redistribution. Here we provide probabilistic sea level <span class="hlt">rise</span> projections for the global coastline with warming above the 2 °C goal. By 2040, with a 2 °C warming under the RCP8.5 scenario, more than 90% of coastal areas will experience sea level <span class="hlt">rise</span> exceeding the global estimate of 0.2 m, with up to 0.4 m expected along the Atlantic coast of North America and Norway. With a 5 °C <span class="hlt">rise</span> by 2100, sea level will <span class="hlt">rise</span> rapidly, reaching 0.9 m (median), and 80% of the coastline will exceed the global sea level <span class="hlt">rise</span> at the 95th percentile upper limit of <span class="hlt">1</span>.8 m. Under RCP8.5, by 2100, New York may expect <span class="hlt">rises</span> of <span class="hlt">1</span>.09 m, Guangzhou may expect <span class="hlt">rises</span> of 0.91 m, and Lagos may expect <span class="hlt">rises</span> of 0.90 m, with the 95th percentile upper limit of 2.24 m, <span class="hlt">1</span>.93 m, and <span class="hlt">1</span>.92 m, respectively. The coastal communities of rapidly expanding cities in the developing world, and vulnerable tropical coastal ecosystems, will have a very limited time after midcentury to adapt to sea level <span class="hlt">rises</span> unprecedented since the dawn of the Bronze Age.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title33-vol1/pdf/CFR-2011-title33-vol1-sec118-85.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title33-vol1/pdf/CFR-2011-title33-vol1-sec118-85.pdf"><span>33 CFR 118.85 - Lights on <span class="hlt">vertical</span> lift bridges.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-07-01</p> <p>... 33 Navigation and Navigable Waters <span class="hlt">1</span> 2011-07-01 2011-07-01 false Lights on <span class="hlt">vertical</span> lift bridges... BRIDGES BRIDGE LIGHTING AND OTHER SIGNALS § 118.85 Lights on <span class="hlt">vertical</span> lift bridges. (a) Lift span lights. The <span class="hlt">vertical</span> lift span of every <span class="hlt">vertical</span> lift bridge shall be lighted so that the center of the...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title33-vol1/pdf/CFR-2010-title33-vol1-sec118-85.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title33-vol1/pdf/CFR-2010-title33-vol1-sec118-85.pdf"><span>33 CFR 118.85 - Lights on <span class="hlt">vertical</span> lift bridges.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-07-01</p> <p>... 33 Navigation and Navigable Waters <span class="hlt">1</span> 2010-07-01 2010-07-01 false Lights on <span class="hlt">vertical</span> lift bridges... BRIDGES BRIDGE LIGHTING AND OTHER SIGNALS § 118.85 Lights on <span class="hlt">vertical</span> lift bridges. (a) Lift span lights. The <span class="hlt">vertical</span> lift span of every <span class="hlt">vertical</span> lift bridge shall be lighted so that the center of the...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70190059','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70190059"><span>Microfossil measures of rapid sea-level <span class="hlt">rise</span>: Timing of response of two microfossil groups to a sudden tidal-flooding experiment in Cascadia</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Horton, B.P.; Milker, Yvonne; Dura, T.; Wang, Kelin; Bridgeland, W.T.; Brophy, Laura S.; Ewald, M.; Khan, Nicole; Engelhart, S.E.; Nelson, Alan R.; Witter, Robert C.</p> <p>2017-01-01</p> <p>Comparisons of pre-earthquake and post-earthquake microfossils in tidal sequences are accurate means to measure coastal subsidence during past subduction earthquakes, but the amount of subsidence is uncertain, because the response times of fossil taxa to coseismic relative sea-level (RSL) <span class="hlt">rise</span> are unknown. We measured the response of diatoms and foraminifera to restoration of a salt marsh in southern Oregon, USA. Tidal flooding following dike removal caused an RSL <span class="hlt">rise</span> of ∼<span class="hlt">1</span> m, as might occur by coseismic subsidence during momentum magnitude (Mw) 8.<span class="hlt">1</span>–8.8 earthquakes on this section of the Cascadia subduction zone. Less than two weeks after dike removal, diatoms colonized low marsh and tidal flats in large numbers, showing that they can record seismically induced subsidence soon after earthquakes. In contrast, low-marsh foraminifera took at least 11 months to appear in sizeable numbers. Where subsidence measured with diatoms and foraminifera differs, their different response times may provide an estimate of postseismic <span class="hlt">vertical</span> deformation in the months following past megathrust earthquakes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009PhRvE..79a1604B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009PhRvE..79a1604B"><span>Capillary <span class="hlt">rise</span> between planar surfaces</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bullard, Jeffrey W.; Garboczi, Edward J.</p> <p>2009-01-01</p> <p>Minimization of free energy is used to calculate the equilibrium <span class="hlt">vertical</span> <span class="hlt">rise</span> and meniscus shape of a liquid column between two closely spaced, parallel planar surfaces that are inert and immobile. States of minimum free energy are found using standard variational principles, which lead not only to an Euler-Lagrange differential equation for the meniscus shape and elevation, but also to the boundary conditions at the three-phase junction where the liquid meniscus intersects the solid walls. The analysis shows that the classical Young-Dupré equation for the thermodynamic contact angle is valid at the three-phase junction, as already shown for sessile drops with or without the influence of a gravitational field. Integration of the Euler-Lagrange equation shows that a generalized Laplace-Young (LY) equation first proposed by O’Brien, Craig, and Peyton [J. Colloid Interface Sci. 26, 500 (1968)] gives an exact prediction of the mean elevation of the meniscus at any wall separation, whereas the classical LY equation for the elevation of the midpoint of the meniscus is accurate only when the separation approaches zero or infinity. When both walls are identical, the meniscus is symmetric about the midpoint, and the midpoint elevation is a more traditional and convenient measure of capillary <span class="hlt">rise</span> than the mean elevation. Therefore, for this symmetric system a different equation is fitted to numerical predictions of the midpoint elevation and is shown to give excellent agreement for contact angles between 15° and 160° and wall separations up to 30mm . When the walls have dissimilar surface properties, the meniscus generally assumes an asymmetric shape, and significant elevation of the liquid column can occur even when one of the walls has a contact angle significantly greater than 90°. The height of the capillary <span class="hlt">rise</span> depends on the spacing between the walls and also on the difference in contact angles at the two surfaces. When the contact angle at one wall is greater</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..17.5136N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..17.5136N"><span>The role of <span class="hlt">vertical</span> land movements on late 19th century sea level <span class="hlt">rise</span> at Cuxhaven, Germany</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Niehüser, Sebastian; Jensen, Jürgen; Wahl, Thomas; Dangendorf, Sönke; Hofstede, Jacobus</p> <p>2015-04-01</p> <p>Tide gauges, located along the world's coastlines, represent one of the most important data sources with information about sea level change back into the 17th century, bridging the gap between paleo proxies and modern remote sensing data sources. While the worldwide coverage of tide gauges has increased considerably since the mid-20th century, there are only a few gauges available providing information about regional sea level changes before 1900. Furthermore, these tide gauge measurements are often contaminated by local <span class="hlt">vertical</span> land movements (VLM) resulting from tectonic processes or local anthropogenic interventions. Such non-climatic effects need to be removed from the raw data to uncover climate signals, which are important, for instance, for answering the question whether and when sea level started to accelerate from the nearly constant rates over the past 2000 years. Here we focus on one of these long tide gauge records: Cuxhaven, which is located in the German Bight and provides uninterrupted digital time series of tidal high and low water levels since 1843. The record has been extensively studied during the past decades with respect to regional and global sea level <span class="hlt">rise</span>. However, a question that still remains is the role of local subsidence before 1900 at the lighthouse of Cuxhaven, located close to the tide gauge. In 1855 Lentz installed a granite height mark at the lighthouse, which was later used as a proxy for VLMs of the tide gauge itself. The height of the control mark was derived by a levelling between Hamburg and Cuxhaven. These levellings were repeated five times between 1855 and 1900 and later evaluated by Siefert and Lassen (1985) with respect to the role of local subsidence. Based on a linear regression of individual levellings Siefert and Lassen (1985) concluded that the lighthouse subsided by an average rate of 2.8 mm/yr (1855-1875: 4.2 mm/yr; 1876-1890: 2 mm/yr; 1890-1900: <span class="hlt">1</span>.2 mm/yr). However, due to the massive uncertainties of these early</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28905897','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28905897"><span>Impact of a global temperature <span class="hlt">rise</span> of <span class="hlt">1</span>.5 degrees Celsius on Asia's glaciers.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kraaijenbrink, P D A; Bierkens, M F P; Lutz, A F; Immerzeel, W W</p> <p>2017-09-13</p> <p>Glaciers in the high mountains of Asia (HMA) make a substantial contribution to the water supply of millions of people, and they are retreating and losing mass as a result of anthropogenic climate change at similar rates to those seen elsewhere. In the Paris Agreement of 2015, 195 nations agreed on the aspiration to limit the level of global temperature <span class="hlt">rise</span> to <span class="hlt">1</span>.5 degrees Celsius ( °C) above pre-industrial levels. However, it is not known what an increase of <span class="hlt">1</span>.5 °C would mean for the glaciers in HMA. Here we show that a global temperature <span class="hlt">rise</span> of <span class="hlt">1</span>.5 °C will lead to a warming of 2.<span class="hlt">1</span> ± 0.<span class="hlt">1</span> °C in HMA, and that 64 ± 7 per cent of the present-day ice mass stored in the HMA glaciers will remain by the end of the century. The <span class="hlt">1</span>.5 °C goal is extremely ambitious and is projected by only a small number of climate models of the conservative IPCC's Representative Concentration Pathway (RCP)2.6 ensemble. Projections for RCP4.5, RCP6.0 and RCP8.5 reveal that much of the glacier ice is likely to disappear, with projected mass losses of 49 ± 7 per cent, 51 ± 6 per cent and 64 ± 5 per cent, respectively, by the end of the century; these projections have potentially serious consequences for regional water management and mountain communities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014GMDD....7.3717C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014GMDD....7.3717C"><span>Verification of a non-hydrostatic dynamical core using horizontally spectral element <span class="hlt">vertically</span> finite difference method: 2-D aspects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Choi, S.-J.; Giraldo, F. X.; Kim, J.; Shin, S.</p> <p>2014-06-01</p> <p>The non-hydrostatic (NH) compressible Euler equations of dry atmosphere are solved in a simplified two dimensional (2-D) slice framework employing a spectral element method (SEM) for the horizontal discretization and a finite difference method (FDM) for the <span class="hlt">vertical</span> discretization. The SEM uses high-order nodal basis functions associated with Lagrange polynomials based on Gauss-Lobatto-Legendre (GLL) quadrature points. The FDM employs a third-order upwind biased scheme for the <span class="hlt">vertical</span> flux terms and a centered finite difference scheme for the <span class="hlt">vertical</span> derivative terms and quadrature. The Euler equations used here are in a flux form based on the hydrostatic pressure <span class="hlt">vertical</span> coordinate, which are the same as those used in the Weather Research and Forecasting (WRF) model, but a hybrid sigma-pressure <span class="hlt">vertical</span> coordinate is implemented in this model. We verified the model by conducting widely used standard benchmark tests: the inertia-gravity wave, <span class="hlt">rising</span> thermal bubble, density current wave, and linear hydrostatic mountain wave. The results from those tests demonstrate that the horizontally spectral element <span class="hlt">vertically</span> finite difference model is accurate and robust. By using the 2-D slice model, we effectively show that the combined spatial discretization method of the spectral element and finite difference method in the horizontal and <span class="hlt">vertical</span> directions, respectively, offers a viable method for the development of a NH dynamical core.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_9 --> <div id="page_10" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="181"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFMED23A0712M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFMED23A0712M"><span>Hi<span class="hlt">RISE</span>: The People's Camera</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McEwen, A. S.; Eliason, E.; Gulick, V. C.; Spinoza, Y.; Beyer, R. A.; HiRISE Team</p> <p>2010-12-01</p> <p>The High Resolution Imaging Science Experiment (Hi<span class="hlt">RISE</span>) camera, orbiting Mars since 2006 on the Mars Reconnaissance Orbiter (MRO), has returned more than 17,000 large images with scales as small as 25 cm/pixel. From it’s beginning, the Hi<span class="hlt">RISE</span> team has followed “The People’s Camera” concept, with rapid release of useful images, explanations, and tools, and facilitating public image suggestions. The camera includes 14 CCDs, each read out into 2 data channels, so compressed images are returned from MRO as 28 long (up to 120,000 line) images that are 1024 pixels wide (or binned 2x2 to 512 pixels, etc.). This raw data is very difficult to use, especially for the public. At the Hi<span class="hlt">RISE</span> operations center the raw data are calibrated and processed into a series of B&W and color products, including browse images and JPEG2000-compressed images and tools to make it easy for everyone to explore these enormous images (see http://hirise.lpl.arizona.edu/). Automated pipelines do all of this processing, so we can keep up with the high data rate; images go directly to the format of the Planetary Data System (PDS). After students visually check each image product for errors, they are fully released just <span class="hlt">1</span> month after receipt; captioned images (written by science team members) may be released sooner. These processed Hi<span class="hlt">RISE</span> images have been incorporated into tools such as Google Mars and World Wide Telescope for even greater accessibility. 51 Digital Terrain Models derived from Hi<span class="hlt">RISE</span> stereo pairs have been released, resulting in some spectacular flyover movies produced by members of the public and viewed up to 50,000 times according to YouTube. Public targeting began in 2007 via NASA Quest (http://marsoweb.nas.nasa.gov/Hi<span class="hlt">RISE</span>/quest/) and more than 200 images have been acquired, mostly by students and educators. At the beginning of 2010 we released HiWish (http://www.uahirise.org/hiwish/), opening Hi<span class="hlt">RISE</span> targeting to anyone in the world with Internet access, and already more</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E3SWC..3301002G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E3SWC..3301002G"><span>Spatially organized «<span class="hlt">vertical</span> city» as a synthesis of tall buildings and airships</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gagulina, Olga; Matovnikov, Sergei</p> <p>2018-03-01</p> <p>The paper explores the compact city concept based on the «spatial» urban development principles and describes the prerequisites and possible methods to move from «horizontal» planning to «<span class="hlt">vertical</span>» urban environments. It highlights the close connection between urban space, high-<span class="hlt">rise</span> city landscape and conveyance options and sets out the ideas for upgrading the existing architectural and urban planning principles. It also conceptualizes the use of airships to create additional spatial connections between urban structure elements and high-<span class="hlt">rise</span> buildings. Functional changes are considered in creating both urban environment and internal space of tall buildings, and the environmental aspects of the new spatial model are brought to light. The paper delineates the prospects for making a truly «spatial» multidimensional city space.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMOS23B1400T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMOS23B1400T"><span>Variability of Relative Sea Level <span class="hlt">Rise</span>: Spatial and Temporal Correlations in Northwest Gulf of Mexico</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tissot, P.; Reisinger, A. S.; Besonen, M. R.</p> <p>2017-12-01</p> <p>While our understanding of global sea level <span class="hlt">rise</span> and its budget has made great progress over the past decade, the spatial and temporal variability of relative sea level <span class="hlt">rise</span> along the coasts still needs to be better understood and quantified. We developed a technique to reduce the confidence intervals associated with relative sea level <span class="hlt">rise</span> (RSLR) estimates for 15 tide gauges located along the Texas coast for the period 1993-2016. Seasonally detrended monthly mean water levels are highly correlated after removal of station-specific RSLR trends, which allows for the quantification of a common, low frequency oceanic signal. RSLR confidence intervals are reduced from over <span class="hlt">1</span>.9 mm/yr, on average 2.3mm, to less than <span class="hlt">1.1</span> mm/yr, on average 0.7 mm/yr after removing this common signal. The resulting RSLR rates range from 3.0 to 8.4 mm/yr. The range is wider than the longer-term rates of 5.3, 3.8 and <span class="hlt">1</span>.9 mm/yr measured from north to south by the three National Water Level Observation Network (NWLON) stations covering the study area (over different and longer time spans). The results emphasize the importance of the spatial variability of the <span class="hlt">vertical</span> land motion component of RSLR. The temporal variability of the coherent oceanic signal is not significantly correlated to the ENSO signal for the study period and is only weakly correlated to the AMO and PDO climate indices. The coherence of the signal is further investigated by comparison with other locations along the Gulf of Mexico and along the Northeast Atlantic coast. The results are discussed while considering strong local processes along the Northwest Gulf of Mexico, such as wind forcing and intermittent eddies and the spatially broader influence of the Gulf Stream. The local significance of the RSLR spatial and temporal differences are discussed in terms of the differences in inundation frequency for nuisance type flooding including comparing the time span to reach a probability of at least one nuisance flood event per</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014GeoRL..41..820K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014GeoRL..41..820K"><span>Evidence for coral island formation during <span class="hlt">rising</span> sea level in the central Pacific Ocean</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kench, Paul S.; Owen, Susan D.; Ford, Murray R.</p> <p>2014-02-01</p> <p>The timing and evolution of Jabat Island, Marshall Islands, was investigated using morphostratigraphic analysis and radiometric dating. Results show the first evidence of island building in the Pacific during latter stages of Holocene sea level <span class="hlt">rise</span>. A three-phase model of development of Jabat is presented. Initially, rapid accumulation of coarse sediments on Jabat occurred 4800-4000 years B.P. across a reef flat higher than present level, as sea level continued to <span class="hlt">rise</span>. During the highstand, island margins and particularly the western margin accreted <span class="hlt">vertically</span> to 2.5-3.0 m above contemporary ridge elevations. This accumulation phase was dominated by sand-size sediments. Phase three involved deposition of gravel ridges on the northern reef, as sea level fell to present position. Jabat has remained geomorphically stable for the past 2000 years. Findings suggest reef platforms may accommodate the oldest reef islands in atoll systems, which may have profound implications for questions of prehistoric migration through Pacific archipelagos.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.3637J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.3637J"><span>Sea level <span class="hlt">rise</span> with warming above 2 degree</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jevrejeva, Svetlana; Jackson, Luke; Riva, Riccardo; Grinsted, Aslak; Moore, John</p> <p>2017-04-01</p> <p>Holding the increase in the global average temperature to below 2 °C above pre-industrial levels, and pursuing efforts to limit the temperature increase to <span class="hlt">1</span>.5 °C, has been agreed by the representatives of the 196 parties of United Nations, as an appropriate threshold beyond which climate change risks become unacceptably high. Sea level <span class="hlt">rise</span> is one of the most damaging aspects of warming climate for the more than 600 million people living in low-elevation coastal areas less than 10 meters above sea level. Fragile coastal ecosystems and increasing concentrations of population and economic activity in coastal areas, are reasons why future sea level <span class="hlt">rise</span> is one of the most damaging aspects of the warming climate. Furthermore, sea level is set to continue to <span class="hlt">rise</span> for centuries after greenhouse gas emissions concentrations are stabilised due to system inertia and feedback time scales. Impact, risk, adaptation policies and long-term decision making in coastal areas depend on regional and local sea level <span class="hlt">rise</span> projections and local projections can differ substantially from the global one. Here we provide probabilistic sea level <span class="hlt">rise</span> projections for the global coastline with warming above the 2 degree goal. A warming of 2°C makes global ocean <span class="hlt">rise</span> on average by 20 cm, but more than 90% of coastal areas will experience greater <span class="hlt">rises</span>, 40 cm along the Atlantic coast of North America and Norway, due to ocean dynamics. If warming continues above 2°C, then by 2100 sea level will <span class="hlt">rise</span> with speeds unprecedented throughout human civilization, reaching 0.9 m (median), and 80% of the global coastline will exceed the global ocean sea level <span class="hlt">rise</span> upper 95% confidence limit of <span class="hlt">1</span>.8 m. Coastal communities of rapidly expanding cities in the developing world, small island states, and vulnerable tropical coastal ecosystems will have a very limited time after mid-century to adapt to sea level <span class="hlt">rises</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28533403','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28533403"><span>Reassessment of 20th century global mean sea level <span class="hlt">rise</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dangendorf, Sönke; Marcos, Marta; Wöppelmann, Guy; Conrad, Clinton P; Frederikse, Thomas; Riva, Riccardo</p> <p>2017-06-06</p> <p>The rate at which global mean sea level (GMSL) rose during the 20th century is uncertain, with little consensus between various reconstructions that indicate rates of <span class="hlt">rise</span> ranging from <span class="hlt">1</span>.3 to 2 mm⋅y -<span class="hlt">1</span> Here we present a 20th-century GMSL reconstruction computed using an area-weighting technique for averaging tide gauge records that both incorporates up-to-date observations of <span class="hlt">vertical</span> land motion (VLM) and corrections for local geoid changes resulting from ice melting and terrestrial freshwater storage and allows for the identification of possible differences compared with earlier attempts. Our reconstructed GMSL trend of <span class="hlt">1.1</span> ± 0.3 mm⋅y -<span class="hlt">1</span> (<span class="hlt">1</span>σ) before 1990 falls below previous estimates, whereas our estimate of 3.<span class="hlt">1</span> ± <span class="hlt">1</span>.4 mm⋅y -<span class="hlt">1</span> from 1993 to 2012 is consistent with independent estimates from satellite altimetry, leading to overall acceleration larger than previously suggested. This feature is geographically dominated by the Indian Ocean-Southern Pacific region, marking a transition from lower-than-average rates before 1990 toward unprecedented high rates in recent decades. We demonstrate that VLM corrections, area weighting, and our use of a common reference datum for tide gauges may explain the lower rates compared with earlier GMSL estimates in approximately equal proportion. The trends and multidecadal variability of our GMSL curve also compare well to the sum of individual contributions obtained from historical outputs of the Coupled Model Intercomparison Project Phase 5. This, in turn, increases our confidence in process-based projections presented in the Fifth Assessment Report of the Intergovernmental Panel on Climate Change.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1349950-thin-polymer-films-continuous-vertically-aligned-nm-pores-fabricated-soft-confinement','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1349950-thin-polymer-films-continuous-vertically-aligned-nm-pores-fabricated-soft-confinement"><span>Thin Polymer Films with Continuous <span class="hlt">Vertically</span> Aligned <span class="hlt">1</span> nm Pores Fabricated by Soft Confinement</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Feng, Xunda; Nejati, Siamak; Cowan, Matthew G.; ...</p> <p>2015-12-03</p> <p>Membrane separations are critically important in areas ranging from health care and analytical chemistry to bioprocessing and water purification. An ideal nanoporous membrane would consist of a thin film with physically continuous and <span class="hlt">vertically</span> aligned nanopores and would display a narrow distribution of pore sizes. However, the current state of the art departs considerably from this ideal and is beset by intrinsic trade-offs between permeability and selectivity. We demonstrate an effective and scalable method to fabricate polymer films with ideal membrane morphologies consisting of submicron thickness films with physically continuous and <span class="hlt">vertically</span> aligned <span class="hlt">1</span> nm pores. The approach is basedmore » on soft confinement to control the orientation of a cross-linkable mesophase in which the pores are produced by self-assembly. The scalability, exceptional ease of fabrication, and potential to create a new class of nanofiltration membranes stand out as compelling aspects.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70012202','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70012202"><span>Volcanic rocks cored on hess <span class="hlt">rise</span>, Western Pacific Ocean</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Vallier, T.L.; Windom, K.E.; Seifert, K.E.; Thiede, Jorn</p> <p>1980-01-01</p> <p>Large aseismic <span class="hlt">rises</span> and plateaus in the western Pacific include the Ontong-Java Plateau, Magellan <span class="hlt">Rise</span>, Shatsky <span class="hlt">Rise</span>, Mid-Pacific Mountains, and Hess <span class="hlt">Rise</span>. These are relatively old features that <span class="hlt">rise</span> above surrounding sea floors as bathymetric highs. Thick sequences of carbonate sediments overlie, what are believed to be, Upper Jurassic and Lower Cretaceous volcanic pedestals. We discuss here petrological and tectonic implications of data from volcanic rocks cored on Hess <span class="hlt">Rise</span>. The data suggest that Hess <span class="hlt">Rise</span> originated at a spreading centre in the late early Cretaceous (Aptian-Albian stages). Subsequent off-ridge volcanism in the late Albian-early Cenomanian stages built a large archipelago of oceanic islands and seamounts composed, at least in part, of alkalic rocks. The volcanic platform subsided during its northward passage through the mid-Cretaceousequatorial zone. Faulting and uplift, and possibly volcanism, occurred in the latest Cretaceous (Campanian-Maastrichtian stages). Since then, Hess <span class="hlt">Rise</span> continued its northward movement and subsidence. Volcanic rocks from holes drilled on Hess <span class="hlt">Rise</span> during IPOD Leg 62 (Fig. <span class="hlt">1</span>) are briefly described here and we relate the petrological data to the origin and evolution of that <span class="hlt">rise</span>. These are the first volcanic rocks reported from Hess <span class="hlt">Rise</span>. ?? 1980 Nature Publishing Group.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AdWR..114..135S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AdWR..114..135S"><span><span class="hlt">Rising</span> tides, <span class="hlt">rising</span> gates: The complex ecogeomorphic response of coastal wetlands to sea-level <span class="hlt">rise</span> and human interventions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sandi, Steven G.; Rodríguez, José F.; Saintilan, Neil; Riccardi, Gerardo; Saco, Patricia M.</p> <p>2018-04-01</p> <p>Coastal wetlands are vulnerable to submergence due to sea-level <span class="hlt">rise</span>, as shown by predictions of up to 80% of global wetland loss by the end of the century. Coastal wetlands with mixed mangrove-saltmarsh vegetation are particularly vulnerable because sea-level <span class="hlt">rise</span> can promote mangrove encroachment on saltmarsh, reducing overall wetland biodiversity. Here we use an ecogeomorphic framework that incorporates hydrodynamic effects, mangrove-saltmarsh dynamics, and soil accretion processes to assess the effects of control structures on wetland evolution. Migration and accretion patterns of mangrove and saltmarsh are heavily dependent on topography and control structures. We find that current management practices that incorporate a fixed gate for the control of mangrove encroachment are useful initially, but soon become ineffective due to sea-level <span class="hlt">rise</span>. Raising the gate, to counteract the effects of sea level <span class="hlt">rise</span> and promote suitable hydrodynamic conditions, excludes mangrove and maintains saltmarsh over the entire simulation period of 100 years</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11829311','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11829311"><span>The psychophysical periphery effect crosses the <span class="hlt">vertical</span> meridian.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kuyk, T; Niculescu, D</p> <p>2001-01-01</p> <p>This study measured the periphery effect and compared its magnitude when the peripheral stimulation was on the same or opposite side of the <span class="hlt">vertical</span> meridian as the test spot. Test thresholds for a <span class="hlt">1</span>.5-deg diameter, 8-ms spot located <span class="hlt">1</span>.75 deg to one side of the <span class="hlt">vertical</span> meridian were elevated by approximately 0.125 log units when a 0.25 cycles/deg (cpd) counterphased grating was presented at a similar eccentric offset on the other side of the <span class="hlt">vertical</span> meridian. The periphery effect disappeared when the test spot was moved outward to 8-deg eccentricity. When the grating and test were presented on the same side of the <span class="hlt">vertical</span> meridian, test thresholds at both retinal locations were elevated by the same amount, 0.2 log units. Consistent with the physiology in cat retina, the periphery effect in humans also crosses over the <span class="hlt">vertical</span> meridian. However, the effect is small and the test spot must be in close proximity to the <span class="hlt">vertical</span> meridian for it to be observed. Also, the crossover periphery effect is reduced in magnitude by 37.5% compared to when the grating and test are presented on the same side of the <span class="hlt">vertical</span> meridian. This suggests there may be a difference in how the underlying neural mechanism that transmits the periphery effect signal laterally is organized for sending the periphery effect signal across the <span class="hlt">vertical</span> meridian as compared to within a retinal hemifield.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MS%26E..338a2010A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MS%26E..338a2010A"><span>Pore Size Control in Aluminium Foam by Standardizing Bubble <span class="hlt">Rise</span> Velocity and Melt Viscosity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Avinash, G.; Harika, V.; Sandeepika, Ch; Gupta, N.</p> <p>2018-03-01</p> <p>In recent years, aluminium foams have found use in a wide range of applications. The properties of these foams, as good structural strength with light weight have made them as a promising structural material for aerospace industry. Foaming techniques (direct and indirect) are used to produce these foams. Direct foaming involves blowing of gas to create gas bubbles in the melt whereas indirect foaming technique uses blowing agents as metallic hydrides, which create hydrogen bubbles. Porosity and its distribution in foams directly affect its properties. This demands for more theoretical studies, to control such cellular structure and hence properties. In present work, we have studied the effect of gas bubble <span class="hlt">rise</span> velocity and melt viscosity, on pore size and its distribution in aluminium foam. A 15 PPI aluminium foam, prepared using indirect foaming technique having porosity ~86 % was used for study. In order to obtain metal foam, the bubble must not escape from the melt and should get entrapped during solidification. Our calculations suggest that bubble <span class="hlt">rise</span> velocity and melt viscosity are responsible for <span class="hlt">vertical</span> displacement of bubble in the melt. It is observed that melt viscosity opposes bubble <span class="hlt">rise</span> velocity and help the bubbles to stay in the melt, resulting in porous structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E3SWC..3301024V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E3SWC..3301024V"><span>Use of high-<span class="hlt">rise</span> structures for sustainable tourism</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vavilova, Tatiana Ya.; Vyshkin, Efim G.</p> <p>2018-03-01</p> <p>The paper deals with such issues as formation and development of the infrastructure of objects for serving tourists in urban environment and specially protected natural areas with particular focus on open tower structures - a type of object which is so popular in Russia. The authors systematize international experience of integrating watchtowers in natural and anthropogenic environment as well as specific features of their modern architectural solutions. A number of examples are given. Summing up the results of the analysis we have come to conclusion that in the field of tourism the most promising tendency in functional use of <span class="hlt">vertical</span> structures is the demonstration of cultural and natural attractions. It is also noted that in national and natural parks objects of the tower type can be built for other purposes, e.g. for conducting research, monitoring weather conditions and emergency situations. It is shown that the development of infrastructure of high-<span class="hlt">rise</span> buildings for educational tourism contributes to sustainable development of territories and settlements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1994WRR....30.3275L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1994WRR....30.3275L"><span>Water movement in glass bead porous media: <span class="hlt">1</span>. Experiments of capillary <span class="hlt">rise</span> and hysteresis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lu, T. X.; Biggar, J. W.; Nielsen, D. R.</p> <p>1994-12-01</p> <p>Experimental observations of capillary <span class="hlt">rise</span> and hysteresis of water or ethanol in glass beads are presented to improve our understanding of those physical processes in porous media. The results provide evidence that capillary <span class="hlt">rise</span> into porous media cannot be fully explained by a model of cylinders. They further demonstrate that the "Ink bottle" model does not provide an adequate explanation of hysteresis. Glass beads serving as a model for ideal soil are enclosed in a rectangular glass chamber model. A TV camera associated with a microscope was used to record the processes of capillary <span class="hlt">rise</span> and drainage. It is clearly shown during capillary <span class="hlt">rise</span> that the fluid exhibits a "jump" behavior at the neck of the pores in an initially dry profile or at the bottom of the water film in an initially wet profile. Under an initially dry condition, the jump initiates at the particle with smallest diameter. The jump process continues to higher elevations until at equilibrium the surface tensile force is balanced by the hydrostatic force. The wetting front at that time is readily observed as flat and saturated. Under an initially wet condition, capillary <span class="hlt">rise</span> occurs as a water film thickening process associated with the jump process. Trapped air behind the wetting front renders the wetting front irregular and unsaturated. The capillary <span class="hlt">rise</span> into an initially wet porous medium can be higher than that into an initially dry profile. During the drying process, large surface areas associated with the gas-liquid interface develop, allowing the porous medium to retain more water than during the wetting process at the same pressure. That mechanism explains better the hysteresis phenomenon in porous media in contrast to other mechanisms that now prevail.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED337607.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED337607.pdf"><span>Voices <span class="hlt">Rising</span>: A Bulletin about Women and Popular Education. Volumes <span class="hlt">1</span>-4. 1987-1990.</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Voices Rising: A Bulletin about Women and Popular Education, 1990</p> <p>1990-01-01</p> <p>This document consists of the six issues of the serial "Voices <span class="hlt">Rising</span>" issued during the four-year period 1987-1990. "Voices <span class="hlt">Rising</span>" is the primary networking tool of the Women's Program of the International Council for Adult Education (ICAE). Articles in these issues include: "Tribute to a Courageous Woman--Nabila Breir"; "Centre for Women's…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5332610','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5332610"><span>Stability of <span class="hlt">vertical</span> magnetic chains</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p></p> <p>2017-01-01</p> <p>A linear stability analysis is performed for a pair of coaxial <span class="hlt">vertical</span> chains made from permanently magnetized balls under the influence of gravity. While one chain <span class="hlt">rises</span> from the ground, the other hangs from above, with the remaining ends separated by a gap of prescribed length. Various boundary conditions are considered, as are situations in which the magnetic dipole moments in the two chains are parallel or antiparallel. The case of a single chain attached to the ground is also discussed. The stability of the system is examined with respect to three quantities: the number of balls in each chain, the length of the gap between the chains, and a single dimensionless parameter which embodies the competition between magnetic and gravitational forces. Asymptotic scaling laws involving these parameters are provided. The Hessian matrix is computed in exact form, allowing the critical parameter values at which the system loses stability and the respective eigenmodes to be determined up to machine precision. A comparison with simple experiments for a single chain attached to the ground shows good agreement. PMID:28293135</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017RSPSA.47360703S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017RSPSA.47360703S"><span>Stability of <span class="hlt">vertical</span> magnetic chains</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schönke, Johannes; Fried, Eliot</p> <p>2017-02-01</p> <p>A linear stability analysis is performed for a pair of coaxial <span class="hlt">vertical</span> chains made from permanently magnetized balls under the influence of gravity. While one chain <span class="hlt">rises</span> from the ground, the other hangs from above, with the remaining ends separated by a gap of prescribed length. Various boundary conditions are considered, as are situations in which the magnetic dipole moments in the two chains are parallel or antiparallel. The case of a single chain attached to the ground is also discussed. The stability of the system is examined with respect to three quantities: the number of balls in each chain, the length of the gap between the chains, and a single dimensionless parameter which embodies the competition between magnetic and gravitational forces. Asymptotic scaling laws involving these parameters are provided. The Hessian matrix is computed in exact form, allowing the critical parameter values at which the system loses stability and the respective eigenmodes to be determined up to machine precision. A comparison with simple experiments for a single chain attached to the ground shows good agreement.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1984Sci...225..333M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1984Sci...225..333M"><span>Adsorption to Fish Sperm of <span class="hlt">Vertically</span> Transmitted Fish Viruses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mulcahy, Dan; Pascho, Ronald J.</p> <p>1984-07-01</p> <p>More than 99 percent of a <span class="hlt">vertically</span> transmitted fish rhabdovirus, infectious hematopoietic necrosis virus, was removed from suspension in less than <span class="hlt">1</span> minute by adsorption to the surface membrane of sperm from two genera of salmonid fishes. The <span class="hlt">vertically</span> transmitted, infectious pancreatic necrosis virus adsorbed to a lesser degree, but no adsorption occurred with a second fish rhabdovirus that is not <span class="hlt">vertically</span> transmitted. Such adsorption may be involved in <span class="hlt">vertical</span> transmission of these viruses.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70162168','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70162168"><span>Adsorption to fish sperm of <span class="hlt">vertically</span> transmitted fish viruses</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Mulcahy, D.; Pascho, R.J.</p> <p>1984-01-01</p> <p>More than 99 percent of a <span class="hlt">vertically</span> transmitted fish rhabdovirus, infectious hematopoietic necrosis virus, was removed from suspension in less than <span class="hlt">1</span> minute by adsorption to the surface membrane of sperm from two genera of salmonid fishes. The <span class="hlt">vertically</span> transmitted, infectious pancreatic necrosis virus adsorbed to a lesser degree, but no adsorption occurred with a second fish rhabdovirus that is not <span class="hlt">vertically</span> transmitted. Such adsorption may be involved in <span class="hlt">vertical</span> transmission of these viruses.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22630356','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22630356"><span>Orienting numbers in mental space: horizontal organization trumps <span class="hlt">vertical</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Holmes, Kevin J; Lourenco, Stella F</p> <p>2012-01-01</p> <p>While research on the spatial representation of number has provided substantial evidence for a horizontally oriented mental number line, recent studies suggest <span class="hlt">vertical</span> organization as well. Directly comparing the relative strength of horizontal and <span class="hlt">vertical</span> organization, however, we found no evidence of spontaneous <span class="hlt">vertical</span> orientation (upward or downward), and horizontal trumped <span class="hlt">vertical</span> when pitted against each other (Experiment <span class="hlt">1</span>). Only when numbers were conceptualized as magnitudes (as opposed to nonmagnitude ordinal sequences) did reliable <span class="hlt">vertical</span> organization emerge, with upward orientation preferred (Experiment 2). Altogether, these findings suggest that horizontal representations predominate, and that <span class="hlt">vertical</span> representations, when elicited, may be relatively inflexible. Implications for spatial organization beyond number, and its ontogenetic basis, are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JMMM..456..300V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JMMM..456..300V"><span>Onset of thermomagnetic convection around a <span class="hlt">vertically</span> oriented hot-wire in ferrofluid</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vatani, Ashkan; Woodfield, Peter Lloyd; Nguyen, Nam-Trung; Dao, Dzung Viet</p> <p>2018-06-01</p> <p>The onset of thermomagnetic convection in ferrofluid in a <span class="hlt">vertical</span> transient hot-wire cell is analytically and experimentally investigated by studying the temperature <span class="hlt">rise</span> of an electrically-heated wire. During the initial stage of heating, the temperature <span class="hlt">rise</span> is found to correspond well to that predicted by conduction only. For high electrical current densities, the initial heating stage is followed by a sudden change in the slope of the temperature <span class="hlt">rise</span> with respect to time as a result of the onset of thermomagnetic convection cooling. The observed onset of thermomagnetic convection was then compared to that of natural convection of deionized water. For the first time, the critical time corresponding to the onset of thermomagnetic convection around an electrically-heated wire is characterized and non-dimensionalized as a critical Fourier number (Foc). We propose an equation for Foc as a function of a magnetic Rayleigh number to predict the time for the onset of thermomagnetic convection. We observed that thermomagnetic convection in ferrofluid occurs earlier than natural convection in non-magnetic fluids for similar experimental conditions. The onset of thermomagnetic convection is dependent on the current supplied to the wire. The findings have important implications for cooling of high-power electronics using ferrofluids and for measuring thermal properties of ferrofluids.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_10 --> <div id="page_11" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="201"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70169087','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70169087"><span>Evaluation of dynamic coastal response to sea-level <span class="hlt">rise</span> modifies inundation likelihood</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Lentz, Erika E.; Thieler, E. Robert; Plant, Nathaniel G.; Stippa, Sawyer R.; Horton, Radley M.; Gesch, Dean B.</p> <p>2016-01-01</p> <p>Sea-level <span class="hlt">rise</span> (SLR) poses a range of threats to natural and built environments<span class="hlt">1</span>, 2, making assessments of SLR-induced hazards essential for informed decision making3. We develop a probabilistic model that evaluates the likelihood that an area will inundate (flood) or dynamically respond (adapt) to SLR. The broad-area applicability of the approach is demonstrated by producing 30 × 30 m resolution predictions for more than 38,000 km2 of diverse coastal landscape in the northeastern United States. Probabilistic SLR projections, coastal elevation and <span class="hlt">vertical</span> land movement are used to estimate likely future inundation levels. Then, conditioned on future inundation levels and the current land-cover type, we evaluate the likelihood of dynamic response versus inundation. We find that nearly 70% of this coastal landscape has some capacity to respond dynamically to SLR, and we show that inundation models over-predict land likely to submerge. This approach is well suited to guiding coastal resource management decisions that weigh future SLR impacts and uncertainty against ecological targets and economic constraints.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AMT....11.2937W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AMT....11.2937W"><span>Derivation of gravity wave intrinsic parameters and <span class="hlt">vertical</span> wavelength using a single scanning OH(3-<span class="hlt">1</span>) airglow spectrometer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wüst, Sabine; Offenwanger, Thomas; Schmidt, Carsten; Bittner, Michael; Jacobi, Christoph; Stober, Gunter; Yee, Jeng-Hwa; Mlynczak, Martin G.; Russell, James M., III</p> <p>2018-05-01</p> <p>For the first time, we present an approach to derive zonal, meridional, and <span class="hlt">vertical</span> wavelengths as well as periods of gravity waves based on only one OH* spectrometer, addressing one vibrational-rotational transition. Knowledge of these parameters is a precondition for the calculation of further information, such as the wave group velocity vector.OH(3-<span class="hlt">1</span>) spectrometer measurements allow the analysis of gravity wave ground-based periods but spatial information cannot necessarily be deduced. We use a scanning spectrometer and harmonic analysis to derive horizontal wavelengths at the mesopause altitude above Oberpfaffenhofen (48.09° N, 11.28° E), Germany for 22 nights in 2015. Based on the approximation of the dispersion relation for gravity waves of low and medium frequencies and additional horizontal wind information, we calculate <span class="hlt">vertical</span> wavelengths. The mesopause wind measurements nearest to Oberpfaffenhofen are conducted at Collm (51.30° N, 13.02° E), Germany, ca. 380 km northeast of Oberpfaffenhofen, by a meteor radar.In order to compare our results, <span class="hlt">vertical</span> temperature profiles of TIMED-SABER (thermosphere ionosphere mesosphere energetics dynamics, sounding of the atmosphere using broadband emission radiometry) overpasses are analysed with respect to the dominating <span class="hlt">vertical</span> wavelength.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1912765H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1912765H"><span>Probabilistic reconstruction of GPS <span class="hlt">vertical</span> ground motion and comparison with GIA models</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Husson, Laurent; Bodin, Thomas; Choblet, Gael; Kreemer, Corné</p> <p>2017-04-01</p> <p>The <span class="hlt">vertical</span> position time-series of GPS stations have become long enough for many parts of the world to infer modern rates of <span class="hlt">vertical</span> ground motion. We use the worldwide compilation of GPS trend velocities of the Nevada Geodetic Laboratory. Those rates are inferred by applying the MIDAS algorithm (Blewitt et al., 2016) to time-series obtained from publicly available data from permanent stations. Because MIDAS filters out seasonality and discontinuities, regardless of their causes, it gives robust long-term rates of <span class="hlt">vertical</span> ground motion (except where there is significant postseismic deformation). As the stations are unevenly distributed, and because data errors are also highly variable, sometimes to an unknown degree, we use a Bayesian inference method to reconstruct 2D maps of <span class="hlt">vertical</span> ground motion. Our models are based on a Voronoi tessellation and self-adapt to the spatially variable level of information provided by the data. Instead of providing a unique interpolated surface, each point of the reconstructed surface is defined through a probability density function. We apply our method to a series of vast regions covering entire continents. Not surprisingly, the reconstructed surface at a long wavelength is dominated by the GIA. This result can be exploited to evaluate whether forward models of GIA reproduce geodetic rates within the uncertainties derived from our interpolation, not only at high latitudes where postglacial rebound is fast, but also in more temperate latitudes where, for instance, such rates may compete with modern sea level <span class="hlt">rise</span>. At shorter wavelengths, the reconstructed surface of <span class="hlt">vertical</span> ground motion features a variety of identifiable patterns, whose geometries and rates can be mapped. Examples are transient dynamic topography over the convecting mantle, actively deforming domains (mountain belts and active margins), volcanic areas, or anthropogenic contributions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1016505','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1016505"><span><span class="hlt">Vertical</span> axis wind turbines</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Krivcov, Vladimir [Miass, RU; Krivospitski, Vladimir [Miass, RU; Maksimov, Vasili [Miass, RU; Halstead, Richard [Rohnert Park, CA; Grahov, Jurij [Miass, RU</p> <p>2011-03-08</p> <p>A <span class="hlt">vertical</span> axis wind turbine is described. The wind turbine can include a top ring, a middle ring and a lower ring, wherein a plurality of <span class="hlt">vertical</span> airfoils are disposed between the rings. For example, three <span class="hlt">vertical</span> airfoils can be attached between the upper ring and the middle ring. In addition, three more <span class="hlt">vertical</span> airfoils can be attached between the lower ring and the middle ring. When wind contacts the <span class="hlt">vertically</span> arranged airfoils the rings begin to spin. By connecting the rings to a center pole which spins an alternator, electricity can be generated from wind.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.G51B..02D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.G51B..02D"><span>Detecting anthropogenic footprints in regional and global sea level <span class="hlt">rise</span> since 1900</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dangendorf, S.; Marcos, M.; Piecuch, C. G.; Jensen, J.</p> <p>2015-12-01</p> <p>While there is scientific consensus that global and local mean sea level (GMSL and LMSL) is <span class="hlt">rising</span> since the late 19th century, it remains unclear how much of this <span class="hlt">rise</span> is due to natural variability or anthropogenic forcing. Distinguishing both contributions requires an extensive knowledge about the persistence of natural high and low stands in GMSL and LMSL. This is challenging, since observational time series represent the superposition of various processes with different spectral properties. Here we provide a probabilistic upper range of long-term persistent natural GMSL/LMSL variability (P=0.99), which in turn determines the minimum/maximum anthropogenic contribution since 1900. To account for different spectral characteristics of various contributing processes, we separate LMSL (corrected for <span class="hlt">vertical</span> land motion) into a slowly varying volumetric (mass and density changes) and a more rapidly changing atmospheric component. Based on a combination of spectral analyses of tide gauge records, barotropic and baroclinic ocean models and numerical Monte-Carlo experiments, we find that in records where transient atmospheric processes dominate the spectra, the persistence of natural volumetric changes tends to be underestimated. If each component is assessed separately, natural centennial trends are locally up to ~<span class="hlt">1</span>.0 mm/yr larger than in case of an integrated assessment, therefore erroneously enhancing the significance of anthropogenic footprints. The GMSL, however, remains unaffected by such biases. On the basis of a model assessment of the separate components, we conclude that it is virtually certain (P=0.99) that at least 45% of the observed increase in GMSL is of anthropogenic origin.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Natur.549..257K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Natur.549..257K"><span>Impact of a global temperature <span class="hlt">rise</span> of <span class="hlt">1</span>.5 degrees Celsius on Asia’s glaciers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kraaijenbrink, P. D. A.; Bierkens, M. F. P.; Lutz, A. F.; Immerzeel, W. W.</p> <p>2017-09-01</p> <p>Glaciers in the high mountains of Asia (HMA) make a substantial contribution to the water supply of millions of people, and they are retreating and losing mass as a result of anthropogenic climate change at similar rates to those seen elsewhere. In the Paris Agreement of 2015, 195 nations agreed on the aspiration to limit the level of global temperature <span class="hlt">rise</span> to <span class="hlt">1</span>.5 degrees Celsius ( °C) above pre-industrial levels. However, it is not known what an increase of <span class="hlt">1</span>.5 °C would mean for the glaciers in HMA. Here we show that a global temperature <span class="hlt">rise</span> of <span class="hlt">1</span>.5 °C will lead to a warming of 2.<span class="hlt">1</span> ± 0.<span class="hlt">1</span> °C in HMA, and that 64 ± 7 per cent of the present-day ice mass stored in the HMA glaciers will remain by the end of the century. The <span class="hlt">1</span>.5 °C goal is extremely ambitious and is projected by only a small number of climate models of the conservative IPCC’s Representative Concentration Pathway (RCP)2.6 ensemble. Projections for RCP4.5, RCP6.0 and RCP8.5 reveal that much of the glacier ice is likely to disappear, with projected mass losses of 49 ± 7 per cent, 51 ± 6 per cent and 64 ± 5 per cent, respectively, by the end of the century; these projections have potentially serious consequences for regional water management and mountain communities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhFl...28i3301J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhFl...28i3301J"><span><span class="hlt">Rise</span> of an argon bubble in liquid steel in the presence of a transverse magnetic field</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jin, K.; Kumar, P.; Vanka, S. P.; Thomas, B. G.</p> <p>2016-09-01</p> <p>The <span class="hlt">rise</span> of gaseous bubbles in viscous liquids is a fundamental problem in fluid physics, and it is also a common phenomenon in many industrial applications such as materials processing, food processing, and fusion reactor cooling. In this work, the motion of a single argon gas bubble <span class="hlt">rising</span> in quiescent liquid steel under an external magnetic field is studied numerically using a Volume-of-Fluid method. To mitigate spurious velocities normally generated during numerical simulation of multiphase flows with large density differences, an improved algorithm for surface tension modeling, originally proposed by Wang and Tong ["Deformation and oscillations of a single gas bubble <span class="hlt">rising</span> in a narrow <span class="hlt">vertical</span> tube," Int. J. Therm. Sci. 47, 221-228 (2008)] is implemented, validated and used in the present computations. The governing equations are integrated by a second-order space and time accurate numerical scheme, and implemented on multiple Graphics Processing Units with high parallel efficiency. The motion and terminal velocities of the <span class="hlt">rising</span> bubble under different magnetic fields are compared and a reduction in <span class="hlt">rise</span> velocity is seen in cases with the magnetic field applied. The shape deformation and the path of the bubble are discussed. An elongation of the bubble along the field direction is seen, and the physics behind these phenomena is discussed. The wake structures behind the bubble are visualized and effects of the magnetic field on the wake structures are presented. A modified drag coefficient is obtained to include the additional resistance force caused by adding a transverse magnetic field.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22598826-rise-argon-bubble-liquid-steel-presence-transverse-magnetic-field','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22598826-rise-argon-bubble-liquid-steel-presence-transverse-magnetic-field"><span><span class="hlt">Rise</span> of an argon bubble in liquid steel in the presence of a transverse magnetic field</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Jin, K.; Kumar, P.; Vanka, S. P., E-mail: spvanka@illinois.edu</p> <p>2016-09-15</p> <p>The <span class="hlt">rise</span> of gaseous bubbles in viscous liquids is a fundamental problem in fluid physics, and it is also a common phenomenon in many industrial applications such as materials processing, food processing, and fusion reactor cooling. In this work, the motion of a single argon gas bubble <span class="hlt">rising</span> in quiescent liquid steel under an external magnetic field is studied numerically using a Volume-of-Fluid method. To mitigate spurious velocities normally generated during numerical simulation of multiphase flows with large density differences, an improved algorithm for surface tension modeling, originally proposed by Wang and Tong [“Deformation and oscillations of a single gasmore » bubble <span class="hlt">rising</span> in a narrow <span class="hlt">vertical</span> tube,” Int. J. Therm. Sci. 47, 221–228 (2008)] is implemented, validated and used in the present computations. The governing equations are integrated by a second-order space and time accurate numerical scheme, and implemented on multiple Graphics Processing Units with high parallel efficiency. The motion and terminal velocities of the <span class="hlt">rising</span> bubble under different magnetic fields are compared and a reduction in <span class="hlt">rise</span> velocity is seen in cases with the magnetic field applied. The shape deformation and the path of the bubble are discussed. An elongation of the bubble along the field direction is seen, and the physics behind these phenomena is discussed. The wake structures behind the bubble are visualized and effects of the magnetic field on the wake structures are presented. A modified drag coefficient is obtained to include the additional resistance force caused by adding a transverse magnetic field.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017GeoJI.211..593H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GeoJI.211..593H"><span>Subduction and <span class="hlt">vertical</span> coastal motions in the eastern Mediterranean</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Howell, Andy; Jackson, James; Copley, Alex; McKenzie, Dan; Nissen, Ed</p> <p>2017-10-01</p> <p>Convergence in the eastern Mediterranean of oceanic Nubia with Anatolia and the Aegean is complex and poorly understood. Large volumes of sediment obscure the shallow structure of the subduction zone, and since much of the convergence is accommodated aseismically, there are limited earthquake data to constrain its kinematics. We present new source models for recent earthquakes, combining these with field observations, published GPS velocities and reflection-seismic data to investigate faulting in three areas: the Florence <span class="hlt">Rise</span>, SW Turkey and the Pliny and Strabo Trenches. The depths and locations of earthquakes reveal the geometry of the subducting Nubian plate NE of the Florence <span class="hlt">Rise</span>, a bathymetric high that is probably formed by deformation of sediment at the surface projection of the Anatolia-Nubia subduction interface. In SW Turkey, the presence of a strike-slip shear zone has often been inferred despite an absence of strike-slip earthquakes. We show that the GPS-derived strain-rate field is consistent with extension on the orthogonal systems of normal faults observed in the region and that strike-slip faulting is not required to explain observed GPS velocities. Further SW, the Pliny and Strabo Trenches are also often interpreted as strike-slip shear zones, but almost all nearby earthquakes have either reverse-faulting or normal-faulting focal mechanisms. Oblique convergence across the trenches may be accommodated either by a partitioned system of strike-slip and reverse faults or by oblique slip on the Aegean-Nubia subduction interface. The observed late-Quaternary <span class="hlt">vertical</span> motions of coastlines close to the subduction zone are influenced by the interplay between: (<span class="hlt">1</span>) thickening of the material overriding the subduction interface associated with convergence, which promotes coastal uplift; and (2) subsidence due to extension and associated crustal thinning. Long-wavelength gravity data suggest that some of the observed topographic contrasts in the eastern</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/889289','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/889289"><span><span class="hlt">VERTICAL</span> BEAM SIZE CONTROL IN TLS AND TPS.</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>KUO, C.C.; CHEN, J.R.; CHOU, P.J.</p> <p>2006-06-26</p> <p><span class="hlt">Vertical</span> beam size control is an important issue in the light source operations. The horizontal-<span class="hlt">vertical</span> betatron coupling and <span class="hlt">vertical</span> dispersion were measured and corrected to small values in the TLS <span class="hlt">1</span>.5 GeV storage ring. Estimated beam sizes are compared with the measured values. By employing an effective transverse damping system, the <span class="hlt">vertical</span> beam blow-up due to transverse coherent instabilities, such as the fast-ion beam instability, was suppressed. As a result, the light source is very stable. In NSRRC we are designing an ultra low emittance 3-GeV storage ring and its designed <span class="hlt">vertical</span> beam size could be as small as amore » few microns. The ground and mechanic vibration effects, and coherent instabilities could spoil the expected photon brightness due to blow-up of the <span class="hlt">vertical</span> beam size if not well taken care of. The contributions of these effects to <span class="hlt">vertical</span> beam size increase will be evaluated and the counter measures to minimize them will be proposed and reported in this paper.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUSM.H41C..07M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUSM.H41C..07M"><span>Effects of Sea Level <span class="hlt">Rise</span> on Groundwater Flow Paths in a Coastal Aquifer System</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Morrissey, S. K.; Clark, J. F.; Bennett, M. W.; Richardson, E.; Stute, M.</p> <p>2008-05-01</p> <p>Changes in groundwater flow in the Floridan aquifer system, South Florida, from the <span class="hlt">rise</span> in sea level at the end of the last glacial period may be indicative of changes coastal aquifers will experience with continued sea level <span class="hlt">rise</span>. As sea level <span class="hlt">rises</span>, the hydraulic head near the coast increases. Coastal aquifers can therefore experience decreased groundwater gradients (increased residence times) and seawater intrusion. Stable isotopes of water, dissolved noble gas temperatures, radiocarbon and He concentrations were analyzed in water collected from 68 wells in the Floridan aquifer system throughout South Florida. Near the recharge area, geochemical data along groundwater flow paths in the Upper Floridan aquifer show a transition from recently recharged groundwater to glacial-aged water. Down gradient from this transition, little variation is apparent in the stable isotopes and noble gas recharge temperatures, indicating that most of the Upper Floridan aquifer contains groundwater recharged during the last glacial period. The rapid 120-meter <span class="hlt">rise</span> in sea level marking the end of the last glacial period increased the hydraulic head in the Floridan aquifer system near the coast, slowing the flow of groundwater from the recharge area to the ocean and trapping glacial-aged groundwater. The raised sea level also flooded half of the Florida platform and caused seawater to intrude into the Lower Floridan. This circulation of seawater in the Lower Floridan continues today as our data indicate that the groundwater is similar to modern seawater with a freshwater component entering <span class="hlt">vertically</span> from the recharge area to the Upper Floridan.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhFl...29e3302L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhFl...29e3302L"><span>Thermal stratification hinders gyrotactic micro-organism <span class="hlt">rising</span> in free-surface turbulence</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lovecchio, Salvatore; Zonta, Francesco; Marchioli, Cristian; Soldati, Alfredo</p> <p>2017-05-01</p> <p>Thermal stratification in water bodies influences the exchange of heat, momentum, and chemical species across the air-water interface by modifying the sub-surface turbulence characteristics. Turbulence modifications may in turn prevent small motile algae (phytoplankton, in particular) from reaching the heated surface. We examine how different regimes of stable thermal stratification affect the motion of these microscopic organisms (modelled as gyrotactic self-propelling cells) in a free-surface turbulent channel flow. This archetypal setup mimics an environmentally plausible situation that can be found in lakes and oceans. Results from direct numerical simulations of turbulence coupled with Lagrangian tracking reveal that <span class="hlt">rising</span> of bottom-heavy self-propelling cells depends strongly on the strength of stratification, especially near the thermocline where high temperature and velocity gradients occur: Here hydrodynamic shear may disrupt directional cell motility and hamper near-surface accumulation. For all gyrotactic re-orientation times considered in this study (spanning two orders of magnitude), we observe a reduction of the cell <span class="hlt">rising</span> speed and temporary confinement under the thermocline: If re-orientation is fast, cells eventually trespass the thermocline within the simulated time span; if re-orientation is slow, confinement lasts much longer because cells align in the streamwise direction and their <span class="hlt">vertical</span> swimming is practically annihilated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5468634','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5468634"><span>Reassessment of 20th century global mean sea level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Dangendorf, Sönke; Marcos, Marta; Wöppelmann, Guy; Conrad, Clinton P.; Frederikse, Thomas; Riva, Riccardo</p> <p>2017-01-01</p> <p>The rate at which global mean sea level (GMSL) rose during the 20th century is uncertain, with little consensus between various reconstructions that indicate rates of <span class="hlt">rise</span> ranging from <span class="hlt">1</span>.3 to 2 mm⋅y−<span class="hlt">1</span>. Here we present a 20th-century GMSL reconstruction computed using an area-weighting technique for averaging tide gauge records that both incorporates up-to-date observations of <span class="hlt">vertical</span> land motion (VLM) and corrections for local geoid changes resulting from ice melting and terrestrial freshwater storage and allows for the identification of possible differences compared with earlier attempts. Our reconstructed GMSL trend of <span class="hlt">1.1</span> ± 0.3 mm⋅y−<span class="hlt">1</span> (<span class="hlt">1</span>σ) before 1990 falls below previous estimates, whereas our estimate of 3.<span class="hlt">1</span> ± <span class="hlt">1</span>.4 mm⋅y−<span class="hlt">1</span> from 1993 to 2012 is consistent with independent estimates from satellite altimetry, leading to overall acceleration larger than previously suggested. This feature is geographically dominated by the Indian Ocean–Southern Pacific region, marking a transition from lower-than-average rates before 1990 toward unprecedented high rates in recent decades. We demonstrate that VLM corrections, area weighting, and our use of a common reference datum for tide gauges may explain the lower rates compared with earlier GMSL estimates in approximately equal proportion. The trends and multidecadal variability of our GMSL curve also compare well to the sum of individual contributions obtained from historical outputs of the Coupled Model Intercomparison Project Phase 5. This, in turn, increases our confidence in process-based projections presented in the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. PMID:28533403</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20165738','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20165738"><span>Influence of loading forces on the <span class="hlt">vertical</span> accuracy of interocclusal records.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ghazal, Muhamad; Kern, Matthias</p> <p>2010-02-01</p> <p>To evaluate the influence of loading forces on the <span class="hlt">vertical</span> discrepancies caused by interocclusal recording materials. A custom-made apparatus was used to simulate the maxilla and mandible. Eight interocclusal records were made in each of the following groups: G<span class="hlt">1</span>-Aluwax (aluminum wax; Aluwax), G2-Beauty Pink wax (hydrocarbon wax compound; Miltex), G3-Futar D, G4-Futar D Fast, G5-Futar Scan (polyvinyl siloxanes; Kettenbach), and G6-Ramitec (polyether; 3M ESPE). The <span class="hlt">vertical</span> discrepancies were measured by an inductive displacement transducer connected to a carrier frequency amplifier after storage of the records for <span class="hlt">1</span> hour at room temperature. Different compressive loading forces up to <span class="hlt">1</span> kg were applied onto the upper part of the apparatus to evaluate the influence on the <span class="hlt">vertical</span> discrepancies of the records. Two-way ANOVA was used for statistical analysis. The compressive loading force had a statistically significant influence on the <span class="hlt">vertical</span> discrepancies (P<.01) (ie, higher forces reduced the <span class="hlt">vertical</span> discrepancies). When a compressive force of <span class="hlt">1</span> kg was applied to the upper part of the apparatus, the mean <span class="hlt">vertical</span> discrepancies for G<span class="hlt">1</span> (11+/-3 microm) and G2 (12+/-3 microm) were statistically significantly higher than in groups G3 (<span class="hlt">1+/-1</span> microm), G4 (2+/-<span class="hlt">1</span> microm), G5 (0+/-<span class="hlt">1</span> microm), and G6 (-2+/-2 microm). A compressive force of <span class="hlt">1</span> kg could be used to stabilize the cast during mounting procedures in an articulator using an interocclusal record made of polyvinyl siloxane without <span class="hlt">vertically</span> changing the interocclusal relationships.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009JPSJ...78c4801H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009JPSJ...78c4801H"><span>Signatures of Currency <span class="hlt">Vertices</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Holme, Petter</p> <p>2009-03-01</p> <p>Many real-world networks have broad degree distributions. For some systems, this means that the functional significance of the <span class="hlt">vertices</span> is also broadly distributed, in other cases the <span class="hlt">vertices</span> are equally significant, but in different ways. One example of the latter case is metabolic networks, where the high-degree <span class="hlt">vertices</span> — the currency metabolites — supply the molecular groups to the low-degree metabolites, and the latter are responsible for the higher-order biological function, of vital importance to the organism. In this paper, we propose a generalization of currency metabolites to currency <span class="hlt">vertices</span>. We investigate the network structural characteristics of such systems, both in model networks and in some empirical systems. In addition to metabolic networks, we find that a network of music collaborations and a network of e-mail exchange could be described by a division of the <span class="hlt">vertices</span> into currency <span class="hlt">vertices</span> and others.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/wsp/2337/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/wsp/2337/report.pdf"><span>Use of temperature profiles beneath streams to determine rates of <span class="hlt">vertical</span> ground-water flow and <span class="hlt">vertical</span> hydraulic conductivity</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Lapham, Wayne W.</p> <p>1989-01-01</p> <p>The use of temperature profiles beneath streams to determine rates of <span class="hlt">vertical</span> ground-water flow and effective <span class="hlt">vertical</span> hydraulic conductivity of sediments was evaluated at three field sites by use of a model that numerically solves the partial differential equation governing simultaneous <span class="hlt">vertical</span> flow of fluid and heat in the Earth. The field sites are located in Hardwick and New Braintree, Mass., and in Dover, N.J. In New England, stream temperature varies from about 0 to 25 ?C (degrees Celsius) during the year. This stream-temperature fluctuation causes ground-water temperatures beneath streams to fluctuate by more than 0.<span class="hlt">1</span> ?C during a year to a depth of about 35 ft (feet) in fine-grained sediments and to a depth of about 50 ft in coarse-grained sediments, if ground-water velocity is 0 ft/d (foot per day). Upward flow decreases the depth affected by stream-temperature fluctuation, and downward flow increases the depth. At the site in Hardwick, Mass., ground-water flow was upward at a rate of less than 0.01 ft/d. The maximum effective <span class="hlt">vertical</span> hydraulic conductivity of the sediments underlying this site is 0.<span class="hlt">1</span> ft/d. Ground-water velocities determined at three locations at the site in New Braintree, Mass., where ground water discharges naturally from the underlying aquifer to the Ware River, ranged from 0.10 to 0.20 ft/d upward. The effective <span class="hlt">vertical</span> hydraulic conductivity of the sediments underlying this site ranged from 2.4 to 17.<span class="hlt">1</span> ft/d. Ground-water velocities determined at three locations at the Dover, N.J., site, where infiltration from the Rockaway River into the underlying sediments occurs because of pumping, were <span class="hlt">1</span>.5 ft/d downward. The effective <span class="hlt">vertical</span> hydraulic conductivity of the sediments underlying this site ranged from 2.2 to 2.5 ft/d. Independent estimates of velocity at two of the three sites are in general agreement with the velocities determined using temperature profiles. The estimates of velocities and conductivities derived from the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19900061998&hterms=Cat+eyes&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DCat%2Beyes','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900061998&hterms=Cat+eyes&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DCat%2Beyes"><span>Influence of gravity on cat <span class="hlt">vertical</span> vestibulo-ocular reflex</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tomko, D. L.; Wall, C., III; Robinson, F. R.; Staab, J. P.</p> <p>1988-01-01</p> <p>The <span class="hlt">vertical</span> vestibulo-ocular reflex (VOR) was recorded in cats using electro-oculography during sinusoidal angular pitch. Peak stimulus velocity was 50 deg/s over a frequency range from 0.01 to 4.0 Hz. To test the effect of gravity on the <span class="hlt">vertical</span> VOR, the animal was pitched while sitting upright or lying on its side. Upright pitch changed the cat's orientation relative to gravity, while on-side pitch did not. The cumulative slow component position of the eye during on-side pitch was less symmetric than during upright pitch. Over the mid-frequency range (0.<span class="hlt">1</span> to <span class="hlt">1</span>.0 Hz), the average gain of the <span class="hlt">vertical</span> VOR was 14.5 percent higher during upright pitch than during on-side pitch. At low frequencies (less than 0.05 Hz) changing head position relative to gravity raised the <span class="hlt">vertical</span> VOR gain and kept the reflex in phase with stimulus velocity. These results indicate that gravity-sensitive mechanisms make the <span class="hlt">vertical</span> VOR more compensatory.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/wri/1987/4147/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/wri/1987/4147/report.pdf"><span>The effects of <span class="hlt">vertical</span> motion on the performance of current meters</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Thibodeaux, K.G.; Futrell, J. C.</p> <p>1987-01-01</p> <p>A series of tests to determine the correction coefficients for Price type AA and Price type OAA current meters, when subjected to <span class="hlt">vertical</span> motion in a towing tank, have been conducted. During these tests, the meters were subjected to <span class="hlt">vertical</span> travel that ranged from <span class="hlt">1</span>.0 to 4.0 ft and <span class="hlt">vertical</span> rates of travel that ranged from 0.33 to <span class="hlt">1</span>.20 ft/sec while being towed through the water at speeds ranging from 0 to 8 ft/sec. The tests show that type AA and type OAA current meters are affected adversely by the rate of <span class="hlt">vertical</span> motion and the distance of <span class="hlt">vertical</span> travel. In addition, the tests indicate that when current meters are moved <span class="hlt">vertically</span>, correction coefficients must be applied to the observed meter velocities to correct for the registration errors that are induced by the <span class="hlt">vertical</span> motion. The type OAA current meter under-registers and the type AA current meter over-registers in observed meter velocity. These coefficients for the type OAA current meter range from 0.99 to <span class="hlt">1</span>.49 and for the type AA current meter range from 0.33 to <span class="hlt">1</span>.07. When making current meter measurements from a boat or a cableway, errors in observed current meter velocity will occur when the bobbing of a boat or cableway places the current meter into <span class="hlt">vertical</span> motion. These errors will be significant when flowing water is < 2 ft/sec and the rate of <span class="hlt">vertical</span> motion is > 0.3 ft/sec. (Author 's abstract)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFDQ22007S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFDQ22007S"><span>Thermally driven film climbing a <span class="hlt">vertical</span> cylinder</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Smolka, Linda</p> <p>2017-11-01</p> <p>The dynamics of a Marangoni driven film climbing the outside of a <span class="hlt">vertical</span> cylinder is examined in numerical simulations of a thin film model. The model has three parameters: the scaled cylinder radius R̂, upstream film height h∞ and downstream precursor film thickness b , and reduces to the model for Marangoni driven film climbing a <span class="hlt">vertical</span> plate when R̂ -> ∞ . The advancing front displays dynamics similar to that along a <span class="hlt">vertical</span> plate where, depending on h∞ , the film forms a Lax shock, an undercompressive double shock or a rarefaction-undercompressive shock. A linear stability analysis of the Lax shock reveals the number of fingers that form along the contact line increases linearly with cylinder circumference while no fingers form below R̂ <span class="hlt">1</span>.15 with b = 0.<span class="hlt">1</span> . The substrate curvature controls the Lax shock height, bounds on h∞ that define the three solutions and the maximum growth rate of perturbations when R̂ = O (<span class="hlt">1</span>) , whereas the shape of solutions and the stability of the Lax shock converge to the behavior on a <span class="hlt">vertical</span> plate when R̂ >= O (10) . The azimuthal curvatures of the base state and perturbation, arising from the annular geometry of the film, promote instability of the advancing contact line.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018TePhL..44...24B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018TePhL..44...24B"><span><span class="hlt">Vertical</span>-Cavity Surface-Emitting <span class="hlt">1</span>.55-μm Lasers Fabricated by Fusion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Babichev, A. V.; Karachinskii, L. Ya.; Novikov, I. I.; Gladyshev, A. G.; Blokhin, S. A.; Mikhailov, S.; Iakovlev, V.; Sirbu, A.; Stepniak, G.; Chorchos, L.; Turkiewicz, J. P.; Voropaev, K. O.; Ionov, A. S.; Agustin, M.; Ledentsov, N. N.; Egorov, A. Yu.</p> <p>2018-01-01</p> <p>The results of studies on fabrication of <span class="hlt">vertical</span>-cavity surface-emitting <span class="hlt">1</span>.55-μm lasers by fusing AlGaAs/GaAs distributed-Bragg-reflector wafers and an active region based on thin In0.74Ga0.26 As quantum wells grown by molecular-beam epitaxy are presented. Lasers with a current aperture diameter of 8 μm exhibit continuous lasing with a threshold current below <span class="hlt">1</span>.5 mA, an output optical power of 6 mW, and an efficiency of approximately 22%. Single-mode lasing with a side-mode suppression ratio of 40-45 dB is observed in the entire operating current range. The effective modulation frequency of these lasers is as high as 9 GHz and is limited by the low parasitic cutoff frequency and self-heating.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_11 --> <div id="page_12" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="221"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.P13D..08P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.P13D..08P"><span><span class="hlt">Vertical</span> and Lateral Electron Content in the Martian Ionosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Paetzold, M. P.; Peter, K.; Bird, M. K.; Häusler, B.; Tellmann, S.</p> <p>2016-12-01</p> <p>The radio-science experiment MaRS (Mars Express Radio Science) on the Mars Express spacecraft sounds the neutral atmosphere and ionosphere of Mars since 2004. Approximately 800 <span class="hlt">vertical</span> profiles of the ionospheric electron density have been acquired until today. The <span class="hlt">vertical</span> electron content (TEC) is easily computed from the <span class="hlt">vertical</span> electron density profile by integrating along the altitude. The TEC is typically a fraction of a TEC unit (<span class="hlt">1</span>E16 m^-2) and depends on the solar zenith angle. The magnitude of the TEC is however fully dominated by the electron density contained in the main layer M2. The contributions by the M<span class="hlt">1</span> layer below M2 or the topside is marginal. MaRS is using two radio frequencies for the sounding of the ionosphere. The directly observed differential Doppler from the two received frequencies is a measure of the lateral electron content that means along the ray path and perpendicular to the <span class="hlt">vertical</span> electron density profile. Combining both the <span class="hlt">vertical</span> electron density profile, the <span class="hlt">vertical</span> TEC and the directly observed lateral TEC describes the lateral electron density distribution in the ionosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19980018859','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19980018859"><span>Aiding <span class="hlt">Vertical</span> Guidance Understanding</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Feary, Michael; McCrobie, Daniel; Alkin, Martin; Sherry, Lance; Polson, Peter; Palmer, Everett; McQuinn, Noreen</p> <p>1998-01-01</p> <p>A two-part study was conducted to evaluate modern flight deck automation and interfaces. In the first part, a survey was performed to validate the existence of automation surprises with current pilots. Results indicated that pilots were often surprised by the behavior of the automation. There were several surprises that were reported more frequently than others. An experimental study was then performed to evaluate (<span class="hlt">1</span>) the reduction of automation surprises through training specifically for the <span class="hlt">vertical</span> guidance logic, and (2) a new display that describes the flight guidance in terms of aircraft behaviors instead of control modes. The study was performed in a simulator that was used to run a complete flight with actual airline pilots. Three groups were used to evaluate the guidance display and training. In the training, condition, participants went through a training program for <span class="hlt">vertical</span> guidance before flying the simulation. In the display condition, participants ran through the same training program and then flew the experimental scenario with the new Guidance-Flight Mode Annunciator (G-FMA). Results showed improved pilot performance when given training specifically for the <span class="hlt">vertical</span> guidance logic and greater improvements when given the training and the new G-FMA. Using actual behavior of the avionics to design pilot training and FMA is feasible, and when the automated <span class="hlt">vertical</span> guidance mode of the Flight Management System is engaged, the display of the guidance mode and targets yields improved pilot performance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19870001022','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19870001022"><span>Climatology of tropospheric <span class="hlt">vertical</span> velocity spectra</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ecklund, W. L.; Gage, K. S.; Balsley, B. B.; Carter, D. A.</p> <p>1986-01-01</p> <p><span class="hlt">Vertical</span> velocity power spectra obtained from Poker Flat, Alaska; Platteville, Colorado; Rhone Delta, France; and Ponape, East Caroline Islands using 50-MHz clear-air radars with <span class="hlt">vertical</span> beams are given. The spectra were obtained by analyzing the quietest periods from the one-minute-resolution time series for each site. The lengths of available <span class="hlt">vertical</span> records ranged from as long as 6 months at Poker Flat to about <span class="hlt">1</span> month at Platteville. The quiet-time <span class="hlt">vertical</span> velocity spectra are shown. Spectral period ranging from 2 minutes to 4 hours is shown on the abscissa and power spectral density is given on the ordinate. The Brunt-Vaisala (B-V) periods (determined from nearby sounding balloons) are indicated. All spectra (except the one from Platteville) exhibit a peak at periods slightly longer than the B-V period, are flat at longer periods, and fall rapidly at periods less than the B-V period. This behavior is expected for a spectrum of internal waves and is very similar to what is observed in the ocean (Eriksen, 1978). The spectral amplitudes vary by only a factor of 2 or 3 about the mean, and show that under quiet conditions <span class="hlt">vertical</span> velocity spectra from the troposphere are very similar at widely different locations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C34B..02B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C34B..02B"><span>Constraining ice sheet history in the Weddell Sea, West Antarctica, using ice fabric at Korff Ice <span class="hlt">Rise</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brisbourne, A.; Smith, A.; Kendall, J. M.; Baird, A. F.; Martin, C.; Kingslake, J.</p> <p>2017-12-01</p> <p>The grounding history of ice <span class="hlt">rises</span> (grounded area of independent flow regime within a floating ice shelf) can be used to constrain large scale ice sheet history: ice fabric, resulting from the preferred orientation of ice crystals due to the stress regime, can be used to infer this grounding history. With the aim of measuring the present day ice fabric at Korff Ice <span class="hlt">Rise</span>, West Antarctica, a multi-azimuth wide-angle seismic experiment was undertaken. Three wide-angle common-midpoint gathers were acquired centred on the apex of the ice <span class="hlt">rise</span>, at azimuths of 60 degrees to one another, to measure variation in seismic properties with offset and azimuth. Both <span class="hlt">vertical</span> and horizontal receivers were used to record P and S arrivals including converted phases. Measurements of the variation with offset and azimuth of seismic traveltimes, seismic attenuation and shear wave splitting have been used to quantify seismic anisotropy in the ice column. The observations cannot be reproduced using an isotropic ice column model. Anisotropic ray tracing has been used to test likely models of ice fabric by comparison with the data. A model with a weak girdle fabric overlying a strong cluster fabric provides the best fit to the observations. Fabric of this nature is consistent with Korff Ice <span class="hlt">Rise</span> having been stable for the order of 10,000 years without any ungrounding or significant change in the ice flow configuration across the ice <span class="hlt">rise</span> for this period. This observation has significant implications for the ice sheet history of the Weddell Sea sector.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120003669','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120003669"><span>Sea-Level Projections from the Sea<span class="hlt">RISE</span> Initiative</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Nowicki, Sophie; Bindschadler, Robert</p> <p>2011-01-01</p> <p>Sea<span class="hlt">RISE</span> (Sea-level Response to Ice Sheet Evolution) is a community organized modeling effort, whose goal is to inform the fifth IPCC of the potential sea-level contribution from the Greenland and Antarctic ice sheets in the 21st and 22nd century. Sea<span class="hlt">RISE</span> seeks to determine the most likely ice sheet response to imposed climatic forcing by initializing an ensemble of models with common datasets and applying the same forcing to each model. Sensitivity experiments were designed to quantify the sea-level <span class="hlt">rise</span> associated with a change in: <span class="hlt">1</span>) surface mass balance, 2) basal lubrication, and 3) ocean induced basal melt. The range of responses, resulting from the multi-model approach, is interpreted as a proxy of uncertainty in our sea-level projections. http://websrv.cs .umt.edu/isis/index.php/Sea<span class="hlt">RISE</span>_Assessment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMGC42A..05K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMGC42A..05K"><span>Impact of a <span class="hlt">1</span>.5 °C Global Temperature <span class="hlt">Rise</span> on the Glaciers of High Mountain Asia</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kraaijenbrink, P. D. A.; Lutz, A. F.; Bierkens, M. F.; Immerzeel, W. W.</p> <p>2017-12-01</p> <p>Glaciers in High Mountain Asia (HMA) play a substantial role in regional water resources. In general, these glaciers have been retreating and losing mass over the last decades, a trend that is most likely to persist under future temperature <span class="hlt">rise</span>. At the 2015 UN Climate Change Conference in Paris, 195 nations signed the "Paris Agreement" and agreed on efforts to limit the global temperature <span class="hlt">rise</span> to <span class="hlt">1</span>.5 °C. It is unknown, however, how much of Asia's ice mass will be lost under a <span class="hlt">1</span>.5 °C increase in temperature, or under more extreme temperature scenarios. To estimate this, we use a mass balance gradient glacier model including mass redistribution for all individual glaciers in HMA larger than 0.4 km2. The model was run transiently up to 2100 using climate change scenarios from the full CMIP5 climate model ensemble, and a Monte Carlo framework was implemented to account for parameter uncertainty. Results show that only a handful of the climate models project a global temperature <span class="hlt">rise</span> of <span class="hlt">1</span>.5 °C, and that under such a scenario HMA warms stronger than the global mean (2.<span class="hlt">1</span>±0.<span class="hlt">1</span> °C). By the end of the 21st century, the <span class="hlt">1</span>.5 °C scenario results in a 36±7% mass loss relative to the estimated present day ice mass in HMA of 4754±350 Gt. Mass losses projected by RCP4.5, RCP6.0 and RCP8.5 model ensembles are considerably higher with 49±7%, 51±6% and 65±6%, respectively. Results also show large regional differences in mass loss, which are primarily caused by regional differences in debris cover, ice mass, observed mass balance, and glacier sensitivity to climate perturbations. We conclude that even under the ambitious <span class="hlt">1</span>.5 °C scenario over one third of HMA's glacier mass will be lost by the end of century, and that the difference with more likely scenarios is large. The <span class="hlt">1</span>.5 °C target is a task of a dimension that the world has never seen before, but it might prove to be very important for the fate of the glaciers and for water resources in the region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/15298','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/15298"><span>Effect of <span class="hlt">vertical</span> integration on the utilization of hardwood resources</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Jan Wiedenbeck</p> <p>2002-01-01</p> <p>The effectiveness of <span class="hlt">vertical</span> integration in promoting the efficient utilization of the hardwood resource in the eastern United States was assessed during a series of interviews with <span class="hlt">vertically</span> integrated hardwood manufacturers in the Appalachian region. Data from 19 companies that responded to the 1996 phone survey indicate that: <span class="hlt">1</span>) <span class="hlt">vertically</span> integrated hardwood...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EPJWC.14009002P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EPJWC.14009002P"><span>Influence of obstacles on bubbles <span class="hlt">rising</span> in water-saturated sand</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Poryles, Raphaël; Varas, Germán; Vidal, Valérie</p> <p>2017-06-01</p> <p>This work investigates the dynamics of air <span class="hlt">rising</span> through a water-saturated sand confined in a Hele- Shaw cell in which a circular obstacle is trapped. The air is injected at constant flow rate through a single nozzle at the bottom center of the cell. Without obstacle, in a similar configuration, previous studies pointed out the existence of a fluidized zone generated by the central upward gas motion which entrains two granular convection rolls on its sides. Here, a circular obstacle which diameter is of the order of the central air channel width is trapped at the <span class="hlt">vertical</span> of the injection nozzle. We analyze the influence of the obstacle location on the size of the fluidized zone and its impact on the morphology of the central air channel. Finally, we quantify the variations of the granular free surface. Two configurations with multiple obstacles are also considered.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27585489','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27585489"><span>GLI<span class="hlt">1</span>+ progenitor cells in the adrenal capsule of the adult mouse give <span class="hlt">rise</span> to heterotopic gonadal-like tissue.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dörner, Julia; Martinez Rodriguez, Verena; Ziegler, Ricarda; Röhrig, Theresa; Cochran, Rebecca S; Götz, Ronni M; Levin, Mark D; Pihlajoki, Marjut; Heikinheimo, Markku; Wilson, David B</p> <p>2017-02-05</p> <p>As certain strains of mice age, hyperplastic lesions resembling gonadal tissue accumulate beneath the adrenal capsule. Gonadectomy (GDX) accelerates this heterotopic differentiation, resulting in the formation of wedge-shaped adrenocortical neoplasms that produce sex steroids. Stem/progenitor cells that reside in the adrenal capsule and retain properties of the adrenogonadal primordium are thought to be the source of this heterotopic tissue. Here, we demonstrate that GLI<span class="hlt">1</span> + progenitors in the adrenal capsule give <span class="hlt">rise</span> to gonadal-like cells that accumulate in the subcapsular region. A tamoxifen-inducible Cre driver (Gli<span class="hlt">1</span>-creER T2 ) and two reporters (R26R-lacZ, R26R-confetti) were used to track the fate of GLI<span class="hlt">1</span> + cells in the adrenal glands of B6D2F2 mice, a strain that develops both GDX-induced adrenocortical neoplasms and age-dependent subcapsular cell hyperplasia. In gonadectomized B6D2F2 mice GLI<span class="hlt">1</span> + progenitors contributed to long-lived adrenal capsule cells and to adrenocortical neoplasms that expressed Gata4 and Foxl2, two prototypical gonadal markers. Pdgfra, a gene expressed in adrenocortical stromal cells, was upregulated in the GDX-induced neoplasms. In aged non-gonadectomized B6D2F2 mice GLI<span class="hlt">1</span> + progenitors gave <span class="hlt">rise</span> to patches of subcapsular cell hyperplasia. Treatment with GANT61, a small-molecule GLI antagonist, attenuated the upregulation of gonadal-like markers (Gata4, Amhr2, Foxl2) in response to GDX. These findings support the premise that GLI<span class="hlt">1</span> + progenitor cells in the adrenal capsule of the adult mouse give <span class="hlt">rise</span> to heterotopic tissue. Copyright © 2016 Elsevier Ireland Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015ApJ...813...81R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015ApJ...813...81R"><span>Radially Magnetized Protoplanetary Disk: <span class="hlt">Vertical</span> Profile</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Russo, Matthew; Thompson, Christopher</p> <p>2015-11-01</p> <p>This paper studies the response of a thin accretion disk to an external radial magnetic field. Our focus is on protoplanetary disks (PPDs), which are exposed during their later evolution to an intense, magnetized wind from the central star. A radial magnetic field is mixed into a thin surface layer, wound up by the disk shear, and pushed downward by a combination of turbulent mixing and ambipolar and ohmic drift. The toroidal field reaches much greater strengths than the seed <span class="hlt">vertical</span> field that is usually invoked in PPD models, even becoming superthermal. Linear stability analysis indicates that the disk experiences the magnetorotational instability (MRI) at a higher magnetization than a <span class="hlt">vertically</span> magnetized disk when both the effects of ambipolar and Hall drift are taken into account. Steady <span class="hlt">vertical</span> profiles of density and magnetic field are obtained at several radii between 0.06 and <span class="hlt">1</span> AU in response to a wind magnetic field Br ˜ (10-4-10-2)(r/ AU)-2 G. Careful attention is given to the radial and <span class="hlt">vertical</span> ionization structure resulting from irradiation by stellar X-rays. The disk is more strongly magnetized closer to the star, where it can support a higher rate of mass transfer. As a result, the inner ˜<span class="hlt">1</span> AU of a PPD is found to evolve toward lower surface density. Mass transfer rates around 10-8 M⊙ yr-<span class="hlt">1</span> are obtained under conservative assumptions about the MRI-generated stress. The evolution of the disk and the implications for planet migration are investigated in the accompanying paper.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19820024465','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19820024465"><span>RSRA <span class="hlt">vertical</span> drag test report. [rotor systems research aircraft</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Flemming, R. J.</p> <p>1981-01-01</p> <p>The Rotor Systems Research Aircraft (RSRA), because of its ability to measure rotor loads, was used to conduct an experiment to determine <span class="hlt">vertical</span> drag, tail rotor blockage, and thrust augmentation as affected by ground clearance and flight velocity. The RSRA was flown in the helicopter configuration at speeds from 0 to 15 knots for wheel heights from 5 to 150 feet, and to 60 knots out of ground effect. The <span class="hlt">vertical</span> drag trends in hover, predicted by theory and shown in model tests, were generally confirmed. The OGE hover <span class="hlt">vertical</span> drag is 4.0 percent, <span class="hlt">1.1</span> percent greater than predicted. The <span class="hlt">vertical</span> drag decreases rapidly as wheel height is reduced, and is zero at a wheel height of 6 feet. The <span class="hlt">vertical</span> drag also decreases with forward speed, approaching zero at sixty knots. The test data show the effect of wheel height and forward speed on thrust, gross weight capability, and power, and provide the relationships for power and collective pitch at constant gross weight required for the simulation of helicopter takeoffs and landings.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E%26ES..113a2128Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E%26ES..113a2128Y"><span>Influence of marine current on <span class="hlt">vertical</span> migration of Pb in marine bay</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yu, Chen; Hong, Ai; Danfeng, Yang; Huijuan, Zhao; Dongfang, Yang</p> <p>2018-02-01</p> <p>This paper analyzed that <span class="hlt">vertical</span> migration of Pb contents waters in Jiaozhou Bay, and revealed the influence of marine current on <span class="hlt">vertical</span> migration process. Results showed that Pb contents in bottom waters of Jiaozhou Bay in April and July 1988 were <span class="hlt">1</span>.49-18.53 μg L-<span class="hlt">1</span> and 12.68/-27.64 μg L-<span class="hlt">1</span>, respectively. The pollution level of Pb in bottom waters was moderate to heavy, and were showing temporal variations and spatial heterogeneity. The <span class="hlt">vertical</span> migration process of Pb in April 1988 included a drifting process from the southwest to the north by means of the marine current was rapid in this region. The <span class="hlt">vertical</span> migration process of Pb in July 1988 in the open waters included no drifting process since the flow rate of marine current was relative low in this region. The <span class="hlt">vertical</span> migration process of Pb was jointly determined by <span class="hlt">vertical</span> water’s effect, source input and water exchange, and the influence of marine current on the <span class="hlt">vertical</span> migration of Pb in marine bay was significant.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..16.4703T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..16.4703T"><span>Sea level <span class="hlt">rise</span> and variability around Peninsular Malaysia</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tkalich, Pavel; Luu, Quang-Hung; Tay, Tze-Wei</p> <p>2014-05-01</p> <p>Peninsular Malaysia is bounded from the west by Malacca Strait and the Andaman Sea, both connected to the Indian Ocean, and from the east by South China Sea being largest marginal sea in the Pacific Basin. As a result, sea level along Peninsular Malaysia coast is assumed to be governed by various regional phenomena associated with the adjacent parts of the Indian and Pacific Oceans. At annual scale, sea level anomalies (SLAs) are generated by the Asian monsoon; interannual sea level variability is determined by the El Niño-Southern Oscillation (ENSO) and the Indian Ocean Dipole (IOD); whilst long term sea level trend is coordinated by the global climate change. To quantify the relative impacts of these multi-scale phenomena on sea level trend and variability surrounding the Peninsular Malaysia, long-term tide gauge record and satellite altimetry are used. During 1984-2011, relative sea level <span class="hlt">rise</span> (SLR) rates in waters of Malacca Strait and eastern Peninsular Malaysia are found to be 2.4 ± 0.8 mm/yr and 2.7 ± 0.6 mm/yr, respectively. Discounting for their <span class="hlt">vertical</span> land movements (0.8 ± 2.6 mm/yr and 0.9 ± 2.2 mm/yr, respectively), their pure SLR rates are <span class="hlt">1</span>.6 ± 3.4 mm/yr and <span class="hlt">1</span>.8 ± 2.8 mm/yr, respectively, which are lower than the global tendency. At interannual scale, ENSO affects sea level over the Malaysian east coast in the range of ± 5 cm with very high correlation coefficient. Meanwhile, IOD modulates sea level anomalies in the Malacca Strait in the range of ± 2 cm with high correlation coefficient. Interannual regional sea level drops are associated with El Niño events and positive phases of the IOD index; while the <span class="hlt">rises</span> are correlated with La Niña episodes and the negative periods of the IOD index. Seasonally, SLAs are mainly monsoon-driven, in the order of 10-25 cm. Geographically, sea level responds differently to the monsoon: two cycles per year are observed in the Malacca Strait, presumably due to South Asian - Indian Monsoon; while single</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AnGeo..36..577L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AnGeo..36..577L"><span>High-resolution <span class="hlt">vertical</span> velocities and their power spectrum observed with the MAARSY radar - Part <span class="hlt">1</span>: frequency spectrum</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Qiang; Rapp, Markus; Stober, Gunter; Latteck, Ralph</p> <p>2018-04-01</p> <p>The Middle Atmosphere Alomar Radar System (MAARSY) installed at the island of Andøya has been run for continuous probing of atmospheric winds in the upper troposphere and lower stratosphere (UTLS) region. In the current study, we present high-resolution wind measurements during the period between 2010 and 2013 with MAARSY. The spectral analysis applying the Lomb-Scargle periodogram method has been carried out to determine the frequency spectra of <span class="hlt">vertical</span> wind velocity. From a total of 522 days of observations, the statistics of the spectral slope have been derived and show a dependence on the background wind conditions. It is a general feature that the observed spectra of <span class="hlt">vertical</span> velocity during active periods (with wind velocity > 10 m s-<span class="hlt">1</span>) are much steeper than during quiet periods (with wind velocity < 10 m s-<span class="hlt">1</span>). The distribution of spectral slopes is roughly symmetric with a maximum at -5/3 during active periods, whereas a very asymmetric distribution with a maximum at around -<span class="hlt">1</span> is observed during quiet periods. The slope profiles along altitudes reveal a significant height dependence for both conditions, i.e., the spectra become shallower with increasing altitudes in the upper troposphere and maintain roughly a constant slope in the lower stratosphere. With both wind conditions considered together the general spectra are obtained and their slopes are compared with the background horizontal winds. The comparisons show that the observed spectra become steeper with increasing wind velocities under quiet conditions, approach a spectral slope of -5/3 at a wind velocity of 10 m s-<span class="hlt">1</span> and then roughly maintain this slope (-5/3) for even stronger winds. Our findings show an overall agreement with previous studies; furthermore, they provide a more complete climatology of frequency spectra of <span class="hlt">vertical</span> wind velocities under different wind conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFDD15007N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFDD15007N"><span>Dynamics of poroelastocapillary <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nasouri, Babak; Elfring, Gwynn</p> <p>2017-11-01</p> <p>The surface-tension-driven <span class="hlt">rise</span> of a liquid between two elastic sheets can result in their deformation or coalescence depending on their flexibility. When the sheets are poroelastic, the flexibility of the immersed parts of the sheets can change considerably thereby altering the dynamical behavior of the system. To better understand this phenomenon, we study the poroelastocapillary <span class="hlt">rise</span> of a wetting liquid between poroelastic sheets. Using the lubrication theory and linear elasticity, we quantify the effects of the change in material properties of the wet sheets on the capillary <span class="hlt">rise</span> and the equilibrium state of the system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24435976','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24435976"><span>Localization of angiotensin-II type <span class="hlt">1</span>(AT<span class="hlt">1</span>) receptors on buffalo spermatozoa: AT<span class="hlt">1</span> receptor activation during capacitation triggers <span class="hlt">rise</span> in cyclic AMP and calcium.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Vedantam, Sivaram; Rani, Rita; Garg, Monica; Atreja, Suresh K</p> <p>2014-01-01</p> <p>The purpose of this study was to determine the role of Ang-II in buffalo spermatozoa; localize angiotensin type <span class="hlt">1</span> (AT<span class="hlt">1</span>) receptors on the sperm surface and understand the signaling mechanisms involved therein. Immunoblotting and immunocytochemistry using polyclonal Rabbit anti-AT<span class="hlt">1</span> (N-10) IgG were performed to confirm the presence of AT<span class="hlt">1</span> receptors. Intracellular levels of cyclic adenosine monophosphate (cAMP) were determined by non-radioactive enzyme immunoassay, while that of Calcium [Ca(2+)] were estimated by fluorimetry using Fura2AM dye. The results obtained showed that AT<span class="hlt">1</span> receptors were found on the post-acrosomal region, neck and tail regions. Immunoblotting revealed a single protein band with molecular weight of 40 kDa. Ang-II treated cells produced significantly higher level of cAMP compared to untreated cells (22.66 ± 2.4 vs. 10.8 ± 0.98 pmol/10(8) cells, p < 0.01). The mean levels of Ca(2+) were also higher in Ang-II treated cells compared to control (117.4 ± 6.<span class="hlt">1</span> vs. 61.15 ± 4.2 nmol/10(8) cells; p < 0.01). The stimulatory effect of Ang-II in both the cases was significantly inhibited in the presence of Losartan (AT<span class="hlt">1</span> antagonist; p < 0.05) indicating the involvement of AT<span class="hlt">1</span> receptors. Further, presence of neomycin (protein kinase C inhibitor) inhibited significantly the Ang-II mediated <span class="hlt">rise</span> in Ca(2+) indicating the involvement of PKC pathway. These findings confirm the presence of AT<span class="hlt">1</span> receptors in buffalo spermatozoa and that Ang-II mediates its actions via the activation of these receptors. Ang-II stimulates the <span class="hlt">rise</span> in intracellular levels of cAMP and Ca(2+) during capacitation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70195364','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70195364"><span>Correcting spacecraft jitter in Hi<span class="hlt">RISE</span> images</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Sutton, S. S.; Boyd, A.K.; Kirk, Randolph L.; Cook, Debbie; Backer, Jean; Fennema, A.; Heyd, R.; McEwen, A.S.; Mirchandani, S.D.; Wu, B.; Di, K.; Oberst, J.; Karachevtseva, I.</p> <p>2017-01-01</p> <p>Mechanical oscillations or vibrations on spacecraft, also called pointing jitter, cause geometric distortions and/or smear in high resolution digital images acquired from orbit. Geometric distortion is especially a problem with pushbroom type sensors, such as the High Resolution Imaging Science Experiment (Hi<span class="hlt">RISE</span>) instrument on board the Mars Reconnaissance Orbiter (MRO). Geometric distortions occur at a range of frequencies that may not be obvious in the image products, but can cause problems with stereo image correlation in the production of digital elevation models, and in measuring surface changes over time in orthorectified images. The Hi<span class="hlt">RISE</span> focal plane comprises a staggered array of fourteen charge-coupled devices (CCDs) with pixel IFOV of <span class="hlt">1</span> microradian. The high spatial resolution of Hi<span class="hlt">RISE</span> makes it both sensitive to, and an excellent recorder of jitter. We present an algorithm using Fourier analysis to resolve the jitter function for a Hi<span class="hlt">RISE</span> image that is then used to update instrument pointing information to remove geometric distortions from the image. Implementation of the jitter analysis and image correction is performed on selected Hi<span class="hlt">RISE</span> images. Resulting corrected images and updated pointing information are made available to the public. Results show marked reduction of geometric distortions. This work has applications to similar cameras operating now, and to the design of future instruments (such as the Europa Imaging System).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25823433','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25823433"><span>Resultant <span class="hlt">vertical</span> prism in toric soft contact lenses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sulley, Anna; Hawke, Ryan; Lorenz, Kathrine Osborn; Toubouti, Youssef; Olivares, Giovanna</p> <p>2015-08-01</p> <p>Rotational stability of toric soft contact lenses (TSCLs) is achieved using a range of designs. Designs utilising prism or peripheral ballast may result in residual prism in the optic zone. This study quantifies the <span class="hlt">vertical</span> prism in the central 6mm present in TSCLs with various stabilisation methods. <span class="hlt">Vertical</span> prism was computed using published refractive index and <span class="hlt">vertical</span> thickness changes in the central optic zone on a full lens thickness map. Thickness maps were measured using scanning transmission microscopy. Designs tested were reusable, silicone hydrogel and hydrogel TSCLs: SofLens(®) Toric, PureVision(®)2 for Astigmatism, PureVision(®) Toric, Biofinity(®) Toric, Avaira(®) Toric, clariti(®) toric, AIR OPTIX(®) for ASTIGMATISM and ACUVUE OASYS(®) for ASTIGMATISM; with eight parameter combinations for each lens (-6.00DS to +3.00DS, -<span class="hlt">1</span>.25DC, 90° and 180° axes). All TSCL designs evaluated had <span class="hlt">vertical</span> prism in the optic zone except one which had virtually none (0.01Δ). Mean prism ranged from 0.52Δ to <span class="hlt">1</span>.15Δ, with three designs having prism that varied with sphere power. <span class="hlt">Vertical</span> prism in ACUVUE OASYS(®) for ASTIGMATISM was significantly lower than all other TSCLs tested. TSCL designs utilising prism-ballast and peri-ballast for stabilisation have <span class="hlt">vertical</span> prism in the central optic zone. In monocular astigmats fitted with a TSCL or those wearing a mix of toric designs, <span class="hlt">vertical</span> prism imbalance could create or exacerbate disturbances in binocular vision function. Practitioners should be aware of this potential effect when selecting which TSCL designs to prescribe, particularly for monocular astigmats with pre-existing binocular vision anomalies, and when managing complaints of asthenopia in monocular astigmats. Copyright © 2015 British Contact Lens Association. Published by Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=294642','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=294642"><span>Macrophage-tropic variants initiate human immunodeficiency virus type <span class="hlt">1</span> infection after sexual, parenteral, and <span class="hlt">vertical</span> transmission.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>van't Wout, A B; Kootstra, N A; Mulder-Kampinga, G A; Albrecht-van Lent, N; Scherpbier, H J; Veenstra, J; Boer, K; Coutinho, R A; Miedema, F; Schuitemaker, H</p> <p>1994-01-01</p> <p>Macrophage-tropic, non-syncytium-inducing, HIV-<span class="hlt">1</span> variants predominate in the asymptomatic phase of infection and may be responsible for establishing infection in an individual exposed to the mixture of HIV-<span class="hlt">1</span> variants. Here, genotypical and phenotypical characteristics of virus populations, present in sexual, parenteral, or <span class="hlt">vertical</span> donor-recipient pairs, were studied. Sequence analysis of the V3 domain confirmed the presence of a homogeneous virus population in recently infected individuals. Biological HIV-<span class="hlt">1</span> clones were further characterized for syncytium inducing capacity on the MT2 cell line and for macrophage tropism as defined by the appearance of proviral DNA upon inoculation of monocyte-derived macrophages. Both sexual and parenteral transmission cases revealed a selective outgrowth in the recipient of the most macrophage-tropic variant(s) present in the donor. In three out of five <span class="hlt">vertical</span> transmission cases, more than one highly macrophage-tropic virus variant was present in the child shortly after birth, suggestive of transmission of multiple variants. In three primary infection cases, homogeneous virus populations of macrophage-tropic, non-syncytium-inducing variants were present prior to seroconversion, thus excluding humoral immunity as the selective pressure in favour of macrophage-tropic variants. These observations may have important implications for vaccine development. PMID:7962552</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002ExFl...32..624E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002ExFl...32..624E"><span>Measurements of air entrainment by <span class="hlt">vertical</span> plunging liquid jets</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>El Hammoumi, M.; Achard, J. L.; Davoust, L.</p> <p>2002-06-01</p> <p>This paper addresses the issue of the air-entrainment process by a <span class="hlt">vertical</span> plunging liquid jet. A non-dimensional physical analysis, inspired by the literature on the stability of free jets submitted to an aerodynamic interaction, was developed and yielded two correlation equations for the laminar and the turbulent plunging jets. These correlation equations allow the volumetric flow rate of the air carryunder represented by the Weber number of entrainment We n to be predicted. The plunging jets under consideration issued from circular tubes long enough to achieve a fully developed flow at the outlet. A sensitive technique based on a <span class="hlt">rising</span> soap meniscus was developed to measure directly the volumetric flow rate of the air carryunder. Our data are compared with other experimental data available in the literature; they also stand as a possible database for future theoretical modelling.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_12 --> <div id="page_13" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="241"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20090026463','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20090026463"><span>Flight Investigation of Effect of Various <span class="hlt">Vertical</span>-Tail Modifications on the Directional Stability and Control Characteristics of the P-63A-<span class="hlt">1</span> Airplane (AAF No. 42-68889)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Johnson, Harold I.</p> <p>1946-01-01</p> <p>Because the results of preliminary flight tests had indicated. the P-63A-<span class="hlt">1</span> airplane possessed insufficient directional stability, the NACA and the manufacturer (Bell Aircraft Corporation) suggested three <span class="hlt">vertical</span>-tail modifications to remedy the deficiencies in the directional characteristics. These modifications included an enlarged <span class="hlt">vertical</span> tail formed by adding a tip extension to the original <span class="hlt">vertical</span> tail, a large sharp-edge ventral fin, and a small dorsal fin. The enlarged <span class="hlt">vertical</span> tail involved only a slight increase in total <span class="hlt">vertical</span>-tail area from 23.73 to 26.58 square feet but a relatively much larger increase in geometric aspect ratio from <span class="hlt">1</span>.24 to <span class="hlt">1</span>.73 based on height and area above the horizontal tail. At the request of the Air Material Command, Army Air Forces, flight tests were made to determine the effect of these modifications and of some combinations of these modifications on the directional stability and control characteristics of the airplane, In all, six different <span class="hlt">vertical</span>-tail. configurations were investigated to determine the lateral and directional oscillation characteristics of the airplane, the sideslip characteristics, the yaw due to ailerons in rudder-fixed rolls from turns and pull-outs, the trim changes due to speed changes; and the trim changes due to power changes. Results of the tests showed that the enlarged <span class="hlt">vertical</span> tail approximately doubled the directional stability of the airplane and that the pilots considered the directional stability provided by the enlarged <span class="hlt">vertical</span> tail to be satisfactory. Calculations based on sideslip data obtained at an indicated airspeed of 300 miles per hour showed that the directional stability of the airplane with the original <span class="hlt">vertical</span> tail corresponded to a value of 0(sub n beta) of -0.00056 whereas for the enlarged <span class="hlt">vertical</span> tail the estimated va<span class="hlt">1</span>ue of C(sub n beta) was -0.00130, The ventral fin was found to increase by a moderate amount the directional stability of the airplane with the original</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22521954-radially-magnetized-protoplanetary-disk-vertical-profile','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22521954-radially-magnetized-protoplanetary-disk-vertical-profile"><span>RADIALLY MAGNETIZED PROTOPLANETARY DISK: <span class="hlt">VERTICAL</span> PROFILE</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Russo, Matthew; Thompson, Christopher</p> <p>2015-11-10</p> <p>This paper studies the response of a thin accretion disk to an external radial magnetic field. Our focus is on protoplanetary disks (PPDs), which are exposed during their later evolution to an intense, magnetized wind from the central star. A radial magnetic field is mixed into a thin surface layer, wound up by the disk shear, and pushed downward by a combination of turbulent mixing and ambipolar and ohmic drift. The toroidal field reaches much greater strengths than the seed <span class="hlt">vertical</span> field that is usually invoked in PPD models, even becoming superthermal. Linear stability analysis indicates that the disk experiencesmore » the magnetorotational instability (MRI) at a higher magnetization than a <span class="hlt">vertically</span> magnetized disk when both the effects of ambipolar and Hall drift are taken into account. Steady <span class="hlt">vertical</span> profiles of density and magnetic field are obtained at several radii between 0.06 and <span class="hlt">1</span> AU in response to a wind magnetic field B{sub r} ∼ (10{sup −4}–10{sup −2})(r/ AU){sup −2} G. Careful attention is given to the radial and <span class="hlt">vertical</span> ionization structure resulting from irradiation by stellar X-rays. The disk is more strongly magnetized closer to the star, where it can support a higher rate of mass transfer. As a result, the inner ∼<span class="hlt">1</span> AU of a PPD is found to evolve toward lower surface density. Mass transfer rates around 10{sup −8} M{sub ⊙} yr{sup −<span class="hlt">1</span>} are obtained under conservative assumptions about the MRI-generated stress. The evolution of the disk and the implications for planet migration are investigated in the accompanying paper.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AIPC.1157...19P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AIPC.1157...19P"><span>Adapting to <span class="hlt">Rising</span> Sea Level: A Florida Perspective</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Parkinson, Randall W.</p> <p>2009-07-01</p> <p>Global climate change and concomitant <span class="hlt">rising</span> sea level will have a profound impact on Florida's coastal and marine systems. Sea-level <span class="hlt">rise</span> will increase erosion of beaches, cause saltwater intrusion into water supplies, inundate coastal marshes and other important habitats, and make coastal property more vulnerable to erosion and flooding. Yet most coastal areas are currently managed under the premise that sea-level <span class="hlt">rise</span> is not significant and the shorelines are static or can be fixed in place by engineering structures. The new reality of sea-level <span class="hlt">rise</span> and extreme weather due to climate change requires a new style of planning and management to protect resources and reduce risk to humans. Scientists must: (<span class="hlt">1</span>) assess existing coastal vulnerability to address short term management issues and (2) model future landscape change and develop sustainable plans to address long term planning and management issues. Furthermore, this information must be effectively transferred to planners, managers, and elected officials to ensure their decisions are based upon the best available information. While there is still some uncertainty regarding the details of <span class="hlt">rising</span> sea level and climate change, development decisions are being made today which commit public and private investment in real estate and associated infrastructure. With a design life of 30 yrs to 75 yrs or more, many of these investments are on a collision course with <span class="hlt">rising</span> sea level and the resulting impacts will be significant. In the near term, the utilization of engineering structures may be required, but these are not sustainable and must ultimately yield to "managed withdrawal" programs if higher sea-level elevations or rates of <span class="hlt">rise</span> are forthcoming. As an initial step towards successful adaptation, coastal management and planning documents (i.e., comprehensive plans) must be revised to include reference to climate change and <span class="hlt">rising</span> sea-level.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19790014533','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19790014533"><span><span class="hlt">Vertical</span> motions in the equatorial middle atmosphere</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Weisman, M. L.</p> <p>1979-01-01</p> <p>A single station <span class="hlt">vertical</span> velocity equation which considers ageostrophic and diabatic effects derived from the first law of thermodynamics and a generalized thermal wind relation is presented. An analysis and verification procedure which accounts for measurement and calculation errors as well as time and space continuity arguments and theoretical predictions are described. <span class="hlt">Vertical</span> velocities are calculated at every kilometer between 25 and 60 km and for approximately every three hours for the above diurnal period at Kourou (French Guiana), Fort Sherman (Panama Canal Zone), Ascension Island, Antigua (British West Indies) and Natal (Brazil). The results, plotted as time series cross sections, suggest <span class="hlt">vertical</span> motions ranging in magnitude from <span class="hlt">1</span> or 2 cm/sec at 30 km to as much as 15 cm/sec at 60 km. Many of the general features of the results agree well with atmospheric tidal predictions but many particular features suggest that both smaller time scale gravity waves (periods less than 6 hours) and synoptic type waves (periods greater than <span class="hlt">1</span> day) may be interacting significantly with the tidal fields. The results suggest that <span class="hlt">vertical</span> motions can be calculated for the equatorial middle atmosphere and must be considered a significant part of the motion for time scales from 8 to 24 hours.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.G11A1059S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.G11A1059S"><span>Six years after the El Mayor-Cucapah earthquake: Transient far-field postseismic <span class="hlt">vertical</span> motion observed by tide gauges and GPS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Smith-Konter, B. R.; Gonzalez-Ortega, J. A.; Merrifield, M. A.; Tong, X.; Sandwell, D. T.; Hardy, S.; Howell, S. M.</p> <p>2016-12-01</p> <p>On April 4, 2010, the El Mayor-Cucapah earthquake (Mw 7.2) ruptured a 120 km long set of faults of the southernmost San Andreas Fault System in northeastern Baja California, Mexico. Near-field coseismic GPS observations revealed up to <span class="hlt">1.1</span> m of horizontal surface slip and 0.6 m of <span class="hlt">vertical</span> subsidence at near-field stations. Early near-field InSAR and GPS time series postseismic observations also suggested several tens of centimeters of afterslip occurred within the first two years, however postseismic transients due to viscoelastic or poroelastic relaxation have also been offered as candidate models. Here we investigate the role of viscoelastic transients from six years of regional far-field ( 200 km from rupture) tide gauge and <span class="hlt">vertical</span> GPS time series observations to further constrain postseismic deformation mechanisms. <span class="hlt">Vertical</span> viscoelastic postseismic models of the El Mayor-Cucapah earthquake suggest alternating quadrants of uplift and subsidence straddling the rupture, with uplift to the north near the Salton Trough and subsidence to the west spanning the San Diego and Ensenada regions. These decaying transient motions are confirmed by both <span class="hlt">vertical</span> postseismic GPS and tide gauge-altimetry observations, in both the near- and far fields. For example, tide gauge data in San Diego, which typically record <span class="hlt">vertical</span> land motions on the order of a few millimeters per year, recorded nearly 30 mm of transient land subsidence over the first 3 years. We find that the magnitude and decay of far-field postseismic subsidence can be attributed to viscoelastic relaxation of the mantle assuming a temporally varying rheology; viscosities as low as 1017 Pa-s for at least the first 6-12 months, followed by an increasing viscosity on the order of 1018 Pa-s in the years following, best fit the data. While transient viscosity anomalies have been previously suggested from GPS data spanning the first <span class="hlt">1</span>.5 years following the earthquake [Pollitz et al., 2012], the combined results from</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2837276','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2837276"><span>Latitude and longitude <span class="hlt">vertical</span> disparity</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Read, Jenny C. A.; Phillipson, Graeme P.; Glennerster, Andrew</p> <p>2010-01-01</p> <p>The literature on <span class="hlt">vertical</span> disparity is complicated by the fact that several different definitions of the term “<span class="hlt">vertical</span> disparity” are in common use, often without a clear statement about which is intended or a widespread appreciation of the properties of the different definitions. Here, we examine two definitions of retinal <span class="hlt">vertical</span> disparity: elevation-latitude and elevation-longitude disparity. Near the fixation point, these definitions become equivalent, but in general, they have quite different dependences on object distance and binocular eye posture, which have not previously been spelt out. We present analytical approximations for each type of <span class="hlt">vertical</span> disparity, valid for more general conditions than previous derivations in the literature: we do not restrict ourselves to objects near the fixation point or near the plane of regard, and we allow for non-zero torsion, cyclovergence and <span class="hlt">vertical</span> misalignments of the eyes. We use these expressions to derive estimates of the latitude and longitude <span class="hlt">vertical</span> disparity expected at each point in the visual field, averaged over all natural viewing. Finally, we present analytical expressions showing how binocular eye position – gaze direction, convergence, torsion, cyclovergence, and <span class="hlt">vertical</span> misalignment – can be derived from the <span class="hlt">vertical</span> disparity field and its derivatives at the fovea. PMID:20055544</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70034763','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70034763"><span>Sea-level <span class="hlt">rise</span> in New Jersey over the past 5000 years: Implications to anthropogenic changes</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Miller, Kenneth G.; Sugarman, Peter J.; Browning, James V.; Horton, Benjamin P.; Stanley, Alissa; Kahn, Alicia; Uptegrove, Jane; Aucott, Michael</p> <p>2009-01-01</p> <p>We present a mid to late Holocene sea-level record derived from drilling the New Jersey coast that shows a relatively constant <span class="hlt">rise</span> of <span class="hlt">1</span>.8??mm/yr from ~ 5000 to 500 calibrated calendar years before present (yrBP). This contrasts with previous New Jersey estimates that showed only 0.5??mm/yr <span class="hlt">rise</span> since 2000??yrBP. Comparison with other Mid-Atlantic sea-level records (Delaware to southern New England) indicates surprising uniformity considering different proximities to the peripheral bulge of the Laurentide ice sheet, with a relative <span class="hlt">rise</span> throughout the region of ~ <span class="hlt">1.7-1</span>.9??mm/yr since ~ 5000??yrBP. This regional sea-level <span class="hlt">rise</span> includes both: <span class="hlt">1</span>) global sea-level (eustatic) <span class="hlt">rise</span>; and 2) far-field geoidal subsidence (estimated as ~ 0.8-<span class="hlt">1</span>.4??mm/yr today) due to removal of the Laurentide ice sheet and water loading. Correcting for geoidal subsidence, the U.S. east coast records suggest a global sea-level (eustatic) <span class="hlt">rise</span> of ~ 0.4-<span class="hlt">1</span>.0??mm/yr (with a best estimate of 0.7 ?? 0.3??mm/yr) since 5000??yrBP. Comparison with other records provides a best estimate of pre-anthropogenic global sea-level <span class="hlt">rise</span> of < <span class="hlt">1</span>.0??mm/yr from 5000 until ~ 200??yrBP. Tide gauge data indicate a 20th century rate of eustatic <span class="hlt">rise</span> of <span class="hlt">1</span>.8??mm/yr, whereas both tide gauge and satellite data suggest an increase in the rate of <span class="hlt">rise</span> to ~ 3.3??mm/yr from 1993-2006 AD. This indicates that the modern <span class="hlt">rise</span> (~ 3.3??mm/yr) is significantly higher than the pre-anthropogenic <span class="hlt">rise</span> (0.7 ?? 0.3??mm/yr). ?? 2008 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ERL.....7b1001R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ERL.....7b1001R"><span>Sea-level <span class="hlt">rise</span>: towards understanding local vulnerability</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rahmstorf, Stefan</p> <p>2012-06-01</p> <p>Projections of global sea-level <span class="hlt">rise</span> into the future have become more pessimistic over the past five years or so. A global <span class="hlt">rise</span> by more than one metre by the year 2100 is now widely accepted as a serious possibility if greenhouse gas emissions continue unabated. That is witnessed by the scientific assessments that were made since the last IPCC report was published in 2007. The Delta Commission of the Dutch government projected up to <span class="hlt">1</span>.10 m as a 'high-end' scenario (Vellinga et al 2009). The Scientific Committee on Antarctic Research (SCAR) projected up to <span class="hlt">1</span>.40 m (Scientific Committee on Antarctic Research 2009), and the Arctic Monitoring and Assessment Programme (AMAP) gives a range of 0.90-<span class="hlt">1</span>.60 m in its 2011 report (Arctic Monitoring and Assessment Programme 2011). And recently the US Army Corps of Engineers recommends using a 'low', an 'intermediate' and a 'high' scenario for global sea-level <span class="hlt">rise</span> when planning civil works programmes, with the high one corresponding to a <span class="hlt">1</span>.50 m <span class="hlt">rise</span> by 2100 (US Army Corps of Engineers 2011). This more pessimistic view is based on a number of observations, most importantly perhaps the fact that sea level has been <span class="hlt">rising</span> at least 50% faster in the past decades than projected by the IPCC (Rahmstorf et al 2007, IPCC 2007). Also, the rate of <span class="hlt">rise</span> (averaged over two decades) has accelerated threefold, from around <span class="hlt">1</span> mm yr-<span class="hlt">1</span> at the start of the 20th century to around 3 mm yr-<span class="hlt">1</span> over the past 20 years (Church and White 2006), and this rate increase closely correlates with global warming (Rahmstorf et al 2011). The IPCC projections, which assume almost no further acceleration in the 20th century, thus look less plausible. And finally the observed net mass loss of the two big continental ice sheets (Van den Broeke et al 2011) calls into question the assumption that ice accumulation in Antarctica would largely balance ice loss from Greenland in the course of further global warming (IPCC 2007). With such a serious sea-level <span class="hlt">rise</span> on the horizon</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15.4646T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15.4646T"><span>Analysis of Sea Level <span class="hlt">Rise</span> in Singapore Strait</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tkalich, Pavel; Luu, Quang-Hung</p> <p>2013-04-01</p> <p>Sea level in Singapore Strait is governed by various scale phenomena, from global to local. Global signals are dominated by the climate change and multi-decadal variability and associated sea level <span class="hlt">rise</span>; at regional scale seasonal sea level variability is caused by ENSO-modulated monsoons; locally, astronomic tides are the strongest force. Tide gauge records in Singapore Strait are analyzed to derive local sea level trend, and attempts are made to attribute observed sea level variability to phenomena at various scales, from global to local. It is found that at annual scale, sea level anomalies in Singapore Strait are quasi-periodic, of the order of ±15 cm, the highest during northeast monsoon and the lowest during southwest monsoon. Interannual regional sea level falls are associated with El Niño events, while the <span class="hlt">rises</span> are related to La Niña episodes; both variations are in the range of ±9 cm. At multi-decadal scale, sea level in Singapore Strait has been <span class="hlt">rising</span> at the rate <span class="hlt">1.2-1</span>.9 mm/year for the period 1975-2009, 2.0±0.3 mm/year for 1984-2009, and <span class="hlt">1</span>.3-4.7 mm/year for 1993-2009. When compared with the respective global trends of 2.0±0.3, 2.4, and 2.8±0.8 mm/year, Singapore Strait sea level <span class="hlt">rise</span> trend was weaker at the earlier period and stronger at the recent decade.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26265123','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26265123"><span>Influence of short incompatible practice on the Simon effect: transfer along the <span class="hlt">vertical</span> dimension and across <span class="hlt">vertical</span> and horizontal dimensions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Conde, Erick F Q; Fraga-Filho, Roberto Sena; Lameira, Allan Pablo; Mograbi, Daniel C; Riggio, Lucia; Gawryszewski, Luiz G</p> <p>2015-11-01</p> <p>In spatial compatibility and Simon tasks, the response is faster when stimulus and response locations are on the same side than when they are on opposite sides. It has been shown that a spatial incompatible practice leads to a subsequent modulation of the Simon effect along the horizontal dimension. It has also been reported that this modulation occurs both along and across <span class="hlt">vertical</span> and horizontal dimensions, but only after intensive incompatible training (600 trials). In this work, we show that this modulatory effect can be obtained with a smaller number of incompatible trials, changing the spatial arrangement of the <span class="hlt">vertical</span> response keys to obtain a stronger dimensional overlap between the spatial codes of stimuli and response keys. The results of Experiment <span class="hlt">1</span> showed that 80 incompatible <span class="hlt">vertical</span> trials abolished the Simon effect in the same dimension. Experiment 2 showed that a modulation of the <span class="hlt">vertical</span> Simon effect could be obtained after 80 horizontal incompatible trials. Experiment 3 explored whether the transfer effect can also occur in a horizontal Simon task after a brief <span class="hlt">vertical</span> spatial incompatibility task, and results were similar to the previous experiments. In conclusion, we suggest that the spatial arrangement between response key and stimulus locations may be critical to establish the short-term memory links that enable the transfer of learning between brief incompatible practices and the Simon effects, both along the <span class="hlt">vertical</span> dimension and across <span class="hlt">vertical</span> and horizontal dimensions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25535809','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25535809"><span>Cone-Beam Computed Tomography Evaluation of Horizontal and <span class="hlt">Vertical</span> Dimensional Changes in Buccal Peri-Implant Alveolar Bone and Soft Tissue: A <span class="hlt">1</span>-Year Prospective Clinical Study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kaminaka, Akihiro; Nakano, Tamaki; Ono, Shinji; Kato, Tokinori; Yatani, Hirofumi</p> <p>2015-10-01</p> <p>This study evaluated changes in the horizontal and <span class="hlt">vertical</span> dimensions of the buccal alveolar bone and soft tissue over a <span class="hlt">1</span>-year period following implant prosthesis. Thirty-three participants with no history of guided bone regeneration or soft tissue augmentation underwent dental implant placement with different types of connections. The dimensions of the buccal alveolar bone and soft tissue were evaluated immediately and at <span class="hlt">1</span> year after prosthesis from reconstructions of cross-sectional cone-beam computed tomography images. The <span class="hlt">vertical</span> and horizontal loss of buccal bone and soft tissue around implants with conical connections were lower than around those with external or internal connections. Statistically significant negative correlations were observed between initial horizontal bone thickness and changes in <span class="hlt">vertical</span> bone and soft tissue height (p < .05), and between initial horizontal soft tissue thickness and the change in <span class="hlt">vertical</span> soft tissue height (p < .05). Implants with a conical connection preserve peri-implant alveolar bone and soft tissue more effectively than other connection types. Furthermore, the initial buccal alveolar bone and soft tissue thickness around the implant platform may influence their <span class="hlt">vertical</span> dimensional changes at <span class="hlt">1</span> year after implant prosthesis. © 2014 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.P23B1714S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.P23B1714S"><span>CRISM/Hi<span class="hlt">RISE</span> Correlative Spectroscopy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Seelos, F. P.; Murchie, S. L.; McGovern, A.; Milazzo, M. P.; Herkenhoff, K. E.</p> <p>2011-12-01</p> <p>The Mars Reconnaissance Orbiter (MRO) Compact Reconnaissance Imaging Spectrometer for Mars (CRISM) and High Resolution Imaging Science Experiment (Hi<span class="hlt">RISE</span>) are complementary investigations with high spectral resolution and broad wavelength coverage (CRISM ~20 m/pxl; ~400 - 4000 nm, 6.55 nm sampling) and high spatial resolution with broadband color capability (Hi<span class="hlt">RISE</span> ~25 cm/pxl; ~500, 700, 900 nm band centers, ~200-300 nm FWHM). Over the course of the MRO mission it has become apparent that spectral variations in the IR detected by CRISM (~1000 nm - 4000 nm) sometimes correlate spatially with visible and near infrared 3-band color variations observed by Hi<span class="hlt">RISE</span>. We have developed a data processing procedure that establishes a numerical mapping between Hi<span class="hlt">RISE</span> color and CRISM VNIR and IR spectral data and provides a statistical evaluation of the uncertainty in the mapping, with the objective of extrapolating CRISM-inferred mineralogy to the Hi<span class="hlt">RISE</span> spatial scale. The MRO mission profile, spacecraft capabilities, and science planning process emphasize coordinated observations - the simultaneous observation of a common target by multiple instruments. The commonalities of CRISM/Hi<span class="hlt">RISE</span> coordinated observations present a unique opportunity for tandem data analysis. Recent advances in the systematic processing of CRISM hyperspectral targeted observations account for gimbal-induced photometric variations and transform the data to a synthetic nadir acquisition geometry. The CRISM VNIR (~400 nm - 1000 nm) data can then be convolved to the Hi<span class="hlt">RISE</span> Infrared, Red, and Blue/Green (IRB) response functions to generate a compatible CRISM IRB product. Statistical evaluation of the CRISM/Hi<span class="hlt">RISE</span> spatial overlap region establishes a quantitative link between the data sets. IRB spectral similarity mapping for each Hi<span class="hlt">RISE</span> color spatial pixel with respect to the CRISM IRB product allows a given Hi<span class="hlt">RISE</span> pixel to be populated with information derived from the coordinated CRISM observation</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29642087','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29642087"><span>Market and organizational factors associated with hospital <span class="hlt">vertical</span> integration into sub-acute care.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hogan, Tory H; Lemak, Christy Harris; Hearld, Larry R; Sen, Bisakha P; Wheeler, Jack R C; Menachemi, Nir</p> <p>2018-04-11</p> <p>Changes in payment models incentivize hospitals to <span class="hlt">vertically</span> integrate into sub-acute care (SAC) services. Through <span class="hlt">vertical</span> integration into SAC, hospitals have the potential to reduce the transaction costs associated with moving patients throughout the care continuum and reduce the likelihood that patients will be readmitted. The purpose of this study is to examine the correlates of hospital <span class="hlt">vertical</span> integration into SAC. Using panel data of U.S. acute care hospitals (2008-2012), we conducted logit regression models to examine environmental and organizational factors associated with hospital <span class="hlt">vertical</span> integration. Results are reported as average marginal effects. Among 3,775 unique hospitals (16,269 hospital-year observations), 25.7% <span class="hlt">vertically</span> integrated into skilled nursing facilities during at least <span class="hlt">1</span> year of the study period. One measure of complexity, the availability of skilled nursing facilities in a county (ME = -<span class="hlt">1</span>.780, p < .001), was negatively associated with hospital <span class="hlt">vertical</span> integration into SAC. Measures of munificence, percentage of the county population eligible for Medicare (ME = 0.018, p < .001) and rural geographic location (ME = 0.069, p < .001), were positively associated with hospital <span class="hlt">vertical</span> integration into SAC. Dynamism, when measured as the change county population between 2008 and 2011 (ME = <span class="hlt">1</span>.19e-06, p < .001), was positively associated with hospital <span class="hlt">vertical</span> integration into SAC. Organizational resources, when measured as swing beds (ME = 0.069, p < .001), were positively associated with hospital <span class="hlt">vertical</span> integration into SAC. Organizational resources, when measured as investor owned (ME = -0.052, p < .<span class="hlt">1</span>) and system affiliation (ME = -0.041, p < .<span class="hlt">1</span>), were negatively associated with hospital <span class="hlt">vertical</span> integration into SAC. Hospital adaption to the changing health care landscape through <span class="hlt">vertical</span> integration varies across market and organizational conditions. Current Centers for Medicare and Medicaid reimbursement programs do not take</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28699419','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28699419"><span>Chair <span class="hlt">rise</span> capacity and associated factors in older home-care clients.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tiihonen, Miia; Hartikainen, Sirpa; Nykänen, Irma</p> <p>2017-07-01</p> <p>The aim of this study was to investigate the ability of older home-care clients to perform the five times chair <span class="hlt">rise</span> test and associated personal characteristics, nutritional status and functioning. The study sample included 267 home-care clients aged ≥75 years living in Eastern and Central Finland. The home-care clients were interviewed at home by home-care nurses, nutritionists and pharmacists. The collected data contained sociodemographic factors, functional ability (Barthel Index, IADL), cognitive functioning (MMSE), nutritional status (MNA), depressive symptoms (GDS-15), medical diagnoses and drug use. The primary outcome was the ability to perform the five times chair <span class="hlt">rise</span> test. Fifty-one per cent ( n=135) of the home-care clients were unable to complete the five times chair <span class="hlt">rise</span> test. Twenty-three per cent ( n=64) of the home-care clients had good chair <span class="hlt">rise</span> capacity (≤17 seconds). In a multivariate logistic regression analysis, fewer years of education (odds ratio [OR] = <span class="hlt">1</span>.11, 95% confidence interval [CI] <span class="hlt">1.04-1</span>.18), lower ADL (OR = <span class="hlt">1</span>.54, 95% CI <span class="hlt">1.34-1</span>.78) and low MNA scores (OR = <span class="hlt">1</span>.12, 95% CI <span class="hlt">1.04-1</span>.20) and a higher number of co-morbidities (OR = <span class="hlt">1</span>.21, 95% CI <span class="hlt">1.02-1</span>.43) were associated with inability to complete the five times chair <span class="hlt">rise</span> test. Poor functional mobility, which was associated with less education, a high number of co-morbidities and poor nutritional status, was common among older home-care clients. To maintain and to prevent further decline in functional mobility, physical training and nutritional services are needed. (NutOrMed, ClinicalTrials.gov Identifier: NCT02214758).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18172495','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18172495"><span><span class="hlt">Vertical</span> structure of recent Arctic warming.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Graversen, Rune G; Mauritsen, Thorsten; Tjernström, Michael; Källén, Erland; Svensson, Gunilla</p> <p>2008-01-03</p> <p>Near-surface warming in the Arctic has been almost twice as large as the global average over recent decades-a phenomenon that is known as the 'Arctic amplification'. The underlying causes of this temperature amplification remain uncertain. The reduction in snow and ice cover that has occurred over recent decades may have played a role. Climate model experiments indicate that when global temperature <span class="hlt">rises</span>, Arctic snow and ice cover retreats, causing excessive polar warming. Reduction of the snow and ice cover causes albedo changes, and increased refreezing of sea ice during the cold season and decreases in sea-ice thickness both increase heat flux from the ocean to the atmosphere. Changes in oceanic and atmospheric circulation, as well as cloud cover, have also been proposed to cause Arctic temperature amplification. Here we examine the <span class="hlt">vertical</span> structure of temperature change in the Arctic during the late twentieth century using reanalysis data. We find evidence for temperature amplification well above the surface. Snow and ice feedbacks cannot be the main cause of the warming aloft during the greater part of the year, because these feedbacks are expected to primarily affect temperatures in the lowermost part of the atmosphere, resulting in a pattern of warming that we only observe in spring. A significant proportion of the observed temperature amplification must therefore be explained by mechanisms that induce warming above the lowermost part of the atmosphere. We regress the Arctic temperature field on the atmospheric energy transport into the Arctic and find that, in the summer half-year, a significant proportion of the <span class="hlt">vertical</span> structure of warming can be explained by changes in this variable. We conclude that changes in atmospheric heat transport may be an important cause of the recent Arctic temperature amplification.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title21-vol2/pdf/CFR-2010-title21-vol2-sec137-180.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title21-vol2/pdf/CFR-2010-title21-vol2-sec137-180.pdf"><span>21 CFR 137.180 - Self-<span class="hlt">rising</span> flour.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-04-01</p> <p>... 21 Food and Drugs 2 2010-04-01 2010-04-01 false Self-<span class="hlt">rising</span> flour. 137.180 Section 137.180 Food... Flours and Related Products § 137.180 Self-<span class="hlt">rising</span> flour. (a) Self-<span class="hlt">rising</span> flour, self-<span class="hlt">rising</span> white flour, self-<span class="hlt">rising</span> wheat flour, is an intimate mixture of flour, sodium bicarbonate, and one or more of the...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C51B0979M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C51B0979M"><span>Using a <span class="hlt">Vertically</span> Integrated Model to Determine the Effects of Seasonal Forcing on the Basal Topography of Ice Shelves</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>MacMackin, C. T.; Wells, A.</p> <p>2017-12-01</p> <p>While relatively small in mass, ice shelves play an important role in buttressing ice sheets, slowing their flow into the ocean. As such, an understanding of ice shelf stability is needed for predictions of future sea level <span class="hlt">rise</span>. Networks of channels have been observed underneath Antarctic ice shelves and are thought to affect their stability. While the origins of channels running parallel to ice flow are thought to be well understood, transverse channels have also been observed and the mechanism for their formation is less clear. It has been suggested that seasonal variations in ice and ocean properties could be a source and we run nonlinear, <span class="hlt">vertically</span> integrated <span class="hlt">1</span>-D simulations of a coupled ice shelf and plume to test this hypothesis. We also examine how these variations might alter the shape of internal radar reflectors within the ice, suggesting a new technique to model their distribution using a <span class="hlt">vertically</span> integrated model of ice flow. We examine a range of sources for seasonal forcing which might lead to channel formation, finding that variability in subglacial discharge results in small variations of ice thickness. Additional mechanisms would be required to expand these into large transverse channels.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E3SWC..3301023I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E3SWC..3301023I"><span>Regional approaches in high-<span class="hlt">rise</span> construction</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Iconopisceva, O. G.; Proskurin, G. A.</p> <p>2018-03-01</p> <p>The evolutionary process of high-<span class="hlt">rise</span> construction is in the article focus. The aim of the study was to create a retrospective matrix reflecting the tasks of the study such as: structuring the most iconic high-<span class="hlt">rise</span> objects within historic boundaries. The study is based on contemporary experience of high-<span class="hlt">rise</span> construction in different countries. The main directions and regional specifics in the field of high-<span class="hlt">rise</span> construction as well as factors influencing the further evolution process are analyzed. The main changes in architectural stylistics, form-building, constructive solutions that focus on the principles of energy efficiency and bio positivity of "sustainable buildings", as well as the search for a new typology are noted. The most universal constructive methods and solutions that turned out to be particularly popular are generalized. The new typology of high-<span class="hlt">rises</span> and individual approach to urban context are noted. The results of the study as a graphical scheme made it possible to represent the whole high-<span class="hlt">rise</span> evolution. The new spatial forms of high-<span class="hlt">rises</span> lead them to new role within the urban environments. Futuristic hyperscalable concepts take the autonomous urban space functions itself and demonstrate us how high-<span class="hlt">rises</span> can replace multifunctional urban fabric, developing it inside their shells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=lifts&id=EJ929531','ERIC'); return false;" href="https://eric.ed.gov/?q=lifts&id=EJ929531"><span>Impact of the A18.<span class="hlt">1</span> ASME Standard on Platform Lifts and Stairway Chairlifts on Accessibility and Usability</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Balmer, David C.</p> <p>2010-01-01</p> <p>This article summarizes the effect of the ASME A18.<span class="hlt">1</span> Standard concerning accessibility and usability of Platform Lifts and their remaining technological challenges. While elevators are currently the most effective means of <span class="hlt">vertical</span> transportation related to speed, capacity, <span class="hlt">rise</span> and usability, their major drawbacks for accessibility are cost and…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MeScT..28k2001M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MeScT..28k2001M"><span>Ultrasound excited thermography: an efficient tool for the characterization of <span class="hlt">vertical</span> cracks</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mendioroz, A.; Celorrio, R.; Salazar, A.</p> <p>2017-11-01</p> <p>Ultrasound excited thermography has gained a renewed interest in the last two decades as a nondestructive testing technique aimed at detecting and characterizing surface breaking and shallow subsurface discontinuities. It is based on measurement of the IR radiation emitted by the specimen surface to detect temperature <span class="hlt">rises</span> produced by the heating of defects under high amplitude ultrasound excitation and is primarily addressed to flaws with contacting faces, such as kissing cracks or tight delaminations. The simplicity of application and the ability to detect small cracks in challenging media makes it an attractive emerging technology, which is still in a development stage. However, it has proven to provide an opportunity for the quantitative characterization of defects, mainly of <span class="hlt">vertical</span> cracks. In this review, we present the principles of the technique and the different experimental implementations, we put it in context with other nondestructive tests and we summarize the work done in order to improve defect detectability and test reliability, with the final goal of determining the probability of detection. Then we review the contributions aimed at characterizing <span class="hlt">vertical</span> cracks, i.e. retrieving the geometry and location of the crack from surface temperature data, generated by ultrasonic excitation.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_13 --> <div id="page_14" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="261"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.2807R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.2807R"><span>The global signature of post-1900 land ice wastage on <span class="hlt">vertical</span> land motion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Riva, Riccardo; Frederikse, Thomas; King, Matt; Marzeion, Ben; van den Broeke, Michiel</p> <p>2017-04-01</p> <p>The amount of ice stored on land has strongly declined during the 20th century, and melt rates showed a significant acceleration over the last two decades. Land ice wastage is well known to be one of the main drivers of global mean sea-level <span class="hlt">rise</span>, as widely discussed in the literature and reflected in the last assessment report of the IPCC. A less obvious effect of melting land ice is the response of the solid earth to mass redistribution on its surface, which, in the first approximation, results in land uplift where the load reduces (e.g., close to the meltwater sources) and land subsidence where the load increases (e.g., under the <span class="hlt">rising</span> oceans). This effect is nowadays well known within the cryospheric and sea level communities. However, what is often not realized is that the solid earth response is a truly global effect: a localized mass change does cause a large deformation signal in its proximity, but also causes a change of the position of every other point on the Earth's surface. The theory of the Earth's elastic response to changing surface loads forms the basis of the 'sea-level equation', which allows sea-level fingerprints of continental mass change to be computed. In this paper, we provide the first dedicated analysis of global <span class="hlt">vertical</span> land motion driven by land ice wastage. By means of established techniques to compute the solid earth elastic response to surface load changes and the most recent datasets of glacier and ice sheet mass change, we show that land ice loss currently leads to <span class="hlt">vertical</span> deformation rates of several tenths of mm per year at mid-latitudes, especially over the Northern Hemisphere where most sources are located. In combination with the improved accuracy of space geodetic techniques (e.g., Global Navigation Satellite Systems), this means that the effect of ice melt is non-negligible over a large part of the continents. In particular, we show how deformation rates have been strongly varying through the last century, which implies</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17053668','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17053668"><span>Movement characteristics of persons with prader-willi syndrome <span class="hlt">rising</span> from supine.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Belt, A B; Hertel, T A; Mante, J R; Marks, T; Rockett, V L; Wade, C; Clayton-Krasinski, D</p> <p>2001-01-01</p> <p>The purposes of this study were to: <span class="hlt">1</span>) determine if previously published descriptors of the supine to stand <span class="hlt">rising</span> task in healthy individuals could be applied to the movements of persons with Prader-Willi Syndrome (PWS); and 2) assess upper extremity (UE), axial region (AX), and lower extremity (LE) movements among subjects with PWS compared with controls. Nine subjects with PWS (seven-36 years of age) and matched controls were videotaped performing 10 <span class="hlt">rising</span> trials. The UE, AX, and LE movements were classified using published descriptors. Occurrence frequencies of movement patterns, duration of movement, and the relationships among body region movement score, BMI, and age were determined. Subjects with PWS utilized developmentally less advanced asymmetrical <span class="hlt">rising</span> patterns, took longer to <span class="hlt">rise</span>, and demonstrated less within subject variability than controls. Categorical descriptors, with minor modifications, can be used to describe <span class="hlt">rising</span> movements in persons with PWS. Knowledge of successful <span class="hlt">rising</span> patterns may assist PTs when examining or planning intervention strategies for teaching the <span class="hlt">rising</span> task.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018CSR...152...27C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018CSR...152...27C"><span>Mean relative sea level <span class="hlt">rise</span> along the coasts of the China Seas from mid-20th to 21st centuries</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Nan; Han, Guoqi; Yang, Jingsong</p> <p>2018-01-01</p> <p>Mean relative sea level (MRSL) <span class="hlt">rise</span> has caused more frequent flooding in many parts of the world. The MRSL <span class="hlt">rise</span> varies substantially from place to place. Here we use tide-gauge data and satellite measurements to examine past MRSL trends for the coasts of the China Seas. We then combine climate model output and satellite observations to provide MRSL projections in the 21st century. The MRSL trend based on tide-gauge data shows substantial regional variations, from <span class="hlt">1</span> to 5 mm/yr. The <span class="hlt">vertical</span> land motion (VLM) based on altimetry and tide-gauge (ATG) data indicates large land subsidence at some tide-gauge locations, consistent with the Global Positioning Systems (GPS)-based VLM but different significantly from small uplift estimated by a Glacial Isostatic Adjustment (GIA) model, which suggests other important factors causing the VLM instead of the GIA process. When GPS- or ATG-based VLM estimates are used, the projected MRSL <span class="hlt">rise</span> between 1986-2005 and 2081-2100 at tide-gauge sites varies from 60 to 130 cm under the Representative Concentration Pathway 8.5 (RCP8.5) scenario of the Intergovernmental Panel on Climate Change (IPCC). Our projections are significantly larger than those of IPCC and other literature, as a result of accounting for the land subsidence derived from observations. Steric and dynamic ocean effects and land-ice melt effects are comparable (about 30 cm each) and do not vary much over the tide-gauge locations. The VLM effect varies from -10 to 60 cm. The projections between 1986-2005 and 2081-2100 under RCP4.5 show a similar spatial distribution to that under RCP8.5, with a smaller amount of <span class="hlt">rise</span> by 18 cm on average for this region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26641954','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26641954"><span>Fast Electrically Driven Capillary <span class="hlt">Rise</span> Using Overdrive Voltage.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hong, Sung Jin; Hong, Jiwoo; Seo, Hee Won; Lee, Sang Joon; Chung, Sang Kug</p> <p>2015-12-29</p> <p>Enhancement of response speed (or reduction of response time) is crucial for the commercialization of devices based on electrowetting (EW), such as liquid lenses and reflective displays, and presents one of the main challenges in EW research studies. We demonstrate here that an overdrive EW actuation gives <span class="hlt">rise</span> to a faster <span class="hlt">rise</span> of a liquid column between parallel electrodes, compared to a DC EW actuation. Here, DC actuation is actually a simple applied step function, and overdrive is an applied step followed by reduction to a lower voltage. Transient behaviors and response time (i.e., the time required to reach the equilibrium height) of the <span class="hlt">rising</span> liquid column are explored under different DC and overdrive EW actuations. When the liquid column <span class="hlt">rises</span> up to a target height by means of an overdrive EW, the response time is reduced to as low as <span class="hlt">1</span>/6 of the response time using DC EW. We develop a theoretical model to simulate the EW-driven capillary <span class="hlt">rise</span> by combining the kinetic equation of capillary flow (i.e., Lucas-Washburn equation) and the dynamic contact angle model considering contact line friction, contact angle hysteresis, contact angle saturation, and the EW effect. This theoretical model accurately predicts the outcome to within a ± 5% error in regard to the <span class="hlt">rising</span> behaviors of the liquid column with a low viscosity, under both DC EW and overdrive actuation conditions, except for the early stage (<about 20 ms).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA09361&hterms=Artistic&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DArtistic','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA09361&hterms=Artistic&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DArtistic"><span>Europa <span class="hlt">Rising</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2007-01-01</p> <p><p/> New Horizons took this image of the icy moon Europa <span class="hlt">rising</span> above Jupiter's cloud tops with its Long Range Reconnaissance Imager (LORRI) at 11:48 Universal Time on February 28, 2007, six hours after the spacecraft's closest approach to Jupiter. <p/> The picture was one of a handful of the Jupiter system that New Horizons took primarily for artistic, rather than scientific, value. This particular scene was suggested by space enthusiast Richard Hendricks of Austin, Texas, in response to an Internet request by New Horizons scientists for evocative, artistic imaging opportunities at Jupiter. <p/> The spacecraft was 2.3 million kilometers (<span class="hlt">1</span>.4 million miles) from Jupiter and 3 million kilometers (<span class="hlt">1</span>.8 million miles) from Europa when the picture was taken. Europa's diameter is 3,120 kilometers (<span class="hlt">1</span>,939 miles). The image is centered on Europa coordinates 5 degrees south, 6 degrees west. In keeping with its artistic intent - and to provide a more dramatic perspective - the image has been rotated so south is at the top.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=security+AND+protection&pg=4&id=EJ632676','ERIC'); return false;" href="https://eric.ed.gov/?q=security+AND+protection&pg=4&id=EJ632676"><span>Safety <span class="hlt">Rises</span> to New Levels.</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Lafo, Joseph; Robillard, Marc</p> <p>2001-01-01</p> <p>Explains how high-<span class="hlt">rise</span> residence halls can provide high-level safety and security at colleges and universities. Boston University is used to illustrate high-<span class="hlt">rise</span> security and fire protection issues. (GR)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017OcMod.113...50S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017OcMod.113...50S"><span><span class="hlt">Vertical</span> resolution of baroclinic modes in global ocean models</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stewart, K. D.; Hogg, A. McC.; Griffies, S. M.; Heerdegen, A. P.; Ward, M. L.; Spence, P.; England, M. H.</p> <p>2017-05-01</p> <p>Improvements in the horizontal resolution of global ocean models, motivated by the horizontal resolution requirements for specific flow features, has advanced modelling capabilities into the dynamical regime dominated by mesoscale variability. In contrast, the choice of the <span class="hlt">vertical</span> grid remains a subjective choice, and it is not clear that efforts to improve <span class="hlt">vertical</span> resolution adequately support their horizontal counterparts. Indeed, considering that the bulk of the <span class="hlt">vertical</span> ocean dynamics (including convection) are parameterized, it is not immediately obvious what the <span class="hlt">vertical</span> grid is supposed to resolve. Here, we propose that the primary purpose of the <span class="hlt">vertical</span> grid in a hydrostatic ocean model is to resolve the <span class="hlt">vertical</span> structure of horizontal flows, rather than to resolve <span class="hlt">vertical</span> motion. With this principle we construct <span class="hlt">vertical</span> grids based on their abilities to represent baroclinic modal structures commensurate with the theoretical capabilities of a given horizontal grid. This approach is designed to ensure that the <span class="hlt">vertical</span> grids of global ocean models complement (and, importantly, to not undermine) the resolution capabilities of the horizontal grid. We find that for z-coordinate global ocean models, at least 50 well-positioned <span class="hlt">vertical</span> levels are required to resolve the first baroclinic mode, with an additional 25 levels per subsequent mode. High-resolution ocean-sea ice simulations are used to illustrate some of the dynamical enhancements gained by improving the <span class="hlt">vertical</span> resolution of a <span class="hlt">1</span>/10° global ocean model. These enhancements include substantial increases in the sea surface height variance (∼30% increase south of 40°S), the barotropic and baroclinic eddy kinetic energies (up to 200% increase on and surrounding the Antarctic continental shelf and slopes), and the overturning streamfunction in potential density space (near-tripling of the Antarctic Bottom Water cell at 65°S).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA266714','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA266714"><span>Low Threshold Voltage Continuous Wave <span class="hlt">Vertical</span>-Cavity Surface-Emitting Lasers</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1993-04-26</p> <p>Data are presented demonstrating a design and fabrication process for the realization of low- threshold , high-output <span class="hlt">vertical</span>-cavity surface-emitting...layers), the low series resistance of the design results in a bias voltage on o <span class="hlt">1</span>.8 V at a threshold current of <span class="hlt">1</span>.9 mA for 10-micrometer-diam devices.... <span class="hlt">Vertical</span>-cavity surface-emitting lasers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19780010630','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19780010630"><span>Users' instructions for the NASA/MSFC cloud-<span class="hlt">rise</span> preprocessor program, version 6, and the NASA/MSFC multilayer diffusion program, version 6: Research version for Univac 1108 system</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bjorklund, J. R.</p> <p>1978-01-01</p> <p>The cloud-<span class="hlt">rise</span> preprocessor and multilayer diffusion computer programs were used by NASA in predicting concentrations and dosages downwind from normal and abnormal launches of rocket vehicles. These programs incorporated: (<span class="hlt">1</span>) the latest data for the heat content and chemistry of rocket exhaust clouds; (2) provision for the automated calculation of surface water pH due to deposition of HCl from precipitation scavenging; (3) provision for automated calculation of concentration and dosage parameters at any level within the <span class="hlt">vertical</span> grounds for which meteorological inputs have been specified; and (4) provision for execution of multiple cases of meteorological data. Procedures used to automatically calculate wind direction shear in a layer were updated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24626415','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24626415"><span>Risk factors of HIV-<span class="hlt">1</span> <span class="hlt">vertical</span> transmission (VT) and the influence of antiretroviral therapy (ART) in pregnancy outcome.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Barral, Maria F M; de Oliveira, Gisele R; Lobato, Rubens C; Mendoza-Sassi, Raul A; Martínez, Ana M B; Gonçalves, Carla V</p> <p>2014-01-01</p> <p>In the absence of intervention, the rate of <span class="hlt">vertical</span> transmission of HIV can range from 15-45%. With the inclusion of antiretroviral drugs during pregnancy and the choice of delivery route this amounts to less than 2%. However ARV use during pregnancy has generated several questions regarding the adverse effects of the gestational and neonatal outcome. This study aims to analyze the risk factors for <span class="hlt">vertical</span> transmission of HIV-<span class="hlt">1</span> seropositive pregnant women living in Rio Grande and the influence of the use of ARVs in pregnancy outcome. Among the 262 pregnant women studied the rate of <span class="hlt">vertical</span> transmission of HIV was found to be 3.8%. Regarding the VT, there was a lower risk of transmission when antiretroviral drugs were used and prenatal care was conducted at the referral service. However, the use of ART did not influence the outcome of pregnancy. However, initiation of prenatal care after the first trimester had an influence on low birth weight, as well as performance of less than six visits increased the risk of prematurity. Therefore, the risk factors analyzed in this study appear to be related to the realization of inadequate pre-natal and maternal behavior.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1436580','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1436580"><span>Abrupt Bølling warming and ice saddle collapse contributions to the Meltwater Pulse <span class="hlt">1</span>a rapid sea level <span class="hlt">rise</span>: North American MWP<span class="hlt">1</span>a Contribution</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Gregoire, Lauren J.; Otto-Bliesner, Bette; Valdes, Paul J.</p> <p></p> <p>Elucidating the source(s) of Meltwater Pulse <span class="hlt">1</span>a, the largest rapid sea level <span class="hlt">rise</span> caused by ice melt (14-18 m in less than 340 years, 14,600 years ago), is important for understanding mechanisms of rapid ice melt and the links with abrupt climate change. Here we quantify how much and by what mechanisms the North American ice sheet could have contributed to Meltwater Pulse <span class="hlt">1</span>a, by driving an ice sheet model with two transient climate simulations of the last 21,000 years. Ice sheet perturbed physics ensembles were run to account for model uncertainties, constraining ice extent and volume with reconstructions ofmore » 21,000 years ago to present. We determine that the North American ice sheet produced 3-4 m global mean sea level <span class="hlt">rise</span> in 340 years due to the abrupt Bølling warming, but this response is amplified to 5-6 m when it triggers the ice sheet saddle collapse.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1436580-abrupt-blling-warming-ice-saddle-collapse-contributions-meltwater-pulse-rapid-sea-level-rise-north-american-mwp1a-contribution','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1436580-abrupt-blling-warming-ice-saddle-collapse-contributions-meltwater-pulse-rapid-sea-level-rise-north-american-mwp1a-contribution"><span>Abrupt Bølling warming and ice saddle collapse contributions to the Meltwater Pulse <span class="hlt">1</span>a rapid sea level <span class="hlt">rise</span>: North American MWP<span class="hlt">1</span>a Contribution</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Gregoire, Lauren J.; Otto-Bliesner, Bette; Valdes, Paul J.; ...</p> <p>2016-08-23</p> <p>Elucidating the source(s) of Meltwater Pulse <span class="hlt">1</span>a, the largest rapid sea level <span class="hlt">rise</span> caused by ice melt (14-18 m in less than 340 years, 14,600 years ago), is important for understanding mechanisms of rapid ice melt and the links with abrupt climate change. Here we quantify how much and by what mechanisms the North American ice sheet could have contributed to Meltwater Pulse <span class="hlt">1</span>a, by driving an ice sheet model with two transient climate simulations of the last 21,000 years. Ice sheet perturbed physics ensembles were run to account for model uncertainties, constraining ice extent and volume with reconstructions ofmore » 21,000 years ago to present. We determine that the North American ice sheet produced 3-4 m global mean sea level <span class="hlt">rise</span> in 340 years due to the abrupt Bølling warming, but this response is amplified to 5-6 m when it triggers the ice sheet saddle collapse.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DPPNO4011M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DPPNO4011M"><span>Improvement of <span class="hlt">vertical</span> stabilization on KSTAR</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mueller, D.; Bak, J. G.; Boyer, M. D.; Eideitis, N.; Hahn, S. H.; Humphreys, D. A.; Kim, H. S.; Jeon, Y. M.; Lanctot, M.; Walker, M. L.</p> <p>2017-10-01</p> <p>The successful control of strongly shaped plasmas on the Korea Superconducting Tokamak Advanced Research (KSTAR) device requires active feedback of fast motion of the plasma <span class="hlt">vertical</span> position by the use of internal normal conducting coils (IVC). This has required new electronics to supply relative flux loop differences, for zp, and voltage loop differences, for dzp/dt, as well as a novel technique (Zfast) to use a high-pass filter, typically <span class="hlt">1</span> Hz, on the error in the signal in the feedback loop. Use of Zfast avoids the potential contention encountered when the internal coil attempts to perform control of the plasma shape which should be controlled by the slower and more powerful superconducting coils. A common problem of this contention is saturation of the IVC and loss of fast <span class="hlt">vertical</span> control. This is eliminated by proper use of the Zfast. A Ziegler-Nichols relay feedback system was used to fine tune the required feedback gains. The selection of the magnetic sensors, filter time constants, control gains and of the Zfast control strategy which allowed <span class="hlt">vertically</span> stable operation at a plasma elongation, kappa. of up to 2.16 at li = <span class="hlt">1</span>.15 and Betap = 2.4 will be discussed which is beyond the design reference of KSTAR of kappa = 2.0 at li = <span class="hlt">1</span>.2 and Betap = <span class="hlt">1</span>.9. Work Supported by U.S.D.O.E. Contract No. DE-AC02-09CH11466 and DE-SC0010685 and the KSTAR project.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A34A..06W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A34A..06W"><span>Retrieving <span class="hlt">Vertical</span> Air Motion and Raindrop Size Distributions from <span class="hlt">Vertically</span> Pointing Doppler Radars</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Williams, C. R.; Chandra, C. V.</p> <p>2017-12-01</p> <p>The <span class="hlt">vertical</span> evolution of falling raindrops is a result of evaporation, breakup, and coalescence acting upon those raindrops. Computing these processes using <span class="hlt">vertically</span> pointing radar observations is a two-step process. First, the raindrop size distribution (DSD) and <span class="hlt">vertical</span> air motion need to be estimated throughout the rain shaft. Then, the changes in DSD properties need to be quantified as a function of height. The change in liquid water content is a measure of evaporation, and the change in raindrop number concentration and size are indicators of net breakup or coalescence in the <span class="hlt">vertical</span> column. The DSD and air motion can be retrieved using observations from two <span class="hlt">vertically</span> pointing radars operating side-by-side and at two different wavelengths. While both radars are observing the same raindrop distribution, they measure different reflectivity and radial velocities due to Rayleigh and Mie scattering properties. As long as raindrops with diameters greater than approximately 2 mm are in the radar pulse volumes, the Rayleigh and Mie scattering signatures are unique enough to estimate DSD parameters using radars operating at 3- and 35-GHz (Williams et al. 2016). <span class="hlt">Vertical</span> decomposition diagrams (Williams 2016) are used to explore the processes acting on the raindrops. Specifically, changes in liquid water content with height quantify evaporation or accretion. When the raindrops are not evaporating, net raindrop breakup and coalescence are identified by changes in the total number of raindrops and changes in the DSD effective shape as the raindrops. This presentation will focus on describing the DSD and air motion retrieval method using <span class="hlt">vertical</span> profiling radar observations from the Department of Energy (DOE) Atmospheric Radiation Measurement (ARM) Southern Great Plains (SGP) central facility in Northern Oklahoma.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27199886','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27199886"><span>A <span class="hlt">1</span>-Diopter <span class="hlt">Vertical</span> Prism Induces a Decrease of Head Rotation: A Pilot Investigation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Matheron, Eric; Zandi, Ava; Wang, Danping; Kapoula, Zoï</p> <p>2016-01-01</p> <p>Clinical studies in non-specific chronic arthralgia and back pain seem to indicate an association between <span class="hlt">vertical</span> heterophoria (VH - latent <span class="hlt">vertical</span> retinal misalignment) and asymmetrical head rotation. Such clinical observations suggest a link between VH and head rotation, but this was never tested. The purpose of this study was to simulate a VH in healthy subjects and examine its influence on the amplitude of active head rotation during 3D motion capture in upright stance. Subjects were asked to rotate their head three times from the straight ahead position and then to the right, back to straight ahead, to the left, and back to the straight ahead again. Three randomized conditions were run: normal viewing, with a <span class="hlt">1</span>-diopter prism base down on the dominant eye, or the non-dominant eye. The most important finding is that the experimental VH whichever the eye with the prism induces a significant decrease in the mean angle of head rotation compared to the normal viewing condition. This decrease was significant for rotation to the left. We suggest that the prism-induced VH modifies the reference posture and thereby affects head rotation; further studies are needed to confirm this effect and to extend to other types of dynamic activities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvP...7b4016O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvP...7b4016O"><span><span class="hlt">Vertical</span> Hole Transport and Carrier Localization in InAs /InAs<span class="hlt">1</span> -xSbx Type-II Superlattice Heterojunction Bipolar Transistors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Olson, B. V.; Klem, J. F.; Kadlec, E. A.; Kim, J. K.; Goldflam, M. D.; Hawkins, S. D.; Tauke-Pedretti, A.; Coon, W. T.; Fortune, T. R.; Shaner, E. A.; Flatté, M. E.</p> <p>2017-02-01</p> <p>Heterojunction bipolar transistors are used to measure <span class="hlt">vertical</span> hole transport in narrow-band-gap InAs /InAs<span class="hlt">1</span> -xSbx type-II superlattices (T2SLs). <span class="hlt">Vertical</span> hole mobilities (μh) are reported and found to decrease rapidly from 360 cm2/V s at 120 K to approximately 2 cm2/V s at 30 K, providing evidence that holes are confined to localized states near the T2SL valence-miniband edge at low temperatures. Four distinct transport regimes are identified: (<span class="hlt">1</span>) pure miniband transport, (2) miniband transport degraded by temporary capture of holes in localized states, (3) hopping transport between localized states in a mobility edge, and (4) hopping transport through defect states near the T2SL valence-miniband edge. Region (2) is found to have a thermal activation energy of ɛ2=36 meV corresponding to the energy range of a mobility edge. Region (3) is found to have a thermal activation energy of ɛ3=16 meV corresponding to the hopping transport activation energy. This description of <span class="hlt">vertical</span> hole transport is analogous to electronic transport observed in disordered amorphous semiconductors displaying Anderson localization. For the T2SL, we postulate that localized states are created by disorder in the group-V alloy of the InAs<span class="hlt">1</span> -xSbx hole well causing fluctuations in the T2SL valence-band energy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29671322','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29671322"><span>Utilizing Interlayer Excitons in Bilayer WS2 for Increased Photovoltaic Response in Ultrathin Graphene <span class="hlt">Vertical</span> Cross-Bar Photodetecting Tunneling Transistors.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhou, Yingqiu; Tan, Haijie; Sheng, Yuewen; Fan, Ye; Xu, Wenshuo; Warner, Jamie H</p> <p>2018-04-19</p> <p>Here we study the layer-dependent photoconductivity in Gr/WS 2 /Gr <span class="hlt">vertical</span> stacked tunneling (VST) cross-bar devices made using two-dimensional (2D) materials all grown by chemical vapor deposition. The larger number of devices (>100) enables a statistically robust analysis on the comparative differences in the photovoltaic response of monolayer and bilayer WS 2 , which cannot be achieved in small batch devices made using mechanically exfoliated materials. We show a dramatic increase in photovoltaic response for Gr/WS 2 (2L)/Gr compared to monolayers because of the long inter- and intralayer exciton lifetimes and the small exciton binding energy (both interlayer and intralayer excitons) of bilayer WS 2 compared with that of monolayer WS 2 . Different doping levels and dielectric environments of top and bottom graphene electrodes result in a potential difference across a ∼<span class="hlt">1</span> nm <span class="hlt">vertical</span> device, which gives <span class="hlt">rise</span> to large electric fields perpendicular to the WS 2 layers that cause band structure modification. Our results show how precise control over layer number in all 2D VST devices dictates the photophysics and performance for photosensing applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JHyd..557..688I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JHyd..557..688I"><span>Comparison of a <span class="hlt">vertically</span>-averaged and a <span class="hlt">vertically</span>-resolved model for hyporheic flow beneath a pool-riffle bedform</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ibrahim, Ahmad; Steffler, Peter; She, Yuntong</p> <p>2018-02-01</p> <p>The interaction between surface water and groundwater through the hyporheic zone is recognized to be important as it impacts the water quantity and quality in both flow systems. Three-dimensional (3D) modeling is the most complete representation of a real-world hyporheic zone. However, 3D modeling requires extreme computational power and efforts; the sophistication is often significantly compromised by not being able to obtain the required input data accurately. Simplifications are therefore often needed. The objective of this study was to assess the accuracy of the <span class="hlt">vertically</span>-averaged approximation compared to a more complete <span class="hlt">vertically</span>-resolved model of the hyporheic zone. The groundwater flow was modeled by either a simple one-dimensional (<span class="hlt">1</span>D) Dupuit approach or a two-dimensional (2D) horizontal/<span class="hlt">vertical</span> model in boundary fitted coordinates, with the latter considered as a reference model. Both groundwater models were coupled with a <span class="hlt">1</span>D surface water model via the surface water depth. Applying the two models to an idealized pool-riffle sequence showed that the <span class="hlt">1</span>D Dupuit approximation gave comparable results in determining the characteristics of the hyporheic zone to the reference model when the stratum thickness is not very large compared to the surface water depth. Conditions under which the <span class="hlt">1</span>D model can provide reliable estimate of the seepage discharge, upwelling/downwelling discharges and locations, the hyporheic flow, and the residence time were determined.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1996JAtS...53.1887B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1996JAtS...53.1887B"><span><span class="hlt">Vertical</span> Motion Characteristics of Tropical Cyclones Determined with Airborne Doppler Radial Velocities.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Black, Micheal L.; Burpee, Robert W.; Marks, Frank D., Jr.</p> <p>1996-07-01</p> <p><span class="hlt">Vertical</span> motions in seven Atlantic hurricanes are determined from data recorded by Doppler radars on research aircraft. The database consists of Doppler velocities and reflectivities from <span class="hlt">vertically</span> pointing radar rays collected along radial flight legs through the hurricane centers. The <span class="hlt">vertical</span> motions are estimated throughout the depth of the troposphere from the Doppler velocities and bulk estimates of particle fallspeeds.Portions of the flight tracks are subjectively divided into eyewall, rainband, stratiform, and `other' regions. Characteristics of the <span class="hlt">vertical</span> velocity and radar structure are described as a function of altitude for the entire dataset and each of the four regions. In all of the regions, more than 70% of the <span class="hlt">vertical</span> velocities range from 2 to 2 m s<span class="hlt">1</span>. The broadest distribution of <span class="hlt">vertical</span> motion is in the eyewall region where 5% of the <span class="hlt">vertical</span> motions are >5 m s<span class="hlt">1</span>. Averaged over the entire dataset, the mean <span class="hlt">vertical</span> velocity is upward at all altitudes. Mean downward motion occurs only in the lower troposphere of the stratiform region. Significant <span class="hlt">vertical</span> variations in the mean profiles of <span class="hlt">vertical</span> velocity and reflectivity are discussed and related to microphysical processes.In the lower and middle troposphere, the characteristics of the Doppler-derived <span class="hlt">vertical</span> motions are similar to those described in an earlier study using flight-level <span class="hlt">vertical</span> velocities, even though the horizontal resolution of the Doppler data is 750 m compared to 125 m from the in situ flight-level measurements. The Doppler data are available at higher altitudes than those reached by turboprop aircraft and provide information on <span class="hlt">vertical</span> as well as horizontal variations. In a <span class="hlt">vertical</span> plane along the radial flight tracks, Doppler up- and downdrafts are defined at each 300-m altitude interval as <span class="hlt">vertical</span> velocities whose absolute values continuously exceed <span class="hlt">1</span>.5 m s<span class="hlt">1</span>, with at least one speed having an absolute value greater than 3.0 m s<span class="hlt">1</span>. The properties of the Doppler</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20060040701&hterms=Ph&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3DPh','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20060040701&hterms=Ph&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3DPh"><span><span class="hlt">Vertical</span> Distribution of PH(sub 3) in Saturn from Observations of the <span class="hlt">1</span>-0 and 3-2 Rotational Lines</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Orton, G.; Serabyn, E.; Lee, Y.</p> <p>1999-01-01</p> <p>Far-infrared Fourier-transform spectrometer measurements of the <span class="hlt">1</span>-0 and 3-2 PH(sub 3) transitions in Saturn's disk near 267 and 800 GHz (8.9 and 26.7 cm(sup -<span class="hlt">1</span>)), respectively, were analyzed simultaneously to derive a global mean profile for the PH(sub 3) <span class="hlt">vertical</span> mixing ratio between 100 and 800 mbar total pressure.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_14 --> <div id="page_15" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="281"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSA53C..05P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSA53C..05P"><span>Using Meteoric Ablation to Constrain <span class="hlt">Vertical</span> Transport in the Upper Mesosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Plane, J. M. C.; Carrillo-Sánchez, J. D.; Nesvorny, D.; Pokorný, P.; Janches, D.</p> <p>2016-12-01</p> <p>Meteoric ablation injects a variety of metals into the upper mesosphere and lower thermosphere, giving <span class="hlt">rise</span> to layers of metal atoms centered around 90 km. The Na, Fe, K and Ca atom densities are measured accurately using resonance lidars. Since the reaction kinetics of many of the chemical reactions which produce these layers have now been studied in the laboratory, chemistry modules for each of the metals have been developed with a reasonable degree of confidence. When these modules are put into a global high-top model such as NCAR's Whole Atmosphere Community Climate Model (WACCM), a major problem emerges: the injection flux of each of the metals, termed the Meteoric Input Function (MIF), has to be reduced substantially in order to model the observed metal atom densities. For instance, the Na and Fe MIFs need to be reduced by factors of 8 and 14, respectively, compared with the MIFs determined from the lidar-measured <span class="hlt">vertical</span> fluxes of Na and Fe atoms. The accumulation of meteoric smoke particles in polar ice cores also indicates that the meteoric ablation flux is significantly larger that can be handled in models where <span class="hlt">vertical</span> transport is solely due to eddy diffusional mixing. Here we derive new Na and Fe MIFs by determining the relative contributions of the known dust sources in the near-Earth environment: Jupiter Family Comets (JFCs), the main asteroid belt, Halley Type comets, and Oort Cloud comets. The mass/velocity/radiant distributions of these cosmic dust populations are Monte Carlo sampled and the elemental ablation rates calculated with the Leeds Chemical Ablation Model. The contribution of each dust source in the Earth's atmosphere is then determined by fitting the measured cosmic spherule accretion rate at the South Pole, and the measured <span class="hlt">vertical</span> Na and Fe fluxes above 86 km. We conclude that JFCs contribute either 85% or 93% to the total incoming mass, depending on whether infra-red observations of the Zodiacal Dust Cloud by the IRAS or Planck</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/37787','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/37787"><span>Arthropod <span class="hlt">vertical</span> stratification in temperate deciduous forests: Implications for conservation oriented management</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Ulyshen Michael</p> <p>2011-01-01</p> <p>Studies on the <span class="hlt">vertical</span> distribution patterns of arthropods in temperate deciduous forests reveal highly stratified (i.e., unevenly <span class="hlt">vertically</span> distributed) communities. These patterns are determined by multiple factors acting simultaneously, including: (<span class="hlt">1</span>) time (forest age, season, time of day); (2) forest structure (height, <span class="hlt">vertical</span> foliage complexity, plant surface...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ApPhL.109z3101C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ApPhL.109z3101C"><span>High-performance PbS quantum dot <span class="hlt">vertical</span> field-effect phototransistor using graphene as a transparent electrode</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Che, Yongli; Zhang, Yating; Cao, Xiaolong; Song, Xiaoxian; Zhang, Haiting; Cao, Mingxuan; Dai, Haitao; Yang, Junbo; Zhang, Guizhong; Yao, Jianquan</p> <p>2016-12-01</p> <p>Solution processed photoactive PbS quantum dots (QDs) were used as channel in high-performance near-infrared <span class="hlt">vertical</span> field-effect phototransistor (VFEpT) where monolayer graphene embedded as transparent electrode. In this <span class="hlt">vertical</span> architecture, the PbS QD channel was sandwiched and naturally protected between the drain and source electrodes, which made the device ultrashort channel length (110 nm) simply the thickness of the channel layer. The VFEpT exhibited ambipolar operation with high mobilities of μe = 3.5 cm2/V s in n-channel operation and μh = 3.3 cm2/V s in p-channel operation at low operation voltages. By using the photoactive PbS QDs as channel material, the VFEpT exhibited good photoresponse properties with a responsivity of 4.2 × 102 A/W, an external quantum efficiency of 6.4 × 104% and a photodetectivity of 2.<span class="hlt">1</span> × 109 Jones at the light irradiance of 36 mW/cm2. Additionally, the VFEpT showed excellent on/off switching with good stability and reproducibility and fast response speed with a short <span class="hlt">rise</span> time of 12 ms in n-channel operation and 10.6 ms in p-channel operation. These high mobilities, good photoresponse properties and simplistic fabrication of our VFEpTs provided a facile route to the high-performance inorganic photodetectors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1012592','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1012592"><span>Cryogenic <span class="hlt">vertical</span> test facility for the SRF cavities at BNL</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Than, R.; Liaw, CJ; Porqueddu, R.</p> <p>2011-03-28</p> <p>A <span class="hlt">vertical</span> test facility has been constructed to test SRF cavities and can be utilized for other applications. The liquid helium volume for the large <span class="hlt">vertical</span> dewar is approximate 2.<span class="hlt">1</span>m tall by <span class="hlt">1</span>m diameter with a clearance inner diameter of 0.95m after the inner cold magnetic shield installed. For radiation enclosure, the test dewar is located inside a concrete block structure. The structure is above ground, accessible from the top, and equipped with a retractable concrete roof. A second radiation concrete facility, with ground level access via a labyrinth, is also available for testing smaller cavities in 2 smaller dewars.more » The cryogenic transfer lines installation between the large <span class="hlt">vertical</span> test dewar and the cryo plant's sub components is currently near completion. Controls and instrumentations wiring are also nearing completion. The <span class="hlt">Vertical</span> Test Facility will allow onsite testing of SRF cavities with a maximum overall envelope of 0.9 m diameter and 2.<span class="hlt">1</span> m height in the large dewar and smaller SRF cavities and assemblies with a maximum overall envelope of 0.66 m diameter and <span class="hlt">1</span>.6 m height.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23107911','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23107911"><span>Muscle activation history at different <span class="hlt">vertical</span> jumps and its influence on <span class="hlt">vertical</span> velocity.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kopper, Bence; Csende, Zsolt; Sáfár, Sándor; Hortobágyi, Tibor; Tihanyi, József</p> <p>2013-02-01</p> <p>In the present study we investigated displacement, time, velocity and acceleration history of center of mass (COM) and electrical activity of knee extensors to estimate the dominance of the factors influencing the <span class="hlt">vertical</span> velocity in squat jumps (SJs), countermovement jumps (CMJs) and drop jumps (DJs) performed with small (40°) and large (80°) range of joint motion (SROM and LROM). The maximum <span class="hlt">vertical</span> velocity (v4) was 23.4% (CMJ) and 7.8% (DJ) greater when the jumps were performed with LROM compared with SROM (p < 0.05). These differences are considerably less than it could be expected from the greater COM and knee angular displacement and duration of active state. This small difference can be attributed to the greater deceleration during eccentric phase (CMJ:32.<span class="hlt">1</span>%, DJ:91.5%) in SROM than that in LROM. v4 was greater for SJ in LROM than for SJ in SROM indicating the significance of the longer active state and greater activation level (p < 0.001). The difference in v4 was greater between SJ and CMJ in SROM (38.6%) than in LROM (9.0%), suggesting that elastic energy storage and re-use can be a dominant factor in the enhancement of <span class="hlt">vertical</span> velocity of CMJ and DJ compared with SJ performed with SROM. Copyright © 2012 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4630640','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4630640"><span>Coastal sea level projections with improved accounting for <span class="hlt">vertical</span> land motion</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Han, Guoqi; Ma, Zhimin; Chen, Nan; Yang, Jingsong; Chen, Nancy</p> <p>2015-01-01</p> <p>Regional and coastal mean sea level projections in the Intergovernmental Panel for Climate Change (IPCC) Fifth Assessment Report (AR5) account only for <span class="hlt">vertical</span> land motion (VLM) associated with glacial isostatic adjustment (GIA), which may significantly under- or over-estimate sea level <span class="hlt">rise</span>. Here we adjust AR5-like regional projections with the VLM from Global Positioning Satellite (GPS) measurements and/or from a combination of altimetry and tide-gauge data, which include both GIA and non-GIA VLM. Our results at selected tide-gauge locations on the North American and East Asian coasts show drastically different projections with and without non-GIA VLM being accounted for. The present study points to the importance of correcting IPCC AR5 coastal projections for the non-GIA VLM in making adaptation decisions. PMID:26526287</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26526287','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26526287"><span>Coastal sea level projections with improved accounting for <span class="hlt">vertical</span> land motion.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Han, Guoqi; Ma, Zhimin; Chen, Nan; Yang, Jingsong; Chen, Nancy</p> <p>2015-11-03</p> <p>Regional and coastal mean sea level projections in the Intergovernmental Panel for Climate Change (IPCC) Fifth Assessment Report (AR5) account only for <span class="hlt">vertical</span> land motion (VLM) associated with glacial isostatic adjustment (GIA), which may significantly under- or over-estimate sea level <span class="hlt">rise</span>. Here we adjust AR5-like regional projections with the VLM from Global Positioning Satellite (GPS) measurements and/or from a combination of altimetry and tide-gauge data, which include both GIA and non-GIA VLM. Our results at selected tide-gauge locations on the North American and East Asian coasts show drastically different projections with and without non-GIA VLM being accounted for. The present study points to the importance of correcting IPCC AR5 coastal projections for the non-GIA VLM in making adaptation decisions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUOSEC34B1182C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUOSEC34B1182C"><span>Anthropogenic sea level <span class="hlt">rise</span> and adaptation in the Yangtze estuary</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cheng, H.; Chen, J.; Chen, Z.; Ruan, R.; Xu, G.; Zeng, G.; Zhu, J.; Dai, Z.; Gu, S.; Zhang, X.; Wang, H.</p> <p>2016-02-01</p> <p>Sea level <span class="hlt">rise</span> is a major projected threat of climate change. There are regional variations in sea level changes, depending on both naturally the tectonic subsidence, geomorphology, naturally changing river inputs and anthropogenic driven forces as artificial reservoir water impoundment within the watershed and urban land subsidence driven by ground water depletion in the river delta. Little is known on regional sea level fall in response to the channel erosion due to the sediment discharge decline by reservoir interception in the upstream watershed, and water level <span class="hlt">rise</span> driven by anthropogenic measures as the land reclamation, deep waterway regulation and fresh water reservoir construction to the sea level change in estuaries. Changing coastal cities are situated in the delta regions expected to be threatened in various degrees. Shanghai belongs to those cities. Here we show that the anthropogenic driven sea level <span class="hlt">rise</span> in the Yangtze estuary from the point of view of the continuous hydrodynamic system consisted of river catchment, estuary and coastal sea. Land subsidence is cited as 4 mm/a (2011-2030). Scour depth of the estuarine channel by upstream engineering as Three Gauge Dam is estimated at 2-10 cm (2011-2030). The <span class="hlt">rise</span> of water level by deep waterway and land reclamation is estimated at 8-10 cm (2011-2030). The relative sea level <span class="hlt">rise</span> will be speculated about 10 -16 cm (2011-2030), which these anthropogenic sea level changes will be imposed into the absolute sea level <span class="hlt">rise</span> 2 mm/a and tectonic subsidence <span class="hlt">1</span> mm/a measured in 1990s. The action guideline to the sea level <span class="hlt">rise</span> strategy in the Shanghai city have been proposed to the Shanghai government as (<span class="hlt">1</span>) recent actions (2012-2015) to upgrade the city water supply and drainage engineering and protective engineering; (2) interim actions (2016-2020) to improve sea level monitoring and early warning system, and then the special, city, regional planning considering sea level <span class="hlt">rise</span>; (3) long term actions (2021</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5572756-bermuda-appalachian-labrador-rises-common-non-hotspot-processes','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5572756-bermuda-appalachian-labrador-rises-common-non-hotspot-processes"><span>Bermuda and Appalachian-Labrador <span class="hlt">rises</span>: Common non-hotspot processes</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Vogt, P.R.</p> <p>1991-01-01</p> <p>Other than the Corner <span class="hlt">Rise</span>-New England seamounts and associated White Mountains, most postbreakup intraplate igneous activity and topographic uplift in the western North Atlantic and eastern North America do not readily conform to simple hotspot models. For examples, the Bermuda <span class="hlt">Rise</span> trends normal to its predicted hotspot trace. On continental crust, Cretaceous-Eocene igneous activity is scattered along a northeast-trending belt {approximately}500-<span class="hlt">1</span>,000 km west of and paralleling the continent-ocean boundary. Corresponding activity in the western Atlantic generated seamounts preferentially clustered in a belt {approximately}<span class="hlt">1</span>,000 km east of the boundary. The Eocene volcanism on Bermuda is paired with coeval magmatism of themore » Shenandoah igneous province, and both magmatic belts are associated with northeast-trending topographic bulges - the Appalachian-Labrador <span class="hlt">Rise</span> to the west and the Bermuda <span class="hlt">Rise</span> (Eocene ) to the east. The above observations suggest the existence of paired asthenosphere upwelling, paralleling and controlled by the deep thermal contrast across the northeast-trending continental margin. Such convection geometry, apparently fixed to the North American plate rather than to hotspots, is consistent with recent convection models by B. Hager. The additional importance of plate-kinematic reorganizations (causing midplate stress enhancement) is suggested by episodic igneous activity ca. 90-100 Ma and 40-45 Ma.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA280389','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA280389"><span>Bistable <span class="hlt">Vertical</span>-Cavity Surface-Emitting Laser. Structures on GaAs and Si Substrates</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1994-06-01</p> <p><span class="hlt">vertical</span> - cavity surface - emitting lasers ( VCSELs ) [<span class="hlt">1</span>,5,6 of publications below], fabrication processes to realize low...May 91 through <span class="hlt">1</span> June 94 R&T Number: Contract / Grant Number: N00014-91-J-1952 Contract / Grant Title: Bistable <span class="hlt">Vertical</span> - Cavity Surface - Emitting Laser ...T.J. Rogers, B.G. Streetman, S.C. Smith, and R.D. Burnham, "Cascadabity of an Optically Iathing <span class="hlt">Vertical</span> - Cavity Surface - Emitting Laser</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12664353','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12664353"><span>[A new method to test <span class="hlt">vertical</span> ocular deviations using perilimbal light reflexes].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Breyer, Armin; Rütsche, Adrian; Gampe, Elisabeth; Mojon, Daniel S</p> <p>2003-03-01</p> <p>To develop a new diagnostic technique to determine <span class="hlt">vertical</span> ocular deviations when the center of the pupil is covered by swollen eyelids in up- and downgaze. In upgaze (downgaze) the reflex of a diagnostic lamp held at about 50 cm distance from the patient is observed on the lower (upper) limbus. In the case of an asymmetric reflex, prisms are used to obtain symmetrical reflexes. The amount of prisms indicates the size of the <span class="hlt">vertical</span> misalignment. In five healthy volunteers, the angles of <span class="hlt">vertical</span> changes of gaze position were plotted against the prism size needed to recenter the perilimbal reflex. There was a linear correlation between the amount of upgaze changes in degrees and the strength of prisms used for compensation in degrees. This linear correlation was also found in downgaze. For both the correlation coefficient was r = 0.98 +/- 0.01. In upgaze the slope of the average regression line was 0.55 +/- 2.3 degrees, in downgaze - 4.<span class="hlt">1</span> +/- 0.8 degrees. A prism of <span class="hlt">1</span> degrees corresponds in upgaze to a <span class="hlt">vertical</span> deviation of about <span class="hlt">1</span>.3 +/- 0.14 degrees, in downgaze to a deviation of about <span class="hlt">1.1</span> +/- 0.07 degrees. These results demonstrate that the perilimbal light reflex test is suitable for measuring simulated <span class="hlt">vertical</span> ocular deviations. Therefore, the test may also be used in patients with <span class="hlt">vertical</span> deviations who cannot be measured with classical methods. The method is more exact for measurements in upgaze.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003APS..DFD.FL004T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003APS..DFD.FL004T"><span>Buoyancy and Pressure Driven Flow of Hot Gases in <span class="hlt">Vertical</span> Shafts with Natural and Forced Ventilation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tamm, Gunnar; Jaluria, Yogesh</p> <p>2003-11-01</p> <p>An experimental investigation has been carried out on the buoyancy and pressure induced flow of hot gases in <span class="hlt">vertical</span> shafts, in order to simulate the propagation of combustion products in elevator shafts due to fire in multilevel buildings. Various geometrical configurations are studied, with regard to natural and forced ventilation imposed at the top or bottom of the <span class="hlt">vertical</span> shaft. The aspect ratio is taken at a fixed value of 6 and the inflow conditions for the hot gases, at a vent near the bottom, are varied in terms of the Reynolds and Grashof numbers. Temperature measurements within the shaft allow a detailed study of the steady state thermal fields, from which optimal means for smoke alleviation in high-<span class="hlt">rise</span> building fires may be developed. Flow visualization is also used to study the flow characteristics. The results obtained indicate a wall plume as the primary transport mechanism. Flow recirculation dominates at high Grashof number flows, while increased Reynolds numbers gives <span class="hlt">rise</span> to greater mixing in the shaft. The development and stability of the flow and its effect on the spread of smoke and hot gases are assessed for the different shaft configurations and inlet conditions. It is found that the fastest smoke removal and lowest shaft temperatures occur for a configuration with natural ventilation at the top and forced ventilation up from the shaft bottom. It is also shown that forced ventilation can be used to arrest smoke spread, as well as to dilute the effects of the fire.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29248166','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29248166"><span>High temporal resolution modeling of the impact of rain, tides, and sea level <span class="hlt">rise</span> on water table flooding in the Arch Creek basin, Miami-Dade County Florida USA.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sukop, Michael C; Rogers, Martina; Guannel, Greg; Infanti, Johnna M; Hagemann, Katherine</p> <p>2018-03-01</p> <p>Modeling of groundwater levels in a portion of the low-lying coastal Arch Creek basin in northern Miami-Dade County in Southeast Florida USA, which is subject to repetitive flooding, reveals that rain-induced short-term water table <span class="hlt">rises</span> can be viewed as a primary driver of flooding events under current conditions. Areas below 0.9m North American <span class="hlt">Vertical</span> Datum (NAVD) elevation are particularly vulnerable and areas below <span class="hlt">1</span>.5m NAVD are vulnerable to exceptionally large rainfall events. Long-term water table <span class="hlt">rise</span> is evident in the groundwater data, and the rate appears to be consistent with local rates of sea level <span class="hlt">rise</span>. Linear extrapolation of long-term observed groundwater levels to 2060 suggest roughly a doubling of the number of days when groundwater levels exceed 0.9m NAVD and a threefold increase in the number of days when levels exceed <span class="hlt">1</span>.5m NAVD. Projected sea level <span class="hlt">rise</span> of 0.61m by 2060 together with increased rainfall lead to a model prediction of frequent groundwater-related flooding in areas<0.9m NAVD. However, current simulations do not consider the range of rainfall events that have led to water table elevations><span class="hlt">1</span>.5m NAVD and widespread flooding of the area in the past. Tidal fluctuations in the water table are predicted to be more pronounced within 600m of a tidally influenced water control structure that is hydrodynamically connected to Biscayne Bay. The inland influence of tidal fluctuations appears to increase with increased sea level, but the principal driver of high groundwater levels under the 2060 scenario conditions remains groundwater recharge due to rainfall events. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA019708','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA019708"><span>Stratospheric Turbulence and <span class="hlt">Vertical</span> Effective Diffusion Coefficients</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1975-09-29</p> <p>UMBER AFCRL-TR-75.-0519 - 4. TILE (moiS."Eti) S. Tlr OF C RP~hT S PESO0 COVERED STRATOSPHERIC TURBULENCE AND <span class="hlt">VERTICAL</span> EFFECTIVE DIFFUSION COEFFICIENTS...that CAT plays a prominent role in <span class="hlt">vertical</span> transport in the stratosphere. I ~<span class="hlt">1</span> Unclassified t FUrs,*Tv C , Uq C ~ml .. at ’r *n he.. a* U I Department...phenomenon. Thorpe himself refers (1973) to underwater K-H as "underwater CAT." ____ ____ ____WE006 SflJGLE ( SPAD M LAVER 4" Ri" i0 15 0t (m’iJr</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/872600','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/872600"><span>Micromachined electrostatic <span class="hlt">vertical</span> actuator</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Lee, Abraham P.; Sommargren, Gary E.; McConaghy, Charles F.</p> <p></p> <p>A micromachined <span class="hlt">vertical</span> actuator utilizing a levitational force, such as in electrostatic comb drives, provides <span class="hlt">vertical</span> actuation that is relatively linear in actuation for control, and can be readily combined with parallel plate capacitive position sensing for position control. The micromachined electrostatic <span class="hlt">vertical</span> actuator provides accurate movement in the sub-micron to micron ranges which is desirable in the phase modulation instrument, such as optical phase shifting. For example, compact, inexpensive, and position controllable micromirrors utilizing an electrostatic <span class="hlt">vertical</span> actuator can replace the large, expensive, and difficult-to-maintain piezoelectric actuators. A thirty pound piezoelectric actuator with corner cube reflectors, as utilized inmore » a phase shifting diffraction interferometer can be replaced with a micromirror and a lens. For any very precise and small amplitudes of motion` micromachined electrostatic actuation may be used because it is the most compact in size, with low power consumption and has more straightforward sensing and control options.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/20013796-micromachined-electrostatic-vertical-actuator','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/20013796-micromachined-electrostatic-vertical-actuator"><span>Micromachined electrostatic <span class="hlt">vertical</span> actuator</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Lee, A.P.; Sommargren, G.E.; McConaghy, C.F.</p> <p></p> <p>A micromachined <span class="hlt">vertical</span> actuator utilizing a levitational force, such as in electrostatic comb drives, provides <span class="hlt">vertical</span> actuation that is relatively linear in actuation for control, and can be readily combined with parallel plate capacitive position sensing for position control. The micromachined electrostatic <span class="hlt">vertical</span> actuator provides accurate movement in the sub-micron to micron ranges which is desirable in the phase modulation instrument, such as optical phase shifting. For example, compact, inexpensive, and position controllable micromirrors utilizing an electrostatic <span class="hlt">vertical</span> actuator can replace the large, expensive, and difficult-to-maintain piezoelectric actuators. A thirty pound piezoelectric actuator with corner cube reflectors, as utilized inmore » a phase shifting diffraction interferometer can be replaced with a micromirror and a lens. For any very precise and small amplitudes of motion, micromachined electrostatic actuation may be used because it is the most compact in size, with low power consumption and has more straightforward sensing and control options.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title46-vol4/pdf/CFR-2010-title46-vol4-sec108-160.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title46-vol4/pdf/CFR-2010-title46-vol4-sec108-160.pdf"><span>46 CFR 108.160 - <span class="hlt">Vertical</span> ladders.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-10-01</p> <p>... Construction and Arrangement Means of Escape § 108.160 <span class="hlt">Vertical</span> ladders. (a) Each <span class="hlt">vertical</span> ladder must have... <span class="hlt">vertical</span> fixed ladders may be made of wood. [CGD 73-251, 43 FR 56808, Dec. 4, 1978, as amended by USCG-2002...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ApPhL.108f3506K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ApPhL.108f3506K"><span>Enhancing photoresponsivity using MoTe2-graphene <span class="hlt">vertical</span> heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kuiri, Manabendra; Chakraborty, Biswanath; Paul, Arup; Das, Subhadip; Sood, A. K.; Das, Anindya</p> <p>2016-02-01</p> <p>MoTe2 with a narrow band-gap of ˜<span class="hlt">1.1</span> eV is a promising candidate for optoelectronic applications, especially for the near-infrared photo detection. However, the photo responsivity of few layers MoTe2 is very small (<<span class="hlt">1</span> mA W-<span class="hlt">1</span>). In this work, we show that a few layer MoTe2-graphene <span class="hlt">vertical</span> heterostructures have a much larger photo responsivity of ˜20 mA W-<span class="hlt">1</span>. The trans-conductance measurements with back gate voltage show on-off ratio of the <span class="hlt">vertical</span> transistor to be ˜(0.5-<span class="hlt">1</span>) × 105. The rectification nature of the source-drain current with the back gate voltage reveals the presence of a stronger Schottky barrier at the MoTe2-metal contact as compared to the MoTe2-graphene interface. In order to quantify the barrier height, it is essential to measure the work function of a few layers MoTe2, not known so far. We demonstrate a method to determine the work function by measuring the photo-response of the <span class="hlt">vertical</span> transistor as a function of the Schottky barrier height at the MoTe2-graphene interface tuned by electrolytic top gating.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27965238','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27965238"><span><span class="hlt">Vertical</span> stratification of the foliar fungal community in the world's tallest trees.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Harrison, Joshua G; Forister, Matthew L; Parchman, Thomas L; Koch, George W</p> <p>2016-12-01</p> <p>The aboveground tissues of plants host numerous, ecologically important fungi, yet patterns in the spatial distribution of these fungi remain little known. Forest canopies in particular are vast reservoirs of fungal diversity, but intracrown variation in fungal communities has rarely been explored. Knowledge of how fungi are distributed throughout tree crowns will contribute to our understanding of interactions between fungi and their host trees and is a first step toward investigating drivers of community assembly for plant-associated fungi. Here we describe spatial patterns in fungal diversity within crowns of the world's tallest trees, coast redwoods (Sequoia sempervirens). We took a culture-independent approach, using the Illumina MiSeq platform, to characterize the fungal assemblage at multiple heights within the crown across the geographical range of the coast redwood. Within each tree surveyed, we uncovered evidence for <span class="hlt">vertical</span> stratification in the fungal community; different portions of the tree crown harbored different assemblages of fungi. We also report between-tree variation in the fungal community within redwoods. Our results suggest the potential for <span class="hlt">vertical</span> stratification of fungal communities in the crowns of other tall tree species and should prompt future study of the factors giving <span class="hlt">rise</span> to this stratification. © 2016 Botanical Society of America.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1082348','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1082348"><span><span class="hlt">Vertical</span> axis wind turbine airfoil</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Krivcov, Vladimir; Krivospitski, Vladimir; Maksimov, Vasili; Halstead, Richard; Grahov, Jurij Vasiljevich</p> <p>2012-12-18</p> <p>A <span class="hlt">vertical</span> axis wind turbine airfoil is described. The wind turbine airfoil can include a leading edge, a trailing edge, an upper curved surface, a lower curved surface, and a centerline running between the upper surface and the lower surface and from the leading edge to the trailing edge. The airfoil can be configured so that the distance between the centerline and the upper surface is the same as the distance between the centerline and the lower surface at all points along the length of the airfoil. A plurality of such airfoils can be included in a <span class="hlt">vertical</span> axis wind turbine. These airfoils can be <span class="hlt">vertically</span> disposed and can rotate about a <span class="hlt">vertical</span> axis.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_15 --> <div id="page_16" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="301"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012EGUGA..14.3941S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012EGUGA..14.3941S"><span>Sensitivity analysis of hydrogeological parameters affecting groundwater storage change caused by sea level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shin, J.; Kim, K.-H.; Lee, K.-K.</p> <p>2012-04-01</p> <p>Sea level <span class="hlt">rise</span>, which is one of the representative phenomena of climate changes caused by global warming, can affect groundwater system. The <span class="hlt">rising</span> trend of the sea level caused by the global warming is reported to be about 3 mm/year for the most recent 10 year average (IPCC, 2007). The rate of sea level <span class="hlt">rise</span> around the Korean peninsula is reported to be 2.30±2.22 mm/yr during the 1960-1999 period (Cho, 2002) and 2.16±<span class="hlt">1</span>.77 mm/yr (Kim et al., 2009) during the 1968-2007 period. Both of these rates are faster than the <span class="hlt">1</span>.8±0.5 mm/yr global average for the similar 1961-2003 period (IPCC, 2007). In this study, we analyzed changes in the groundwater environment affected by the sea level <span class="hlt">rise</span> by using an analytical methodology. We tried to find the most effective parameters of groundwater amount change in order to estimate the change in fresh water amount in coastal groundwater. A hypothetical island model of a cylindrical shape in considered. The groundwater storage change is bi-directional as the sea level <span class="hlt">rises</span> according to the natural and hydrogeological conditions. Analysis of the computation results shows that topographic slope and hydraulic conductivity are the most sensitive factors. The contributions of the groundwater recharge rate and the thickness of aquifer below sea level are relatively less effective. In the island with steep seashore slopes larger than <span class="hlt">1</span>~2 degrees or so, the storage amount of fresh water in a coastal area increases as sea level <span class="hlt">rises</span>. On the other hand, when sea level drops, the storage amount decreases. This is because the groundwater level also <span class="hlt">rises</span> with the <span class="hlt">rising</span> sea level in steep seashores. For relatively flat seashores, where the slope is smaller than around <span class="hlt">1</span>-2 degrees, the storage amount of coastal fresh water decreases when the sea level <span class="hlt">rises</span> because the area flooded by the <span class="hlt">rising</span> sea water is increased. The volume of aquifer fresh water in this circumstance is greatly reduced in proportion to the flooded area with the sea</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMGC11D1167H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMGC11D1167H"><span>Healthy coral reefs may assure coastal protection in face of climate change related sea level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Harris, D. L.; Rovere, A.; Parravicini, V.; Casella, E.; Canavesio, R.; Collin, A.</p> <p>2016-12-01</p> <p>Coral reefs are diverse ecosystems that support millions of people worldwide providing crucial services, of which, coastal protection is one of the most relevant. The efficiency of coral reefs in protecting coastlines and dissipating waves is directly linked to the cover of living corals and three dimensional reef structural complexity. Climate change and human impacts are leading to severe global reductions in live coral cover, posing serious concerns regarding the capacity of degraded reef systems in protecting tropical coastal regions. Although it is known that the loss of structurally complex reefs may lead to greater erosion of coastlines, this process has rarely been quantified and it is still unknown whether the maintenance of healthy reefs through conservation will be enough to guarantee coastal protection during <span class="hlt">rising</span> sea levels. We show that a significant loss of wave dissipation and a subsequent increase in back-reef wave height (up to 5 times present wave height) could occur even at present sea level if living corals are lost and reef structural complexity is reduced. Yet we also show that healthy reefs, measured by structural complexity and efficiency of <span class="hlt">vertical</span> reef accretion, may maintain their present capacity of wave dissipation even under <span class="hlt">rising</span> sea levels. Our results indicate that the health of coral reefs and not sea level <span class="hlt">rise</span> will be the major determinant of the coastal protection services provided by coral reefs and calls for investments into coral reef conservation to ensure the future protection of tropical coastal communities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70035295','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70035295"><span>Predicting the retreat and migration of tidal forests along the northern Gulf of Mexico under sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Doyle, T.W.; Krauss, K.W.; Conner, W.H.; From, A.S.</p> <p>2010-01-01</p> <p>Tidal freshwater forests in coastal regions of the southeastern United States are undergoing dieback and retreat from increasing tidal inundation and saltwater intrusion attributed to climate variability and sea-level <span class="hlt">rise</span>. In many areas, tidal saltwater forests (mangroves) contrastingly are expanding landward in subtropical coastal reaches succeeding freshwater marsh and forest zones. Hydrological characteristics of these low-relief coastal forests in intertidal settings are dictated by the influence of tidal and freshwater forcing. In this paper, we describe the application of the Sea Level Over Proportional Elevation (SLOPE) model to predict coastal forest retreat and migration from projected sea-level <span class="hlt">rise</span> based on a proxy relationship of saltmarsh/mangrove area and tidal range. The SLOPE model assumes that the sum area of saltmarsh/mangrove habitat along any given coastal reach is determined by the slope of the landform and <span class="hlt">vertical</span> tide forcing. Model results indicated that saltmarsh and mangrove migration from sea-level <span class="hlt">rise</span> will vary by county and watershed but greater in western Gulf States than in the eastern Gulf States where millions of hectares of coastal forest will be displaced over the next century with a near meter <span class="hlt">rise</span> in relative sea level alone. Substantial losses of coastal forests will also occur in the eastern Gulf but mangrove forests in subtropical zones of Florida are expected to replace retreating freshwater forest and affect regional biodiversity. Accelerated global eustacy from climate change will compound the degree of predicted retreat and migration of coastal forests with expected implications for ecosystem management of State and Federal lands in the absence of adaptive coastal management.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6176981-vertical-electromagnetic-profiling-vemp','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6176981-vertical-electromagnetic-profiling-vemp"><span><span class="hlt">Vertical</span> electromagnetic profiling (VEMP)</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Lytle, R.J.</p> <p>1984-08-01</p> <p><span class="hlt">Vertical</span> seismic profiling (VSP) is based upon reception measurements performed in a borehole with a source near the ground surface. This technology has seen a surge in application and development in the last decade. The analogous concept of <span class="hlt">vertical</span> electromagnetic profiling (VEMP) consists of reception measurements performed in a borehole with a source near the ground surface. Although the electromagnetic concept has seen some application, this technology has not been as systematically developed and applied as VSP. <span class="hlt">Vertical</span> electromagnetic profiling provides distinct and complementary data due to sensing different physical parameters than seismic profiling. Certain of the advantages of VEMPmore » are presented. 28 references, 7 figures.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1356943','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1356943"><span>Triphasic 2D Materials by <span class="hlt">Vertically</span> Stacking Laterally Heterostructured 2H-/<span class="hlt">1</span>T'-MoS 2 on Graphene for Enhanced Photoresponse</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Cui, Weili; Xu, Shanshan S.; Yan, Bo</p> <p></p> <p>Recently the applications of two-dimensional (2D) materials have been broadened by engineering their mechanical, electronic, and optical properties through either lateral or <span class="hlt">vertical</span> hybridization. Along with this line, we report the successful design and fabrication of a novel triphasic 2D material by <span class="hlt">vertically</span> stacking lateral 2H-/<span class="hlt">1</span>T'-molybdenum disulfide (MoS 2) heterostructures on graphene with the assistance of supercritical carbon dioxide. This triphasic structure is experimentally shown to significantly enhance the photocurrent densities for hydrogen evolution reactions. First-principles theoretical analyses reveal that the improved photoresponse should be ascribed to the beneficial band alignments of the triphasic heterostructure. More specifically, electrons can efficientlymore » hop to the <span class="hlt">1</span>T'-MoS 2 phase via the highly conductive graphene layer as a result of their strong <span class="hlt">vertical</span> interfacial electronic coupling. Subsequently, the electrons acquired on the <span class="hlt">1</span>T'-MoS 2 phase are exploited to fill the photoholes on the photo-excited 2H-MoS 2 phase through the lateral heterojunction structure, thereby suppressing the recombination process of the photo-induced charge carriers on the 2H-MoS 2 phase. This novel triphasic concept promises to open a new avenue to widen the molecular design of 2D hybrid materials for photonics-based energy conversion applications.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1356943-triphasic-materials-vertically-stacking-laterally-heterostructured-mos2-graphene-enhanced-photoresponse','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1356943-triphasic-materials-vertically-stacking-laterally-heterostructured-mos2-graphene-enhanced-photoresponse"><span>Triphasic 2D Materials by <span class="hlt">Vertically</span> Stacking Laterally Heterostructured 2H-/<span class="hlt">1</span>T'-MoS 2 on Graphene for Enhanced Photoresponse</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Cui, Weili; Xu, Shanshan S.; Yan, Bo; ...</p> <p>2017-05-11</p> <p>Recently the applications of two-dimensional (2D) materials have been broadened by engineering their mechanical, electronic, and optical properties through either lateral or <span class="hlt">vertical</span> hybridization. Along with this line, we report the successful design and fabrication of a novel triphasic 2D material by <span class="hlt">vertically</span> stacking lateral 2H-/<span class="hlt">1</span>T'-molybdenum disulfide (MoS 2) heterostructures on graphene with the assistance of supercritical carbon dioxide. This triphasic structure is experimentally shown to significantly enhance the photocurrent densities for hydrogen evolution reactions. First-principles theoretical analyses reveal that the improved photoresponse should be ascribed to the beneficial band alignments of the triphasic heterostructure. More specifically, electrons can efficientlymore » hop to the <span class="hlt">1</span>T'-MoS 2 phase via the highly conductive graphene layer as a result of their strong <span class="hlt">vertical</span> interfacial electronic coupling. Subsequently, the electrons acquired on the <span class="hlt">1</span>T'-MoS 2 phase are exploited to fill the photoholes on the photo-excited 2H-MoS 2 phase through the lateral heterojunction structure, thereby suppressing the recombination process of the photo-induced charge carriers on the 2H-MoS 2 phase. This novel triphasic concept promises to open a new avenue to widen the molecular design of 2D hybrid materials for photonics-based energy conversion applications.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/circ/1392/pdf/circ1392.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/circ/1392/pdf/circ1392.pdf"><span>Land subsidence and relative sea-level <span class="hlt">rise</span> in the southern Chesapeake Bay region</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Eggleston, Jack; Pope, Jason</p> <p>2013-01-01</p> <p>The southern Chesapeake Bay region is experiencing land subsidence and <span class="hlt">rising</span> water levels due to global sea-level <span class="hlt">rise</span>; land subsidence and <span class="hlt">rising</span> water levels combine to cause relative sea-level <span class="hlt">rise</span>. Land subsidence has been observed since the 1940s in the southern Chesapeake Bay region at rates of <span class="hlt">1.1</span> to 4.8 millimeters per year (mm/yr), and subsidence continues today. This land subsidence helps explain why the region has the highest rates of sea-level <span class="hlt">rise</span> on the Atlantic Coast of the United States. Data indicate that land subsidence has been responsible for more than half the relative sea-level <span class="hlt">rise</span> measured in the region. Land subsidence increases the risk of flooding in low-lying areas, which in turn has important economic, environmental, and human health consequences for the heavily populated and ecologically important southern Chesapeake Bay region. The aquifer system in the region has been compacted by extensive groundwater pumping in the region at rates of <span class="hlt">1</span>.5- to 3.7-mm/yr; this compaction accounts for more than half of observed land subsidence in the region. Glacial isostatic adjustment, or the flexing of the Earth’s crust in response to glacier formation and melting, also likely contributes to land subsidence in the region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5364142','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5364142"><span>Eye-Tracking Reveals that the Strength of the <span class="hlt">Vertical</span>-Horizontal Illusion Increases as the Retinal Image Becomes More Stable with Fixation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Chouinard, Philippe A.; Peel, Hayden J.; Landry, Oriane</p> <p>2017-01-01</p> <p>The closer a line extends toward a surrounding frame, the longer it appears. This is known as a framing effect. Over 70 years ago, Teodor Künnapas demonstrated that the shape of the visual field itself can act as a frame to influence the perceived length of lines in the <span class="hlt">vertical</span>-horizontal illusion. This illusion is typically created by having a <span class="hlt">vertical</span> line <span class="hlt">rise</span> from the center of a horizontal line of the same length creating an inverted T figure. We aimed to determine if the degree to which one fixates on a spatial location where the two lines bisect could influence the strength of the illusion, assuming that the framing effect would be stronger when the retinal image is more stable. We performed two experiments: the visual-field and <span class="hlt">vertical</span>-horizontal illusion experiments. The visual-field experiment demonstrated that the participants could discriminate a target more easily when it was presented along the horizontal vs. <span class="hlt">vertical</span> meridian, confirming a framing influence on visual perception. The <span class="hlt">vertical</span>-horizontal illusion experiment determined the effects of orientation, size and eye gaze on the strength of the illusion. As predicted, the illusion was strongest when the stimulus was presented in either its standard inverted T orientation or when it was rotated 180° compared to other orientations, and in conditions in which the retinal image was more stable, as indexed by eye tracking. Taken together, we conclude that the results provide support for Teodor Künnapas’ explanation of the <span class="hlt">vertical</span>-horizontal illusion. PMID:28392764</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22350880-continuous-wave-single-transverse-mode-emission-from-edge-emitting-lasers-vertically-extended-lasing-area','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22350880-continuous-wave-single-transverse-mode-emission-from-edge-emitting-lasers-vertically-extended-lasing-area"><span><span class="hlt">1</span>.9 W continuous-wave single transverse mode emission from 1060 nm edge-emitting lasers with <span class="hlt">vertically</span> extended lasing area</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Miah, M. J., E-mail: jarez.miah@tu-berlin.de; Posilovic, K.; Kalosha, V. P.</p> <p>2014-10-13</p> <p>High-brightness edge-emitting semiconductor lasers having a <span class="hlt">vertically</span> extended waveguide structure emitting in the 1060 nm range are investigated. Ridge waveguide (RW) lasers with 9 μm stripe width and 2.64 mm cavity length yield highest to date single transverse mode output power for RW lasers in the 1060 nm range. The lasers provide <span class="hlt">1</span>.9 W single transverse mode optical power under continuous-wave (cw) operation with narrow beam divergences of 9° in lateral and 14° (full width at half maximum) in <span class="hlt">vertical</span> direction. The beam quality factor M{sup 2} is less than <span class="hlt">1</span>.9 up to <span class="hlt">1</span>.9 W optical power. A maximum brightness of 72 MWcm{sup −2}sr{supmore » −<span class="hlt">1</span>} is obtained. 100 μm wide and 3 mm long unpassivated broad area lasers provide more than 9 W optical power in cw operation.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004AGUSM.G51B..01V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004AGUSM.G51B..01V"><span>A Gravimetric Geoid Model for <span class="hlt">Vertical</span> Datum in Canada</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Veronneau, M.; Huang, J.</p> <p>2004-05-01</p> <p>The need to realize a new <span class="hlt">vertical</span> datum for Canada dates back to 1976 when a study group at Geodetic Survey Division (GSD) investigated problems related to the existing <span class="hlt">vertical</span> system (CGVD28) and recommended a redefinition of the <span class="hlt">vertical</span> datum. The US National Geodetic Survey and GSD cooperated in the development of a new North American <span class="hlt">Vertical</span> Datum (NAVD88). Although the USA adopted NAVD88 in 1993 as its datum, Canada declined to do so as a result of unexplained discrepancies of about <span class="hlt">1</span>.5 m from east to west coasts (likely due to systematic errors). The high cost of maintaining the <span class="hlt">vertical</span> datum by the traditional spirit leveling technique coupled with budgetary constraints has forced GSD to modify its approach. A new attempt (project) to modernize the <span class="hlt">vertical</span> datum is currently in process in Canada. The advance in space-based technologies (e.g. GPS, satellite radar altimetry, satellite gravimetry) and new developments in geoid modeling offer an alternative to spirit leveling. GSD is planning to implement, after stakeholder consultations, a geoid model as the new <span class="hlt">vertical</span> datum for Canada, which will allow space-based technology users access to an accurate and uniform datum all across the Canadian landmass and surrounding oceans. CGVD28 is only accessible through a limited number of benchmarks, primarily located in southern Canada. The new <span class="hlt">vertical</span> datum would be less sensitive to geodynamic activities (post-glacial rebound and earthquake), local uplift and subsidence, and deterioration of the benchmarks. The adoption of a geoid model as a <span class="hlt">vertical</span> datum does not mean that GSD is neglecting the current benchmarks. New heights will be given to the benchmarks by a new adjustment of the leveling observations, which will be constrained to the geoid model at selected stations of the Active Control System (ACS) and Canadian Base Network (CBN). This adjustment will not correct <span class="hlt">vertical</span> motion at benchmarks, which has occurred since the last leveling observations</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..MARF32009Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..MARF32009Z"><span>Ultrasensitive near-infrared photodetectors based on graphene-MoTe2-graphene <span class="hlt">vertical</span> van der Waals heterostructure</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Kun; Ye, Yu; Dai, Lun; School of Physics, Peking University Team</p> <p></p> <p>Two-dimensional (2D) materials have rapidly established themselves as exceptional building blocks for optoelectronic applications, due to their unique properties and atomically thin nature. Nevertheless, near-infrared (NIR) photodetectors based on layered 2D semiconductors are rarely realized. In this work, we fabricate graphene-MoTe2-graphene <span class="hlt">vertical</span> vdWs heterostructure by a facile and reliable site controllable transfer method, and apply it for photodetection from visible to the NIR wavelength range. Compared to the 2D semiconductor based photodetectors reported thus far, the graphene-MoTe2-graphene photodetector has superior performance, including high photoresponsivity (110 mA W-<span class="hlt">1</span> at 1064 nm and 205 mA W-<span class="hlt">1</span> at 473 nm), high external quantum efficiency (EQE, 12.9% at 1064 nm and 53.8% at 473 nm), rapid response and recovery processes (<span class="hlt">rise</span> time of 24 μs, fall time of 46 μs under 1064 nm illumination), and free from an external source-drain power supply. The all-2D-materials heterostructure has promising applications in future novel high responsivity, high speed and flexible NIR devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25629092','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25629092"><span>Probabilistic reanalysis of twentieth-century sea-level <span class="hlt">rise</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hay, Carling C; Morrow, Eric; Kopp, Robert E; Mitrovica, Jerry X</p> <p>2015-01-22</p> <p>Estimating and accounting for twentieth-century global mean sea level (GMSL) <span class="hlt">rise</span> is critical to characterizing current and future human-induced sea-level change. Several previous analyses of tide gauge records--employing different methods to accommodate the spatial sparsity and temporal incompleteness of the data and to constrain the geometry of long-term sea-level change--have concluded that GMSL rose over the twentieth century at a mean rate of <span class="hlt">1</span>.6 to <span class="hlt">1</span>.9 millimetres per year. Efforts to account for this rate by summing estimates of individual contributions from glacier and ice-sheet mass loss, ocean thermal expansion, and changes in land water storage fall significantly short in the period before 1990. The failure to close the budget of GMSL during this period has led to suggestions that several contributions may have been systematically underestimated. However, the extent to which the limitations of tide gauge analyses have affected estimates of the GMSL rate of change is unclear. Here we revisit estimates of twentieth-century GMSL <span class="hlt">rise</span> using probabilistic techniques and find a rate of GMSL <span class="hlt">rise</span> from 1901 to 1990 of <span class="hlt">1</span>.2 ± 0.2 millimetres per year (90% confidence interval). Based on individual contributions tabulated in the Fifth Assessment Report of the Intergovernmental Panel on Climate Change, this estimate closes the twentieth-century sea-level budget. Our analysis, which combines tide gauge records with physics-based and model-derived geometries of the various contributing signals, also indicates that GMSL rose at a rate of 3.0 ± 0.7 millimetres per year between 1993 and 2010, consistent with prior estimates from tide gauge records.The increase in rate relative to the 1901-90 trend is accordingly larger than previously thought; this revision may affect some projections of future sea-level <span class="hlt">rise</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.U23A..04K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.U23A..04K"><span>The Climate Science Special Report: <span class="hlt">Rising</span> Seas and Changing Oceans</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kopp, R. E.</p> <p>2017-12-01</p> <p>GMSL has risen by about 16-21 cm since 1900. Ocean heat content has increased at all depths since the 1960s, and global mean sea-surface temperature increased 0.7°C/century between 1900 to 2016. Human activity contributed substantially to generating a rate of GMSL <span class="hlt">rise</span> since 1900 faster than during any preceding century in at least 2800 years. A new set of six sea-level <span class="hlt">rise</span> scenarios, spanning a range from 30 cm to 250 cm of 21st century GMSL <span class="hlt">rise</span>, were developed for the CSSR. The lower scenario is based on linearly extrapolating the past two decades' rate of <span class="hlt">rise</span>. The upper scenario is informed by literature estimates of maximum physically plausible values, observations indicating the onset of marine ice sheet instability in parts of West Antarctica, and modeling of ice-cliff and ice-shelf instability mechanisms. The new scenarios include localized projections along US coastlines. There is significant variability around the US, with rates of <span class="hlt">rise</span> likely greater than GMSL <span class="hlt">rise</span> in the US Northeast and the western Gulf of Mexico. Under scenarios involving extreme Antarctic contributions, regional <span class="hlt">rise</span> would be greater than GMSL <span class="hlt">rise</span> along almost all US coastlines. Historical sea-level <span class="hlt">rise</span> has already driven a 5- to 10-fold increase in minor tidal flooding in several US coastal cities since the 1960s. Under the CSSR's Intermediate sea-level <span class="hlt">rise</span> scenario (<span class="hlt">1</span>.0 m of GMSL <span class="hlt">rise</span> in 2100) , a majority of NOAA tide gauge locations will by 2040 experience the historical 5-year coastal flood about 5 times per year. Ocean changes are not limited to <span class="hlt">rising</span> sea levels. Ocean pH is decreasing at a rate that may be unparalleled in the last 66 million years. Along coastlines, ocean acidification can be enhanced by changes in the upwelling (particularly along the US Pacific Coast); by episodic, climate change-enhanced increases in freshwater input (particularly along the US Atlantic Coast); and by the enhancement of biological respiration by nutrient runoff. Climate models project</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19880014727','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19880014727"><span>Three-dimensional models of conventional and <span class="hlt">vertical</span> junction laser-photovoltaic energy converters</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Heinbockel, John H.; Walker, Gilbert H.</p> <p>1988-01-01</p> <p>Three-dimensional models of both conventional planar junction and <span class="hlt">vertical</span> junction photovoltaic energy converters have been constructed. The models are a set of linear partial differential equations and take into account many photoconverter design parameters. The model is applied to Si photoconverters; however, the model may be used with other semiconductors. When used with a Nd laser, the conversion efficiency of the Si <span class="hlt">vertical</span> junction photoconverter is 47 percent, whereas the efficiency for the conventional planar Si photoconverter is only 17 percent. A parametric study of the Si <span class="hlt">vertical</span> junction photoconverter is then done in order to describe the optimum converter for use with the <span class="hlt">1</span>.06-micron Nd laser. The efficiency of this optimized <span class="hlt">vertical</span> junction converter is 44 percent at <span class="hlt">1</span> kW/sq cm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014APS..DFD.G4003R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..DFD.G4003R"><span>Rupture of <span class="hlt">vertical</span> soap films</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rio, Emmanuelle</p> <p>2014-11-01</p> <p>Soap films are ephemeral and fragile objects. They tend to thin under gravity, which gives <span class="hlt">rise</span> to the fascinating variations of colors at their interfaces but leads systematically to rupture. Even a child can create, manipulate and admire soap films and bubbles. Nevertheless, the reason why it suddenly bursts remains a mystery although the soap chosen to stabilize the film as well as the humidity of the air seem very important. One difficulty to study the rupture of <span class="hlt">vertical</span> soap films is to control the initial solution. To avoid this problem we choose to study the rupture during the generation of the film at a controlled velocity. We have built an experiment, in which we measure the maximum length of the film together with its lifetime. The generation of the film is due to the presence of a gradient of surface concentration of surfactants at the liquid/air interface. This leads to a Marangoni force directed toward the top of the film. The film is expected to burst only when its weight is not balanced anymore by this force. We will show that this leads to the surprising result that the thicker films have shorter lifetimes than the thinner ones. It is thus the ability of the interface to sustain a surface concentration gradient of surfactants which controls its stability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ERL....12l4010B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ERL....12l4010B"><span>Global mean sea-level <span class="hlt">rise</span> in a world agreed upon in Paris</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bittermann, Klaus; Rahmstorf, Stefan; Kopp, Robert E.; Kemp, Andrew C.</p> <p>2017-12-01</p> <p>Although the 2015 Paris Agreement seeks to hold global average temperature to ‘well below 2 °C above pre-industrial levels and to pursue efforts to limit the temperature increase to <span class="hlt">1</span>.5 °C above pre-industrial levels’, projections of global mean sea-level (GMSL) <span class="hlt">rise</span> commonly focus on scenarios in which there is a high probability that warming exceeds <span class="hlt">1</span>.5 °C. Using a semi-empirical model, we project GMSL changes between now and 2150 CE under a suite of temperature scenarios that satisfy the Paris Agreement temperature targets. The projected magnitude and rate of GMSL <span class="hlt">rise</span> varies among these low emissions scenarios. Stabilizing temperature at <span class="hlt">1</span>.5 °C instead of 2 °C above preindustrial reduces GMSL in 2150 CE by 17 cm (90% credible interval: 14-21 cm) and reduces peak rates of <span class="hlt">rise</span> by <span class="hlt">1</span>.9 mm yr-<span class="hlt">1</span> (90% credible interval: <span class="hlt">1</span>.4-2.6 mm yr-<span class="hlt">1</span>). Delaying the year of peak temperature has little long-term influence on GMSL, but does reduce the maximum rate of <span class="hlt">rise</span>. Stabilizing at 2 °C in 2080 CE rather than 2030 CE reduces the peak rate by 2.7 mm yr-<span class="hlt">1</span> (90% credible interval: 2.0-4.0 mm yr-<span class="hlt">1</span>).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70033879','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70033879"><span>A Bayesian network to predict coastal vulnerability to sea level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Gutierrez, B.T.; Plant, N.G.; Thieler, E.R.</p> <p>2011-01-01</p> <p>Sea level <span class="hlt">rise</span> during the 21st century will have a wide range of effects on coastal environments, human development, and infrastructure in coastal areas. The broad range of complex factors influencing coastal systems contributes to large uncertainties in predicting long-term sea level <span class="hlt">rise</span> impacts. Here we explore and demonstrate the capabilities of a Bayesian network (BN) to predict long-term shoreline change associated with sea level <span class="hlt">rise</span> and make quantitative assessments of prediction uncertainty. A BN is used to define relationships between driving forces, geologic constraints, and coastal response for the U.S. Atlantic coast that include observations of local rates of relative sea level <span class="hlt">rise</span>, wave height, tide range, geomorphic classification, coastal slope, and shoreline change rate. The BN is used to make probabilistic predictions of shoreline retreat in response to different future sea level <span class="hlt">rise</span> rates. Results demonstrate that the probability of shoreline retreat increases with higher rates of sea level <span class="hlt">rise</span>. Where more specific information is included, the probability of shoreline change increases in a number of cases, indicating more confident predictions. A hindcast evaluation of the BN indicates that the network correctly predicts 71% of the cases. Evaluation of the results using Brier skill and log likelihood ratio scores indicates that the network provides shoreline change predictions that are better than the prior probability. Shoreline change outcomes indicating stability (-<span class="hlt">1</span> <span class="hlt">1</span> m/yr) was not well predicted. We find that BNs can assimilate important factors contributing to coastal change in response to sea level <span class="hlt">rise</span> and can make quantitative, probabilistic predictions that can be applied to coastal management decisions. Copyright ?? 2011 by the American Geophysical Union.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1981AdSpR...1...35P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1981AdSpR...1...35P"><span><span class="hlt">Vertical</span> sounding balloons for stratospheric photochemistry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pommereau, J. P.</p> <p></p> <p>The use of <span class="hlt">vertical</span> sounding balloons for stratospheric photochemistry studies is illustrated by the use of a <span class="hlt">vertical</span> piloted gas balloon for the search of NO2 diurnal variations. It is shown that the use of montgolfieres (hot air balloons) can enhance the <span class="hlt">vertical</span> sounding technique. Particular attention is given to a sun-heated montgolfiere and to the more sophisticated infrared montgolfiere that is able to perform three to four <span class="hlt">vertical</span> excursions per day and to remain aloft for weeks or months.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA44A..06K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA44A..06K"><span>Climatology of Neutral <span class="hlt">vertical</span> winds in the midlatitude thermosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kerr, R.; Kapali, S.; Riccobono, J.; Migliozzi, M. A.; Noto, J.; Brum, C. G. M.; Garcia, R.</p> <p>2017-12-01</p> <p>More than one thousand measurements of neutral <span class="hlt">vertical</span> winds, relative to an assumed average of 0 m/s during a nighttime period, have been made at Arecibo Observatory and the Millstone Hill Optical Facility since 2012, using imaging Fabry-Perot interferometers. These instruments, tuned to the 630 nm OI emission, are carefully calibrated for instrumental frequency drift using frequency stabilized lasers, allowing isolation of Doppler motion in the zenith with <span class="hlt">1</span>-2 m/s accuracy. As one example of the results, relative <span class="hlt">vertical</span> winds at Arecibo during quiet geomagnetic conditions near winter solstice 2016, range ±70 m/s and have a one standard deviation statistical variability of ±34 m/s. This compares with a ±53 m/s deviation from the average meridional wind, and a ±56 m/s deviation from the average zonal wind measured during the same period. <span class="hlt">Vertical</span> neutral wind velocities for all periods range from roughly 30% - 60% of the horizontal velocity domain at Arecibo. At Millstone Hill, the <span class="hlt">vertical</span> velocities relative to horizontal velocities are similar, but slightly smaller. The midnight temperature maximum at Arecibo is usually correlated with a surge in the upward wind, and <span class="hlt">vertical</span> wind excursions of more than 80 m/s are common during magnetic storms at both sites. Until this compilation of <span class="hlt">vertical</span> wind climatology, <span class="hlt">vertical</span> motions of the neutral atmosphere outside of the auroral zone have generally been assumed to be very small compared to horizontal transport. In fact, excursions from small <span class="hlt">vertical</span> velocities in the mid-latitude thermosphere near the F2 ionospheric peak are common, and are not isolated events associated with unsettled geomagnetic conditions or other special dynamic conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007JCrGr.304..127W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007JCrGr.304..127W"><span>Thermoelectric properties of Ge <span class="hlt">1</span>-xSn xTe crystals grown by <span class="hlt">vertical</span> Bridgman method</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, C. C.; Ferng, N. J.; Gau, H. J.</p> <p>2007-06-01</p> <p>Single crystals of Ge <span class="hlt">1</span>-xSn xTe compounds with x=0, 0.8, 0.9 and <span class="hlt">1</span>.0 were grown by <span class="hlt">vertical</span> Bridgman method. The crystalline phase and stochiometry for these crystals were investigated by X-ray diffraction, metallographic microscope as well as electron-probe microanalysis (EPMA). Electrical property of the as-grown samples was characterized using room temperature resistivity and Hall measurements. The thermoelectric behaviors for the Ge <span class="hlt">1</span>-xSn xTe crystals were studied by means of thermal and carrier transport measurements. Temperature dependences of resistivity, Seebeck coefficient and thermal conductivity for the various compositions of Ge <span class="hlt">1</span>-xSn xTe were analyzed. A two-valence band model was proposed to describe the temperature dependence of thermoelectric property of the Ge <span class="hlt">1</span>-xSn xTe crystals. The dimensionless thermoelectric figure of merit ZT for the alloys was evaluated and discussed.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_16 --> <div id="page_17" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="321"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018EaFut...6..583B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018EaFut...6..583B"><span>Quantifying Land and People Exposed to Sea-Level <span class="hlt">Rise</span> with No Mitigation and <span class="hlt">1</span>.5°C and 2.0°C <span class="hlt">Rise</span> in Global Temperatures to Year 2300</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brown, S.; Nicholls, R. J.; Goodwin, P.; Haigh, I. D.; Lincke, D.; Vafeidis, A. T.; Hinkel, J.</p> <p>2018-03-01</p> <p>We use multiple synthetic mitigation sea-level scenarios, together with a non-mitigation sea-level scenario from the Warming Acidification and Sea-level Projector model. We find sea-level <span class="hlt">rise</span> (SLR) continues to accelerate post-2100 for all but the most aggressive mitigation scenarios indicative of <span class="hlt">1</span>.5°C and 2.0°C. Using the Dynamic Interactive Vulnerability Assessment modeling framework, we project land and population exposed in the <span class="hlt">1</span> in 100 year coastal flood plain under SLR and population change. In 2000, the flood plain is estimated at 540 × 103 km2. By 2100, under the mitigation scenarios, it ranges between 610 × 103 and 640 × 103 km2 (580 × 103 and 700 × 103 km2 for the 5th and 95th percentiles). Thus differences between the mitigation scenarios are small in 2100. However, in 2300, flood plains are projected to increase to between 700 × 103 and 960 × 103 km2 in 2300 (610 × 103 and 1290 × 103 km2) for the mitigation scenarios, but 1630 × 103 km2 (1190 × 103 and 2220 × 103 km2) for the non-mitigation scenario. The proportion of global population exposed to SLR in 2300 is projected to be between <span class="hlt">1</span>.5% and 5.4% (<span class="hlt">1</span>.2%-7.6%) (assuming no population growth after 2100) for the aggressive mitigation and the non-mitigation scenario, respectively. Hence over centennial timescales there are significant benefits to climate change mitigation and temperature stabilization. However, sea-levels will continue to <span class="hlt">rise</span> albeit at lower rates. Thus potential impacts will keep increasing necessitating adaptation to existing coastal infrastructure and the careful planning of new coastal developments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20180002526','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20180002526"><span>Metal Oxide <span class="hlt">Vertical</span> Graphene Hybrid Supercapacitors</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Meyyappan, Meyya (Inventor)</p> <p>2018-01-01</p> <p>A metal oxide <span class="hlt">vertical</span> graphene hybrid supercapacitor is provided. The supercapacitor includes a pair of collectors facing each other, and <span class="hlt">vertical</span> graphene electrode material grown directly on each of the pair of collectors without catalyst or binders. A separator may separate the <span class="hlt">vertical</span> graphene electrode materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..1615137S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..1615137S"><span>Coastal Vulnerability Due to Sea-level <span class="hlt">Rise</span> Hazard in the Bangladesh Delta</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shum, Ck; Ballu, Valérie; Calmant, Stéphane; Duan, Jianbin; Guo, Junyi; Hossain, Fasial; Jenkins, Craig; Haque Khan, Zahirul; Kim, Jinwoo; Kuhn, Michael; Kusche, Jürgen; Papa, Fabrice; Tseng, Kuohsin; Wan, Junkun</p> <p>2014-05-01</p> <p>Approximately half of the world's population or 3.2 billion people lives within 200 km of coastlines and many of them in the world's deltaic plains. Sea-level <span class="hlt">rise</span>, widely recognized as one of consequences resulting from anthropogenic climate change, has induced substantial coastal vulnerability globally and in particular, in the deltaic regions, such as coastal Bangladesh, and Yangtze Delta. Bangladesh, a low-lying, one of the most densely populated countries in the world located at the Bay of Bengal, is prone to transboundary monsoonal flooding, potentially aggravated by more frequent and intensified cyclones resulting from anthropogenic climate change. Sea-level <span class="hlt">rise</span>, along with tectonic, sediment load and groundwater extraction induced land uplift/subsidence, have exacerbated Bangladesh's coastal vulnerability. Here we describe the physical science component of the integrated approach based on both physical and social sciences to address the adaption and potential mitigation of coastal Bangladesh vulnerability. The objective is to quantify the estimates of spatial varying sea-level trend separating the <span class="hlt">vertical</span> motion of the coastal regions using geodetic and remote-sensing measurements (tide gauges, 1950-current; satellite altimetry, 1992-present, GRACE, 2003-present, Landsat/MODIS), reconstructed sea-level trends (1950-current), and GPS and InSAR observed land subsidence. Our goal is to conduct physically based robust projection of relative sea-level change at the end of the 21st century for the Bangladesh Delta to enable quantitative measures of social science based adaption and possible mitigation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5121450','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5121450"><span>Assessment of Temperature <span class="hlt">Rise</span> and Time of Alveolar Ridge Splitting by Means of Er:YAG Laser, Piezosurgery, and Surgical Saw: An Ex Vivo Study</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Flieger, Rafał; Dominiak, Marzena</p> <p>2016-01-01</p> <p>The most common adverse effect after bone cutting is a thermal damage. The aim of our study was to evaluate the bone temperature <span class="hlt">rise</span> during an alveolar ridge splitting, rating the time needed to perform this procedure and the time to raise the temperature of a bone by 10°C, as well as to evaluate the bone carbonization occurrence. The research included 60 mandibles (n = 60) of adult pigs, divided into 4 groups (n = 15). Two <span class="hlt">vertical</span> and one horizontal cut have been done in an alveolar ridge using Er:YAG laser with set power of 200 mJ (G<span class="hlt">1</span>), 400 mJ (G2), piezosurgery unit (G3), and a saw (G4). The temperature was measured by K-type thermocouple. The highest temperature gradient was noted for piezosurgery on the buccal and lingual side of mandible. The temperature <span class="hlt">rises</span> on the bone surface along with the increase of laser power. The lower time needed to perform ridge splitting was measured for a saw, piezosurgery, and Er:YAG laser with power of 400 mJ and 200 mJ, respectively. The temperature <span class="hlt">rise</span> measured on the bone over 10°C and bone carbonization occurrence was not reported in all study groups. Piezosurgery, Er:YAG laser (200 mJ and 400 mJ), and surgical saw are useful and safe tools in ridge splitting surgery. PMID:27957502</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27957502','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27957502"><span>Assessment of Temperature <span class="hlt">Rise</span> and Time of Alveolar Ridge Splitting by Means of Er:YAG Laser, Piezosurgery, and Surgical Saw: An Ex Vivo Study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Matys, Jacek; Flieger, Rafał; Dominiak, Marzena</p> <p>2016-01-01</p> <p>The most common adverse effect after bone cutting is a thermal damage. The aim of our study was to evaluate the bone temperature <span class="hlt">rise</span> during an alveolar ridge splitting, rating the time needed to perform this procedure and the time to raise the temperature of a bone by 10°C, as well as to evaluate the bone carbonization occurrence. The research included 60 mandibles ( n = 60) of adult pigs, divided into 4 groups ( n = 15). Two <span class="hlt">vertical</span> and one horizontal cut have been done in an alveolar ridge using Er:YAG laser with set power of 200 mJ (G<span class="hlt">1</span>), 400 mJ (G2), piezosurgery unit (G3), and a saw (G4). The temperature was measured by K-type thermocouple. The highest temperature gradient was noted for piezosurgery on the buccal and lingual side of mandible. The temperature <span class="hlt">rises</span> on the bone surface along with the increase of laser power. The lower time needed to perform ridge splitting was measured for a saw, piezosurgery, and Er:YAG laser with power of 400 mJ and 200 mJ, respectively. The temperature <span class="hlt">rise</span> measured on the bone over 10°C and bone carbonization occurrence was not reported in all study groups. Piezosurgery, Er:YAG laser (200 mJ and 400 mJ), and surgical saw are useful and safe tools in ridge splitting surgery.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19970018786','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19970018786"><span>Documentation of Atmospheric Conditions During Observed <span class="hlt">Rising</span> Aircraft Wakes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Zak, J. Allen; Rodgers, William G., Jr.</p> <p>1997-01-01</p> <p>Flight tests were conducted in the fall of 1995 off the coast of Wallops Island, Virginia in order to determine characteristics of wake vortices at flight altitudes. A NASA Wallops Flight Facility C130 aircraft equipped with smoke generators produced visible wakes at altitudes ranging from 775 to 2225 m in a variety of atmospheric conditions, orientations (head wind, cross wind), and airspeeds. Meteorological and aircraft parameters were collected continuously from a Langley Research Center OV-10A aircraft as it flew alongside and through the wake vortices at varying distances behind the C130. Meteorological data were also obtained from special balloon observations made at Wallops. Differential GPS capabilities were on each aircraft from which accurate altitude profiles were obtained. Vortices were observed to <span class="hlt">rise</span> at distances beyond a mile behind the C130. The maximum altitude was 150 m above the C130 in a near neutral atmosphere with significant turbulence. This occurred from large <span class="hlt">vertical</span> oscillations in the wakes. There were several cases when vortices did not descend after a very short initial period and remained near generation altitude in a variety of moderately stable atmospheres and wind shears.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ThCFD..30..387B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ThCFD..30..387B"><span>Computational analysis of <span class="hlt">vertical</span> axis wind turbine arrays</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bremseth, J.; Duraisamy, K.</p> <p>2016-10-01</p> <p>Canonical problems involving single, pairs, and arrays of <span class="hlt">vertical</span> axis wind turbines (VAWTs) are investigated numerically with the objective of understanding the underlying flow structures and their implications on energy production. Experimental studies by Dabiri (J Renew Sustain Energy 3, 2011) suggest that VAWTs demand less stringent spacing requirements than their horizontal axis counterparts and additional benefits may be obtained by optimizing the placement and rotational direction of VAWTs. The flowfield of pairs of co-/counter-rotating VAWTs shows some similarities with pairs of cylinders in terms of wake structure and vortex shedding. When multiple VAWTs are placed in a column, the extent of the wake is seen to spread further downstream, irrespective of the direction of rotation of individual turbines. However, the aerodynamic interference between turbines gives <span class="hlt">rise</span> to regions of excess momentum between the turbines which lead to significant power augmentations. Studies of VAWTs arranged in multiple columns show that the downstream columns can actually be more efficient than the leading column, a proposition that could lead to radical improvements in wind farm productivity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26298490','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26298490"><span>A reduction of the saddle <span class="hlt">vertical</span> force triggers the sit-stand transition in cycling.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Costes, Antony; Turpin, Nicolas A; Villeger, David; Moretto, Pierre; Watier, Bruno</p> <p>2015-09-18</p> <p>The purpose of the study was to establish the link between the saddle <span class="hlt">vertical</span> force and its determinants in order to establish the strategies that could trigger the sit-stand transition. We hypothesized that the minimum saddle <span class="hlt">vertical</span> force would be a critical parameter influencing the sit-stand transition during cycling. Twenty-five non-cyclists were asked to pedal at six different power outputs from 20% (<span class="hlt">1</span>.6 ± 0.3 W kg(-<span class="hlt">1</span>)) to 120% (9.6 ± <span class="hlt">1</span>.6 W kg(-<span class="hlt">1</span>)) of their spontaneous sit-stand transition power obtained at 90 rpm. Five 6-component sensors (saddle tube, pedals and handlebars) and a full-body kinematic reconstruction were used to provide the saddle <span class="hlt">vertical</span> force and other force components (trunk inertial force, hips and shoulders reaction forces, and trunk weight) linked to the saddle <span class="hlt">vertical</span> force. Minimum saddle <span class="hlt">vertical</span> force linearly decreased with power output by 87% from a static position on the bicycle (5.30 ± 0.50 N kg(-<span class="hlt">1</span>)) to power output=120% of the sit-stand transition power (0.68 ± 0.49 N kg(-<span class="hlt">1</span>)). This decrease was mainly explained by the increase in instantaneous pedal forces from 2.84 ± 0.58 N kg(-<span class="hlt">1</span>) to 6.57 ± <span class="hlt">1</span>.02 N kg(-<span class="hlt">1</span>) from 20% to 120% of the power output corresponding to the sit-stand transition, causing an increase in hip <span class="hlt">vertical</span> forces from -0.17 N kg(-<span class="hlt">1</span>) to 3.29 N kg(-<span class="hlt">1</span>). The emergence of strategies aiming at counteracting the elevation of the trunk (handlebars and pedals pulling) coincided with the spontaneous sit-stand transition power. The present data suggest that the large decrease in minimum saddle <span class="hlt">vertical</span> force observed at high pedal reaction forces might trigger the sit-stand transition in cycling. Copyright © 2015 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5594671','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5594671"><span>Exposure to <span class="hlt">rising</span> inequality shapes Americans’ opportunity beliefs and policy support</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>McCall, Leslie; Burk, Derek; Laperrière, Marie; Richeson, Jennifer A.</p> <p>2017-01-01</p> <p>Economic inequality has been on the <span class="hlt">rise</span> in the United States since the 1980s and by some measures stands at levels not seen since before the Great Depression. Although the strikingly high and <span class="hlt">rising</span> level of economic inequality in the nation has alarmed scholars, pundits, and elected officials alike, research across the social sciences repeatedly concludes that Americans are largely unconcerned about it. Considerable research has documented, for instance, the important role of psychological processes, such as system justification and American Dream ideology, in engendering Americans’ relative insensitivity to economic inequality. The present work offers, and reports experimental tests of, a different perspective—the opportunity model of beliefs about economic inequality. Specifically, two convenience samples (study <span class="hlt">1</span>, n = 480; and study 2, n = <span class="hlt">1</span>,305) and one representative sample (study 3, n = <span class="hlt">1</span>,501) of American adults were exposed to information about <span class="hlt">rising</span> economic inequality in the United States (or control information) and then asked about their beliefs regarding the roles of structural (e.g., being born wealthy) and individual (e.g., hard work) factors in getting ahead in society (i.e., opportunity beliefs). They then responded to policy questions regarding the roles of business and government actors in reducing economic inequality. Rather than revealing insensitivity to <span class="hlt">rising</span> inequality, the results suggest that <span class="hlt">rising</span> economic inequality in contemporary society can spark skepticism about the existence of economic opportunity in society that, in turn, may motivate support for policies designed to redress economic inequality. PMID:28831007</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28831007','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28831007"><span>Exposure to <span class="hlt">rising</span> inequality shapes Americans' opportunity beliefs and policy support.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>McCall, Leslie; Burk, Derek; Laperrière, Marie; Richeson, Jennifer A</p> <p>2017-09-05</p> <p>Economic inequality has been on the <span class="hlt">rise</span> in the United States since the 1980s and by some measures stands at levels not seen since before the Great Depression. Although the strikingly high and <span class="hlt">rising</span> level of economic inequality in the nation has alarmed scholars, pundits, and elected officials alike, research across the social sciences repeatedly concludes that Americans are largely unconcerned about it. Considerable research has documented, for instance, the important role of psychological processes, such as system justification and American Dream ideology, in engendering Americans' relative insensitivity to economic inequality. The present work offers, and reports experimental tests of, a different perspective-the opportunity model of beliefs about economic inequality. Specifically, two convenience samples (study <span class="hlt">1</span>, n = 480; and study 2, n = <span class="hlt">1</span>,305) and one representative sample (study 3, n = <span class="hlt">1</span>,501) of American adults were exposed to information about <span class="hlt">rising</span> economic inequality in the United States (or control information) and then asked about their beliefs regarding the roles of structural (e.g., being born wealthy) and individual (e.g., hard work) factors in getting ahead in society (i.e., opportunity beliefs). They then responded to policy questions regarding the roles of business and government actors in reducing economic inequality. Rather than revealing insensitivity to <span class="hlt">rising</span> inequality, the results suggest that <span class="hlt">rising</span> economic inequality in contemporary society can spark skepticism about the existence of economic opportunity in society that, in turn, may motivate support for policies designed to redress economic inequality.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFMED41A0715D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFMED41A0715D"><span>Sea-Level <span class="hlt">Rise</span> and Flood Potential along the California Coast</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Delepine, Q.; Leung, C.</p> <p>2013-12-01</p> <p>Sea-level <span class="hlt">rise</span> is becoming an ever-increasing problem in California. Sea-level is expected to <span class="hlt">rise</span> significantly in the next 100 years, which will raise flood elevations in coastal communities. This will be an issue for private homeowners, businesses, and the state. One study suggests that Venice Beach could lose a total of at least $440 million in tourism spending and tax dollars from flooding and beach erosion if sea level <span class="hlt">rises</span> <span class="hlt">1</span>.4 m by 2100. In addition, several airports, such as San Francisco International Airport, are located in coastal regions that have flooded in the past and will likely be flooded again in the next 30 years, but sea-level <span class="hlt">rise</span> is expected to worsen the effects of flooding in the coming decades It is vital for coastal communities to understand the risks associated with sea-level <span class="hlt">rise</span> so that they can plan to adapt to it. By obtaining accurate LiDAR elevation data from the NOAA Digital Coast Website (http://csc.noaa.gov/dataviewer/?keyword=lidar#), we can create flood maps to simulate sea level <span class="hlt">rise</span> and flooding. The data are uploaded to ArcGIS and contour lines are added for different elevations that represent future coastlines during 100-year flooding. The following variables are used to create the maps: <span class="hlt">1</span>. High-resolution land surface elevation data - obtained from NOAA 2. Local mean high water level - from USGS 3. Local 100-year flood water level - from the Pacific Institute 4. Sea-level <span class="hlt">rise</span> projections for different future dates (2030, 2050, and 2100) - from the National Research Council The values from the last three categories are added to represent sea-level <span class="hlt">rise</span> plus 100-year flooding. These values are used to make the contour lines that represent the projected flood elevations, which are then exported as KML files, which can be opened in Google Earth. Once these KML files are made available to the public, coastal communities will gain an improved understanding of how flooding and sea-level <span class="hlt">rise</span> might affect them in the future</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DPPP11041B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DPPP11041B"><span>Analysis of <span class="hlt">vertical</span> stability limits and <span class="hlt">vertical</span> displacement event behavior on NSTX-U</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Boyer, Mark; Battaglia, Devon; Gerhardt, Stefan; Menard, Jonathan; Mueller, Dennis; Myers, Clayton; Sabbagh, Steven; Smith, David</p> <p>2017-10-01</p> <p>The National Spherical Torus Experiment Upgrade (NSTX-U) completed its first run campaign in 2016, including commissioning a larger center-stack and three new tangentially aimed neutral beam sources. NSTX-U operates at increased aspect ratio due to the larger center-stack, making <span class="hlt">vertical</span> stabilization more challenging. Since ST performance is improved at high elongation, improvements to the <span class="hlt">vertical</span> control system were made, including use of multiple up-down-symmetric flux loop pairs for real-time estimation, and filtering to remove noise. Similar operating limits to those on NSTX (in terms of elongation and internal inductance) were achieved, now at higher aspect ratio. To better understand the observed limits and project to future operating points, a database of <span class="hlt">vertical</span> displacement events and <span class="hlt">vertical</span> oscillations observed during the plasma current ramp-up on NSTX/NSTX-U has been generated. Shots were clustered based on the characteristics of the VDEs/oscillations, and the plasma parameter regimes associated with the classes of behavior were studied. Results provide guidance for scenario development during ramp-up to avoid large oscillations at the time of diverting, and provide the means to assess stability of target scenarios for the next campaign. Results will also guide plans for improvements to the <span class="hlt">vertical</span> control system. Work supported by U.S. D.O.E. Contract No. DE-AC02-09CH11466.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017FrEaS...5...78B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017FrEaS...5...78B"><span>Retrograde accretion of a Caribbean fringing reef controlled by hurricanes and sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Blanchon, Paul; Richards, Simon; Bernal, Juan Pablo; Cerdeira-Estrada, Sergio; Ibarra, M. Socrates; Corona-Martínez, Liliana; Martell-Dubois, Raúl</p> <p>2017-10-01</p> <p>Predicting the impact of sea-level (SL) <span class="hlt">rise</span> on coral reefs requires reliable models of reef accretion. Most assume that accretion results from <span class="hlt">vertical</span> growth of coralgal framework, but recent studies show that reefs exposed to hurricanes consist of layers of coral gravel rather than in-place corals. New models are therefore needed to account for hurricane impact on reef accretion over geological timescales. To investigate this geological impact, we report the configuration and development of a 4-km-long fringing reef at Punta Maroma along the northeast Yucatan Peninsula. Satellite-derived bathymetry shows the crest is set-back a uniform distance of 315 ±15 m from a mid-shelf slope break, and the reef-front decreases 50% in width and depth along its length. A 12-core drill transect constrained by multiple 230Th ages shows the reef is composed of an 2-m thick layer of coral clasts that has retrograded 100 m over its back-reef during the last 5.5 ka. These findings are consistent with a hurricane-control model of reef development where large waves trip and break over the mid-shelf slope break, triggering rapid energy dissipation and thus limiting how far upslope individual waves can fragment corals and redistribute clasts. As SL <span class="hlt">rises</span> and water depth increases, energy dissipation during wave-breaking is reduced, extending the clast-transport limit, thus leading to reef retrogradation. This hurricane model may be applicable to a large sub-set of fringing reefs in the tropical Western-Atlantic necessitating a reappraisal of their accretion rates and response to future SL <span class="hlt">rise</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27281277','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27281277"><span>The Medial Stitch in Transosseous-Equivalent Rotator Cuff Repair: <span class="hlt">Vertical</span> or Horizontal Mattress?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Montanez, Anthony; Makarewich, Christopher A; Burks, Robert T; Henninger, Heath B</p> <p>2016-09-01</p> <p>Despite advances in surgical technique, rotator cuff repair retears continue to occur at rates of 10%, 22%, and 57% for small, medium, and large tears, respectively. A common mode of failure in transosseous-equivalent rotator cuff repairs is tissue pullout of the medial mattress stitch. While the medial mattress stitch has been studied extensively, no studies have evaluated a <span class="hlt">vertical</span> mattress pattern placed near the musculotendinous junction in comparison with a horizontal mattress pattern. <span class="hlt">Vertical</span> mattress stitches will have higher load to failure and lower gapping compared with horizontal mattress stitches in a transosseous-equivalent rotator cuff repair. Controlled laboratory study. Double-row transosseous-equivalent rotator cuff repairs were performed in 9 pairs of human male cadaveric shoulders (mean age ± SD, 58 ± 10 years). One shoulder in each pair received a medial-row suture pattern using a <span class="hlt">vertical</span> mattress stitch, and the contralateral shoulder received a horizontal mattress. Specimens were mounted in a materials testing machine and tested in uniaxial tensile deformation for cyclic loading (500 cycles at <span class="hlt">1</span> Hz to <span class="hlt">1</span>.0 MPa of effective stress), followed by failure testing carried out at a rate of <span class="hlt">1</span> mm/s. Construct gapping and applied loads were monitored continuously throughout the testing. <span class="hlt">Vertical</span> mattress sutures were placed in 5 right and 4 left shoulders. Peak cyclic gapping did not differ between <span class="hlt">vertical</span> (mean ± SD, 2.8 ± <span class="hlt">1.1</span> mm) and horizontal mattress specimens (3.0 ± <span class="hlt">1</span>.2 mm) (P = .684). <span class="hlt">Vertical</span> mattress sutures failed at higher loads compared with horizontal mattress sutures (568.9 ± 140.3 vs 451.<span class="hlt">1</span> ± 174.3 N; P = .025); however, there was no significant difference in failure displacement (8.0 ± <span class="hlt">1</span>.6 vs 6.0 ± 2.<span class="hlt">1</span> mm; P = .092). Failure stiffness did not differ between the suture patterns (P = .204). In transosseous-equivalent rotator cuff repairs near the musculotendinous junction, a <span class="hlt">vertical</span> mattress suture used as the medial stitch</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPRS..139...14U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPRS..139...14U"><span>Extraction and height estimation of artificial <span class="hlt">vertical</span> structures based on the wrapped interferometric phase difference within their layovers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Uemoto, Jyunpei; Nadai, Akitsugu; Kojima, Shoichiro; Kobayashi, Tatsuharu; Umehara, Toshihiko; Matsuoka, Takeshi; Uratsuka, Seiho; Satake, Makoto</p> <p>2018-05-01</p> <p>The geometric modulation of synthetic aperture radar (SAR) imagery such as radar shadow, foreshortening, and layover often complicates image interpretation while it contains useful information about targets. Recently, some methods for automatic building detection utilizing a peculiar pattern of phase differences (PDs) within building layovers on SAR interferograms have been proposed. One of the merits of these methods is the capability to detect buildings even taller than the height of ambiguity without incorporating any external data. In this paper, we propose a new method that has achieved the following improvements while maintaining the merit mentioned above. The first improvement is freedom from the dependence of target heights; without changing any parameters and thresholds, the proposed method can detect low-<span class="hlt">rise</span> apartments to skyscrapers. The second one is the prevention of the false grouping of <span class="hlt">vertical</span> structure constituents by considering relationships between their PDs. In addition, the method can measure the height of <span class="hlt">vertical</span> structures without assuming their shape to be simple ones such as a parallelogram. These improvements have been verified by applying the method to real datasets acquired from an airborne X-band SAR. The quantitative assessment for apartment complexes has demonstrated the high performance of the method; the correctness and completeness are 94% and 83%, respectively. The mean error in the measured height is -0.2 m, while the standard deviation is <span class="hlt">1</span>.8 m. The verification using real datasets has revealed at the same time that the performance of the method can be degraded due to the crowdedness in dense urban areas including skyscrapers and owing to the poor discriminability between artificial <span class="hlt">vertical</span> structures and trees. Overcoming these limitations is necessary in future studies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMMR13C..01P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMMR13C..01P"><span>Complex Contact Angles Calculated from Capillary <span class="hlt">Rise</span> Measurements on Rock Fracture Faces</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Perfect, E.; Gates, C. H.; Brabazon, J. W.; Santodonato, L. J.; Dhiman, I.; Bilheux, H.; Bilheux, J. C.; Lokitz, B. S.</p> <p>2017-12-01</p> <p>Contact angles for fluids in unconventional reservoir rocks are needed for modeling hydraulic fracturing leakoff and subsequent oil and gas extraction. Contact angle measurements for wetting fluids on rocks are normally performed using polished flat surfaces. However, such prepared surfaces are not representative of natural rock fracture faces, which have been shown to be rough over multiple scales. We applied a variant of the Wilhelmy plate method for determining contact angle from the height of capillary <span class="hlt">rise</span> on a <span class="hlt">vertical</span> surface to the wetting of rock fracture faces by water in the presence of air. Cylindrical core samples (5.05 cm long x 2.54 cm diameter) of Mancos shale and 6 other rock types were investigated. Mode I fractures were created within the cores using the Brazilian method. Each fractured core was then separated into halves exposing the fracture faces. One fracture face from each rock type was oriented parallel to a collimated neutron beam in the CG-<span class="hlt">1</span>D imaging instrument at ORNL's High Flux Isotope Reactor. Neutron radiography was performed using the multi-channel plate detector with a spatial resolution of 50 μm. Images were acquired every 60 s after a water reservoir contacted the base of the fracture face. The images were normalized to the initial dry condition so that the upward movement of water on the fracture face was clearly visible. The height of wetting at equilibrium was measured on the normalized images using ImageJ. Contact angles were also measured on polished flat surfaces using the conventional sessile drop method. Equilibrium capillary <span class="hlt">rise</span> on the exposed fracture faces was up to 8.5 times greater than that predicted for polished flat surfaces from the sessile drop measurements. These results indicate that rock fracture faces are hyperhydrophilic (i.e., the height of capillary <span class="hlt">rise</span> is greater than that predicted for a contact angle of zero degrees). The use of complex numbers permitted calculation of imaginary contact angles for</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26808841','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26808841"><span>Acute Improvement of <span class="hlt">Vertical</span> Jump Performance After Isometric Squats Depends on Knee Angle and <span class="hlt">Vertical</span> Jumping Ability.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tsoukos, Athanasios; Bogdanis, Gregory C; Terzis, Gerasimos; Veligekas, Panagiotis</p> <p>2016-08-01</p> <p>Tsoukos, A, Bogdanis, GC, Terzis, G, and Veligekas, P. Acute improvement of <span class="hlt">vertical</span> jump performance after isometric squats depends on knee angle and <span class="hlt">vertical</span> jumping ability. J Strength Cond Res 30(8): 2250-2257, 2016-This study examined the acute effects of maximum isometric squats at 2 different knee angles (90 or 140°) on countermovement jump (CMJ) performance in power athletes. Fourteen national-level male track and field power athletes completed 3 main trials (2 experimental and <span class="hlt">1</span> control) in a randomized and counterbalanced order <span class="hlt">1</span> week apart. Countermovement jump performance was evaluated using a force-plate before and 15 seconds, 3, 6, 9, and 12 minutes after 3 sets of 3 seconds maximum isometric contractions with <span class="hlt">1</span>-minute rest in between, from a squat position with knee angle set at 90 or 140°. Countermovement jump performance was improved compared with baseline only in the 140° condition by 3.8 ± <span class="hlt">1</span>.2% on the 12th minute of recovery (p = 0.027), whereas there was no change in CMJ height in the 90° condition. In the control condition, there was a decrease in CMJ performance over time, reaching -3.6 ± <span class="hlt">1</span>.2% (p = 0.049) after 12 minutes of recovery. To determine the possible effects of baseline jump performance on subsequent CMJ performance, subjects were divided into 2 groups ("high jumpers" and "low jumpers"). The baseline CMJ values of "high jumpers" and "low jumpers" differed significantly (CMJ: 45.<span class="hlt">1</span> ± 2.2 vs. 37.<span class="hlt">1</span> ± 3.9 cm, respectively, p = 0.001). Countermovement jump was increased only in the "high jumpers" group by 5.4 ± <span class="hlt">1</span>.4% (p = 0.001) and 7.4 ± <span class="hlt">1</span>.2% (p = 0.001) at the knee angles of 90 and 140°, respectively. This improvement was larger at the 140° angle (p = 0.049). Knee angle during isometric squats and <span class="hlt">vertical</span> jumping ability are important determinants of the acute CMJ performance increase observed after a conditioning activity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70169351','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70169351"><span><span class="hlt">Vertical</span> deformation associated with normal fault systems evolved over coseismic, postseismic, and multiseismic periods</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Thompson, George A.; Parsons, Thomas E.</p> <p>2016-01-01</p> <p><span class="hlt">Vertical</span> deformation of extensional provinces varies significantly and in seemingly contradictory ways. Sparse but robust geodetic, seismic, and geologic observations in the Basin and Range province of the western United States indicate that immediately after an earthquake, <span class="hlt">vertical</span> change primarily occurs as subsidence of the normal fault hanging wall. A few decades later, a ±100 km wide zone is symmetrically uplifted. The preserved topography of long-term rifting shows bent and tilted footwall flanks <span class="hlt">rising</span> high above deep basins. We develop finite element models subjected to extensional and gravitational forces to study time-varying deformation associated with normal faulting. We replicate observations with a model that has a weak upper mantle overlain by a stronger lower crust and a breakable elastic upper crust. A 60° dipping normal fault cuts through the upper crust and extends through the lower crust to simulate an underlying shear zone. Stretching the model under gravity demonstrates that asymmetric slip via collapse of the hanging wall is a natural consequence of coseismic deformation. Focused flow in the upper mantle imposed by deformation of the lower crust localizes uplift under the footwall; the breakable upper crust is a necessary model feature to replicate footwall bending over the observed width ( < 10 km), which is predicted to take place within <span class="hlt">1</span>-2 decades after each large earthquake. Thus the best-preserved topographic signature of rifting is expected to occur early in the postseismic period. The relatively stronger lower crust in our models is necessary to replicate broader postseismic uplift that is observed geodetically in subsequent decades.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29463787','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29463787"><span>Committed sea-level <span class="hlt">rise</span> under the Paris Agreement and the legacy of delayed mitigation action.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mengel, Matthias; Nauels, Alexander; Rogelj, Joeri; Schleussner, Carl-Friedrich</p> <p>2018-02-20</p> <p>Sea-level <span class="hlt">rise</span> is a major consequence of climate change that will continue long after emissions of greenhouse gases have stopped. The 2015 Paris Agreement aims at reducing climate-related risks by reducing greenhouse gas emissions to net zero and limiting global-mean temperature increase. Here we quantify the effect of these constraints on global sea-level <span class="hlt">rise</span> until 2300, including Antarctic ice-sheet instabilities. We estimate median sea-level <span class="hlt">rise</span> between 0.7 and <span class="hlt">1</span>.2 m, if net-zero greenhouse gas emissions are sustained until 2300, varying with the pathway of emissions during this century. Temperature stabilization below 2 °C is insufficient to hold median sea-level <span class="hlt">rise</span> until 2300 below <span class="hlt">1</span>.5 m. We find that each 5-year delay in near-term peaking of CO 2 emissions increases median year 2300 sea-level <span class="hlt">rise</span> estimates by ca. 0.2 m, and extreme sea-level <span class="hlt">rise</span> estimates at the 95th percentile by up to <span class="hlt">1</span> m. Our results underline the importance of near-term mitigation action for limiting long-term sea-level <span class="hlt">rise</span> risks.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014GMD.....7.2717C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014GMD.....7.2717C"><span>Verification of a non-hydrostatic dynamical core using the horizontal spectral element method and <span class="hlt">vertical</span> finite difference method: 2-D aspects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Choi, S.-J.; Giraldo, F. X.; Kim, J.; Shin, S.</p> <p>2014-11-01</p> <p>The non-hydrostatic (NH) compressible Euler equations for dry atmosphere were solved in a simplified two-dimensional (2-D) slice framework employing a spectral element method (SEM) for the horizontal discretization and a finite difference method (FDM) for the <span class="hlt">vertical</span> discretization. By using horizontal SEM, which decomposes the physical domain into smaller pieces with a small communication stencil, a high level of scalability can be achieved. By using <span class="hlt">vertical</span> FDM, an easy method for coupling the dynamics and existing physics packages can be provided. The SEM uses high-order nodal basis functions associated with Lagrange polynomials based on Gauss-Lobatto-Legendre (GLL) quadrature points. The FDM employs a third-order upwind-biased scheme for the <span class="hlt">vertical</span> flux terms and a centered finite difference scheme for the <span class="hlt">vertical</span> derivative and integral terms. For temporal integration, a time-split, third-order Runge-Kutta (RK3) integration technique was applied. The Euler equations that were used here are in flux form based on the hydrostatic pressure <span class="hlt">vertical</span> coordinate. The equations are the same as those used in the Weather Research and Forecasting (WRF) model, but a hybrid sigma-pressure <span class="hlt">vertical</span> coordinate was implemented in this model. We validated the model by conducting the widely used standard tests: linear hydrostatic mountain wave, tracer advection, and gravity wave over the Schär-type mountain, as well as density current, inertia-gravity wave, and <span class="hlt">rising</span> thermal bubble. The results from these tests demonstrated that the model using the horizontal SEM and the <span class="hlt">vertical</span> FDM is accurate and robust provided sufficient diffusion is applied. The results with various horizontal resolutions also showed convergence of second-order accuracy due to the accuracy of the time integration scheme and that of the <span class="hlt">vertical</span> direction, although high-order basis functions were used in the horizontal. By using the 2-D slice model, we effectively showed that the combined spatial</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_17 --> <div id="page_18" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="341"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E3SWC..3303025Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E3SWC..3303025Y"><span>Strategic advantages of high-<span class="hlt">rise</span> construction</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yaskova, Natalya</p> <p>2018-03-01</p> <p>Traditional methods to assess the competitiveness of different types of real estate in the context of huge changes of new technological way of life don't provide building solutions that would be correct from a strategic perspective. There are many challenges due to changes in the consumers' behavior in the housing area. A multiplicity of life models, a variety of opportunities and priorities, traditions and new trends in construction should be assessed in terms of prospective benefits in the environment of the emerging new world order. At the same time, the mane discourse of high-<span class="hlt">rise</span> construction mainly relates to its design features, technical innovations, and architectural accents. We need to clarify the criteria for economic evaluation of high-<span class="hlt">rise</span> construction in order to provide decisions with clear and quantifiable contexts. The suggested approach to assessing the strategic advantage of high-<span class="hlt">rise</span> construction and the prospects for capitalization of high-<span class="hlt">rise</span> buildings poses new challenges for the economy to identify adequate quantitative assessment methods of the high-<span class="hlt">rise</span> buildings economic efficiency, taking into account all stages of their life cycle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998JGR...103.7549H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998JGR...103.7549H"><span>Bifurcation of the Kuroshio Extension at the Shatsky <span class="hlt">Rise</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hurlburt, Harley E.; Metzger, E. Joseph</p> <p>1998-04-01</p> <p>A <span class="hlt">1</span>/16° six-layer Pacific Ocean model north of 20°S is used to investigate the bifurcation of the Kuroshio Extension at the main Shatsky <span class="hlt">Rise</span> and the pathway of the northern branch from the bifurcation to the subarctic front. Upper ocean-topographic coupling via a mixed barotropic-baroclinic instability is essential to this bifurcation and to the formation and mean pathway of the northern branch as are several aspects of the Shatsky <span class="hlt">Rise</span> complex of topography and the latitude of the Kuroshio Extension in relation to the topography. The flow instabilities transfer energy to the abyssal layer where it is constrained by geostrophic contours of the bottom topography. The topographically constrained abyssal currents in turn steer upper ocean currents, which do not directly impinge on the bottom topography. This includes steering of mean pathways. Obtaining sufficient coupling requires very fine resolution of mesoscale variability and sufficient eastward penetration of the Kuroshio as an unstable inertial jet. Resolution of <span class="hlt">1</span>/8° for each variable was not sufficient in this case. The latitudinal extent of the main Shatsky <span class="hlt">Rise</span> (31°N-36°N) and the shape of the downward slope on the north side are crucial to the bifurcation at the main Shatsky <span class="hlt">Rise</span>, with both branches passing north of the peak. The well-defined, relatively steep and straight eastern edge of the Shatsky <span class="hlt">Rise</span> topographic complex (30°N-42°N) and the southwestward abyssal flow along it play a critical role in forming the rest of the Kuroshio northern branch which flows in the opposite direction. A deep pass between the main Shatsky <span class="hlt">Rise</span> and the rest of the ridge to the northeast helps to link the northern fork of the bifurcation at the main <span class="hlt">rise</span> to the rest of the northern branch. Two <span class="hlt">1</span>/16° "identical twin" interannual simulations forced by daily winds 1981-1995 show that the variability in this region is mostly nondeterministic on all timescales that could be examined (up to 7 years in these 15-year</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=America%27s+AND+military&pg=4&id=EJ866709','ERIC'); return false;" href="https://eric.ed.gov/?q=America%27s+AND+military&pg=4&id=EJ866709"><span>Choices for a <span class="hlt">Rising</span> Generation</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Obama, Barack</p> <p>2008-01-01</p> <p>This article presents an essay by the 2008 Democratic Party Presidential Nominee. This essay focuses on the role of the <span class="hlt">rising</span> generation in bringing about real change in America. The author contends that, at this historic moment, Americans must ask their <span class="hlt">rising</span> generation to serve their country as Americans always have--by working on a political…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20100017220','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20100017220"><span>Mars Structural and Stratigraphic Mapping along the Coprates <span class="hlt">Rise</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Saunders, R Stephen</p> <p>2009-01-01</p> <p>This geologic mapping project supports a topical study of structures in east Thaumasia associated with the Coprates <span class="hlt">rise</span>. The study examines cuesta-like features on the east flank of the Coprates <span class="hlt">rise</span> first identified by Saunders et al. [<span class="hlt">1</span>]. Mapping combines detailed local stratigraphy, structural geology and topography. Hogbacks and cuestas indicate erosion of tilted rock units. The extent of the erosion will be determined in the course of the mapping. The region of interest lies along the eastern margin of Thaumasia bounded by latitudes -15 and -35 and longitudes 50 to 70 W (Figure <span class="hlt">1</span>). Three MTM geologic quadrangles are being compiled for publication by the USGS (-20057, -25057, -30057). All existing data sources are used including THEMIS, MOC, CTX, Hi<span class="hlt">RISE</span>, MOLA and gravity, as well as higher level data available through the PDS data nodes at ASU, UA and Washington University. The extremely valuable ASU JMARS tools are used for analysis of many of the data sets. ArcGIS software has been obtained and is being learned for the map compilation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70125421','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70125421"><span>Comment [on “Sea level <span class="hlt">rise</span> shown to drive coastal erosion”</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Pilkey, Orrin H.; Young, Robert S.; Bush, David M.</p> <p>2000-01-01</p> <p>Leatherman et al. [2000] (Eos, Trans., AGU, February 8, 2000, p.55) affirm that global eustatic sea-level <span class="hlt">rise</span> is driving coastal erosion. Furthermore, they argue that the long-term average rate of shoreline retreat is 150 times the rate of sea-level <span class="hlt">rise</span>. This rate, they say, is more than a magnitude greater than would be expected from a simple response to sea-level <span class="hlt">rise</span> through inundation of the shoreline. We agree that sea-level <span class="hlt">rise</span> is the primary factor causing shoreline retreat in stable coastal areas.This is intuitive. We also believe, however, that the Leatherman et al. [2000] study has greatly underestimated the rate of coastal recession along most low slope shorelines. Slopes along the North Carolina continental shelf/coastal plain approach 10,000:<span class="hlt">1</span>. To us, this suggests that we should expect rates of shoreline recession 10,000 times the rate of sea-level <span class="hlt">rise</span> through simple inundation of the shoreline.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27885566','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27885566"><span><span class="hlt">Vertical</span> zonation of soil fungal community structure in a Korean pine forest on Changbai Mountain, China.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ping, Yuan; Han, Dongxue; Wang, Ning; Hu, Yanbo; Mu, Liqiang; Feng, Fujuan</p> <p>2017-01-01</p> <p>Changbai Mountain, with intact montane <span class="hlt">vertical</span> vegetation belts, is located at a sensitive area of global climate change and a central distribution area of Korean pine forest. Broad-leaved Korean pine mixed forest (Pinus koraiensis as an edificator) is the most representative zonal climax vegetation in the humid region of northeastern China; their <span class="hlt">vertical</span> zonation is the most intact and representative on Changbai Mountain. In this study, we analyzed the composition and diversity of soil fungal communities in the Korean pine forest on Changbai Mountain at elevations ranging from 699 to 1177 m using Illumina High-throughput sequencing. We obtained a total 186,663 optimized sequences, with an average length of 268.81 bp. We found soil fungal diversity index was decreased with increasing elevation from 699 to 937 m and began to <span class="hlt">rise</span> after reaching 1044 m; the richness and evenness indices were decreased with an increase in elevation. Soil fungal compositions at the phylum, class and genus levels varied significantly at different elevations, but with the same dominant fungi. Beta-diversity analysis indicated that the similarity of fungal communities decreased with an increased <span class="hlt">vertical</span> distance between the sample plots, showing a distance-decay relationship. Variation partition analysis showed that geographic distance (mainly elevation gradient) only explained 20.53 % of the total variation of fungal community structure, while soil physicochemical factors explained 69.78 %.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JPhCS.744a2071B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JPhCS.744a2071B"><span>Modeling and control of a cable-suspended robot for inspection of <span class="hlt">vertical</span> structures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barry, Nicole; Fisher, Erin; Vaughan, Joshua</p> <p>2016-09-01</p> <p>In this paper, a cable-driven system is examined for the application of inspection of large, <span class="hlt">vertical</span>-walled structures such as chemical storage tanks, large ship hulls, and high-<span class="hlt">rise</span> buildings. Such cable-driven systems are not commonly used for these tasks due to vibration, which decreases inspection accuracy and degrades safety. The flexible nature of the cables make them difficult to control. In this paper, input shaping is implemented on a cable-driven system to reduce vibration. To design the input shapers, a model of the cable-driven system was developed. Analysis of the dominant dynamics and changes in them over the large workspace are also presented. The performance improvements provided by the input shaping controller are quantified through a series of simulations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DPPU11054H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DPPU11054H"><span>Measurement of <span class="hlt">vertical</span> stability metrics in KSTAR</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hahn, Sang-Hee; Humphreys, D. A.; Mueller, D.; Bak, J. G.; Eidietis, N. W.; Kim, H.-S.; Ko, J. S.; Walker, M. L.; Kstar Team</p> <p>2017-10-01</p> <p>The paper summarizes results of multi-year ITPA experiments regarding measurement of the <span class="hlt">vertical</span> stabilization capability of KSTAR discharges, including most recent measurements at the highest achievable elongation (κ 2.0 - 2.<span class="hlt">1</span>). The measurements of the open-loop growth rate of VDE (γz) and the maximum controllable <span class="hlt">vertical</span> displacement (ΔZmax) are done by the release-and-catch method. The dynamics of the <span class="hlt">vertical</span> movement of the plasma is verified by both relevant magnetic reconstructions and non-magnetic diagnostics. The measurements of γz and ΔZmax were done for different plasma currents, βp, internal inductances, elongations and different configurations of the vessel conductors that surround the plasma as the first wall. Effects of control design choice and diagnostics noise are discussed, and comparison with the axisymmetric plasma response model is given for partial accounting for the measured control capability. This work supported by Ministry of Science, ICT, and Future Planning under KSTAR project.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28772073','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28772073"><span>Extremely Black <span class="hlt">Vertically</span> Aligned Carbon Nanotube Arrays for Solar Steam Generation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yin, Zhe; Wang, Huimin; Jian, Muqiang; Li, Yanshen; Xia, Kailun; Zhang, Mingchao; Wang, Chunya; Wang, Qi; Ma, Ming; Zheng, Quan-Shui; Zhang, Yingying</p> <p>2017-08-30</p> <p>The unique structure of a <span class="hlt">vertically</span> aligned carbon nanotube (VACNT) array makes it behave most similarly to a blackbody. It is reported that the optical absorptivity of an extremely black VACNT array is about 0.98-0.99 over a large spectral range of 200 nm-200 μm, inspiring us to explore the performance of VACNT arrays in solar energy harvesting. In this work, we report the highly efficient steam generation simply by laminating a layer of VACNT array on the surface of water to harvest solar energy. It is found that under solar illumination the temperature of upper water can significantly increase with obvious water steam generated, indicating the efficient solar energy harvesting and local temperature <span class="hlt">rise</span> by the thin layer of VACNTs. We found that the evaporation rate of water assisted by VACNT arrays is 10 times that of bare water, which is the highest ratio for solar-thermal-steam generation ever reported. Remarkably, the solar thermal conversion efficiency reached 90%. The excellent performance could be ascribed to the strong optical absorption and local temperature <span class="hlt">rise</span> induced by the VACNT layer, as well as the ultrafast water transport through the VACNT layer due to the frictionless wall of CNTs. Based on the above, we further demonstrated the application of VACNT arrays in solar-driven desalination.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28155062','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28155062"><span>Bipedicle <span class="hlt">Vertical</span> Mammoplasty Associated with Liposuction/Lipotunnelization.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Maia, Manuel; Pinto, Cristina</p> <p>2017-06-01</p> <p><span class="hlt">Vertical</span> techniques have become more popular and versatile in breast reduction and mastopexy procedures. The authors introduce a combination of <span class="hlt">vertical</span> mammoplasty with <span class="hlt">vertical</span> bipedicle technique, presenting some innovations concerning pedicle design, glandular dissection pattern, and the role of liposuction. In this article, we describe a personal surgical technique and analyze the results on 73 mastopexy and breast reduction patients, operated on between 2012 and 2014 by the senior author. The most important aspects of this technique are as follows: <span class="hlt">1</span>. the concept of lipotunnelization, where parenchymal tunnels are made with cannulas without suction; 2. the systematic use of liposuction/lipotunnelization to recreate the inframammary fold and decrease pillars height; 3. the use of a thick and narrow <span class="hlt">vertical</span> pedicle (bipedicled); 4. no skin undermining. Twenty-five patients underwent mastopexy and 48 breast reduction. Results were evaluated by clinical examination and patient photographs. Good shape and projection of the breast and correct nipple elevation were achieved, correlating with a high level of patient satisfaction. No major complications occurred. This lipomammoplasty technique is a reliable and effective option, suitable for a wide range of symptomatic macromastia/ptosis. Adjunctive use of liposuction/lipotunnelization in <span class="hlt">vertical</span> breast reduction accomplishes effective contouring of the breast with low associated complications and significantly reduces the revision rates. This journal requires that authors assign a level of evidence to each article. For a full description of these evidence-based medicine ratings, please refer to the Table of Contents or the online Instructions to Authors www.springer.com/00266 .</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15882104','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15882104"><span><span class="hlt">Vertical</span> integration of medical education: Riverland experience, South Australia.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rosenthal, D R; Worley, P S; Mugford, B; Stagg, P</p> <p>2004-01-01</p> <p><span class="hlt">Vertical</span> integration of medical education is currently a prominent international topic, resulting from recent strategic initiatives to improve medical education and service delivery in areas of poorly met medical need. In this article, <span class="hlt">vertical</span> integration of medical education is defined as 'a grouping of curricular content and delivery mechanisms, traversing the traditional boundaries of undergraduate, postgraduate and continuing medical education, with the intent of enhancing the transfer of knowledge and skills between those involved in the learning-teaching process'. Educators closely involved with <span class="hlt">vertically</span> integrated teaching in the Riverland of South Australia present an analytical description of the educational dynamics of this system. From this analysis, five elements are identified which underpin the process of successful <span class="hlt">vertical</span> integration: (<span class="hlt">1</span>) raised educational stakes; (2) local ownership; (3) broad university role; (4) longer attachments; and (5) shared workforce vision. Given the benefits to the Riverland medical education programs described in this paper, it is not surprising that <span class="hlt">vertical</span> integration of medical education is a popular goal in many rural regions throughout the world. Although different contexts will result in different functional arrangements, it could be argued that the five principles outlined in this article can be applied in any region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006ECSS...67..475M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006ECSS...67..475M"><span>Seagrass ( Posidonia oceanica) <span class="hlt">vertical</span> growth as an early indicator of fish farm-derived stress</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Marbà, Núria; Santiago, Rocío; Díaz-Almela, Elena; Álvarez, Elvira; Duarte, Carlos M.</p> <p>2006-04-01</p> <p>The usefulness of <span class="hlt">vertical</span> rhizome growth as an early indicator of fish farm impacts to Posidonia oceanica meadows was tested by comparing annual estimates of <span class="hlt">vertical</span> rhizome growth, quantified retrospectively, at distances ranging between 5 and 1200 m from fish cages at four Mediterranean locations (Cyprus, Greece, Italy and Spain). The studied fish farms had been operating since, at least, 1997, producing between 150 and 1150 tons yr -<span class="hlt">1</span> of sea bream ( Sparus aurata) and sea bass ( Dicentrachus labrax), and, at Italy, also sharpsnout sea bream ( Diplodus puntazzo). The reconstructed <span class="hlt">vertical</span> rhizome growth spanned from 19 to 25 years of growth, depending on sites, and the average <span class="hlt">vertical</span> rhizome growth before the onset of fish farm operations ranged between 4.48 and 8.79 mm yr -<span class="hlt">1</span>. The <span class="hlt">vertical</span> rhizome growth after the onset of farming activities declined significantly ( t-test, P < 0.05) from the control station (at >800 m from the farm; <span class="hlt">vertical</span> growth rate averaged 6.79, 5.52, 3.89 and 3.70 mm yr -<span class="hlt">1</span> at Cyprus, Greece, Italy and Spain control stations, respectively) to the impacted one (at 5-300 m from the farm; <span class="hlt">vertical</span> growth rate was 4.82, 3.52, 2.77 and <span class="hlt">1</span>.92 mm yr -<span class="hlt">1</span> at Cyprus, Greece, Italy and Spain impacted stations, respectively) at each farm. Moreover, <span class="hlt">vertical</span> growth significantly ( t-test, P < 0.05) declined by about twofold following the onset of fish farm operations for the extant meadow nearest to the cages, as well as those supporting intermediate impacts at distances 35-400 m from the cages. <span class="hlt">Vertical</span> rhizome growth was not significantly affected after the onset of fish farm operations for the meadows located more than 800 m from the farm, except in those from the Italian site, the largest farm. Examination of the time course of <span class="hlt">vertical</span> growth for individual rhizomes in the areas of the meadow nearest to the farms, except for those at Cyprus, showed that the decline in <span class="hlt">vertical</span> growth was initiated within the year of the onset of farming</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006PhDT.......113K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006PhDT.......113K"><span>A search for thermospheric composition perturbations due to <span class="hlt">vertical</span> winds</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Krynicki, Matthew P.</p> <p></p> <p>The thermosphere is generally in hydrostatic equilibrium, with winds blowing horizontally along stratified constant-pressure surfaces, driven by the dayside-to-nightside pressure gradient. A marked change in this paradigm resulted after Spencer et al. [1976] reported <span class="hlt">vertical</span> wind measurements of 80 m·s-<span class="hlt">1</span> from analyses of AE-C satellite data. It is now established that the thermosphere routinely supports large-magnitude (˜30-150 m·s-<span class="hlt">1</span>) <span class="hlt">vertical</span> winds at auroral latitudes. These <span class="hlt">vertical</span> winds represent significant departure from hydrostatic and diffusive equilibrium, altering locally---and potentially globally---the thermosphere's and ionosphere's composition, chemistry, thermodynamics and energy budget. Because of their localized nature, large-magnitude <span class="hlt">vertical</span> wind effects are not entirely known. This thesis presents ground-based Fabry-Perot Spectrometer OI(630.0)-nm observations of upper-thermospheric <span class="hlt">vertical</span> winds obtained at Inuvik, NT, Canada and Poker Flat, AK. The wind measurements are compared with <span class="hlt">vertical</span> displacement estimates at ˜104 km2 horizontal spatial scales determined from a new modification to the electron transport code of Lummerzheim and Lilensten [1994] as applied to FUV-wavelength observations by POLAR spacecraft's Ultraviolet Imager [Torr et al. , 1995]. The modification, referred to as the column shift, simulates <span class="hlt">vertical</span> wind effects such as neutral transport and disruption of diffusive equilibrium by <span class="hlt">vertically</span> displacing the Hedin [1991] MSIS-90 [O2]/[N2] and [O]/([N2]+[O2]) mixing ratios and subsequently redistributing the O, O2, and N 2 densities used in the transport code. Column shift estimates are inferred from comparisons of UVI OI(135.6)-nm auroral observations to their corresponding modeled emission. The modeled OI(135.6)-nm brightness is determined from the modeled thermospheric response to electron precipitation and estimations of the energy flux and characteristic energy of the precipitation, which are inferred from UVI</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23838700','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23838700"><span>Trauma of the upper cervical spine: focus on <span class="hlt">vertical</span> atlantoaxial dislocation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pissonnier, M L; Lazennec, J Y; Renoux, J; Rousseau, M A</p> <p>2013-10-01</p> <p>Traumatic ligament injuries of the craniovertebral junction, either isolated or associated with bone avulsion or fracture, often lead to death. These injuries are rare and underrated but are increasingly seen in emergency departments due to the improvement in initial on-scene management of accidents. <span class="hlt">Vertical</span> atlantoaxial dislocation (AAD) is a specific lesion that was barely reported. Based on our experience, our goal was to systematically investigate the prevalence and prognosis of traumatic <span class="hlt">vertical</span> AAD and discuss its management. All cervical CT scans performed at our institution between 2006 and 2010 for cervical trauma in adults were retrospectively reviewed. Based on the measurement of lateral mass index (LMI), defined as the gap between C<span class="hlt">1</span> and C2 articular facets, we identified three cases of traumatic <span class="hlt">vertical</span> AAD in 300 CT scans. Their medical records were investigated. The incidence of <span class="hlt">vertical</span> AAD was <span class="hlt">1</span>% in the exposed population. One case was an isolated <span class="hlt">vertical</span> AAD and two were associated with a type II odontoid fracture. We report the first case in the literature of unilateral <span class="hlt">vertical</span> AAD. Two patients died rapidly; the survivor was treated with occipitocervical fixation. Specific maneuvers were used for immobilization and reduction. This study found a not insignificant incidence of <span class="hlt">vertical</span> AAD and a high lethality rate. LMI appears to be a relevant radiological criterion for this diagnosis, for which traction is contraindicated. Associated neurological or vascular damage should be suspected and investigated. In our experience, spinal surgical fixation is required because of major instability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2013-title33-vol1/pdf/CFR-2013-title33-vol1-sec84-03.pdf','CFR2013'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2013-title33-vol1/pdf/CFR-2013-title33-vol1-sec84-03.pdf"><span>33 CFR 84.03 - <span class="hlt">Vertical</span> positioning and spacing of lights.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2013&page.go=Go">Code of Federal Regulations, 2013 CFR</a></p> <p></p> <p>2013-07-01</p> <p>... 33 Navigation and Navigable Waters <span class="hlt">1</span> 2013-07-01 2013-07-01 false <span class="hlt">Vertical</span> positioning and spacing... <span class="hlt">Vertical</span> positioning and spacing of lights. (a) On a power-driven vessel of 20 meters or more in length the... is carried, then that light, at a height above the hull of not less than 5 meters, and, if the...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2014-title33-vol1/pdf/CFR-2014-title33-vol1-sec84-03.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title33-vol1/pdf/CFR-2014-title33-vol1-sec84-03.pdf"><span>33 CFR 84.03 - <span class="hlt">Vertical</span> positioning and spacing of lights.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2014&page.go=Go">Code of Federal Regulations, 2014 CFR</a></p> <p></p> <p>2014-07-01</p> <p>... 33 Navigation and Navigable Waters <span class="hlt">1</span> 2014-07-01 2014-07-01 false <span class="hlt">Vertical</span> positioning and spacing... <span class="hlt">Vertical</span> positioning and spacing of lights. (a) On a power-driven vessel of 20 meters or more in length the... is carried, then that light, at a height above the hull of not less than 5 meters, and, if the...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title33-vol1/pdf/CFR-2012-title33-vol1-sec84-03.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title33-vol1/pdf/CFR-2012-title33-vol1-sec84-03.pdf"><span>33 CFR 84.03 - <span class="hlt">Vertical</span> positioning and spacing of lights.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2012&page.go=Go">Code of Federal Regulations, 2012 CFR</a></p> <p></p> <p>2012-07-01</p> <p>... 33 Navigation and Navigable Waters <span class="hlt">1</span> 2012-07-01 2012-07-01 false <span class="hlt">Vertical</span> positioning and spacing... <span class="hlt">Vertical</span> positioning and spacing of lights. (a) On a power-driven vessel of 20 meters or more in length the... is carried, then that light, at a height above the hull of not less than 5 meters, and, if the...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title33-vol1/pdf/CFR-2011-title33-vol1-sec84-03.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title33-vol1/pdf/CFR-2011-title33-vol1-sec84-03.pdf"><span>33 CFR 84.03 - <span class="hlt">Vertical</span> positioning and spacing of lights.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-07-01</p> <p>... 33 Navigation and Navigable Waters <span class="hlt">1</span> 2011-07-01 2011-07-01 false <span class="hlt">Vertical</span> positioning and spacing... <span class="hlt">Vertical</span> positioning and spacing of lights. (a) On a power-driven vessel of 20 meters or more in length the... is carried, then that light, at a height above the hull of not less than 5 meters, and, if the...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Energy+AND+Prices&pg=7&id=ED499913','ERIC'); return false;" href="https://eric.ed.gov/?q=Energy+AND+Prices&pg=7&id=ED499913"><span>Wrestling <span class="hlt">Rising</span> Costs with Innovation. Policy Matters. Volume 4, Number <span class="hlt">1</span>, January 2007</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Markowitz, Melissa</p> <p>2007-01-01</p> <p>While tuition costs are likely the most talked about topic in higher education, focusing on the institutional finance is equally important. The growing expenses associated with educating students is often a catalyst for <span class="hlt">rising</span> tuition and fees, and they play a large role as educators plan for the future of their institutions. Although higher…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMED53C0174D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMED53C0174D"><span>Preparing Norfolk Area Students for America's Second Highest Sea Level <span class="hlt">Rise</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dunbar, R. R.</p> <p>2017-12-01</p> <p>The nonprofit Elizabeth River Project located in Hampton Roads, Virginia was awarded a 3-year national NOAA Environmental Literacy award 2016-2019 to teach 21,000 K-12 youth how to help restore one of the most polluted rivers on the Chesapeake Bay and to help create a resilient community that is facing impacts from the <span class="hlt">rising</span> seas and changing climate. Through a community collaboration, partners are also creating perhaps the nation's first Youth Resilience Strategy with a vision, goals, best practices and resources on engaging youth to help create resilient cities facing environmental and economic changes. During Year <span class="hlt">1</span>, 7,000 elementary students held field investigations aboard the floating classroom Learning Barge and at Paradise Creek Nature Park and helped restore wetland restoration sites. Students performed inquiry based investigations, learned stewardship actions to help create resilience and showed a 40% increase in knowledge. Year <span class="hlt">1</span> best practices in teaching resilience include youth: getting out of the classroom, discovering how rain water travels, performing bioblitzes and water quality testing, engaging in hands-on GreenSTEM activities, using investigation tools, creating innovative solutions to retain and reuse rain water, creating art and voicing their opinions on creating a resilient community.Lessons learned include developing engaging inquiry questions based on creating a resilient community. These included: "What are the impact of <span class="hlt">rising</span> tides?", "How can sea level <span class="hlt">rise</span> affect river animals?", "How can we be safe and prepare for extreme weather and flooding as the sea level <span class="hlt">rises</span>?", "How has the way people worked with the Elizabeth River changed?", "How could sea level <span class="hlt">rise</span> affect the Elizabeth River's water quality?", "How hot might the air temperature get by 2050 and what can we do to keep it cooler?", "What does this park show us about sea level <span class="hlt">rise</span> and other ways our climate is changing?", "How do trees help make our park and community</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_18 --> <div id="page_19" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="361"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title21-vol2/pdf/CFR-2011-title21-vol2-sec137-185.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title21-vol2/pdf/CFR-2011-title21-vol2-sec137-185.pdf"><span>21 CFR 137.185 - Enriched self-<span class="hlt">rising</span> flour.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-04-01</p> <p>... 21 Food and Drugs 2 2011-04-01 2011-04-01 false Enriched self-<span class="hlt">rising</span> flour. 137.185 Section 137... Cereal Flours and Related Products § 137.185 Enriched self-<span class="hlt">rising</span> flour. Enriched self-<span class="hlt">rising</span> flour... carbon dioxide evolved under ordinary conditions of use of the enriched self-<span class="hlt">rising</span> flour is not less...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2013-title21-vol2/pdf/CFR-2013-title21-vol2-sec137-185.pdf','CFR2013'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2013-title21-vol2/pdf/CFR-2013-title21-vol2-sec137-185.pdf"><span>21 CFR 137.185 - Enriched self-<span class="hlt">rising</span> flour.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2013&page.go=Go">Code of Federal Regulations, 2013 CFR</a></p> <p></p> <p>2013-04-01</p> <p>... 21 Food and Drugs 2 2013-04-01 2013-04-01 false Enriched self-<span class="hlt">rising</span> flour. 137.185 Section 137... Cereal Flours and Related Products § 137.185 Enriched self-<span class="hlt">rising</span> flour. Enriched self-<span class="hlt">rising</span> flour... carbon dioxide evolved under ordinary conditions of use of the enriched self-<span class="hlt">rising</span> flour is not less...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title21-vol2/pdf/CFR-2012-title21-vol2-sec137-185.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title21-vol2/pdf/CFR-2012-title21-vol2-sec137-185.pdf"><span>21 CFR 137.185 - Enriched self-<span class="hlt">rising</span> flour.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2012&page.go=Go">Code of Federal Regulations, 2012 CFR</a></p> <p></p> <p>2012-04-01</p> <p>... 21 Food and Drugs 2 2012-04-01 2012-04-01 false Enriched self-<span class="hlt">rising</span> flour. 137.185 Section 137... Cereal Flours and Related Products § 137.185 Enriched self-<span class="hlt">rising</span> flour. Enriched self-<span class="hlt">rising</span> flour... carbon dioxide evolved under ordinary conditions of use of the enriched self-<span class="hlt">rising</span> flour is not less...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2014-title21-vol2/pdf/CFR-2014-title21-vol2-sec137-185.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title21-vol2/pdf/CFR-2014-title21-vol2-sec137-185.pdf"><span>21 CFR 137.185 - Enriched self-<span class="hlt">rising</span> flour.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2014&page.go=Go">Code of Federal Regulations, 2014 CFR</a></p> <p></p> <p>2014-04-01</p> <p>... 21 Food and Drugs 2 2014-04-01 2014-04-01 false Enriched self-<span class="hlt">rising</span> flour. 137.185 Section 137... Cereal Flours and Related Products § 137.185 Enriched self-<span class="hlt">rising</span> flour. Enriched self-<span class="hlt">rising</span> flour... carbon dioxide evolved under ordinary conditions of use of the enriched self-<span class="hlt">rising</span> flour is not less...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/2004/1021/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/2004/1021/"><span>Coastal vulnerability assessment of Olympic National Park to sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Pendleton, Elizabeth A.; Hammar-Klose, Erika S.; Thieler, E. Robert; Williams, S. Jeffress</p> <p>2004-01-01</p> <p>A coastal vulnerability index (CVI) was used to map the relative vulnerability of the coast to future sea-level <span class="hlt">rise</span> within Olympic National Park (OLYM), Washington. The CVI scores the following in terms of their physical contribution to sea-level <span class="hlt">rise</span>-related coastal change: geomorphology, regional coastal slope, rate of relative sea-level <span class="hlt">rise</span>, shoreline change rates, mean tidal range and mean wave height. The rankings for each variable were combined and an index value calculated for <span class="hlt">1</span>-minute grid cells covering the park. The CVI highlights those regions where the physical effects of sea-level <span class="hlt">rise</span> might be the greatest. This approach combines the coastal system's susceptibility to change with its natural ability to adapt to changing environmental conditions, yielding a quantitative, although relative, measure of the park's natural vulnerability to the effects of sea-level <span class="hlt">rise</span>. The CVI provides an objective technique for evaluation and long-term planning by scientists and park managers. The Olympic National Park coast consists of rocky headlands, pocket beaches, glacial-fluvial features, and sand and gravel beaches. The Olympic coastline that is most vulnerable to sea-level <span class="hlt">rise</span> are beaches in gently sloping areas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17347316','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17347316"><span>Does plyometric training improve <span class="hlt">vertical</span> jump height? A meta-analytical review.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Markovic, Goran</p> <p>2007-06-01</p> <p>The aim of this study was to determine the precise effect of plyometric training (PT) on <span class="hlt">vertical</span> jump height in healthy individuals. Meta-analyses of randomised and non-randomised controlled trials that evaluated the effect of PT on four typical <span class="hlt">vertical</span> jump height tests were carried out: squat jump (SJ); countermovement jump (CMJ); countermovement jump with the arm swing (CMJA); and drop jump (DJ). Studies were identified by computerised and manual searches of the literature. Data on changes in jump height for the plyometric and control groups were extracted and statistically pooled in a meta-analysis, separately for each type of jump. A total of 26 studies yielding 13 data points for SJ, 19 data points for CMJ, 14 data points for CMJA and 7 data points for DJ met the initial inclusion criteria. The pooled estimate of the effect of PT on <span class="hlt">vertical</span> jump height was 4.7% (95% CI <span class="hlt">1</span>.8 to 7.6%), 8.7% (95% CI 7.0 to 10.4%), 7.5% (95% CI 4.2 to 10.8%) and 4.7% (95% CI 0.8 to 8.6%) for the SJ, CMJ, CMJA and DJ, respectively. When expressed in standardised units (ie, effect sizes), the effect of PT on <span class="hlt">vertical</span> jump height was 0.44 (95% CI 0.15 to 0.72), 0.88 (95% CI 0.64 to <span class="hlt">1</span>.11), 0.74 (95% CI 0.47 to <span class="hlt">1</span>.02) and 0.62 (95% CI 0.18 to <span class="hlt">1</span>.05) for the SJ, CMJ, CMJA and DJ, respectively. PT provides a statistically significant and practically relevant improvement in <span class="hlt">vertical</span> jump height with the mean effect ranging from 4.7% (SJ and DJ), over 7.5% (CMJA) to 8.7% (CMJ). These results justify the application of PT for the purpose of development of <span class="hlt">vertical</span> jump performance in healthy individuals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED076828.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED076828.pdf"><span>Fire Problems in High-<span class="hlt">Rise</span> Buildings. California Fire Service Training Program.</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>California State Dept. of Education, Sacramento. Bureau of Industrial Education.</p> <p></p> <p>Resulting from a conference concerned with high-<span class="hlt">rise</span> fire problems, this manual has been prepared as a fire department training manual and as a reference for students enrolled in fire service training courses. Information is provided for topics dealing with: (<span class="hlt">1</span>) Typical Fire Problems in High-<span class="hlt">Rise</span> Buildings, (2) Heat, (3) Smoke and Fire Gases, (4)…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JEMat..47..655N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JEMat..47..655N"><span>Temporal Response of Dilute Nitride Multi-Quantum-Well <span class="hlt">Vertical</span> Cavity Enhanced Photodetector</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nordin, M. S.; Sarcan, F.; Gunes, M.; Boland-Thoms, A.; Erol, A.; Vickers, A. J.</p> <p>2018-01-01</p> <p>The temporal response characteristics of a GaInNAs-based <span class="hlt">vertical</span> resonant cavity enhanced photodetector device are presented for operation at λ ≈ <span class="hlt">1</span>.3 μm. The absorption layers of the device are composed of nine 7-nm-thick Ga0.65In0.35N0.02As0.98 quantum wells and are sandwiched between top and bottom AlGaAs/GaAs distributed Bragg reflectors (DBRs). The temperature dependence of the transient photoconductivity (TPC) under different light intensities and bias voltages is reported. Photoluminescence measurements were also performed on structures with and without the top DBR to determine their optical response under continuous illumination. The response time was measured using excitation from a 1047-nm pulsed neodymium-doped yttrium lithium fluoride laser with pulse width of 500 ps and repetition rate of <span class="hlt">1</span> kHz. The <span class="hlt">rise</span> time of the TPC was 2.27 ns at T = 50 K, decreasing to <span class="hlt">1</span>.79 ns at T = 300 K. The TPC decay time was 25.44 ns at T = 50 K, decreasing to 16.58 ns at T = 300 K. With detectivity of 2.28 × 10^{10} {cm}√ {Hz} / {W} and noise-equivalent power of 2.45 × 10^{ - 11} {W/}√ {Hz} , the proposed device is faster and more sensitive with better signal-to-noise ratio compared with other GaInNAs-based resonant cavity enhanced photodetectors (RCEPDs) for operation at <span class="hlt">1</span>.3 μm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15042085','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15042085"><span>Mass and volume contributions to twentieth-century global sea level <span class="hlt">rise</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Miller, Laury; Douglas, Bruce C</p> <p>2004-03-25</p> <p>The rate of twentieth-century global sea level <span class="hlt">rise</span> and its causes are the subjects of intense controversy. Most direct estimates from tide gauges give <span class="hlt">1</span>.5-2.0 mm yr(-<span class="hlt">1</span>), whereas indirect estimates based on the two processes responsible for global sea level <span class="hlt">rise</span>, namely mass and volume change, fall far below this range. Estimates of the volume increase due to ocean warming give a rate of about 0.5 mm yr(-<span class="hlt">1</span>) (ref. 8) and the rate due to mass increase, primarily from the melting of continental ice, is thought to be even smaller. Therefore, either the tide gauge estimates are too high, as has been suggested recently, or one (or both) of the mass and volume estimates is too low. Here we present an analysis of sea level measurements at tide gauges combined with observations of temperature and salinity in the Pacific and Atlantic oceans close to the gauges. We find that gauge-determined rates of sea level <span class="hlt">rise</span>, which encompass both mass and volume changes, are two to three times higher than the rates due to volume change derived from temperature and salinity data. Our analysis supports earlier studies that put the twentieth-century rate in the <span class="hlt">1</span>.5-2.0 mm yr(-<span class="hlt">1</span>) range, but more importantly it suggests that mass increase plays a larger role than ocean warming in twentieth-century global sea level <span class="hlt">rise</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17030050','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17030050"><span>Factors accounting for the <span class="hlt">rise</span> in health-care spending in the United States: the role of <span class="hlt">rising</span> disease prevalence and treatment intensity.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Thorpe, Kenneth E</p> <p>2006-11-01</p> <p>To examine the factors responsible for the <span class="hlt">rise</span> in health- care spending in the United States over the past 15 years. Nationally representative survey data from 1987 and 2003 were used to examine the top medical conditions accounting for the <span class="hlt">rise</span> in spending. I also estimate how much of the <span class="hlt">rise</span> is traced to <span class="hlt">rising</span> treated disease prevalence and <span class="hlt">rising</span> spending per case. The study finds most of the <span class="hlt">rise</span> in spending is linked to <span class="hlt">rising</span> rates of treated disease prevalence. The <span class="hlt">rise</span> in prevalence is associated with the doubling of obesity in the US and changes in clinical thresholds for treating asymptomatic patients with certain cardiovascular risk factors. Most of the policy solutions offered in the US to slow the growth in spending do not address the fundamental factors accounting for spending growth. More aggressive efforts for slowing the growth in obesity among adults and children should be centre-stage in the efforts to slow the <span class="hlt">rise</span> in health-care spending.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19800052962&hterms=chemical+fertilizer&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dchemical%2Bfertilizer','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19800052962&hterms=chemical+fertilizer&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dchemical%2Bfertilizer"><span>The <span class="hlt">vertical</span> distribution of tropospheric ammonia</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Levine, J. S.; Hoell, J. M.; Augustsson, T. R.</p> <p>1980-01-01</p> <p>A one-dimensional tropospheric photochemical model is used to simulate measured profiles of NH3 obtained with the Infrared Heterodyne Radiometer. The relative roles of homogeneous loss, heterogeneous loss, and <span class="hlt">vertical</span> eddy transport are discussed in terms of selecting parameters which best fit the measurements. The best fit was obtained for a <span class="hlt">vertical</span> eddy diffusion coefficient of 200,000/sq cm per sec or greater (corresponding to a characteristic <span class="hlt">vertical</span> transport time in excess of about 35 days), and a characteristic heterogeneous loss time in excess of 10 days. The characteristic homogeneous chemical loss time was found to be about 40 days at the surface and decreased to about 180 days at 10 km, and not very sensitive to model chemical perturbations. Increased ground-level concentrations of NH3 to about 10 ppb, compared to background surface concentrations of about <span class="hlt">1</span> ppb, were measured several weeks after application of ammonium nitrate fertilizer. This suggests that the volatilization of ammonium nitrate fertilizer is rapid, and an important source of NH3. Because of the characteristic times for the loss mechanisms, synoptic time-scale phenomena may play an important role in determining the tropospheric distribution of NH3 concentrations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24158607','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24158607"><span>Human sensitivity to <span class="hlt">vertical</span> self-motion.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nesti, Alessandro; Barnett-Cowan, Michael; Macneilage, Paul R; Bülthoff, Heinrich H</p> <p>2014-01-01</p> <p>Perceiving <span class="hlt">vertical</span> self-motion is crucial for maintaining balance as well as for controlling an aircraft. Whereas heave absolute thresholds have been exhaustively studied, little work has been done in investigating how <span class="hlt">vertical</span> sensitivity depends on motion intensity (i.e., differential thresholds). Here we measure human sensitivity for <span class="hlt">1</span>-Hz sinusoidal accelerations for 10 participants in darkness. Absolute and differential thresholds are measured for upward and downward translations independently at 5 different peak amplitudes ranging from 0 to 2 m/s(2). Overall <span class="hlt">vertical</span> differential thresholds are higher than horizontal differential thresholds found in the literature. Psychometric functions are fit in linear and logarithmic space, with goodness of fit being similar in both cases. Differential thresholds are higher for upward as compared to downward motion and increase with stimulus intensity following a trend best described by two power laws. The power laws' exponents of 0.60 and 0.42 for upward and downward motion, respectively, deviate from Weber's Law in that thresholds increase less than expected at high stimulus intensity. We speculate that increased sensitivity at high accelerations and greater sensitivity to downward than upward self-motion may reflect adaptations to avoid falling.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=challenges+AND+facilities&id=EJ1017856','ERIC'); return false;" href="https://eric.ed.gov/?q=challenges+AND+facilities&id=EJ1017856"><span>APPA Thought Leaders Report, 2013. Part <span class="hlt">1</span>: The <span class="hlt">Rising</span> Cost of Higher Education</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Facilities Manager, 2013</p> <p>2013-01-01</p> <p>While many issues in higher education are only discussed among members of the education community, the sharp <span class="hlt">rise</span> in costs is no longer a topic solely for academia. Parents and politicians alike are fuming over the apparently unstoppable climb of the cost of a college education. It seems every day a new magazine article or newspaper story bemoans…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E3SWC..3303065E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E3SWC..3303065E"><span>Socioecological Aspects of High-<span class="hlt">rise</span> Construction</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Eichner, Michael; Ivanova, Zinaida</p> <p>2018-03-01</p> <p>In this article, the authors consider the socioecological problems that arise in the construction and operation of high-<span class="hlt">rise</span> buildings. They study different points of view on high-<span class="hlt">rise</span> construction and note that the approaches to this problem are very different. They also analyse projects of modern architects and which attempts are made to overcome negative impacts on nature and mankind. The article contains materials of sociological research, confirming the ambivalent attitude of urban population to high-<span class="hlt">rise</span> buildings. In addition, one of the author's sociological survey reveals the level of environmental preparedness of the university students, studying in the field of "Construction of unique buildings and structures", raising the question of how future specialists are ready to take into account socioecological problems. Conclusion of the authors: the construction of high-<span class="hlt">rise</span> buildings is associated with huge social and environmental risks, negative impact on the biosphere and human health. This requires deepened skills about sustainable design methods and environmental friendly construction technologies of future specialists. Professor M. Eichner presents in the article his case study project results on implementation of holistic eco-sustainable construction principles for mixed-use high-<span class="hlt">rise</span> building in the metropolis of Cairo.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70010837','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70010837"><span>Igneous rocks of the East Pacific <span class="hlt">Rise</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Engel, A.E.J.; Engel, C.G.</p> <p>1964-01-01</p> <p>The apical parts of large volcanoes along the East Pacific <span class="hlt">Rise</span> (islands and seamounts) are encrusted with rocks of the alkali volcanic suite (alkali basalt, andesine- and oligoclase-andesite, and trachyte). In contrast, the more submerged parts of the <span class="hlt">Rise</span> are largely composed of a tholeiitic basalt which has low concentrations of K, P, U, Th, Pb, and Ti. This tholeiitic basalt is either the predominant or the only magma generated in the earth's mantle under oceanic ridges and <span class="hlt">rises</span>. It is at least 1000-fold more abundant than the alkali suite, which is probably derived from tholeiitic basalt by magmatic differentiation in and immediately below the larger volcanoes. Distinction of oceanic tholeiites from almost all continental tholeiites is possible on the simple basis of total potassium content, with the discontinuity at 0.3 to 0.5 percent K2O by weight. Oceanic tholeiites also are readily distinguished from some 19 out of 20 basalts of oceanic islands and seamount cappings by having less than 0.3 percent K2O by weight and more than 48 percent SiO2. Deep drilling into oceanic volcanoes should, however, core basalts transitional between the oceanic tholeiites and the presumed derivative alkali basalts.The composition of the oceanic tholeiites suggests that the mantle under the East Pacific <span class="hlt">Rise</span> contains less than 0.10 percent potassium oxide by weight; 0.<span class="hlt">1</span> part per million of uranium and 0.4 part of thorium; a potassium:rubidium ratio of about 1200 and a potassium: uranium ratio of about 104.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24830257','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24830257"><span>[Vulnerability assessment on the coastal wetlands in the Yangtze Estuary under sea-level <span class="hlt">rise</span>].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cui, Li-Fang; Wang, Ning; Ge, Zhen-Ming; Zhang, Li-Quan</p> <p>2014-02-01</p> <p>To study the response of coastal wetlands to climate change, assess the impacts of climate change on the coastal wetlands and formulate feasible and practical mitigation strategies are the important prerequisite for securing coastal ecosystems. In this paper, the possible impacts of sea level <span class="hlt">rise</span> caused by climate change on the coastal wetlands in the Yangtze Estuary were analyzed by the Source-Pathway-Receptor-Consequence (SPRC) model and IPCC definition on the vulnerability. An indicator system for vulnerability assessment was established, in which sea-level <span class="hlt">rise</span> rate, subsidence rate, habitat elevation, inundation threshold of habitat and sedimentation rate were selected as the key indicators. A quantitatively spatial assessment method based on the GIS platform was established by quantifying each indicator, calculating the vulnerability index and grading the vulnerability index for the assessment of coastal wetlands in the Yangtze Estuary under the scenarios of sea-level <span class="hlt">rise</span>. The vulnerability assessments on the coastal wetlands in the Yangtze Estuary in 2030 and 2050 were performed under two sea-level <span class="hlt">rise</span> scenarios (the present sea-level <span class="hlt">rise</span> trend over recent 30 years and IPCC A<span class="hlt">1</span>F<span class="hlt">1</span> scenario). The results showed that with the projection in 2030 under the present trend of sea-level <span class="hlt">rise</span> (0.26 cm x a(-<span class="hlt">1</span>)), 6.6% and 0.<span class="hlt">1</span>% of the coastal wetlands were in the low and moderate vulnerabilities, respectively; and in 2050, 9.8% and 0.2% of the coastal wetlands were in low and moderate vulnerabilities, respectively. With the projection in 2030 under the A<span class="hlt">1</span>F<span class="hlt">1</span> scenario (0.59 cm x a(-<span class="hlt">1</span>)), 9.0% and 0.<span class="hlt">1</span>% of the coastal wetlands were in the low and moderate vulnerabilities, respectively; and in 2050, 9.5%, <span class="hlt">1</span>.0% and 0.3% of the coastal wetlands were in the low, moderate and high vulnerabilities, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004OcMod...7..285H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004OcMod...7..285H"><span>Evaluation of <span class="hlt">vertical</span> coordinate and <span class="hlt">vertical</span> mixing algorithms in the HYbrid-Coordinate Ocean Model (HYCOM)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Halliwell, George R.</p> <p></p> <p><span class="hlt">Vertical</span> coordinate and <span class="hlt">vertical</span> mixing algorithms included in the HYbrid Coordinate Ocean Model (HYCOM) are evaluated in low-resolution climatological simulations of the Atlantic Ocean. The hybrid <span class="hlt">vertical</span> coordinates are isopycnic in the deep ocean interior, but smoothly transition to level (pressure) coordinates near the ocean surface, to sigma coordinates in shallow water regions, and back again to level coordinates in very shallow water. By comparing simulations to climatology, the best model performance is realized using hybrid coordinates in conjunction with one of the three available differential <span class="hlt">vertical</span> mixing models: the nonlocal K-Profile Parameterization, the NASA GISS level 2 turbulence closure, and the Mellor-Yamada level 2.5 turbulence closure. Good performance is also achieved using the quasi-slab Price-Weller-Pinkel dynamical instability model. Differences among these simulations are too small relative to other errors and biases to identify the "best" <span class="hlt">vertical</span> mixing model for low-resolution climate simulations. Model performance deteriorates slightly when the Kraus-Turner slab mixed layer model is used with hybrid coordinates. This deterioration is smallest when solar radiation penetrates beneath the mixed layer and when shear instability mixing is included. A simulation performed using isopycnic coordinates to emulate the Miami Isopycnic Coordinate Ocean Model (MICOM), which uses Kraus-Turner mixing without penetrating shortwave radiation and shear instability mixing, demonstrates that the advantages of switching from isopycnic to hybrid coordinates and including more sophisticated turbulence closures outweigh the negative numerical effects of maintaining hybrid <span class="hlt">vertical</span> coordinates.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3774540','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3774540"><span>Zucchini yellow mosaic virus (ZYMV, Potyvirus): <span class="hlt">Vertical</span> transmission, seed infection and cryptic infections</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Simmons, H.E.; Dunham, J.P.; Zinn, K. E.; Munkvold, G.P.; Holmes, E.C.; Stephenson, A.G.</p> <p>2013-01-01</p> <p>The role played by seed transmission in the evolution and epidemiology of viral crop pathogens remains unclear. We determined the seed infection and <span class="hlt">vertical</span> transmission rates of zucchini yellow mosaic virus (ZYMV), in addition to undertaking Illumina sequencing of nine <span class="hlt">vertically</span> transmitted ZYMV populations. We previously determined the seed-to-seedling transmission rate of ZYMV in Cucurbita pepo ssp. texana (a wild gourd) to be <span class="hlt">1</span>.6%, and herein observed a similar rate (<span class="hlt">1</span>.8%) in the subsequent generation. We also observed that the seed infection rate is substantially higher (21.9%) than the seed-to-seedling transmission rate, suggesting that a major population bottleneck occurs during seed germination and seedling growth. In contrast, that two thirds of the variants present in the horizontally transmitted inoculant population were also present in the <span class="hlt">vertically</span> transmitted populations implies that the bottleneck at <span class="hlt">vertical</span> transmission may not be particularly severe. Strikingly, all of the <span class="hlt">vertically</span> infected plants were symptomless in contrast to those infected horizontally, suggesting that <span class="hlt">vertical</span> infection may be cryptic. Although no known virulence determining mutations were observed in the <span class="hlt">vertically</span> infected samples, the 5’ untranslated region was highly variable, with at least 26 different major haplotypes in this region compared to the two major haplotypes observed in the horizontally transmitted population. That the regions necessary for vector transmission are retained in the <span class="hlt">vertically</span> infected populations, combined with the cryptic nature of <span class="hlt">vertical</span> infection, suggests that seed transmission may be a significant contributor to the spread of ZYMV. PMID:23845301</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23845301','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23845301"><span>Zucchini yellow mosaic virus (ZYMV, Potyvirus): <span class="hlt">vertical</span> transmission, seed infection and cryptic infections.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Simmons, H E; Dunham, J P; Zinn, K E; Munkvold, G P; Holmes, E C; Stephenson, A G</p> <p>2013-09-01</p> <p>The role played by seed transmission in the evolution and epidemiology of viral crop pathogens remains unclear. We determined the seed infection and <span class="hlt">vertical</span> transmission rates of zucchini yellow mosaic virus (ZYMV), in addition to undertaking Illumina sequencing of nine <span class="hlt">vertically</span> transmitted ZYMV populations. We previously determined the seed-to-seedling transmission rate of ZYMV in Cucurbita pepo ssp. texana (a wild gourd) to be <span class="hlt">1</span>.6%, and herein observed a similar rate (<span class="hlt">1</span>.8%) in the subsequent generation. We also observed that the seed infection rate is substantially higher (21.9%) than the seed-to-seedling transmission rate, suggesting that a major population bottleneck occurs during seed germination and seedling growth. In contrast, that two thirds of the variants present in the horizontally transmitted inoculant population were also present in the <span class="hlt">vertically</span> transmitted populations implies that the bottleneck at <span class="hlt">vertical</span> transmission may not be particularly severe. Strikingly, all of the <span class="hlt">vertically</span> infected plants were symptomless in contrast to those infected horizontally, suggesting that <span class="hlt">vertical</span> infection may be cryptic. Although no known virulence determining mutations were observed in the <span class="hlt">vertically</span> infected samples, the 5' untranslated region was highly variable, with at least 26 different major haplotypes in this region compared to the two major haplotypes observed in the horizontally transmitted population. That the regions necessary for vector transmission are retained in the <span class="hlt">vertically</span> infected populations, combined with the cryptic nature of <span class="hlt">vertical</span> infection, suggests that seed transmission may be a significant contributor to the spread of ZYMV. Copyright © 2013 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006JGeo...41...39L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006JGeo...41...39L"><span><span class="hlt">Vertical</span> and horizontal seismometric observations of tides</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lambotte, S.; Rivera, L.; Hinderer, J.</p> <p>2006-01-01</p> <p>Tidal signals have been largely studied with gravimeters, strainmeters and tiltmeters, but can also be retrieved from digital records of the output of long-period seismometers, such as STS-<span class="hlt">1</span>, particularly if they are properly isolated. Horizontal components are often noisier than the <span class="hlt">vertical</span> ones, due to sensitivity to tilt at long periods. Hence, horizontal components are often disturbed by local effects such as topography, geology and cavity effects, which imply a strain-tilt coupling. We use series of data (duration larger than <span class="hlt">1</span> month) from several permanent broadband seismological stations to examine these disturbances. We search a minimal set of observable signals (tilts, horizontal and <span class="hlt">vertical</span> displacements, strains, gravity) necessary to reconstruct the seismological record. Such analysis gives a set of coefficients (per component for each studied station), which are stable over years and then can be used systematically to correct data from these disturbances without needing heavy numerical computation. A special attention is devoted to ocean loading for stations close to oceans (e.g. Matsushiro station in Japon (MAJO)), and to pressure correction when barometric data are available. Interesting observations are made for <span class="hlt">vertical</span> seismometric components; in particular, we found a pressure admittance between pressure and data 10 times larger than for gravimeters for periods larger than <span class="hlt">1</span> day, while this admittance reaches the usual value of -3.5 nm/s 2/mbar for periods below 3 h. This observation may be due to instrumental noise, but the exact mechanism is not yet understood.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_19 --> <div id="page_20" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="381"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..18.9403B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..18.9403B"><span><span class="hlt">Vertical</span> microbial community variability of carbonate-based cones may provide insight into ancient conical stromatolite formation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bradley, James; Daille, Leslie; Trivedi, Christopher; Bojanowski, Caitlin; Nunn, Heather; Stamps, Blake; Johnson, Hope; Stevenson, Bradley; Berelson, Will; Corsetti, Frank; Spear, John</p> <p>2016-04-01</p> <p>Stromatolite morphogenesis is poorly understood, and the process by which microbial mats become mineralized is a primary question in microbialite formation. Ancient conical stromatolites are primarily carbonate-based whereas the few modern analogues in hot springs are either non-mineralized or mineralized by silica. A team from the 2015 International GeoBiology Course investigated carbonate-rich microbial cones from near Little Hot Creek (LHC), Long Valley Caldera, California, to investigate how conical stromatolites might form in a hot spring carbonate system. The cones <span class="hlt">rise</span> up from a layered microbial mat on the east side of a 45° C pool with very low flow that is super-saturated with respect to CaCO3. Cone structures are 8-30 mm in height, are rigid and do not deform when removed from the pool. Morphological characterization through environmental scanning electronic microscopy revealed that the cone structure is maintained by a matrix of intertwining microbial filaments around carbonate grains. This matrix gives <span class="hlt">rise</span> to cone-filaments that are arranged <span class="hlt">vertically</span> or horizontally, and provides further stability to the cone. Preliminary 16S rRNA gene analysis indicated variability of community composition between different <span class="hlt">vertical</span> levels of the cone. The cone tip had comparatively greater abundance of filamentous cyanobacteria including Leptolingbya, Phormidium and Isosphaera and fewer heterotrophs (e.g. Chloroflexi) compared to the cone bottom. This supports the hypothesis that cone formation may depend on the differential abundance of the microbial community and their potential functional roles. Metagenomic analyses of the cones revealed potential genes related to chemotaxis and motility. Specifically, a genomic bin identified as a member of the genus Isosphaera contained an hmp chemotaxis operon implicated in gliding motility in the cyanobacterium Nostoc punctiforme. Isosphaera is a Planctomycete shown to have phototactic capabilities, and may play a role in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/20630567-rise-time-simulated-veritas-davies-cotton-reflector','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/20630567-rise-time-simulated-veritas-davies-cotton-reflector"><span><span class="hlt">Rise</span> Time of the Simulated VERITAS 12 m Davies-Cotton Reflector</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>White, Richard J.</p> <p></p> <p>The Very Energetic Radiation Imaging Telescope Array System (VERITAS) will utilise Imaging Atmospheric Cherenkov Telescopes (IACTs) based on a Davies-Cotton design with f-number f/<span class="hlt">1</span>.0 to detect cosmic gamma-rays. Unlike a parabolic reflector, light from the Davies-Cotton does not arrive isochronously at the camera. Here the effect of the telescope geometry on signal <span class="hlt">rise</span>-time is examined. An almost square-pulse arrival time profile with a <span class="hlt">rise</span> time of <span class="hlt">1</span>.7 ns is found analytically and confirmed through simulation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.8814S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.8814S"><span>GGOS Focus Area 3: Understanding and Forecasting Sea-Level <span class="hlt">Rise</span> and Variability</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schöne, Tilo; Shum, Ck; Tamisiea, Mark; Woodworth, Philip</p> <p>2017-04-01</p> <p>Sea level and its change have been measured for more than a century. Especially for coastal nations, deltaic regions, and coastal-oriented industries, observations of tides, tidal extremes, storm surges, and sea level <span class="hlt">rise</span> at the interannual or longer scales have substantial impacts on coastal vulnerability towards resilience and sustainability of world's coastal regions. To date, the observed global sea level <span class="hlt">rise</span> is largely associated with climate related changes. To find the patterns and fingerprints of those changes, and to e.g., separate the land motion from sea level signals, different monitoring techniques have been developed. Some of them are local, e.g., tide gauges, while others are global, e.g., satellite altimetry. It is well known that sea level change and land <span class="hlt">vertical</span> motion varies regionally, and both signals need to be measured in order to quantify relative sea level at the local scale. The Global Geodetic Observing System (GGOS) and its services contribute in many ways to the monitoring of the sea level. These includes tide gauge observations, estimation of gravity changes, satellite altimetry, InSAR/Lidar, GNSS-control of tide gauges, providing ground truth sites for satellite altimetry, and importantly the maintenance of the International Reference Frame. Focus Area 3 (Understanding and Forecasting Sea-Level <span class="hlt">Rise</span> and Variability) of GGOS establishes a platform and a forum for researchers and authorities dealing with estimating global and local sea level changes in a 10- to 30-year time span, and its project to the next century or beyond. It presents an excellent opportunity to emphasize the global, through to regional and local, importance of GGOS to a wide range of sea-level related science and practical applications. Focus Area 3 works trough demonstration projects to highlight the value of geodetic techniques to sea level science and applications. Contributions under a call for participation (http://www.ggos.org/Applications/theme3_SL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPC.1826b0034P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1826b0034P"><span>Effect of oxygen concentration on fire growth of various types of cable bending in horizontal and <span class="hlt">vertical</span> orientations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pangaribuan, Adrianus; Dhiputra, I. M. K.; Nugroho, Yulianto S.</p> <p>2017-03-01</p> <p>Electrical cable is a whole of the material including metal (cooper) conductor and its insulation, when an electrical cable is flowed by electric current, based on its own capacity, the temperature of cable conductor increases gradually. If the current flows above the cable carrying capacity, then an extreme temperature <span class="hlt">rises</span> are expected. When temperature increase, the electric current flow inside cable conductor will decrease gradually related to the resistance and could occur repeatedly in a period. Since electrical faults on electrical cable system are often suspected as the cause of fires, thus this research aims to investigate measures of preventing the fire to start by means of controlling oxygen concentration in a cable compartment. The experimental work was conducted in laboratory by using electrical power cable of <span class="hlt">1</span>.5 mm2 size. Two transparent chambers were applied for studying the effect of <span class="hlt">vertical</span> and horizontal orientations on the cable temperature <span class="hlt">rise</span>, under various oxygen concentration of the gas streams. In the present work, the electrical was maintained at a constant level during a typical test run. Parametric studies reported in the paper include the use of a bare and insulated cables as well as the bending shape of the cable lines of a straight cable, coiled cable and randomly bent cable which were loaded with the same electric load and oxygen concentration in the gas supply.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/20933306-classical-matrix-su-vertical-bar-super-yang-mills-spin-chain','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/20933306-classical-matrix-su-vertical-bar-super-yang-mills-spin-chain"><span>Classical r matrix of the su(2 <span class="hlt">vertical</span> bar 2) super Yang-Mills spin chain</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Torrielli, Alessandro</p> <p>2007-05-15</p> <p>In this note we straightforwardly derive and make use of the quantum R matrix for the su(2 <span class="hlt">vertical</span> bar 2) super Yang-Mills spin chain in the manifest su(<span class="hlt">1</span> <span class="hlt">vertical</span> bar 2)-invariant formulation, which solves the standard quantum Yang-Baxter equation, in order to obtain the correspondent (undressed) classical r matrix from the first order expansion in the 'deformation' parameter 2{pi}/{radical}({lambda}) and check that this last solves the standard classical Yang-Baxter equation. We analyze its bialgebra structure, its dependence on the spectral parameters, and its pole structure. We notice that it still preserves an su(<span class="hlt">1</span> <span class="hlt">vertical</span> bar 2) subalgebra, thereby admitting anmore » expression in terms of a combination of projectors, which spans only a subspace of su(<span class="hlt">1</span> <span class="hlt">vertical</span> bar 2)xsu(<span class="hlt">1</span> <span class="hlt">vertical</span> bar 2). We study the residue at its simple pole at the origin and comment on the applicability of the classical Belavin-Drinfeld type of analysis.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19760009699','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19760009699"><span>Brainstem auditory evoked responses in man. <span class="hlt">1</span>: Effect of stimulus <span class="hlt">rise</span>-fall time and duration</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hecox, K.; Squires, N.; Galambos, R.</p> <p>1975-01-01</p> <p>Short latency (under 10 msec) evoked responses elicited by bursts of white noise were recorded from the scalp of human subjects. Response alterations produced by changes in the noise burst duration (on-time) inter-burst interval (off-time), and onset and offset shapes are reported and evaluated. The latency of the most prominent response component, wave V, was markedly delayed with increases in stimulus <span class="hlt">rise</span>-time but was unaffected by changes in fall-time. The amplitude of wave V was insensitive to changes in signal <span class="hlt">rise</span>-and-fall times, while increasing signal on-time produced smaller amplitude responses only for sufficiently short off-times. It is concluded that wave V of the human auditory brainstem evoked response is solely an onset response.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5523029-equatorial-waves-stratospheric-gcm-effects-vertical-resolution-gcm-general-circulation-model','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5523029-equatorial-waves-stratospheric-gcm-effects-vertical-resolution-gcm-general-circulation-model"><span>Equatorial waves in a stratospheric GCM: Effects of <span class="hlt">vertical</span> resolution. [GCM (general circulation model)</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Boville, B.A.; Randel, W.J.</p> <p>1992-05-01</p> <p>Equatorially trapped wave modes, such as Kelvin and mixed Rossby-gravity waves, are believed to play a crucial role in forcing the quasi-biennial oscillation (QBO) of the lower tropical stratosphere. This study examines the ability of a general circulation model (GCM) to simulate these waves and investigates the changes in the wave properties as a function of the <span class="hlt">vertical</span> resolution of the model. The simulations produce a stratopause-level semiannual oscillation but not a QBO. An unfortunate property of the equatorially trapped waves is that they tend to have small <span class="hlt">vertical</span> wavelengths ([le] 15 km). Some of the waves, believed to bemore » important in forcing the QBO, have wavelengths as short as 4 km. The short <span class="hlt">vertical</span> wavelengths pose a stringent computational requirement for numerical models whose <span class="hlt">vertical</span> grid spacing is typically chosen based on the requirements for simulating extratropical Rossby waves (which have much longer <span class="hlt">vertical</span> wavelengths). This study examines the dependence of the equatorial wave simulation of <span class="hlt">vertical</span> resolution using three experiments with <span class="hlt">vertical</span> grid spacings of approximately 2.8, <span class="hlt">1</span>.4, and 0.7 km. Several Kelvin, mixed Rossby-gravity, and 0.7 km. Several Kelvin, mixed Rossby-gravity, and inertio-gravity waves are identified in the simulations. At high <span class="hlt">vertical</span> resolution, the simulated waves are shown to correspond fairly well to the available observations. The properties of the relatively slow (and <span class="hlt">vertically</span> short) waves believed to play a role in the QBO vary significantly with <span class="hlt">vertical</span> resolution. <span class="hlt">Vertical</span> grid spacings of about <span class="hlt">1</span> km or less appear to be required to represent these waves adequately. The simulated wave amplitudes are at least as large as observed, and the waves are absorbed in the lower stratosphere, as required in order to force the QBO. However, the EP flux divergence associated with the waves is not sufficient to explain the zonal flow accelerations found in the QBO. 39 refs., 17 figs., <span class="hlt">1</span> tab.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010EGUGA..12.5480R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010EGUGA..12.5480R"><span>Understanding the Effects of Sea-Level <span class="hlt">Rise</span> on Coastal Wetlands: The Human Dimension</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Reed, Denise</p> <p>2010-05-01</p> <p> factors, and plant growth is an additional factor influencing the survival of more organic marshes. Salt marsh surfaces are frequently considered to be in an equilibrium relationship with local mean sea level but the projection of salt marsh sustainability under future climate scenarios is a complex issue and depends on: the relative importance of organic matter to marsh <span class="hlt">vertical</span> development; the complexities governing organic matter accumulation during <span class="hlt">rising</span> sea level; the importance of subsurface processes in determining surface elevation change; and the role of storm events and hydrologic changes in controlling sediment deposition, soil conditions and plant growth. The effects of global change, both climate and human induced, on coastal wetlands will be manifest differently among various geomorphic settings but their vulnerability to global change in the 21st century should be taken seriously by coastal managers and policy-makers alike.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title45-vol1/pdf/CFR-2011-title45-vol1-sec77-3.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title45-vol1/pdf/CFR-2011-title45-vol1-sec77-3.pdf"><span>45 CFR 77.3 - Conditions that may give <span class="hlt">rise</span> to remedial actions.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-10-01</p> <p>... 45 Public Welfare <span class="hlt">1</span> 2011-10-01 2011-10-01 false Conditions that may give <span class="hlt">rise</span> to remedial actions. 77.3 Section 77.3 Public Welfare DEPARTMENT OF HEALTH AND HUMAN SERVICES GENERAL ADMINISTRATION REMEDIAL ACTIONS APPLICABLE TO LETTER OF CREDIT ADMINISTRATION § 77.3 Conditions that may give <span class="hlt">rise</span> to...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title45-vol1/pdf/CFR-2010-title45-vol1-sec77-3.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title45-vol1/pdf/CFR-2010-title45-vol1-sec77-3.pdf"><span>45 CFR 77.3 - Conditions that may give <span class="hlt">rise</span> to remedial actions.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-10-01</p> <p>... 45 Public Welfare <span class="hlt">1</span> 2010-10-01 2010-10-01 false Conditions that may give <span class="hlt">rise</span> to remedial actions. 77.3 Section 77.3 Public Welfare DEPARTMENT OF HEALTH AND HUMAN SERVICES GENERAL ADMINISTRATION REMEDIAL ACTIONS APPLICABLE TO LETTER OF CREDIT ADMINISTRATION § 77.3 Conditions that may give <span class="hlt">rise</span> to...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMNH34B..06M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMNH34B..06M"><span>The impact of local land subsidence and global sea level <span class="hlt">rise</span> on flood severity in Houston-Galveston caused by Hurricane Harvey</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Miller, M. M.; Shirzaei, M.</p> <p>2017-12-01</p> <p>Category-4 Hurricane Harvey had devastating socioeconomic impacts to Houston, with flooding far past the 100-year flood zones published by FEMA. In recent decades, frequency and intensity of coastal flooding are escalating, correlated with sea level <span class="hlt">rise</span> (SLR). Moreover, Local land subsidence (LLS) due to groundwater and hydrocarbon extraction and natural compaction changes surface elevation and slope, potentially altering drainage patterns. GPS data show a mm broad co-cyclonic subsidence due to elastic loading from the water mass measured by GPS, which is inverted to solve for the total fluid volume of 2.73x1010 m3. We additionally investigate the joint impact of an SLR and pre-cyclonic LLS on the flooding of Houston-Galveston during Hurricane Harvey. We examine <span class="hlt">vertical</span> land motion within North American <span class="hlt">Vertical</span> Datum 2012 for the period 2007 until the cyclone by investigating SAR imaged acquired by ALOS and Sentinel-<span class="hlt">1</span>A/B radar satellites combined with GPS data. We find patchy, LLS bowls resulting in sinks where floodwater can collect. We map the flooding extent by comparing amplitudes of Sentinal<span class="hlt">1</span>-A/B pixels' backscattered radar signal from pre- and post-Harvey acquisitions and estimate 782 km2 are submerged within the area of 3478 km2 of pixels covered by Sentinel frame. Comparing with the LLS map, 89% of the flooded pixels exhibit -3 mm/yr or greater <span class="hlt">vertical</span> motion. Flooding attributed to the storm surge is determined with high-resolution LiDAR digital elevation models (DEM) and a 0.75 m storm tide inundation model, which engulfs only 195 km2 and nearby the shorelines. We estimate future inundation hazard by combining LiDAR DEMs with our InSAR derived subsidence map, projecting LLS rates forward 100 years, and modeling projected SLR from 0.4 to <span class="hlt">1</span>.2 meters. Were subsidence to continue unabated, the total flooded area is 281 km2 with a 0.4 m and 394 km2 with a <span class="hlt">1</span>.2 m SLR. Next, we add a modest storm tide (0.752 m), which increases the flooded area to 389 - 480</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA619945','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA619945"><span>Controlled Synthesis and Functionalization of <span class="hlt">Vertically</span>-Aligned Carbon Nanotubes for Multifunctional Applications</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2015-05-07</p> <p>6 <span class="hlt">1</span>.6 Lithium - Ion Batteries Based on <span class="hlt">Vertically</span>-Aligned Carbon Nanotube Electrodes and Ionic...Cl, Br, or I) Prepared by Ball-Milling and Used as Anode Materials for Lithium - Ion Batteries ……………....................23 3.4 Well-Defined Two...9 <span class="hlt">1</span>.6 Lithium - Ion Batteries Based on <span class="hlt">Vertically</span>-Aligned Carbon Nanotube Electrodes and Ionic Liquid Electrolytes</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004JFM...504..229G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004JFM...504..229G"><span><span class="hlt">Vertical</span> length scale selection for pancake vortices in strongly stratified viscous fluids</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Godoy-Diana, Ramiro; Chomaz, Jean-Marc; Billant, Paul</p> <p>2004-04-01</p> <p>The evolution of pancake dipoles of different aspect ratio is studied in a stratified tank experiment. Two cases are reported here for values of the dipole initial aspect ratio alpha_0 = L_v/L_h (where L_v and L_h are <span class="hlt">vertical</span> and horizontal length scales, respectively) of alpha_0 = 0.4 (case I) and alpha_0 = <span class="hlt">1</span>.2 (case II). In the first case, the usual decay scenario is observed where the dipole diffuses slowly with a growing thickness and a decaying circulation. In case II, we observed a regime where the thickness of the dipole decreases and the circulation in the horizontal mid-plane of the vortices remains constant. We show that this regime where the <span class="hlt">vertical</span> length scale decreases can be explained by the shedding of two boundary layers at the top and bottom of the dipole that literally peel off vorticity layers. Horizontal advection and <span class="hlt">vertical</span> diffusion cooperate in this regime and the decrease towards the viscous <span class="hlt">vertical</span> length scale delta = L_hRe(-<span class="hlt">1</span>/2) occurs on a time scale alpha_0 Re(<span class="hlt">1</span>/2) T_A, T_A being the advection time L_h/U. From a scaling analysis of the equations for a stratified viscous fluid in the Boussinesq approximation, two dominant balances depending on the parameter R = ReF_h(2) are discussed, where F_h = U/NL_h is the horizontal Froude number and Re = UL_h/nu is the Reynolds number, U, N and nu being, respectively, the translation speed of the dipole, the Brunt Väisälä frequency and the kinematic viscosity. When R≫ <span class="hlt">1</span> the <span class="hlt">vertical</span> length scale is determined by buoyancy effects to be of order L_b = U/N. The experiments presented in this paper pertain to the case of small R, where viscous effects govern the selection of the <span class="hlt">vertical</span> length scale. We show that if initially L_v ≤ delta, the flow diffuses on the <span class="hlt">vertical</span> (case I), while if L_v ≫ delta (case II), <span class="hlt">vertically</span> sheared horizontal advection decreases the <span class="hlt">vertical</span> length scale down to delta. This viscous regime may explain results from experiments and numerical simulations on</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21330225','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21330225"><span>Rumination syndrome: when the lower oesophageal sphincter <span class="hlt">rises</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gourcerol, Guillaume; Dechelotte, Pierre; Ducrotte, Philippe; Leroi, Anne Marie</p> <p>2011-07-01</p> <p>Rumination syndrome is an uncommon condition characterised by the self-induced regurgitation from the stomach to the mouth of recently ingested meal that is chewed and reswallowed. Rumination is caused by a voluntary <span class="hlt">rise</span> in intra-abdominal and intra-gastric pressure leading to the reflux of the gastric content into the oesophagus. However, the precise mechanisms preventing reflux at the gastro-oesophageal junction during the <span class="hlt">rise</span> in intra-gastric pressure remains unknown. In 5 patients, rumination episodes were monitored using combined multiple intra-luminal impedance monitoring, high resolution manometry, and video-fluoroscopic recording. We showed that the gastro-oesophageal junction moved from the abdominal cavity into the thorax creating a "pseudo-hernia". This occurred at a range of <span class="hlt">1</span>.4 ± 0.3 s before the <span class="hlt">rise</span> in intra-oesophageal pressure and the gastro-oesophageal reflux. This displacement of the gastro-oesophageal junction into thorax, rather than a lower oesophageal sphincter opening, explains the mechanism of voluntary regurgitations occurring during rumination syndrome. Copyright © 2011 Editrice Gastroenterologica Italiana S.r.l. Published by Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15252352','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15252352"><span>Reading speed benefits from increased <span class="hlt">vertical</span> word spacing in normal peripheral vision.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chung, Susana T L</p> <p>2004-07-01</p> <p>Crowding, the adverse spatial interaction due to proximity of adjacent targets, has been suggested as an explanation for slow reading in peripheral vision. The purposes of this study were to (<span class="hlt">1</span>) demonstrate that crowding exists at the word level and (2) examine whether or not reading speed in central and peripheral vision can be enhanced with increased <span class="hlt">vertical</span> word spacing. Five normal observers read aloud sequences of six unrelated four-letter words presented on a computer monitor, one word at a time, using rapid serial visual presentation (RSVP). Reading speeds were calculated based on the RSVP exposure durations yielding 80% correct. Testing was conducted at the fovea and at 5 degrees and 10 degrees in the inferior visual field. Critical print size (CPS) for each observer and at each eccentricity was first determined by measuring reading speeds for four print sizes using unflanked words. We then presented words at 0.8x or <span class="hlt">1</span>.4x CPS, with each target word flanked by two other words, one above and one below the target word. Reading speeds were determined for <span class="hlt">vertical</span> word spacings (baseline-to-baseline separation between two <span class="hlt">vertically</span> separated words) ranging from 0.8x to 2x the standard single-spacing, as well as the unflanked condition. At the fovea, reading speed increased with <span class="hlt">vertical</span> word spacing up to about <span class="hlt">1</span>.2x to <span class="hlt">1</span>.5x the standard spacing and remained constant and similar to the unflanked reading speed at larger <span class="hlt">vertical</span> word spacings. In the periphery, reading speed also increased with <span class="hlt">vertical</span> word spacing, but it remained below the unflanked reading speed for all spacings tested. At 2x the standard spacing, peripheral reading speed was still about 25% lower than the unflanked reading speed for both eccentricities and print sizes. Results from a control experiment showed that the greater reliance of peripheral reading speed on <span class="hlt">vertical</span> word spacing was also found in the right visual field. Increased <span class="hlt">vertical</span> word spacing, which presumably decreases the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70171365','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70171365"><span>Characterizing seasonal and diel <span class="hlt">vertical</span> movement and habitat use of lake whitefish (Coregonus clupeaformis) in Clear Lake, Maine</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Zydlewski, Joseph D.; Gorsky, Dimitry; Balsey, David</p> <p>2016-01-01</p> <p>Seasonal and daily <span class="hlt">vertical</span> activity of lake whitefish Coregonus clupeaformis was studied in Clear Lake, Maine (253 ha), using acoustic telemetry from November 2004 to June 2009. Twenty adult lake whitefish were tagged with acoustic tags that had either a depth sensor or both depth and temperature sensors to assess <span class="hlt">vertical</span> habitat use at a seasonal and daily resolution. <span class="hlt">Vertical</span> habitat selection varied seasonally and was strongly influenced by temperature. Between December and April, when the lake was covered with ice, surface temperature was below 2°C and tagged individuals occupied deep areas of the lake (∼15 m). After ice-out, fish ascended into shallow waters (∼5 m), responding to increased water temperature and possibly to greater foraging opportunity. When surface water temperatures exceeded 20°C, fish descended below the developing thermocline (∼9 m), where they remained until surface temperatures fell below 20°C; fish then ascended into shallower depths, presumably for feeding and spawning. Through the winter, fish remained in thermal habitats that were warmer than the surface temperatures; in the summer, they selected depths with thermal habitats below 15°C. Though the amplitude varied greatly across seasons, lake whitefish displayed a strong diurnal pattern of activity as measured by <span class="hlt">vertical</span> velocities. Fish were twofold more active during spring, summer, and fall than during winter. Lake whitefish exhibited diel <span class="hlt">vertical</span> migrations, <span class="hlt">rising</span> in the water column during nighttime and occupying deeper waters during the day. This pattern was more pronounced in the spring and fall and far less prominent during winter and summer. The strong linkage between temperature and habitat use may limit the current range of lake whitefish and may be directly impacted by climatic change.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70043010','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70043010"><span><span class="hlt">Rising</span> sea level may cause decline of fringing coral reefs</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Field, Michael E.; Ogston, Andrea S.; Storlazzi, Curt D.</p> <p>2011-01-01</p> <p>Coral reefs are major marine ecosystems and critical resources for marine diversity and fisheries. These ecosystems are widely recognized to be at risk from a number of stressors, and added to those in the past several decades is climate change due to anthropogenically driven increases in atmospheric concentrations of greenhouse gases. Most threatening to most coral reefs are elevated sea surface temperatures and increased ocean acidity [e.g., Kleypas et al., 1999; Hoegh-Guldberg et al., 2007], but sea level <span class="hlt">rise</span>, another consequence of climate change, is also likely to increase sedimentary processes that potentially interfere with photosynthesis, feeding, recruitment, and other key physiological processes (Figure <span class="hlt">1</span>). Anderson et al. [2010] argue compellingly that potential hazardous impacts to coastlines from 21st-century sea level <span class="hlt">rise</span> are greatly underestimated, particularly because of the rapid rate of <span class="hlt">rise</span>. The Intergovernmental Panel on Climate Change estimates that sea level will <span class="hlt">rise</span> in the coming century (1990–2090) by 2.2–4.4 millimeters per year, when projected with little contribution from melting ice [Meehl et al., 2007]. New studies indicate that rapid melting of land ice could substantially increase the rate of sea level <span class="hlt">rise</span> [Grinsted et al., 2009; Milne et al., 2009].</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009PMB....54.3393L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009PMB....54.3393L"><span>Assessment of the computational uncertainty of temperature <span class="hlt">rise</span> and SAR in the eyes and brain under far-field exposure from <span class="hlt">1</span> to 10 GHz</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Laakso, Ilkka</p> <p>2009-06-01</p> <p>This paper presents finite-difference time-domain (FDTD) calculations of specific absorption rate (SAR) values in the head under plane-wave exposure from <span class="hlt">1</span> to 10 GHz using a resolution of 0.5 mm in adult male and female voxel models. Temperature <span class="hlt">rise</span> due to the power absorption is calculated by the bioheat equation using a multigrid method solver. The computational accuracy is investigated by repeating the calculations with resolutions of <span class="hlt">1</span> mm and 2 mm and comparing the results. Cubically averaged 10 g SAR in the eyes and brain and eye-averaged SAR are calculated and compared to the corresponding temperature <span class="hlt">rise</span> as well as the recommended limits for exposure. The results suggest that 2 mm resolution should only be used for frequencies smaller than 2.5 GHz, and <span class="hlt">1</span> mm resolution only under 5 GHz. Morphological differences in models seemed to be an important cause of variation: differences in results between the two different models were usually larger than the computational error due to the grid resolution, and larger than the difference between the results for open and closed eyes. Limiting the incident plane-wave power density to smaller than 100 W m-2 was sufficient for ensuring that the temperature <span class="hlt">rise</span> in the eyes and brain were less than <span class="hlt">1</span> °C in the whole frequency range.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28118306','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28118306"><span>Acute Effects of Stretching on Leg and <span class="hlt">Vertical</span> Stiffness During Treadmill Running.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pappas, Panagiotis T; Paradisis, Giorgos P; Exell, Timothy A; Smirniotou, Athanasia S; Tsolakis, Charilaos K; Arampatzis, Adamantios</p> <p>2017-12-01</p> <p>Pappas, PT, Paradisis, GP, Exell, TA, Smirniotou, AS, Tsolakis, CK, and Arampatzis, A. Acute effects of stretching on leg and <span class="hlt">vertical</span> stiffness during treadmill running. J Strength Cond Res 31(12): 3417-3424, 2017-The implementation of static (SS) and dynamic (DS) stretching during warm-up routines produces significant changes in biological and functional properties of the human musculoskeletal system. These properties could affect the leg and <span class="hlt">vertical</span> stiffness characteristics that are considered important factors for the success of athletic activities. The aim of this study was to investigate the influence of SS and DS on selected kinematic variables, and leg and <span class="hlt">vertical</span> stiffness during treadmill running. Fourteen men (age: 22.58 ± <span class="hlt">1</span>.05 years, height: <span class="hlt">1</span>.77 ± 0.05 m, body mass: 72.74 ± 10.04 kg) performed 30-second running bouts at 4.44 m·s, under 3 different stretching conditions (SS, DS, and no stretching). The total duration in each stretching condition was 6 minutes, and each of the 4 muscle groups was stretched for 40 seconds. Leg and <span class="hlt">vertical</span> stiffness values were calculated using the "sine wave" method, with no significant differences in stiffness found between stretching conditions. After DS, <span class="hlt">vertical</span> ground reaction force increased by <span class="hlt">1</span>.7% (p < 0.05), which resulted in significant (p < 0.05) increases in flight time (5.8%), step length (2.2%), and <span class="hlt">vertical</span> displacement of the center of mass (4.5%) and a decrease in step rate (2.2%). Practical durations of SS and DS stretching did not influence leg or <span class="hlt">vertical</span> stiffness during treadmill running. However, DS seems to result in a small increase in lower-limb force production which may influence running mechanics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28874485','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28874485"><span>A Physician's Perspective On <span class="hlt">Vertical</span> Integration.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Berenson, Robert A</p> <p>2017-09-01</p> <p><span class="hlt">Vertical</span> integration has been a central feature of health care delivery system change for more than two decades. Recent studies have demonstrated that <span class="hlt">vertically</span> integrated health care systems raise prices and costs without observable improvements in quality, despite many theoretical reasons why cost control and improved quality might occur. Less well studied is how physicians view their newfound partnerships with hospitals. In this article I review literature findings and other observations on five aspects of <span class="hlt">vertical</span> integration that affect physicians in their professional and personal lives: patients' access to physicians, physician compensation, autonomy versus system support, medical professionalism and culture, and lifestyle. I conclude that the movement toward physicians' alignment with and employment in <span class="hlt">vertically</span> integrated systems seems inexorable but that policy should not promote such integration either intentionally or inadvertently. Instead, policy should address the flaws in current payment approaches that reward high prices and excessive service use-outcomes that <span class="hlt">vertical</span> integration currently produces. Project HOPE—The People-to-People Health Foundation, Inc.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_20 --> <div id="page_21" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="401"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017E%26ES..100a2070Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017E%26ES..100a2070Y"><span>Different stages and status of <span class="hlt">vertical</span> transporting process of Cu in Jiaozhou Bay</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Dongfang; Li, Haixia; Wang, Qi; Zhang, Xiaolong; Ding, Jun</p> <p>2017-12-01</p> <p>Understanding the stages and status of <span class="hlt">vertical</span> transporting process of pollutant in marine bay is essential to pollution control. This paper analyzed the stages and status of Cu’s <span class="hlt">vertical</span> transporting process in waters in Jiaozhou Bay. Results showed that the <span class="hlt">vertical</span> transporting process in waters in Jiaozhou Bay included four stages of <span class="hlt">1</span>) Cu was imported to the bay by major sources, 2) Cu was transported to surface waters, 3) Cu was transported from surface waters to sediment in sea bottom, and 4) Cu was fixed and buried in sediment. Furthermore, Cu’s <span class="hlt">vertical</span> transporting process could be divided into seven status in detail, and he characteristics of the <span class="hlt">vertical</span> transport process of Cu were also analyzed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.A53D0207S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.A53D0207S"><span>Short Lived Climate Pollutants cause a Long Lived Effect on Sea-level <span class="hlt">Rise</span>: Analyzing climate metrics for sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sterner, E.; Johansson, D. J.</p> <p>2013-12-01</p> <p>Climate change depends on the increase of several different atmospheric pollutants. While long term global warming will be determined mainly by carbon dioxide, warming in the next few decades will depend to a large extent on short lived climate pollutants (SLCP). Reducing emissions of SLCPs could contribute to lower the global mean surface temperature by 0.5 °C already by 2050 (Shindell et al. 2012). Furthermore, the warming effect of one of the most potent SLCPs, black carbon (BC), may have been underestimated in the past. Bond et al. (2013) presents a new best estimate of the total BC radiative forcing (RF) of <span class="hlt">1.1</span> W/m2 (90 % uncertainty bounds of 0.17 to 2.<span class="hlt">1</span> W/m2) since the beginning of the industrial era. BC is however never emitted alone and cooling aerosols from the same sources offset a majority of this RF. In the wake of calls for mitigation of SLCPs it is important to study other aspects of the climate effect of SLCPs. One key impact of climate change is sea-level <span class="hlt">rise</span> (SLR). In a recent study, the effect of SLCP mitigation scenarios on SLR is examined. Hu et al (2013) find a substantial effect on SLR from mitigating SLCPs sharply, reducing SLR by 22-42% by 2100. We choose a different approach focusing on emission pulses and analyse a metric based on sea level <span class="hlt">rise</span> so as to further enlighten the SLR consequences of SLCPs. We want in particular to understand the time dynamics of SLR impacts caused by SLCPs compared to other greenhouse gases. The most commonly used physical based metrics are GWP and GTP. We propose and evaluate an additional metric: The global sea-level <span class="hlt">rise</span> potential (GSP). The GSP is defined as the sea level <span class="hlt">rise</span> after a time horizon caused by an emissions pulse of a forcer to the sea level <span class="hlt">rise</span> after a time horizon caused by an emissions pulse of a CO2. GSP is evaluated and compared to GWP and GTP using a set of climate forcers chosen to cover the whole scale of atmospheric perturbation life times (BC, CH4, N2O, CO2 and SF6). The study</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21768892','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21768892"><span>Effect of hang cleans or squats paired with countermovement <span class="hlt">vertical</span> jumps on <span class="hlt">vertical</span> displacement.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Andrews, Tedi R; Mackey, Theresa; Inkrott, Thomas A; Murray, Steven R; Clark, Ida E; Pettitt, Robert W</p> <p>2011-09-01</p> <p>Complex training is characterized by pairing resistance exercise with plyometric exercise to exploit the postactivation potentiation (PAP) phenomenon, thereby promising a better training effect. Studies on PAP as measured by human power performances are equivocal. One issue may be the lack of analyses across multiple sets of paired exercises, a common practice used by athletes. We evaluated countermovement <span class="hlt">vertical</span> jump (CMJ) performance in 19 women, collegiate athletes in 3 of the following trials: (a) CMJs-only, where <span class="hlt">1</span> set of CMJs served as a conditioning exercise, (b) heavy-load, back squats paired with CMJs, and (c) hang cleans paired with CMJs. The CMJ <span class="hlt">vertical</span> displacement (3-attempt average), as measured with digital video, served as the dependent variable of CMJ performance. Across 3 sets of paired-exercise regimens, CMJ-only depreciated <span class="hlt">1</span>.6 cm and CMJ paired with back squats depreciated 2.0 cm (main effect, p < 0.05). Conversely, CMJ paired with hang cleans depreciated 0.30 cm (interaction, p < 0.05). Thus, the best complex training scheme was achieved by pairing CMJs with hang cleans in comparison to back squats or CMJs in and of themselves. Future research on exercise modes of complex training that best help athletes preserve and train with the highest power possible, in a given training session, is warranted.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-201103190002HQ.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-201103190002HQ.html"><span>Super Moon <span class="hlt">Rises</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-03-19</p> <p>The full moon is seen as it <span class="hlt">rises</span> near the National Mall, Saturday, March 19, 2011, in Washington. The full moon tonight is called a "Super Moon" since it is at its closest to Earth. Photo Credit: (NASA/Paul E. Alers)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017E%26ES...95e2003B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017E%26ES...95e2003B"><span>Temperature and Relative Humidity <span class="hlt">Vertical</span> Profiles within Planetary Boundary Layer in Winter Urban Airshed</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bendl, Jan; Hovorka, Jan</p> <p>2017-12-01</p> <p>The planetary boundary layer is a dynamic system with turbulent flow where horizontal and <span class="hlt">vertical</span> air mixing depends mainly on the weather conditions and geomorphology. Normally, air temperature from the Earth surface decreases with height but inversion situation may occur, mainly during winter. Pollutant dispersion is poor during inversions so air pollutant concentration can quickly <span class="hlt">rise</span>, especially in urban closed valleys. Air pollution was evaluated by WHO as a human carcinogen (mostly by polycyclic aromatic hydrocarbons) and health effects are obvious. Knowledge about inversion layer height is important for estimation of the pollution impact and it can give us also information about the air pollution sources. Temperature and relative humidity <span class="hlt">vertical</span> profiles complement ground measurements. Ground measurements were conducted to characterize comprehensively urban airshed in Svermov, residential district of the city of Kladno, about 30 km NW of Prague, from the 2nd Feb. to the 3rd of March 2016. The Svermov is an air pollution hot-spot for long time benzo[a]pyrene (B[a]P) limit exceedances, reaching the highest B[a]P annual concentration in Bohemia - west part of the Czech Republic. Since the Svermov sits in a shallow valley, frequent <span class="hlt">vertical</span> temperature inversion in winter and low emission heights of pollution sources prevent pollutant dispersal off the valley. Such orography is common to numerous small settlements in the Czech Republic. Ground measurements at the sports field in the Svermov were complemented by temperature and humidity <span class="hlt">vertical</span> profiles acquired by a Vaisala radiosonde positioned at tethered He-filled balloon. Total number of 53 series of <span class="hlt">vertical</span> profiles up to the height of 300 m was conducted. Meteorology parameters were acquired with 4 Hz frequency. The measurements confirmed frequent early-morning and night formation of temperature inversion within boundary layer up to the height of 50 m. This rather shallow inversion had significant</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AdSpR..61.1628S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AdSpR..61.1628S"><span>Post-midnight equatorial irregularity distributions and <span class="hlt">vertical</span> drift velocity variations during solstices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Su, S.-Y.; Liu, C. H.; Chao, C.-K.</p> <p>2018-04-01</p> <p>Longitudinal distributions of post-midnight equatorial ionospheric irregularity occurrences observed by ROCSAT-<span class="hlt">1</span> (<span class="hlt">1</span>st satellite of the Republic of China) during moderate to high solar activity years in two solstices are studied with respect to the <span class="hlt">vertical</span> drift velocity and density variations. The post-midnight irregularity distributions are found to be similar to the well-documented pre-midnight ones, but are different from some published distributions taken during solar minimum years. Even though the post-midnight ionosphere is sinking in general, longitudes of frequent positive <span class="hlt">vertical</span> drift and high density seems to coincide with the longitudes of high irregularity occurrences. Large scatters found in the <span class="hlt">vertical</span> drift velocity and density around the dip equator in different ROCSAT-<span class="hlt">1</span> orbits indicate the existence of large and frequent variations in the <span class="hlt">vertical</span> drift velocity and density that seem to be able to provide sufficient perturbations for the Rayleigh-Taylor (RT) instability to cause the irregularity occurrences. The need of seeding agents such as gravity waves from atmospheric convective clouds to initiate the Rayleigh-Taylor instability may not be necessary.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008APS..DFD.LT004E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008APS..DFD.LT004E"><span><span class="hlt">Vertical</span> structures in vibrated wormlike micellar solutions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Epstein, Tamir; Deegan, Robert</p> <p>2008-11-01</p> <p><span class="hlt">Vertically</span> vibrated shear thickening particulate suspensions can support a free-standing interfaces oriented parallel to gravity. We find that shear thickening worm-like micellar solutions also support such <span class="hlt">vertical</span> interfaces. Above a threshold in acceleration, the solution spontaneously accumulates into a labyrinthine pattern characterized by a well-defined <span class="hlt">vertical</span> edge. The formation of <span class="hlt">vertical</span> structures is of interest because they are unique to shear-thickening fluids, and they indicate the existence of an unknown stress bearing mechanism.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ISPAr62W1..277H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ISPAr62W1..277H"><span>Acceleration of Sea Level <span class="hlt">Rise</span> Over Malaysian Seas from Satellite Altimeter</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hamid, A. I. A.; Din, A. H. M.; Khalid, N. F.; Omar, K. M.</p> <p>2016-09-01</p> <p>Sea level <span class="hlt">rise</span> becomes our concern nowadays as a result of variously contribution of climate change that cause by the anthropogenic effects. Global sea levels have been <span class="hlt">rising</span> through the past century and are projected to <span class="hlt">rise</span> at an accelerated rate throughout the 21st century. Due to this change, sea level is now constantly <span class="hlt">rising</span> and eventually will threaten many low-lying and unprotected coastal areas in many ways. This paper is proposing a significant effort to quantify the sea level trend over Malaysian seas based on the combination of multi-mission satellite altimeters over a period of 23 years. Eight altimeter missions are used to derive the absolute sea level from Radar Altimeter Database System (RADS). Data verification is then carried out to verify the satellite derived sea level <span class="hlt">rise</span> data with tidal data. Eight selected tide gauge stations from Peninsular Malaysia, Sabah and Sarawak are chosen for this data verification. The pattern and correlation of both measurements of sea level anomalies (SLA) are evaluated over the same period in each area in order to produce comparable results. Afterwards, the time series of the sea level trend is quantified using robust fit regression analysis. The findings clearly show that the absolute sea level trend is <span class="hlt">rising</span> and varying over the Malaysian seas with the rate of sea level varies and gradually increase from east to west of Malaysia. Highly confident and correlation level of the 23 years measurement data with an astonishing root mean square difference permits the absolute sea level trend of the Malaysian seas has raised at the rate 3.14 ± 0.12 mm yr-<span class="hlt">1</span> to 4.81 ± 0.15 mm yr-<span class="hlt">1</span> for the chosen sub-areas, with an overall mean of 4.09 ± 0.12 mm yr-<span class="hlt">1</span>. This study hopefully offers a beneficial sea level information to be applied in a wide range of related environmental and climatology issue such as flood and global warming.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApPhL.112p3501M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApPhL.112p3501M"><span>Formation of <span class="hlt">1</span>.4 MeV runaway electron flows in air using a solid-state generator with 10 MV/ns voltage <span class="hlt">rise</span> rate</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mesyats, G. A.; Pedos, M. S.; Rukin, S. N.; Rostov, V. V.; Romanchenko, I. V.; Sadykova, A. G.; Sharypov, K. A.; Shpak, V. G.; Shunailov, S. A.; Ul'masculov, M. R.; Yalandin, M. I.</p> <p>2018-04-01</p> <p>Fulfillment of the condition that the voltage <span class="hlt">rise</span> time across an air gap is comparable with the time of electron acceleration from a cathode to an anode allows a flow of runaway electrons (REs) to be formed with relativistic energies approaching that determined by the amplitude of the voltage pulse. In the experiment described here, an RE energy of <span class="hlt">1</span>.4 MeV was observed by applying a negative travelling voltage pulse of 860-kV with a maximum <span class="hlt">rise</span> rate of 10 MV/ns and a <span class="hlt">rise</span> time of 100-ps. The voltage pulse amplitude was doubled at the cathode of the 2-cm-long air gap due to the delay of conventional pulsed breakdown. The above-mentioned record-breaking voltage pulse of ˜120 ps duration with a peak power of 15 GW was produced by an all-solid-state pulsed power source utilising pulse compression/sharpening in a multistage gyromagnetic nonlinear transmission line.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Nanot..28n5201S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Nanot..28n5201S"><span>Graphene and PbS quantum dot hybrid <span class="hlt">vertical</span> phototransistor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Song, Xiaoxian; Zhang, Yating; Zhang, Haiting; Yu, Yu; Cao, Mingxuan; Che, Yongli; Dai, Haitao; Yang, Junbo; Ding, Xin; Yao, Jianquan</p> <p>2017-04-01</p> <p>A field-effect phototransistor based on a graphene and lead sulfide quantum dot (PbS QD) hybrid in which PbS QDs are embedded in a graphene matrix has been fabricated with a <span class="hlt">vertical</span> architecture through a solution process. The n-type Si/SiO2 substrate (gate), Au/Ag nanowire transparent source electrode, active layer and Au drain electrode are <span class="hlt">vertically</span> stacked in the device, which has a downscaled channel length of 250 nm. Photoinduced electrons in the PbS QDs leap into the conduction band and fill in the trap states, while the photoinduced holes left in the valence band transfer to the graphene and form the photocurrent under biases from which the photoconductive gain is evaluated. The graphene/QD-based <span class="hlt">vertical</span> phototransistor shows a photoresponsivity of 2 × 103 A W-<span class="hlt">1</span>, and specific detectivity up to 7 × 1012 Jones under 808 nm laser illumination with a light irradiance of 12 mW cm-2. The solution-processed <span class="hlt">vertical</span> phototransistor provides a new facile method for optoelectronic device applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E%26ES..126a2023S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E%26ES..126a2023S"><span>Dynamic facade module prototype development for solar radiation prevention in high <span class="hlt">rise</span> building</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sega Sufia Purnama, Muhammad; Sutanto, Dalhar</p> <p>2018-03-01</p> <p>Solar radiation is an aspect that high <span class="hlt">rise</span> building must avoid. The problem is, if high <span class="hlt">rise</span> building facade can’t overcome, the solar thermal will come in the building, and its affects on the increasing of room temperature above comfort range. A type of additional facade element that could solve solar thermal in high <span class="hlt">rise</span> building is adding a sun shading. A dynamic facade is a shade plane in high <span class="hlt">rise</span> building that can moved or changed on outside condition such as solar movement and intensity. This research will discuss the dynamic facade module prototype development in high <span class="hlt">rise</span> building in Jakarta. This research will be finish through some step. (<span class="hlt">1</span>) Static shading shadow simulation. (2) Dynamic facade concept design development. (3) Dynamic shading shadow simulation. (4) Making of dynamic facade module prototype. (5) Field test for the dynamic facade module prototype. The dynamic facade in Jakarta case will be effective to solve solar transmission in high <span class="hlt">rise</span> building rather than static facade.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016FrEaS...4...36K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016FrEaS...4...36K"><span>Global DEM Errors Underpredict Coastal Vulnerability to Sea Level <span class="hlt">Rise</span> and Flooding</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kulp, Scott; Strauss, Benjamin</p> <p>2016-04-01</p> <p>Elevation data based on NASA's Shuttle Radar Topography Mission (SRTM) have been widely used to evaluate threats from global sea level <span class="hlt">rise</span>, storm surge, and coastal floods. However, SRTM data are known to include large <span class="hlt">vertical</span> errors in densely urban or densely vegetated areas. The errors may propagate to derived land and population exposure assessments. We compare assessments based on SRTM data against references employing high-accuracy bare-earth elevation data generated from lidar data available for coastal areas of the United States. We find that both <span class="hlt">1</span>-arcsecond and 3-arcsecond horizontal resolution SRTM data systemically underestimate exposure across all assessed spatial scales and up to at least 10m above the high tide line. At 3m, <span class="hlt">1</span>-arcsecond SRTM underestimates U.S. population exposure by more than 60%, and under-predicts population exposure in 90% of coastal states, 87% of counties, and 83% of municipalities. These fractions increase with elevation, but error medians and variability fall to lower levels, with national exposure underestimated by just 24% at 10m. Results using 3-arcsecond SRTM are extremely similar. Coastal analyses based on SRTM data thus appear to greatly underestimate sea level and flood threats, especially at lower elevations. However, SRTM-based estimates may usefully be regarded as providing lower bounds to actual threats. We additionally assess the performance of NOAA's Global Land One-km Base Elevation Project (GLOBE), another publicly-available global DEM, but do not reach any definitive conclusion because of the spatial heterogeneity in its quality.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/1968/0146/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/1968/0146/report.pdf"><span><span class="hlt">Vertical</span> mass transfer in open channel flow</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Jobson, Harvey E.</p> <p>1968-01-01</p> <p>The <span class="hlt">vertical</span> mass transfer coefficient and particle fall velocity were determined in an open channel shear flow. Three dispersants, dye, fine sand and medium sand, were used with each of three flow conditions. The dispersant was injected as a continuous line source across the channel and downstream concentration profiles were measured. From these profiles along with the measured velocity distribution both the <span class="hlt">vertical</span> mass transfer coefficient and the local particle fall velocity were determined.The effects of secondary currents on the <span class="hlt">vertical</span> mixing process were discussed. Data was taken and analyzed in such a way as to largely eliminate the effects of these currents on the measured values. A procedure was developed by which the local value of the fall velocity of sand sized particles could be determined in an open channel flow. The fall velocity of the particles in the turbulent flow was always greater than their fall velocity in quiescent water. Reynolds analogy between the transfer of momentum and marked fluid particles was further substantiated. The turbulent Schmidt number was shown to be approximately <span class="hlt">1</span>.03 for an open channel flow with a rough boundary. Eulerian turbulence measurements were not sufficient to predict the <span class="hlt">vertical</span> transfer coefficient. <span class="hlt">Vertical</span> mixing of sediment is due to three semi-independent processes. These processes are: secondary currents, diffusion due to tangential velocity fluctuations and diffusion due to the curvature of the fluid particle path lines. The diffusion coefficient due to tangential velocity fluctuations is approximately proportional to the transfer coefficient of marked fluid particles. The proportionality constant is less than or equal to <span class="hlt">1</span>.0 and decreases with increasing particle size. The diffusion coefficient due to the curvature of the fluid particle path lines is not related to the diffusion coefficient for marked fluid particles and increases with particle size, at least for sediment particles in the sand size</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A43F2537Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A43F2537Z"><span>Modeling Emissions and <span class="hlt">Vertical</span> Plume Transport of Crop Residue Burning Experiments in the Pacific Northwest</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhou, L.; Baker, K. R.; Napelenok, S. L.; Pouliot, G.; Elleman, R. A.; ONeill, S. M.; Urbanski, S. P.; Wong, D. C.</p> <p>2017-12-01</p> <p>Crop residue burning has long been a common practice in agriculture with the smoke emissions from the burning linked to negative health impacts. A field study in eastern Washington and northern Idaho in August 2013 consisted of multiple burns of well characterized fuels with nearby surface and aerial measurements including trace species concentrations, plume <span class="hlt">rise</span> height and boundary layer structure. The chemical transport model CMAQ (Community Multiscale Air Quality Model) was used to assess the fire emissions and subsequent <span class="hlt">vertical</span> plume transport. The study first compared assumptions made by the 2014 National Emission Inventory approach for crop residue burning with the fuel and emissions information obtained from the field study and then investigated the sensitivity of modeled carbon monoxide (CO) and PM2.5 concentrations to these different emission estimates and plume <span class="hlt">rise</span> treatment with CMAQ. The study suggests that improvements to the current parameterizations are needed in order for CMAQ to reliably reproduce smoke plumes from burning. In addition, there is enough variability in the smoke emissions, stemming from variable field-specific information such as field size, that attempts to model crop residue burning should use field-specific information whenever possible.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.A53H0240N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.A53H0240N"><span>A Direct Detection <span class="hlt">1</span>.6μm DIAL with Three Wavelengths for Measurements of <span class="hlt">Vertical</span> CO2 Concentration and Temperature Profiles in the Atmosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nagasawa, C.; Abo, M.; Shibata, Y.; Nagai, T.; Tsukamoto, M.</p> <p>2012-12-01</p> <p>We report the new <span class="hlt">1</span>.6 μm DIAL system that can measure the temperature profiles with the CO2 concentration profiles in the atmosphere because of improvement of measurement accuracy of the CO2 density and mixing ratio (ppm). We have developed a direct detection <span class="hlt">1</span>.6 μm differential absorption lidar (DIAL) technique to perform range-resolved measurements of <span class="hlt">vertical</span> CO2 concentration profiles in the atmosphere [Sakaizawa et al. 2009]. Our <span class="hlt">1</span>.6 μm DIAL system consists of the Optical Parametric Generator (OPG) transmitter that excited by the LD pumped Nd:YAG laser with high repetition rate (500 Hz) and the receiving optics that included the near-infrared photomultiplier tube with high quantum efficiency operating at the photon counting mode and the telescope with larger aperture than that of the coherent detection method. Laser beams of three wavelengths around a CO2 absorption line is transmitted alternately to the atmosphere for measurements of CO2 concentration and temperature profiles. Moreover, a few retrieval algorithms of CO2-DIAL are also performed for improvement of measurement accuracy. The accurate <span class="hlt">vertical</span> CO2 profiles in the troposphere are highly desirable in the inverse techniques to improve quantification and understanding of the global budget of CO2 and also global climate changes [Stephens et al. 2007]. In comparison with the ground-based monitoring network, CO2 measurements for <span class="hlt">vertical</span> profiles in the troposphere have been limited to campaign-style aircraft and commercial airline observations with the limited spatial and temporal coverage. This work was financially supported by the System Development Program for Advanced Measurement and Analysis of the Japan Science and Technology Agency. References Sakaizawa, D., C. Nagasawa, T. Nagai, M. Abo, Y. Shibata, H. Nagai, M. Nakazato, and T. Sakai, Development of a <span class="hlt">1</span>.6μm differential absorption lidar with a quasi-phase-matching optical parametric oscillator and photon-counting detector for the <span class="hlt">vertical</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1335974-imaging-taurine-central-nervous-system-using-chemically-specific-ray-fluorescence-imaging-sulfur-edge','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1335974-imaging-taurine-central-nervous-system-using-chemically-specific-ray-fluorescence-imaging-sulfur-edge"><span>Imaging Taurine in the Central Nervous System Using Chemically Specific X-ray Fluorescence Imaging at the Sulfur K-Edge</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Hackett, Mark J.; Paterson, Phyllis G.; Pickering, Ingrid J.</p> <p></p> <p>A method to image taurine distributions within the central nervous system and other organs has long been sought. Since taurine is small and mobile, it cannot be chemically “tagged” and imaged using conventional immuno-histochemistry methods. Combining numerous indirect measurements, taurine is known to play critical roles in brain function during health and disease and is proposed to act as a neuro-osmolyte, neuro-modulator, and possibly a neuro-transmitter. Elucidation of taurine’s neurochemical roles and importance would be substantially enhanced by a direct method to visualize alterations, due to physiological and pathological events in the brain, in the local concentration of taurine atmore » or near cellular spatial resolution in vivo or in situ in tissue sections. We thus have developed chemically specific X-ray fluorescence imaging (<span class="hlt">XFI</span>) at the sulfur K-edge to image the sulfonate group in taurine in situ in ex vivo tissue sections. To our knowledge, this represents the first undistorted imaging of taurine distribution in brain at 20 μm resolution. We report quantitative technique validation by imaging taurine in the cerebellum and hippocampus regions of the rat brain. Further, we apply the technique to image taurine loss from the vulnerable CA<span class="hlt">1</span> (cornus ammonis <span class="hlt">1</span>) sector of the rat hippocampus following global brain ischemia. The location-specific loss of taurine from CA<span class="hlt">1</span> but not CA3 neurons following ischemia reveals osmotic stress may be a key factor in delayed neurodegeneration after a cerebral ischemic insult and highlights the significant potential of chemically specific <span class="hlt">XFI</span> to study the role of taurine in brain disease.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22663684-vertical-structure-radiation-pressure-dominated-thin-disks-link-between-vertical-advection-convective-stability','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22663684-vertical-structure-radiation-pressure-dominated-thin-disks-link-between-vertical-advection-convective-stability"><span><span class="hlt">Vertical</span> Structure of Radiation-pressure-dominated Thin Disks: Link between <span class="hlt">Vertical</span> Advection and Convective Stability</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Gong, Hong-Yu; Gu, Wei-Min, E-mail: guwm@xmu.edu.cn</p> <p>2017-04-20</p> <p>In the classic picture of standard thin accretion disks, viscous heating is balanced by radiative cooling through the diffusion process, and the radiation-pressure-dominated inner disk suffers convective instability. However, recent simulations have shown that, owing to the magnetic buoyancy, the <span class="hlt">vertical</span> advection process can significantly contribute to energy transport. In addition, in comparing the simulation results with the local convective stability criterion, no convective instability has been found. In this work, following on from simulations, we revisit the <span class="hlt">vertical</span> structure of radiation-pressure-dominated thin disks and include the <span class="hlt">vertical</span> advection process. Our study indicates a link between the additional energy transportmore » and the convectively stable property. Thus, the <span class="hlt">vertical</span> advection not only significantly contributes to the energy transport, but it also plays an important role in making the disk convectively stable. Our analyses may help to explain the discrepancy between classic theory and simulations on standard thin disks.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..17.6886S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..17.6886S"><span>The effect of sediment thermal conductivity on <span class="hlt">vertical</span> groundwater flux estimates</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sebok, Eva; Müller, Sascha; Engesgaard, Peter; Duque, Carlos</p> <p>2015-04-01</p> <p>The interaction between groundwater and surface water is of great importance both from ecological and water management perspective. The exchange fluxes are often estimated based on <span class="hlt">vertical</span> temperature profiles taken from shallow sediments assuming a homogeneous standard value of sediment thermal conductivity. Here we report on a field investigation in a stream and in a fjord, where <span class="hlt">vertical</span> profiles of sediment thermal conductivity and temperatures were measured in order to, (i) define the <span class="hlt">vertical</span> variability in sediment thermal conductivity, (ii) quantify the effect of heterogeneity in sediment thermal conductivity on the estimated <span class="hlt">vertical</span> groundwater fluxes. The study was carried out at field sites located in Ringkøbing fjord and Holtum stream in Western Denmark. Both locations have soft, sandy sediments with an upper organic layer at the fjord site. First 9 and 12 <span class="hlt">vertical</span> sediment temperature profiles up to 0.5 m depth below the sediment bed were collected in the fjord and in the stream, respectively. Later sediment cores of 0.05 m diameter were removed at the location of the temperature profiles. Sediment thermal conductivity was measured in the sediment cores at 0.<span class="hlt">1</span> m intervals with a Decagon KD2 Pro device. A <span class="hlt">1</span>D flow and heat transport model (HydroGeoSphere) was set up and <span class="hlt">vertical</span> groundwater fluxes were estimated based on the measured <span class="hlt">vertical</span> sediment temperature profiles by coupling the model with PEST. To determine the effect of heterogeneity in sediment thermal conductivity on estimated <span class="hlt">vertical</span> groundwater fluxes, the model was run by assigning (i) a homogeneous thermal conductivity for all sediment layers, calculated as the average sediment thermal conductivity of the profile, (ii) measured sediment thermal conductivities to the different model layers. The field survey showed that sediment thermal conductivity over a 0.5 m profile below the sediment bed is not uniform, having the largest variability in the fjord where organic sediments were also</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25897963','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25897963"><span>Increased <span class="hlt">vertical</span> impact forces and altered running mechanics with softer midsole shoes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Baltich, Jennifer; Maurer, Christian; Nigg, Benno M</p> <p>2015-01-01</p> <p>To date it has been thought that shoe midsole hardness does not affect <span class="hlt">vertical</span> impact peak forces during running. This conclusion is based partially on results from experimental data using homogeneous samples of participants that found no difference in <span class="hlt">vertical</span> impact peaks when running in shoes with different midsole properties. However, it is currently unknown how apparent joint stiffness is affected by shoe midsole hardness. An increase in apparent joint stiffness could result in a harder landing, which should result in increased <span class="hlt">vertical</span> impact peaks during running. The purpose of this study was to quantify the effect of shoe midsole hardness on apparent ankle and knee joint stiffness and the associated <span class="hlt">vertical</span> ground reaction force for age and sex subgroups during heel-toe running. 93 runners (male and female) aged 16-75 years ran at 3.33 ± 0.15 m/s on a 30 m-long runway with soft, medium and hard midsole shoes. The <span class="hlt">vertical</span> impact peak increased as the shoe midsole hardness decreased (mean(SE); soft: <span class="hlt">1</span>.70BW(0.03), medium: <span class="hlt">1</span>.64BW(0.03), hard: <span class="hlt">1</span>.54BW(0.03)). Similar results were found for the apparent ankle joint stiffness where apparent stiffness increased as the shoe midsole hardness decreased (soft: 2.08BWm/º x 100 (0.05), medium: <span class="hlt">1</span>.92 BWm/º x 100 (0.05), hard: <span class="hlt">1</span>.85 BWm/º x 100 (0.05)). Apparent knee joint stiffness increased for soft (<span class="hlt">1</span>.06BWm/º x 100 (0.04)) midsole compared to the medium (0.95BWm/º x 100 (0.04)) and hard (0.96BWm/º x 100 (0.04)) midsoles for female participants. The results from this study confirm that shoe midsole hardness can have an effect on <span class="hlt">vertical</span> impact force peaks and that this may be connected to the hardness of the landing. The results from this study may provide useful information regarding the development of cushioning guidelines for running shoes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4405580','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4405580"><span>Increased <span class="hlt">Vertical</span> Impact Forces and Altered Running Mechanics with Softer Midsole Shoes</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Baltich, Jennifer; Maurer, Christian; Nigg, Benno M.</p> <p>2015-01-01</p> <p>To date it has been thought that shoe midsole hardness does not affect <span class="hlt">vertical</span> impact peak forces during running. This conclusion is based partially on results from experimental data using homogeneous samples of participants that found no difference in <span class="hlt">vertical</span> impact peaks when running in shoes with different midsole properties. However, it is currently unknown how apparent joint stiffness is affected by shoe midsole hardness. An increase in apparent joint stiffness could result in a harder landing, which should result in increased <span class="hlt">vertical</span> impact peaks during running. The purpose of this study was to quantify the effect of shoe midsole hardness on apparent ankle and knee joint stiffness and the associated <span class="hlt">vertical</span> ground reaction force for age and sex subgroups during heel-toe running. 93 runners (male and female) aged 16-75 years ran at 3.33 ± 0.15 m/s on a 30 m-long runway with soft, medium and hard midsole shoes. The <span class="hlt">vertical</span> impact peak increased as the shoe midsole hardness decreased (mean(SE); soft: <span class="hlt">1</span>.70BW(0.03), medium: <span class="hlt">1</span>.64BW(0.03), hard: <span class="hlt">1</span>.54BW(0.03)). Similar results were found for the apparent ankle joint stiffness where apparent stiffness increased as the shoe midsole hardness decreased (soft: 2.08BWm/º x 100 (0.05), medium: <span class="hlt">1</span>.92 BWm/º x 100 (0.05), hard: <span class="hlt">1</span>.85 BWm/º x 100 (0.05)). Apparent knee joint stiffness increased for soft (<span class="hlt">1</span>.06BWm/º x 100 (0.04)) midsole compared to the medium (0.95BWm/º x 100 (0.04)) and hard (0.96BWm/º x 100 (0.04)) midsoles for female participants. The results from this study confirm that shoe midsole hardness can have an effect on <span class="hlt">vertical</span> impact force peaks and that this may be connected to the hardness of the landing. The results from this study may provide useful information regarding the development of cushioning guidelines for running shoes. PMID:25897963</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_21 --> <div id="page_22" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="421"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28403012','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28403012"><span>Can <span class="hlt">Vertical</span> Integration Reduce Hospital Readmissions? A Difference-in-Differences Approach.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lopes, Sílvia; Fernandes, Óscar B; Marques, Ana Patrícia; Moita, Bruno; Sarmento, João; Santana, Rui</p> <p>2017-05-01</p> <p><span class="hlt">Vertical</span> integration is expected to improve communication and coordination between inpatient care and care after discharge. Despite being used across health systems worldwide, evidence about its impact on readmissions is sparse and contradictory. To assess the impact of <span class="hlt">vertical</span> integration on hospital readmissions. Using difference-in-differences we compared readmissions before and after <span class="hlt">vertical</span> integration in 6 Portuguese hospitals for years 2004-2013. A control group with 6 similar hospitals not integrated was utilized. Considered outcome was 30-day unplanned readmission. We used logistic regression at the admission level and accounted for patients' risk factors using claims data. Analyses for each hospital and selected conditions were also run. Our results suggest that readmissions decreased overall after <span class="hlt">vertical</span> integration [odds ratio (OR)=0.900; 95% confidence interval (CI), 0.812-0.997]. Hospital analysis indicated that there was no impact for 2 hospitals (OR=0.960; 95% CI, 0.848-<span class="hlt">1</span>.087 and OR=0.944; 95% CI, 0.857-<span class="hlt">1</span>.038), and a positive effect in 4 hospitals (greatest effect: OR=0.811; 95% CI, 0.736-0.894). A positive evolution was observed for a limited number of conditions, with better results for diabetes with complications (OR=0.689; 95% CI, 0.525-0.904), but no impact regarding congestive heart failure (OR=<span class="hlt">1</span>.067; 95% CI, 0.827-<span class="hlt">1</span>.377). Merging acute and primary care providers was associated with reduced readmissions, even though improvements were not found for all institutions or condition-specific groups. There are still challenges to be addressed regarding the success of <span class="hlt">vertical</span> integration in reducing 30-day hospital readmissions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/864967','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/864967"><span>Amplitude- and <span class="hlt">rise</span>-time-compensated filters</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Nowlin, Charles H.</p> <p>1984-01-01</p> <p>An amplitude-compensated <span class="hlt">rise</span>-time-compensated filter for a pulse time-of-occurrence (TOOC) measurement system is disclosed. The filter converts an input pulse, having the characteristics of random amplitudes and random, non-zero <span class="hlt">rise</span> times, to a bipolar output pulse wherein the output pulse has a zero-crossing time that is independent of the <span class="hlt">rise</span> time and amplitude of the input pulse. The filter differentiates the input pulse, along the linear leading edge of the input pulse, and subtracts therefrom a pulse fractionally proportional to the input pulse. The filter of the present invention can use discrete circuit components and avoids the use of delay lines.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120016516','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120016516"><span>The <span class="hlt">Rise</span> and Fall of Star Formation Histories of Blue Galaxies at Redshifts 0.2 < z < <span class="hlt">1</span>.4</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Pacifici, Camilla; Kassin, Susan A.; Weiner, Benjamin; Charlot, Stephane; Gardner, Jonathan P.</p> <p>2012-01-01</p> <p>Popular cosmological scenarios predict that galaxies form hierarchically from the merger of many progenitor, each with their own unique star formation history (SFH). We use the approach recently developed by Pacifici et al. to constrain the SFHs of 4517 blue (presumably star-forming) galaxies with spectroscopic redshifts in the range O.2 < z < <span class="hlt">1</span>:4 from the All-Wavelength Extended Groth Strip International Survey (AEGIS). This consists in the Bayesian analysis of the observed galaxy spectral ' energy distributions with a comprehensive library of synthetic spectra assembled using state-of-the-art models of star formation and chemical enrichment histories, stellar population synthesis, nebular emission and attenuation by dust. We constrain the SFH of each galaxy in our sample by comparing the observed fluxes in the B, R,l and K(sub s) bands and rest-frame optical emission-line luminosities with those of one million model spectral energy distributions. We explore the dependence of the resulting SFH on galaxy stellar mass and redshift. We find that the average SFHs of high-mass galaxies <span class="hlt">rise</span> and fall in a roughly symmetric bell-shaped manner, while those of low-mass galaxies <span class="hlt">rise</span> progressively in time, consistent with the typically stronger activity of star formation in low-mass compared to high-mass galaxies. For galaxies of all masses, the star formation activity <span class="hlt">rises</span> more rapidly at high than at low redshift. These findings imply that the standard approximation of exponentially declining SFHs wIdely used to interpret observed galaxy spectral energy distributions is not appropriate to constrain the physical parameters of star-forming galaxies at intermediate redshifts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007ACPD....712751V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007ACPD....712751V"><span><span class="hlt">Vertical</span> distribution of ozone and VOCs in the low boundary layer of Mexico City</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Velasco, E.; Márquez, C.; Bueno, E.; Bernabé, R. M.; Sánchez, A.; Fentanes, O.; Wöhrnschimmel, H.; Cárdenas, B.; Kamilla, A.; Wakamatsu, S.; Molina, L. T.</p> <p>2007-08-01</p> <p>The evolution of ozone and 13 volatile organic compounds (VOCs) in the boundary layer of Mexico City was investigated during 2000-2004 to improve our understanding of the complex interactions between those trace gases and meteorological variables, and their influence on the air quality of a polluted megacity. A tethered balloon, fitted with electrochemical and meteorological sondes, was used to obtain detailed <span class="hlt">vertical</span> profiles of ozone and meteorological parameters up to 1000 m above ground during part of the diurnal cycle (02:00-18:00 h). VOCs samples were collected up to 200 m by pumping air to canisters with a Teflon tube attached to the tether line. Overall, features of these profiles were found to be consistent with a simple picture of nighttime trapping of ozone in an upper residual layer and of VOCs in a shallow unstable layer above the ground. After sunrise an ozone balance is determined by photochemical production, entrainment from the upper residual layer and destruction by titration with NO, delaying the ground-level ozone <span class="hlt">rise</span> by 2 h. The subsequent evolution of the conductive boundary layer and <span class="hlt">vertical</span> distribution of pollutants are discussed in terms of the energy balance, the presence of turbulence and the atmospheric stability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20090023136','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20090023136"><span><span class="hlt">Vertical</span> Lift - Not Just For Terrestrial Flight</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Young, Larry A</p> <p>2000-01-01</p> <p>Autonomous <span class="hlt">vertical</span> lift vehicles hold considerable potential for supporting planetary science and exploration missions. This paper discusses several technical aspects of <span class="hlt">vertical</span> lift planetary aerial vehicles in general, and specifically addresses technical challenges and work to date examining notional <span class="hlt">vertical</span> lift vehicles for Mars, Titan, and Venus exploration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.H33F1459S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.H33F1459S"><span>Temporal Variability in <span class="hlt">Vertical</span> Groundwater Fluxes and the Effect of Solar Radiation on Streambed Temperatures Based on <span class="hlt">Vertical</span> High Resolution Distributed Temperature Sensing</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sebok, E.; Karan, S.; Engesgaard, P. K.; Duque, C.</p> <p>2013-12-01</p> <p>Due to its large spatial and temporal variability, groundwater discharge to streams is difficult to quantify. Methods using <span class="hlt">vertical</span> streambed temperature profiles to estimate <span class="hlt">vertical</span> fluxes are often of coarse <span class="hlt">vertical</span> spatial resolution and neglect to account for the natural heterogeneity in thermal conductivity of streambed sediments. Here we report on a field investigation in a stream, where air, stream water and streambed sediment temperatures were measured by Distributed Temperature Sensing (DTS) with high spatial resolution to; (i) detect spatial and temporal variability in groundwater discharge based on <span class="hlt">vertical</span> streambed temperature profiles, (ii) study the thermal regime of streambed sediments exposed to different solar radiation influence, (iii) describe the effect of solar radiation on the measured streambed temperatures. The study was carried out at a field site located along Holtum stream, in Western Denmark. The 3 m wide stream has a sandy streambed with a cobbled armour layer, a mean discharge of 200 l/s and a mean depth of 0.3 m. Streambed temperatures were measured with a high-resolution DTS system (HR-DTS). By helically wrapping the fiber optic cable around two PVC pipes of 0.05 m and 0.075 m outer diameter over <span class="hlt">1</span>.5 m length, temperature measurements were recorded with 5.7 mm and 3.8 mm <span class="hlt">vertical</span> spacing, respectively. The HR-DTS systems were installed 0.7 m deep in the streambed sediments, crossing both the sediment-water and the water-air interface, thus yielding high resolution water and air temperature data as well. One of the HR-DTS systems was installed in the open stream channel with only topographical shading, while the other HR-DTS system was placed 7 m upstream, under the canopy of a tree, thus representing the shaded conditions with reduced influence of solar radiation. Temperature measurements were taken with 30 min intervals between 16 April and 25 June 2013. The thermal conductivity of streambed sediments was calibrated in a <span class="hlt">1</span>D flow</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=brainstem&pg=2&id=EJ986631','ERIC'); return false;" href="https://eric.ed.gov/?q=brainstem&pg=2&id=EJ986631"><span>Neural Control of <span class="hlt">Rising</span> and Falling Tones in Mandarin Speakers Who Stutter</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Howell, Peter; Jiang, Jing; Peng, Danling; Lu, Chunming</p> <p>2012-01-01</p> <p>Neural control of <span class="hlt">rising</span> and falling tones in Mandarin people who stutter (PWS) was examined by comparing with that which occurs in fluent speakers [Howell, Jiang, Peng, and Lu (2012). Neural control of fundamental frequency <span class="hlt">rise</span> and fall in Mandarin tones. "Brain and Language, 121"(<span class="hlt">1</span>), 35-46]. Nine PWS and nine controls were scanned. Functional…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26854286','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26854286"><span>Effects of <span class="hlt">Vertical</span> Direction and Aperture Size on the Perception of Visual Acceleration.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mueller, Alexandra S; González, Esther G; McNorgan, Chris; Steinbach, Martin J; Timney, Brian</p> <p>2016-02-06</p> <p>It is not well understood whether the distance over which moving stimuli are visible affects our sensitivity to the presence of acceleration or our ability to track such stimuli. It is also uncertain whether our experience with gravity creates anisotropies in how we detect <span class="hlt">vertical</span> acceleration and deceleration. To address these questions, we varied the <span class="hlt">vertical</span> extent of the aperture through which we presented <span class="hlt">vertically</span> accelerating and decelerating random dot arrays. We hypothesized that observers would better detect and pursue accelerating and decelerating stimuli that extend over larger than smaller distances. In Experiment <span class="hlt">1</span>, we tested the effects of <span class="hlt">vertical</span> direction and aperture size on acceleration and deceleration detection accuracy. Results indicated that detection is better for downward motion and for large apertures, but there is no difference between <span class="hlt">vertical</span> acceleration and deceleration detection. A control experiment revealed that our manipulation of <span class="hlt">vertical</span> aperture size affects the ability to track <span class="hlt">vertical</span> motion. Smooth pursuit is better (i.e., with higher peak velocities) for large apertures than for small apertures. Our findings suggest that the ability to detect <span class="hlt">vertical</span> acceleration and deceleration varies as a function of the direction and <span class="hlt">vertical</span> extent over which an observer can track the moving stimulus. © The Author(s) 2016.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Classical+AND+Perspectives&id=EJ1014611','ERIC'); return false;" href="https://eric.ed.gov/?q=Classical+AND+Perspectives&id=EJ1014611"><span>Measuring Growth with <span class="hlt">Vertical</span> Scales</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Briggs, Derek C.</p> <p>2013-01-01</p> <p>A <span class="hlt">vertical</span> score scale is needed to measure growth across multiple tests in terms of absolute changes in magnitude. Since the warrant for subsequent growth interpretations depends upon the assumption that the scale has interval properties, the validation of a <span class="hlt">vertical</span> scale would seem to require methods for distinguishing interval scales from…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/AD1011801','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/AD1011801"><span>Study of Charge Transport in <span class="hlt">Vertically</span> Aligned Nitride Nanowire Based Core Shell P-I-N Junctions</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2016-07-01</p> <p><span class="hlt">Vertically</span>- Aligned Nitride Nanowire Based Core Shell P-I-N Junctions Distribution Statement A. Approved for public release; distribution is...Study of Charge Transport in <span class="hlt">Vertically</span>- Aligned Nitride Nanowire Based Core Shell P-I-N Junctions Grant Number: HDTRA<span class="hlt">1-14-1</span>-0003 Principal...Investigator: Abhishek Motayed University of Maryland DISTRIBUTION A: Public Release Study of Charge Transport in <span class="hlt">Vertically</span>-Aligned Nitride Nanowire</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E%26ES..140a2071S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E%26ES..140a2071S"><span>Effect of Lime Stabilization on <span class="hlt">Vertical</span> Deformation of Laterite Halmahera Soil</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Saing, Zubair; Djainal, Herry</p> <p>2018-04-01</p> <p>In this paper, the study was conducted to determine the lime effect on <span class="hlt">vertical</span> deformation of road base physical model of laterite Halmahera soil. The samples of laterite soil were obtained from Halmahera Island, North Maluku Province, Indonesia. Soil characteristics were obtained from laboratory testing, according to American Standard for Testing and Materials (ASTM), consists of physical, mechanical, minerals, and chemical. The base layer of physical model testing with the dimension; 2m of length, 2m of width, and <span class="hlt">1</span>.5m of height. The addition of lime with variations of 3, 5, 7, an 10%, based on maximum dry density of standard Proctor test results and cured for 28 days. The model of lime treated laterite Halmahera soil with 0,<span class="hlt">1</span>m thickness placed on subgrade layer with <span class="hlt">1</span>,5m thickness. Furthermore, the physical model was given static <span class="hlt">vertical</span> loading. Some dial gauge is placed on the lime treated soil surface with distance interval 20cm, to read the <span class="hlt">vertical</span> deformation that occurs during loading. The experimentals data was analyzed and validated with numerical analysis using finite element method. The results showed that the <span class="hlt">vertical</span> deformation reduced significantly on 10% lime content (three times less than untreated soil), and qualify for maximum deflection (standard requirement L/240) on 7-10% lime content.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUOSME14D0629F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUOSME14D0629F"><span>Effects of light conditions and temperature gradients on <span class="hlt">vertical</span> migration behavior of larval Arctic cod (Boreogadus saida) and walleye pollock (Gadus chalcogramma)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Flanders, K. R.; Laurel, B.</p> <p>2016-02-01</p> <p>Early life stages of marine fishes must maximize growth while minimizing vulnerability to predators. Larval stages in particular are subject to ocean currents, but encounter favorable habitats by adjusting their <span class="hlt">vertical</span> position in the water column. The investigation of environmental cues that change larval fish behavior is therefore crucial to understanding larval drift and dispersal modeling, and subsequently population structure and connectivity. In this study, the behavioral responses of larval Arctic cod (Boreogadus saida) and walleye pollock (Gadus chalcogramma) in a <span class="hlt">vertical</span> water column were examined. Two prominent environmental variables, light and temperature, were manipulated over 3 h during observational trials. Light intensity was studied at two levels (<span class="hlt">1</span>.484 x 101 μE m-2 s-<span class="hlt">1</span> ; 2.54 x102 μE m-2 s-<span class="hlt">1</span>), and a diel effect was studied through the removal of light after 2 h. Light intensity did not significantly impact the position of either species in a <span class="hlt">vertical</span> water column. However, a significant difference by species was apparent when all light levels were considered: the mean position of Arctic cod was closer to the surface of the water than that of walleye pollock. The effect of temperature through the introduction of a thermocline (range 5.6°C - <span class="hlt">1</span>.5°C) was limited to walleye pollock given the Arctic cod larvae were surface oriented across all light treatments. However, the thermocline did not significantly impact the relative change in position from light to dark in walleye pollock, likely because they were also surface oriented in control treatments. These results could be incorporated into future larval dispersal and survival models, particularly in Alaskan and Arctic waters, to investigate changes in species distributions resulting from global warming impacts. These results also indicate population structures of Arctic cod and walleye pollock could be affected, which may be reflected in ecosystem and trophic interactions. Because Arctic cod</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014SJRUE..14....5S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014SJRUE..14....5S"><span>Thermal Impacts of <span class="hlt">Vertical</span> Greenery Systems</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Safikhani, Tabassom; Abdullah, Aminatuzuhariah Megat; Ossen, Dilshan Remaz; Baharvand, Mohammad</p> <p>2014-12-01</p> <p>- Using <span class="hlt">vertical</span> greenery systems to reduce heat transmission is becoming more common in modern architecture. <span class="hlt">Vertical</span> greenery systems are divided into two main categories; green facades and living walls. This study aims to examine the thermal performance of <span class="hlt">vertical</span> greenery systems in hot and humid climates. An experimental procedure was used to measure indoor temperature and humidity. These parameters were also measured for the gap between the <span class="hlt">vertical</span> greenery systems and wall surfaces. Three boxes were used as small-scale rooms. Two boxes were provided with either a living wall or a green facade and one box did not have any greenery (benchmark). Blue Trumpet Vine was used in the <span class="hlt">vertical</span> greenery systems. The data were recorded over the course of three sunny days in April 2013. An analyses of the results showed that the living wall and green facade reduced indoor temperature up to 4.0 °C and 3.0 °C, respectively. The living wall and green facade also reduced cavity temperatures by 8.0 °C and 6.5 °C, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20090026469','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20090026469"><span>Rough-water Landings of a 0.<span class="hlt">1</span>-Size Powered Dynamic Model of the XP5Y-<span class="hlt">1</span> Flying Boat with Two Types of Afterbody - Langley Tank Model 228 (TED No. NACA DE309)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Garrison, Charlie C.</p> <p>1949-01-01</p> <p>A 0.<span class="hlt">1</span>-size powered dynamic model of a large, high-speed flying boat was landed in Langley tank no. <span class="hlt">1</span> into oncoming waves 4 feet high (full size). The model was tested with two afterbodies of differing lengths (4.12 and 6.63 beams). The short afterbody had a constant angle of dead <span class="hlt">rise</span> of 22.5deg and a keel angle of 6.5deg. The long afterbody had warped dead <span class="hlt">rise</span> and a keel angle of 8.5deg. The <span class="hlt">vertical</span> accelerations were slightly greater and the maximum angular accelerations and maxim= trims were slightly less for the model with the long afterbody than for the model with -the short afterbody. A wave length of 210 feet (full size) imposed the highest accelerations on the model with either the long or the short afterbody.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/14604562','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/14604562"><span>Why did employee health insurance contributions <span class="hlt">rise</span>?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gruber, Jonathan; McKnight, Robin</p> <p>2003-11-01</p> <p>We explore the causes of the dramatic <span class="hlt">rise</span> in employee contributions to health insurance over the past two decades. In 1982, 44% of those who were covered by their employer-provided health insurance had their costs fully financed by their employer, but by 1998 this had fallen to 28%. We discuss the theory of why employers might shift premiums to their employees, and empirically model the role of four factors suggested by the theory. We find that there was a large impact of falling tax rates, <span class="hlt">rising</span> eligibility for insurance through the Medicaid system, <span class="hlt">rising</span> medical costs, and increased managed care penetration. Overall, this set of factors can explain more than one-half of the <span class="hlt">rise</span> in employee premiums over the 1982-1996 period.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27917328','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27917328"><span>GPS Imaging of <span class="hlt">vertical</span> land motion in California and Nevada: Implications for Sierra Nevada uplift.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hammond, William C; Blewitt, Geoffrey; Kreemer, Corné</p> <p>2016-10-01</p> <p>We introduce Global Positioning System (GPS) Imaging, a new technique for robust estimation of the <span class="hlt">vertical</span> velocity field of the Earth's surface, and apply it to the Sierra Nevada Mountain range in the western United States. Starting with <span class="hlt">vertical</span> position time series from Global Positioning System (GPS) stations, we first estimate <span class="hlt">vertical</span> velocities using the MIDAS robust trend estimator, which is insensitive to undocumented steps, outliers, seasonality, and heteroscedasticity. Using the Delaunay triangulation of station locations, we then apply a weighted median spatial filter to remove velocity outliers and enhance signals common to multiple stations. Finally, we interpolate the data using weighted median estimation on a grid. The resulting velocity field is temporally and spatially robust and edges in the field remain sharp. Results from data spanning 5-20 years show that the Sierra Nevada is the most rapid and extensive uplift feature in the western United States, <span class="hlt">rising</span> up to 2 mm/yr along most of the range. The uplift is juxtaposed against domains of subsidence attributable to groundwater withdrawal in California's Central Valley. The uplift boundary is consistently stationary, although uplift is faster over the 2011-2016 period of drought. Uplift patterns are consistent with groundwater extraction and concomitant elastic bedrock uplift, plus slower background tectonic uplift. A discontinuity in the velocity field across the southeastern edge of the Sierra Nevada reveals a contrast in lithospheric strength, suggesting a relationship between late Cenozoic uplift of the southern Sierra Nevada and evolution of the southern Walker Lane.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5114868','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5114868"><span>GPS Imaging of <span class="hlt">vertical</span> land motion in California and Nevada: Implications for Sierra Nevada uplift</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Blewitt, Geoffrey; Kreemer, Corné</p> <p>2016-01-01</p> <p>Abstract We introduce Global Positioning System (GPS) Imaging, a new technique for robust estimation of the <span class="hlt">vertical</span> velocity field of the Earth's surface, and apply it to the Sierra Nevada Mountain range in the western United States. Starting with <span class="hlt">vertical</span> position time series from Global Positioning System (GPS) stations, we first estimate <span class="hlt">vertical</span> velocities using the MIDAS robust trend estimator, which is insensitive to undocumented steps, outliers, seasonality, and heteroscedasticity. Using the Delaunay triangulation of station locations, we then apply a weighted median spatial filter to remove velocity outliers and enhance signals common to multiple stations. Finally, we interpolate the data using weighted median estimation on a grid. The resulting velocity field is temporally and spatially robust and edges in the field remain sharp. Results from data spanning 5–20 years show that the Sierra Nevada is the most rapid and extensive uplift feature in the western United States, <span class="hlt">rising</span> up to 2 mm/yr along most of the range. The uplift is juxtaposed against domains of subsidence attributable to groundwater withdrawal in California's Central Valley. The uplift boundary is consistently stationary, although uplift is faster over the 2011–2016 period of drought. Uplift patterns are consistent with groundwater extraction and concomitant elastic bedrock uplift, plus slower background tectonic uplift. A discontinuity in the velocity field across the southeastern edge of the Sierra Nevada reveals a contrast in lithospheric strength, suggesting a relationship between late Cenozoic uplift of the southern Sierra Nevada and evolution of the southern Walker Lane. PMID:27917328</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JGRB..121.7681H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JGRB..121.7681H"><span>GPS Imaging of <span class="hlt">vertical</span> land motion in California and Nevada: Implications for Sierra Nevada uplift</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hammond, William C.; Blewitt, Geoffrey; Kreemer, Corné</p> <p>2016-10-01</p> <p>We introduce Global Positioning System (GPS) Imaging, a new technique for robust estimation of the <span class="hlt">vertical</span> velocity field of the Earth's surface, and apply it to the Sierra Nevada Mountain range in the western United States. Starting with <span class="hlt">vertical</span> position time series from Global Positioning System (GPS) stations, we first estimate <span class="hlt">vertical</span> velocities using the MIDAS robust trend estimator, which is insensitive to undocumented steps, outliers, seasonality, and heteroscedasticity. Using the Delaunay triangulation of station locations, we then apply a weighted median spatial filter to remove velocity outliers and enhance signals common to multiple stations. Finally, we interpolate the data using weighted median estimation on a grid. The resulting velocity field is temporally and spatially robust and edges in the field remain sharp. Results from data spanning 5-20 years show that the Sierra Nevada is the most rapid and extensive uplift feature in the western United States, <span class="hlt">rising</span> up to 2 mm/yr along most of the range. The uplift is juxtaposed against domains of subsidence attributable to groundwater withdrawal in California's Central Valley. The uplift boundary is consistently stationary, although uplift is faster over the 2011-2016 period of drought. Uplift patterns are consistent with groundwater extraction and concomitant elastic bedrock uplift, plus slower background tectonic uplift. A discontinuity in the velocity field across the southeastern edge of the Sierra Nevada reveals a contrast in lithospheric strength, suggesting a relationship between late Cenozoic uplift of the southern Sierra Nevada and evolution of the southern Walker Lane.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SPIE10019E..0AZ','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SPIE10019E..0AZ"><span>High efficiency single transverse mode photonic band crystal lasers with low <span class="hlt">vertical</span> divergence</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhao, Shaoyu; Qu, Hongwei; Liu, Yun; Li, Lunhua; Chen, Yang; Zhou, Xuyan; Lin, Yuzhe; Liu, Anjin; Qi, Aiyi; Zheng, Wanhua</p> <p>2016-10-01</p> <p>High efficiency 980 nm longitudinal photonic band crystal (PBC) edge emitting laser diodes are designed and fabricated. The calculated results show that eight periods of Al0.<span class="hlt">1</span>Ga0.9As and Al0.25Ga0.75As layer pairs can reduce the <span class="hlt">vertical</span> far field divergence to 10.6° full width at half maximum (FWHM). The broad area (BA) lasers show a very high internal quantum efficiency ηi of 98% and low internal loss αi of <span class="hlt">1</span>.92 cm-<span class="hlt">1</span>. Ridge waveguide (RW) lasers with 3 mm cavity length and 5um strip width provide 430 mW stable single transverse mode output at 500 mA injection current with power conversion efficiency (PCE) of 47% under continuous wave (CW) mode. A maximum PCE of 50% is obtained at the 300 mA injection current. A very low <span class="hlt">vertical</span> far field divergence of 9.4° is obtained at 100 mA injection. At 500 mA injection, the <span class="hlt">vertical</span> far field divergence increases to 11°, the beam quality factors M2 values are <span class="hlt">1</span>.707 in <span class="hlt">vertical</span> direction and <span class="hlt">1</span>.769 in lateral direction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=gain+AND+function&id=EJ1019617','ERIC'); return false;" href="https://eric.ed.gov/?q=gain+AND+function&id=EJ1019617"><span>The Gains from <span class="hlt">Vertical</span> Scaling</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Briggs, Derek C.; Domingue, Ben</p> <p>2013-01-01</p> <p>It is often assumed that a <span class="hlt">vertical</span> scale is necessary when value-added models depend upon the gain scores of students across two or more points in time. This article examines the conditions under which the scale transformations associated with the <span class="hlt">vertical</span> scaling process would be expected to have a significant impact on normative interpretations…</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_22 --> <div id="page_23" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="441"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29785352','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29785352"><span>Diversity and community structure of marine microbes around the Benham <span class="hlt">Rise</span> underwater plateau, northeastern Philippines.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gajigan, Andrian P; Yñiguez, Aletta T; Villanoy, Cesar L; San Diego-McGlone, Maria Lourdes; Jacinto, Gil S; Conaco, Cecilia</p> <p>2018-01-01</p> <p>Microbes are central to the structuring and functioning of marine ecosystems. Given the remarkable diversity of the ocean microbiome, uncovering marine microbial taxa remains a fundamental challenge in microbial ecology. However, there has been little effort, thus far, to describe the diversity of marine microorganisms in the region of high marine biodiversity around the Philippines. Here, we present data on the taxonomic diversity of bacteria and archaea in Benham <span class="hlt">Rise</span>, Philippines, Western Pacific Ocean, using 16S V4 rRNA gene sequencing. The major bacterial and archaeal phyla identified in the Benham <span class="hlt">Rise</span> are Proteobacteria, Cyanobacteria, Actinobacteria, Bacteroidetes, Marinimicrobia, Thaumarchaeota and, Euryarchaeota. The upper mesopelagic layer exhibited greater microbial diversity and richness compared to surface waters. <span class="hlt">Vertical</span> zonation of the microbial community is evident and may be attributed to physical stratification of the water column acting as a dispersal barrier. Canonical Correspondence Analysis (CCA) recapitulated previously known associations of taxa and physicochemical parameters in the environment, such as the association of oligotrophic clades with low nutrient surface water and deep water clades that have the capacity to oxidize ammonia or nitrite at the upper mesopelagic layer. These findings provide foundational information on the diversity of marine microbes in Philippine waters. Further studies are warranted to gain a more comprehensive picture of microbial diversity within the region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5960264','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5960264"><span>Diversity and community structure of marine microbes around the Benham <span class="hlt">Rise</span> underwater plateau, northeastern Philippines</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Gajigan, Andrian P.; Yñiguez, Aletta T.; Villanoy, Cesar L.; San Diego-McGlone, Maria Lourdes; Jacinto, Gil S.</p> <p>2018-01-01</p> <p>Microbes are central to the structuring and functioning of marine ecosystems. Given the remarkable diversity of the ocean microbiome, uncovering marine microbial taxa remains a fundamental challenge in microbial ecology. However, there has been little effort, thus far, to describe the diversity of marine microorganisms in the region of high marine biodiversity around the Philippines. Here, we present data on the taxonomic diversity of bacteria and archaea in Benham <span class="hlt">Rise</span>, Philippines, Western Pacific Ocean, using 16S V4 rRNA gene sequencing. The major bacterial and archaeal phyla identified in the Benham <span class="hlt">Rise</span> are Proteobacteria, Cyanobacteria, Actinobacteria, Bacteroidetes, Marinimicrobia, Thaumarchaeota and, Euryarchaeota. The upper mesopelagic layer exhibited greater microbial diversity and richness compared to surface waters. <span class="hlt">Vertical</span> zonation of the microbial community is evident and may be attributed to physical stratification of the water column acting as a dispersal barrier. Canonical Correspondence Analysis (CCA) recapitulated previously known associations of taxa and physicochemical parameters in the environment, such as the association of oligotrophic clades with low nutrient surface water and deep water clades that have the capacity to oxidize ammonia or nitrite at the upper mesopelagic layer. These findings provide foundational information on the diversity of marine microbes in Philippine waters. Further studies are warranted to gain a more comprehensive picture of microbial diversity within the region. PMID:29785352</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27315016','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27315016"><span>HIV-<span class="hlt">1</span> <span class="hlt">Vertical</span> Transmission in Zimbabwe in 622 Mother and Infant Pairs: Rethinking the Contribution of Mannose Binding Lectin Deficiency in Africa.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zinyama-Gutsire, Rutendo B L; Christiansen, Michael; Hedley, Paula L; Rusakaniko, Simbarashe; Hagen, Christian; Stray-Pedersen, Babill; Buzdugan, Raluca; Cowan, Frances; Chasela, Charles</p> <p>2016-07-01</p> <p><span class="hlt">Vertical</span> transmission of human immunodeficiency virus (HIV) remains a major global health problem. We assessed the association of mannose binding lectin (MBL) deficiency and <span class="hlt">vertical</span> transmission of HIV. Novel diagnostics would be a major breakthrough in this regard. MBL is a liver-derived protein and a key component of the innate immune system. MBL levels may be classified as normal, intermediate, or deficient in the plasma and can use MBL2 haplotypes as a proxy. These haplotypes comprise polymorphisms in the MBL2 gene and promoter region and are known to result in varying levels of MBL deficiency. MBL deficiency can be defined as presence of A/O and O/O genotypes in the mothers and their children. MBL deficiency leads to defective opsonization activities of the innate immune system and increased susceptibility to several infections, including HIV-<span class="hlt">1</span>. We determined the prevalence of MBL deficiency, using MBL2 haplotypes among 622 HIV-positive Zimbabwean mothers and their children aged 9-18 months old, in relation to the HIV-<span class="hlt">1</span> <span class="hlt">vertical</span> transmission risk. The median age of the mothers was 30 (26-34, interquartile range [IQR]) years, and the babies' median age was 13 (11-15, IQR) months old at the time of enrollment. From the sample of 622 mothers who were HIV-<span class="hlt">1</span> infected, 574 babies were HIV negative and 48 were HIV-<span class="hlt">1</span>-positive babies, giving a transmission rate of 7.7%. MBL2 normal structural allele A and variants B (codon 5 A>G), C (codon 57 A>G), and promoter region SNPs -550(H/L) and -221(X/Y) were detected. Prevalence of haplotype-predicted MBL deficiency was 34% among the mothers and 32% among the children. We found no association between maternal MBL2 deficiency and HIV-<span class="hlt">1</span> transmission to their children. We found no difference in the distribution of HIV-<span class="hlt">1</span> infected and uninfected children between the MBL2 genotypes of the mothers and those of the children. Taken together, the present study in a large sample of mother-infant pairs in Zimbabwe adds to the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70039314','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70039314"><span>Population dynamics of Hawaiian seabird colonies vulnerable to sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Hatfield, Jeff S.; Reynolds, Michelle H.; Seavy, Nathaniel E.; Krause, Crystal M.</p> <p>2012-01-01</p> <p>Globally, seabirds are vulnerable to anthropogenic threats both at sea and on land. Seabirds typically nest colonially and show strong fidelity to natal colonies, and such colonies on low-lying islands may be threatened by sea-level <span class="hlt">rise</span>. We used French Frigate Shoals, the largest atoll in the Hawaiian Archipelago, as a case study to explore the population dynamics of seabird colonies and the potential effects sea-level <span class="hlt">rise</span> may have on these rookeries. We compiled historic observations, a 30-year time series of seabird population abundance, lidar-derived elevations, and aerial imagery of all the islands of French Frigate Shoals. To estimate the population dynamics of 8 species of breeding seabirds on Tern Island from 1980 to 2009, we used a Gompertz model with a Bayesian approach to infer population growth rates, density dependence, process variation, and observation error. All species increased in abundance, in a pattern that provided evidence of density dependence. Great Frigatebirds (Fregata minor), Masked Boobies (Sula dactylatra), Red-tailed Tropicbirds (Phaethon rubricauda), Spectacled Terns (Onychoprion lunatus), and White Terns (Gygis alba) are likely at carrying capacity. Density dependence may exacerbate the effects of sea-level <span class="hlt">rise</span> on seabirds because populations near carrying capacity on an island will be more negatively affected than populations with room for growth. We projected 12% of French Frigate Shoals will be inundated if sea level <span class="hlt">rises</span> <span class="hlt">1</span> m and 28% if sea level <span class="hlt">rises</span> 2 m. Spectacled Terns and shrub-nesting species are especially vulnerable to sea-level <span class="hlt">rise</span>, but seawalls and habitat restoration may mitigate the effects of sea-level <span class="hlt">rise</span>. Losses of seabird nesting habitat may be substantial in the Hawaiian Islands by 2100 if sea levels <span class="hlt">rise</span> 2 m. Restoration of higher-elevation seabird colonies represent a more enduring conservation solution for Pacific seabirds.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1176036','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1176036"><span><span class="hlt">Vertically</span> aligned nanostructure scanning probe microscope tips</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Guillorn, Michael A.; Ilic, Bojan; Melechko, Anatoli V.; Merkulov, Vladimir I.; Lowndes, Douglas H.; Simpson, Michael L.</p> <p>2006-12-19</p> <p>Methods and apparatus are described for cantilever structures that include a <span class="hlt">vertically</span> aligned nanostructure, especially <span class="hlt">vertically</span> aligned carbon nanofiber scanning probe microscope tips. An apparatus includes a cantilever structure including a substrate including a cantilever body, that optionally includes a doped layer, and a <span class="hlt">vertically</span> aligned nanostructure coupled to the cantilever body.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E3SWC..3301042R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E3SWC..3301042R"><span>Trend analysis of modern high-<span class="hlt">rise</span> construction</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Radushinsky, Dmitry; Gubankov, Andrey; Mottaeva, Asiiat</p> <p>2018-03-01</p> <p>The article reviews the main trends of modern high-<span class="hlt">rise</span> construction considered a number of architectural, engineering and technological, economic and image factors that have influenced the intensification of construction of high-<span class="hlt">rise</span> buildings in the 21st century. The key factors of modern high-<span class="hlt">rise</span> construction are identified, which are associated with an attractive image component for businessmen and politicians, with the ability to translate current views on architecture and innovations in construction technologies and the lobbying of relevant structures, as well as the opportunity to serve as an effective driver in the development of a complex of national economy sectors with the achievement of a multiplicative effect. The estimation of the priority nature of participation of foreign architectural bureaus in the design of super-high buildings in Russia at the present stage is given. The issue of economic expediency of construction of high-<span class="hlt">rise</span> buildings, including those with only a residential function, has been investigated. The connection between the construction of skyscrapers as an important component of the image of cities in the marketing of places and territories, the connection of the availability of a high-<span class="hlt">rise</span> center, the City, with the possibilities of attracting a "creative class" and the features of creating a large working space for specialists on the basis of territorial proximity and density of high-<span class="hlt">rise</span> buildings.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6456638-cenozoic-seismic-stratigraphy-sw-bermuda-rise','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6456638-cenozoic-seismic-stratigraphy-sw-bermuda-rise"><span>Cenozoic seismic stratigraphy of the SW Bermuda <span class="hlt">Rise</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Mountain, G.S.; Driscoll, N.W.; Miller, K.G.</p> <p>1985-01-01</p> <p>The seismic Horizon A-Complex (Tucholke, 1979) readily explains reflector patterns observed along the western third of the Bermuda <span class="hlt">Rise</span>; farther east, basement is much more rugged and gravity flows shed from local topographic highs complicate the stratigraphy. Distal turbidites on the southwestern Bermuda <span class="hlt">Rise</span> onlap reflector A* from the west, suggesting early Paleocene mass wasting of the North American margin. Locally erosive bottom currents cut into the middle Eocene section of the SW Bermuda <span class="hlt">Rise</span>; these northward flowing currents preceded those that formed reflector Au along the North American margin near the Eocene-Oligocene boundary. Southward flowing currents swift enough tomore » erode the sea floor and to form reflector Au did not reach as far east as the SW Bermuda <span class="hlt">Rise</span>. Instead, the main effect of these Au currents was to pirate sediment into contour-following geostrophic flows along the North American margin and to deprive the deep basin and the Bermuda <span class="hlt">Rise</span> of sediment transported down-slope. Consequently, post-Eocene sediments away from the margin are fine-grained muds. Deposition of these muds on the SW Bermuda <span class="hlt">Rise</span> was controlled by northward flowing bottom currents. The modern Hatteras Abyssal Plain developed in the late Neogene as turbidites once again onlapped the SW Bermuda <span class="hlt">Rise</span>. Today, these deposits extend farthest east in fracture zone valleys and in the swales between sediment waves. Northward flowing currents continue at present to affect sediment distribution patterns along the western edge of the Bermuda <span class="hlt">Rise</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/862980','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/862980"><span><span class="hlt">Vertically</span> stabilized elongated cross-section tokamak</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Sheffield, George V.</p> <p>1977-01-01</p> <p>This invention provides a <span class="hlt">vertically</span> stabilized, non-circular (minor) cross-section, toroidal plasma column characterized by an external separatrix. To this end, a specific poloidal coil means is added outside a toroidal plasma column containing an endless plasma current in a tokamak to produce a rectangular cross-section plasma column along the equilibrium axis of the plasma column. By elongating the spacing between the poloidal coil means the plasma cross-section is <span class="hlt">vertically</span> elongated, while maintaining <span class="hlt">vertical</span> stability, efficiently to increase the poloidal flux in linear proportion to the plasma cross-section height to achieve a much greater plasma volume than could be achieved with the heretofore known round cross-section plasma columns. Also, <span class="hlt">vertical</span> stability is enhanced over an elliptical cross-section plasma column, and poloidal magnetic divertors are achieved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20050000753&hterms=cat+behavior&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dcat%2Bbehavior','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20050000753&hterms=cat+behavior&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dcat%2Bbehavior"><span>Response of pontomedullary reticulospinal neurons to vestibular stimuli in <span class="hlt">vertical</span> planes. Role in <span class="hlt">vertical</span> vestibulospinal reflexes of the decerebrate cat</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bolton, P. S.; Goto, T.; Schor, R. H.; Wilson, V. J.; Yamagata, Y.; Yates, B. J.</p> <p>1992-01-01</p> <p><span class="hlt">1</span>. To investigate the neural substrate of vestibulospinal reflexes in decerebrate cats, we studied the responses of pontomedullary reticulospinal neurons to natural stimulation of the labyrinth in <span class="hlt">vertical</span> planes. Our principal aim was to determine whether reticulospinal neurons that terminate in, or are likely to give off collaterals to, the upper cervical segments had properties similar to those of the vestibulocollic reflex (VCR). 2. Antidromic stimulation was used to determine whether the neurons projected to the neck, lower cervical, thoracic, or lumbar levels. Dynamics of the responses of spontaneously firing neurons were studied with sinusoidal stimuli delivered at 0.05-<span class="hlt">1</span> Hz and aligned to the plane of body rotation, that produced maximal modulation of the neuron (response vector orientation). Each neuron was assigned a vestibular input classification of otolith, <span class="hlt">vertical</span> canal, otolith + canal, or spatial-temporal convergence (STC). 3. We found, in agreement with previous studies, that the largest fraction of pontomedullary reticulospinal neurons projected to the lumbar cord, and that only a small number ended in the neck segments. Neurons projecting to all levels of the spinal cord had similar responses to labyrinth stimulation. 4. Reticulospinal neurons that received only <span class="hlt">vertical</span> canal inputs were rare (<span class="hlt">1</span> of 67 units). Most reticulospinal neurons (48%) received predominant otolith inputs, 18% received otolith + canal input, and only 9% had STC behavior. These data are in sharp contrast to the results of our previous studies of vestibulospinal neurons. A considerable portion of vestibulospinal neurons receives <span class="hlt">vertical</span> canal input (38%), fewer receive predominantly otolith input (22%), whereas the proportion that have otolith + canal input or STC behavior is similar to our present reticulospinal data. 5. The response vector orientations of our reticulospinal neurons, particularly those with canal inputs (canal, otolith + canal, STC) were predominantly in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20647944','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20647944"><span>Validity and reliability of Optojump photoelectric cells for estimating <span class="hlt">vertical</span> jump height.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Glatthorn, Julia F; Gouge, Sylvain; Nussbaumer, Silvio; Stauffacher, Simone; Impellizzeri, Franco M; Maffiuletti, Nicola A</p> <p>2011-02-01</p> <p><span class="hlt">Vertical</span> jump is one of the most prevalent acts performed in several sport activities. It is therefore important to ensure that the measurements of <span class="hlt">vertical</span> jump height made as a part of research or athlete support work have adequate validity and reliability. The aim of this study was to evaluate concurrent validity and reliability of the Optojump photocell system (Microgate, Bolzano, Italy) with force plate measurements for estimating <span class="hlt">vertical</span> jump height. Twenty subjects were asked to perform maximal squat jumps and countermovement jumps, and flight time-derived jump heights obtained by the force plate were compared with those provided by Optojump, to examine its concurrent (criterion-related) validity (study <span class="hlt">1</span>). Twenty other subjects completed the same jump series on 2 different occasions (separated by <span class="hlt">1</span> week), and jump heights of session <span class="hlt">1</span> were compared with session 2, to investigate test-retest reliability of the Optojump system (study 2). Intraclass correlation coefficients (ICCs) for validity were very high (0.997-0.998), even if a systematic difference was consistently observed between force plate and Optojump (-<span class="hlt">1</span>.06 cm; p < 0.001). Test-retest reliability of the Optojump system was excellent, with ICCs ranging from 0.982 to 0.989, low coefficients of variation (2.7%), and low random errors (±2.81 cm). The Optojump photocell system demonstrated strong concurrent validity and excellent test-retest reliability for the estimation of <span class="hlt">vertical</span> jump height. We propose the following equation that allows force plate and Optojump results to be used interchangeably: force plate jump height (cm) = <span class="hlt">1</span>.02 × Optojump jump height + 0.29. In conclusion, the use of Optojump photoelectric cells is legitimate for field-based assessments of <span class="hlt">vertical</span> jump height.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20040031848','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20040031848"><span>High-Resolution Simulation of Hurricane Bonnie (1998). Part <span class="hlt">1</span>; The Organization of <span class="hlt">Vertical</span> Motion</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Braun, Scott A.; Montgomery, Michael T.; Pu, Zhaoxia</p> <p>2003-01-01</p> <p>Hurricanes are well known for their strong winds and heavy rainfall, particularly in the intense rainband (eyewall) surrounding the calmer eye of the storm. In some hurricanes, the rainfall is distributed evenly around the eye so that it has a donut shape on radar images. In other cases, the rainfall is concentrated on one side of the eyewall and nearly absent on the other side and is said to be asymmetric. This study examines how the <span class="hlt">vertical</span> air motions that produce the rainfall are distributed within the eyewall of an asymmetric hurricane and the factors that cause this pattern of rainfall. We use a sophisticated numerical forecast model to simulate Hurricane Bonnie, which occurred in late August of 1998 during a special NASA field experiment designed to study hurricanes. The simulation results suggest that <span class="hlt">vertical</span> wind shear (a rapid change in wind speed or direction with height) caused the asymmetric rainfall and <span class="hlt">vertical</span> air motion patterns by tilting the hurricane vortex and favoring upward air motions in the direction of tilt. Although the rainfall in the hurricane eyewall may surround more than half of the eye, the updrafts that produce the rainfall are concentrated in very small-scale, intense updraft cores that occupy only about 10% of the eyewall area. The model simulation suggests that the timing and location of individual updraft cores are controlled by intense, small-scale vortices (regions of rapidly swirling flow) in the eyewall and that the updrafts form when the vortices encounter low-level air moving into the eyewall.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70031495','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70031495"><span><span class="hlt">Vertical</span> motions of the Puerto Rico Trench and Puerto Rico and their cause</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>ten Brink, Uri S.</p> <p>2005-01-01</p> <p>The Puerto Rico trench exhibits great water depth, an extremely low gravity anomaly, and a tilted carbonate platform between (reconstructed) elevations of +1300 m and -4000 m. I argue that these features are manifestations of large <span class="hlt">vertical</span> movements of a segment of the Puerto Rico trench, its forearc, and the island of Puerto Rico that took place 3.3 m.y. ago over a time period as short as 14-40 kyr. I explain these <span class="hlt">vertical</span> movements by a sudden increase in the slab's descent angle that caused the trench to subside and the island to <span class="hlt">rise</span>. The increased dip could have been caused by shearing or even by a complete tear of the descending North American slab, although the exact nature of this deformation is unknown. The rapid (14-40 kyr) and uniform tilt along a 250 km long section of the trench is compatible with scales of mantle flow and plate bending. The proposed shear zone or tear is inferred from seismic, morphological, and gravity observations to start at the trench at 64.5??W and trend southwestwardly toward eastern Puerto Rico. The tensile stresses necessary to deform or tear the slab could have been generated by increased curvature of the trench following a counterclockwise rotation of the upper plate and by the subduction of a large seamount.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20010016292&hterms=TRANSPORT+AIR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DTRANSPORT%2BAIR','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20010016292&hterms=TRANSPORT+AIR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DTRANSPORT%2BAIR"><span>Age-of-Air, Tape Recorder, and <span class="hlt">Vertical</span> Transport Schemes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lin, S.-J.; Einaudi, Franco (Technical Monitor)</p> <p>2000-01-01</p> <p>A numerical-analytic investigation of the impacts of <span class="hlt">vertical</span> transport schemes on the model simulated age-of-air and the so-called 'tape recorder' will be presented using an idealized <span class="hlt">1</span>-D column transport model as well as a more realistic 3-D dynamical model. By comparing to the 'exact' solutions of 'age-of-air' and the 'tape recorder' obtainable in the <span class="hlt">1</span>-D setting, useful insight is gained on the impacts of numerical diffusion and dispersion of numerical schemes used in global models. Advantages and disadvantages of Eulerian, semi-Lagrangian, and Lagrangian transport schemes will be discussed. <span class="hlt">Vertical</span> resolution requirement for numerical schemes as well as observing systems for capturing the fine details of the 'tape recorder' or any upward propagating wave-like structures can potentially be derived from the <span class="hlt">1</span>-D analytic model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4696803','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4696803"><span>The Subjective Visual <span class="hlt">Vertical</span> and the Subjective Haptic <span class="hlt">Vertical</span> Access Different Gravity Estimates</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Fraser, Lindsey E.; Makooie, Bobbak; Harris, Laurence R.</p> <p>2015-01-01</p> <p>The subjective visual <span class="hlt">vertical</span> (SVV) and the subjective haptic <span class="hlt">vertical</span> (SHV) both claim to probe the underlying perception of gravity. However, when the body is roll tilted these two measures evoke different patterns of errors with SVV generally becoming biased towards the body (A-effect, named for its discoverer, Hermann Rudolph Aubert) and SHV remaining accurate or becoming biased away from the body (E-effect, short for Entgegengesetzt-effect, meaning “opposite”, i.e., opposite to the A-effect). We compared the two methods in a series of five experiments and provide evidence that the two measures access two different but related estimates of gravitational <span class="hlt">vertical</span>. Experiment <span class="hlt">1</span> compared SVV and SHV across three levels of whole-body tilt and found that SVV showed an A-effect at larger tilts while SHV was accurate. Experiment 2 found that tilting either the head or the trunk independently produced an A-effect in SVV while SHV remained accurate when the head was tilted on an upright body but showed an A-effect when the body was tilted below an upright head. Experiment 3 repeated these head/body configurations in the presence of vestibular noise induced by using disruptive galvanic vestibular stimulation (dGVS). dGVS abolished both SVV and SHV A-effects while evoking a massive E-effect in the SHV head tilt condition. Experiments 4 and 5 show that SVV and SHV do not combine in an optimally statistical fashion, but when vibration is applied to the dorsal neck muscles, integration becomes optimal. Overall our results suggest that SVV and SHV access distinct underlying gravity percepts based primarily on head and body position information respectively, consistent with a model proposed by Clemens and colleagues. PMID:26716835</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26716835','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26716835"><span>The Subjective Visual <span class="hlt">Vertical</span> and the Subjective Haptic <span class="hlt">Vertical</span> Access Different Gravity Estimates.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fraser, Lindsey E; Makooie, Bobbak; Harris, Laurence R</p> <p>2015-01-01</p> <p>The subjective visual <span class="hlt">vertical</span> (SVV) and the subjective haptic <span class="hlt">vertical</span> (SHV) both claim to probe the underlying perception of gravity. However, when the body is roll tilted these two measures evoke different patterns of errors with SVV generally becoming biased towards the body (A-effect, named for its discoverer, Hermann Rudolph Aubert) and SHV remaining accurate or becoming biased away from the body (E-effect, short for Entgegengesetzt-effect, meaning "opposite", i.e., opposite to the A-effect). We compared the two methods in a series of five experiments and provide evidence that the two measures access two different but related estimates of gravitational <span class="hlt">vertical</span>. Experiment <span class="hlt">1</span> compared SVV and SHV across three levels of whole-body tilt and found that SVV showed an A-effect at larger tilts while SHV was accurate. Experiment 2 found that tilting either the head or the trunk independently produced an A-effect in SVV while SHV remained accurate when the head was tilted on an upright body but showed an A-effect when the body was tilted below an upright head. Experiment 3 repeated these head/body configurations in the presence of vestibular noise induced by using disruptive galvanic vestibular stimulation (dGVS). dGVS abolished both SVV and SHV A-effects while evoking a massive E-effect in the SHV head tilt condition. Experiments 4 and 5 show that SVV and SHV do not combine in an optimally statistical fashion, but when vibration is applied to the dorsal neck muscles, integration becomes optimal. Overall our results suggest that SVV and SHV access distinct underlying gravity percepts based primarily on head and body position information respectively, consistent with a model proposed by Clemens and colleagues.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title21-vol2/pdf/CFR-2011-title21-vol2-sec137-290.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title21-vol2/pdf/CFR-2011-title21-vol2-sec137-290.pdf"><span>21 CFR 137.290 - Self-<span class="hlt">rising</span> yellow corn meal.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-04-01</p> <p>... 21 Food and Drugs 2 2011-04-01 2011-04-01 false Self-<span class="hlt">rising</span> yellow corn meal. 137.290 Section 137... Cereal Flours and Related Products § 137.290 Self-<span class="hlt">rising</span> yellow corn meal. Self-<span class="hlt">rising</span> yellow corn meal conforms to the definition and standard of identity prescribed by § 137.270 for self-<span class="hlt">rising</span> white corn meal...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title21-vol2/pdf/CFR-2010-title21-vol2-sec137-290.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title21-vol2/pdf/CFR-2010-title21-vol2-sec137-290.pdf"><span>21 CFR 137.290 - Self-<span class="hlt">rising</span> yellow corn meal.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-04-01</p> <p>... 21 Food and Drugs 2 2010-04-01 2010-04-01 false Self-<span class="hlt">rising</span> yellow corn meal. 137.290 Section 137... Cereal Flours and Related Products § 137.290 Self-<span class="hlt">rising</span> yellow corn meal. Self-<span class="hlt">rising</span> yellow corn meal conforms to the definition and standard of identity prescribed by § 137.270 for self-<span class="hlt">rising</span> white corn meal...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22078348-rise-fall-star-formation-histories-blue-galaxies-redshifts','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22078348-rise-fall-star-formation-histories-blue-galaxies-redshifts"><span>THE <span class="hlt">RISE</span> AND FALL OF THE STAR FORMATION HISTORIES OF BLUE GALAXIES AT REDSHIFTS 0.2 < z < <span class="hlt">1</span>.4</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Pacifici, Camilla; Kassin, Susan A.; Gardner, Jonathan P.</p> <p>2013-01-01</p> <p>Popular cosmological scenarios predict that galaxies form hierarchically from the merger of many progenitors, each with their own unique star formation history (SFH). We use a sophisticated approach to constrain the SFHs of 4517 blue (presumably star-forming) galaxies with spectroscopic redshifts in the range 0.2 < z < <span class="hlt">1</span>.4 from the All-Wavelength Extended Groth Strip International Survey. This consists in the Bayesian analysis of the observed galaxy spectral energy distributions with a comprehensive library of synthetic spectra assembled using realistic, hierarchical star formation, and chemical enrichment histories from cosmological simulations. We constrain the SFH of each galaxy in our samplemore » by comparing the observed fluxes in the B, R, I, and K{sub s} bands and rest-frame optical emission-line luminosities with those of one million model spectral energy distributions. We explore the dependence of the resulting SFHs on galaxy stellar mass and redshift. We find that the average SFHs of high-mass galaxies <span class="hlt">rise</span> and fall in a roughly symmetric bell-shaped manner, while those of low-mass galaxies <span class="hlt">rise</span> progressively in time, consistent with the typically stronger activity of star formation in low-mass compared to high-mass galaxies. For galaxies of all masses, the star formation activity <span class="hlt">rises</span> more rapidly at high than at low redshift. These findings imply that the standard approximation of exponentially declining SFHs widely used to interpret observed galaxy spectral energy distributions may not be appropriate to constrain the physical parameters of star-forming galaxies at intermediate redshifts.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950037899&hterms=vertical+height&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dvertical%2Bheight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950037899&hterms=vertical+height&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dvertical%2Bheight"><span><span class="hlt">Vertical</span> velocity in oceanic convection off tropical Australia</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lucas, Christopher; Zipser, Edward J.; Lemone, Margaret A.</p> <p>1994-01-01</p> <p>Time series of <span class="hlt">1</span>-Hz <span class="hlt">vertical</span> velocity data collected during aircraft penetrations of oceanic cumulonimbus clouds over the western Pacific warm pool as part of the Equatorial Mesoscale Experiment (EMEX) are analyzed for updraft and downdraft events called cores. An updraft core is defined as occurring whenever the <span class="hlt">vertical</span> velocity exceeds <span class="hlt">1</span> m/sec for at least 500 m. A downdraft core is defined analogously. Over 19,000 km of straight and level flight legs are used in the analysis. Five hundred eleven updraft cores and 253 downdraft cores are included in the dataset. Core properties are summarized as distributions of average and maximum <span class="hlt">vertical</span> velocity, diameter, and mass flux in four altitude intervals between 0.2 and 5.8 km. Distributions are approximately lognormal at all levels. Examination of the variation of the statistics with height suggests a maximum in <span class="hlt">vertical</span> velocity between 2 and 3 km; slightly lower or equal <span class="hlt">vertical</span> velocity is indicated at 5 km. Near the freezing level, virtual temperature deviations are found to be slightly positive for both updraft and downdraft cores. The excess in updraft cores is much smaller than that predicted by parcel theory. Comparisons with other studies that use the same analysis technique reveal that EMEX cores have approximately the same strength as cores of other oceanic areas, despite warmer sea surface temperatures. Diameter and mass flux are greater than those in the Global Atmospheric Research Program (GATE) but smaller than those in hurricane rainbands. Oceanic cores are much weaker and appear to be slightly smaller than those observed over land during the Thunderstorm Project. The markedly weaker oceanic <span class="hlt">vertical</span> velocities below 5.8 km (compared to the continental cores) cannot be attributed to smaller total convective available potential energy or to very high water loading. Rather, it is suggested that water loading, although less than adiabatic, is more effective in reducing buoyancy of oceanic cores</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5398905','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5398905"><span>Can <span class="hlt">Vertical</span> Integration Reduce Hospital Readmissions? A Difference-in-Differences Approach</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Fernandes, Óscar B.; Marques, Ana Patrícia; Moita, Bruno; Sarmento, João; Santana, Rui</p> <p>2017-01-01</p> <p>Background: <span class="hlt">Vertical</span> integration is expected to improve communication and coordination between inpatient care and care after discharge. Despite being used across health systems worldwide, evidence about its impact on readmissions is sparse and contradictory. Objective: To assess the impact of <span class="hlt">vertical</span> integration on hospital readmissions. Research Design, Subjects, and Measures: Using difference-in-differences we compared readmissions before and after <span class="hlt">vertical</span> integration in 6 Portuguese hospitals for years 2004–2013. A control group with 6 similar hospitals not integrated was utilized. Considered outcome was 30-day unplanned readmission. We used logistic regression at the admission level and accounted for patients’ risk factors using claims data. Analyses for each hospital and selected conditions were also run. Results: Our results suggest that readmissions decreased overall after <span class="hlt">vertical</span> integration [odds ratio (OR)=0.900; 95% confidence interval (CI), 0.812–0.997]. Hospital analysis indicated that there was no impact for 2 hospitals (OR=0.960; 95% CI, 0.848–<span class="hlt">1</span>.087 and OR=0.944; 95% CI, 0.857–<span class="hlt">1</span>.038), and a positive effect in 4 hospitals (greatest effect: OR=0.811; 95% CI, 0.736–0.894). A positive evolution was observed for a limited number of conditions, with better results for diabetes with complications (OR=0.689; 95% CI, 0.525–0.904), but no impact regarding congestive heart failure (OR=<span class="hlt">1</span>.067; 95% CI, 0.827–<span class="hlt">1</span>.377). Conclusions: Merging acute and primary care providers was associated with reduced readmissions, even though improvements were not found for all institutions or condition-specific groups. There are still challenges to be addressed regarding the success of <span class="hlt">vertical</span> integration in reducing 30-day hospital readmissions. PMID:28403012</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_23 --> <div id="page_24" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="461"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16404243','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16404243"><span>Effect of pregnancy and breast-feeding on <span class="hlt">vertical</span> mammaplasty.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cruz-Korchin, Norma; Korchin, Leo</p> <p>2006-01-01</p> <p>A retrospective study was performed to evaluate the effect of pregnancy and breast-feeding on the breasts of women who had undergone <span class="hlt">vertical</span> reduction mammaplasty. The study group consisted of 57 women who had pregnancies after their <span class="hlt">vertical</span> reduction mammaplasty. Of this group, 24 breast-fed. The control group consisted of 103 women who had <span class="hlt">vertical</span> mammaplasty but no subsequent pregnancies. An evaluation form was completed that included the age, body mass index, amount of tissue removed per breast, pregnancies after the mammaplasty, history of breast-feeding, and breast measurements. All patients had breast measurements routinely performed postoperatively at 2 weeks and again at 2 years. The following measurements were obtained: mid-clavicle to nipple, and inframammary fold to inferior areola. No significant difference was found between the control and the study group regarding age (27 +/- 12 versus 29 +/- 10), body mass index (26 +/- 5 versus 27 +/- 4), and grams of tissue excised per breast (610 +/- 201 versus 598 +/- 279). The breast measurement from the mid-clavicle to nipple was not significantly altered by pregnancy with or without breast-feeding (p > 0.05). The distance between the inframammary fold and the inferior margin of the areola was significantly (p < 0.05) increased by pregnancy both with breast-feeding (4.<span class="hlt">1</span> +/- 2.3 cm) and without (3.5 +/- 2.6 cm) when compared with the control group (<span class="hlt">1</span>.2 +/- <span class="hlt">1</span>.5 cm). The <span class="hlt">vertical</span> mammaplasty has less tendency for pseudoptosis (bottoming out), but the alterations of breast volume brought about by pregnancy and breast-feeding may affect the final outcome of even this good reduction mammaplasty method.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21066061-bottlenecks-aggravate-rising-construction-costs','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21066061-bottlenecks-aggravate-rising-construction-costs"><span>Bottlenecks aggravate <span class="hlt">rising</span> construction costs</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>NONE</p> <p>2008-05-15</p> <p><span class="hlt">Rising</span> demand for power in developing countries combined with concerns about carbon emissions from coal-fired power plants in developed countries have created a bonanza for carbon-light technologies, including nuclear, renewables and natural gas plants. This, in turn, has put upward pressure on the price of natural gas in key markets while resulting in shortages in critical components for building renewables and nuclear reactors. Globalization of the power industry means that pressures in one segment or one region translate into shortages and <span class="hlt">rising</span> prices everywhere else.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17401560','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17401560"><span><span class="hlt">Vertical</span> migration of aggregated aerobic and anaerobic ammonium oxidizers enhances oxygen uptake in a stagnant water layer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Vlaeminck, Siegfried E; Dierick, Katleen; Boon, Nico; Verstraete, Willy</p> <p>2007-07-01</p> <p>Ammonium can be removed as dinitrogen gas by cooperating aerobic and anaerobic ammonium-oxidizing bacteria (AerAOB and AnAOB). The goal of this study was to verify putative mutual benefits for aggregated AerAOB and AnAOB in a stagnant freshwater environment. In an ammonium fed water column, the biological oxygen consumption rate was, on average, 76 kg O(2) ha(-<span class="hlt">1</span>) day(-<span class="hlt">1</span>). As the oxygen transfer rate of an abiotic control column was only 17 kg O(2) ha(-<span class="hlt">1</span>) day(-<span class="hlt">1</span>), biomass activity enhanced the oxygen transfer. Increasing the AnAOB gas production increased the oxygen consumption rate with more than 50% as a result of enhanced <span class="hlt">vertical</span> movement of the biomass. The coupled decrease in dissolved oxygen concentration increased the diffusional oxygen transfer from the atmosphere in the water. Physically preventing the biomass from <span class="hlt">rising</span> to the upper water layer instantaneously decreased oxygen and ammonium consumption and even led to the occurrence of some sulfate reduction. Floating of the biomass was further confirmed to be beneficial, as this allowed for the development of a higher AerAOB and AnAOB activity, compared to settled biomass. Overall, the results support mutual benefits for aggregated AerAOB and AnAOB, derived from the biomass uplifting effect of AnAOB gas production.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20160004095&hterms=sea&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dsea','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20160004095&hterms=sea&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dsea"><span>Evaluation of Dynamic Coastal Response to Sea-level <span class="hlt">Rise</span> Modifies Inundation Likelihood</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lentz, Erika E.; Thieler, E. Robert; Plant, Nathaniel G.; Stippa, Sawyer R.; Horton, Radley M.; Gesch, Dean B.</p> <p>2016-01-01</p> <p>Sea-level <span class="hlt">rise</span> (SLR) poses a range of threats to natural and built environments, making assessments of SLR-induced hazards essential for informed decision making. We develop a probabilistic model that evaluates the likelihood that an area will inundate (flood) or dynamically respond (adapt) to SLR. The broad-area applicability of the approach is demonstrated by producing 30x30m resolution predictions for more than 38,000 sq km of diverse coastal landscape in the northeastern United States. Probabilistic SLR projections, coastal elevation and <span class="hlt">vertical</span> land movement are used to estimate likely future inundation levels. Then, conditioned on future inundation levels and the current land-cover type, we evaluate the likelihood of dynamic response versus inundation. We find that nearly 70% of this coastal landscape has some capacity to respond dynamically to SLR, and we show that inundation models over-predict land likely to submerge. This approach is well suited to guiding coastal resource management decisions that weigh future SLR impacts and uncertainty against ecological targets and economic constraints.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15.2389S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15.2389S"><span>Comparison of upconing under <span class="hlt">vertical</span> and horizontal wells in freshwater lenses: sand-box experiments and numerical modeling</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stoeckl, Leonard; Stefan, Loeffler; Houben, Georg</p> <p>2013-04-01</p> <p>Freshwater lenses on islands and in inland areas are often the primary freshwater resource there. The fragile equilibrium between saline and fresh groundwater can be disrupted by excessive pumping, leading to an upward migration of the saline water underneath the well. Sand-box experiments were conducted to compare the upconing at <span class="hlt">vertical</span> and horizontal wells pumping from a freshwater lens. Results were then compared to numerical simulations. To simulate the cross-section of an "infinite strip island", an acrylic box with a spacing of 5 cm was filled with coarse sand. After saturating the model with degassed saltwater from bottom to top, freshwater recharge was applied from above. By coloring the infiltrating freshwater with different tracer colors using uranine and indigotine we were able to visualize flow paths during pumping. A horizontal and a <span class="hlt">vertical</span> well were placed at the left and right side of the symmetric island. Both had equal diameter, screen length, depth of placement, and distance to shore. Three increasing pumping rates were applied to each well successively and the electrical conductivity of the abstracted water was continuously measured using a through-flow cell. Results show that no saltwater entered the wells when pumping at the lowest rate. Still, slight saltwater upconing and a shift of the freshwater divide in the island were observed. At the second rate a clear saltwater breakthrough into the <span class="hlt">vertical</span> well occurred, while the electrical conductivity remained nearly unchanged in the horizontal well. Applying the third (highest) abstraction rate to each of the wells saltwater entered both wells, exceeding drinking water standards in the <span class="hlt">vertical</span> well. The described behavior indicates the advantage of horizontal over <span class="hlt">vertical</span> wells on islands and in coastal zones prone to saltwater up-coning. Numerical simulations show similar patterns, even though deviations exist between the second and the third pumping rate, which are under and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008APS..MARH36002K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008APS..MARH36002K"><span>A Low Temperature Scanning Force Microscope with a <span class="hlt">Vertical</span> Cantilever and Interferometric Detection Scheme</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, Jeehoon; Williams, T. L.; Chu, Sang Lin; Korre, Hasan; Chalfin, Max; Hoffman, J. E.</p> <p>2008-03-01</p> <p>We have developed a fiber-optic interferometry system with a <span class="hlt">vertical</span> cantilever for scanning force microscopy. A lens, mounted on a Pan-type walker, was used to collect the interference signal in the cavity between the cantilever and the single mode fiber. This <span class="hlt">vertical</span> geometry has several advantages: (<span class="hlt">1</span>) it is directly sensitive to lateral forces; (2) low spring constant <span class="hlt">vertical</span> cantilevers may allow increased force sensitivity by solving the ``snap-in'' problem that occurs with soft horizontal cantilevers. We have sharpened <span class="hlt">vertical</span> cantilevers by focused ion beam (FIB), achieving a tip radius of 20 nm. We will show test results of a magnetic force microscope (MFM) with this <span class="hlt">vertical</span> cantilever system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23480272','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23480272"><span>Temperature <span class="hlt">rises</span> during application of Er:YAG laser under different primary dentin thicknesses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hubbezoglu, Ihsan; Unal, Murat; Zan, Recai; Hurmuzlu, Feridun</p> <p>2013-05-01</p> <p>The present study investigated the effects of the Er:YAG laser's different pulse repetition rates on temperature <span class="hlt">rise</span> under various primary dentin thicknesses. The Er:YAG laser can be used for restorative approaches in clinics and is used to treat dental caries. There are some reports that explain the temperature <span class="hlt">rise</span> effect of the Er:YAG laser. Recently, the Er:YAG laser has been found to play an important role in temperature <span class="hlt">rises</span> during the application on dentin. Caries-free primary mandibular molars were prepared to obtain dentin discs with 0.5, <span class="hlt">1</span>, <span class="hlt">1</span>.5, and 2 mm thicknesses (n=10). These discs were placed between the Teflon mold cylinders of a temperature test apparatus. We preferred three pulse repetition rates of 10, 15, and 20 Hz with an energy density of 12.7 J/cm2 and a 230 μs pulse duration. All dentin discs were irradiated for 30 sec by the Er:YAG laser. Temperature <span class="hlt">rises</span> were recorded using an L-type thermocouple and universal data loggers/scanners (E-680, Elimko Co., Turkey). Data were analyzed by two-way ANOVA and Tukey tests. Whereas the lowest temperature <span class="hlt">rise</span> (0.44±0.09 °C) was measured from a 10 Hz pulse repetition rate at a dentin thickness of 2 mm, the highest temperature <span class="hlt">rise</span> (3.86±0.43 °C) was measured from a 20 Hz pulse repetition rate at a 0.5 mm dentin thickness. Temperature <span class="hlt">rise</span> did not reach critical value for pulpal injury in any primary dentin thicknesses irradiated by a high repetition rate of the Er:YAG laser.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27653154','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27653154"><span>Validation of the iPhone app using the force platform to estimate <span class="hlt">vertical</span> jump height.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Carlos-Vivas, Jorge; Martin-Martinez, Juan P; Hernandez-Mocholi, Miguel A; Perez-Gomez, Jorge</p> <p>2018-03-01</p> <p><span class="hlt">Vertical</span> jump performance has been evaluated with several devices: force platforms, contact mats, Vertec, accelerometers, infrared cameras and high-velocity cameras; however, the force platform is considered the gold standard for measuring <span class="hlt">vertical</span> jump height. The purpose of this study was to validate an iPhone app called My Jump, that measures <span class="hlt">vertical</span> jump height by comparing it with other methods that use the force platform to estimate <span class="hlt">vertical</span> jump height, namely, <span class="hlt">vertical</span> velocity at take-off and time in the air. A total of 40 sport sciences students (age 21.4±<span class="hlt">1</span>.9 years) completed five countermovement jumps (CMJs) over a force platform. Thus, 200 CMJ heights were evaluated from the <span class="hlt">vertical</span> velocity at take-off and the time in the air using the force platform, and from the time in the air with the My Jump mobile application. The height obtained was compared using the intraclass correlation coefficient (ICC). Correlation between APP and force platform using the time in the air was perfect (ICC=<span class="hlt">1</span>.000, P<0.001). Correlation between APP and force platform using the <span class="hlt">vertical</span> velocity at take-off was also very high (ICC=0.996, P<0.001), with an error margin of 0.78%. Therefore, these results showed that application, My Jump, is an appropriate method to evaluate the <span class="hlt">vertical</span> jump performance; however, <span class="hlt">vertical</span> jump height is slightly overestimated compared with that of the force platform.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880062503&hterms=vertical+height&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dvertical%2Bheight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880062503&hterms=vertical+height&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dvertical%2Bheight"><span>Observations of <span class="hlt">vertical</span> velocities in the tropical upper troposphere and lower stratosphere using the Arecibo 430-MHz radar</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cornish, C. R.</p> <p>1988-01-01</p> <p>The first clear-air observations of <span class="hlt">vertical</span> velocities in the tropical upper troposphere and lower stratosphere (8-22 km) using the Arecibo 430-MHz radar are presented. Oscillations in the <span class="hlt">vertical</span> velocity near the Brunt-Vaisala period are observed in the lower stratosphere during the 12-hour observation period. Frequency power spectra from the <span class="hlt">vertical</span> velocity time series show a slope between -0.5 and -<span class="hlt">1</span>.0. <span class="hlt">Vertical</span> wave number spectra computed from the height profiles of <span class="hlt">vertical</span> velocities have slopes between -<span class="hlt">1</span>.0 and -<span class="hlt">1</span>.5. These observed slopes do not agree well with the slopes of +<span class="hlt">1</span>/3 and -2.5 for frequency and <span class="hlt">vertical</span> wave number spectra, respectively, predicted by a universal gravity-wave spectrum model. The spectral power of wave number spectra of a radial beam directed 15 deg off-zenith is enhanced by an order of magnitude over the spectral power levels of the <span class="hlt">vertical</span> beam. This enhancement suggests that other geophysical processes besides gravity waves are present in the horizontal flow. The steepening of the wave number spectrum of the off-<span class="hlt">vertical</span> beam in the lower stratosphere to near -2.0 is attributed to a quasi-inertial period wave, which was present in the horizontal flow during the observation period.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017E%26PSL.457..292H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017E%26PSL.457..292H"><span><span class="hlt">Vertical</span> tectonics at an active continental margin</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Houlié, N.; Stern, T. A.</p> <p>2017-01-01</p> <p>Direct observations of <span class="hlt">vertical</span> movements of the earth's surface are now possible with space-based GPS networks, and have applications to resources, hazards and tectonics. Here we present data on <span class="hlt">vertical</span> movements of the Earth's surface in New Zealand, computed from the processing of GPS data collected between 2000 and 2015 by 189 permanent GPS stations. We map the geographical variation in <span class="hlt">vertical</span> rates and show how these variations are explicable within a tectonic framework of subduction, volcanic activity and slow slip earthquakes. Subsidence of >3 mm/yr is observed along southeastern North Island and is interpreted to be due to the locked segment of the Hikurangi subduction zone. Uplift of <span class="hlt">1</span>-3 mm/yr further north along the margin of the eastern North Island is interpreted as being due to the plate interface being unlocked and underplating of sediment on the subduction thrust. The Volcanic Plateau of the central North Island is being uplifted at about <span class="hlt">1</span> mm/yr, which can be explained by basaltic melts being injected in the active mantle-wedge at a rate of ∼6 mm/yr. Within the Central Volcanic Region there is a 250 km2 area that subsided between 2005 and 2012 at a rate of up to 14 mm/yr. Time series from the stations located within and near the zone of subsidence show a strong link between subsidence, adjacent uplift and local earthquake swarms.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=mutual+AND+fund&pg=3&id=EJ247851','ERIC'); return false;" href="https://eric.ed.gov/?q=mutual+AND+fund&pg=3&id=EJ247851"><span><span class="hlt">Rising</span> College Costs.</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>USA Today, 1981</p> <p>1981-01-01</p> <p>Focuses on ways in which parents of school-age children can offset the <span class="hlt">rising</span> costs of college, including encouraging students to get summer and part-time jobs, putting savings toward students' education in accounts in students' names to save taxes, investigating cooperative work/education plans, and investing in mutual funds. (DB)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFMED13B..08G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFMED13B..08G"><span>Online Citizen Science with Clickworkers & MRO Hi<span class="hlt">RISE</span> E/PO</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gulick, V. C.; Deardorff, G.; Kanefsky, B.; HiRISE Science Team</p> <p>2010-12-01</p> <p>The High-Resolution Imaging Science Experiment’s E/PO has fielded several online citizen science projects. Our efforts are guided by Hi<span class="hlt">RISE</span> E/PO’s philosophy of providing innovative opportunities for students and the public to participate in the scientific discovery process. Hi<span class="hlt">RISE</span> Clickworkers, a follow-on to the original Clickworkers crater identification and size diameter marking website, provides an opportunity for the public to identify & mark over a dozen landform feature types in Hi<span class="hlt">RISE</span> images, including dunes, gullies, patterned ground, wind streaks, boulders, craters, layering, volcanoes, etc. In Hi<span class="hlt">RISE</span> Clickworkers, the contributor views several sample images showing variations of different landforms, and simply marks all the landform types they could spot while looking at a small portion of a Hi<span class="hlt">RISE</span> image. Contributors then submit their work & once validated by comparison to the output of other participants, results are then added to geologic feature databases. Scientists & others will eventually be able to query these databases for locations of particular geologic features in the Hi<span class="hlt">RISE</span> images. Participants can also mark other features that they find intriguing for the Hi<span class="hlt">RISE</span> camera to target. The original Clickworkers website pilot study ran from November 2000 until September 2001 (Kanefsky et al., 2001, LPSC XXXII). It was among the first online Citizen Science efforts for planetary science. In its pilot study, we endeavored to answer two questions: <span class="hlt">1</span>) Was the public willing & able to help science, & 2) Can the public produce scientifically useful results? Since its inception over 3,500,000 craters have been identified, & over 350,000 of these craters have been classified. Over 2 million of these craters were marked on Viking Orbiter image mosaics, nearly 800,000 craters were marked on Mars Orbiter Camera (MOC) images. Note that these are not counts of distinct craters. For example, each crater in the Viking orbiter images was counted by about 50</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3491795','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3491795"><span>Neglected locked <span class="hlt">vertical</span> patellar dislocation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Gupta, Rakesh Kumar; Gupta, Vinay; Sangwan, Sukhbir Singh; Kamboj, Pradeep</p> <p>2012-01-01</p> <p>Patellar dislocations occurring about the <span class="hlt">vertical</span> and horizontal axis are rare and irreducible. The neglected patellar dislocation is still rarer. We describe the clinical presentation and management of a case of neglected <span class="hlt">vertical</span> patellar dislocation in a 6 year-old boy who sustained an external rotational strain with a laterally directed force to his knee. Initially the diagnosis was missed and 2 months later open reduction was done. The increased tension generated by the rotation of the lateral extensor retinaculum kept the patella locked in the lateral gutter even with the knee in full extension. Traumatic patellar dislocation with rotation around a <span class="hlt">vertical</span> axis has been described earlier, but no such neglected case has been reported to the best of our knowledge. PMID:23162154</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018Geomo.304...64L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018Geomo.304...64L"><span>Implications of sea-level <span class="hlt">rise</span> in a modern carbonate ramp setting</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lokier, Stephen W.; Court, Wesley M.; Onuma, Takumi; Paul, Andreas</p> <p>2018-03-01</p> <p>This study addresses a gap in our understanding of the effects of sea-level <span class="hlt">rise</span> on the sedimentary systems and morphological development of recent and ancient carbonate ramp settings. Many ancient carbonate sequences are interpreted as having been deposited in carbonate ramp settings. These settings are poorly-represented in the Recent. The study documents the present-day transgressive flooding of the Abu Dhabi coastline at the southern shoreline of the Arabian/Persian Gulf, a carbonate ramp depositional system that is widely employed as a Recent analogue for numerous ancient carbonate systems. Fourteen years of field-based observations are integrated with historical and recent high-resolution satellite imagery in order to document and assess the onset of flooding. Predicted rates of transgression (i.e. landward movement of the shoreline) of 2.5 m yr- <span class="hlt">1</span> (± 0.2 m yr- <span class="hlt">1</span>) based on global sea-level <span class="hlt">rise</span> alone were far exceeded by the flooding rate calculated from the back-stepping of coastal features (10-29 m yr- <span class="hlt">1</span>). This discrepancy results from the dynamic nature of the flooding with increased water depth exposing the coastline to increased erosion and, thereby, enhancing back-stepping. A non-accretionary transgressive shoreline trajectory results from relatively rapid sea-level <span class="hlt">rise</span> coupled with a low-angle ramp geometry and a paucity of sediments. The flooding is represented by the landward migration of facies belts, a range of erosive features and the onset of bioturbation. Employing Intergovernmental Panel on Climate Change (Church et al., 2013) predictions for 21st century sea-level <span class="hlt">rise</span>, and allowing for the post-flooding lag time that is typical for the start-up of carbonate factories, it is calculated that the coastline will continue to retrograde for the foreseeable future. Total passive flooding (without considering feedback in the modification of the shoreline) by the year 2100 is calculated to likely be between 340 and 571 m with a flooding rate of 3</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JHyd..414...72B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JHyd..414...72B"><span>Solving <span class="hlt">vertical</span> and horizontal well hydraulics problems analytically in Cartesian coordinates with <span class="hlt">vertical</span> and horizontal anisotropies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Batu, Vedat</p> <p>2012-01-01</p> <p>SummaryA new generalized three-dimensional analytical solution is developed for a partially-penetrating <span class="hlt">vertical</span> rectangular parallelepiped well screen in a confined aquifer by solving the three-dimensional transient ground water flow differential equation in x- y- z Cartesian coordinates system for drawdown by taking into account the three principal hydraulic conductivities ( Kx, Ky, and Kz) along the x- y- z coordinate directions. The fully penetrating screen case becomes equivalent to the single <span class="hlt">vertical</span> fracture case of Gringarten and Ramey (1973). It is shown that the new solution and Gringarten and Ramey solution (1973) match very well. Similarly, it is shown that this new solution for a horizontally tiny fully penetrating parallelepiped rectangular parallelepiped screen case match very well with Theis (1935) solution. Moreover, it is also shown that the horizontally tiny partially-penetrating parallelepiped rectangular well screen case of this new solution match very well with Hantush (1964) solution. This new analytical solution can also cover a partially-penetrating horizontal well by representing its screen interval with <span class="hlt">vertically</span> tiny rectangular parallelepiped. Also the solution takes into account both the <span class="hlt">vertical</span> anisotropy ( azx = Kz/ Kx) as well as the horizontal anisotropy ( ayx = Ky/ Kx) and has potential application areas to analyze pumping test drawdown data from partially-penetrating <span class="hlt">vertical</span> and horizontal wells by representing them as tiny rectangular parallelepiped as well as line sources. The solution has also potential application areas for a partially-penetrating parallelepiped rectangular <span class="hlt">vertical</span> fracture. With this new solution, the horizontal anisotropy ( ayx = Ky/ Kx) in addition to the <span class="hlt">vertical</span> anisotropy ( azx = Kz/ Kx) can also be determined using observed drawdown data. Most importantly, with this solution, to the knowledge of the author, it has been shown the first time in the literature that some well-known well hydraulics</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.C31C..02N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.C31C..02N"><span>Constraining Future Sea Level <span class="hlt">Rise</span> Estimates from the Amundsen Sea Embayment, West Antarctica</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nias, I.; Cornford, S. L.; Edwards, T.; Gourmelen, N.; Payne, A. J.</p> <p>2016-12-01</p> <p>The Amundsen Sea Embayment (ASE) is the primary source of mass loss from the West Antarctic Ice Sheet. The catchment is particularly susceptible to grounding line retreat, because the ice sheet is grounded on bedrock that is below sea level and deepening towards its interior. Mass loss from the ASE ice streams, which include Pine Island, Thwaites and Smith glaciers, is a major uncertainty on future sea level <span class="hlt">rise</span>, and understanding the dynamics of these ice streams is essential to constraining this uncertainty. The aim of this study is to construct a distribution of future ASE sea level contributions from an ensemble of ice sheet model simulations and observations of surface elevation change. A 284 member ensemble was performed using BISICLES, a <span class="hlt">vertically</span>-integrated ice flow model with adaptive mesh refinement. Within the ensemble parameters associated with basal traction, ice rheology and sub-shelf melt rate were perturbed, and the effect of bed topography and sliding law were also investigated. Initially each configuration was run to 50 model years. Satellite observations of surface height change were then used within a Bayesian framework to assign likelihoods to each ensemble member. Simulations that better reproduced the current thinning patterns across the catchment were given a higher score. The resulting posterior distribution of sea level contributions is narrower than the prior distribution, although the central estimates of sea level <span class="hlt">rise</span> are similar between the prior and posterior. The most extreme simulations were eliminated and the remaining ensemble members were extended to 200 years, using a simple melt rate forcing.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhDT.......227R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhDT.......227R"><span>Radial-<span class="hlt">vertical</span> profiles of tropical cyclone derived from dropsondes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ren, Yifang</p> <p></p> <p>The scopes of this thesis research are two folds: the first one is to the construct the intensity-based composite radial-<span class="hlt">vertical</span> profiles of tropical cyclones (TC) using GPS-based dropsonde observations and the second one is to identify the major deficiencies of Mathur vortices against the dropsonde composites of TCs. The intensity-based dropsonde composites of TCs advances our understanding of the dynamic and thermal structure of TCs of different intensity along the radial direction in and above the boundary layer where lies the devastating high wind that causes property damages and storm surges. The identification of the major deficiencies of Mathur vortices in representing the radial-<span class="hlt">vertical</span> profiles of TC of different intensity helps to improve numerical predictions of TCs since most operational TC forecast models need to utilize bogus vortices, such as Mathur vortices, to initialize TC forecasts and simulations. We first screen all available GPS dropsonde data within and round 35 named TCs over the tropical Atlantic basin from 1996 to 2010 and pair them with TC parameters derived from the best-track data provided by the National Hurricane Center (NHC) and select 1149 dropsondes that have continuous coverage in the lower troposphere. The composite radial-<span class="hlt">vertical</span> profiles of tangential wind speed, temperature, mixing ratio and humidity are based for each TC category ranging from "Tropical Storm" (TS) to "Hurricane Category <span class="hlt">1</span>" (H<span class="hlt">1</span>) through "Hurricane Category 5" (H5). The key findings of the dropsonde composites are: (i) all TCs have the maximum tangential wind within <span class="hlt">1</span> km above the ground and a distance of <span class="hlt">1</span>-2 times of the radius of maximum wind (RMW) at the surface; (ii) all TCs have a cold ring surrounding the warm core near the boundary layer at a distance of <span class="hlt">1</span>-3 times of the RMW and the cold ring structure gradually diminishes at a higher elevation where the warm core structure prevails along the radial direction; (iii) the existence of such shallow cold</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001488.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001488.html"><span>Glaciers and Sea Level <span class="hlt">Rise</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>Small valley glacier exiting the Devon Island Ice Cap in Canada. To learn about the contributions of glaciers to sea level <span class="hlt">rise</span>, visit: www.nasa.gov/topics/earth/features/glacier-sea-<span class="hlt">rise</span>.html Credit: Alex Gardner, Clark University NASA image use policy. NASA Goddard Space Flight Center enables NASA’s mission through four scientific endeavors: Earth Science, Heliophysics, Solar System Exploration, and Astrophysics. Goddard plays a leading role in NASA’s accomplishments by contributing compelling scientific knowledge to advance the Agency’s mission. Follow us on Twitter Like us on Facebook Find us on Instagram</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28149384','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28149384"><span>Effects of ethnicity on the relationship between <span class="hlt">vertical</span> jump and maximal power on a cycle ergometer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rouis, Majdi; Coudrat, Laure; Jaafar, Hamdi; Attiogbé, Elvis; Vandewalle, Henry; Driss, Tarak</p> <p>2016-06-01</p> <p>The aim of this study was to verify the impact of ethnicity on the maximal power-<span class="hlt">vertical</span> jump relationship. Thirty-one healthy males, sixteen Caucasian (age: 26.3 ± 3.5 years; body height: 179.<span class="hlt">1</span> ± 5.5 cm; body mass: 78.<span class="hlt">1</span> ± 9.8 kg) and fifteen Afro-Caribbean (age: 24.4 ±2.6 years; body height: 178.9 ± 5.5 cm; body mass: 77.<span class="hlt">1</span> ± 10.3 kg) completed three sessions during which <span class="hlt">vertical</span> jump height and maximal power of lower limbs were measured. The results showed that the values of <span class="hlt">vertical</span> jump height and maximal power were higher for Afro-Caribbean participants (62.92 ± 6.7 cm and 14.70 ± <span class="hlt">1</span>.75 W∙kg-<span class="hlt">1</span>) than for Caucasian ones (52.92 ± 4.4 cm and 12.75 ± <span class="hlt">1</span>.36 W∙kg-<span class="hlt">1</span>). Moreover, very high reliability indices were obtained on <span class="hlt">vertical</span> jump (e.g. 0.95 < ICC < 0.98) and maximal power performance (e.g. 0.75 < ICC < 0.97). However, multiple linear regression analysis showed that, for a given value of maximal power, the Afro-Caribbean participants jumped 8 cm higher than the Caucasians. Together, these results confirmed that ethnicity impacted the maximal power-<span class="hlt">vertical</span> jump relationship over three sessions. In the current context of cultural diversity, the use of <span class="hlt">vertical</span> jump performance as a predictor of muscular power should be considered with caution when dealing with populations of different ethnic origins.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFMGC21B..02A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFMGC21B..02A"><span>Projecting Future Sea Level <span class="hlt">Rise</span> for Water Resources Planning in California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Anderson, J.; Kao, K.; Chung, F.</p> <p>2008-12-01</p> <p>Sea level <span class="hlt">rise</span> is one of the major concerns for the management of California's water resources. Higher water levels and salinity intrusion into the Sacramento-San Joaquin Delta could affect water supplies, water quality, levee stability, and aquatic and terrestrial flora and fauna species and their habitat. Over the 20th century, sea levels near San Francisco Bay increased by over 0.6ft. Some tidal gauge and satellite data indicate that rates of sea level <span class="hlt">rise</span> are accelerating. Sea levels are expected to continue to <span class="hlt">rise</span> due to increasing air temperatures causing thermal expansion of the ocean and melting of land-based ice such as ice on Greenland and in southeastern Alaska. For water planners, two related questions are raised on the uncertainty of future sea levels. First, what is the expected sea level at a specific point in time in the future, e.g., what is the expected sea level in 2050? Second, what is the expected point of time in the future when sea levels will exceed a certain height, e.g., what is the expected range of time when the sea level <span class="hlt">rises</span> by one foot? To address these two types of questions, two factors are considered: (<span class="hlt">1</span>) long term sea level <span class="hlt">rise</span> trend, and (2) local extreme sea level fluctuations. A two-step approach will be used to develop sea level <span class="hlt">rise</span> projection guidelines for decision making that takes both of these factors into account. The first step is developing global sea level <span class="hlt">rise</span> probability distributions for the long term trends. The second step will extend the approach to take into account the effects of local astronomical tides, changes in atmospheric pressure, wind stress, floods, and the El Niño/Southern Oscillation. In this paper, the development of the first step approach is presented. To project the long term sea level <span class="hlt">rise</span> trend, one option is to extend the current rate of sea level <span class="hlt">rise</span> into the future. However, since recent data indicate rates of sea level <span class="hlt">rise</span> are accelerating, methods for estimating sea level <span class="hlt">rise</span></p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_24 --> <div id="page_25" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="481"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22391585-vertical-deformation-western-part-sumatra','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22391585-vertical-deformation-western-part-sumatra"><span><span class="hlt">Vertical</span> deformation at western part of Sumatra</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Febriyani, Caroline, E-mail: caroline.fanuel@students.itb.ac.id; Prijatna, Kosasih, E-mail: prijatna@gd.itb.ac.id; Meilano, Irwan, E-mail: irwan.meilano@gd.itb.ac.id</p> <p>2015-04-24</p> <p>This research tries to make advancement in GPS signal processing to estimate the interseismic <span class="hlt">vertical</span> deformation field at western part of Sumatra Island. The data derived by Continuous Global Positioning System (CGPS) from Badan Informasi Geospasial (BIG) between 2010 and 2012. GPS Analyze at Massachusetts Institute of Technology (GAMIT) software and Global Kalman Filter (GLOBK) software are used to process the GPS signal to estimate the <span class="hlt">vertical</span> velocities of the CGPS station. In order to minimize noise due to atmospheric delay, Vienna Mapping Function <span class="hlt">1</span> (VMF<span class="hlt">1</span>) is used as atmospheric parameter model and include daily IONEX file provided by themore » Center for Orbit Determination in Europe (CODE) as well. It improves GAMIT daily position accuracy up to 0.8 mm. In a second step of processing, the GLOBK is used in order to estimate site positions and velocities in the ITRF08 reference frame. The result shows that the uncertainties of estimated displacement velocity at all CGPS stations are smaller than <span class="hlt">1</span>.5 mm/yr. The subsided deformation patterns are seen at the northern and southern part of west Sumatra. The <span class="hlt">vertical</span> deformation at northern part of west Sumatra indicates postseismic phase associated with the 2010 and 2012 Northern Sumatra earthquakes and also the long-term postseismic associated with the 2004 and 2005 Northern Sumatra earthquakes. The uplifted deformation patterns are seen from Bukit Tinggi to Seblat which indicate a long-term interseismic phase after the 2007 Bengkulu earthquake and 2010 Mentawai earthquake. GANO station shows a subsidence at rate 12.25 mm/yr, indicating the overriding Indo-Australia Plate which is dragged down by the subducting Southeast Asian Plate.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1914925G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1914925G"><span>Gravitational salt tectonics above a <span class="hlt">rising</span> basement plateau offshore Algeria</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gaullier, Virginie; Vendeville, Bruno C.; Besème, Grégoire; Legoux, Gaetan; Déverchère, Jacques; Lymer, Gaël</p> <p>2017-04-01</p> <p>Seismic data (survey "MARADJA <span class="hlt">1</span>", 2003) offshore the Algerian coast have imaged an unexpected deformation pattern of the Messinian salt (Mobile Unit; MU) and its sedimentary overburden (Messinian Upper Unit and Plio-Quaternary) above an actively <span class="hlt">rising</span> plateau in the subsalt basement. From a geodynamic point of view, the region is undergoing crustal convergence, as attested by the Boumerdes earthquake (2003, magnitude 6.8). The <span class="hlt">rise</span> of this plateau, forming a 3D promontory restricted to the area offshore Algiers, is associated with that geodynamic setting. The seismic profiles show several subsalt thrusts (Domzig et al. 2006). The data provided additional information on the deformation of the Messinian mobile evaporitic unit and its Plio-Quaternary overburden. Margin-perpendicular profiles show mostly compressional features (anticlines and synclines) that had little activity during Messinian times, then grew more during Plio-Quaternary times. A few normal faults are also present, but are not accompanied by salt <span class="hlt">rise</span>. By contrast, margin-parallel profiles clearly show that extensional, reactive salt diapiric ridges (symptomatic with their triangular shape in cross section) formed early, as early as the time of deposition of the Messinian Upper Unit, as recorded by fan-shaped strata. These ridges have recorded E-W, thin-skinned gravity gliding above the Messinian salt, as a response to the <span class="hlt">rise</span> of the basement plateau. We tested this hypothesis using two analogue models, one where we assumed that the <span class="hlt">rise</span> of the plateau started after Messinian times (initially tabular salt across the entire region), the second model assumed that the plateau had already risen partially as the Messininan Mobile Unit was deposited (salt initially thinner above the plateau than in the adjacent regions). In both experiments, the <span class="hlt">rise</span> of the plateau generated preferential E-W extension above the salt, combined with N-S shortening. Extension was caused by gravity gliding of the salt from</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4914910','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4914910"><span>Influence of maxillary posterior discrepancy on upper molar <span class="hlt">vertical</span> position and facial <span class="hlt">vertical</span> dimensions in subjects with or without skeletal open bite</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Aliaga-Del Castillo, Aron; Pérez-Vargas, Luis Fernando; Flores-Mir, Carlos</p> <p>2016-01-01</p> <p>Summary Objectives: To determine the influence of maxillary posterior discrepancy on upper molar <span class="hlt">vertical</span> position and dentofacial <span class="hlt">vertical</span> dimensions in individuals with or without skeletal open bite (SOB). Materials and methods: Pre-treatment lateral cephalograms of 139 young adults were examined. The sample was divided into eight groups categorized according to their sagittal and <span class="hlt">vertical</span> skeletal facial growth pattern and maxillary posterior discrepancy (present or absent). Upper molar <span class="hlt">vertical</span> position, overbite, lower anterior facial height and facial height ratio were measured. Independent t-test was performed to determine differences between the groups considering maxillary posterior discrepancy. Principal component analysis and MANCOVA test were also used. Results: No statistically significant differences were found comparing the molar <span class="hlt">vertical</span> position according to maxillary posterior discrepancy for the SOB Class I group or the group with adequate overbite. Significant differences were found in SOB Class II and Class III groups. In addition, an increased molar <span class="hlt">vertical</span> position was found in the group without posterior discrepancy. Limitations: Some variables closely related with the individual’s intrinsic craniofacial development that could influence the evaluated <span class="hlt">vertical</span> measurements were not considered. Conclusions and implications: Overall maxillary posterior discrepancy does not appear to have a clear impact on upper molar <span class="hlt">vertical</span> position or facial <span class="hlt">vertical</span> dimensions. Only the SOB Class III group without posterior discrepancy had a significant increased upper molar <span class="hlt">vertical</span> position. PMID:26385786</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23538832','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23538832"><span>Electrical image of passive mantle upwelling beneath the northern East Pacific <span class="hlt">Rise</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Key, Kerry; Constable, Steven; Liu, Lijun; Pommier, Anne</p> <p>2013-03-28</p> <p>Melt generated by mantle upwelling is fundamental to the production of new oceanic crust at mid-ocean ridges, yet the forces controlling this process are debated. Passive-flow models predict symmetric upwelling due to viscous drag from the diverging tectonic plates, but have been challenged by geophysical observations of asymmetric upwelling that suggest anomalous mantle pressure and temperature gradients, and by observations of concentrated upwelling centres consistent with active models where buoyancy forces give <span class="hlt">rise</span> to focused convective flow. Here we use sea-floor magnetotelluric soundings at the fast-spreading northern East Pacific <span class="hlt">Rise</span> to image mantle electrical structure to a depth of about 160 kilometres. Our data reveal a symmetric, high-conductivity region at depths of 20-90 kilometres that is consistent with partial melting of passively upwelling mantle. The triangular region of conductive partial melt matches passive-flow predictions, suggesting that melt focusing to the ridge occurs in the porous melting region rather than along the shallower base of the thermal lithosphere. A deeper conductor observed east of the ridge at a depth of more than 100 kilometres is explained by asymmetric upwelling due to viscous coupling across two nearby transform faults. Significant electrical anisotropy occurs only in the shallowest mantle east of the ridge axis, where high <span class="hlt">vertical</span> conductivity at depths of 10-20 kilometres indicates localized porous conduits. This suggests that a coincident seismic-velocity anomaly is evidence of shallow magma transport channels rather than deeper off-axis upwelling. We interpret the mantle electrical structure as evidence that plate-driven passive upwelling dominates this ridge segment, with dynamic forces being negligible.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70033070','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70033070"><span>Elevated CO2 enhances biological contributions to elevation change in coastal wetlands by offsetting stressors associated with sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Cherry, J.A.; McKee, K.L.; Grace, J.B.</p> <p>2009-01-01</p> <p><span class="hlt">1</span>. Sea-level <span class="hlt">rise</span>, one indirect consequence of increasing atmospheric CO2, poses a major challenge to long-term stability of coastal wetlands. An important question is whether direct effects of elevated CO 2 on the capacity of marsh plants to accrete organic material and to maintain surface elevations outweigh indirect negative effects of stressors associated with sea-level <span class="hlt">rise</span> (salinity and flooding). 2. In this study, we used a mesocosm approach to examine potential direct and indirect effects of atmospheric CO2 concentration, salinity and flooding on elevation change in a brackish marsh community dominated by a C3 species, Schoenoplectus americanus, and a C4 grass, Spartina patens. This experimental design permitted identification of mechanisms and their role in controlling elevation change, and the development of models that can be tested in the field. 3. To test hypotheses related to CO2 and sea-level <span class="hlt">rise</span>, we used conventional anova procedures in conjunction with structural equation modelling (SEM). SEM explained 78% of the variability in elevation change and showed the direct, positive effect of S. americanus production on elevation. The SEM indicated that C3 plant response was influenced by interactive effects between CO2 and salinity on plant growth, not a direct CO2 fertilization effect. Elevated CO2 ameliorated negative effects of salinity on S. americanus and enhanced biomass contribution to elevation. 4. The positive relationship between S. americanus production and elevation change can be explained by shoot-base expansion under elevated CO 2 conditions, which led to <span class="hlt">vertical</span> soil displacement. While the response of this species may differ under other environmental conditions, shoot-base expansion and the general contribution of C3 plant production to elevation change may be an important mechanism contributing to soil expansion and elevation gain in other coastal wetlands. 5. Synthesis. Our results revealed previously unrecognized interactions and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/9083959','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/9083959"><span>Lincoln's <span class="hlt">vertical</span> strabismus.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Goldstein, J H</p> <p>1997-01-01</p> <p>The <span class="hlt">vertical</span> strabismus manifested by Abraham Lincoln has been noted. This article reviews the historical findings and provides a specific diagnosis. Previous reports of symptoms and history relating to Lincoln's left hypertropia were reviewed. A series of photographs were reviewed. Lincoln's own description of his symptoms is provided. Previous history indicates an intermittent left hypertropia. A family history of <span class="hlt">vertical</span> strabismus was noted with regard to Mr Lincoln's cousin. There also is a history of trauma to the left frontal area and life-mask evidence of fracture over the left eye. The findings include a history of head tilt and diplopia, presumably most readily in downgaze. Given the history and findings, the diagnosis of left superior oblique paresis of either congenital or traumatic origin seems appropriate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28277199','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28277199"><span><span class="hlt">Vertical</span> Transmission of Zika Virus by Aedes aegypti and Ae. albopictus Mosquitoes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ciota, Alexander T; Bialosuknia, Sean M; Ehrbar, Dylan J; Kramer, Laura D</p> <p>2017-05-01</p> <p>To determine the potential role of <span class="hlt">vertical</span> transmission in Zika virus expansion, we evaluated larval pools of perorally infected Aedes aegypti and Ae. albopictus adult female mosquitoes; ≈<span class="hlt">1</span>/84 larvae tested were Zika virus-positive; and rates varied among mosquito populations. Thus, <span class="hlt">vertical</span> transmission may play a role in Zika virus spread and maintenance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17983295','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17983295"><span>What's "up" with God? <span class="hlt">Vertical</span> space as a representation of the divine.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Meier, Brian P; Hauser, David J; Robinson, Michael D; Friesen, Chris Kelland; Schjeldahl, Katie</p> <p>2007-11-01</p> <p>"God" and "Devil" are abstract concepts often linked to <span class="hlt">vertical</span> metaphors (e.g., "glory to God in the highest," "the Devil lives down in hell"). It is unknown, however, whether these metaphors simply aid communication or implicate a deeper mode of concept representation. In 6 experiments, the authors examined the extent to which the <span class="hlt">vertical</span> dimension is used in noncommunication contexts involving God and the Devil. Experiment <span class="hlt">1</span> established that people have implicit associations between God-Devil and up-down. Experiment 2 revealed that people encode God-related concepts faster if presented in a high (vs. low) <span class="hlt">vertical</span> position. Experiment 3 found that people's memory for the <span class="hlt">vertical</span> location of God- and Devil-like images showed a metaphor-consistent bias (up for God; down for Devil). Experiments 4, 5a, and 5b revealed that people rated strangers as more likely to believe in God when their images appeared in a high versus low <span class="hlt">vertical</span> position, and this effect was independent of inferences related to power and likability. These robust results reveal that <span class="hlt">vertical</span> perceptions are invoked when people access divinity-related cognitions. (c) 2007 APA, all rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25187243','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25187243"><span>Relationships between explosive and maximal triple extensor muscle performance and <span class="hlt">vertical</span> jump height.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chang, Eunwook; Norcross, Marc F; Johnson, Sam T; Kitagawa, Taichi; Hoffman, Mark</p> <p>2015-02-01</p> <p>The purpose of this study was to examine the relationships between maximum <span class="hlt">vertical</span> jump height and (a) rate of torque development (RTD) calculated during 2 time intervals, 0-50 milliseconds (RTD50) and 0-200 milliseconds (RTD200) after torque onset and (b) peak torque (PT) for each of the triple extensor muscle groups. Thirty recreationally active individuals performed maximal isometric voluntary contractions (MVIC) of the hip, knee and ankle extensors, and a countermovement <span class="hlt">vertical</span> jump. Rate of torque development was calculated from 0 to 50 (RTD50) and 0 to 200 (RTD200) milliseconds after the onset of joint torque. Peak torque was identified and defined as the maximum torque value during each MVIC trial. Greater <span class="hlt">vertical</span> jump height was associated with greater knee and ankle extension RTD50, RTD200, and PT (p ≤ 0.05). However, hip extension RTD50, RTD200, and PT were not significantly related to maximal <span class="hlt">vertical</span> jump height (p > 0.05). The results indicate that 47.6 and 32.5% of the variability in <span class="hlt">vertical</span> jump height was explained by knee and ankle extensor RTD50, respectively. Knee and ankle extensor RTD50 also seemed to be more closely related to <span class="hlt">vertical</span> jump performance than RTD200 (knee extensor: 28.<span class="hlt">1</span>% and ankle extensor: 28.<span class="hlt">1</span>%) and PT (knee extensor: 31.4% and ankle extensor: 13.7%). Overall, these results suggest that training specifically targeted to improve knee and ankle extension RTD, especially during the early phases of muscle contraction, may be effective for increasing maximal <span class="hlt">vertical</span> jump performance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19810019546','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19810019546"><span>A Mathematical Model for <span class="hlt">Vertical</span> Attitude Takeoff and Landing (VATOL) Aircraft Simulation. Volume <span class="hlt">1</span>; Model Description Application</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fortenbaugh, R. L.</p> <p>1980-01-01</p> <p>A mathematical model of a high performance airplane capable of <span class="hlt">vertical</span> attitude takeoff and landing (VATOL) was developed. An off line digital simulation program incorporating this model was developed to provide trim conditions and dynamic check runs for the piloted simulation studies and support dynamic analyses of proposed VATOL configuration and flight control concepts. Development details for the various simulation component models and the application of the off line simulation program, <span class="hlt">Vertical</span> Attitude Take-Off and Landing Simulation (VATLAS), to develop a baseline control system for the Vought SF-121 VATOL airplane concept are described.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvD..97j6021K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvD..97j6021K"><span>Cubic interactions of massless bosonic fields in three dimensions. II. Parity-odd and Chern-Simons <span class="hlt">vertices</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kessel, Pan; Mkrtchyan, Karapet</p> <p>2018-05-01</p> <p>This work completes the classification of the cubic <span class="hlt">vertices</span> for arbitrary-spin massless bosons in three dimensions started in a previous companion paper by constructing parity-odd <span class="hlt">vertices</span>. Similarly to the parity-even case, there is a unique parity-odd vertex for any given triple s<span class="hlt">1</span>≥s2≥s3≥2 of massless bosons if the triangle inequalities are satisfied (s<span class="hlt">1</span><s2+s3 ) and none otherwise. These <span class="hlt">vertices</span> involve two (three) derivatives for odd (even) values of the sum s<span class="hlt">1</span>+s2+s3. A nontrivial relation between parity-even and parity-odd <span class="hlt">vertices</span> is found. Similarly to the parity-even case, the scalar and Maxwell matter can couple to higher spins through current couplings with higher derivatives. We comment on possible lessons for two-dimensional conformal field theory. We also derive both parity-even and parity-odd <span class="hlt">vertices</span> with Chern-Simons fields and comment on the analogous classification in two dimensions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=ocean+AND+climate+AND+changes&pg=3&id=EJ912346','ERIC'); return false;" href="https://eric.ed.gov/?q=ocean+AND+climate+AND+changes&pg=3&id=EJ912346"><span><span class="hlt">Rising</span> Sea Levels: Truth or Scare?</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Peacock, Alan</p> <p>2007-01-01</p> <p>When "ITV News" ran an item that shocked the author, about <span class="hlt">rising</span> sea levels that will have caused the entire evacuation of the islands by the end of this year, he began to wonder whether the Pacific Ocean is really <span class="hlt">rising</span> as fast as this. The media reporting of such things can be a double-edged sword. On the one hand, it brought to the author's…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=203865&keyword=Scheme&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=203865&keyword=Scheme&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>Diel <span class="hlt">Vertical</span> Migration Thresholds of Karenia brevis (Dinophyceae).</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>Light and nutrient availability change throughout dinoflagellate diel <span class="hlt">vertical</span> migration (DVM) and/or with subpopulation location in the water column along the west Florida shelf. Typically, the <span class="hlt">vertical</span> depth of the shelf is greater than the distance a subpopulation can <span class="hlt">vertical</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940026128','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940026128"><span>Methods of testing parameterizations: <span class="hlt">Vertical</span> ocean mixing</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tziperman, Eli</p> <p>1992-01-01</p> <p>The ocean's velocity field is characterized by an exceptional variety of scales. While the small-scale oceanic turbulence responsible for the <span class="hlt">vertical</span> mixing in the ocean is of scales a few centimeters and smaller, the oceanic general circulation is characterized by horizontal scales of thousands of kilometers. In oceanic general circulation models that are typically run today, the <span class="hlt">vertical</span> structure of the ocean is represented by a few tens of discrete grid points. Such models cannot explicitly model the small-scale mixing processes, and must, therefore, find ways to parameterize them in terms of the larger-scale fields. Finding a parameterization that is both reliable and plausible to use in ocean models is not a simple task. <span class="hlt">Vertical</span> mixing in the ocean is the combined result of many complex processes, and, in fact, mixing is one of the less known and less understood aspects of the oceanic circulation. In present models of the oceanic circulation, the many complex processes responsible for <span class="hlt">vertical</span> mixing are often parameterized in an oversimplified manner. Yet, finding an adequate parameterization of <span class="hlt">vertical</span> ocean mixing is crucial to the successful application of ocean models to climate studies. The results of general circulation models for quantities that are of particular interest to climate studies, such as the meridional heat flux carried by the ocean, are quite sensitive to the strength of the <span class="hlt">vertical</span> mixing. We try to examine the difficulties in choosing an appropriate <span class="hlt">vertical</span> mixing parameterization, and the methods that are available for validating different parameterizations by comparing model results to oceanographic data. First, some of the physical processes responsible for <span class="hlt">vertically</span> mixing the ocean are briefly mentioned, and some possible approaches to the parameterization of these processes in oceanographic general circulation models are described in the following section. We then discuss the role of the <span class="hlt">vertical</span> mixing in the physics of the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27580268','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27580268"><span><span class="hlt">Vertical</span> Transmission of Hepatozoon in the Garter Snake Thamnophis elegans.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kauffman, Kiera L; Sparkman, Amanda; Bronikowski, Anne M; Palacios, Maria G</p> <p>2017-01-01</p> <p><span class="hlt">Vertical</span> transmission of blood parasites has been demonstrated in humans and some domestic species, but it has not been well documented in wild populations. We assessed whether Hepatozoon blood parasites are <span class="hlt">vertically</span> transmitted in naturally infected individuals of the viviparous western terrestrial garter snake ( Thamnophis elegans ). Blood smears were taken from nine wild-caught gravid female snakes at capture, preparturition, and postparturition, and then from their laboratory-born offspring at age 2 mo and <span class="hlt">1</span> yr. All infected offspring were born to four infected females, although not all offspring in a given litter were necessarily infected. Parasites were not detected in offspring born to the five uninfected mothers. The highest parasite loads were found in neonates at 2 mo of age. Parasite prevalence did not vary between sexes in offspring, but females showed higher loads than did males when 2 mo old. This study supports <span class="hlt">vertical</span> transmission of Hepatozoon in naturally infected viviparous snakes and suggests that <span class="hlt">vertical</span> transmission of hematozoan parasites might be an overlooked mode of transmission in wildlife.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DFDM39008H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DFDM39008H"><span>Measurements of fluid transport by controllable <span class="hlt">vertical</span> migrations of plankton</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Houghton, Isabel A.; Dabiri, John O.</p> <p>2016-11-01</p> <p>Diel <span class="hlt">vertical</span> migration of zooplankton has been proposed to be a significant contributor to local and possibly large-scale fluid transport in the ocean. However, studies of this problem to date have been limited to order-of-magnitude estimates based on first principles and a small number of field observations. In this work, we leverage the phototactic behavior of zooplankton to stimulate controllable <span class="hlt">vertical</span> migrations in the laboratory and to study the associated fluid transport and mixing. Building upon a previous prototype system, a laser guidance system induces <span class="hlt">vertical</span> swimming of brine shrimp (Artemia salina) in a 2.<span class="hlt">1</span> meter tall, density-stratified water tank. The animal swimming speed and spacing during the controlled <span class="hlt">vertical</span> migration is characterized with video analysis. A schlieren imaging system is utilized to visualize density perturbations to a stable stratification for quantification of fluid displacement length scales and restratification timescales. These experiments can add to our understanding of the dynamics of active particles in stratified flows. NSF and US-Israel Binational Science Foundation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22246885-social-values-risk-from-sea-level-rise','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22246885-social-values-risk-from-sea-level-rise"><span>The social values at risk from sea-level <span class="hlt">rise</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Graham, Sonia, E-mail: sonia.graham@unimelb.edu.au; Barnett, Jon, E-mail: jbarn@unimelb.edu.au; Fincher, Ruth, E-mail: r.fincher@unimelb.edu.au</p> <p></p> <p>Analysis of the risks of sea-level <span class="hlt">rise</span> favours conventionally measured metrics such as the area of land that may be subsumed, the numbers of properties at risk, and the capital values of assets at risk. Despite this, it is clear that there exist many less material but no less important values at risk from sea-level <span class="hlt">rise</span>. This paper re-theorises these multifarious social values at risk from sea-level <span class="hlt">rise</span>, by explaining their diverse nature, and grounding them in the everyday practices of people living in coastal places. It is informed by a review and analysis of research on social values frommore » within the fields of social impact assessment, human geography, psychology, decision analysis, and climate change adaptation. From this we propose that it is the ‘lived values’ of coastal places that are most at risk from sea-level <span class="hlt">rise</span>. We then offer a framework that groups these lived values into five types: those that are physiological in nature, and those that relate to issues of security, belonging, esteem, and self-actualisation. This framework of lived values at risk from sea-level <span class="hlt">rise</span> can guide empirical research investigating the social impacts of sea-level <span class="hlt">rise</span>, as well as the impacts of actions to adapt to sea-level <span class="hlt">rise</span>. It also offers a basis for identifying the distribution of related social outcomes across populations exposed to sea-level <span class="hlt">rise</span> or sea-level <span class="hlt">rise</span> policies.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19750020624','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19750020624"><span>Brainstem auditory evoked responses in man. <span class="hlt">1</span>: Effect of stimulus <span class="hlt">rise</span>-fall time and duration</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hecox, K.; Squires, N.; Galambos, R.</p> <p>1975-01-01</p> <p>Short latency (under 10 msec) responses elicited by bursts of white noise were recorded from the scalps of human subjects. Response alterations produced by changes in the noise burst duration (on-time), inter-burst interval (off-time), and onset and offset shapes were analyzed. The latency of the most prominent response component, wave V, was markedly delayed with increases in stimulus <span class="hlt">rise</span> time but was unaffected by changes in fall time. Increases in stimulus duration, and therefore in loudness, resulted in a systematic increase in latency. This was probably due to response recovery processes, since the effect was eliminated with increases in stimulus off-time. The amplitude of wave V was insensitive to changes in signal <span class="hlt">rise</span> and fall times, while increasing signal on-time produced smaller amplitude responses only for sufficiently short off-times. It was concluded that wave V of the human auditory brainstem evoked response is solely an onset response.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26385786','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26385786"><span>Influence of maxillary posterior discrepancy on upper molar <span class="hlt">vertical</span> position and facial <span class="hlt">vertical</span> dimensions in subjects with or without skeletal open bite.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Arriola-Guillén, Luis Ernesto; Aliaga-Del Castillo, Aron; Pérez-Vargas, Luis Fernando; Flores-Mir, Carlos</p> <p>2016-06-01</p> <p>To determine the influence of maxillary posterior discrepancy on upper molar <span class="hlt">vertical</span> position and dentofacial <span class="hlt">vertical</span> dimensions in individuals with or without skeletal open bite (SOB). Pre-treatment lateral cephalograms of 139 young adults were examined. The sample was divided into eight groups categorized according to their sagittal and <span class="hlt">vertical</span> skeletal facial growth pattern and maxillary posterior discrepancy (present or absent). Upper molar <span class="hlt">vertical</span> position, overbite, lower anterior facial height and facial height ratio were measured. Independent t-test was performed to determine differences between the groups considering maxillary posterior discrepancy. Principal component analysis and MANCOVA test were also used. No statistically significant differences were found comparing the molar <span class="hlt">vertical</span> position according to maxillary posterior discrepancy for the SOB Class I group or the group with adequate overbite. Significant differences were found in SOB Class II and Class III groups. In addition, an increased molar <span class="hlt">vertical</span> position was found in the group without posterior discrepancy. Some variables closely related with the individual's intrinsic craniofacial development that could influence the evaluated <span class="hlt">vertical</span> measurements were not considered. Overall maxillary posterior discrepancy does not appear to have a clear impact on upper molar <span class="hlt">vertical</span> position or facial <span class="hlt">vertical</span> dimensions. Only the SOB Class III group without posterior discrepancy had a significant increased upper molar <span class="hlt">vertical</span> position. © The Author 2015. Published by Oxford University Press on behalf of the European Orthodontic Society. All rights reserved. For permissions, please email: journals.permissions@oup.com.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUOSME24B0706A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUOSME24B0706A"><span>Characterization of <span class="hlt">vertical</span> mixing in oscillatory vegetated flows</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Abdolahpour, M.; Ghisalberti, M.; Lavery, P.; McMahon, K.</p> <p>2016-02-01</p> <p>Seagrass meadows are primary producers that provide important ecosystem services, such as improved water quality, sediment stabilisation and trapping and recycling of nutrients. Most of these ecological services are strongly influenced by the <span class="hlt">vertical</span> exchange of water across the canopy-water interface. That is, <span class="hlt">vertical</span> mixing is the main hydrodynamic process governing the large-scale ecological and environmental impact of seagrass meadows. The majority of studies into mixing in vegetated flows have focused on steady flow environments whereas many coastal canopies are subjected to oscillatory flows driven by surface waves. It is known that the rate of mass transfer will vary greatly between unidirectional and oscillatory flows, necessitating a specific investigation of mixing in oscillatory canopy flows. In this study, we conducted an extensive laboratory investigation to characterise the rate of <span class="hlt">vertical</span> mixing through a <span class="hlt">vertical</span> turbulent diffusivity (Dt,z). This has been done through gauging the evolution of <span class="hlt">vertical</span> profiles of concentration (C) of a dye sheet injected into a wave-canopy flow. Instantaneous measurement of the variance of the <span class="hlt">vertical</span> concentration distribution ( allowed the estimation of a <span class="hlt">vertical</span> turbulent diffusivity (). Two types of model canopies, rigid and flexible, with identical heights and frontal areas, were subjected to a wide and realistic range of wave height and period. The results showed two important mechanisms that dominate <span class="hlt">vertical</span> mixing under different conditions: a shear layer that forms at the top of the canopy and wake turbulence generated by the stems. By allowing a coupled contribution of wake and shear layer mixing, we present a relationship that can be used to predict the rate of <span class="hlt">vertical</span> mixing in coastal canopies. The results further showed that the rate of <span class="hlt">vertical</span> mixing within flexible vegetation was always lower than the corresponding rigid canopy, confirming the impact of plant flexibility on canopy</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_25 --> <div class="footer-extlink text-muted" style="margin-bottom:1rem; text-align:center;">Some links on this page may take you to non-federal websites. Their policies may differ from this site.</div> </div><!-- container --> <footer><a id="backToTop" href="#top"> </a><nav><a id="backToTop" href="#top"> </a><ul class="links"><a id="backToTop" href="#top"> </a><li><a id="backToTop" href="#top"></a><a href="/sitemap.html">Site Map</a></li> <li><a href="/members/index.html">Members Only</a></li> <li><a href="/website-policies.html">Website Policies</a></li> <li><a href="https://doe.responsibledisclosure.com/hc/en-us" target="_blank">Vulnerability Disclosure Program</a></li> <li><a href="/contact.html">Contact Us</a></li> </ul> <div class="small">Science.gov is maintained by the U.S. Department of Energy's <a href="https://www.osti.gov/" target="_blank">Office of Scientific and Technical Information</a>, in partnership with <a href="https://www.cendi.gov/" target="_blank">CENDI</a>.</div> </nav> </footer> <script type="text/javascript"><!-- // var lastDiv = ""; function showDiv(divName) { // hide last div if (lastDiv) { document.getElementById(lastDiv).className = "hiddenDiv"; } //if value of the box is not nothing and an object with that name exists, then change the class if (divName && document.getElementById(divName)) { document.getElementById(divName).className = "visibleDiv"; lastDiv = divName; } } //--> </script> <script> /** * Function that tracks a click on an outbound link in Google Analytics. * This function takes a valid URL string as an argument, and uses that URL string * as the event label. */ var trackOutboundLink = function(url,collectionCode) { try { h = window.open(url); setTimeout(function() { ga('send', 'event', 'topic-page-click-through', collectionCode, url); }, 1000); } catch(err){} }; </script> <!-- Google Analytics --> <script> (function(i,s,o,g,r,a,m){i['GoogleAnalyticsObject']=r;i[r]=i[r]||function(){ (i[r].q=i[r].q||[]).push(arguments)},i[r].l=1*new Date();a=s.createElement(o), m=s.getElementsByTagName(o)[0];a.async=1;a.src=g;m.parentNode.insertBefore(a,m) })(window,document,'script','//www.google-analytics.com/analytics.js','ga'); ga('create', 'UA-1122789-34', 'auto'); ga('send', 'pageview'); </script> <!-- End Google Analytics --> <script> showDiv('page_1') </script> </body> </html>