Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 18 de 18
Filter
Add more filters










Publication year range
1.
Neuroscience ; 156(4): 801-18, 2008 Oct 28.
Article in English | MEDLINE | ID: mdl-18786618

ABSTRACT

Spatial updating is the means by which we keep track of the locations of objects in space even as we move. Four decades of research have shown that humans and non-human primates can take the amplitude and direction of intervening movements into account, including saccades (both head-fixed and head-free), pursuit, whole-body rotations and translations. At the neuronal level, spatial updating is thought to be maintained by receptive field locations that shift with changes in gaze, and evidence for such shifts has been shown in several cortical areas. These regions receive information about the intervening movement from several sources including motor efference copies when a voluntary movement is made and vestibular/somatosensory signals when the body is in motion. Many of these updating signals arise from brainstem regions that monitor our ongoing movements and subsequently transmit this information to the cortex via pathways that likely include the thalamus. Several issues of debate include (1) the relative contribution of extra-retinal sensory and efference copy signals to spatial updating, (2) the source of an updating signal for real life, three-dimensional motion that cannot arise from brain areas encoding only two-dimensional commands, and (3) the reference frames used by the brain to integrate updating signals from various sources. This review highlights the relevant spatial updating studies and provides a summary of the field today. We find that spatial constancy is maintained by a highly evolved neural mechanism that keeps track of our movements, transmits this information to relevant brain regions, and then uses this information to change the way in which single neurons respond. In this way, we are able to keep track of relevant objects in the outside world and interact with them in meaningful ways.


Subject(s)
Eye Movements/physiology , Head Movements/physiology , Space Perception/physiology , Vision, Ocular/physiology , Animals , Humans , Orientation , Photic Stimulation/methods , Visual Fields/physiology
2.
Curr Opin Neurobiol ; 13(6): 655-62, 2003 Dec.
Article in English | MEDLINE | ID: mdl-14662365

ABSTRACT

Although the eyes and head can potentially rotate about any three-dimensional axis during orienting gaze shifts, behavioral recordings have shown that certain lawful strategies--such as Listing's law and Donders' law--determine which axis is used for a particular sensory input. Here, we review recent advances in understanding the neuromuscular mechanisms for these laws, the neural mechanisms that control three-dimensional head posture, and the neural mechanisms that coordinate three-dimensional eye orientation with head motion. Finally, we consider how the brain copes with the perceptual consequences of these motor acts.


Subject(s)
Brain/physiology , Eye Movements/physiology , Head Movements/physiology , Psychomotor Performance/physiology , Animals , Humans
4.
Nat Neurosci ; 4(6): 627-32, 2001 Jun.
Article in English | MEDLINE | ID: mdl-11369944

ABSTRACT

The superior colliculus (SC) has a topographic map of visual space, but the spatial nature of its output command for orienting gaze shifts remains unclear. Here we show that the SC codes neither desired gaze displacement nor gaze direction in space (as debated previously), but rather, desired gaze direction in retinal coordinates. Electrical micro-stimulation of the SC in two head-free (non-immobilized) monkeys evoked natural-looking, eye-head gaze shifts, with anterior sites producing small, fixed-vector movements and posterior sites producing larger, strongly converging movements. However, when correctly calculated in retinal coordinates, all of these trajectories became 'fixed-vector.' Moreover, our data show that this eye-centered SC command is then further transformed, as a function of eye and head position, by downstream mechanisms into the head- and body-centered commands for coordinated eye-head gaze shifts.


Subject(s)
Fixation, Ocular/physiology , Head Movements/physiology , Neurons/physiology , Retina/physiology , Saccades/physiology , Superior Colliculi/physiology , Animals , Darkness , Light , Macaca fascicularis , Mesencephalon/physiology , Motor Activity , Photic Stimulation
5.
J Neurophysiol ; 81(4): 1760-82, 1999 Apr.
Article in English | MEDLINE | ID: mdl-10200211

ABSTRACT

The purpose of this investigation was to describe the neural constraints on three-dimensional (3-D) orientations of the eye in space (Es), head in space (Hs), and eye in head (Eh) during visual fixations in the monkey and the control strategies used to implement these constraints during head-free gaze saccades. Dual scleral search coil signals were used to compute 3-D orientation quaternions, two-dimensional (2-D) direction vectors, and 3-D angular velocity vectors for both the eye and head in three monkeys during the following visual tasks: radial to/from center, repetitive horizontal, nonrepetitive oblique, random (wide 2-D range), and random with pin-hole goggles. Although 2-D gaze direction (of Es) was controlled more tightly than the contributing 2-D Hs and Eh components, the torsional standard deviation of Es was greater (mean 3.55 degrees ) than Hs (3.10 degrees ), which in turn was greater than Eh (1.87 degrees ) during random fixations. Thus the 3-D Es range appeared to be the byproduct of Hs and Eh constraints, resulting in a pseudoplanar Es range that was twisted (in orthogonal coordinates) like the zero torsion range of Fick coordinates. The Hs fixation range was similarly Fick-like, whereas the Eh fixation range was quasiplanar. The latter Eh range was maintained through exquisite saccade/slow phase coordination, i.e., during each head movement, multiple anticipatory saccades drove the eye torsionally out of the planar range such that subsequent slow phases drove the eye back toward the fixation range. The Fick-like Hs constraint was maintained by the following strategies: first, during purely vertical/horizontal movements, the head rotated about constantly oriented axes that closely resembled physical Fick gimbals, i.e., about head-fixed horizontal axes and space-fixed vertical axes, respectively (although in 1 animal, the latter constraint was relaxed during repetitive horizontal movements, allowing for trajectory optimization). However, during large oblique movements, head orientation made transient but dramatic departures from the zero-torsion Fick surface, taking the shortest path between two torsionally eccentric fixation points on the surface. Moreover, in the pin-hole goggle task, the head-orientation range flattened significantly, suggesting a task-dependent default strategy similar to Listing's law. These and previous observations suggest two quasi-independent brain stem circuits: an oculomotor 2-D to 3-D transformation that coordinates anticipatory saccades with slow phases to uphold Listing's law, and a flexible "Fick operator" that selects head motor error; both nested within a dynamic gaze feedback loop.


Subject(s)
Head Movements/physiology , Psychomotor Performance/physiology , Saccades/physiology , Algorithms , Animals , Conditioning, Psychological/physiology , Fixation, Ocular/physiology , Macaca fascicularis , Photic Stimulation , Torsion Abnormality , Visual Perception/physiology
6.
J Neurophysiol ; 80(5): 2274-94, 1998 Nov.
Article in English | MEDLINE | ID: mdl-9819243

ABSTRACT

A recent theoretical investigation has demonstrated that three-dimensional (3-D) eye position dependencies in the geometry of retinal stimulation must be accounted for neurally (i.e., in a visuomotor reference frame transformation) if saccades are to be both accurate and obey Listing's law from all initial eye positions. Our goal was to determine whether the human saccade generator correctly implements this eye-to-head reference frame transformation (RFT), or if it approximates this function with a visuomotor look-up table (LT). Six head-fixed subjects participated in three experiments in complete darkness. We recorded 60 degrees horizontal saccades between five parallel pairs of lights, over a vertical range of +/-40 degrees (experiment 1), and 30 degrees radial saccades from a central target, with the head upright or tilted 45 degrees clockwise/counterclockwise to induce torsional ocular counterroll, under both binocular and monocular viewing conditions (experiments 2 and 3). 3-D eye orientation and oculocentric target direction (i.e., retinal error) were computed from search coil signals in the right eye. Experiment 1: as predicted, retinal error was a nontrivial function of both target displacement in space and 3-D eye orientation (e.g., horizontally displaced targets could induce horizontal or oblique retinal errors, depending on eye position). These data were input to a 3-D visuomotor LT model, which implemented Listing's law, but predicted position-dependent errors in final gaze direction of up to 19.8 degrees. Actual saccades obeyed Listing's law but did not show the predicted pattern of inaccuracies in final gaze direction, i.e., the slope of actual error, as a function of predicted error, was only -0. 01 +/- 0.14 (compared with 0 for RFT model and 1.0 for LT model), suggesting near-perfect compensation for eye position. Experiments 2 and 3: actual directional errors from initial torsional eye positions were only a fraction of those predicted by the LT model (e. g., 32% for clockwise and 33% for counterclockwise counterroll during binocular viewing). Furthermore, any residual errors were immediately reduced when visual feedback was provided during saccades. Thus, other than sporadic miscalibrations for torsion, saccades were accurate from all 3-D eye positions. We conclude that 1) the hypothesis of a visuomotor look-up table for saccades fails to account even for saccades made directly toward visual targets, but rather, 2) the oculomotor system takes 3-D eye orientation into account in a visuomotor reference frame transformation. This transformation is probably implemented physiologically between retinotopically organized saccade centers (in cortex and superior colliculus) and the brain stem burst generator.


Subject(s)
Psychomotor Performance/physiology , Saccades/physiology , Visual Perception/physiology , Adult , Feedback , Female , Fixation, Ocular/physiology , Humans , Male , Models, Neurological , Retina/physiology , Vision, Binocular/physiology , Vision, Monocular/physiology
7.
J Neurosci ; 18(4): 1583-94, 1998 Feb 15.
Article in English | MEDLINE | ID: mdl-9454863

ABSTRACT

Establishing a coherent internal reference frame for visuospatial representation and maintaining the integrity of this frame during eye movements are thought to be crucial for both perception and motor control. A stable headcentric representation could be constructed by internally comparing retinal signals with eye position. Alternatively, visual memory traces could be actively remapped within an oculocentric frame to compensate for each eye movement. We tested these models by measuring errors in manual pointing (in complete darkness) toward briefly flashed central targets during three oculomotor paradigms; subjects pointed accurately when gaze was maintained on the target location (control paradigm). However, when steadily fixating peripheral locations (static paradigm), subjects exaggerated the retinal eccentricity of the central target by 13.4 +/- 5.1%. In the key "dynamic" paradigm, subjects briefly foveated the central target and then saccaded peripherally before pointing toward the remembered location of the target. Our headcentric model predicted accurate pointing (as seen in the control paradigm) independent of the saccade, whereas our oculocentric model predicted misestimation (as seen in the static paradigm) of an internally shifted retinotopic trace. In fact, pointing errors were significantly larger than were control errors (p /= 0.25) from the static paradigm errors. Scatter plots of pointing errors (dynamic vs static paradigm) for various final fixation directions showed an overall slope of 0.97, contradicting the headcentric prediction (0. 0) and supporting the oculocentric prediction (1.0). Varying both fixation and pointing-target direction confirmed that these errors were a function of retinotopically shifted memory traces rather than eye position per se. To reconcile these results with previous pointing experiments, we propose a "conversion-on-demand" model of visuomotor control in which multiple visual targets are stored and rotated (noncommutatively) within the oculocentric frame, whereas only select targets are transformed further into head- or bodycentric frames for motor execution.


Subject(s)
Arm/physiology , Brain/physiology , Memory/physiology , Psychomotor Performance/physiology , Space Perception/physiology , Visual Perception/physiology , Adult , Female , Humans , Male , Models, Neurological , Movement/physiology , Ocular Physiological Phenomena , Retina/physiology , Saccades/physiology
8.
Cas Lek Cesk ; 128(11): 332-5, 1989 Mar 10.
Article in Czech | MEDLINE | ID: mdl-2731199

ABSTRACT

The authors explain the principle of cochlear neuroprostheses, an electronic device which makes it possible to mediate acoustic perceptions to completely deaf patients. They describe the Czechoslovak cochlear neuroprosthesis, developed and manufactured in the Laboratory of electronic sensory prostheses. For clinical application preoperative and postoperative examination of the patients is essential. For these examinations in the Laboratory a computer-controlled testing system was developed. The authors describe tests for the selection of candidates for implantation and for evaluation of the assets of the neuroprosthesis.


Subject(s)
Cochlear Implants , Humans
9.
J Biomed Eng ; 10(5): 470-1, 1988 Oct.
Article in English | MEDLINE | ID: mdl-3236876

ABSTRACT

A number of failures of extracochlear implants are caused by a bad contact between the active electrode and the tissue. A novel and simple method has been developed to enable the impedance of the active electrode to be measured before completion of surgery.


Subject(s)
Cochlear Implants , Electric Conductivity , Humans
11.
Zentralbl Gynakol ; 102(11): 628-30, 1980.
Article in German | MEDLINE | ID: mdl-6256998

ABSTRACT

Reported in this paper is a case of androblastoma of the right ovary, four years from first parturition. Signs of virilisation developed concomitantly but faded away very soon after tumour extirpation. The second birth took place, less than four years later. No symptoms so far have been recordable in two-a-half year follow-up checks.


Subject(s)
Ovarian Neoplasms/surgery , Sertoli-Leydig Cell Tumor/surgery , Adult , Female , Follow-Up Studies , Humans , Ovarian Neoplasms/pathology , Pregnancy , Sertoli-Leydig Cell Tumor/pathology
12.
Zentralbl Gynakol ; 100(24): 1619-20, 1978.
Article in German | MEDLINE | ID: mdl-742256

ABSTRACT

Report on a rapidly growing large retroperitonally placed tumor which after fifteen months was extirpated by applying a Wertheim-radical operation and following ray-therapy. The tumor weighed 6 kg and turned out to be a good-natured lipom.


Subject(s)
Lipoma/pathology , Postoperative Complications/pathology , Retroperitoneal Neoplasms/pathology , Aged , Carcinoma, Basal Cell/surgery , Female , Humans , Hysterectomy , Lipoma/surgery , Postoperative Complications/surgery , Retroperitoneal Neoplasms/surgery , Uterine Cervical Neoplasms/surgery
SELECTION OF CITATIONS
SEARCH DETAIL
...