RETURNING MATERIALS: )V{ESI_J P1ace in book drop to LJBRARJES remove this checkout from W your record. FMES win \. be charged if book is returned after the date stamped below. CEREBROSPINAL FLUID (CSF) TRANSIENT RESPONSES INDUCED BY HYPERCAPNIA BY Marye Jill Fisher A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Physiology 1984 n.-. ABSTRACT CEREBROSPINAL FLUID (CS?) TRANSIENT RESPONSES INDUCED BY EYPERCAPNIA 3}! larye Jill Fisher Intracranial CSF volume and brain blood volume (BBV) are reciprocally related. This dynamic equilibrium is transiently disrupted when an animal breathes C02. Both its brain blood flow and BBV increase to displace competitively a CSF volume (Vd). ‘Ventriculocisternal perfusion studies using tracers have shown that not only is there a transient increase in CSF outflow rate with hypercapnia. but also a transient decrease in CSF effluent tracer concentration (Cd). These CSF transient responses (Vd and Cd) during early respiratory acidosis are attributed to two phenomena. One explanation assumes a homogeneous distribution of tracer throughout the CSF spaces and that CSF formation (vf) episodically increases during C02 breathing. The second explanation allows heterogeneous dispersion of tracer in CSF, an unchanging 9f, and is based on a redistribution of craniospinal fluid volumes with C02 breathing. Both phenomena may contribute to the short-term displacement of a dilute CSF during hypercapnia. Harye Jill Fisher CSF transient responses to C02 inhalation were measured before and after facilitated perfusate flow through sub- arachnoid spaces of anesthetized cats during ventriculocis- ternal perfusion with artificial CSF containing 14C-dextran. Convective mixing of perfusate in subarachnoid spaces was augmented while infusion was constant, either by impeding cisternal efflux of perfusate by raising the cisternal out- flow cannula (high CSF pressure), or by preventing CSF outflow by clamping the cisternal outflow cannula (stop- flow; S-F). CSF transients were also measured before and after systemic administration of phenoxybenzamine (P82) in order to evaluate the contribution of sympatho-adrenergic activity to craniospinal CSF redistribution and mixing. Results from high CSF pressure and S-F experiments indicate that unequilibrated CSF contributes significantly to the reduced tracer concentration in Vd, since Cd was decreased after subarachnoid facilitated flow. Further, results from 8-? studies indicate that at least 50% of Cd is due to craniospinal fluid redistribution, a process which, along with CSF outflow transients, was unaffected by PBZ. Conversely, PBz administration decreased steady state CSF formation and absorption through alpha-mediated cerebrovas- cular responses and/or through beta-adrenoceptor inhibition of metabolism of CSF secretory epithelium. ACKNOWLEDGEMENTS I wish to thank my mentors, Dr. S. Richard Heisey, and Dr. Thomas Adams for their guidance and support throughout my graduate training. Also, thanks to Deborah Traxinger and David Manner for their collaborative and technical contributions to this research project. ii TABLE OF CONTENTS Page LIST OF TABLES O O O O O O O O O O O O O O O O O O O 0 v i 1 LIST OF FIGURES O O O O O O O O O O O O O O O O O O O Vii 1 CHAPTER I. INTRODUCTION . . . . . . . . . . . . . . . . 1 II. LITERATURE REVIEW . . . . . . . . . . . . . . 6 A. CEREBROSPINAL MENINGES . . . . . . . . . 6 1. Dura Mater . . . . . . . . . . . . . 6 2. Pia-Arachnoid . . . . . . . . . . . . 7 a. Arachnoid Villi . . . . . . . . . 8 B. CEREBRAL CIRCULATION . . . . . . . . . . 9 1. Arterial Supply of the Brain . . . . 9 a. Cephalic Arteries . . . . . . . . 9 b. Extracerebral (Pial) Arteries . . 11 c. Intracerebral (Parenchymal) Arteries . . . . . . . . . . . . 11 d. Choroid Plexuses . . . . . . . . 12 2. Venous Drainage of the Brain . . . . 14 a. External Cerebral Veins . . . . . 14 b. Internal Cerebral Veins . . . . . 15 3. Cerebrovascular Innervation . . . . . 16 a. Cephalic Vessels . . . . . . . . 16 b. Pial Vessels . . . . . . . . . . 16 c. Parenchymal Vessels . . . . . . . 19 d. Choroid Plexuses . . . . . . . . 22 iii cnaprsn Page II. LITERATURE nsvrsw (Cont.) c. CEREBROSPINAL FLUID . . . . . . . . . . . 25 1. CSF Composition and Homeostatic neChanisms O O O O O I O O O O O O O 25 2. CSF Secretory Mechanism . . . . . . . 29 3. Methods of Study . . . . . . . . . . 34 a. Sampling of Choroidal Venous Blood from in sitn Choroid Plexus . . . . . . . . . . . . . 35 b. Extracorporeal Perfusion of the Isolated Choroid Plexus . . . . . 37 c. Collection of Nascent Fluid from the Surface of Exposed Choroidal Tissue I I O O O O I O O O I O O 39 d. Ventriculocisternal Perfusion . . 40 4. Factors Affecting CSF Formation Rate 42 a. Effects of Temperature . . . . . 43 b. Effects of Osmotic Gradients . . 45 c. Effects of Blood Flow . . . . . . 49 d. Effects of Hydrostatic Pressure Gradients . . . . . . . . . . . . 53 e. Effects of C02 Tension . . . . . 54 f. Effects of Carbonic Anhydrase Inhibitors . . . . . . . . . . . 57 9. Effects of Cardiac Glycosides . . 60 h. Effects of the Autonomic Nervous System . . . . . . . . . . . . . 61 5. Factors Affecting CSF Absorption Rate 67 C. CRANIOSPINAL VOLUME PRESSURE RELATIONSHIP 68 III. HATERIALS AND NETBODS . . . . . . . . . . . . 73 A. ANESTHESIA . . . . . . . . . . . . . . . 73 B. SURGICAL AND EXPERIMENTAL PROCEDURES . . 73 C. BRAIN VENTRICULAR AND CISTERNAL PUNCTURES 75 iv CHAPTER Page III. NATERIALS AND IETEODS (Canto) D. VENTRICULOCISTERNAL PERFUSION EXPERIMENTS 76 1. Protocol A: High Pressure Perfusions 77 2. Protocol B: StOp-Flow Perfusions . . 80 3. Protocol C: Phenoxybenzamine PerfUSions O O O I O O O O O O O O O 80 E. DETERMINATION OF PERFUSION FLOW CONCENTRATIONS AND RATES . . . . . . . . 81 F. MATHEMATICAL ANALYSES AND CALCULATIONS . 81 l. Steady State Analysis . . . . . . . . 81 2. Transient Response Analysis . . . . . 83 G. STATISTICAL ANALYSES . . . . . . . . . . 87 Iv. RESULTS 0 O O O O O O O O O O O I O O O O O 0 8 9 A. PROTOCOL A: EFFECTS OF HIGH CSF PRESSURE ON CSF STEADY STATE AND TRANSIENT RESPONSES TO HYPERCAPNIA . . . . . . . . 93 B. PROTOCOL B: EFFECTS OF STOP-FLOW ON CSF STEADY STATE AND TRANSIENT RESPONSES TO HYPERCAPNIA O O O O O O O O O O O O O O O 9 9 C. PROTOCOL C: EFFECTS OF PHENOXYBENZAMINE ON CSF STEADY STATE AND TRANSIENT RESPONSES TO HYPERCAPNIA . . . . . . . . 106 v. DISCUSSION 0 O O O O O O O O O O O O O O O O 1 1 5 VI 0 meUSIONS O O O O O O O O O O O O O O O O O l 4 0 APPENDICES A. DEFINITION OF TERMS O O O O O O O O O O O O B. PREPARATION OF DIAL-URETHANE ANESTHETIC SOLUTION O O C O C O O O O O O O C O O O O C. ARTERIAL pH AND PC02 MEASUREMENTS . . . . . D. RESPIRATORY MEASUREMENTS AND CALCULATIONS . E. ANALYSIS OF CARBON DIOXIDE IN ALVEOLAR AND MIXED EXPIRED GAS C C I O O O C C C O O O O F. COMPOSITION AND PREPARATION OF ARTIFICIAL CAT CEREBROSPINAL FLUID (CSF) . . . . . . . G. DETERMINATION OF INFLOW AND OUTFLOW PERFUS ION RATES O O O O O O C O O O O O O O H. BETA RADIATION COUNTING . . . . . . . . . . BIBLImRAPBY O O O O O O O O O O O O O O O O O O O 0 vi Page 141 143 144 147 150 154 156 158 163 LIST OF TABLES Steady State Respiratory and Cardiovascular Data (Protocol A) . . . . . . . . . . . . . Steady State CSF Data (Protocol A) . . . . Transient CSF Outflow Data (Protocol A) . . Steady State Respiratory and Cardiovascular Data (Protocol B) . . . . . . . . . . . . . Steady State CSF Data (Protocol B) . . . . Transient CSF Outflow Data (Protocol B) . . Steady State Respiratory and Cardiovascular Data (Protocol C) . . . . . . . . . . . . . Steady State CSF Data (Protocol C) . . . . Transient CSF Outflow Data (Protocol C) . . vii Page 94 95 100 101 103 107 108 110 114 LIST OF FIGURES FIGURE Page 1. Schematic of Ventriculocisternal Perfusion Experimental Apparatus . . . . . . . . . . . . . 79 2. Representative Time Courses for CSF Transient Responses to Hypercapnia . . . . . . . . . . . . 85 3. Data From a Representative Ventriculocisternal PerfUSion O O O O O O O O O O O O O O O O O O O 92 4. Results of a Representative High CSF Pressure Ventriculocisternal Perfusion (Protocol A) . . . 98 5. Results of a Representative StOp-Flow Perfusion (PrOtOCOI B) o o o o o o o o o o o o o o o o o o 105 6. Results of a Representative Phenoxybenzamine Ventriculocisternal Perfusion (Protocol C) . . . 112 7. CSF Flow Patterns 0 O O O O O O I O O O O O O O 123 8. Data From Prolonged Step-Flow Perfusion . . . . 126 viii INTRODUCTION Cerebrospinal fluid (CSF) comprises all the fluid in the brain's ventricles and in the subarachnoid spaces surrounding both the brain and spinal cord. Nearly 70% of this fluid is formed from blood flowing through the ventricular choroid plexuses (Katzman and Pappius, 1973) which are a single layer of epithelial cells on a basement membrane and a highly vascularized stroma of connective tissue. The remaining 30% of CSF is formed by less well- defined intracranial extrachoroidal sites (Bering, 1959; Welch, 1963; Pollay and Curl, 1967). The primary process for CSF synthesis is the active movement of sodium from the interstitium across the epithelium into the ventricle (Wright, 1978; Davson, 1967; Pollay, 1975) which occurs in two steps; First, sodium is filtered through the Choroidal capillaries and diffuses across the basolateral membranes of the Choroidal epithelial cells down its electrochemical gradient. Second, sodium is pumped out into the ventricle by an energy dependent sodium-potassium exchange pump on the apical surface of the epithelial cells. Chloride and bicarbonate are also secreted into the ventricles and potassium is absorbed from CSF back into the blood. Water movement is coupled to the active salt secretion by osmotic 2 obligation (Wright, 1978). This newly formed Choroidal fluid is slightly alkaline compared to blood, and has a higher sodium concentration and a lower potassium concentration than expected in a simple ultrafiltrate of serum (Davson, 1956; De Rougemont et a1., 1960). It mixes with the CSF within the ventricles and circulates along a tortuous path through the ventricular cavities and into the subarachnoid spaces. During its transit, CSF is continually exposed to energy-requiring transport processes. Further modifications are made in its composition (de Rougemont et a1., 1960), which is insensitive to wide fluctuations in electrolyte concentrations in blood (Goldstein et a1., 1979; Reed and Yen, 1978). Once in the subarachnoid space, bulk absorption of CSF into venous blood occurs by the cranial and spinal arachnoid villi (Pollay and Welch, 1962; Hammerstad et a1., 1969; Davson et a1., 1970; Marmarou et a1., 1975). These absorp- tive ports are evaginations of the subarachnoid space into the lumen of the large dural sinuses, and operate like one- way valves. When CSF pressure exceeds that in the venous sinuses, the channels are distended, particles up to 7 u.in diameter pass freely (Welch and Friedman, 1960), and CSF is absorbed into the blood. When CSF pressure falls below dural sinus pressure, however, the pressure gradient is reversed, the villi collapse and CSF absorption ceases. Colloid osmotic pressure gradients between CSF and blood, unlike hydrostatic pressure gradients, have little effect on 3 CSF absorption (Pollay and Welch, 1962; Davson et a1., 1970). How CSF secretion is controlled remains unknown despite years of study. Steady state CSF formation rate is reported to increase (Ames et a1., 1965), decrease (Martins et a1., 1976), or remain unchanged (Oppelt et a1., 1963) during bouts of respiratory acidosis. Also, there are large dif- ferences in CSF production rate within a species despite control of such influential factors as blood pressure (Carey and Vela, 1974), body temperature (Snodgrass and Lorenzo, 1972), and osmotic gradients between blood, brain and CSF (DiMattio et a1., 1975; Wald et a1., 1976). Moreover, autonomic neural activity influences CSF synthesis (Edvinsson et a1., 1971; Edvinsson et a1., 1972; Lindvall et a1., 1978a; Lindvall et a1., 1978b; Lindvall et a1., 1979). Also, Choroidal epithelium is innervated by cholinergic fibers (Edvinsson et a1., 1973; Lindvall et a1., 1978) and by adrenergic nerve fibers originating from the superior cervical ganglia (Edvinsson, 1977) and from other less well- defined sources (Lindvall et a1., 1978). These sympathetic nerve terminals also innervate vascular resistance vessels (Edvinsson et a1., 1975; Nakamura and Milhorat, 1978), so that reported changes in CSF formation rate after denerva- tion or electrical stimulation of cervical sympathetic ganglia (Lindvall et a1., 1978) may be the result of either direct effects on the secretory epithelium (Edvinsson et a1., 1975; Lindvall et a1., 1982) and/or changes in blood 4 perfusion of the choroid plexuses (Lindvall and mean, 1981). Other sources of variation for CSF production were first identified by Martins and co-workers (1977). They are related to the method by which CSF formation rate is measured, and to the anatomy and physiology of the CSF spaces. Ventriculocisternal perfusion is a precise tech- nique for estimating rates of bulk fluid production and reabsorption of CSF, yet it is not error-free. Multi- chambered and convoluted CSF spaces inhibit rapid and complete mixing of perfusion fluid with the entire CSF volume. Since craniospinal CSF and blood volumes are in a dynamic equilibrium, pools of CSF unequilibrated with per- fusion fluid may be mobilized when craniospinal blood volume changes. This may explain why CSF formation rate appears to increase transiently during acute respiratory acidosis (Tschirgi et a1., 1954; Songer et a1., 1980; Fisher et a1., 1983). CSF formation rate may also increase in response to C02 breathing. Equations which assess CSF dynamics during ventriculocisternal perfusion are accurate only if cranio- spinal fluid volumes are in a steady state. (Heisey et a1., 1983; Fisher et a1., 1983). The present study was designed to define phenomena underlying the apparent transient increase in CSF formation rate during acute hypercapnia using a modified ventriculo- cisternal perfusion method. It was also designed to investigate the influence of alpha adrenergic blockade on 5 CSF transient phenomena and on steady state rates of CSF formation and absorption. LITERATURE REVIEW The soft, gelatinous brain and the slightly firmer spinal cord are rigidly confined within the skull and vertebral column, while cushioned in a jacket of secreted clear cerebrospinal fluid (CSF). The central nervous system (CNS) is further supported and protected from mechanical trauma by two connective tissue layers which lie between the honey encasements and the nervous tissue mass; the firm dura mater externally, and the delicate, thin pia arachnoid layer which adheres to the brain and spinal cord. CEBEBBQSRINAL_MENIN§ES Dura_nater The dura mater is a dense, thick layer of collagenous connective tissue. The spinal dura is tubular, pierced by spinal nerve roots, and, in humans, extends from the foramen magnum to the second sacral segment. It is separated from the wall of the spinal canal by an epidural space which contains adipose tissue and a venous plexus. The veins which drain the spinal cord empty into the epidural venous plexus, which in turn drains into an external vertebral plexus via conduits in the intervertebral foramina. The cranial dura mater is firmly affixed to the periosteum of 6 7 the cranial vault, except where it is reflected to form dural septae. The spaces formed between the septa and the periosteum.accommodate»the endothelial-lined dural venous sinuses. The external and internal cerebral veins, which lie in the subarachnoid space on all surfaces of the hemispheres and in the central core of the cerebrum, respec- tively, empty into the venous sinuses through which blood reaches the internal jugular veins. In some non-human species, the internal jugular veins are small and venous drainage from the brain mixes with blood from other structures. Pia-:Arachnoid The delicate arachnoid layer is held to the dura by the surface tension of a thin layer of fluid between the two membranes, while the thin pia mater adheres to the surface of the brain and spinal cord following all their contours. These pia-arachnoid layers constitute the leptomeninges ("slender membranes"). They originate from a single layer of mesodermal tissue, as attested by the rich trabecular meshwork passing between them; both surfaces of the arachnoid and the outer surface of the pia are covered by mesenchymal epithelium. The subarachnoid spaces within these two layers are filled with CSF and are narrow over the vertices of the gyri, but wide at the base of the brain (called the cisterna magna) and in the lumbar region of the spinal canal. 8 While the arachnoid is avascular, the pia mater contains many fine blood vessels. Large blood vessels go to the brain surface through the subarachnoid spaces before penetrating into the parenchyma, the bulk of the brain tissue, taking with them a sleeve of pia mater. Within the sleeve, the subarachnoid space and its contained CSF extend into the Virchow-Robin spaces, becoming progressively attenuated as the vessels become increasingly smaller, until finally the perivascular spaces are only a fraction of a micron long around the capillaries. The subarachnoid space also extends into and communicates with the subdural space through the arachnoid villi, which are labyrinths of coapted tubes that connect with each other and allow open communication between the subarachnoid and venous channels of the dura. W11. The arachnoid villi, which may be the main site of CSF absorption into venous blood (Welch, 1975; Marmarou et a1., 1975), are composed of a thin cellular layer which is derived from arachnoid mesenchymal epithelium and the endothelium of the sinus. Its cells have a sparse amount of granular cytOplasm which surrounds round or oval nuclei. As the cytoplasm extends from the nucleus it becomes flattened and forms a delicate tubular wall. The tubular projections enclose an extension of the subarachnoid space containing trabeculae. The arachnoid tubes form a communication channel between the subarachnoid space and 9 venous sinus (Welch and Friedman, 1960; Pollay and Welch, 1962). Descriptions of the architecture of the arachnoid villi vary depending on how it was examined. ‘Welch and Friedman (1960) fixed arachnoid villi from the African green monkey by arterial or venous perfusion of formalin, and then performed serial sections after the tissue had been dissected and embedded in paraffin. They found that if the fixation occurred during low cerebrospinal fluid pressure, brought about by cisternal drainage, the normal tubular structure of the arachnoid villi collapsed. The diameters of the tubular villi decreased, and the villi cells were less flattened and bulged into the tubular lumen. Conversely, they found that if the villi.were fixed by perfusion of formalin into the subarachnoid space under unphysiologically high pressures, the villi distended and deformed, while the cytoplasm of the cells was greatly attenuated. Under normal CSF and subarachnoid pressures, the villi were found to be tubular, i.e., "open“ but not distended, and communication was free between the subarachnoid and dural spaces. CEBEBRAL.QIBQULAIIQN WW Cephalig_brt§rigs, Except for a minor contribution from the anterior spinal artery to the medulla, blood is carried to the brain via two routes: the common carotid and 10 the vertebral arteries. The common carotid artery divides into the external and internal carotid arteries. 'The external carotid artery gives off several branches to supply blood to the extracranial tissues. The distal part of the internal carotid artery or the beginning of the middle cerebral artery (a collateral branch of the internal carotid) gives rise to the anterior Choroidal artery, which has a rather long subarachnoid course and which supplies blood to the lateral ventricular choroid plexuses as well as some other cerebral structures. In the adult cat, however, the internal carotid artery is vestigial and the brain is supplied with blood through the external carotid, vertebral and pharyngeal arteries by way of the internal maxillary artery and the carotid plexus (Crouch, 1949). The vertebral arteries unite to form the basilar artery on the ventral side of the spinal cord. This artery gives off branches, including the posterior cerebral artery and the posterior choroidal artery, which supplies the choroid plexuses of a small portion of the rostral end of the lateral ventricle and the third ventricle. The posterior cerebral artery finally joins the posterior part of the Circle of Willis, which is a network of communicating arteries that provides blood to the ventral side of the brain in the region of the pituitary. The blood is further dispersed from these arteries to the other extracerebral bed (pial vessels), which forms a reticulation on the outer 11 surface of the brain, and then proceeds to the intracerebral vessels (in the brain parenchyma). W. The pial vessels, which lie in the subarachnoid space, are bounded on their outer surface by a thin adventitial layer which is composed of a layer of loose connective tissue intermingled with a dense plexus of adrenergic and cholinergic nerves. A basement membrane surrounds this layer and confines the adventitial space from the perivascular space. Toward the center, the adventitial space is adjoined to a muscular cell layer, the tunica media, which, in turn, adjoins an innermost single layer of endothelial cells, the tunica intima. The pial vessels distribute blood to the vessels which penetrate the cerebral cortex, the intracerebral or parenchymal vessels. W. The parenchymal vessels, which have a short extracerebral course before penetrating the brain parenchyma, become smaller as they move inward from the brain surface. There are three main types: arterioles, capillaries, and venules (Cervos-Navarro and Matakas, 1973). These vessels have a number of unique anatomic features which distinguish them from the peripheral microvasculature and which are considered to constitute the morphological blood-brain barrier (mean and Hardebo, 1982). In most regions of the brain, the vascular endothelial cells are bounded by a basement membrane, are non-fenestrated, are fused to each other by tight junctions (i.e., intercellular 12 clefts are sealed shut), and demonstrate a paucity of pinocytotic vesicles in their cytoplasm (Oldendorf, 1975; Edvinsson, 1975); vesicular transport is presumed to be virtually absent (Oldendorf, 1975). Further, the brain capillaries are sometimes ensheathed by the membrane of astrocytic end-feet (pericytes) which are apposed to the capillary basement membrane. Whether or not a molecule in solution in plasma will penetrate the capillary cell wall (traverse the blood-brain barrier) depends on its oil/water partition coefficient, and its relative affinity for plasma and membrane carrier proteins (Oldendorf, 1975). The integrity of the blood-brain barrier is enhanced by special enzymatic systems found in the cytoplasm of brain capillary endothelial cells (mean and Hardebo, 1982). Specifically, monoamine oxidase (MAO), catechol-o- methyltransferase (CONT), DOPA-decarboxylase, as well as other enzymes, show considerable activity in brain capillaries (Hardebo et a1., 1980). Monoamines such as dopamine, norepinephrine, epinephrine and S-hydroxytrypta- mine, and other less well-known neurotransmitters may be deaminated in the walls of brain vessels (Hardebo and mean, 1982). Choroid_£lexnses, The choroid plexuses, a special part of the pial vascular system, also have MAO activity, although the bulk of the enzyme activity is present in choroidal epithelial cells rather than in the endothelial lining of the choroidal capillaries (Lindvall and mean, 1980). ‘The l3 plexuses in each of the lateral ventricles and in both the third and fourth ventricles are richly vascularized epithelial tissue formed by an invagination of the pia mater (tela choroidea). The lateral ventricle is formed by an invagination of pia mater on the medial surface of the cerebral hemisphere, while the plexuses of the third and fourth ventricles are similarly formed by invagination of pia mater attached to the roofs of these ventricles. During the invagination process, the connective tissue acquires a covering of epithelium from the ependymal lining of the ventricles. Three distinct elements comprise the choroid plexus: (l) epithelium of ependymal origin which has secretory characteristics; (2) a layer of leptomeningeal tissue which is ectodermal in origin and is characterized by large aqueous spaces, consistent with the suggestion that these spaces are fundamentally extentions of the subarachnoid space (Maxwell and Pease, 1956); and (3) a dense vascular stroma. The choroidal vascular stroma contains capillaries which are anatomically'distinct from.most intracerebral capillaries. 'The thin endothelium, which rests on basement membrane like that of the peritubular capillaries of the kidney (Maxwell and Pease, 1956), has "leaky-tight“ junctions and fenestrations or pores closed by thin diaphragms (Wright, 1979). These increase the permeability in these capillaries more than is normally found in capillaries which supply nervous tissue. In fact, there is 14 free exchange of elements between the blood and the interstitial fluid bathing the basolateral surface of the choroidal epithelium. The morphological blood-brain barrier is lacking (Wright, 1979). Cerebrospinal fluid (CSF) is ultimately derived from blood flowing through these choroidal capillaries. Electrolytes and water cross the capillary endothelial fenestrations and enter the extracellular fluid space of the choroidal epithelial cells where they are selectively'and actively secreted into the ventricular cavities (Wright,“ 1979). However, choroidal blood flow, of which 25% of the plasma volume becomes CSF (Cserr, 1971), is nearly twelve times higher than cortical blood flow in normoxic, normotensive sheep (Page et a1., 1980) and can be reduced as much as 50% without affecting CSF formation rate (Ff) (Nakamura and Hochwald, 1983). W The cerebral hemispheres are drained by both an external and an internal venous system. The external cerebral veins course through the subarachnoid space and lie on all surfaces of the cerebral hemispheres. The internal cerebral veins which are located beneath the corpus callosum in the transverse fissure drain the central core of the cerebrum. W. The superior cerebral veins and the superficial middle cerebral vein, which are external 15 cerebral veins, course upward over the dorsolateral surface of the hemispheres, and downward and forward along the sylvian fissure, respectively. They pierce the arachnoid and dura mater, and empty into the superior sagittal, cavernous, and transverse sinuses, and the lateral lacunae adjacent to the superior sagittal sinus. The basal vein, formed by the union of the deep middle and anterior cerebral veins, runs backward along the base of the brain and eventually empties into the Great Cerebral Vein of Galen. This latter vein, which is considered to be an internal cerebral vein, drains into the straight sinus. Many more external veins, in addition to the few mentioned, drain circumscribed regions of the brain and channel into adjacent dural sinuses. W. The internal venous system forms beneath the floor of the lateral ventricles and continues along under the corpus callosum. The choroidal vein, a rather tortuous vessel, passes along the choroid plexus of the lateral ventricle and drains the plexus as well as a few neighboring structures. This vein unites with the terminal vein, which drains the corpus striatum, internal capsule, fornix, and septum pellucidum, to form the internal cerebral vein. The paired internal cerebral veins course posteriorly in the tela choroidea and unite beneath the corpus callosum to form the Great Cerebral Vein of Galen. 16 The cerebral venous system has larger radii and thinner walls, and is more compliant than cerebral arteries (Marmarou et a1., 1975). The cerebral blood volume, determined primarily by the volume of blood contained in the compliant cerebral venous system (Rich et a1., 1953), exists in a dynamic equilibrium with the cranial CSF volume. Sudden changes in one of these volumes elicits a rapid reciprocal change in the other fluid volume (Ryder et a1., 1953). Since the cerebral veins, in particular the external cerebral veins contained in the CSF—filled subarachnoid space, are highly compliant compared to the cerebral arteries (Marmarou et a1., 1975) and brain (Sklar et a1., 1977), intracranial pressure changes induced by sudden changes in cranial CSF or blood volume are damped primarily by the rapid reciprocal interaction of these two cranial fluid volumes (Marmarou et a1., 1975; Weed et a1., 1932;, Weed et a1., 1933; Ryder et a1., 1953). W Cephalie_!£ssels. Nerve fibers were first described on the anterior and posterior cerebral arteries by Willis in 1664 (cited by Edvinsson, 1975) using silver-impregnated tissue sections. Confirming studies have traced the origin of many of these fibers to the superior cervical ganglion (Edvinsson and MacKenzie, 1975). W. Pial vessels of many species of animals (Nielsen et a1., 1971), including the cat (Edvinsson et a1., 17 1972), have well-developed plexi of sympathetic adrenergic nerves and cholinergic nerves, as well as some tryptaminergic and peptidergic fibers (mean and Edvinsson, 1977). The sympathetic innervation, which arises from the superior cervical ganglia (Edvinsson, 1975) is comparable in density, distribution, and organization to that found in other richly innervated peripheral vascular beds (Nielsen et a1., 1971; Edvinsson and MacKenzie, 1977; Edvinsson, 1975). Electron.microscopy studies have revealed that the sympathetic nerves are less dense in cerebral veins, compared to cerebral arteries (Edvinsson and MacKenzie, 1977), and in the phylogenetically older regions of the brain (the posterior part of the cerebrum and brain stem), as compared to the younger parts of the brain (neocortex) (Edvinsson, 1975). The extent and distribution of cholinergic nerves to the pial vessels is similar to that of the adrenergic supply (Edvinsson et a1., 1972). In fact, cholinergic and adrenergic nerves frequently run together surrounded by the same Schwann cell, terminate in juxtaposition, and form a typical ground plexus in the adventitia immediately adjacent to the muscular layer of the media. This provides an opportunity for an interactive mechanism between the two types of axon terminals (Nielsen et a1., 1971). The origin of the cholinergic fibers, however, is unknown. It has been suggested that the majority of cholinergic fibers track in the facial nerve, course off in the region of the geniculate 18 ganglion, and continue in the greater superficial petrosal nerve to the plexus of the internal carotid artery where they form ganglia and project to the pial vessels (Edvinsson, 1975). The predominant adrenergic receptor on pial vessels is the alpha-adrenoceptor, mediating vasoconstriction. Noradrenaline induces dose-dependent vasoconstrictor responses in isolated pial vessels which can be reduced or blocked by specific alpha-receptor blockers, phentolamine or phenoxybenzamine (Edvinsson and mean, 1974). The cerebral vessel alpha-adrenoceptors, however, may be distinct from the alpha-adrenoceptors associated with peripheral arteries. The cerebral receptors are only weakly sensitive to noradrenaline and phenylephrine, and demonstrate an atypical biphasic response in the dose-response curve for noradrenaline (Duckles and Bevan, 1976). The weak sensitivity of pial vessels to noradrenaline may be due either to a limited capacity to introduce extracellular calcium and/or to mobilize intracellular calcium, or to an increased activity of the neurotransmitter uptake system in the sympathetic nerve endings (Marin and Rivilla, 1982). Beta-adrenoceptors have also been identified in pial vessels and have been implicated in the vasodilation seen when isoproterenol is introduced into isolated cerebral arteries pre-treated with serotonin or prostaglandins; propranolol (10"8 - 10'7 M) blocks the vasodilation (Edvinsson and mean, 1974; Sercombe et a1., 1975). 19 25W. Like pial vessels, parenchymal vessels, including intracerebral capillaries and their associated pericytes (Rennels and Nelson, 1975), are supplied with adrenergic nerves (Kawamurs et a1., 1972; Cervos- Navarro and Matakas, 1973). These nerves, however, may arise from a source other than the peripheral autonomic ganglia. In 1972, Hartman and Udenfriend suggested that small blood vessels are innervated by a central adrenergic system, which is part of an intra-axial autonomic system that is isolated by the blood-brain barrier (Raichle et a1., 1975) and distinct from the peripheral sympathetic system, but having analogous functions. Their hypothesis was based on the immunohistochemical demonstration that bilateral cervical ganglionectomy in rats did not disrupt the close association between dOpaminejd-hydroxylase (DBH)-containing fibers and capillaries found in normal rats. They suggested that since there are no (DBH)-containing cell bodies rostral to the pons, the vascular innervation in the rostral brain arrives via ascending noradrenergic fiber pathways from the locus coeruleus and other clusters of noradrenergic cell bodies in the lower brainstem. Subsequently, several studies investigating the influence of the central adrenergic system on cerebral circulation have been conducted. Raichle et a1. chemically' (1975) and electrically (1976) stimulated the locus coeruleus, a major cell group of the central adrenergic 20 system (Hartman and Udenfriend, 1972), and measured the effects on capillary permeability and regional cerebral blood flow. When stimulating the locus coeruleus either electrically'(l976) or by microinjections of carbachol (1975), there was a marked and rapid increase in brain capillary permeability to water and a reduction in blood flow; intraventricular administration of phentolamine, a rapidly acting alpha adrenergic blocker, produced the opposite effect (1975). Preskorn and associates (1978) demonstrated similar phenomena by measuring the extraction fraction of a diffusion-limited test tracer both before and after the acute intraperitoneal administration of amitriptyline (AMI) to rodents and monkeys. AMI, a tricyclic antidepressant, acts indirectly as an adrenergic agonist by blocking norepinephrine reuptake. They found that after AMI treatment, the extraction fraction of both water and ethanol increased; pre-treatment with phenoxybenzamine or intraventricular administration of 6-hydroxyd0pamine blocked the effects. In a later study, Preskorn et a1. (1982) provided further evidence that the AMI-induced increase in the extraction fraction of water is mediated by the effect of the drug on central adrenergic neurons rather than through the known serotonergic, anticholinergic, and antihistaminergic actions of AMI, by themselves: central serotonergic abalation by p-chloroamphetamine, and treatment 21 with atropine and hydroxyzine separately did not alter the extraction fraction of water. These data imply a neural control of the blood-brain barrier at the level of the capillary (Preskorn et a1., 1978; Grubb and Raichle, 1981; Preskorn et a1., 1981; Preskorn et a1., 1982). Further evidence for central adrenergic effects on the cerebral microcirculation was provided by immunohistochemical demonstration of actomyosin in pericytes and endothelial cells in the brain capillary wall (mean et a1., 1977). The integrity of the blood-brain barrier may also be affected by the peripheral sympathetic nervous system. Sadoshima and co-workers (1982) showed that unilateral superior cervical ganglionectomy in stroke-prone spontaneously hypertensive rats (SHRSP) increases the permeability of the blood-brain barrier to the Evans blue- albumin complex and/or increases the incidence of cerebral infarction or hemorrhage in the denervated hemisphere. They concluded that the peripheral sympathetic nervous system protects the blood-brain barrier during chronic hypertension, and that disruption of the blood-brain barrier may play a role in the pathogenesis of cerebral infarction by allowing vasoactive substances to enter the brain and produce focal areas of cerebral ischemia. Sadoshima et a1. speculated that the mechanisms of this protective effect may be through a trophic effect of sympathetic nerves which may accentuate cerebral vascular wall thickening (hypertrOphy) 22 during chronic hypertension. Hypertrophy of large cerebral arteries may augment the structural contribution to cerebral resistance, reduce pressure in small cerebral vessels, and therefore protect the cerebral microcirculation. Further, hypertrOphy of small resistance vessels may reduce the wall stress of these vessels. Hardebo et al. (1977a) extended the postulated role of the peripheral sympathetic nervous system in protecting the cerebral circulation to include the modification of cerebral autoregulatory limits during hypertensive episodes. They showed, as have others (Eklof et a1., 1971; Hernandez et a1., 1971/72; Hernandez-Perez et a1., 1975; Waltz et a1., 1971; Stone et alr, 1974), that superior cervical ganglionectomy does not abolish autoregulation, but causes the denervated hemisphere to autoregulate at a higher blood flow level. QhQLQid_£1£anes, .Morphological evidence suggests extrinsic neural regulation of choroidal secretion of CSF. The choroid plexuses, like the pial vascular system, are well innervated with autonomic nerves (Edvinsson et a1., 1973; Edvinsson et a1., 1974; Edvinsson et al., 1975b; Lindvall et al., 1978b; Edvinsson and Lindvall, 1978; Nakmura and Milhorat, 1978). The axons enclose not only the choroidal arteries, but also form a reticulation in the choroid plexus and around the veins which drain the tissue (Edvinsson and Lindvall, 1978L. Using a histo-fluorescence method in combination with superior cervical ganglionectomy, Edvinsson and associates (1974) demonstrated the existence 23 of nerve fibers showing a fluorescence typical for that of catecholamine fluor0phores in the choroid plexuses of various mammalian species. The nerve fibers formed networks around both the choroidal arterial and venous vessels, while others coursed between the base of the choroidal epithelial cells and the underlying vascular wall. Within one week after bilateral cervical sympathectomy, the nerve fluorescence disappeared from the choroid plexuses of the lateral and the third ventricles, while a small number of fibers were still visible in the plexus of the fourth ventricle (Lindvall et al., 1978b; Edvinsson et al., 1975a). In a subsequent study, Edvinsson and co-workers (1975b) found corroborative ultrastructural evidence of autonomic innervation of the choroid plexus from fluorescent and electron microscopy of cat and rabbit plexuses after administration of S-hydroxydopamine. They found thick bundles of nonmyelinated axons in the proximal portions of the plexuses where they accompanied small arterioles, and smaller axon bundles located contiguous to the walls of small arterioles as well as to therbase of the epithelial cells, between adjacent cells, or within cellular invaginations. The distance between the membrane of the perivascular nerves and the membrane of the smooth muscle cells of the arterioles was less than 100 nm. ‘Axonal- epithelial cell distances were sometimes as small as 30 nm. The nerve terminals had no neurilemmal sheath but had an abundance of electron-dense vesicles (indicative of 24 adrenergic terminals) or of electron-lucent vesicles (suggestive of cholinergic terminals) of the same size. They concluded that the sympathetic vasomotor innervation of the choroid plexus resembles that of the pial and intracerebral vascular beds, and that the close apposition of nerve cells to plexus epithelia provides morphological evidence for true autonomic innervation. Histochemical studies of the choroid plexuses of superior cervical ganglionectomized rabbits and cats confirmed the presence of numerous acetylcholinesterase- containing nerve terminals (Edvinsson et a1., 1973). The conditions under which the cholinesterase staining developed in the plexus preparations under study (i.e., preincubated in 4 uM solution of the cholinesterase inhibitor Mipalox and incubated for 4 to 6 hours in the presence of acetylthiocholine) suggested that the cholinergic nerves were parasympathetic rather than sensory. The cholinergic nerve fibers were found to form a wide-meshed network around most vessels supplying and draining the plexus, and to ramify among epithelial cells without obvious relation to the subjacent vascular system. Edvinsson and co-workers concluded that the distribution of the cholinergic nerves resembled that reported for the adrenergic system (Edvinsson et a1., 1974) and that, in fact, the overall density of nerve contribution to plexus tissue was even greater for the cholinergic system. 25 CEBEBBQSEINAL_ELHID_1£5£1. CSF is known to be produced by the choroid plexuses of all four ventricles (de Rougemont et a1., 1960; Welch, 1963; Ames et a1., 1965; Wright, 1978), the ependymal surface of the brain (Pollay and Curl, 1967), and perhaps in the cranial (Sato and Bering, 1967) but not the spinal subarachnoid spaces (Hamer and Sahar, 1978; Lux and Fenstermacher, 1975). In experimental animals, 55 to 65% of the CSF formed within the ventricles is reportedly produced by the choroid plexus in the fourth ventricle (Wright, 1978). CSF, which is slightly hypertonic to plasma (Wright, 1979 ; Cserr, 1971), flows presumably along regional pressure gradients established by fluid secretion from the lateral ventricles through the third ventricle to the fourth ventricle where it enters subarachnoid cisterns surrounding the brain via the foramina of Luschka and Magendie. It then flows over the brain convexities and down the spinal cord where it is absorbed through the cranial arachnoid villi and spinal nerve roots, respectively'(Pollay and Welch, 1962; Hammerstad et a1., 1969; Davson et a1., 1970; Marmarou et a1., 1975). ESE: 'I‘ in III Mechanism CSF closely resembles an ultrafiltrate of plasma, however, both anatomical and functional evidence clearly indicate that CSF production is an active mechanism and not a simple ultrafiltration process. Several characteristics 26 of the choroid plexus epithelial cell seen in electron micrographs suggest that the production of CSF is at least in part an energy-requiring active process. There is a large spherical nucleus, numerous and randomly arranged mitochondria, and a slightly polarized system of vesicles (representing end0plasmic reticulum) in an abundant cytoplasm; the vesicles are both more numerous and often large near the apex of the cell (Maxwell and Pease, 1956). Further, the choroidal epithelial plasma membrane, which has a low hydraulic conductivity (Wright, 1978), has the lateral aspects of its apical surface tightly adjoined to adjacent cells (Wright, 1979) forming a barrier to solutes between the CSF and the cellular interstitial spaces; filtration cannot play a significant role in the passage of fluid across this layer. The apical surfaces of the epithelial cells are lavishly folded, as.are the basal surfaces, and specialized into a fairly well-developed brush border (Maxwell and Pease, 1956). ‘This folding vastly increases surface area, and is common to epithelial cells involved in active secretory processes. Chemical analyses of CSF support the idea that CSF elaboration involves active processes. In 1960, de Rougemont et a1. published the first of a series of papers in which they analyzed newly formed choroidal fluid, and compared it to that of serum and to that of CSF obtained from sites progressively further along its circulation course (ldh, cisternal magna and cisterna pericallosa). 27 Their technique for sampling from the choroid plexus involved the surgical exposure and subsequent filling of the lateral ventricles of a cat with PantOpaque warmed to 37°. Since the choroid plexus and its secretions have lower specific gravities than PantOpaque, they rose in the pool of oil; permitting their collection of newly formed CSF droplets with a fine glass pipette. This fluid had a lower potassium and a higher concentration of sodium and chloride than expected in a simple ultrafiltrate of serum. Further, although there are no detectable differences in Na+ concentration among samples from the ventricular and cisternal collection sites, fluid collected from the cisternae contained even higher chloride than plexus fluid, and potassium concentration progressively declines from the choroid plexus to cisterna magna to the cisterna pericallosa. de Rougemont and associates interpreted these findings to indicate that CSF is actively elaborated from serum by transport processes located at the site of its bulk formation, and that CSF continues to be exposed to the action of energy-requiring transport processes as it circulates along a course between the ventricles and the subarachnoid spaces. These active mechanisms produce further modifications in CSF in the same direction as those initiated during its choroidal formation process. They postulated that extrachoroidal transport mechanisms reside in cells that have a similar embryological origin to that of the choroid plexus. 28 Ames et a1. (1964) also examined the composition of CSF collected from the exposed ventricular choroid plexus to that which had contacted the ependyma and pia mater. Their data agreed with those published by de Rougemont: CSF collected from the choroid plexus and the cisterna magna have concentration differences in four ions (chloride, magnesium, calcium and potassium) while sodium concentration remained constant between the two sites. Chloride, magnesium, and calcium concentrations increase as CSF traverses the ventricular system, while the potassium concentration declines. Ames et al. also estimated bicarbonate concentrations based on charge balance and determined that choroidal bicarbonate is greater than cisternal bicarbonate concentration. These regional differences in concentrations of potassium, magnesium, and calcium in CSF stay constant despite wide fluctuations in plasma electrolyte composition (Goldstein et a1., 1979; Reed and Yen, 1978). Ames et a1. (1965) varied plasma potassium concentration from 1.59 to 11.08 mM/kg H20 and found that CSF potassium changed less than 2 mM. Reed and Yen (1978) investigated the regulation of CSF magnesium concentration using the cat choroid plexus isolated in a chamber in_sitn, and varied the magnesium concentration from 0 to 4.8 mEq/liter in the plasma and/or the solution bathing the plexus. They observed that chamber magnesium concentrations were rapidly returned to the control value of 1.8 mEq/liter, and concluded that the 29 choroid plexus is directly involved in maintaining the constancy of CSF magnesium concentration by sensing changes in the normal CSF magnesium concentration and altering the rate of active magnesium secretion. Subsequently, Goldstein et a1. (1979) arrived at similar conclusions about the role of the choroid plexus in CSF calcium regulation. They simultaneously measured total and ionized calcium and inorganic phosphorous in blood and CSF in human subjects and dogs with and without parathyroid glands, in dogs which had been treated three days with either parathyroid extract or phosphate infusion, or in dogs which were made and maintained hypercalcemic for 3 months with oral administration of vitamin D. IDespite marked variations in the blood concentrations of total calcium, ionized calcium, inorganic phosphorous, and of the calcium- phosphorous product, these concentrations and products varied only slightly in CSF under all conditions tested. These studies show that CSF is at least partly actively secreted, and suggest that CSF is contiguous with the immediate environment of neural cells, the brain extracellular fluid. CSF electrolyte levels are actively maintained constant providing a critial step in the overall process which prevents changes in brain composition (Wright, 1978). W Considering the architecture of the choroidal capillaries and the overlying epithelium, filtration may be 30 an important process in moving fluid from capillary lumen to interstitial space of the plexus and a necessary prelude to an active transepithelial secretion (Welch, 1975). Pollay (1975) develOped a model which provided a basis for the current understanding of the CSF production process. He applied the standing gradient hypothesis of solute and water transport (Diamond and Bossert, 1967) to the choroidal ependyma. Considering the tight junctions connecting adjacent choroidal ependymal cells as well as the large fenestrated capillaries in the plexus, he proposed that the first step in the formation of CSF is the transcapillary movement of an ultrafiltrate of plasma under the influence of a hydrostatic gradient. The filtrate, which is isotonic to blood, would then traverse the basal foldings of the choroidal cell, or between choroidal cells proximal to the apical tight junctions; the large particles are filtered by the basement membrane. Active transport of sodium from these channels renders the local interstitial fluid hypotonic as it approaches the blind end of the channel, and a favorable osmotic gradient is established for water flow into the cell. A similar mechanism would couple solute and solvent.movement into the ventricular system through the apical microvilli channels. Pollay proposed that CSF synthesis involves both filtration and the active transport of sodium as primary processes in CSF secretion (Wright, 1979). The problem with his theory, however, rests in the fact that it predicts basolateral 31 sodium pumping and explains water flux upon the osmotic gradient induced by these pumps. Wright (1978) using autoradiographic techniques studied the binding of 3H- ouabain to its natural substrate, sodium—potassium adenosine triphosphatase (Na/K-ATPase) which is now recognized as the sodium pumping mechanism (Wright, 1978). He determined that Na/K-ATPase, which represents one-fourth to one-third of the total ATPase activity in the plexus (Lindvall et a1., 1982), was largely confined to the apical cell membrane, and not to the basolateral membrane as Pollay“s theory would require. In fact, the ouabain-sensitive influx of sodium across the serosal surface of the plexus has been demonstrated to be only 7% of that across the apical plasma membrane (Wright, 1978). .Adenylate cyclase, an enzyme which converts adenosine triphosphate (ATP) to adenosine 3', 5“- monophosphate (cyclic AMP), is involved in the mediation of both alpha— and beta-adrenergic effects (Exton, 1982; Nathanson, 1979; Epstein et a1., 1977), and alkaline phosphatase (Lindvall et a1., 1982) has also been localized on the apical plasma membrane of choroidal epithelial cells. The basolateral membranes, though relatively devoid of sodium pumps, are not inactive. Guanylate cyclase, an enzyme that converts guanine triphosphate (GTP) to guanosine 3', 5'-monophosphate (cyclic GMP), a compound which has been shown to decrease the binding of agonists to cortical alphaz—receptors (Exton, 1982), and glutamyl transpeptidase, an enzyme active in transmembrane amino acid transport, 32 have been localized on the basolateral membrane CLindvall et a1., 1982). The entry of sodium into the epithelium across the basolateral membrane does not need to be active. The electrochemical potential gradient for sodium influx, in fact, favors its passive movement into plexus cells (i.e., .‘fiua+ is approximately 150 mv) (Wright, 1978; Wright, 1979; Welch, 19750. Once within the epithelium, however, sodium is pumped out in the ventricle by Na/K-ATPase (Wright, 1979). The Na/K pump is ouabain-sensitive (Wright, 1978), and is electrogenic. It contributes approximately 10 mv to the electrical potential difference across the apical membrane of the plexus; 3 sodium ions are pumped out of the cell into the ventricles in exchange for two potassium ions. Potassium accumulation within the plexus is against an electrochemical gradient (approximately 57 mv) (Wright, 1979), and its leakage across the basolateral membrane accounts for the observed net absorption of potassium from CSF to blood (Wright, 1978; Wright, 1979). The rate of sodium-potassium pumping is limited by the entry of sodium into the epithelium across the basolateral membrane. Sodium influx is accelerated by the presence of an increased supply of intracellular protons which couple with sodium in an antiport process, and of intracellular buffer anions, especially'bicarbonate (Wright, 1978), and chloride to accompany sodium transport out of the cell across the apical membrane to maintain electroneutrality 33 (Curl and Pollay, 1967; Wright, 1978; Wright, 1979; Welch, 1975). Bicarbonate can enter the cell in the form of C02 where it is hydrated, a process catalyzed by carbonic anhydrase, to form carbonic acid. Carbonic acid is an unstable compound and forms bicarbonate and hydrogen ions. Acetazolamide, a carbonic anhydrase inhibitor, causes a decrease in CSF formation rate (Ames et a1., 1965), possibly by limiting the number of intracellular protons to exchange for sodium at the apical epithelial membrane (Wright, 1979). With pumps only on the apical membrane, the link between active salt transport and fluid secretion remained an enigma. Wright (1977) isolated the fourth ventricular choroid plexus of the frog, removed all the solution from the ventricular surface of the plexus, covered it with oil, and collected the newly secreted fluid from the tissue/oil interface at regular intervals of time. He found that when the serosal side of the frog choroid plexus was exposed to a Ringer solution containing 25 mM sodium glycodiazine and equilibrated with 100% 02 (viz., control conditions), the nascent CSF was hypertonic to the serosal bathing medium, due primarily to an increase in the sodium concentration of the newly formed CSF. Further, Wright discovered that when the osmolarity of the bathing solution was varied from 110 to 290 mOsm/liter by varying the sodium chloride concentration, the osmolarity of the newly secreted CSF remained hypertonic to the bathing solution by approximately 30 mOsm. 34 A local osmotic pressure gradient, established by active sodium transport, may provide the link between solute and water transport (Wright, 1975). The local gradient may be perpetuated by the presence of a large “unstirred“ CSF layer adjacent to the ventricular surface of the tissue (Wright and Prather, 1970). Active salt transport across the epithelium, then, increases the osmotic pressure in this unstirred layer, thereby passively pulling water from the serosal side of the cell. The exact mechanism of solute- solvent coupling remains the subject of debate in this and other epithelia specialized for water transport. MethodsJLSthdx Numerous in gitrg and in vigg methods have been used to study choroidal and brain ependymal transport processes as well as to estimate rates of CSF formation and absorption. Principal methods for assessing CSF fluid dynamics include: 1) sampling of choroidal venous blood from in_aitn choroid plexus (Welch, 1963); 2) extracorporeal perfusion of the isolated plexus (Pollay et a1., 1972); 3) collection of nascent fluid from the surface of exposed choroidal tissue (deRougemont et a1., 1960); and 4) ventriculocisternal perfusion (Pappenheimer et a1., 1962; Heisey et a1., 1962). Of these, the ventriculocisternal perfusion system has been the most widely used means of studying rates of CSF formation (Vf) and absorption (Va) (Cserr, 1977). 35 1. Sampling of Choroidal Venous Blood from 1n_aitn Choroid Plexus In 1963, Welch developed an approach for determining CSF formation rate which involves the estimation of the blood flow through the choroid plexus and the loss of plasma fluid volume during the transit of blood. He observed that the vascularization of the rabbit}s lateral ventricular choroid plexus favors the direct measurement of venous blood flow, since the venous drainage route is chiefly through one large vein, the internal cerebral vein. He surmised that if red blood cells do not appreciably change in volume as they pass through the choroid plexus, then the volume of red blood cells entering and leaving the plexus during any period are necessarily constant. Hcta x Qa = Hctv x Qv (l-l) where, Hct = hematocrit of choroidal blood Q = choroidal blood flow (ul/mg-min'l) arterial a V venous Welch assumed that the fluid loss from the blood in its transit through the rabbit plexus, (i.ea, the difference between choroidal arterial and venous flow)*would represent CSF formation (Vf; ul/min). 6: = Oa - 0v (1-2) 36 Substituting Eq. 1-1 into Eq. 1—2 and rearranging, CSF formation rate can be estimated if choroidal venous blood flow (0v), and choroidal arterial (Beta) and venous (Hctv) hematocrit are measured. Hct Hct Vf = 0v [ v - 1] (1-3) a By the use of an eyepiece micrometer, the diameter and a length of the main choroidal vein were measured. A micropipette filled with 1-octanol was introduced into the vein and spherules of 1-octanol were injected. The linear velocity of the spherules was determined using a cinematographic technique from film projections on graph paper. Blood flow (Ov) in the main choroidal vein was then determined by: Qv = (nr2 x l/t) - (ngfir3) (1-4) where, wrz = cross-sectional area of the vein (cmz) l/t = the linear velocity (cm/min) n a the number of spherules injected during one unit of time (number/min) gwr3 = volume of l-octanol spherules injected (cm3) Venous blood from the plexus and blood from the aorta were collected in capillary tubes and hematocrits determined by centrifugation and microscopic inspection. This method permits the in 111g measurement of choroidal CSF formation rate, yet may have inherent sources 37 of error in at least one of its assumptions. Welch sampled from the aorta to determine anterior choroid arterial hematocrit;(Hcta)*when it is known that small cerebral arteries have a smaller hematocrit than that of aortic blood (Cserr, 1971); this would lead to an underestimation of transependymal volume flow. 2. Extracorporeal Perfusion of the Isolated Choroid Plexus In 1972, Pollay and his associates determined that the choroid plexus of sheep can be maintained satisfactorily in an extracorporeal perfusion system for up to 7 hours. They procured heads from freshly decapitated sheep, rapidly entered the skull and removed the brain, being careful to section the supraclinoid portion of the internal carotid artery. 4All branches of the carotid artery were ligated, except for the branch to the anterior choroidal artery, so that arterial inflow to the plexus could be controlled by carotid perfusion. Venous outflow from the plexus was collected from a cannula introduced into the Great Vein of Galen. The vascular perfusate consisted of donor sheep blood which had been diluted with a salt solution containing dextran (10% solution); osmolality'was adjusted with sucrose while the hematocrit remained approximately one-half that of whole blood. The fluid bathing the plexus was artificial CSF similar in composition to that of sheep CSF. 38 The method of analysis and determination of CSF formation rate (Vf) is similar to that used by Welch (1963), and depends on the constancy of red cell volume between choroidal arterial and venous blood. However, Pollay and co-workers could measure blood flow into (Oa) and out of (Ov) the perfused plexus by determining the time required to fill 100 pl precision volumetric pipets, and measured choroidal arterial and venous hematocrits. Although it was possible to measure directly 0a, Pollay chose to compute it since there was uncertainty as to the quantity of blood lost to brain from the anterior choroidal artery prior to its entrance into the inferior tip of the choroid plexus. The actual blood flow into choroidal tissue from the anterior choroidal artery can be calculated from the venous outflow and arterial (Hcta) and venous (Hctv) hematocrits as: . . Hctv = v Qa Q Hcta (2-1) Further, the volume flow rate of CSF is the inequality between arterial (Qa) and venous (Qv) inflow and outflow, respectively. CSF formation rate (Vf;‘u1/min) can be calculated from: Hct Hct V Vf = Qa - Qv = Qv [ - 1] (2-2) a Although this procedure incorporates the direct measurement of choroidal arterial and venous hematocrit and 39 permits the control and indirect measurement of choroidal blood flow, errors in collecting these data might occur since the posterior choroidal artery supplies a portion of the choroid plexus, yet is not perfused in this preparation. Further, if Oa is measured instead of calculated, and all deep branches of the anterior choroidal artery (which supply brain areas other than the choroid plexus) are not electrocoagulated, some inflowing blood will be lost through these vessels, and Vf will be overestimated. These errors, however, are probably small. Less than 5% of the plexus by weight receives blood from the posterior choroidal artery, and less than 5% of the inflowing blood is lost from deep anterior choroidal arterial branches as determined volumetrically (Pollay et a1., 1972). Clearly the strength of this method is that it allows the separation of the role of the choroid plexus in the formation and modification of CSF from the complex solvent and solute exchanges that occur between CSF and brain tissue. 3. Collection of Nascent Fluid from the Surface of Exposed Choroidal Tissue The technique of collecting freshly formed CSF from cat choroid plexus developed by de Rougemont in 1960 has been previously discussed with regards to its use in determining the composition of nascent CSF. However, the direct sampling procedure has proven useful in estimating the rate of choroid plexus fluid formation (Vf) as well. 40 The determination of choroidal Vf is based on the measurement of the maximum rate at which newly secreted fluid can be collected with a standardized micropipette reproducibly positioned on the surface of the plexus. The pipettes are made of tubing of uniform bore and shape, and are held horizontally to avoid the development of‘a hydrostatic pressure as the fluid is collected; capillary action provides the force for fluid collection. This technique is most useful in predicting changes in choroidal Vf between control and test conditions since the area sampled represents only a small portion of the whole choroid plexus. 'These changes, however, may also reflect those in whole choroid plexus Vf. 4. Ventriculocisternal Perfusion Perfusion of the ventricular system to the cisterna magna was originally introduced as a means of recovering molecules of cerebral origin and as a method for studying the effects of ions and drugs on behavior (Cserr, 1971; Welch, 1975). Pappenheimer and his associates (1962) developed quantitative methods for investigating the rates of molecular exchange between the perfusion fluid, plasma and brain, and for estimating rates of bulk CSF production and reabsorption. This method involves the perfusion of an artificial CSF, similar in composition to normal CSF, into one or both lateral ventricles at a known and constant rate and with a known concentration of a tracer material, and the 41 quantitative collection of perfusate samples from another site, i.e., usually the cisterna magna. Alternatively, inflow or outflow cannulae may be placed in the cortical or lumbar subarachnoid spaces. Expressions for the steady state rates of bulk CSF production (Vf; nl/min) and absorption (Va; ul/min) were developed from the law of conservation of mass, assuming that all CSF formation occurs upstream from the outflow cannula. A reference compound of large molecular size, such that it is unlikely to diffuse from the system, is mixed with artificial CSF and perfused into the system at a known rate (Vi; ul/min) and concentration (ci; quantity/ul). Perfusate outflow rate (V0; ul/min) and concentration (co; quantity/DI) are the other measured variables. Given that tracer loss by diffusion is negligible, then the dilution of the tracer molecule that occurs during its transit through the CSF system can be used to calculate the steady state rate at which newly formed (tracer-free) CSF is added to the perfusate. Vf 8 Vi (ci - co) co (4—1) Steady state CSF absorption rate, ine., the rate at which fluid is absorbed in the subarachnoid space, can be estimated by the clearance of the non-diffusible tracer molecule. Va 8 Vi ci - Vo co (4_2) co 42 Finally, a mass fluid balance expression must be satisfied in order to demonstrate definitively the existence of a perfusion steady state during sampling periods. 6: + vi = 6. + 6. (4-3) As pointed out by Heisey et a1. (1983), the apprOpri- ateness and accuracy of these equations for estimating Vf and Va are rigidly dependent upon the stability of Vi, Vo, ci and co. 'This is because tracer molecules may not be uniformly distributed throughout all portions of the CSF system which are otherwise in communication. Martins et a1. (1977) suggested that the ventriculocisternal perfusion method possesses inherent potential sources of error linked directly with perfusion inhomogeneities which may exist. He prOposed that erroneous estimates of CSF formation rate may be a consequence of the mobilization of sequestrated and incompletely equilibrated CSF when craniospinal blood volume is suddenly changed by either an increased PaC02 or an increased central venous pressure. Ventriculocisternal perfusion equations (Eqs. 4-1 to 4-3) derived for the steady state analyses of CSF formation and absorption rate are not reliable indicators of these quantities during transient changes in the CSF fluid system (Heisey et a1., 1983; Fisher et a1., 1983). E | EEE l' ESE E l' E | CSF production is a two step process that includes a passive transudation of fluid and electrolytes from brain 43 capillaries and an active secretory process which involves the expenditure of energy (Wright, 1979). Consequently, factors that interfere either with the delivery of solutes and solvent to the secretory epithelium, or‘with local supplies of metabolic energy will affect the rate at which CSF is formed. ‘Although results of experiments are sometimes contradictory and mechanisms underlying changes in secretory rate are poorly understood, a consistent pattern emerges which relates various physical, chemical and biological factors to CSF formation. 1. Effects of Temperature Several investigations have been conducted on the effect of temperature on the secretion of CSF. Davson and Spaziani (1962) compared the rates of penetration of ethyl alcohol, ethyl thiourea, and Na24 from blood into brain and CSF in anesthetized rabbits at 37°C (normal body temperature) and at 25°C (hypothermia). The rate of penetration of Na24 into CSF is considered to be a measure of CSF turnover rate. The rates at which lipid-soluble ethyl alcohol and ethyl thiourea enter CSF, however, are largely affected by diffusion rates from the capillaries of the central nervous parenchyma. ‘Davson et a1. discovered that cooling the rabbit to 25°C significantly slowed the penetration of Na24 and ethyl thiourea, but not ethyl alcohol, into CSF and brain tissue. They postulated that the decreased Na24 and ethyl thiourea turnover rates are due to 44 a decrease in the primary secretion rate of CSF and/or to a reduction in choroidal capillary circulation. The small effects of cooling on the penetration of ethyl alcohol, they proposed, is due to the fact that ethyl alcohol is normally consumed quite rapidly by nervous tissue; cooling the animal markedly reduced the consumption and thereby obscured a fall in penetration rate that had, in fact, occurred. Another study which elucidated the effects of temperature on CSF formation rate was conducted by Fenstermacher and colleagues in 1969. They selectively cooled the brains of anesthetized cats to 15°C by perfusing cooled blood through the right carotid artery which was exposed and connected by polyethylene tubing to an extracorporeal system with a heat exchanger. CSF formation rate (Vf) was measured using ventriculocisternal perfusion methods during both normo- and hypothermic conditions. They found that CSF formation rate of the cat whose brain was made hypothermic was about one-sixth that of the cat with a normothermic brain. In fact, in some of their experiments, CSF formation appeared to cease at 15°C. Snodgrass and Lorenzo (1972) concurred with these reports. Like Fenstermacher et a1. (1969), they measured CSF formation rate in anesthetized cats using ventriculocisternal perfusion, but varied temperature between 31 and 41°C by heating the whole animal with a heating blanket or by allowing it to cool. Cats were divided into two groups: 1) an ”ascending group' in which 45 each steady state was at a higher temperature than its predecessor, and 2) a “descending group“ in which the first temperature was the highest. Within the temperature range studied, a change in 1°C in rectal temperature in either direction produced a matching 11% change in CSF formation rate. They concluded that the thermal effects on CSF formation were not due to injury, but rather resembled the relationship between rate of a chemical reaction and temperature. 2. Effects of Osmotic Gradients The effects of ventricular fluid osmolality on the bulk flow of nascent CSF into the cerebral ventricles has been investigated. Heisey et a1. (1962) used the ventriculocisternal perfusion method to determine CSF secretion rates in unanesthetized goats provided with chronically implanted ventricular and cisternal cannulae. The osmolarity of the perfusion fluid was varied by altering sodium chloride concentrations while the concentrations of all other ions were maintained constant. Osmolality was determined by freezing point or by calculation from sodium analyses. ‘When the perfusion fluid was hypotonic to plasma, CSF production decreased. Hypertonic perfusion fluid caused CSF synthesis to increase. They concluded that since CSF secretion continued despite an adverse gradient (chemical potential of water greater in CSF than in blood), CSF formation must be dependent upon a mechanism other than 46 osmotic flow secondary to the ependymal cell secretion of a hypertonic sodium chloride solution. They also concluded that the ventricular ependyma must significantly contribute to CSF secretion, at least when the CSF was hypertonic to blood, since abnormally high rates of formation were observed; if the choroid plexus were the sole source of CSF, choroidal blood flow would have to be on the order of 3 ml/min per gram tissue if 20% of the plasma flow were secreted as CSF. In a later study it was determined that choroid plexus blood flow in anesthetized sheep is, in fact, 6 ml/min per gram tissue (Page et a1., 1980). In a later study, Wald et a1. (1976) determined not only the effects of osmolality gradients between CSF and blood on CSF production rate, but also attempted to differentiate between the sources of the nascent fluid during the osmotic tests. 'They measured the bulk flow rate of newly formed CSF in anesthetized cats during ventriculocisternal perfusion with solutions containing sucrose, urea, and sodium chloride in different concentrations. Subsequently, they inhibited the active secretion of the choroid plexus with acetazolamide while osmotic gradients were induced using urea and sodium chloride solutions and again measured the bulk flow rate of nascent CSF. The difference between the normal CSF formation.(measured while perfusing with mock CSF) and that measured during perfusion with anisosmotic fluids was linearly related to corresponding differences in osmolality _ 47 of the effluent from the ventricles. The slope of this relationship is the coefficient of osmotic flow. 'The coefficient for sodium chloride and sucrose were similar, and these molecules were considered to have equal reflection coefficients in the CSF compartment. However, a larger osmotic gradient was required to inhibit bulk flow completely when the perfusion flow contained sucrose instead of sodium chloride, and perfusion with isotonic sucrose solution rather than isotonic sodium chloride increased bulk flow of nascent fluid. They explained this difference between sucrose, a nonelectrolyte, and sodium chloride, an electrolyte normally found in nascent CSF, to the fact that when a sodium-free solution is perfused through the‘ ventricles (i.e., a sucrose solution), a sodium gradient between the perfusion fluid and the blood and brain is established. The sodium gradient results in a net sodium influx from the blood into the perfusion fluid within the ventricles, which by coupling between solute and solvent, increases bulk flow rate. When perfusing with urea and sodium chloride, the coefficient of osmotic flow and the reflection coefficient for urea in the CSF system were smaller than for sodium chloride, while the y-intercepts (idh, the osmotic gradient required to inhibit nascent fluid bulk flow completely) for both urea and sodium chloride perfusions were not significantly different. The bulk flow rates of nascent CSF with either of the test solutions were not different from 48 the nascent fluid bulk flow rate when mock CSF was perfused. After intravenous administration of acetazolamide, however,. both the slopes and the y-intercepts for the two isotonic test solutions were equal. Wald and co-workers interpreted these data to indicate that acetazolamide inhibits a source of fluid which is responsive to differences in the reflection coefficient, while another mechanism, related solely to osmotic gradient, still remains active. Since CSF production is an active process at the choroid plexus, they interpreted their data to indicate that the choroid plexuses contriubte significantly to the increase in CSF production observed during hypertonic perfusions. More recently, Nakamura and Hochwald (1983) using anesthetized cats, measured choroidal and regional cerebral blood flow with microspheres during ventriculocisternal perfusion with anisosmotic mannitol solutions. They confirmed that perfusion of the cerebral ventricles with anisosmotic solutions alters CSF formation. When the perfusion fluid was 720 mOsm/liter, CSF formation rate increased by 174%, but during perfusion with a 60 mOsm/liter solution of mannitol, there was an 80% reduction. How CSF volume flow rate increases or decreases during anisosmotic perfusions remains unknown since neither choroidal nor regional cerebral blood flow changed in response to the induced changes in perfusion fluid osmolarity. 49 3. Effects of Blood Flow Blood flow to choroidal or extrachoroidal sites of CSF production may be influential in regulating CSF formation rate (Heisey et a1., 1962; Ames et a1., 1965; Carey and Vela, 1974; Haywood and Vogh, 1979; Martins et a1., 1976; Davson and Spaziani, 1962; Snodgrass and Lorenzo, 1972; Lindvall et a1., 1979; Deane and Segal, 1978). It has been suggested that blood flow places an upper limit on the rate of CSF formation because as fluid is extracted from CSF secretory cells, the concentration of plasma proteins increase. An increased plasma protein concentraton will increase colloid osmotic pressure in the capillaries and slow filtration (Cserr, 1971). Carey and Vela (1974) provided support for the idea that blood flow to CSF secretory sites alters CSF production when they found significant reductions in formation rate in anesthetized dogs whose systemic blood pressures were lowered to 60 Torr by exsanguination. 4Although neither choroidal nor cerebral blood flow was measured in these experiments, Carey and Vela speculated that the decrease in Vf with decreased blood pressure was due to a diminished blood flow to CSF secretory sites. Weiss and Wertman (1978) found results similar to Carey and Vela when they measured the rate of CSF production in anesthetized cats while cerebral perfusion pressure (CPP; CPP = mean arterial blood pressure - CSF pressure) was decreased either by decreasing systemic arterial blood 50 pressure, or by raising CSF pressure. Only when CPP was subsequently reduced (50 mmHg) was there an inverse relationship between CSF formation rate and CSF pressure. They suggested that variations in CSF formation rate pertain to the maintenance of constant blood flow through the choroid plexus under conditions of normal CPP (i.e., when cerebral blood flow autoregulation is functional). When the capacity for compensation by cerebral autoregulation is exceeded (when CPP falls below 55 mmHg; Drake, 1977, cited in Weiss and Wertman, 1978), adequate choroidal perfusion is not.maintained and ultrafiltration at the level of the choroidal capillary is decreased which, in turn, decreases CSF production. In 1980, Sklar and co-workers corroborated that CSF formation rate varies with CSF pressure below limits of cerebral blood flow autoregulation. They found that CSF dynamics were altered when cerebral perfusion pressure (CPP) transiently fell below the lower limits of autoregulation (spontaneous hypotensive bouts), and CSF formation rate decreased with increasing CSF pressure. Sklar hypothesized that because intact autoregulation maintains constant cerebral blood flow and perhaps choroidal blood flow, CSF formation may normally be independent of CPP. With a disturbance of blood flow autoregulation, CSF formation may follow cerebral blood flow passively with changes in CPP. Heisey et a1. (1983) found a direct relationship between CSF formation rate and CPP over the range of 70 to 51 105 Torr, and also suggested that cerebral blood flow autoregulatory mechanisms may have been disturbed in their anesthetized cats, possibly by anesthetization. They suggested that both changes in CPP and regional brain blood flow would predictably affect CSF formation rate by altering the rate of plasma ultrafiltration and the amount of fluid and solutes presented to the epithelial cells of the choroid plexuses as predicted by the standing gradient hypothesis (Diamond and Bossert, 1967). Pollay (1975) was the first to relate the standing gradient hypothesis of solute and water transport (Diamond and Bossert, 1967) to the relationship between choroidal Vf and cerebral blood flow. This theory predicts that Vf should be regulated both by the metabolically driven transport systems found at the apical and basal aspects of the choroidal epithelium, and by the blood flow supplying these secretory cells with ultrafiltered fluid. If cerebral blood flow is restricted, as it would be were systemic blood pressure below the level of autoregulation (Carey and Vela, 1974; Weiss and Wertman, 1978), then transudation of plasma ultrafiltrate would be decreased. 'This could limit the delivery of available solute and water to the choroidal transport systems and result in a decreased rate of CSF formation. Even with increased cortical and choridal blood flow (as found with an increased PaCOz (Page et a1., 1980; Ames et a1., 1965; Martins et a1., 1976)), and increased transudation of solutes and water from the choroidal 52 capillaries, Vf may not necessarily increase, because with adequate blood flow, the limiting process in the formation of CSF may be metabolic (Sklar et a1., 1980). Unless metabolism is changed during increased transudation, Vf may remain stable despite an increase in choroidal blood flow. Bering (1959) investigated the relationships between cerebral blood flow and the rate of CSFformation by simul- taneously measuring in the same animal global cerebral blood flow and cerebral oxygen consumption using the nitrous-oxide Kety-Schmidt method, and CSF formation rate (Vf) using timed-collections of CSF drainage. He found that Vf (ml/min per 100 mg brain weight) was linearly correlated with cere- bral oxygen consumption (ml/min per 100 gm brain weight), and Vf (ml/min per 100 mg choroid plexus tissue) was correlated with cerebral blood flow'(ml/min per 100 gm brain weight), and concluded that while both cerebral metabolism and cerebral blood flow are important influences on the net rate of CSF formation, there are differences between these processes. Cerebral oxygen consumption is a primary regula- tory mechanism for CSF production from brain ependyma, whereas cerebral blood flow is preeminent in determining the rate of CSF synthesis from choroidal sources. Cerebral blood flow and CSF formation rate were measured later by Nakamura and Hochwald (1983). They used microspheres to measure blood flow in cerebral cortex and choroid plexus of anesthetized cats during normocapnia, and 53 after Paco2 was increased by 300% or decreased by 50%. Blood flow measurements were also made during perfusion of the ventricular system with mock CSF. They found that blood flow in the choroid plexus, unlike that in the cortex, decreased by almost 50% during hypercapnia while CSF forma- tion rate remained constant. Both choroid plexus and corti- cal blood flow decreased and CSF formation rate stayed constant during hypocapnia. Nakamura and Hochwald did not explain the decreased choroidal blood flow during hypercapnia, but concluded that choroid plexus blood flow does not limit or even affect the volume flow rate of CSF from the choroid plexus. 4. Effects of Hydrostatic Pressure Gradients Heisey et a1. (1962) studied the influence of acute changes of hydrostatic pressure on the rate of ventricular fluid formation in goats during ventriculocisternal perfusion. CSF pressure was varied over a range of -10 to +30 cmHZO by lowering and raising the height of the cisternal outflow cannula. They found that although CSF steady state absorption rate varies linearly with CSF pressure, steady state CSF formation rate is independent of acute changes in CSF pressure over the range tested. Hochwald and‘Wallenstein (1967), and Sklar et a1. (1980) found similar results from experiments conducted on anesthetized cats and dogs. 54 Bering and Sato (1963) investigated the effects of chronic increases in CSF pressure on CSF formation and absorption rates by inducing hydrocephalus in anesthetized dogs and then measuring these rates using either a lateral ventricle-lateral ventricle or lateral ventricle-fourth ventricle perfusion systems Effects of acute increases in CSF pressure were also studied in normal dogs using these perfusion systems, as well as those for ventriculocisternal, subarachnoid-cisternal, and ventriculo-aqueductal perfusion. The formation of CSF was found to be constant, independent of intracranial pressure, and unaffected by the development of hydrocephalus. (Absorption, however, was linearly and directly related to the hydrostatic pressure in the ventricular and subarachnoid spaces for both the normal and hydrocephalic dog. Decreases in CSF formation rate with increasing CSF pressure (20 to 25 cmHZO) however, have been reported for anesthetized normal rabbits and cats (Hochwald and Sahar, 1971) and anesthetized cats made hydrocephalic by cisternal instillation of kaolin (Eisenberg et a1., 1974). Sklar and his associates (1980) suggested that CSF formation rate is influenced by CSF pressure through impaired cerebral blood flow autoregulation. 5. Effects of C02 Tension Hypercapnia induces cerebral vasodilation with an increase in cerebral blood flow and volume (Reivich, 1964; 55 Rich et a1., 1953; Adams et a1., 1980; Grubb et a1., 1974) and hypocapnia, a vasoconstriction with a corresponding decline in cerebral blood flow (Reivich, 1964; Rich et a1., 1953). In 1954, Tschirgi and colleagues administered a 70% 02-. 30% C02 gas mixture to anesthetized cats and rabbits before and after the administration of acetazolamide, a carbonic anhydrase inhibitor. CSF formation rate was measured by timed-collections of CSF draining from a cisternal cannula. C02 inhalation caused only a transient increase in the rate of CSF outflow which they attributed to a rapid change in the volume of the intracranial vascular compartment due to the increased C02 tension. The inhalation of 30% C02 following acetazolamide (Diamox) administration, which alone caused CSF flow to decline over a period of an hour to a new steady rate approximately one-third its previous value, still produced a transient, but smaller, rise in CSF flow. CSF flow rate gradually returned to its prehypercapnic level. Tschirgi concluded that large changes in inspired C02 do not effect steady state changes in CSF flow and that acetazolamide must inhibit CSF flow through some mechanism other than through an increased 002 tension of intracranial fluids. Others (Oppelt et a1., 1963; Hochwald et a1., 1973; Martins et a1., 1976; Fisher et a1., 1983) conclude similar- ly to Tschirgi that although the inhalation of C02 produces transient alterations in CSF flow rate, it does not produce consistent changes in steady state CSF formation rate. 56 Conversely, Ames et a1. (1965) reported that both the inhalation of 10% C02 and changing the PC02 in the Pantopaque medium covering the exposed lateral ventricular choroid plexus caused a rise in the maximum rate at which choroidal fluid formed on the surface of the plexus. They observed.the choroidal vasodilation during hypercapnia, and suggested that an increase in blood flow may have been primarily responsible for the changes in fluid production. In ventriculocisternal perfusion studies (Oppelt et a1., 1963; Oppelt et a1., 1964; Hochwald et a1., 1973; Davson and Segal, 1970), including those in which average cerebral blood flow changed with increased Paco2 (Martins et a1., 1976; Nakamura and Hochwald, 1982), 6: was found to be unaffected by increases in PC02. This apparent contradiction may be partly explained on the basis of methodological differences between Ames' in Sitll choroid plexus preparation and the ventriculocisternal perfusion method. First, the former experimental technique requires more extensive surgery than does the ventriculocisternal pefusion technique, and may damage the blood-brain barrier and cerebral autoregulation. Secondly, data from in sitn studies report only choroidal production of CSF, whereas data from ventriculocisternal perfusions report net changes in Vf independently of their source, which may be more realistic, since there are two sources of Vf; the choroid plexuses and the brain ependyma (Hochwald and Wallenstein, 1967; Bering and Sato, 1963; Welch, 1963; Pollay and Curl, 57 1967; Pollay, 1975; Bering, 1959; Curl and Pollay, 1968; Davson and Luck, 1957; Davson and Segal, 1970). 6. Effects of Carbonic Anydrase Inhibitors Common to the kidney, stomach, pancreas, and other exocrine structures, the choroid plexus has a high concentration of carbonic anhydrase (Maren, 1972; Vogh, 1980) which is also found in brain glial cells (Vogh, 19801. This enzyme catalyzes the formation of bicarbonate from gaseous C02, and is intimately involved with the movement of fluid across epithelial tissues. In fact, the first substance found to affect materially the rate of CSF formation was acetazolamide, which inhibits carbonic anhydrase. Carbonic anhydrase inhibitors markedly depress CSF formation rate (Ames et a1., m 1965; Oppelt et a1., 1963; Pollay, 1975; Cserr, 1971; Welch, 1975; Welch, 1963; Oppelt et a1., 1964; Knopp et a1., 1957; Davson and Luck, 1957; Davson and Segal, 1970; Maren, 1972; Pollay et a1., 1972; Curl and Pollay, 1968), however, the mechanism is unknown. Tschirgi and co-workers (1954) administered acetazolamide (150 mg/kg;iv) to anesthetized cats and rabbits, and compared the rate at which CSF drained freely from a cisternal cannula before and after the drug treatment. They observed a 30% decrease in CSF pressure in the intact ventricular system, (which was later confirmed by Knopp et a1. (1957)), and a sustained 3- to lS-fold (70%) decrease in the CSF drainage rate after acetazolamide 58 treatment. They proposed that acetazolamide inhibits CSF synthesis by decreasing the rate of carbonic acid production from metabolically produced C02. They postulated also that choroidal and extrachoroidal CSF production is dependent upon an ion exchange between plasma and CSF. Transport processes located on choroidal epithelium and throughout brain glial processes exchange hydrogen and bicarbonate derived from the hydration of C02 for sodium and chloride. One mole of metabolically produced C02 is exchanged for one mole each of sodium and chloride which is transported from plasma to CSF. To maintain osmotic equilibrium, it was proposed that water passively follows transported sodium and chloride into CSF. Carbonic anhydrase, then decreases CSF formation rate by decreasing the amount of hydrogen and bicarbonate available for exchange. Davson and Luck (1957) refuted Tschirgiis proposal that acetazolamide inhibits extrachoroidal CSF formation when they found that Na24 turnover (an index of CSF production) decreases in CSF but not in brain after intravenous adminis- tration of acetazolamide. IDavson and Luck agreed that acetazolamide may be interfering with ion-exchange at the choroid plexus, an idea supported by Pollay et a1. (1972) who demonstrated using the isolated perfused choroid plexus that the addition of acetazolamide to the arterial blood (20 pg/ml) or to CSF (10"3 M) stopped choroidal CSF secretion. Pollay and his co-workers also measured choroidal blood flow and found that it was not affected by acetazolamide. 59 If an ion-exchange mechanism were operating at the choroid plexus, reasoned Welch (1963), the content of C02 in choroidal venous blood would have to be enormous. At rates of choroidal CSF formation measured by him, approximately 24 mEq of sodium ions are secreted, an equivalent amount of C02 would be expected to be recovered in choroidal venous blood if the ion-exchange mechanism were operating. He measured the C02 content of venous blood from the plexus and found it insufficient to account for the sodium chloride flux into CSF. A more recent view of the role of carbonic anhydrase in CSF secretion, and the mechanism whereby acetazolamide inhibits CSF formation was presented by Maren in 1972. He discovered that intravenous acetazolamide inhibited the transependymal flux of bicarbonate to the CSF similarly to that of sodium and water. He proposed that plexus secretory cells, through the protolysis of water, produce hydroxyl ions which react with C02, with or perhaps without carbonic anhydrase. Bicarbonate and hydrogen ions are formed as a result of this reaction. The proton would subsequently enter the bloodstream and be buffered while a bicarbonate gradient was established. Sodium would then serve as a counter ion for transport with bicarbonate into the CSF system. Maren proposed, therefore, that carbonic anhydrase inhibitors act by interrupting the necessary coupling of bicarbonate and sodium which, in turn, prevents the osmotically obligated movement of water into the CSF and CSF 60 formation declines. Further, he hypothesized that since sodium and bicarbonate are intimately coupled, anything that would interfere with carbonic anhydrase or ATPase would produce similar changes in sodium and water transport. 7. Effects of Cardiac Glycosides The choroid plexuses have transport systems for inorganic ions as well as for organic acids and bases (Wright, 1978). Since ouabain can inhibit these processes, they have been implicated to be dependent on Na/K-ATPase (Lindvall et a1., 1982; Vates et a1., 1964). Likewise, CSF formation is dependent upon this enyzme system since ouabain produces sharp declines in CSF production when administered intraventricularly (Cserr, 1971; Vates et a1., 1964), intrathecally (Oppelt et a1., 1964), or when applied tepically to the choroid plexus (Welch, 1963). In fact, if ouabain is added to a medium which bathes the plexus, the elaboration of CSF is almost completely inhibited (Welch, 1963). Unfortunately, ouabain is not a useful in 1119, research tool because, unlike acetazolamide which is well tolerated at doses required to achieve maximal effects, it is extremely toxic producing both marked circulatory and respiratory disturbances. The mode of action of cardiac glycosides on CSF formation rate is better defined than that of carbonic anhydrase inhibitors (Pollay, 1975; Cserr, 1971; Ames et a1., 1965; Curl and Pollay, 1968; Davson and Segal, 1970). 61 Ouabain decreases CSF formation by inhibiting the Na/K- ATPase enzyme system (Vates et a1., 1964). The Na/K pumps localized on the apical surface of choroidal epithelial cells (Wright, 1978) and found throughout brain tissue are fueled by the hydrolysis of intracellular ATP by the enzyme Na/K-ATPase. Inhibiting the pump prevents the active secretion of sodium into the ventricle (Vates, 1964), which is considered by Wright to be the primary process in CSF secretion. 8. Effects of the Autonomic Nervous System Possible neural influences on the production rate of CSF has been suggested by histological studies since 1874, when Benedikt demonstrated the presence of nerves in the choroid plexuses. It is now well known that both sympathetic and cholinergic nerve terminals within the plexus innervate both the resistance vessels and the secretory epithelium (Edvinsson et a1., 1973; Edvinsson et a1., 1974; Edvinsson et al., 1975a; Edvinsson et al., 1975b; Lindvall et a1., 1982). In 1971, Edvinsson and co-workers helped to establish that the choroid plexus innervation was indeed functional. They measured the activity of carbonic anhydrase from rabbit choroid plexus homogenates before and after cervical sympathectomy by incubating the plexus homogenates in phosphate buffer containing NaH14CO3 and the 14C02 formed was collected in Protosol-soaked filter paper strips. The radioactivity of the strips was measured by 62 liquid scintillation spectrometry. They found that at one and three days post-sympathectomy, carbonic anhydrase activity had approximately doubled; at 14 days it was reduced by 60% compared with unoperated controls. Subsequently, Edvinsson and colleagues (1975) were able to confirm their previous observations in homogenates of blood-free rabbit choroid plexus. They found that carbonic anhydrase activity was higher in the choroid plexus from the fourth ventricle than from the lateral ventricles, and that depletion of norepinephrine from the adrenergic nerves by reserpine treatment increased carbonic anhydrase activity. Two weeks after sympathectomy, however, the carbonic anhydrase activity had returned to normal which they attributed to the development of denervation supersensi- tivity of the adrenergic receptors to circulating catechola- mines. They concluded that the sympathetic nervous system inhibits carbonic anhydrase activity and CSF production. Lindvall et al. (1978b) directly measured the influence of sympathetic nerves on the rate of CSF production. One group of anesthetized rabbits underwent ventriculocisternal perfusion one week after superior cervical ganglionectomy. Another group of animals had their cervical sympathetic trunks electrically stimulated during ventriculocisternal perfusion. The first group of animals showed a 33% increase in the rate of bulk CSF production compared to the second group before electrical stimulation (control). In the intact group, sympathetic activation produced a reduction in 63 the rate of CSF formation by 32%, but systemic blood pressure effects were negligible. Based on the results from this study and from the carbonic anhydrase study on blood- free choroidal tissue (Edvinsson et a1., 1975), they concluded that the sympathetic nerves in the choroid plexus inhibit CSF production primarily by their effects on the plexus epithelium and its secretory functions. Little was known in 1978 about the nature of the local adrenoceptors mediating the sympathetic-nerve induced effect on CSF formation. Lindvall and associates (1979) pharmacologically characterized these receptors by infusing sympathomimetic agonists and corresponding specific receptor antagonists via the intraventricular or vascular routes during ventriculocisternal perfusion of anesthetized rabbits. They found that intraventricular or intravenous administration of norepinephrine produced a dose-related decrease in CSF production up to 50% which could be counteracted by both phentolamine (an alpha-antagonist) and intravenously» bur.not intraventricularly administered prOpranolol.(a beta antagonistL. Systemic arterial blood pressure remained stable during intraventricular infusions of norepinephrine, but transiently increased with intravenous infusions of norepinephrine. Further, intraventricular administration of a betal-receptor agonist, HBO/62, reduced CSF formation; practolol (a betal- antagonist) blocked this effect. Conversely, terbutaline, a betaz-agonist had little effect on CSF synthesis. They 64 suggested that sympathomimetic reduction of CSF production is a complex function of both betal-receptor-mediated inhibition of plexus epithelium secretion as well as a reduced choroidal blood flow resulting from stimulation of the vascular alpha-receptors. Biochemical evidence supporting the suppostion that beta-receptors exist on choroidal epithelium was provided by Nathanson in 1979. Binding of beta-adrenergic agonists to such receptors leads to the intracellular synthesis of 3', 5'-mon0phosphate (cyclic AMP) from adenosine triphosphate (ATP) (Edvinsson and MacKenzie, 1977), a reaction catalyzed by adenylate cyclase. Nathanson investigated the possible existence of beta-adrenoceptors by enzymatically determining the synthesis of cyclic AMP from ATP in broken cell prepara- tions of cat, rabbit, calf, and dog choroid plexus. Cell preparations were exposed to various concentrations of biogenic amines and cyclic AMP activity was measured. He discovered beta-adrenergic-sensitive adenylate cyclase in broken cell preparations of both whole choroid plexus and choroidal epithelium which had been dissociated from its subjacent vasculature. A substantial portion of the enzyme was present in the epithelial cell fractions. Further, he showed that l-isoproteronol-induced stimulation of adenylate cyclase was greater in epithelial cell fractions than in choroidal vessel fractions, and that the cat fourth ventricular choroid plexus, which contains relatively'more secretory epithelium than the lateral ventricular plexus, 65 shows a twofold greater response in adenylate cyclase to isoproteronol. L-norepinephrine, a mixed beta— and alpha- adrenergic agonist, stimulated enzyme activity nearly as much as isOproteronol (310% of control levels) but was less potent (Ka - 3 x 10"5 M); l-phenylephrine, an alpha- adrenergic agonist, was both less potent (Ra = l x 10" M) and less effective (185% of control levels). Likewise, alpha-adrenergic mediated vasomotor reactivity in the choroid plexus was confirmed from in 11:19, studies on anterior choroidal artery strips. Edvinsson and Lindvall (1978) dissected choroidal arteries from freshly slaughtered cows, and mounted two pieces of the arteries as cylinders in an organ bath containing a CSF buffer solution. Circular vasomotor activity was registered isometrically and reflected by the voltage output from force-displacement transducers. Measurements were made after various vasoactive pharmacological agents were added to the organ bath. They found no spontaneous contractions of the arterial strips, but found weak contractile responses to epinephrine, norepinephrine, phenylephrine, and isOproterenol in decreasing order of potency; Phentolamine (alpha-antagonist) inhibited these responses. Other agents such as histamine, prostaglandin F2“, and saturated potassium chloride produced strong contractions. Dilator responses could also be elicited, provided the choroidal arteries were constricted with prostaglandin F2“ and contractile effects blocked by phenoxybenzamine (an alpha- 66 antagonist). The relative potency for dilator sympathomimetic agents were, in decreasing order, isoproterenol, norepinephrine, epinephrine, terbutaline. The dilator effect was inhibited by propranolol (a beta- antagonist) which indicated the dilator beta-adrenergic receptors coexist with alpha-adrenergic receptors on choroidal arteries. Cholinergic effects on CSF formation rate have also been directly studied and appear to have an influence. Lindvall et a1, (1978a) measured CSF production in anesthetized rabbits by ventriculocisternal perfusion. Intraventricular or intravenous infusion of‘carbamylcholine (a parasympathomimetic) or acetylcholine produced a dose- related reduction in CSF formation by as much as 55%. AtrOpine (a muscarinic cholinergic antagonist) antagonized this response while hexamethonium (a nicotinic cholinergic blocking agent) did not. Intraventricular carbachol (para- sympathomimetic) potentiated the reduction in CSF secretion induced by electrical activation of the cervical sympathetic nerve trunks. They concluded that muscarinic cholinergic mechanisms markedly reduce CSF production, though the site of the parasympathomimetic agents is unknown. Since CSF responses were observed after the agents were administered intraventricularly, Lindvall proposed that the effects were mediated through the secretory epithelium rather than through the vascular system. 67 W Although CSF may be absorbed by the brain (Bering and Sato, 1963; Rosenberg et a1., 1978; Rosenberg et a1., 1982) and by cranial and spinal nerve roots (Pollay and Welch, 1962; Hammerstand et a1., 1969; Davson et a1., 1970; Marmarou et a1., 1975), its main absorption into venous blood is normally'through the arachnoid villi.(Welch, 1975) which project into dural venous sinuses. Welch and Friedman (1960) and Pollay and Welch (1962) investigated the function of the arachnoid villi in monkeys and dogs. They mounted a disc of dura mater between two chambers as an occluding membrane. The two chambers were then filled with saline, and one chamber was attached to a reservoir containing Ringer lactate solution. The other chamber was attached to a small polyethylene tube, and flow was measured by monitoring bubble movement. Flow through the arachnoid villi, seen when the pressure difference across the disc was approximately 30 mm of perfusion fluid, was only from the subarachnoid to the sinus side. ‘Above 30 mm, the critical opening pressure, flow increased non- linearly with increasing pressure. (At pressures just higher than the critical opening pressure, the lepe of the pressure-flow curve was steeper than that for high pres- sures, presumably due to drainage of CSF through valved elastic tubes. Others have confirmed that CSF absorption through the arachnoid villi is a pressure-dependent phenomenon in many 68 mammals (Welch and Friedman, 1960; Pollay and Welch, 1962; Heisey et a1., 1962; Bering and Sato, 1963; Davson et a1., 1970; Hochwald and Sahar, 1971; Marmarou et a1., 1978), and that colloid osmotic pressure differences between CSF and venous blood in the dural sinuses are inconsequential to the rate of CSF absorption through the villi.(Pollay and Welch, 1962; Davson et a1., 1970; Pollay and Davson, 1963). The ‘villi seem to allow an unrestricted passage ofjproteins through their pores, as corroborated by Pollay and Welch (1962) who showed that 8 n.diameter polystyrene spheres and canine erythrocytes pass freely through canine arachnoid villir Also, Mann and co-workers (1979) demonstrated in the rat that.0.5 n.polystyrene beads and inulin flow through arachnoid villi. The arachnoid villi are a major efflux route for CSF and for macromolecular and particulate substances. W The relationship between the volume of CSF and its pressure has been investigated clinically and in the laboratory for many years. Weed et a1. (1932) studied CSF pressure changes produced by changes in body position and determined that when an animal was tilted from a horizontal to a vertical (head-down or tail-down) position, the volume of CSF displaced from the CSF system through an Open-end manometer was linearly related to CSF pressure. They sug- gested that the central nervous system enclosures behaved like an elastic system which permits relocation of 69 craniospinal fluid. .Although cranial dura mater is less elastic because of its close apposition to the skull, the spinal dura and the compressible, distensible spinal and cranial venous beds permit CSF movement. weed and his colleagues later investigated whether the cranial venous system is an elastic component of the central nervous system. lThey measured subarachnoid and saggital sinus pressures in etherized dogs during venous compression and during cranial subarachnoid fluid volume expansion due to cisternal infusions of CSF. Any intracranial venous pressure change affected CSF pressure, while saggital sinus venous pressure was uninfluenced by changes in subarachnoid pressure. They proposed that the pressure-volume relationship between cerebral veins and CSF is that of two fluids of unequal volume but normally'of equal pressure which are separated by an elastic membrane. The two fluids have different pressures because the cerebral blood volume is not confined to the cranium, but CSF volume is confined in rigid craniospinal membranes, which it can leave only slowly. They concluded that abrupt changes in CSF volume are permitted mainly by a concomitant compression of the cerebral venous bed. Ryder et a1. (1953) provided a similar assessment of the volume-pressure relationshiplbetween the vascular and CSF volumes based on CSF pressure recordings taken from patients with different intracranial disorders. They observed in all patients that lateral ventricular or lumbar 70 pressures were stable. Subarachnoid volume additions or withdrawals of CSF did not affect CSF pressure. When CSF pressure was recorded during the withdrawal or addition of several milliliters of fluid from or to the CSF space, return of the displaced pressure to its pre-disturbance level occured within minutes. ‘When large'volumes of fluid were added or withdrawn, it was necessary to withdraw or add fluid to the CSF space in order to re-establish pressure within minutes. When fluid was added to the CSF space, however, less was needed to be withdrawn in order to re- establish CSF pressure. Likewise, after volume withdrawals, a lesser CSF was necessary to restore pressure. They con- cluded that if intracranial volume is constant at a given pressure, then a change in the volume of the cerebrospinal fluid space induces an immediate reciprocal change in the craniospinal vascular bed, as well as a concomitant net seepage of CSF into or out of the CSF system. In 1972, Martins and co-workers characterized the intracranial pressure-volume relationship slightly differently. Using human volunteers, they observed the movement of a PantOpaque column in the spinal subarachnoid space during myelography. Each of the volunteers performed, in turn, a‘Valsalva.maneuver, hyperventilated for two minutes, breathed 10% C02 in oxygen for two minutes, and then repeated the Valsalva maneuver. Finally, the volunteers underwent fifteen seconds of jugular compression. During both the forced expiration with glottis closed 71 (Valsalva maneuver) and during hyperventilation, the spinal dural sac partially collapsed, as indicated by the Pantopaque column narrowing and elongating. Conversely, during 10% C02 inhalation and during bilateral jugular compression, the spinal sac distended, as indicated by the column widening as it moved caudally. to fill the lumbosacral nerve roots. They concluded that the spinal dural sac serves as a dynamic reservoir that can change its capacity in response to prevailing transmural pressure gradients. Since the cranial and spinal CSF spaces freely communicate through the foramen magnum, CSF could rapidly move between these two compartments, and volume-induced sudden changes in CSF pressure could be effectively- buffered. They concluded also that the spinal dural sac together with its loose fatty tissue and epidural venous plexus provide the major source of elasticity which underlies the apparent reciprocal relationship between cranial blood and CSF volumes. Marmarou et a1. (1975), however, disagree with Martinfis hypothesis that the majority of central nervous system elasticity is in the spinal axis. They conducted a series of experiments on anesthetized cats designed to apportion the total central nervous system elasticity into cranial and spinal components. Cerebral and spinal CSF components were isolated by an inflated balloon positioned epidurally'at the C-6 spinal level. CSF volume additions were made, in turn, into the two compartments, while the change in CSF volume in 72 respect to CSF pressure (dV/dP) as well as the change in CSF volume per time (dV/dt) were measured. These ratios were used to define CSF compliance and absorption rate, respectively. They found that 68% of total compliance is attributable to the cerebral compartment, and 32% to the spinal axis. They also found that 84% of CSF absorption is in the cranial compartment. They showed that central nervous system elasticity is due to the compressibility, and distensibility of cranial venous vessels, but recognized that the spinal axis provides an additional buffering system against sudden volume-induced changes in CSF pressure. Cranial, and to a lesser extent spinal, CSF absorptive mechanisms were more important in buffering “steady-state" pressure disturbances, such as those associated with expanding extra- or intracerebral masses. MATERIALS AND METHODS ANESIHESIA Adult cats (Fells domestigns) of either sex, weighing 2.5-4.5 kg were anesthetized either with sodium pentobarbital (25 mg/kg iv or 36 mg/kg ip; Nembutal, Abbott Laboratories, Chicago, IL) or with Dial-urethane (0.6 ml/kg ip; Appendix B). .Anesthetic depth was assessed periodically by evaluating jaw tension, palpebral, corneal and nociceptor reflexes, and supplemental iv doses of sodium pentobarbital were given as needed to maintain a stable anesthetic plane. SUBQICAL_AND_EXEEBIMENIAL_RRQQEDQBE Systemic arterial blood pressure was monitored from a cannula (PE-60; Clay-Adams, Parsippany, NY) introduced into the aorta via the femoral artery, which was attached to a pressure transducer (Statham model P23Dc; Grass Instruments, Quincy, MA) in series with a low level DC preamplifier (model 5P1; Grass Instruments, Quincy, MA), and displayed on a oscillograph (model 5DWC1; Grass Instruments, Quincy, MA). Arterial blood samples were periodically collected anaerobi- cally for pH and PC02 determinations (Appendix C) using either Radiometer thermostatted electrodes (model PHA 927b; Radiometer A/S Copenhagen, Denmark) or a Corning blood gas analyzer (model 165/2; Corning Medical Instruments, 73 74 Medfield, MA). .A venous cannula (PE 60) was introduced into the inferior vena cava via the femoral vein for injecting supplemental doses of anesthetic and for test drug infu- sions. Body temperature was monitored with a rectal therm- istor probe attached to a Wheatstone bridge circuit and meter (model 400; Yellow Springs Instrument Co., Yellow Springs, OH) and maintained by a manually adjusted heating pad on which the animal rested. The trachea of each animal was intubated with an inflatable cuffed cannula (4.0-5.0 mm ID.; American Hospital Supply, McGraw Park, IL) which was affixed to a low resist- ance, low deadspace respiratory valve assembly, through which the cat breathed spontaneously. The expiration port of the respiratory'valve was attached to a Fleisch pneumota- chograph (model 7318; Dynasciences Medical Products, Blue Bell, PA) and a low level DC preamplifier (model 5P1; Grass Instruments, Quincy, MA) for monitoring respiratory frequency and tidal volume (Appendix D). End tidal C02 (FACOZ) was monitored from a needle cannula in the tracheal cannula which led to an infra-red C02 gas analyzer (LB-2 model; Beckman Instruments, Inc., Anaheim, CA; Appendix E). By means of amplified signals from the pneumotachograph and C02 analyzer, tidal volume and FACOZ, respectively. were displayed on a calibrated strip chart recorder (model 7100 IBM; H-P Moseley, Pasadena, CA). Mean steady state tidal volume (VT) and minute ventilation (VB) were calculated as described in Appendix D. 75 BBAIN_yENIB1QflLAB_AND_£ISIBBHAL_EHNQIHRES A midline skin incision was made from the skull's interparietal bone to the spinous process of the second cervical vertebra. Underlying neck muscles and fascia were cleared by blunt dissection to expose the atlanto-occipital membrane. A cisternal needle (21 ga.;:metal hubless needle; Vita Needle Co., Needham, MA) secured in a micromanipulator (model MM-3; Eric Sabotka Co., Inc., Farmingdale, NY) was positioned with its point on the atlanto-occipital membrane surface, near the dorsal edge of the axis opening, and in a plane parallel to the stereotaxic frame; polyethylene tubing (PE 90; nom. 25 cm in length) was attached to it as the outflow cannula. The needle was advanced 6 mm through the atlanto-occipital membrane and underlying dura. A successful cisternal puncture was confirmed by the free flow of clear cerebrospinal fluid from the outflow cannula. The puncture site was then reinforced with collodion (Mallinckrodt Chemical Works, St. Louis, MO), and the outflow cannula attached to a reservoir containing artificial cat cerebrospinal fluid (CSF; Appendix F). A 4 to 6 cm midline incision was made in the skin of the head from a point 4 cm caudal to the supraorbital ridge to a point 2 cm posterior to the intra-aural line. The skin was retracted, and the underlying muscle and connective tissue were cleared to expose a 3 cm2 area of frontal and parietal bone. Two 3 mm diameter holes (13.5 mm anterior to the intra-aural line; 2.5 mm lateral to the midline; were 76 drilled through the skull using a dental burr and drill (model 4N; Emesco Dental Co., Inc., New York, NY), exposing the dura over each cerebral hemisphere. Two metal needles (20 9a.; Becton, Dickinson 5 Co., Rutherford, NJ) heldin electrode carriers (model 1204; David KOpf Instruments, Tujunga, CA) were connected by a T-connector to a pressure transducer (Statham model P23Ac; Grass Instruments, Quincy, MA) for measuring intraventricular pressure (PCSF) and to two, 20 cc syringes on a constant speed drive pump (model 975; Harvard Apparatus Co., Dover, MA) for infusion of artificial cat CSF (Appendix F). The inflow needles were centered over the skull openings and lowered separately through brain tissue into the lateral ventricles; correct needle placement was indicated by a sudden drop in inflow pressure, and was further confirmed when changes in cister- nal reservoir height were reflected by similar changes in CSF pressure (PCSF). yENIRICflLQQISIEBNAL_REBEHSIQN_EXEEBIMENIS The infusion rate (V1) of mock cerebrospinal fluid (Appendix F) which contained a trace amount (nom. 7.5 nCi) of 14C-1abelled dextran (mw=70,000 D; New England Nuclear Corp., Boston, MA) was held constant during each experiment; Vi was 70 ill/min in some experiments and 140 nl/min in other experiments. The cisternal outflow line was disconnected from the CSF reservoir and joined to an adjustable height photocell 77 drop detector attached to a fraction collector (model 568; Instrumentation Specialties Company, Lincoln, NA) for cisternal effluent sampling. Two drop effluent samples (nom. 100 pl) were collected continuously during ventriculocisternal perfusion. The height of the outflow line could be adjusted relative to the zero level at the animal's intra—aural line in order to regulate intraventricular pressure (PCSF). In addition to systemic arterial pressure and respiration, intraventricular pressure and outflow rate were monitored and displayed continuously using a polygraph (model 5P; Grass Instrument Co., Quincy, MA). The arrangement of apparatus used for the ventriculocisternal perfusions is shown in Figure l. E l J E' H' l E E E . Ventricular perfusion was continuous while each animal initially breathed room air (Phase I) for 30 to 40 minutes, and then breathed a 7% COZ-in-air gas mixture (Phase II) for 30 minutes; during these two initial phases, CSF pressure was maintained between 0 and 5 cm of water (relative to the level of the auditory meatus). The animal was then returned to room air breathing for approximately 40 to 45 minutes. During this period (Phase III), PCSF was slowly increased and maintained between 20 and 25 cm of water by gradually raising and keeping the cisternal outflow line above the level of the intra-aural line. Phases I and II were then repeated under the high CSF pressure conditions. 78 .mHmAHccm mac cooHn can we now couooHHoo maamofioowuwm mum moacemm cooah Hmauouuc .Acsonm uocv awesomcmbu musmmoum m ou cwuowccoo mfisccmo Hmfiuwuum HmuoEou a sm50unu cmuouflcoe ma musmmmum pooan cam Aczonm uocv poxcmah mcfiummn counsnpm adamscme m an ucmumcoo pawn m“ can encum noumaEuwnu m cues cmuouficoa m« on:LMuomEmu Hmuoom .wocobcoum muoumuwmmou one much 30am has mcficuoowu Lou umoscmcmLu ousmmoum Hmfiucwuommac m on wouooccoo ma gowns accumonomu055ocm nomfioam a smsounu concouflp ma 0>Hs> may soda mom cmuficxm one .mHsccmo Hamsomuu m 0» conomuum wahsmmmm m>Hm> mommmcmoo 30H .mocmumfimmu 30H m caucus» Anacmmouomhnv musuxwe mum basicalmoo m no AcacQMOOEuocv ufimIEoou donned monumwuh aawsomcmucomm anagcm one .masomaos house» on» Low commando mum mmu ucosammm mo mmHmEmm can .uoucsoo coup m magma consumes m« Ao¢v mumu soamuso .mcmme acumumfio on» ouch couummcfi wasccmo m an“: .umoscmcmuu wusmmmum m an omuouficoe ma wusmmoum mmo was made m>wuclmmcwu>m m an ucmumcoo cams we Aa>v mumu scams“ mmu .mmaowuucm> Hauhmumo Hmuouma on» one“ occasc ma Accuuxoc Ioewv masomaos mmuma m mo aufiucmsc women m mewcwmucoo Ammo. pagan Hmcflmmoanumo Hmaofim«uum cm dawns Acsonm Docv messy onmuomuwum 8 ca monsoon ma umo cmufiuwnummcm as no cams one .usumummns Houcoafluwmwo cofimnmuom Hmcuoumwoousouuusob mo causewaum .H MKDUHM 79 mwoaom—‘(mh wmnmmmcn mmODDm2(1.—. 304... m>4<> ><3Iw IIIJIIJI E< omEaxm D _ _ IJDI. 45> w4m2 0, then Vo > Vi and co/ci < 1. With induction of hypercapnia, Vo transiently increases (at tiv) and co/ci transiently decreases (at tic), but both converge (at tfv and tfc, respectively) on their room air breathing levels. These transient responses, identified by shading in Figure 2, can be quantified by determining their areas; the amount of transiently displaced CSF (Vd) is defined as: t . Vo + Vo Vd = fV Vodt - [ €021 ra1 (tf - t- t 2 V ) 1v 1v - [Vora (l + tiv) - Vocoztiv] (6) where: Vd = volume of displaced CSF (pl) 84 FIGURE 2. Representative time courses for CSF transient responses to hypercapnia. Shown are Vo (cisternal outflow rate, ul/min), Vi (perfusion inflow rate, ul/min), and co/ci (ratio between inflow and outflow tracer concentrations, dimensionless) for room-air (Phase I) and C02 (Phase II) breathing. Transient responses of V0 (Vd; ml) and co/ci (Cd; min'l) to hyper- capnia are defined over time interval tiv (initial time for expression of transient Vo response and onset of C02 breath- ing, min) to tfv (final time for expression of V0 transient response, min), tic (initial time for co/ci transient re- sponse, min) to tfc (final time for experession of co/ci transient response). Cr (min) is the residual area of co/ci over the time interval tic to tfc. 85 v MI )(I )(I (II V0.» ~—---/ . . W» a a CT E E I i m e p 7;L 1 1 ° 'iv 'fv IROOM AIR—+i596002 ,< l - C O/ci M ____ i 0 l #f/ O FIGURE 2 . 86 tiv' tfv a time for onset of C02 breathing and V0 transient, and for the end of V0 transient, respectively (min) vora, Voco2 s steady state perfusion effluent rate for room air and for C02 breathing (ul/min) and, (co/ci) + (co/ci) Ct '-‘ ra £2 (tf " t1 ) (7) 2 C C where: Ct = total area of co/ci assuming no dilution during tic to tfc (min) ti , tf = times for initial and final co/ci c c transient (min) (co/ci)ra, (co/ci)co2 8 steady state perfusion tracer concentration ratios for room air and C02 breathing and, (co/ci) + (co/ci) Cd 3 [ °°22 ra1 (tfc - tic) (8) tr - , _ c (co/ci)ra (1 + tlc) -/:ic (co/ci)dt where: Cd . area of co/ci dilution during transient period (min) 87 and, Cr = Ct - Cd (9) where: Cr = residual co/ci area during transient period (min) and, CSth = Vd (Cd/Ct) (10) where: CSth = amount of unequilibrated or tracer-free CSF in Vd (n1) Cd/Ct = area of transient dilution of co/ci expressed as a fraction of the total area of co/ci if no dilution occurred during tic to tfc (dimensionless) Integrations in equations 5 and 7 were executed by digitiz- ing (model GP—3; Science Accessories Corp., Southport, CT) the graphical forms of the outflow data and processing these results on the PDP 11/34 computer (Digital Equipment Corp", Maynard, MA; Appendix D). SIAIISIICAL.ANALXEES Tabular data are presented as means and standard errors of the mean. .A paired t-test was used to evaluate differences between group means of CSF transient response data. Steady state data were evaluated using a t-test 88 designed for testing differences among group means having unequal sample sizes; a t-statistic was computed using the difference between group means, and the best estimate of the true population variance (a pooled variance). A 0.05 level of significance was chosen to test the null hypothesis and was determined prior to experimentation. RESULTS Intracranial fluid volume is defined by a dynamic recip- rocal relationship between intracranial cerebrospinal fluid (CSF) volume and brain blood volume (BBV). Since the skull rigidly confines these intracranial fluid volumes, any sud- den change in one of the volumes will necessitate a rapid reciprocal adjustment of the other. C02 inhalation is a potent cerebral vasodilator, and is known to increase brain blood flow. Ventriculocisternal perfusion studies using tracers have shown that C02 breathing also causes immediate intracranial CSF volume and pressure changes, presumably due to increases in BBV. As shown in Figure 2, when the animal breathes room air, and CSF pressure (PCSF) is less than approx 10 cmHZO, ventriculocisternal perfusion outflow rate (V0) is larger than inflow rate (Vi) because CSF is being formed continu- ously within the brain's ventricles. Since the nascent CSF is free of tracer molecules, it dilutes the tracer in the perfusate such that co/ci, the ratio between outflow (co) and inflow (ci) tracer concentration, is less than 1. With- in moments of breathing C02, although perfusion inflow rate (Vi) is constant, there is an immediate and short-term 89 90 increase in PCSF and V0, and a transient decrease in CSF effluent tracer concentration (co/oi). Figure 3 presents data from one animal which show these characteristic changes in cisternal effluent tracer concentration (co/ci), perfusate volume flow rate (V0), and CSF pressure (PCSF) during ventriculocisternal perfusion when the animal spontaneously breathes room air and then a COz-in-air gas mixture. In this particular experiment when 7.2% C02 was inhaled at time - 28 minutes, Vo immediately increased from 155 ul/min to 340 ul/min and PCSF increased from 5 cmHZO to 7 cmHZO, and both gradually returned (approx 10 min and 3 min, respectively) to their pre-hypercapnic levels. Further, after a brief delay (approx 2 min) following hypercapnia induction, co/ci declined from 0.81 to 0.65 and returned to the pre-hypercapnic level within 12 min. These data are interpreted to indicate that respiratory acidosis has no effect on steady state rates of CSF production (Vf) and absorption (Va), but does displace a volume of CSF (Vd) which is relatively tracer-free (CSth). This transient expulsion of dilute CSF could be due to an episodic increase in Vf. Alternatively, if tracer molecules are heterogeneously dispersed in the CSF spaces, it could be due to a shift in intracranial fluid volumes, with consequent expulsion of tracer-free CSF into perfusion effluent. Data from similar hypercapnic tests conducted in experiments designed to investigate the generation of the CSF transients are presented in the following sections. 91 .cofiuosccu mficmmouoman meadow AcumEo b on my mummuocfi unmanned» unmwam m bosonm mmum .MHcmmouoman mo cofiuosccw on» nouns cfia a haoumefixOLQQm bonusooo ucowmcmuu fi0\oo on» no mcwccammn one .cfie NH away“: Hm>mH sachOLomanlwum as» ou codename can mm.o ou Hm.o soda pocaaomc «0\oo .o> c“ mafia coppsm was» ccfihcmceoooc .Hm>ma mficmm0ummasloum on» on Ange ca xouccmv nusuou Hannahm m an boonHou cfiE\H1 own on cfiE\Hd mmH scum o> c“ ammouocfl oumapmssfi cm was mean» Ange mm u mefiuv mficcwouommn mo cofiuosccfi spas .NOO ww.> conusmun assess any nuances on 0» mm Eoum «has Econ cmzumouh Ame umov Hmeficm on» oaflns cmumwxm mumum memoum cowmsmuom m nauseae mm ou o scum .QEoc m>fiuclmmcHu>m m an A:AE\H1 mma .Eocv ucmumcoo cam: was AH>V mums scams“ acawsmumm .mous:HE ca beau ma mmmfiowhm one .AONmEo «mmumv musmmouc mmu can «A:HE\H1 ao¢v mums ucosfimmo mmu «AH0\ooV soaucd :oamswuom on» a“ away no :ofiuomum m we commoucxm ucmsammw mmo ca c0wumuucoocoo cmuuxop “Asouuoh on do» Eoumv mum oumcwcuo on» so cmucmmmucom .mmanmfium> cofimsmuwc Ammuv madam HacficmOLQwuoo co umwu owccmoumcmn a mo muomwmm oumum momwum can ucmfimcmLu on» one noncommuc .:0nm=uuom Hmcuoumw00H50wuuco> ObnumucmmoHQOH o scum sumo .m MKDGHm 92 on omen 0v mm mm F QM NM 0 '0 mm ONO. O 1.].in . . . A656. 92.... ‘ q 4‘ i q q d 1 - -----.... '1 .8 .x. me. A... I NI.Eu. umul .0 to £00. .562... LJJJJAALAALIIAJAJJ+JLJJJAJ .m 956: w. 00. 0?. OO- CNN owN Don own 93 Analyses were made not only for CSF transient responses (refer to MATERIAL AND METHODS: MATHEMATICAL ANALYSIS AND CALCULATIONS; TRANSIENT RESPONSE ANALYSIS), but also for steady state respiratory, cardiovascular, and CSF responses (refer to MATERIALS AND METHODS: MATHEMATICAL ANALYSIS AND CALCULATIONS; STEADY STATE ANALYSIS). PROTOCOL A: Effects of High CSF Pressure on CSF Steady State and Transient Responses to Hypercapnia In this series of perfusion experiments, animals spontaneously breathed room air and then a 7.2% COz-in-air gas mixture first while CSF pressure (PCSF) was normal (nom. 0 to 5 cmHZO), and then again after their PCSF was experimentally raised (nom. 18 cmHZO). Minute ventilation (VB) and Paco2 increased and arterial pH decreased significantly (P<0.05) indicating that respiratory acidosis was established during C02 breathing both before and after PCSF was increased (Table 1). The increase in VB was due to a significant increase in tidal volume (VT, p0.05). Blood pressure (MABP), cerebral perfusion pressure (CPP; CPP a MABP - PCSF), and body temperature (Tr) were unaffected (P>0.05) by either the respiratory acidosis or the increased PCSF. Analyses of the steady state CSF responses to room air and C02 breathing under normal and high CSF pressure conditions (Table 2) revealed that increasing PCSF to mcoduwccoo ouswmoum mmo anon Lou mo.o u NooHa can: «nos» scum .mo.o v av neonauuse saueeoauseman .~.n a woven can: neafln> a managed m you saw H menus but mosHm> m o.~ « and... 86 a n8... mm. a no; Em H nmdm e H on .6 .« ..am a u no. a w a: an.» . ~82 ad « Twm 86 a 3;. me H «mm a... a «.2. m u - «a a 9% Z a 8. S a a: 8.0 . ~82 anon new: A. 9 p... w n~.mm no... a. as... -~ .« a8... a... H and: a u. on ~.o a 9% a u on. a H m... an... . ~82 M... a 98 «a... u an... 8 u 3... a... H Tm... m H mm «a .« 9% 2 u :— 2 u a: 86 . ~82 eaaasee. Argosy Anaemnaa-xnnv awakens-V A.e«a\nesaoenv Aces Aeneas Antony Noon. aza us as e as see ans: .sammumv ousmmmum Emu so“: car finance um c0umsmuoc an: nuwumfiooasoauuco> mcausp Am~.h u mouHmv maccmouocmn can Amo.o u Noummv IoEuoc ou noncommwm .A< HoooDOLm. numb uoasomm>oflcumo can muoumLficmwu mumum zcsmum .H mqm¢9 95 whammmum mmo Hmsuo: mo mmosu scum Amo.o v my ucmummuwc maucmofiuficmfim mosam> whammmum mmo scam n waamofiumnucmumm csonm mameficm mo umhfisc s 0 :mm + memos mum mos~m> m vao.. .u o.—m Amvm.s .« ..=m ASN.3 H —.mm Ago; H v.9. 33$ H o.mm Amvp.e .« o.e~ 33.5. H Tam “Neo.~ .« o.om Anvee.o.« .o.o varo.o.fl mo.o Amvoo.o.fl mp.o Amvwoé H N05 3:6 H m. pmp Agofi. H n.NNP 93.0 H mim- SZ..o H 2.35 Asvm.o.n s.am. asvm.e.u s.am. “5mg. H {amp 33.9.0 H 3.2: Ames.o.« s.c. “mvo.o.u m.e_ ASN.9 H 9... 73¢... H EN u~.~ u wouHm N uo.o u oUHa mmom :3”: an.» n wooHn n ae.o u ooHa swam a ea «exec no» «> names .efiamumv enannend emu can: can HmEuo: um cofimsmuwm Hanuwumwooasofluucm> ccwuso Awm.h u NOUHmV scammOmewn can “wo.o n NoonV noELoc on nmncomnem .Aa Hooouonav when emu eueun seeeum .N mamas 96 approximately 17 cmHZO significantly decreased perfusion outflow rate (Vo) during both normocapnia and hypercapnia (P<0.05). Moreover, there was a tendency for steady state CSF absorption rate (Va) to increase under the high PCSF conditions. Steady state Vf and co/ci, however, were unaffected (P>0.05) by either the respiratory acidosis or the increased PCSF (Table 2). Figure 4 presents data from one animal which are representative of the CSF steady state and transient responses to hypercapnia under the two pressure conditions. The normocapnic animal was in a perfusion steady state (i.e., co/ci and V0 were stable) after which hypercapnia was induced (FICO2 = 7.2%; time = 28 min). When PCSF was normal (light tracing), V0 immediately increased from 155 ul/min to 340 ul/min and PCSF from 5 cmHZO to 7 cmHZO, and both then gradually returned (approx 10 min and 8 min, respectively) to their prehypercapnic levels. Further, after a delay of approximately 2 min when FICO2 was increased, co/ci declined from 0.81 to 0.65 and returned to the prehypercapnic level within 12 minutes. Under the high PCSF conditions, Vo increased from 139 nl/min to 260 ul/min and PCSF increased from 16 cmHZO to 18 cmHZO, followed by a gradual return (approx 10 min and 3 min, respectively) to their prehypercapnic levels. Two minutes after hypercapnic induction, co/ci declined from 0.85 to 0.80 and returned to the prehypercapnic level within 13 minutes. 97 .nmfin was a“ cons cmnu HmEuoc was mmom cons uwummuc mums H0\oo ca ammouomp can o> CH mmmmLOCH ucmwmcmuu may .cwmcmaocs mums mmmusoo 65“» Leona coconua< .ucosflumm emu on» cw cowumuucwocoo cmuuch CH ammouomp ucwfimcmLu a new .o> cg ommmuocfi unmanned» cam mumfiwwEEH cm .mmoc c“ antenna“ ucmfimcmLu cam Hanan m maco boos©0um mficchLwcmn .sma: no HMEuoc was mmom umnuwnz .ONmEo ma 0» ONmEo N scum comamu was mmum cons unnumcoo nonwmewu w0\oo can commouooc o> mumum wcmouw .oacmmUOEuoc mm; assess any cons .Acfie mm u mefiuv ~00 wm.h conummun.amswcm on» cogs pounce“ was decadence»: scans bonus “awe mm 0» c oumum museum cofimsmbmc m :w was Ame umov HmE«cm OMCQmUOEuoc one .mEsm w>wuplmmcfiu>m m wn “afiE\H: mmav ucmumcoo camp was ~H>v muse soamcfi scamsmem .Ammmfiomhm «away mafia mo cofiuocsm m mm cmuuofic can Asouuon ou co» Eoumv mumcwcuo on» no coucmwmucmu mum AONmEo «mmocv musmmmuc mmo.©cm .ACHE\H1 ao>v muse ucmsammw mmo .AHO\oov scams“ scamsmumc on» Ca umnu mo :ofiuomum m mm commwucxw ucwsammm mmo a“ :ofiumLucmocoo swuuxmo .Ammcfia xumo “Ommeo may amen on on cwumsnpm waamucmefiuwcxm was DH :mn3 can» can Amwcfia unmflfi «ONmEo m 0» av HMEuo: was Ammomv musmmwuc mmu mafia; umbwm cmuospcoo mums mummy oficmmoumc>m .At Hooououmv scamsmuum Accumumwooasowuucob unannoum mmu can: ObaumucomOHQOH a mo measmom .v mmDUHm 98 NO OO 00 On OO. 0% 00 N 11‘! 4 1 d d ildi d 4 0n No On 0 \BOCON O ‘ d. ‘ ‘ ‘ d ‘ ‘ ..... 1r 9. coon ov mm mm en mm on am one. o all-1.1T. . . .4 “.225. NSC. LII .32 .9550: wcammwca umu LI in ‘ 14-: dilqlidl \eJIJIdIJ I {III 83...... umua «00$ «.5 A... :9 E2: Hi... 3:62.: LlllljllllAlliLllAjjlllLlllJ .: 950E O O. ON 00. O'- OO. ONN OON OOn OVn a. h. 99 Data reported in Table 3 indicate that although the total transient co/ci area (Ct; Eq. 7), the time for the V0 transient (tfv - tiv) and the co/ci transient (tfc - tic), and the time between the onset of Cd and the initiation of C02 breathing (tic - tiv) were unaffected by increasing the PCSF (P>0.05), the amount of displaced CSF during the V0 transient (Vd; Eq. 6) and the area of the transient co/ci dilution (Cd; Eq. 8) were reduced (P<0.05) by 60% and 50%, respectively. Under the normal PCSF conditions, an average of 81.4 ml (CSth) of the expelled CSF was unequilibrated or tracer-free while only 19.1 ul of Vd remained tracer free after PCSF was increased to 18 cmHZO. This corresponds to an 11% reduction in co/ci when PCSF was normal and a 5% reduction when PCSF was high (Cd/Ct; Table 3). PROTOCOL B: Effects of StOp-Flow on CSF Steady State and Transient Responses to Hypercapnia In this series of perfusion experiments, animals spontaneously breathed room air and then a 4.5% COz-in-air gas mixture before and after a 45 min period of stop-flow (S-F) perfusion. Minute ventilation (VB) and Paco2 increased and arterial pH decreased significantly (P<0.05), indicating that respiratory acidosis was established during C02 breathing both before and after S-F (Table 4). The increase in VB was due to a significant increase in tidal volume (VT; P<0.05) while respiratory frequency (f) did not change (P>0.05). Blood pressure (MABP), cerebral perfusion 100 anon Hesse: do each acne Amo.o v av heeneuuae saueaoamaeman eases anon gene a mHHmoHumnucmumm csonm macawcm mo Logan: «2mm H memos mum mosam> m “~36 H —.o— sync... H mo.o 32.6 H -.o .mvoéh H 3.8 33o... H :6 3:6 H a.» Amvmé H 0.0 Amvmé H m.~ Amvmé H ~.~p 8:6 H m.o Clo.— H m.: 3:.me H pJom amon— :8: ago; H o6 25.9 H m.~ “Cm; H 5.3 Amvmd H o.— 2.3.0 H v.3 33.09 H 991. anon 4:202 Anny Anus. Dunn—mo .625 .10 A55 .53 25.5 253 A55 :3 > O 0 O > > so as n as as .. Cs n8 as u e» as, moo mm.» mo :ofiumamncfi on .mammomv musmmmuc mmo ace; cam HmEuo: Lops: noncommmm .A< Honou0umv muse sonuso emu ucmfimcmLe .m mqm<9 11()l .HIAw euwumme was .enoueh. such new wo.o u «OOH. ewes emcee seem Amo.o v EL sentences saueeoausemfin sm.e n NooHn gene neaae> h haamowuonucmumm czosm manages mo Cohan: s a :mm + wcmwe mum mosam> m .av ..~.« am.o. any ~o.e.« a.~.s Tosca—m.u emmp. .o..o...« on... 1o.vm.« an A..v~.o_« ~.om Am_es_n so .m..m.« ma um.e u mooHa has m.m.« a.sm Ame ~e.o_« mM.. .e.vmo .« Lee Lopes.. a. ~.o~ A.FV~.« em as.vp.o.« m.om an..m.« ea Ampvm_u ma no.o . NooHe a.» ensue .m_v..~.« np.ms Awpvwo.o.« nm~.~ Aspeem_.u asmm. nervo.m.« no.om .e.L~.u mm Aspep.o.« F.mm As_vm.« Lop Aepcm_q mo. um.. . Noon. 83a; « e.~m 3:86 a on.» 3:: « mmo 33...... A TE 2:... H 9.... 23.6 H Nam 2:: a no 25.. a no no... ... moo: mum mzonmm Agree. .maem.=«-\sav Anaemxsav A.ea-\naaeoeav Loos Aeneas Agrees .mamlmv soHMImoum nouns cam ouommh cofimsmumc as: N“ lumumfiopooHuuco> mcfiusc Awm.¢ moonv macmmuuocmn 0cm Awo.o u ouHm. IoEuo: ou noncommmm .Am Hooououmv sump Hmasomm>oH©umo can muoumuficmmu mumum acmoum .v mam<9 102 pressure (CPP) and body temperature (Tr) were unaffected (P>0.05) by either the respiratory acidosis or the stop-flow procedure. Analysis of the steady state CSF responses to room air and C02 breathing before and after S-F (Table 5) revealed that steady state perfusion outflow rate (V0), dextran dilution (co/0i), CSF pressure (PCSF) and rates of formation (Vf) and absorption (Va) were maintained at the same levels (P>0.05) for normocapnic and hypercapnic periods both before and after S-F. ' Data in Figure 5 from one animal are representative of the CSF steady state (i.e., V0 and 00/01 were stable) and transient responses to hypercapnia before and after S-F. The normocapnic animal was in a perfusion steady state after which hypercapnia was induced (FICO2 a 4.5%; time = 0 min). Before S-F, V0 immediately increased from 87 nl/min to 146 nl/min and then gradually'returned (approx 10 min) to the prehypercapnic level. Approximately 2 minutes after hypercapnia induction, co/ci declined from 0.80 to 0:72 and returned to the prehypercapnic level within 12 minutes. After S-F, V0 immediately increased from 86 ul/min to 140 ul/min, followed by a gradual return (approx 10 min) to the prehypercapnic level. There was again a delay in the dilution of the dextran marker but co/ci was reduced from a normocapnic value of 0:78 to a minimum of 0:74 after which it returned to its prehypercapnic level within approximately 14 minutes. Although the transient change in V0 was the 103 2mm H memos mum mosam> m m; H m... fin H .2: no... H 85 En H TS is H ..:. To H Tm m um... u ~82 ~2— H m.m Em H m4; mo.o H $6 ed H Ema :6 H .32. m.o H m.~ m 36 u moo: 2am zmam< o.—.H ..o o...H u..~ ~o.o.H sp.o ..~.H o.mo m.o.H m..s m.o.H m.~ _. um.e u «ovum m...H o.o o...H m.o~ ~o.o.H ss.o a...H m.~a m.m.H m.ae ~.o.H o.~ o. no.0 u NooHa alwmfixfim 2.323 2.323 Guile 2:51: Emmaev as ea Hexoo o> u> amen e .elaumv scum noun neuue cam ououmn :0uwsuu0m Hmcuoumu00H20uuu20> ocuusc Awm.v u Noonv mucmmouocmn new Amo.o mouHmv loanoc ou oncocmom .Am Hooououmv sump Emu mumum acmoum .m mqm<9 104 .Amoum pmcuuum mnu >2 czonm may mum ououon cmnu uouum mmou was .u0>030n .u0\oo nu oncouooc ucoumcmuu one .muspooouc mum mnu an omuoouumcs mums u0\00 can o> mo mHm>0H mumum wbmoum onu can .momcocmou ucoumcmuu onu mo momusoo osuu onu .o> Cu masouocfi ucoumcmLu one .u0\0o cu ommouooo ucoumcmuu m .ncue N xoumcmv amaob umuuh m nouum .ccm o> cu 0m00u0Cu mucuceru cm comsmo uu was new couummcH mnu ou moo wm.e acumen >h Anus o u mauuv coosccu mm: nucmmouocxm .mnm umumm can muomoh nuon Anus o ou mlv oumum acmmum coumsmuoc 0 cu was Ahmu uwov unaucm ouccmooeuoc one .cEsc m>uuplomCHu>m m an AcuE\H: «by ucmumcoo pawn was Auov oumu soaucu :Oumsuumm .Ammmuomhw «quay mEuu mo couuocsu H mm oouuoac mum Aoumcucuo Eouuon “CHE\H1 uo>v mumL ucmsamum mmu mam AmumcuCLO mou «u0\oov 30HMCH c0um=muwm onu Cu umnu uo couuomuu m mm commoucxo ucosauuo mmv nu couuMLucwocoo cmLuxoa .Amlmv soaulmoum uo couuoc sue me a AmocHu unabv umuum can Amocfla uncuav ououoh pwuospcoo mums mumou oucmmou0c>m .Am H000u0um. :0umsuuom DOHHImoum obuumucomoumou m we muusmom .m MMDOHb 105 35833:» .m "22on ON 0. Op Ifi fl 1 an ——---#O aOo 8.3.1 :- see. . , loo. . on. .0:( ll 0.0.00 ..|ll " " low. 302 no; . " 2.5.3 0> - O . L . . J . . i. . . . 'nll, . . 1“ E! no " _0\00 I 106 same before and after S-F, that for co/ci was less after S-F when compared with before S-F. Data reported in Table 6 summarize results obtained using st0p-flow in 8 animals. They indicate that although the volume of transiently displaced CSF (Vd), the total transient co/ci area (Ct), the transit times for Vd and Cd [(tfv - tiv) and (tfc:- tic), respectively], and the time between the onset of Cd and the initiation of C02 breathing (tic - tiv) were unaffected by the S-F procedure (P>0.05), the dextran dilution (Cd) was reduced (P<0.05) by 50%. Before S-F, an average of 24.1‘u1 (CSth) of the expelled CSF was unequilibrated or tracer-free. Only 10.3 ul of Vd remained tracer free after stop-flow. This corresponds to an 8% reduction in co/ci before S-F and a 4% reduction after S-F (Cd/Ct; Table 6). PROTOCOL C: Effects of Phenoxybenzamine on CSF Steady State and Transient Responses to Hypercapnia In this series of perfusion experiments, animals spon- taneously breathed room air and then a 7.2% COz-in-air gas mixture before and after an intravenous injection of phenoxybenzamine (P82) (2 mg/kg). Minute ventilation (VB) and Paco2 increased and arterial pH decreased significantly (P<0.05) indicating that respiratory acidosis was estab- lished during C02 breathing both before and after PBz (Table 7). The increase in VB was due to a significant increase (P<0.05) in tidal volume (VT) and respiratory frequency (f). 107 .mlm cOHOMOD: MO “$5.... EOHM Amoco V mv HCOHOMMHG ewHUCwOHMHCDfim OSHO> .mlm :HOUM‘: D .mumeucm w you :mm H memos mum mosHm> m s.m .H m.o. .o.o.H so.o o...H ..p ....H a.» m.o.H m.m s...H m.a ..o.H =.o m...H m.m s.pm.H m.mmm mum emcee o.o..H P.s~ mo.o.H oo.o p.o.H a.» o.o.H o.w m.o.H m.~ ..F.H a.o. m.o.H m.e s.._H a.» F.2m.H s.mmm mum mzoamm 31v 253v 253 23.5 25.5 $3.5 25.5 :43 > O O O > > Chemo euoxeo no so us - us H» u as see as u e» as .mumlmv soawlcoum Loumm cam ououon N8 em.e mo couueueeeu on neneoenem .Im Honouonav when aouuuao emu neeuneene .G mamas 108 Nmm cuwumflc HO OWOSH EOHN AmO.o V my ucwummmwc >HUCMOHMfir—Gflm MODHM> Nmnw cwuowwm: U Nmm cumum0e was .enoumn. anon Lou mo.o u mouHm cons muonu EOLu .mo.o v my ucououufic maucmofiuucmum mm.h n mouHm cons mos~m> n aaamowuonucoumm czonm mHmecm uo amoeba «2mm + memos mum mosHm> m Team...« a.... .p.~a.a.« an... A.c....« ...m. Ase... .H as... Lev1.« as. use..a.« _.e. .e.. .H a. use. .a no .~.e . Na... “use...H ¢.mm upvno.o.H o~.~ “mean .H see Imam.. .H o.mm has..H m~ Lev—.e.H o.mm .ev. .H on Lee. .H ea ue.o . ~ooH2 Nam zuam< .m.....« as... Am.ma.a.H an... Amvsmn.u amoa~ Ance.~..u 2.... .mv~.H a_m .m...o.H ..em Ames..H o~. ance,.H an. u~.p . ~ooH. neee.~.H ..mm .scma.a.H n... nev~e .H on. Leea.m .H a... Levm.H an .eee.a.H o.em .5.o..H an. 123°..H .m. ea.a . nova. Nam uxomum Annohv Ammannculxuev Ammhmnuav A.:«I\nnumonpv Aoov Annohv antoav ~82 J... as as L a. 9.8 one... .maammv oeuEmnconwxoconm Louum cam ouomon c0umsuuom HmcuoumuooHsOuuuco> mauusm Aw~.> u ouHmV muccmouocan cam Amo.o u Noonv IoEuo: ou noncommom .AU HooouOLmv sump umusomm>ofi©umo cam auoumLucmou oumum acmoum .h agree 109 Both normocapnic and hypercapnic blood pressure (MABP) and cerebral perfusion pressure (CPP) were reduced (P<0.05) by 30% after PBZ, while body temperature (Tr) was unaffected (P>0.05) by either the respiratory acidosis or the P32 treatment. Analysis of the steady state CSF responses to room air and C02 breathing before and after PBz (Table 8) revealed that after PBz the 14C-dextran perfusate concentration (co/ci) was increased (P<0.05), and the rates of CSF formation (Vf) and absorption (Va) were significantly decreased (P<0.05), while PCSF and perfusion outflow rate (V0) were unaffected (P>0.05) by either the respiratory acidosis or the P82 treatment. Data shown in Figure 6 are from one animal, and are representative of the CSF steady state and transient responses to hypercapnia before and after PBz. The normocapnic animal was in a perfusion steady state (i.e., V0 and co/ci were stable; time -5 to 0 minutes) and hypercapnia was induced (FICO2 = 7.2%) at time = 0 minutes. Before PBZ, V0 increased form 160 ul/min during room air breathing to 395 ul/min following hypercapnia induction, and then gradually returned (approx 9 min) to the prehypercapnic level. Further, after a brief delay (approx 2 min), co/ci declined from 0:79 to a minimum of 0.50 and returned to the prehypercapnic level within 12 minutes. After PBZ, V0 increased from a normocapnic value of 158 ul/min to 370 ul/min immediately after the onset of hypercapnia and then 110 Nmo .ooouoo. mo onoao eoou Ame.o v as oeooouuuo sauceouuueoun mosses umo Cumumcpp D waamowuonucouom czosm manages uo Embasc «2mm + memos oum mo=Hm> m AGO—.~.H.m.o onp.~.H m.P~ Amvmo.o.H mo.o onm.m.H p.9mu “ovm.o.H p.5m. “eva.o.H a.~ n~.s u moon 23.. H m.» E~.~ H n.- 2:86 H 85 2:; H Tr... Gem... H 92. Send H a; no... ... N82 22 22...: 3:2 H ET 3;... H 22 2:86 H 85 222 H GET 3:... H 92. 5.6 H ~.~ 2... u ~82 2%.. H Tm. 2:2 H ~..m 2:36 H ~26 3.0m H 3...; 826 H 2.9 326 H ..~ 86 u ~82 uma seamen 2:53 3253 351C 2255 on... 4:» $38 2.. 2., 282 .oammmv ocHEmuconmxocosm uoumm cam ououon conomuom HmcuoumfiooHsofiuuco> cowusc Aw~.b u moonv owcmmouomhn can Amo.o u ~00Ha. loanoe oo monsoonom .20 Hooouonnv when amo ououn seeoum .m mamas 111 .mp0uuoc owccmouomws com oficmmooEuoc nuon mauusp Aem.o ou mh.ov Nmm uouum uoumouo >HucooHMHcmHm ouoz .uo>030n .fi0\oo mo mHo>oH ououm mcmoum one .Aumm ououohv mooao> uncommon ucoumcmuu H0uucoo onu Eouu ucououmau amucmowuficmum uoc mums momusoo osuu was mocsuwcmme uuosu .Nmm uouum onuaoou ou wococcou m cm: uo\oo com o> nu mowcocmou ucoumcmLu onu smsonua< .wo\oo cu ommouoop ucoumcmuu o .Acue N soummmv amaoc uofiun m nouns can .o> cu omoouocH oumHmoEEu cm comomo uu can mom woufimmcfl osu ou Nou mm.v caucus an Ange o u oefiuv cocoon“ mos sucmwouommm .c0uumuumwcuecm Nmm Louum new ououoh nuon Anus o ou mlv oumum acmoum coumsuuoc m CH mos Amu umov HoEucm OHCQMUOEuo: one .cEsm o>wuuloccfiu>m m an A2HE\H1 hmav ucmumcoo cam: mos au>v ouou soHMCH coumsuuom .Aommuomhm «away oEuu uo couuocsu m an pouuoam mum noumcficuo Eouuon «cuE\H1 ao>v oumu ucosauuo emu cam Aoumcfiwuo cou «u0\oov soaucu coumsuuoc onu cfl umnu uo cofiuoouu o no commoucxo ucosauuo mmu cu couumuucoocoo couuxoa .A>u .mx\me m .Nmmv oeuemuconwxoconc mo couumuumucuecm Amocua xumcv Louuo cam “monua uzmfiav ououon pouosccoo.ou03 mumou oflcmm0uom>m .AU Hooououm. :0uusuuom Hmcuouuwoou=o«uuco> onwamusonhwocoam o>uumucom0umou a mo muHsmom .@ NMDUHh 112 A 9.25 m2: o~ o. o. o o o- 0.- «I 4 H q q A J _ .8 an. All ... a... . — J . .400. 1 n on. IJ h o- — IA — 1 O0“ - 1 . l 00» . 1 . 4 010 L...( . . 22.0 A 55:3 4 . . 1 can uzmazuc» . 2., L wz.a<~zwn>xozuxa _ _ l O. l .. _ .030 L . m mmDUHh 113 gradually returned (approx 15 min) to the prehypercapnic level. Following a 2 minute delay, co/ci declined from 0.84 to 0.60 and returned to the prehypercapnic level within 17 minutes. CSF transient response data collected from 5 animals that were injected with PBZ (2mg/kg: iv) are summarized in Table 9. The data indicate that the total dextran transient area (Ct). the time between the onset of Cd and the induction of hypercapnia (tiC - tiv). as well as the volume (Vd) and dilution (Cd) transients and their respective time courses [(tfv - tiv) and (tfc - tic)] were unaffected (P>0.05) by the PBZ treatment. Before PBZ, an average of 119.2 pl (CSth) of the expelled CSF was unequilibrated or tracer-free, representing 15% of Vd (Cd/Ct; Table 9). After P32, 11% of Vd, or an average of 83.9 pl, was tracer—free. Like the Vd and Cd values, the values for CSth and Cd/Ct tended to be reduced after P82; in all but one animal, these values did decrease to some degree after PBz. These differences were small, however, and the variability in CSF transient responses among animals was large. Apparent differences in CSF transient data from before and after PBz were insignificant (P>0.05). 114 waamowuwnucwumm csoam mameficm mo umnEs: «2mm H mcwwe mum mosam> m A.c=.o~.« o.mo Azcmo.o.« ...c ..vo.o.« ...— Aac~...« o.~F “mv~.o « m.. Amcm.._« m... “mom.o_u 2.. Anya...» c.,. Am....~._u p.mmo no; cause gym-.mw H N5: Asvmoé H .36 Anvm; H 06 33'.— H ~.~.. Amvmé H 0.. 2;... H 0.9 33.0 H ~2— AmvhJ H 0.: “30.29 H m6: Nam neommm Aaav Asa-V Ana-V Asa-V “cu-v Ann-v nan-V A—av t O O O > > guano ao\vo to so «a u «a «a u u» no «a - c» as .wANmmV wcfiEmncmbmxoconm nouns can muomob Nou w~.~ mo cofiumflmacfi 0» momcommmm .Ao Hooououmv some soauuso mmo ucmfim:Mue .m mamas DISCUSSION In 1954, Tschirgi et a1. administered a C02 gas mixture to anesthetized cats and observed a rapid and transient rise in the rate at which CSF drained from a cannula in the cisterna magna. Subsequent ventriculocisternal perfusion investigations using tracers have shown both an immediate transient increase in perfusion outflow rate (60) with hypercapnia induction, and a decrease in CSF effluent tracer concentration (co), even though perfusion inflow rate (Ti) and tracer concentration (ci) were stable (Martins, et a1., 1977; Songer et a1., 1980; Fisher et a1., 1983). This displacement of a relatively dilute CSF during early respiratory acidosis is attributed to two phenomena. One explanation assumes a homogeneous distribution of tracer throughout the CSF spaces and that CSF formation rate (6f) episodically increases during C02 breathing. Both the‘ volume and composition transients would be accounted for were there a volume of newly formed CSF injected into the test perfusate. The second explanation allows heterogeneous dispersion of tracer in the CSF, an unchanging 6f, and is based on a redistribution of craniospinal fluid volumes with C02 breathing (Martins et a1., 1977, Songer et a1., 1983). The increase in brain blood flow and volume during 115 116 hypercapnia (Reivich, 1964; Rich et a1., 1953; Grubb et a1., 1974; Adams et a1., 1980) empties CSF into the perfusate from spaces that were inadequately mixed by convection and diffusion. Both phenomena, of course, may contribute to the short-term displacement of a dilute CSF during hypercapnia. A major goal for the present eXperiments was to establish a data base to select between these explanations. The mass balance equations developed for ventriculocisternal perfusion tests (See METHODS; Equations l-S) accurately assess CSF dynamics only when the rates and tracer concentrations of perfusion inflow and outflow are stable and when craniospinal fluid volumes are unchanging. CSF transient responses (see Figure 2) to stimuli which perturb the craniospinal fluid balance must be analyzed using alternative methods. In the studies reported here, equations were developed to assess transient CSF responses to hypercapnia during ventriculocisternal perfusion. These equations, which were derived from the relationship between and the time courses of CSF outflow rate (To) and tracer concentration (co/ci), permit the quantification of the volume of perfusate (Vd) and tracer-free CSF (CSth) transiently extruded from the cisternal cannula during early respiratory acidosis, and define the amount of transient dilution (Cd) associated with Vd. Ventriculocisternal perfusion, which is widely accepted for measuring the rates of CSF formation (6f) and absorption (Va), was used in these experiments. This technique 117 involves placing cannulas into the lateral ventricles for artificial CSF inflow, and positioning a large-diameter needle in the cisterna magna for measuring outflow (Figure 78). Because the cisternal cannula provides an alternative flow route of low resistance in parallel with the subarachnoid spaces, normal CSF flow distribution (Figure 7A) is disrupted. During ventriculocisternal perfusion at normal pressure, both newly formed CSF and perfusion flow are shunted preferentially along the highest conductance course, that through the needle in the cisterna magna. Flow rate through the subarachnoid spaces increases, however, when CSF pressure rises, since high CSF pressure increases CSF bulk flow through the arachnoid villi (Welch and Friedman, 1960; Pollay and Welch, 1962). Because mixing of tracer in the total CSF volume during ventriculocisternal perfusion studies depends primarily on convection, enhanced subarachnoid CSF flow will increase the rate at which tracers are distributed by mass flow in these spaces. CSF transient responses to hypercapnia were measured before and after a period of facilitated flow through the subarachnoid space. Subarachnoid perfusate flow was augmented either by impeding cisternal efflux of perfusate by raising the level of the cisternal outflow cannula (i.e., high CSF pressure; Protocol A), or by preventing CSF outflow by clamping the cisternal outflow cannula (i.e., stop-flow perfusion; Protocol B). 118 The question of whether incomplete mixing of tracer throughout the CSF spaces and/or an increased 6f accounts for the CSF transient responses to hypercapnia also was approached in another way. CSF transients were assessed before and after the intravenous administration of a vasodilating agent, phenoxybenzamine, which is a competitive alpha-antagonist that binds irreversibly to both alphal and alphaz adrenoceptors (Goodman et a1., 1980). Chemical denervation was chosen over surgical denervation as a method of achieving cerebral vasodilation since it has been demonstrated that not all adrenergic fibers supplying brain parenchymal vessels and the choroid plexuses degenerate after denervation of sympathetic ganglia (Lindvall et a1., 1978a; Edvinsson et a1., 1975), and because there is a source of adrenergic fibers originating in the locus coeruleus (Hartman et a1., 1972; Raichle et a1., 1975; Rennels and Nelson, 1975). Cerebral vasodilation was produced by phenoxybenzamine prior to a hypercapnic test in order to evaluate the contribution of sympatho-adrenergic activity to the craniospinal redistribution and mixing pro- cess to the CSF transient response thereby unmasking any transient change in 6f associated with early respiratory acidosis. Results from stop-flow (S-F) experiments (Table 6) indicate that at least 50% of the transient dilution of cisternal outflow associated with C02 breathing during ventriculocisternal perfusion is due to the craniospinal 119 shunting of CSF unequilibrated with tracer molecules. The transient dilution of dextran (Cd) decreased by 50% after S-F, even though the total transient co/ci area (Ct) and the volume of transiently displaced CSF (Vd) remained constant. Further, Cd decreased during high CSF pressure perfusion (Table 3). This supports the S-F finding that mixing of tracer in the CSF spaces is incomplete during ventriculocisternal perfusion at normal CSF pressure. The time course for Vd and Cd [(tfv - tiv) and (tfc - tic)' respectively], and the delay between the beginning of Cd and the induction of hypercapnia [(tic - tiv); Figure 2) were unaffected by both S-F (Table 6) and high CSF pressure perfusions (Table 3). These data indicate that while the s-F and high CSF pressure treatments enhanced convective mixing of tracer throughout the CSF spaces, they did not change either the reciprocal relationship between CSF and brain blood volume or the time course of craniospinal fluid redistribution during acute hypercapnia. The decrease in Cd during high CSF pressure perfusion, however, in addition to being due to convective mixing may also be a function of a decrease in Vd (Table 3). This may be due to a high CSF pressure-induced decrease in transmural pressure of the pial vessels, which go through the subarachnoid space. A decreased transmural pressure in the highly compliant pial veins, might decrease venous distensibility so that the C02 stimulus effects a smaller increase in total brain blood volume which is smaller than 120 when CSF pressure and venous transmural pressure are normal. A smaller increase in brain blood volume would induce a smaller compensatory shunting of CSF from the cranium (i.e., Vd would decrease). Although this explanation may be valid when intracranial pressure is markedly increased, cerebrovascular responses to inhaled C02 (FICO2 = 5%) are undisturbed when intracranial pressures are less than about 45 Torr (Avezaat et a1., 1980). Moreover, neither sequential bouts of C02 inhalation (Avezaat et a1., 1980; Lofgren, 1973) nor physiological shifts of resting intracranial pressure (Marmarou et a1., 1973) independently alter the craniospinal fluid redistribution process. An alternative explanation for the decrease in Vd during high CSF pressure perfusion is that the elevated CSF pressure, produced by raising the height of the outflow line, increased CSF absorption (Ta) and increased cisternal outflow pressure so that CSF outflow rate (To) decreased. Although data in Table 2 show that To decreased during high CSF pressure, they do not show a significant increase in Ta. This may be due to the small number of experiments (n a 2). Other studies demonstrate that increasing CSF pressure causes an increase in 6a (Heisey et a1., 1962; Hochwald and Sahar, 1971; Mann et a1., 1979; Davson et a1., 1970; Marmarou et a1., 1978; Welch and Friedman, 1960). If there were an increase in 6a during high CSF pressure perfusion, the total volume of CSF leaving the cranium during acute 121 hypercapnia may be unchanged, but there may be more cranial CSF flowing through the arachnoid villi, and less flowing through the cisternal cannula than when the hypercapnic test was conducted at normal CSF pressure. These data indicate that overestimates of steady state 6f would result if measurements are made prematurely after perturbing the craniospinal fluid balance (Haywood and Vogh, 1978). This does not invalidate, however, the use of mass- balance equations for calculating steady state 6f and ya (Tables 2, 5, and 8), because they are appropriate for stable conditions of dynamic nonhomogeneity (Heisey et a1., 1962; Heisey et a1., 1983) as they are for steady states when perfusate mixing is complete and tracer distribution is uniform. Nonuniform tracer distribution occurs during ventriculocisternal perfusion because of the technique and the anatomy of the CSF spaces. The craniospinal CSF system is a multi-chambered labyrinth in which CSF circulates along a tortuous path from its sites of production in the ventricles to where it is absorbed in the craniospinal subarachnoid spaces. The normal CSF flow pattern (Figure 7A) is disrupted in ventriculocisternal perfusion because the perfusate mainstream is directed from the cerebral ventricles to the cisterna magna (Figure 7B). The perfusate flow and newly formed CSF are directed away from the sub- arachnoid spaces which reduces mixing in these spaces. 122 .ham>auoommmu .mmocmm mmo cfinuws can ou Hanuouxo mcofiuooufic 30am ucomoumou mzouum cwmoHo can some .moommm cwocsomHmnsm a“ mcfixfia can cmsousu 30am nuon mommmuocfi o AI: o> umsu om Hmccwso ucwsammm cofimsmumm on» msamEmHo no .wommm cwocnocumnsm on» a“ mcfixwe o>fiuoo>coo cocoon cfisos o> cmumfl xamumcofiuquOummww < .Ao>v mummsmuwm ucosammo nomHHoo ou mamas maumumfio oucfi condoms“ canoe: oocmumfimwu 30H m cmsousu omam pan .Amm> can om>v mousou cofiumuomnmou assoc: >9 moommm mmo w>mmH ou maoccmno Hmficcuomuucfi can unasowuuco> smsouzu mmmm won» umnuooou can .mm> can om> no pan» 0» upon much 30am wasao> muH .moaowuucw> Hmuwuca oucfl Aw>v mums accumcoo m an comssm ma moasowaoe accouu mcficfimucoo mmo HmonMfluum .consmucm HmcuoumfiooHsofiuucm> mewusa um .Amu>v mmufim HmcfiOuosomuuxm um can nom>v momsxmam QHOuono an coauou m“ a“ gowns um mmonu cu Hmsvm mound um mwommm owocnocumnsm Amm>v Hmcfimm can Aom>v anaembo scum conuomnmwu >Hm>fimmmm ma mmu waamEuoz "d .mCuouumm scam mmo oh MfiDUHh .m. magi s .> 0 Snow: — ouonu\ @ 32230 205.9052.» 0 f 3ch A R; :3... AH - .ofiuzuotxo [I u u . 0:32.. (A333... 30:5 3965 cocoa N/I/ u_occoo.oa=n\\ 3:38 R\R\ 124 When CSF pressure is increasing during ventriculocis- ternal perfusion either by raising the height of the cister- nal outflow line (Protocol A), or by occluding the cisternal cannula (Protocol B),icisternal efflux (60) diminishes (Table 2) or stops, respectively, and circulation of perfusate through the subarachnoid spaces increases. Con- vective mixing is increased and the transient dilution of cisternal effluent during early respiratory acidosis (Cd) is reduced (Tables 3 and 6). Data from the high CSF pressure and S-F perfusions do not explain why there was residual Cd after convective mixing was increased. It is unlikely that mixing was inadequate, since the time for the initiation of dilution [(tic - tiv); Figure 2] was the same before and after periods of facilitated flow distal to the cisternal cannula (Tables 3 and 6). Transit time for unequilibrated CSF would be expected to vary as a function of its distance from the cisternal outflow needle (Figure 7B). In addition, data from an auxiliary study (n = l) in which an animal underwent 70 minutes of S-F, and then a hypercapnic test (FICO2 = 7&2%) during ventriculocisternal perfusion (Figure 8), indi- cate that a transient dilution of dextran was still detect- able even after nearly doubling the duration of the S-F treatment. The percentage of Cd remaining could not be determined since there was no hypercapnic test preceeding the S-F period. 125 .wo\oo c“ owmcuocc ucomeMuu m .Acfie N xoummmv umaow «camp a sound .ccm o> aw wmmouocfi ucofim:Muu can mumficmeew cm comsmo mane .mwm ocufimmcfi may on Nov m~.h weapon an Acfie o u oefiuv pounce“ mm3 cacmmouwamn con3 Ache o o» mlv mumum acmwum c0wmsmuom w ca mc3 Ava“ umov Hmeficm UHCQMUOEuoc on» .mlm moum< .mEsm w>fiuolomau>m w an ACHE\H1 chv ucmumcoo mac: mm3 AH>V much 30amca scamsmumm .Ammmwomnm «:fiEv 08“» mo cofiuocsm m mm kuuoam mum Aouwcacuo Eouuon «cfiE\H1 «o>v much ucosammm mmo can AmuMchuo mo» «fio\oov BOHMCH cofimsuuwm on» ca can» no :ofiuomum 0 mm pommwumxm unusammw mmu cfi cofiuMuucoocoo swuuxmn .cofimsmumm Hanumumfioofisofiuucm> m mcfiusc omuosocoo ummu ofismmouomas mnu wommomum Amlmv soHMImoum mo cofiumm cfiE on c .sofimsuumm toHulmoum womGOHOHQ a scum mama om NMDUHh 126 35va2: 6. on on ma o... m. o. n o n- I 1 d I d I q d - 1 _ _ . . . «CO INK]. :: Eco. . . . . . _ 1 E525 . . O9 . _ . . . . . - . . - hlLlIlllul.‘ r. . . llLllL .oxoo J .m mmDUHh cap on. OON 127 One way to define mixing time would be to complete a series of experiments similar to Protocol B, but use stOp- flow periods with different durations. The transient CSF responses induced by hypercapnia both before and after the assigned S-F period would be analyzed for the volumes of displaced CSF (Vd) and the associated dilutions of dextran in the cisternal effluent (Cd). For those animals in which Vd remained a stable and independent variable, a difference between the before and after Cd values (dCd) would be deter- mined for each experiment and plotted as a function of S-F time. It is anticipated that for relatively short stop-flow periods, dCd will be small and vary directly as a function of S—F time. JFurther, it seems likely that as stop-flow perfusion time is extended, the dCd will increase to some maximum value and become independent of time. ‘At a certain S-F interval, complete mixing of tracer in the CSF spaces will be achieved and dCd will be at a maximum; further prolongation of stop-flow*will no longer have an effect on dCd. When dCd is maximized, the associated duration of S-F would be equivalent to the time it takes to establish a state of perfusion homogeneity. Then any transient dilution of tracer (Cd) occurring with hypercapnia during ventriculo- cisternal perfusion, and after a period of S-F, would neces- sarily be attributed to sources other than the contamination of the perfusion mainstream with poorly mixed CSF. CSF transient responses to hypercapnia may not be due entirely to craniospinal fluid redistribution. Increases in 128 CSF formation rate (9f) contribute to both Vd and Cd and account for the residual dilution after mixing (Tables 3 and 6). It has not been demonstrated, however, that 6f transiently increases with C02 breathing, partly because equations for CSF hydrodynamics require steady states (Heisey et a1., 1962; Heisey et a1., 1983). Also, the transient 6f response to C02 has not been investigated because steady state responses are assumed to reflect accurately transient responses. Portnoy et al. (1980) studied the interrelationships among systemic arterial, CSF, and superior sagittal sinus pressures, and the pulse waves of these pressures during normo- and hypercapnia. They discovered that for the ”closed skull" preparation (idh, no ventriculocisternal perfusion), transient and steady state cerebrovascular responses to hypercapnia are importantly different. During normocapnic and hypercapnic steady states, the CSF pressure pulse wave, which reflects the cerebral venous pulse wave, is transmitted to the superior sagittal sinus unchanged. This suggests that impedance to cerebral venous outflow provided by the intradural venous channels (lateral lacunae) is minimal. Impedance to blood flow from the cerebral veins to the superior sagittal sinus depends on the degree of compression of the lateral lacunaeu Conversely, during hypercapnia induction, the shape of the sagittal sinus pressure pulse wave is different from that of the CSF. This suggests an increased impedance to cerebral venous 129 outflow and an increased compression of the lateral lacunae. Portnoy and co-workers ascribe the transient compression of the lateral lacunae during hypercapnia induction to the transient increase in CSF pressure associated with hyper- capnia induction. CSF pressure increases because vascular expansion is more rapid than CSF absorption (Ga), and the increase in brain blood volume exceeds the capacity of the spinal CSF reservoir to accommodate shunted cranial CSF (Rich et a1., 1953; Marmarou et a1., 1978). {As hypercapnia continues and 6a progressively diminishes the total CSF volume (Martins et a1., 1972) to compensate for the expanded cranial blood volume, CSF pressure begins to fall and lateral lacunae compression is relieved. Compression of the lateral lacunae during hypercapnia induction retards drainage from its tributary intracerebral veins (Portnoy et a1., 1980) causing venous congestion. Venous congestion, in turn, is associated with an increased extravasation of Evans blue from venules located upstream to the congestion (Auer et a1., 1980). This suggests that hypercapnia induction may be associated with a transient ' disturbance of the blood-brain barrier as well as an in- creased capillary filtration rate. The increased rate of plasma filtration may effect a transient and significant increase in brain interstitial fluid flow into the brain ventricles (Rosenberg et a1., 1982) and this may be an extrachoroidal source of CSF (Rosenberg et a1., 1980). CSF production at extrachoroidal sites may be regulated 130 primarily by the rate of capillary filtration while choroidal CSF production is primarily an active process since acetazolamide effects nearly a 100% cessation of CSF production in the isolated perfused choroid plexus and only a 60% decrease in 6f in 111g_(Pollay, 1975). Whether or not the lateral lacunae are significantly compressed in the I'open skull' preparation (i.e., during ventriculocisternal perfusion) cannot be answered from data presented here, nor has it been addressed in other studies. Future investigations using systems analysis of CSF and sagittal sinus pressure pulse waves during ventriculocister- nal perfusion may provide valuable data regarding the mechanisms underlying transient responses of the CSF system to hypercapnia. The CSF transient responses to hypercapnia, derived from the craniospinal redistribution of CSF and possibly from an episodic burst in 9f, do not depend on the activation of alpha-adrenoceptors. Data in Table 9 indicate that all aspects of the CSF transient phenomena were the same before and after phenoxybenzamine (PBZ) administration. These data indicate that cerebral alpha- adrenoceptors do not influence either the transient behavior of 6f or the craniospinal fluid redistribution process associated with the onset of hypercapnia. Since these phenomena are intimately related to cerebral hemodynamics, it is not surprising that investigations of the cerebrovascular C02 response have yielded similar results. 131 Both surgical and pharmacological treatments to eliminate brain perivascular sympathetic nerves and circulating catecholamines (Dahlgren and Siesjo, 1981; IFAlecy et a1., 1979; Hardebo et a1., 1982; Davis and Sundt, 1980; Waltz et a1., 1971; Lopez de Pablo et al., 1982a; Kawamura et a1., 1974) have left the COz-induced immediate and sustained elevation of cerebral blood flow unaffected. Moreover, other studies report a decreased response of cerebral vessels to adrenergic stimuli (Page and Olmstead, 1975; Lopez de Pablo et al., 1982b; Auer et a1., 1982) as well as to some nonadrenergic agents (Altura and Altura, 1977) during hypercapnia. Apparently hypercapnia can overwhelm neural mechanisms and the influences of some non—neural vasoactive substances such that an elevated Paco2 becomes preeminent in determining cerebral hemodynamics. The cerebrovascular C02 response is not dependent upon an intact cerebral sympathetic innervation. The volume of CSF transiently displaced from the cranium to effect spatial compensation for an expanded intracranial blood volume would therefore not depend upon alpha-adrenoceptor activation. Since both Vd and Cd did not change after PBZ (Table 9) any episodic increase in .Vf is not due to an alpha- adrenergically mediated process. Conversely, the steady state rates of both CSF formation (Vf) and absorption (Va) were influenced by alpha— adrenoceptor activity; Vf and Va declined after intravenous PBZ (Table 8). The decrease in Va may reflect either a 132 decrease in the pressure gradient between the CSF subarachnoid space and the dural venous sinus and/or an increase in resistance to CSF outflow at the arachnoid villi. Since CSF pressure remained constant after PBZ (Table 8), venous dural sinus pressure may have increased causing a decreased hydrostatic gradient favoring CSF absorption. Moreover, an increased cerebral venous pressure without a concomitant increase in CSF pressure could cause Va to decrease via another mechanism. Since cerebral venous pressure is that which surrounds the villi, and CSF pressure is that within these valvular structures (Welch and Friedman, 1960), a decrease in the trans-villus hydrostatic pressure gradient may have occurred to cause a narrowing of the diameter of the villi, much like the passive change in blood vessel diameter effected by a decreased transmural pressure. The decrease in villi diameter in turn, would effect an increase in resistance to absorption and thereby cause a decrease in Va. A mechanism underlying such an increase in sagittal sinus pressure after PBZ has not been proposed. It is known, however, that venous sinus obstruction and conditions such as subarachnoid hemorrhage or expanding cerebral tumors are often accompanied by impaired absorption of CSF (Marmarou et a1., 1975). Unlike Vf and Va, CSF pressure (PCSF) and v.0 remained stable after PBZ (Table 8). Perhaps the decrease in Va, which would tend to increase PCSF and V0, was offset by the 133 concomitant decrease in Vf, which would tend to decrease CSF pressure and V0. The decline in Vf after PBZ may be explained on the basis of at least three different phenomena. The first explanation supposes that the PBZ-induced decrease in MABP (Table 7) caused a significant decrease in blood flow to CSF secretory sites which, in turn, caused a decrease in steady state Vf (Table 8). This hypothesis stems from a study conducted by Davis and Sundt (1980) in which they examined the relationship of cerebral blood flow to cardiac output, mean arterial blood pressure, and blood volume. ‘Using Xenon133 clearance as an indicator of cerebral blood flow, they found that the administration of PBZ (1.5 mg/kg) to normotensive cats caused a 24% and a 22% decrease in MABP and cerebral blood flow, respectively: the majority of their data points fell above a MABP level of 90 Torr, and cerebral blood flow remained stable to changes in MABP induced by intravenous angiotensin. They explained the fall in cerebral blood flow with administration of PBZ by speculating that, "Within the range of autoregulation, cere- bral blood flow is relatively insensitive to changes in MABP; However, increasing or decreasing MABP may still have some effect on cerebral blood flow.“ Eklof and his colleagues (1971) found similar results, reporting a 29% decrease in cerebral blood flow after bilateral chronic cervical ganglionectomy while autoregulation mechanisms 134 remained intact. They failed to speculate on the mechanism responsible for the decline in cerebral blood flow. In the present study, the administration of PBZ may have decreased cerebral blood flow, even though MABP, which was lowered by 30% after PBZ (Table 7), remained within a pressure range over which cerebral blood flow normally' autoregulates (Harper, 1966; MacKenzie et a1., 1976; Page et a1., 1980). Consequently, the proposed decrease in cerebral and perhaps choroidal blood flow (Page et a1., 1980) critically limited the passive transudation of plasma ultrafiltrate across choroidal and/or brain parenchymal capillaries. The delivery of solute and water to metabolic transport mechanisms of CSF secretory sites is crucial for sustaining Vf (Pollay, 1975; Sklar et a1., 1980). The present data (Table 8) may show the effects of an inadequate capillary transudation process on Vf and represent a situa- tion in which cerebral blood flow limits CSF formation rate. The second explanation for the lowered Vf after PBZ is also based on the cerebrovascular response to the alpha blockade but ascribes the decreased Vf to a decreased brain capillary permeability; It is‘well known that a relatively rich network of noradrenergic fibers from peripheral ganglia accompany the extracerebral vessels (i.e., pial vessels) and intracerebral vessels including arterioles and cerebral veins (Edvinsson et a1., 1971/72; mean et a1., 1974); the main receptor involved appears to be alpha-adrenergic, mediating vasoconstriction (Edvinsson and mean, 1974). 135 More recent evidence indicates that central adrenergic neurons, the cell bodies of which are densely assembled in the locus coeruleus, may directly innervate the intraparenchymal microvessels (Hartman et a1., 1972; McCulloch et a1., 1982; Swanson et a1., 1977) and regulate microcirculatory hemodynamics and transcapillary transport mechanisms: cerebral hemispheric blood flow decreases when central noradrenergic cell bodies in the locus coeruleus are stimulated by carbachol (Raichle et a1., 1977), angiotensin II (Grubb and Raichle, 1981), and electrical pulses (Raichle et a1., 1976), and brain vascular permeability concomitantly increases (Raichle et a1., 1977; Preskorn et a1., 1980). These effects can be blocked by the intraventricular (but not systemic) administration of phentolamine (Raichle et a1., 1977) or by bilateral locus coeruleus lesions (Bates et a1., 1977). Since PBZ blocks all alpha-adrenoceptors regardless of their location and origin, it can effect a centrally- mediated decrease in capillary permeability. The lowered capillary'permeability'may decrease capillary filtration rate, thereby imposing a limitation on the availability of plasma ultrafiltrate to choroidal and/or extrachoroidal CSF secretory sites. Vf declines (Table 8) according to this theory because of an inadequate delivery of fluid and electrolytes to secretory sites. The final explanation for the Vf values measured after PBZ (Table 8) rests primarily on the direct effects of PBZ 136 on the active component of CSF synthesis, the metabolic apparatus. Electron microscopic studies reveal that autonomic axon terminals in the choroid plexus have a close association not only with the smooth muscle wall of small arterioles, but also with the secretory epithelial cells (Edvinsson et a1., 1975; Edvinsson et a1., 1977). Further, investigations of the local adrenoceptors mediating the sympathetic nerve-induced effect on the CSF secretory sites reveal that although(alpha-receptor-mediated local vasoconstriction may assist in reducing CSF production by decreasing plexus blood flow (Cserr, 1971; Page et a1., 1981), beta-receptor stimulation can decrease Vf presumably by direct inhibitory actions on the plexus epithelium (Lindvall et al., 1978a; Lindvall et a1., 1978b). Phenoxybenzamine exerts its effects primarily through inhibitory actions on both pre- and post-junctional alpha- adrenoceptors (i.e., alphal and alphaz). Recent evidence, however, disputes the validity of ascribing the net effects of PBZ administration solely to the non-specific blockade of alpha-adrenoceptors. Rather, the non-specificity of the blockade imposed by such antagonists as PBZ and phentolamine may actually cause the expression of sympathomimetic actions (Patel et a1., 1981; Saeed et a1., 1982). More than a decade ago it was shown that the normal vasoconstrictor responses to sympathetic nerve stimulation in the splanchnic bed were inverted to a beta-receptor-mediated vasodilation after PBZ, yet the normal vasoconstrictor response was not 137 changed by prOpranolol, a non-specific beta-adrenoceptor antagonist (Greenway and Stark, 1970). It was suggested that the beta-receptors were not innervated by the sympathetic nerves but, after PBZ, were stimulated by an increased overflow of noradrenaline; PBZ inhibits the re-uptake of the transmitter. Further, it is now known that blockade of prejunctional alpha-adrenoceptors increases the release of noradrenaline from sympathetic nerve terminals by dis-inhibiting transmitter release (Langer, 1977). Saeed et a1. (1982) administered phentolamine systemically to conscious dogs at rest and found an excess of plasma catecholamines, as well as hypotension, vasodilation, tachycardia, and an increase in cardiac output. After subsequent beta-adrenoceptor blockade, however, the vasodilation was not observed. It was concluded that when both alpha—adrenoceptor subtypes are blocked, as they are after PBZ, catecholamine release is enhanced and re-uptake is inhibited, such that the amount of plasma catecholamines increases. The larger amounts of circulating catecholamines allow diffusion in sufficient concentration to stimulate adjacent beta-adrenoceptors (Patel et a1., 1981). In the present studies, PBZ presumably blocked both alpha-receptor subtypes and in the process directly deranged MABP (Table 8), and perhaps changed cerebral blood flow (Davis and Sundt, 1980; Eklof et a1., 1971) and/or the capillary transudation process (Raichle et a1., 1977; Preskorn et a1., 1980). Even more importantly, PBZ may have 138 indirectly caused the activation of beta-adrenoceptors by increasing the level of plasma catecholamines. Since beta- adrenoceptors are associated with both cerebral vessels (Edvinsson and mean, 1974) and choroidal secretory epithelial cells (Lindvall et al., 1978b: Nathanson, 1979), and may be accessible to circulating catecholamines which traverse leaky choroidal capillaries (Wright, 1979), these receptors may have been activated. In turn, and perhaps in conjunction with alpha-mediated cardiovascular responses, beta adrenoceptor stimulation caused a significant decrease in vf (Table 8) by direct inhibition of the metabolic machinery of the secretory epithelium (Lindvall.et a1., 1978a; Lindvall et al., 1978b; Nathanson, 1979). Steady state CSF formation rate (Vf), however, remained insensitive to increases in CSF pressure (Table 2) and/or in Paco2 (Tables 2, 5, and 8). That CSF pressure is without effect on Vf has been demonstrated by Heisey et a1. (1962) in goats, Hochwald and‘Wallenstein (1967) in cats, Eisenberg and co-workers (1974) in hydrocephalic cats, and Sklar et a1. (1980) in dogs. Likewise, data from a number of ventriculocisternal perfusion studies show that increases in Paco2 are without effect on Vf (Oppelt et a1., 1963; Oppelt et a1., 1964; Hochwald et a1., 1973; Tschirgi et a1., 1954; Davson and Segal, 1970). It has been hypothesized that only when cerebral blood flow autoregulatory mechanisms are disrupted does Vf become susceptible to increases in CSF pressure and Paco2 (Sklar et 139 a1., 1980; Weiss and Wertman, 1978; Heisey et a1., 1983). Cerebral autoregulation can be disturbed by the inhalation of a C02 gas-in-air mixture of 9% or more (Raichle and Stone, 1972), by extensive cerebral surgical intervention (Ames et a1., 1965), and by hypotensive episodes in which blood pressure draps below 60 Torr (Harper, 1966; MacKenzie et a1., 1976); the latter condition has been shown to be associated with a depressed Vf (Carey and Vela, 1974; Sklar et a1., 1980). During all phases of every experiment MABP levels remained within a pressure range (Tables 1, 4, and 7) over which cerebral blood flow normally'autoregulates (Harper, 1966; MacKenzie et a1., 1976; Page et a1., 1980). Further, hypercapnic tests were conducted using C02 gas-in-air mixtures that were less than 9%. Vf was not sensitive to either increases in PCSF (Table 2) or Paco (Tables 2, 5, 2 and 8), suggesting that cerebral blood flow autoregulatory capacity was not impaired during these studies. CONCLUSIONS Mixing of tracer is incomplete in CSF spaces during ventriculocisternal perfusion. Craniospinal fluid redistribution significantly contributes to the CSF transient responses to hypercapnia. Overestimates of CSF formation rate will result if measurements are made prematurely after perturbing the craniospinal fluid balance. The craniospinal fluid redistribution process and CSF transient responses are independent of alpha- adrenoceptor activity. CSF rates of formation and absorption are decreased after phenoxybenzamine treatment due to alpha-mediated cerebrovascular responses and/or beta adrenoceptor inhibition of the metabolism of CSF secretory epithelium. Steady state CSF formation is independent of increases in CSF pressure or PaC02° 140 APPENDICES BBV Cd Cd/Ct co/ci CPP Cr CSF CSth Ct APPENDIX A DEFINITION OF TERMS brain blood volume area of co/ci dilution during transient period (min) area of transient dilution of co/ci expressed as a fraction of the total transient area of co/ci if no dilution occurred during the period tic to tfc (dimensionless) perfusion effluent concentration (co) of 14C- dextran expressed as a ratio to perfusion inflow concentration (ci; dimensionless) cerebral perfusion pressure (CPP a MABP - PCSF; Torr) residual co/ci area during transient period (Cr = Ct - Cd: min) cerebrospinal fluid amount of unequilibrated or tracer-free CSF in V6 (:11) total area of co/ci assuming no dilution during t. 1c to tfc (min) respiratory frequency (breaths/min) fraction of C02 in alveolar (end-tidal) gas (% (:02) 141 FIC02 MABP PBZ PCSF Tr Va Vd 142 fraction of inspired C02 (% C02) mean arterial blood pressure (Torr) phenoxybenzamine hydrochloride partial pressure of C02 in arterial blood (Torr) cerebrospinal fluid pressure (cmHZO) (logarithmic) hydrogen ion concentration stop-flow perfusion method time of onset of co/ci transient, and of the end of the co/ci transient, respectively (min) time of onset of C02 breathing and V0 transient, and of the end of the V0 transient, respectively (min) rectal temperature (°C) CSF bulk absorption rate (pl/min) volume of displaced CSF during Vo transient period (ul) minute ventilation (ml/min; BTPS) CSF formation rate (ul/min) perfusion inflow rate (ul/min) perfusion outflow rate (ul/min) tidal volume (ml; BTPS) APPENDIX B PREPARATION OF DIAL-URETHANE ANESTHETIC SOLUTION REAGENTS: l. Diallyl barbituric acid (crystalline; K a K Laboratories, Inc., Plainview, NY) 2. Disodium calcium ethylene diamine tetra acetate trihydrate (Pfaltz & Bauer, Inc., Flushing, NY) 3. Monoethyl urea (Pfaltz & Bauer, Inc., Flushing, NY) 4. Urethane (Aldrich Chemical Co., Milwaukee, WI) METHOD: Add 10 g reagent #1, 50 mg reagent #2, 40 g reagent #3, and 40 g reagent #4 to 10 cc of deionized water. Heat the mixture in a water bath to dissolve the chemicals. Cool to room temperature and store in a stoppered dark glass bottle at room temperature. 143 APPENDIX C ARTERIAL pH AND P002 NEASURENENTS Reference: Radiometer Instruction Manual, Model PHA 927b, Radiometer A/S, Copenhagen, Denmark OPERATING PRINCIPLE: pH Electrode: The electrode (model E5021; London Co., Cleveland, OH) is a H+-sensitive glass capillary surrounded by a glass tube containing a reference fluid. .A test solution is drawn into the glass capillary and contact between the glass electrode, sample and calomel electrode is made through a saturated KCl bridge» The glass electrode develops a potential which varies directly with the H+ concentration in the test sample; the calomel electrode potential is constant. The potential difference between the glass and calomel electrodes is amplified and displayed on a pH meter (model PHM27; The London Co., Cleveland, OH) in mv or pH units. PCO2 Electrode: The PCO2 electrode (model E5036; The London Co., Cleveland, OH), a specialized H+-sensitive glass electrode with a built-in Ag-AgCl electrode, measures the pH of a NaHCO3 + NaCl solution in which it is immersed. 'The 144 145 solution is separated from a test sample by a Teflon membrane which is permeable to C02. The C02 in the test solution or calibration gas diffuses into the buffer solution until a equilibrium is established; the amount of C02 diffusing from the test solution to the buffer will depend on the Pco2 of the sample. The (H+) of the buffer solution changes and the H+ sensitive glass electrode deve10ps a potential proportional to this change which is referenced to the potential of the silver chloride electrode. The potential difference is amplified and displayed on an indicator scale calibrated in mmHg PC02 (model PHM27; The London Co., Cleveland, OH). CALIBRATION: The C02 electrode and the pH electrode are calibrated using two samples of known C02 concentrations (5% and 12% C02; The London Co., Cleveland, OH) and commercially prepared buffer solutions (pH = 6.840 and pH = 7.383; Scientific Products Inc., Allen Park, MI) at 38 1 0.2°C, respectively. SAMPLE HANDLING: Arterial blood samples (nom. lcc) were collected anaerobically in syringes which had their dead space filled with heparinized (1000 U.S.P. units/ml; Heparin Sodium, The Upjohn Company, Kalamazoo, MI) saline “Ll cc heparin per m1 of 0.9% saline solution). The pH and PCO2 were measured* 146 and accepted if duplicate readings were 1 0.01 pH units or 1, 1.0 Torr, respectively. Measurements were corrected for the difference between the electrode and animal's body temperature ('TCORR'; HP67 programmable calculator algorithm) which was measured at the time the blood sample was taken. * Radiometer electrodes were used for some pH and PCO2 meas- urements and for the Operating principles. Other pH and PCO2 measurements were made using a Corning blood gas analyzer (model 165/2; Corning Medical Instruments, Medfield, MA). APPENDIX D RESPIRATORY NEASURENENTS AND CALCULATIONS PRINCIPLES OF THE PNEUMOTACHOGRAPH: The pneumotachograph (Fleisch flowtransducer; Dyna- sciences, Blue Bell, PA) is a cylindrical flow transducer containing a small metal grid which is heated to prevent water condensation, and which offers constant resistance to gas flow through it. A pressure difference develops across the grid (constant resistance) which is proportional to the rate of gas flow through it. A differential manometer (model PT 5A; Grass Instruments, Quincy, MA) is connected by tubing to a port on either side of the grid and transduces the pressure drOp to a prOportional voltage output which is directed through a low level DC preamplifier (model 5P1; Grass Instruments, Quincy, MA) to a strip chart recorder (model 7100 BM; H-P Moseley, Pasadena, CA) calibrated in l/minm Calibration is accomplished by directing known gas flows through the pneumotachograph from a calibrated gas flow meter (tube #604-1A; Matheson Gas Products, East Rutherford, NJ). During experiments expiratory flow rate was monitored and displayed on the chart recorder by attach- ing the pneumotachograph downstream from the respiratory valve assembly. 147 148 The following measurements and calculations were made from the flow rate tracings: Respiratory Frequency Every 10 minutes respiratory frequency (f; breaths/min) was determined by counting the number of expirations occur- ring during the 30 sec period and multiplying it by 2. Tidal Volume At 10 minute intervals during respiratory steady states, the speed of the strip chart (recording expiratory flow rate) was increased from 1.25 cm per min to 2.5 cm per sec and 4 to 5 expiratory flow rate tracings were recorded. These flow vs. time tracings were analyzed using 'TIDVOL'*, a program written for use on the digitizer (model GP-3; Science Accessories Corp», Southport, CT) interfaced with the PDP 11/34 computer (Digital Equipment Corp., Maynard, MA). “TIDVOL' determined the area under the respiratory flow curve (VT; m1; ATPS (ambient temperature and pressure, saturated with water vapor)) by approximating a definite integral using the trapezoidal rule (Eulem's method): where, VT a (0.5 Y0 + Y1 + Y2 + . . . + Yn_1 + 0.5 Yn) dx (D-l) * The 'TIDVOL routine" is also appropriate for approximating the definite integrals in equations 6 and 8 in the Materials and Methods section. 149 and, VT = area under one respiratory flow curve (m1) successive expiratory flow rates Y0, Y1, coo Yn (ml/min) dx = time for one expiration (min); derived from recorder chart speed The mean of 4 - 5 tracings, the standard deviation, and the standard error of the mean (SE) were calculated. If the SE was i 1.0 ml the data were accepted. Minute Ventilation Another PDP 11/34 algorithm (“TIDBITS") converted VT in ATPS units to BTPS (body temperature, ambient pressure, and saturated with water vapor) and to STPD (standard temperature and pressure, dry), and calculated minute ventilation (VB; ml/min in BTPS and STPD), where: VB 3 VT x f (D‘Z) APPENDIX E ANALYSIS OF CARBON DIOXIDE IN ALVEOLAR AND NIX. EXPIRED GAS SAMPLE gELL —ub—— wmoow SAMPLE GAS f? ETECTOR CELL INFRARED A ,z I sounce ' co, MOLECULES / I R BEAM I a BEAM 0 METAL DIAPHRAGM INFRARED1 SOURCE DETECTOR CELL (REFERENCE HALF) T t - ZERO CONTROL L__ . -_—————-——nerensncs CELL FIGURE E-l . Reference: Beckman Instruction Manual, LB-2 Medical Gas Analyzer, Beckman Instruments, Inc., Anaheim, CA 150 151 OPERATING PRINCIPLES: Carbon dioxide, unlike nitrogen or oxygen, absorbs energy from a narrow band of the IR spectrum; the amount of energy absorbed will vary directly with the C02 concentra- tion. The gas analyzer is equipped with two matched IR sources and 3 gas cells: a sample cell (containing the gas to be analyzed); a reference cell (containing no C02); and a detector cell (containing C02). The detector cell is divided by a deformable metal membrane and is located behind the other 2 cells (see Figure E-l). IR radiation passes from the 2 matched sources through the sample and reference cells and into the two halves of the detector cell. When C02 molecules within the closed chambers of the detector cell absorb IR radiation, their kinetic energy and the pressure within the cell proportionally increase. If there is no C02 in the sample cell, the amount of radiation reaching the sample side of the detector cell will equal the amount reaching the reference side, the pressure in each half of the detector cell will be equal, and the membrane will not deform. If the test gas contains C02, the amount of radiation reaching the sample side of the detector cell will cause the dividing membrane to bow toward the sample side changing the capacitance between the metal diaphragm and a fixed plate within the detector cell. The resulting change in voltage is directly related to the percentage of C02 in the test gas sample and is displayed on the console indicator in C02 percentage units. 152 CALIBRATION PROCEDURES: The instrument contains a pump which continuously draws gas through the sample cell at an adjustable rate; the recommended air flow through the cell is 500 ml/min. there are two reasons for this suggested flow rate: first, a particular pressure in the sample cell is associated with a given inflow rate and resistance (inflow resistance is a function of the length and diameter of the inlet tubing); second, the absorption of energy from the sample IR beam is not a linear function of the C02 concentration in the sample. Instead, this relationship is inherently nonlinear and requires a complimentary linearizing circuit. This circuit is preset to correct for nonlinearity only when pressure conditions associated with an inflow rate of 500 ml/min (standard conditions) exist in the sample cell. The device can be operated under nonstandard conditions (flow rate < 500 ml/min) and linearity retained if the pressure in the sample cell under standard conditions is duplicated: sample cell pressure will be unchanged if inflow rate and resistance are concurrently and proportionately decreased and increased, respectively; In these experiments it was necessary to operate the instrument under nonstandard conditions (inflow rate = nom. 200 ml/min) since an anesthetized cat breathing air may have a minute ventilation less than 500 ml/min. A.two-stage calibration procedure was devised to maintain linearity under measuring (nonstandard) conditions. 153 Standard Conditions (flow rate = 500 ml/min) 1. Depress POWER and OPERATE buttons and allow the instrument to equilibrate for at least 30 minutes prior to calibration. 2. Adjust ZERO control such that the panel meter reads +0.03% C02 when ambient air is drawn through the heated inlet tubing and the sample cell. 3. Sample from a Haldane analyzed C02 gas mixture (> 8% C02; < 10% C02). Adjust the GAIN control so that the panel meter reads the calibration gas concentra- tion. 4. Sample from another analyzed C02 mixture (between 0% and 8% 002). The panel meter will read this concen- tration if the response is linear. Nonstandard Conditions (flow rate < 500 ml/min) 1. Attach a 23 ga needle onto the heated inlet tubing to increase inflow resistance. 2. Sample from same calibration gas as in A-3 and adjust FLOW (decrease inflow rate) until the panel meter records the same value as in A-3. 3. Repeat A-4. APPENDIX P COMPOSITION AND PREPARATION OF ARTIFICIAL CAT CEREBROSPINAL FLUID (CS?) ELECTROLYTE COMPOSITION OF CSF (from Vogh and Maren, Am. J. Physiol. 228: 673-683, 1975): Substance mull Na+ 151.0 x+ 2.5 Ca++ 2.5 Mg++ 1.0 Cl' 138.0 Hco3' 20.0 apo4' 1.0 304' 1.0 REAGENTS USED IN PREPARING ARTIFICIAL CAT CSF: 1. NaCl, Analytic reagent (A.R.) 2. NaHCO3, (A.R.) 3. NaH2P04'H20, (A.R.) 4. NaZHPO4'H20, (A.R.) 5. KCl, (A.R.) 6. MgSO4'7H20, (A.R.) 7. CaC12, (A.R.) 154 155 SOLUTIONS: A. Dissolve 7.6 g of reagent #1, 1.68 g of reagent #2, 169.0 mg of reagent #3, 134.0 mg of reagent #4, and 186.4 mg. of reagent #5 in deionized water; q.s. to one liter. B. Dissolve 24.65 g of reagent #6 in deionized water; q.s. to 100 ml. C. Dissolve 27.75 g of reagent #7 in deionized water; q.s. to 100 m1. METHOD: Fifty m1 of solution A is bubbled at room temperature (nom. 15 minutes) with 5% COz-95% 02 gas mixture to adjust the pH to approximately 7.40. Then 50 ul each of solutions B and C are added to 50 ml of solution A to establish the normal ionic composition of cat CSF. APPENDIX C DETERNINATION OP INPLOW AND OUTPLOW PERPUSION RATES The inflow and outflow drop volumes were determined gravimetrically'by collecting a known number of drOps of fluid in tared vials from PE tubing connected to the inflow syringe (inflow) or to the cisternal cannula (outflow). Assuming a specific gravity of 1.0, the volume per drOp (V; pl) could be determined by: V = (G-l) 5 IS where: W = net weight of fluid (mg) n = number of drops During an experiment the cisternal outflow cannula was directed through a photocell drop detector attached to a fraction collector controller (model 568; Instrumentation Specialties Co., Lincoln, NA). The controller was programmed to advance the collector after a pre-set number of drOps were detected; normally 2 drop samples (nom. 100 pl) were collected. Further, the controller circuit was modified to provide a voltage input to the polygraph (model 5P; Grass Instrument Co., Quincy, MA) triggering an event 156 157 mark at the end of each sample. The polygraph record was analyzed using a digitizer (model GP-3; Science Accessories Corp”, Maynard, MA) for measuring distance between successive outflow sample events. A computer program ('OUTFLOW') was used to calculate outflow rate by: {, ._. (33.3331 (0-2) D where, V = flow rate (ul/min) V = volume per drop (pl; see equation G-l) S a polygraph chart speed (mm/min) D a distance between event marks on polygraph record (mm) At the completion of each experiment, the outflow cannula was disconnected from the drOp detector and replaced by the inflow cannulae for determining inflow rate using equations 1 and 2. APPENDIX E BETA RADIATION COUNTING Reference: Beckman Liquid Scintillation Counter Manual, Model 3150P, Beckman Instruments Inc., Fullerton, CA. OPERATING PRINCIPLES: Liquid scintillation (LS) counting is a method for detecting beta particle emission from decaying nuclei of unstable isotopes. When a radioisotope such as Carbon-l4 emits beta particles and is suspended in a photoluminescent medium (iJL, LS cocktail; a mixture of organic solvent and fluor molecules), the medium absorbs energy from the beta particles and releases it in the form of light. The LS counter houses special vacuum tubes (photomultiplier tubes; PMT) which are light-sensitive and which produce voltage pulses in response to photons that strike their surfaces; the number of voltage pulses produced is proportional to the number of fluorescent events detected. These voltage pulses are summed, amplified, directed to a sealer, and recorded as “counts.“ Counting rate (counts per min; CPM) is based on the total number of counts accumulated in a one-minute time period. 158 159 The liquid scintillation counting technique relys on critical reactions between the radioactive sample and the light-producing chemical to indicate the presence of a radioisotoPe. Anything which promotes the interaction of the emitted particle with the fluorescing chemical will enhance the efficiency of detection; anything which interferes with it will reduce the photon output of the scintillation fluor mixture and the detection efficiency. This interference process is known as I'quenching" and can be caused by any of the following agents or conditions: color in the sample; the presence of salts in the medium which do not fluoresce; incomplete dispersion of the radioactive sample in the fluorescing mixture; or inadequate concentration of fluor or solvent in the LS cocktail. Quenching reduces the final count (“counting efficiency;' EFF) of a radioactive sample by shifting the isotOpes light emission spectrum toward lower pulse heights. The external standard channels ratio (ESCR) method, based on the effects of gamma radiation on a 14C-quenched standard set (Amersham/Searle Corp», Boston, MA), was used to detect quenching in the samples and to provide a means of correcting for it. The LS counter is equipped with a gamma radiation source (137-cesium) which is brought in proximity to a series of sealed standards containing equal amounts of radioactive matter (carbon-l4) but with varying amounts of chemical quencher. The gamma rays interact with matter and discharge compton electrons which interact with solvent 160 molecules in the LS cocktail in the same way as beta particles to produce light scintillations. .Additionallyy compton and beta emission spectra exhibit similar quench- induced shifts. Each standard is analyzed in two channels which count over two different portions of the energy spectrum both before and after exposure to gamma radiation (to determine sample counts due to beta emission, and total counts due to beta compton emission, respectively). The counts from each channel due to gamma radiation (the difference between total and sample counts) are determined and expressed as a ratio (ESCR) which is indicative of the amount of shift in the compton spectrum and inversely proportional to the amount of quenching in the sample. Counting efficiency (EFF) of the standards was determined as: (CPM - background CPM) x 100 DPM EFF = (H-l) where, CPM a counting rate due to beta emission (counts/min) DPM = known activity of 14C-standards (disintegrations/ min) A quench correction curve is constructed by plotting EFF for the standards as a function of their corresponding ESCR values. The curve can be described as: Y = mX + b (H-2) 161 where, Y = counting efficiency (EFF) for the standards x = ESCR value for the standards EFF (standard #1) - EFF (standard #2) ESCR (standard #1) - ESCR (standard #2) b = counting efficiency for the standards when ESCR = 0 Each sample undergoes the same counting procedure as described for the 14C-standards, and their CPM and ESCR values determined. From the quench correction curve or from the mathematical expression describing it (equation H-2), sample counting efficiencies (EFF) are determined. Finally, the corrected activity of each sample (DPM; disintegrations/min) is calculated as: sample CPM DPM == sample EFF (H-3) CALCULATIONS or 14c-DEXTRAN CONCENTRATIONS: The disintegration rates (DPM) and concentrations (DPM/ul) for 14C-dextran in samples from the inflow syringe and cisternal effluent were calculated using the PDP 11/34 computer (Digital Equipment Corp., Maynard, MA). A paper tape punch (Epson 611; Shinshu Seiki Co., Ltd., Japan) attached to the LS Counter (model 3150P; Beckman Instruments Inc., Fullerton, CA) provided an output which contained ESCR and counts per min (CPM) for quenched 14C-standards, background, perfusion inflow (Ci) and cisternal effluent (co) samples. These data were entered into the computer 162 through a program called 'SDPMl' using a paper tape reader, as were inflow and outflow sample volumes (ul) which were manually entered. A second program ('SDPMZ') using the output data from 'SDPMl,‘ calculated the following information for each standard and sample: the corrected activity (DPM) (see equations H-l to H-3); the 14C-dextran inflow and outflow concentrations (Ci and co, respectively; DPM/pl); and a ratio of outflow to inflow 14C-dextran concentration (co/Ci; dimensionless). .A final program, 'SDPM3,‘ a versatile plotting routine, accepted the stored output from 'SDPM2,‘ combined these data with the manuallyb entered collection time and outflow rate (V0; Appendix G) data for each sample, and plotted CO/Ci and V0 as a function of time. BIBLIOGRAPHY BIBLIOGRAPHY Adams, T., S. R. Heisey, M. C. Smith, M. A. Steinmetz, J. C. Hartman, and H. K. Fry. Thermodynamic technique for the quantification of regional blood flow. Am. J. Physiol. 238: 682-696, 1980. Altura, B. M. and B. T. Altura. Vascular smooth muscle and neurohypophyseal hormones. Fed. Proc. 36: 1853-1860, 1977. Ames, A., M. Sakanoue, and S. Endo. Na, K, Ca, Mg, and Cl concentrations in Choroid plexus fluid and cisternal fluid compared with plasm ultrafiltrate. {L Neuro— physiol. 27: 672-681, 1964. Ames, A. III, R. Higashi, and F. Nesbitt. Effects of PC02, acetazolamide and ouabain on volume and composition of Choroid-plexus fluid. Am. J. Physiol. 229: 415-419, 1965. Auer, L. M. and B. B. Johansson. Pial venous constriction during cervical sympathetic stimulation in the cat. Acta. Physiol. Scand. 110: 203-205, 1980. Auer, L., B. Johansson, and E. T. MacKenzie. Cerebral venous pressure during actively induced hypertension and hypercapnia in cats. Stroke 11: 180-183, 1980. Auer, L. M., W. Kuschinsky, B. B. Johansson, and L. Edvinsson. Sympatho-adrenergic influence on pial veins and arteries in the cat. In: Cerebral Blood Flow: Effects of Nerves and Neurotransmitters. Heistad, D. IL, and M. L. Marcus (edsJ. Elsevier North Holland, Inc. pp. 292-300, 1982. Avezaat, C. J. J., D. H. M. Eijndhoven, and D. J. Wyper. Effects of hypercapnia and arterial hypotension and hypertension on cerebrospinal fluid pulse pressure and intracranial volume-pressure relationships. CL Neurol. Neurosurg. and Psychiat. 43: 222-234, 1980. 163 164 Bates, D., R. M. Weinshiboum, R. J. Campbell, and T. M. Sundt Jr. The effect of lesions in the locus coeruleus on the physiological responses of the crebral blood vessels in cats. Brain Res. 136: 431-443, 1977. Bering, E.1L Cerebrospinal fluid production and its rela- tionship to cerebral metabolism and cerebral blood flow. Am. J. Physiol. 197: 825-828, 1959. Bering, E. A. and O. Sato. Hydrocephalus: changes in formation and absorption of cerebrospinal fluid within the cerebral ventricles. J. Neurosurg. 20: 1050-1063, 1963. Carey, M. E. and A. R. Vela. Effect of systemic arterial hypotension on the rate of CSF formation in dogs. J. Neurosurg. 41: 350-355, 1974. Cervos-Navarro, J. and F. Matakas. Electron microsc0pic evidence for innervation of intracerebral arterioles in the cat. Neurology 23: 282-286, 1973. Cserr, H. Physiology of the Choroid plexus. Physiol. Rev. 51: 273-311, 1971. Curl, F. D. and M. Pollay. Transport of water and electro- lytes between brain and ventricular fluid in the rab- bit. Exp. Neurol. 20: 558-574, 1968. D'Alecy, L. G., C. J. Rose, and S. A. Sellers. Sympathetic modulation of hypercapnic cerebral vasodilation in dogs. Circ. Res. 45: 771-775, 1979. Dahlgren, N. and B. K. Siesjo. Cerebral blood flow and oxygen consumption in normocapnia and hypercapnia: modulation influence of paravertebral sympathetic blockade at the low thoracic level. Acta. Anesth. Scand. 45: 331-336, 1981. Davis, D. H. and T. M. Sundt, Jr. Relationship of cerebral blood flow to cardiac output, mean arterial pressure, blood volume, and alpha and beta blockade in cats. J. Neurosurg. 52: 745-754, 1980. Davson, H. Physiology of the Ocular and Cerebrospinal Fluids. Little and Brown, Boston, 1956. 165 Davson, H. and C. P. Luck. The effect of acetazolamide on the chemical composition of the aqueous humor and cerebrospinal fluid of some mammalian species and on the rate of turnover of 24Na in these fluids. J. Physiol. 137: 279-293, 1957. Davson, H. and E. Spaziani. Effect of hypothermia on certain aspects of the cerebrospinal fluid. Exp. Neurol. 6: 118-128, 1962. Davson, H. Physiology of the Cerebrospinal Fluid. J. and A. Churchill Ltd., London, 1967. Davson, H., G. Hollington, and M. B. Segal. The mechanism of drainage of the cerebrospinal fluid. Brain 93: 665- 678, 1970. Davson, H. and M. B. Segal. The effects of some inhibitors and accelerators of sodium transport on the turnover of 2Na in the cerebrospinal fluid and the brain. J. Physiol. 209: 131-153, 1970. Deane, R. and M. Segal. The effect of vascular perfusion of the Choroid plexus on the secretion of cerebrospinal fluid. Proc. Physiol. Soc. 285: 18-19, 1978. Diamond, J. M. and W. H. Bossert. Standing-gradient osmotic flow. A mechanism for coupling of water and solute transport in epithelia. J. Gen. Physiol. 50: 2061- 2083, 1967. DiMattio, J}, G. M. Hochwald, and C. Malhan. The effects Of the hydrocephalic process on cerebral blood flow in the cat. In: Cerebral Circulation and Metabolism. Langfitt, T. W., L. C. McHenry, Jr., M. Reivich, and H. Wollman, (eds.). Springer-Verlag, New York. pp. 223- 227,1975. Duckles, S. P. and J. A. Bevan. Pharmacological characteri- zation of adrenergic receptors of rabbit cerebral artery in vitro. J. Pharmacol. Exp. Ther. 197: 371- 378, 1976. Edvinsson, L., R. Hakanson, Ch. mean, and K. A. West. Sympathetic influence on carbonic anhydrase activity in choroid plexus. Exp. Cell Res. 67: 245, 1971. 166 Edvinsson, L., Ch. mean, E. Rosengren, and K. A. West. Concentration of noradrenaline in pial vessels, choroid plexus, and iris during 2 weeks after sympathetic ganglionectomy or decentralization. Acta Physiol. Scand. 85: 201-206, 1972. Edvinsson, L., K. C. Nielsen, Ch. mean, and K. A. West. Sympathetic adrenergic influence on brain vessels as studied by change in cerebral blood volume of mice. Europ. Neurol. 6: 193-202, 1971-72. Edvinsson, L. and Ch. mean. Pharmacological characteriza- tion of adrenergic alpha and beta receptors mediating the vasomotor reSponses of cerebral arteries in vitro. Circ. Res. 35: 835-849, 1974. Edvinsson, L., K. C. Nielsen, Ch. mean, and K. A. West. Adrenergic innervation of the mammalian choroid plexus. Amer. J. Anat. 139: 299-308, 1974. Edvinsson, L. Neurogenic mechanism in the cerebrovascular bed. Autonomic nerves, amine receptors and their effects on cerebral blood flow. .Acta Physiol. Scand. 427: 1-36, 1975. Edvinsson, L., P. Aubineau, Ch. mean, R. Sercombe, and J. Seylaz. Sympathetic innervation of cerebral arteries: Prejunctional supersensitivity to noradrenaline after sympathectomy or cocaine treatment. Stroke 6: 525- 530, 1975a. Edvinsson. In, R. Hakanson, M. Lindvall, Ch. mean, and K. (L Svenson. Ultrastructural and biochemical evidence for a sympathetic neural influence on the choroid plexus. Exp. Neurol. 48: 241-251, 1975b. Edvinsson, L. and E. T. MacKenzie. Amine mechanism in the cerebral circulation. Pharmacol. Rev. 28: 275-347, 1977. Edvinsson, L. and M. Lindvall. Autonomic vascular innervation and vasomotor reactivity in the choroid plexus. Exp. Neurol. 62: 394-404, 1978. Eklof, B., D. H. Ingvar, E. Kagstron, and T. Olin. Persist- ence'of'cerebral blood flow autoregulation following chronic bilateral cervical sympathectomy in the monkey. Acta Physiol. Scand. 82: 172-176, 1971. 167 Eisenberg, H. M., J. E. McLennan, and J. Welch. Ventricular perfusion in cats with kaolin-induced hydrocephalus. J. Neurosurg. 41: 20-28, 1974. Epstein, M. H., A. M. Feldman, and S. W. Bradbury. Science 196: 1012-1977. Exton, J. Molecular mechanisms involved in alpha-adrenergic responses. Trends Pharm. Sciences 3: 111-115, 1982. Fenstermacher, J., C. Li, and V. Levin. Extracellular space of the cerebral cortex of normothermic and hypothermic cats. Exp. Neurol. 27: 101-114, 1969. Fisher, M. J., S. R. Heisey, and T. Adams. Cerebrospinal fluid transients induced by hypercapnia. Am. J. Physiol. 245: 701-705, 1983. Goldstein, D. A., M. Romoff, E. Bogin, and S. G. Massry. Relationship between the concentrations of calcium and phosphorous in blood and cerebrospinal fluid. J. Clin. Endocr. Metab. 49: 58-62, 1979. Goodman, L. S. and A. Gilman. The Pharmacological Basis of Therapeutics. Macmillan Publishing Co., Inc., New York, 1975. Greenway, C. V. and R. D. Starke. The vascular responses of the spleen to intravenous infusion of cats, angio- tensin, and vasopressin in the anesthetized cat. Br. J. Pharmacol. 38: 583-592, 1970. Grubb, R. L. and M. E. Raichle, Intravenous Angiotensin II increases brain vascular permeability. Brain Res. 210: 426-430, 1981. Grubb, R. L., M. E. Raichle, and J. O. Eichling. The ef- fects of changes in PaCOZ on cerebral blood volume, blood flow, and vascular mean transit time. .Stroke 5: 630-639, 1974. Hamer, A. and A. Sahar. Absence of formation of cerebro- spinal fluid in the spinal subarachnoid space of the cat. Israel J. Med. Sci. 14: 741-744, 1978. Hammerstad, J. P., A. V. Lorenzo, and R. W. P. Cutler. Iodide transport from the spinal subarachnoid fluid in the cat. Am. J. Physiol. 216: 353-358, 1969. 168 Hardebo, J. E., L. Edvinsson, and Ch. mean. Influence of the cerebrovascular sympathetic innervation on blood- brain barrier function. Acta. Physiol. Scand. 452: 65-68, 1977a. Hardebo, J. E., L. Edvinsson, P. C. Emson, and Ch. mean. Isolated brain microvessels: enzymes related to adrenergic and cholinergic functions. In: Neurogenic Control of the Brain Circulation. mean, C. and L. Edvinsson, (eds.). Pergamon Press, oxford. pp. 105- 113, 1977b. Hardebo, J. E., P. C. Emson, B. Falk, Ch. mean, and E. Rosengren. Enzymes related to monoamine metabolism in brain microvessels. J. Neurochem. 35: 1388-1393, 1980. Hardebo, J. E. and Ch. mean. Barrier mechanisms for neurotransmitter monamines at the blood brain interface. Ann. Neurol. 9: 1-6, 1980. Hardebo, J. E., O. Lindvall, and B. Nilsson. On the possible influence of adrenergic and cholinergic mechanisms in normo- and hypercapnia. In: Cerebral Blood Flow: Effects Of Nerves and Neurotransmitters. Heistad, D. D. and M. L. Marcus (eds.). pp. 378-383. Harper, A. M. Autoregulation of cerebral blood flow: influence of the arterial blood pressure on the blood flow through the cerebral cortex. J. Neurol. Neuro- surg. Psychiat. 29: 398-403, 1966. Hartman, B. K., D. Zide, and S. Udenfriend. The use of dopamine beta-hydroxylase as a marker for the central noradrenergic nervous system in rat brain. Proc. Nat. Acad. Sci. 69: 2722-2726, 1972. Haywood, J. R. and B. P. Vogh. Some measurements of autonomic nervous system influences on production of cerebrospinal fluid in the cat. J. Pharmacol. Exp. Ther. 208: 341-346, 1978. Heisey, S. R., D. Held, and J. R. Pappenheimer. Bulk flow and diffusion in the cerebrospinal fluid system of the goat. Am. J. Physiol. 291: 1564-1567, 1962. 169 Heisey, S. R., T. Adams, M. J. Fischer, and W. Dang. Effect of hypercapnia and cerebral perfusion pressure on cerebrospinal fluid production in the cat. Am. J. Physiol. 244: 224-227, 1983. Hernandez, M. J., M. E. Raichle, and H. L. Stone. The role of the sympathetic nervous system in cerebral blood flow autoregulation. Europ. Neurol. 6: 175-179, 1971- 72. Hernandez-Perez, M. J., M. E. Raichle, and H. L. Stone. The role of the peripheral sympathetic nervous system in cerebral blood flow autoregulation. Stroke 6: 284- 292, 1975. Hochwald, G. M. and Wallenstein. Exchange of albumin be- tween blood, cerebrospinal fluid and brain in the cat. Am. J. Physiol. 212: 1199-1204, 1967. Hochwald, G. M. and A. Sahar. Effect of spinal fluid pressure on cerebrospinal fluid formation. Exp. Neurol. 32: 30-40, 1971. Hochwald, G. M., C. Malhan, and J. Brown. Effect of hypercapnia on CSF turnover and blood-CSF barrier to protein. Arch. Neurol. 28: 150-155, 1973. Katzman, R. and H. M. Pappius. Brain Electrolytes and Fluid Metabolism. Williams and Wilkins Co., Baltimore, 1973. Kawamura, J., M. Rennels, and E. Nelson. The innervation of intracerebral arteries of the human brain: an electron microscopic study. AMNA 15-17, 1972. KnOpp, L. M., J. R. Atkinson, and A. A. Ward, Jr. Effect of Diamox on cerebrospinal fluid pressure of cat and monkey. Neurol. 7: 119-123, 1957. Langer, S. 2. Presynaptic receptors and their role in the regulation of transmitter release. Br. J. Pharmacol. 60: 481-497, 1977. Lindvall, M., L. Edvinsson, and Ch. mean. Histochemical, ultrastructural, and functional evidence for a neurogenic control of cerebrospinal fluid production. Acta Physiol. Scand. Suppl. 452: 77-86, 1977. 170 Lindvall. 1L, L. Edvinsson, and Ch. mean. Reduced cerebro- spinal fluid formation through cholinergic mechanisms. Neuroscience Letters 10: 311-316, 1978a. Lindvall, M., L. Edvinsson, and Ch. mean. Sympathetic nervous control of cerebrospinal fluid production from the choroid plexus. Science 201: 1767-1768, 1978b. Lindvall. IL, L. Edvinsson, and Ch. mean. Effect of sym- pathomimetic drugs and corresponding receptor antago- nists on the rate of cerebrospinal fluid production. Exp. Neurol. 64: 132-145, 1979. Lindvall, M. and Ch. mean. Evidence for the presence of two types of monoamine oxidase in rabbit choroid plexus and their role in breakdown of amines influencing cere- brospinal fluid formation. J. Neurochem. 34: 518-522, 1980. Lindvall, M. and Ch. mean. Automonic nerves in the mammalian choroid plexus and their influences on the formation of cerebrospinal fluid. CL Cereb. Blood Flow and Metab., 1981. Lindvall, M., Ch. mean, and B. Winbladh. Effect of sympathetic denervation on various aspects of transport function in the mammalia Choroid plexus. In: Cerebral Blood Flow: Effects of Nerves and Neurotransmitters. Heistad, D. D., and M. L. Marcus, (eds.). Elsevier North Holland, Inc. pp. 393-401, 1982. Lofgren, J. Effects of variations in arterial pressure and PaCO2 tensions on the cerebrospinal fluid pressure- volume relationships. .Acta Neurol. Scandinav. 49: 586-598, 1973. LOpez de Pablo, A. L., M. C. Gonzalez, G. Dieguez, B. Gomez, and S. Lluch. Reduction of cerebrovascular reactivity during hypercapnia. Am. J. Physiol. 242: 441-446, 1982a. Lopez de Pablo, A. L., M. C. Gonzalez, G. Dieguez, B. Gomez, and S. Lluch. Cerebrovascular responses to C02 after inhibition of sympathetic activity. J. Appl. Physiol.: Respirat. Environ. Exercise Physiol. 53: 873-878, 1982b. 171 Lux, W. and J. Fenstermacher. Cerebrospinal fluid formation in ventricles and spinal subarachnoid space of the rhesus monkey. J. Neurosurg. 42: 674-678, 1975. MacKenzie, E. T., S. Strandgaard, D. I. Graham, J. V. Jones, A. M. Hargper, and J. K. Farrar. Effects of acutely induced hypertension in cats on pial arteriolar cali- ber, local cerebral blood flow, and the blood-brain barrier. Circ. Res. 39: 33-41, 1976. Mann, D. D., A. B. Butler, R. N. Johnson, and N. H. Bass. Clearance of macromolecular and particulate substances from the cerebrospinal fluid system of the rat. J. Neurosurg. 50: 343-348, 1979. Maren, T. Bicarbonate formation in cerebrospinal fluid: role in sodium transport and pH regulation. Am. J. Physiol. 222: 885-899, 1972. Marin, J. and F. Rivilla. Nerve endings and pharmacological receptors in cerebral vessels. Gen. Pharmacol. 13: 361-368, 1982. Marmarou, A., K. Shulman, and J. LaMogese. Compartmental analysis of compliance and outflow resistnace of the cerebrospinal fluid system. J. Neurosurg. 48: 332- 344, 1978. Martins, A. N., J. K. Wiley, and P. W. Myers. Dynamics of the cerebrospinal fluid and the spinal dura mater. J. Neurol. Neurosurg. and Psychiat. 35: 468-473, 1972. Martins, A. N., T. Doyle, and N. Newby. PC02 and rate of formation of cerebrospinal fluid in monkey. Am. J. Physiol. 231: 127-131, 1976. Martins, A. N., N. Newby, and T. F. Doyle. Sources of error in measuring cerebrospinal fluid formation by ventriculocisternal perfusion. J. Neruol. Neurosurg. and Psychiat. 40: 645-650, 1977. Masuoka, D. and E. Hansson. Autoradiographic distribution studies of adrenergic blocking agents. I. 14C- phenoxybenzamine (Bensylyt NFN), an alpha-receptor-type blocking agent. Acta Pharmacol. Toxicol. 25: 113-122, 1967. 172 Maxwell, Pu S. and D. C. Pease. The electron microscopy of the choroid plexus. J. Biophys. Biochem. Cytol. 2: 467-474, 1956. McCulloch, J., H. E. Savake, and W. Angerson. Regional water permeability in CNS of conscious rats: effects of hypercapnia and locus coeruleus lesions. In: Cerebral Blood Flow: Effects of Nerves and Neurotransmitters. Heistad, D. D. and M. L. Marcus (eds.). Elsevier North Holland, Inc. pp. 509-516, 1982. Milhorat, T. H., M. K. Hammock, J. D. Fenstermacher, D. P. Rall, and V: A. Levin. Cerebrospinal fluid production by the choroid plexus and brain. Science 173: 330- 332, 1971. Millen, J. W. and G. E. Rogers. An electron microsCOpic study of the choroid plexus in the rabbit. J. Biophys. Biochem. Cytol. 2: 467-474, 1956. Nakamura, S. and T. H. Milhorat. Nerve endings in the choroid plexus of the fourth ventricle of the rat: electronic microsc0pic study. Brain Res. 153: 285- 293, 1978. Nakamura, S. and G. M. Hochwald. Effects of arterial PC02 and cerebrospinal fluid volume flow rate changes on choroid plexus and cerebral blood flow in normal and experimental hydrocephalic cats. {L Cerebral Blood Flow and Metab. 3: 369-375, 1983. Nathanson, J.IL B-Adrenergic-sensitive adenylate cyclase in secretory cells of the choroid polexus. Science 204: 843-844, 1979. Nelson, E. and M. Rennels. Innervation of intracranial arteries. Brain 93: 475-490, 1970. Nielsen, K. C., Ch. mean, and B. Sporrong. Ultrastructure of the autonomic innervation apparatus in the main pial arteries of rats and cats. Brain Research 27: 25-32, 1971. Oldendorf, W. The Nervous System. In: The Basic Neuro- sciences. 1. Raven Press, New York. pp. 279-289, 1975. 173 Oppelt, W. W., T. Maren, E. S. Owens, and D. P. Rall. Effects of acid-base alterations on cerebrospinal fluid formation. Proc. Soc. Exp. Biol. Med. 114: 86- 89, 1963. Oppelt, W. W., C. S. Patlak, and D. P. Rall. Effect of certain drugs on cerebrospinal fluid production in the dog. Am. J. Physiol. 206: 247-250, 1964. mean, C., L. Edvinsson, and K. C. Nielsen. Autonomic neuroceptor mechanisms in brain vessels. Blood Vessels 11: 2-31, 1974. mean, C. and L. Edvinsson. Histochemical and pharmacologi- cal approach to the investigation of neurotransmitters, with particular_regard to the cerebrovascular bed. In: Neurogenic Control of Brain Circulation mean, C. and L. Edvinsson (edsJ. Pergamon Press, Oxford. pp. 15- 38, 1977. mean, C., L. Edvinsson, and J. E. Hardebo. Pharmacological in 113:9 analysis of amine-mediated vasomotor functions in the intracranial and extracranial vascular beds. Blood Vessels. 15: 128-147, 1978. mean, C. and J. E. Hardebo. Functional aspects of the blood brain barrier with particular regard to effects of Circulating vasoactive:neurotransmitters. In: Cere- bral Blood Flow: Effects of Nerves and Neurotransmit- ters. Heistad, D. D. and M. L. Marcus (eds.). pp. 119-128, 1982. Page, I. H. and F. Olmsted. Influence of respiratory gas mixtures on arterial pressure and vascular reactivity in normal and hypertensive dogs. Circulation 3: 801- 819, 1975. Page, R. H., D. J. Funsch, R. W. Brennan, and M. J. Hernandez. Choroid plexus blood flow in the sheep. Brain Res. 197: 532-537, 1980. Pappenheimer, J. R., S. R. Heisey, E. F. Jordan, and J. Downer. Perfusion of the cerebral ventricular system in unanesthetized goats. Am. J. Physiol. 203: 763- 774, 1962. 174 Patel, P., D. Bose, and C. Grenway. Effects of prazosin and phenoxybenzamine andd- andp-receptor-mediated re- sponses in intestinal resistance and capacitance ves- sels. J. Cardiovasc. Pharmacol. 3: 1050-1059, 1981. Pollay, M. and K. Welch. The function and structure of canine arachnoid villi. J. Surg. Res. 2: 307-311, 1962. Pollay, M. and H. Davson. The passage of certain substances out of the cerebrospinal fluid. Brain 86: 137-150, 1963. Pollay, M. and F. Curl. Secretion of cerebrospinal fluid by the ventricular ependyma of the rabbit. Am. J. Physiol. 213: 1031-1038, 1967. Pollay, M., A. Stevens, E. Estrada, and R. Kaplan. Extra- corporeal perfusion of choroid plexus. J. Appl. Physiol. 32: 612-617, 1972. Pollay, M. Formation of cerebrospinal fluid. J. Neurosurg. 42: 665-673, 1975. Portnoy, H. D., M. ChoPP: C. Branch, and M. B. Shannon. Cerebrospinal fuid pulse wave form as an indicator of cerebral autoregulation. J. Neurosurg. 56: 666-678, 1982. Preskorn, S. H., B. K. Hartman, M. E. Raichle, L. W. Swanson, and H. B. Clark. Central adrenergic regula- tion of cerebral microvascular permeability and blood flow: pharmacological evidence. Adv. Exp. Biol. Med. 131: 127-138, 1978. Preskorn, S. H., G. H. Irwin, S. Simpson, D. Firesen, J. Rinne, and G. Jerovich. Medical therapies for mood disorders alter the blood-brain barrier. Science 213: 469-471, 1981. Preskorn, S. H., B. K. Hartman, G. H. Irwin, and C. W. Highes. Role of the central adrenergic system in medi- ating Amitriptyline-induced alteration in the mammalian blood-brain barrier in vivo. J. Pharm. Exp. Ther. 223: 388-395, 1982. 175 Raichle, M. E. and H. L. Stone. Cerebral blood flow autoregulation and graded hypercapnia. In: Cerebral Blood Flow and Intracranial Pressure. I Europ. Neurol. 6: 1-5, 1971-72. Raichle, M. E., B. K. Hartman, J. O. Eichling, and L. G. Sharpe. Central noradrenergic regulation of cerebral blood flow and vascular permeability. Proc. Nat. Acad. Sci. 72: 3726-3730, 1975. Raichle, M. E., J. O. Eichling, M. G. Straatman, M. J. Welch, K. B. Larson, and M. M. TerPogossian. Blood- brain barrier permeability of 11C-labeled alcoholics and ISO-labeled water. Am. J. Physiol. 230: 543-552, 1976. Raichle, M. E., R. L. Grubb, and J. O. Eichling. Neural and hormonal regulation of brain water permeability. In: Neurogenic Control of Brain Circulation. mean, C. and L. Edvinsson (edsJ. Pergamon Press, Oxford. pp. 465- 470, 1977. Reed, D. J. and M. H. Yen. The role of the cat choroid plexus in regulating cerebrospinal fluid magnesium. J. Physiol. 281: 477-485, 1978. Reivich, M. Arterial PC02 and cerebral hemodynamics. Am. J. Physiol. 206: 25-35, 1964. Rennels, M. L. and E. Nelson. Capillary innervation in the mammalian central nervous system: an electron micro- scopic demonstration. Amer. J. Anat. 144: 233-241, 1975. Rich, M., P. Scheinberg, and M. S. Belle. Relationship between cerebrospinal fluid pressure and cerebral blood pressure. Circ. Res. 1: 389-395, 1953. Rosenberg, G., L. Wolfson, and R. Katzman. Pressure- dependent bulk flow of cerebrospinal fluid into brain. Exp. Neurol. 60: 267-276, 1978. Rosenberg, G. A., W. T. Kyner, and E. Estrada. Bulk flow of brain interstitial fluid under normal and hyperosmolar conditions. Amer. J. Physiol. 238: 42-49, 1980. 176 Rosenberg, G., W. Kyner, and E. Estrada. The effect of increased CSF pressure on interstitial fluid flow during ventriculocisternal perfusion in the cat. Brain Res. 232: 141-150, 1982. de Rougemont, J., A. Ames III, F. B. Nesbitt, and H. F. Hofmann. Fluid formed by choroid plexus. A technique for its collection and a comparison of its electrolyte composition with serum and cisternal fluids. J. Neuro- physiol. 23: 485-495, 1960. Ryder, H. W., F. F. Epsey, F. D. Kimbell, E. J. Penka, A. Rosenauer, B. Podolsky, and J. P. Evans. The mechanism of the change in cerebrospinal fluid pressure following an induced change in the volume of the fluid space. J. Lab. Clin. Med. 41: 428-435, 1953. Sadoshima, S., M. Thames, and D. Heistad. Cerebral blood flow during elevation of intracranial pressure: role of sympathetic nerves. Am. J. Physiol. 241: 78-84, 1981. Sadoshima, S. and D. Heistad. Sympathetic nerves protect the blood-brain barrier in stroke-prone spontaneously hypertensive rats. Hypertension 4: 904-907, 1982. Sadoshima, S., D. W. Busija, and D. D. Heistad. Mechanism of protection against stroke in stroke-prone spon- taneously hypertensive rats. Am. J. Physiol. 244: 406-412, 1983. Saeed, M., O. Sommer, J. Haltz, and E. Bassenge. -Adreno- ceptor blockade by phentolamine causesfi -adrenergic vasodilation by increased catecholamine release due to presynapticcfl-blockade. J. Cardiovasc. Pharmacol. 4: 44-52, 1982. Sata, O. and E. A. Bering. Extra-ventricular formation of cerebrospinal fluid. Brain Nerve (Tokyo) 19: 883-885, 1967. Sercombe, R., P. Aubineau, L. Edvinsson, H. Mamo, C. mean, E. Pinard, and J. Seylaz. Neurogenic influence on local cerebral blood flow: Effect of catecholamines or sympathetic stimulation as correlated with the sympa- thetic innervation. Neurol. 25: 954-963, 1975. 177 Sklar, F.1L and.I. Elashvili. The pressure-volume function of brain elasticity. J. Neurosurg. 47: 670-679, 1977. Sklar, F. H., J. Reisch, I. Elashvil i, T. Smith, and D. M. Long. Effects of pressure on cerebrospinal fluid for- mation: nonsteady-state measurements in dogs. Am. J. Physiol. 239: 277-284, 1980. Snodgrass, S. R. and A. V. Lorenzo. Temperature and cerebrospinal fluid production rate. Am. J. Physiol. 222: 1524-1527, 1972. Songer, M. J., W. Dang, S. R. Heisey, T. Adams, and H. K. Fry. .Redistribution of intracranial fluid volumes during hypercapnia (Abstract). Physiologist 23: 98, 1980. Starke, K. :Regulation of noradrenaline release by presynap- tic receptor system. Rev. Physiol. Biochem. Pharmacol. 77: 1-124, 1977. Stone, H. L., M. E. Raichle, and M. Hernandez. The effect of sympathetic denervation on cerebral C02 sensitivity. Stroke 5: 13-18, 1974. Swanson, L. W., M. A. Connel ly, and B. K. Hartman. Ultra- structural evidence for central monoaminergic innerva- tion of blood vessels in the paraventricular nucleus Of the hypothalamus. Brain Res. 136: 166-173, 1977. Tschirgi, R. D., R. W. Frost, and J. L. Taylor. Inhibition of cerebrospinal fluid formation by a carbonic anhy- drase inhibitor, 2-acetylamino-l,3,4-thiadiazole-5- sulfonamide (Diamox). Proc. Soc. Exp. Biol. and Med. 87: 373-376, 1954. Vates, T. 8., S. L. Bonting, and W. W. Oppelt. Na-K acti- vated adenosine triphosphatase formation of cerebro- spinal fluid in the cat. Am. J. Physiol. 206: 1165- 1172, 1964. Vogh, B. 'The relation of choroid plexus carbonic anhydrase activity to cerebrospinal fluid formation: study of three inhibitors in cat with extrapolation to man. J. Pharmacol. Exp. Ther. 213: 321-331, 1980. 178 Wald, A., G. M. Hochwald, and C. Malhan. The effects of ventricular fluid osmolality'on bulk flow'of fluid into cerebral ventricles of cats. Exp. Brain Res. 25: 157- 167, 1976. Waltz, A. G., T. Yamaguchi, and F. Regli. Regulatory re- sponses of cerebral vasculature after sympathetic denervation. Am. J. Physiol. 221: 298-302, 1971. Weed, L. H., L. B. Flexner, and J. H. Clark. Effect of dislocation of cerebrospinal fluid upon its pressure. Am. J. Physiol. 100: 246-261, 1932. Weed, L. H. and L. B. Flexner. The relations of the intra- cranial pressures. Am. J. Physiol. 105: 266-272, 1933. Weiss, M. H. and N.‘Wertman. Modulation of CSF production by alterations in cerebral perfusion pressure. Arch. Neurol. 35: 527-529, 1978. Welch, L. and V. Friedman. The cerebrospinal fluid valves. Brain 83: 454-469, 1960. Welch. K. Secretion of cerebrospinal fluid by the choroid plexus of the rabbit. Am. J. Physiol. 205: 617-624, 1963. Welch, K. The principles of physiology of the cerebrospinal fluid in relation to hydrocephalus. In: Advances in Neurology. I W. J. Friedlander (ed.). Raven Press, New York, 1975. Wright, E. and J. Prather. The permeability of the frog choroid plexus to non-electrolytes. J. Membrane Biol. Wright, E., G.‘Wiedner, and G. Rumrich. Fluid secretion by the frog choroid plexus. Exp. Eye Res. 25: 149-155, 1977. Wright, E. M. Transport process in the formation of the cerebrospinal fluid. Rev. Physiol. Biochem. Pharmacol. 83: 1'34, 1978. Wright, E. Relations between the Choroid plexuses and the cerebrospinal fluid. TINS 1: 13-15, 1979.