Bryostatin-1 - Wiley Online Library

2 downloads 360 Views 76KB Size Report
Institute, 9601 Medical Center Drive, Academic and Research Bldg., Room 319 .... limited antidepressant efficacy of the
CNS Drug Reviews Vol. 12, No. 1, pp. 1–8 © 2006 The Authors Journal compilation © 2006 Blackwell Publishing Inc.

Bryostatin-1: Pharmacology and Therapeutic Potential as a CNS Drug Miao-Kun Sun and Daniel L. Alkon Blanchette Rockefeller Neurosciences Institute, Rockville, MD, USA Keywords: Alzheimer’s disease — Antidepressants — Bryostatin-1 — Cognition activation — Depression — Learning and memory — Mood — Protein kinase C.

ABSTRACT Bryostatin-1 is a powerful protein kinase C (PKC) agonist, activating PKC isozymes at nanomolar concentrations. Pharmacological studies of bryostatin-1 have mainly been focused on its action in preventing tumor growth. Emerging evidence suggests, however, that bryostatin-1 exhibits additional important pharmacological activities. In preclinical studies bryostatin-1 has been shown at appropriate doses to have cognitive restorative and antidepressant effects. The underlying pharmacological mechanisms may involve an activation of PKC isozymes, induction of synthesis of proteins required for long-term memory, restoration of stress-evoked inhibition of PKC activity, and reduction of neurotoxic amyloid accumulation and tau protein hyperphosphorylation. The therapeutic potential of bryostatin-1 as a CNS drug should be further explored.

INTRODUCTION The bryostatins, a family of at least 20 marine natural products with a chemical structure of macrocyclic polyketides, were first isolated from the bryozoan Bugula neritina by Pettit et al. in 1968 and characterized chemically in 1982 (33). Their main pharmacological mechanism of action is modulation of protein kinase C (PKC) activity. The prototypic member, bryostatin-1 (Fig. 1) possesses a unique pharmacological profile as a cancer chemotherapeutic. It is currently in phases I and II trials against a variety of cancers (8,13,19,27) and produces behavioral and cognitive effects when appropriate drug levels are achieved in the brain (14,39). It inhibits tumor invasion, tumor growth in vitro and in vivo, and angiogenesis. It also synergizes with other anticancer agents (2), reverses multidrug resistance (3,37), stimulates the immune system (11), enhances cognition, and has antidepressant effects. The focus of this short review is on pharmacology of bryostatin-1 and evaluation of its therapeutic potential in the treatment of dementia and depression. Address correspondence and reprint requests to: Miao-Kun Sun, Ph.D., Blanchette Rockefeller Neurosciences Institute, 9601 Medical Center Drive, Academic and Research Bldg., Room 319, Rockville, MD 20850, USA. Phone: +1 (301) 294-7181, Fax: +1 (301) 294-7007, E-mail: [email protected]

1

2

M-K. SUN AND D. L. ALKON OH

O

OAc

11

O

B

O

O

A

1

3

OH 19

O

H

OH

C

O

O

26

O

OH O

O O

FIG. 1. Chemical structure of bryostatin-1.

PHARMACODYNAMICS Bryostatin-1 is a partial agonist of several members of the PKC family (see below). Its pharmacological effects appear to depend on its action on the PKC isozymes. Bryostatin-1 binds PKC isozymes with IC50 subnanomolar concentrations and a dissociation half-time of several hours [12]. The binding of bryostatin-1 to PKC results in PKC activation, autophosphorylation, and translocation to the cell membrane. Bryostatin-1-bound PKC is then downregulated by ubiquitination and degradation in proteasomes. Downregulation is most significant when the PKC isozymes are exposed to high and/or prolonged high concentrations of bryostatin-1. It should be, however, mentioned here that modulation of PKC activity may not be the only biological action of bryostatins (25). For example, in B16/F10 melanoma cells (26) epi-bryostatin-1, a bryostatin-1 stereoisomer, with much lower affinity for PKC, exhibits the same potency in inhibiting cell growth as bryostatin-1 (41), suggesting the involvement of aPKC activity-independent action. Nevertheless, modulation of PKC activity is the only well established biological mechanism by which bryostatin-1 produces a variety of biological effects. Bryostatin-1 can modulate the activity of two of the three subgroups of PKC isozymes. PKC activity, with the active site of isozymes usually located at the C-terminus, is involved in the phosphorylation of serine and threonine residues. PKC isozymes are divided into three subgroups, which differ in their molecular structures and co-factor requirements: classical PKC (cPKC; á, âI, âII, and ã isoforms), novel PKC (nPKC; ä, å, å¢, ç, è, and ì isoforms), and atypical PKC (aPKC; æ and ë/i). The cPKC and nPKC contain regulatory C1 domains (C1A and C1B) and are activated by diacylglycerol, phorbol esters, and bryostatins. The difference between cPKC and nPKC is that cPKC has a C2 region corresponding to the Ca2+ binding site and requires Ca2+ as a co-factor for activation, whereas the nPKC does not contain the C2 region and, therefore, does not require Ca2+ as a cofactor for activation. The third group, aPKC, does not bind phorbol esters or bryostatins

CNS Drug Reviews, Vol. 12, No. 1, 2006

BRYOSTATIN-1

3

and contains no C2 region and one of the repeated cysteine-rich zinc finger binding motifs within the C1 domain. Bryostatin-1 is, therefore, expected to modulate nPKC activity, independent of a Ca2+ signal. It activates cPKC only when associated with Ca2+ signaling. aPKC activity is not sensitive to bryostatin-1 administration. Interestingly, Ca2+ signals play an important role in synaptic transmission and information processing. Bryostatin-1 thus possesses a unique action profile: it will not affect cPKC activity in neurons that are not functioning as an active part of the signaling processing circuit with significant Ca2+ influx and intracellular Ca2+ release. Studies of structure-activity relations of bryostatin-1 (Fig. 1) for PKC isoforms (44–46) indicate that the intact 20-membered macrolactone ring is needed for good PKC binding activity while the A-ring and the B-ring exocyclic olefin can be deleted from the 20-membered structure without inducing much change in PKC-binding affinity. The C[26] free hydroxyl is essential for a good interaction with the PKC isozymes, while the C[1] carbonyl group is also important for a high affinity (44). The C[19] hydroxyl group may be involved in the interaction with the lipid bilayer during binding, while the C[3] hydroxyl group plays an important role in the conformation of the molecule. The A-ring region can be extensively modified without affecting binding affinity. Thus the C[9] region could be modified as needed to tune ADME (absorption, distribution, metabolism, and excretion) and pharmacokinetic characteristics (46). The C[20] region is a non-pharmacophoric site and can also be modified to tune analogs for function and physical properties without significantly affecting their binding affinity for the PKC isozymes (44). These positions can thus be used to change for improvement of physical properties and selectivity, and for incorporation of additional functionality, such as a fluorophore. Bryostatin-1 can mimic certain effects of phorbol esters in some biological systems. However, several other properties of bryostatin-1 are distinct from those of the phorbol esters (22). Of critical importance and unlike the phorbol esters, this compound lacks tumor-promoting capabilities and actually counteracts tumor promotion induced by phorbol esters (17), an important pharmacological property when clinical potential is considered.

Role in Alzheimer’s Disease Numerous reports (10,16,24) imply a critical role of deficient functions of PKC isozymes in the pathogenesis of Alzheimer’s disease (AD). Intensive efforts have, therefore, been focused on the development of bryostatin-1 and its analogs to treat dementia of the Alzheimer’s type. AD, a devastating and progressive decline of memory and other cognitive functions (e.g., a reduced ability to learn, loss of memory, decreased attention, judgment, and decision-making), robs the affected individuals of quality of life. The main histopathological hallmarks of AD brain are extracellular senile plaques and intracellular neurofibrillary tangles at late stages of the disorder, particularly in the hippocampus and related neural structures that play an essential role in memory formation. Three pharmacological profiles favor potential use of bryostatin-1 and its analogs to restore cognitive dysfunction especially that associated with AD. 1) PKC isozymes are activated by synaptic inputs and intracellular signals that are involved in information processing and play an important role in various types of learning and memory (4,6). Reduced PKC activity is associated with AD (10,16), at least partially due to a direct inhibition of PKC activity by neurotoxic Aâ (24). Activation of PKC with bryostatin-1 induces the de novo synthesis of proteins necessary and sufficient for subse-

CNS Drug Reviews, Vol. 12, No. 1, 2006

4

M-K. SUN AND D. L. ALKON

quent long-term memory consolidation in Hermissenda (5). PKC inhibition, on the other hand, impairs learning and memory. Mice deficient in PKCb have been found to show normal brain anatomy and normal hippocampal synaptic transmission, paired facilitation, and long-term potentiation of the glutamatergic synaptic responses but a loss of learning in both cued and contextual fear conditioning (42). Bryostatin-1 selectively activates cPKC isozymes in those neurons that are active, as part of the information processing network with Ca2+ signaling. Activation of PKC with bryostatin-1 promotes stabilization of growth-associated protein 43 (GAP-43) mRNA, resulting in an increased GAP-43 protein level in human neuroblastoma cells (34). In addition, evidence has been provided that bryostatin-1, at appropriate doses (bilateral intracerebroventricular 0.64 or 2 pmoles/site), directly enhances learning and memory in Wistar rats (39). 2) Activation of certain PKC isozymes has therapeutic values in reducing the formation of extracellular senile plaques, one of the AD’s main histopathological hallmarks. Bryostatin-1 activates á-secretase, probably through PKC á, ä, or å (14,21,47,50). At subnanomolar concentrations, it enhances the secretion of á-secretase product soluble amyloid precursor protein (sAPP)á in fibroblasts from AD patients and reduces brain Aâ40 and Aâ42 accumulation in AD double-transgenic mice (14). In APP[V7171] transgenic mice, PKC activation has also been found to reduce Aâ40 accumulation in the brain (14). 3) PKC activation inhibits glycogen synthase 3 kinase (15,23) and thereby reduces ô protein hyperphosphorylation (7) and intracellular neurofibrillary tangles, another main histopathological hallmark of AD. The combination of memory-enhancing action and reduction in brain amyloid burden and ô protein hyperphosphorylation may represent a promising multi-target strategy with one agent and an effective therapeutic approach against AD.

Role in Depression Mood disorders are among the most prevalent forms of mental illness. They are recurrent, life threatening (due to the risk of suicide), and a major cause of morbidity worldwide. Several forms of depression affect 2–5% of the U.S. population, and up to 20% of this population suffers from milder forms of the illness. Efforts to enhance monoamine activity in the brain pharmacologically have led to the development of several clinically effective antidepressants. Unfortunately, progress in the development of new and improved antidepressants has been limited. The new drugs developed so far have essentially the same mechanism of action as the older ones and, as a result, the efficacy of the newer agents and the ranges of depressed patients they treat are no better than those of the older drugs. Today’s treatment thus remains suboptimal, with only ~50% of all patients demonstrating complete remission, although many more (up to 80%) show partial response. In addition, the current pharmacological treatments, including the use of the second generation of antidepressants, selective serotonin reuptake inhibitors, are associated in some cases with a disturbing “impulsive” tendency for suicide. The adverse effects as well as limited antidepressant efficacy of the available drugs call for new antidepressants. It is of interest that dementia and depression appear to share common structural and signal pathway injuries (38). An involvement of PKC activity in depression and its treatment has been implicated in several studies. PKC may be involved in 5-hydroxytryptamine (5-HT2A) receptor desensitization (35). Activation of the hypothalamic-pituitaryadrenal axis causes downregulation of PKC isozymes in rat brain (30). Reduced PKC activity occurs in the prefrontal cortex and the hippocampus in suicide victims (28,31,32)

CNS Drug Reviews, Vol. 12, No. 1, 2006

BRYOSTATIN-1

5

and in fibroblasts cultured from skin biopsies from patients with melancholic depression (1). The human plasma membrane serotonin transporter is a substrate for PKC. PKC activation or phosphatase inhibition downregulate uptake of 5-HT. This effect is probably mediated by reduction of the expression of 5-HT transporter (36). Imipramine, a tricyclic antidepressant, but not citalopram, has been shown to increase PKC levels in primary rat neuronal cultures (29). Along this line of an important role of PKC activity in mood regulation is the observation that bryostatin-1, at appropriate doses, has antidepressant activity in rats (39). In an open space swim model of induced depressive behavior, bryostatin-1, at 100 nmoles/kg i.v., significantly reduced non-searching immobility in rats. This antidepressant effect of bryostatin-1 is largely abolished by co-administration of 1-(5-isoquinolinesulfonyl)-2-methylpiperazine (H-7), a PKC inhibitor, suggesting the involvement of PKC activation in its antidepressant action. It is, therefore, likely that bryostatin-1 and/or its analogs could be developed as new antidepressants.

PHARMACOKINETICS In clinical trials as an antitumor agent, bryostatin-1 is given either as a bolus intravenous injection or a continuous infusion. It is difficult to perform pharmacokinetic analysis of bryostatin-1 in humans as its in vivo blood levels are not easily detectable by mass spectrometry or high-performance liquid chromatography. Trial design has been hampered by lack of human pharmacokinetic data. Efforts have been made to develop more specific and sensitive methods for detection of bryostatin-1 in human plasma (49). These efforts may lead to an improvement in the design of future trials. Pharmacokinetics of bryostatin-1 has been, however, studied in animals. In mice, studies with [C26-3H]bryostatin-1, i.v. or i.p. reveal a wide distribution of the drug with high levels in lungs, liver, gastrointestinal tract, and fat (48), whereas its oral bioavailability, to our knowledge, has not been clearly established. Bryostatin-1 is relatively stable to metabolic breakdown in vivo. High-performance liquid chromatography analysis indicates that radioactivity is mainly associated with the intact molecule, suggesting that the compound remained largely intact. Following intravenous administration to mice, the plasma disappearance curve of bryostatin-1 fits a two compartment model, with half-lives of 1.05 and 22.97 h, respectively. In contrast, the plasma disappearance curve for bryostatin-1 administered i.p. is better described by a first order absorption one-compartment model, with an absorption half-life of 0.81 h and an elimination half-life of 28.76 h, respectively. The majority of radioactivity in plasma was associated with the intact drug for up to 24 h after dosing. During the first 12 h after i.v. administration of bryostatin-1, urinary excretion represented the major pathway of drug elimination, with 23.0 ± 1.9% (mean ± S.D.) of the administered dose excreted. Within 72 h after i.v. administration, approximately equal amounts of radioactivity (40%) were excreted in feces and in urine. After i.v. administration of bryostatin-1 to mice, a brief increase in its brain levels was observed, reaching several nanomoles after an i.v. dose of 40 mg/kg (48). These brain levels were sufficient to activate PKC isozymes that were sensitive to bryostatin-1, but the brain levels rapidly declined after administration of the drug. After i.p. injection of the same dose to mice, on the other hand, the peak brain levels were much lower, but the increases were longer-lasting than those observed after an i.v. dose. These data indicate that bryostatin-1 can pass through the blood-brain barrier, although the brain levels of the drug were much lower than its plasma levels (48).

CNS Drug Reviews, Vol. 12, No. 1, 2006

6

M-K. SUN AND D. L. ALKON

ADVERSE EFFECTS AND HUMAN TOXICITY The maximum tolerated dose of bryostatin-1 in humans has been found to be about 25 ìg/m2/week, given intravenously over up to 8 weeks (9,20,26). Toxic and adverse reactions are rare and generally mild. Myalgia is the dose-limiting toxicity in humans. Myalgia occurs at one to two days after infusion and tends to get worse with repeated administration. The symptoms are eased by exercise but return on resting. The calves, thighs and extraocular muscles are affected first but myalgia becomes more generalized as therapy continues. Myalgia affecting the muscles of hypopharynx may result in frontal headache and odynophasia. Lasting impairment of oxidative metabolism in muscle mitochondria may be responsible for myalgia (18). Other reported adverse reactions in humans include fatigue and lethargy; they are common but generally mild. Less common adverse effects include low-grade pyrexia, nausea and anorexia. In a toxicological study in rats, lethargy, unsteadiness, and hematuria have also been observed following bryostatin-1 treatment. Adverse hematological toxicities in humans are not common, although thrombocytopenia and leucopenia have been reported. Mild abnormalities of liver function have also been reported. Bryostatin-1 toxicity appears to be age-dependent. Children can tolerate higher doses of the drug. The maximal tolerated dose of bryotaxin-1 in children has been reported to be 44 ìg/m2 (43). Myalgia and photophobia are the dose limiting toxic effects of bryostatin-1 in children. It remains to be determined whether the age-dependent toxicity of bryostatin-1 would limit its utility in aged AD patients.

CONCLUSIONS Several studies have suggested that at appropriate doses bryostatin-1 can produce beneficial effects on cognition and mood, through modulation of PKC activity. Activation of PKC isozymes represents a potential therapeutic strategy in improving memory and mood, although we do not know precisely which PKC isozyme(s) may mediate the observed effects. PKC isozymes are involved in a variety of vital functions and neurological disorders so that a long-term PKC activation may cause wide and severe reactions. For instance, all forms of PKC isozymes are sensitive to oxidative stress. It remains to be studied whether a sustained increase in the plasma levels of bryostatin-1 would further sensitize PKC isozymes to oxidants. However, intermittent PKC activation with lower doses over an extended period may avoid such reactions. Multiple but intermittent doses of bryostatin-1 or its analogs, administered chronically, are expected to result in brief but multiple “peaks” of PKC activation with smaller doses, maximizing effects of PKC cascade activation but minimizing potential “tolerance” or PKC downregulation that is associated with the exposure to high and/or long-lasting high concentrations of bryostatin-1. Based on the data available in clinical trials, bryostatin-1 is well tolerated as an antitumor agent. This means that its adverse side effects are rare, generally mild, and reversible. It possesses several pharmacological actions, including functional restoration of a reduced PKC signal cascade that occurs in dementia and depression, a reduction in the accumulation of neurotoxic amyloid, and an inhibition of tau protein hyperphosphorylation. Structural modifications of bryostatin-1 may lead to compounds that possess bryostatin-1 activity but can pass through the blood-brain barrier much easier, so that they may achieve the desired brain effects at smaller doses. Bryostatin-1 and its analogs (44,46) may be developed as therapeutic drugs for the treatment of dementia, depression, and/or disorders such as attention-deficit/hyper-

CNS Drug Reviews, Vol. 12, No. 1, 2006

BRYOSTATIN-1

7

activity disorder in which cognitive enhancement is desirable and co-morbid depression is prominent.

REFERENCES 1. Akin D, Hal Manier D, Sanders-Bush E, Shelton RC. Signal transduction abnormalities in melancholic depression. Int J Neuropsychopharmacol 2005;8:5–16. 2. Ali S, Aranha O, Li YW, Pettit GR, Sarkar FH, Philip PA. Sensitization of human breast cancer cells to gemcitabine by the protein kinase C modulator bryostatin-1. Cancer Chemother Pharm 2003;52:235–246. 3. Al-Katib AM, Smith MR, Kamanda WS, et al. Bryostatin-1 down-regulates mdr1 and potentiates vincristine cytotoxicity in diffuse large cell lymphoma xenografts. Clin Cancer Res 1998;4:1305–1314. 4. Alkon DL, Nelson TJ, Zhao WQ, Cavallaro S. Time domains of neuronal Ca2+ signaling and associative memory: Steps through a calexcitin, ryanodine receptor, K+ channel cascade. Trends Neurosci 1998;21: 529–537. 5. Alkon DL, Epstein H, Kuzirian A, Bennett M. C, Nelson TJ. Protein synthesis required for long-term memory is induced by PKC activation on days before associative learning. Proc Natl Acad Sci USA 2005; 102:16432–16437. 6. Bank B, DeWeer A, Kuzirian AM, Rasmussen H, Alkon DL. Classical conditioning induces long-term translocation of protein kinase C in rabbit hippocampal CA1 cells. Proc Natl Acad Sci USA 1988;85: 1988–1992. 7. Cho JH, Johnson GV. Glycogen synthase kinase 3 beta induces caspase-cleaved tau aggregation in situ. J Biol Chem 2004;279:54716–54723. 8. Clamp A, Jayson GC. The clinical development of the bryostatins. Anti-Cancer Drugs 2002;13:673–683. 9. Clamp AR, Blackhall FH, Vasey P, et al. A phase II trial of bryostatin-1 administered by weekly 24-hour infusion in recurrent epithelial ovarian carcinoma. Br J Cancer 2003;89:1152–1154. 10. Cole G, Dobkins KR, Hansen LA, Terry RD, Saitoh T. Decreased levels of protein kinase C in Alzheimer brain. Brain Res 1988;452:165–174. 11. Curiel RE, Garcia CS, Farooq L, Aguero MF, Espinoza-Dagado I. Bryostatin-1 and IL-2 synergize to induce IFN-gamma expression in human peripheral blood T cells: Implications for cancer immunotherapy. J Immunol 2001;167:4828–4837. 12. de Vries D, Herald CL, Pettit HG, Blumberg PM. Demonstration of sub-nanomolar affinity of bryostatin-1 for the phorbol ester receptor in rat brain. Biochem Pharmacol 1988;37:4069–4073. 13. Dowlati A, Lazarus HM, Hartman P, et al. Phase I and correlative study of combination bryostatin-1 and vincristine in relapsed B-cell malignancies. Clin Cancer Res 2003;9:5929–5935. 14. Etcheberrigaray R, Tan M, Dewachter I, et al. Therapeutic effects of PKC activators in Alzheimer’s disease transgenic mice. Proc Natl Acad Sci USA 2004;101:11141–11146. 15. Fang X, Yu S, Tanyi JL, Lu Y, Woodgett JR, Mills GB. Convergence of multiple signaling cascades at glycogen synthase kinase 3: Edg receptor-mediated phsophorylation and inactivation by lysophosphatidic acid through a protein kinase C-dependent intracellular pathway. Mol Cell Biol 2002;22:2099–2110. 16. Favit A, Grimaldi M, Nelson TJ, Alkon DL. Alzheimer’s-specific effects of soluble â-amyloid on protein kinase C-á and -ã degradation in human fibroblasts. Proc Natl Acad Sci USA 1998;10:5562–5567. 17. Hennings H, Blumberg PM, Pettit GR, Herald CL, Shores R, Yuspa SH. Bryostatin-1, an activator of protein kinase C, inhibits tumor promotion by phorbol esters in SENCAR mouse skin. Carcinogenesis 1987;8: 1343–1346. 18.18. Hickman PF, Kemp GJ, Thompson CH, et al. Bryostatin-1, a novel antineoplastic agent and protein kinase C activator, induces human myalgia and muscle metabolic defects: A 31P magnetic resonance spectroscopic study. Br J Cancer 1995;72:998–1003. 19. Hofmann J Protein kinase C isozymes as potential targets for anticancer therapy. Curr Cancer Drug Targets 2004;4:125–146. 20. Jayson GC, Crowther D, Prendiville J, et al. A phase I trial of bryostatin-1 in patients with advanced malignancy using a 24 hour intravenous infusion. Br J Cancer 1995;72:461–468. 21. Kinouchi T, Sorimachi H, Maruyama K, et al. Conventional protein kinase C (PKC)-alpha and novel PKC epsilon, but not –delta, increase the secretion of an N-terminal fragment of Alzheimer’s disease amyloid precursor protein from PKC cDNA transfected 3Y1 fibroblasts. FEBS Lett 1995;364:203–206. 22. Kraft AS, Reeves JA, Ashendel CL. Differing modulation of protein kinase C by bryostatin-1 and phorbol esters in JB6 mouse epidermal cells. J Biol Chem 1988;263:8437. 23. Lavoie L, Band CJ, Kong M, Bergeron JJM, Posner BI. Regulation of glycogen synthase in rat hepatocytes. Evidence for multiple signaling pathway. J Biol Chem 1999;274:28279–28285.

CNS Drug Reviews, Vol. 12, No. 1, 2006

8

M-K. SUN AND D. L. ALKON

24. Lee W, Boo JH, Jung MW, et al. Amyloid beta peptide directly inhibits PKC activation. Mol Cell Neurosci 2004;26:222–231. 25. Matthews SA, Pettit GR, Rozengurt E. Bryostatin-1 induces biphasic activation of protein kinase D in intact cells. J Biol Chem 1997;272:20245–20250. 26. Mutter R, Wills M. Chemistry and clinical biology of the bryostatins. Bioorg Med Chem 2000;8:1841–1860. 27. Nezhat F, Wadler S, Muggia F, et al. Phase II trial of the combination of bryostatin-1 and cisplatin in advanced or recurrent carcinoma of the cervix: A New York Gynecologic Oncology Group Study. Gynecol Oncol 2004;93:144–148. 28. Pacheco MA, Stockmeier C, Meltzer NY, Overholser JC, Dilley GE, Jope RS. Alterations in phosphoinositide signaling and G-protein levels in depressed suicide brain. Brain Res 1996;723:37–45. 29. Pákáski M, Bjelik A, Hugyecz M, Kása P, Janka Z, Kálmán J. Imipramine and citalopram facilitate amyloid precursor protein secretion in vitro. Neurochem Intern 2005;47:190–195. 30. Pandey GN, Dwivedi Y. Effects of adrenal glucocorticoid on protein kinase C (PKC) binding sites, PKC activity and expression of PKC isozymes in the brain. J Neurochem 2000;74:S21D. 31. Pandey GN, Dwivedi Y, Ren X, et al. Altered expression and phosphorylation of myristoylated alanine-rich C kinase (MARCKS) in postmortem brain of suicide victims with or without depression. J Psychiatr Res 2003;37:421–432. 32. Pandey GN, Ewivedi Y, Rizavi HS, Ren X, Conley RR. Decreased catalytic activity and expression of protein kinase C isozymes in teenage suicide victims. Arch Gen Psychiatry 2004;61:685–693. 33. Pettit GR, Herald CL, Doubek DL, Arnold E, Clardy J. Isolation and structures of bryostatin-1. J Am Chem Soc 1982;104:6846–6848. 34. Pascale A, Amadio M, Scapagnini G, et al. Neuronal ELAV proteins enhances mRNA stability by a PKCádependent pathway. Proc Natl Acad Sci USA 2005;102:12065–12070. 35. Rahimian R, Hrdina PD. Possible role of protein kinase C in regulation of 5-hydroxytryptamine 2A receptors in rat brain. Can J Physiol Pharmacol 1995;73:1686–1691. 36. Ramamoorthy S, Giovanetti E, Qian Y, Blakely RD. Phosphorylation and regulation of antidepressant-sensitive serotonin transporters. J Biol Chem 1998;273:2458–2466. 37. Spitaler M, Utz I, Hilbe W, Hofmann J, Grunicke HH. PKC-independent modulation of multidrug resistance in cells with mutant (V185) but not wild-type (G185) P-glycoprotein by bryostatin-1. Biochem Pharm 1998; 56:861–869. 38. Sun M-K, Alkon DL. Depressed or demented: Common CNS drug targets?! Curr Drug Targets. CNS Neurol Disord 2002;1:573–589. 39. Sun M-K, Alkon DL. Dual effects of bryostatin-1 on spatial memory and depression. Eur J Pharmacol 2005;512:43–51. 40. Sun M-K, Alkon DL. Protein kinase C isozymes: Memory therapeutic potential. Curr Drug Targets. CNS Neurol Disord 2005;4:541–52. 41. Szallasi Z, Du L, Levine R, et al. The bryostatins inhibit growth of B16/F10 melanoma cells in vitro through a protein kinase C-independent mechanism: Dissociation of activities using 26-epi-bryostatin-1. Cancer Res 1996;56:2105–2111. 42. Weeber EJ, Atkins CM, Selcher JC, et al. A role for the b isoform of protein kinase C in fear conditioning. J Neurosci 2000;20:5906–5914. 43. Weitman S, Langevin AM, Berkow RL, et al. A phase I trial of bryostatin-1 in children with refractory solid tumors: A pediatric oncology group study. Clin Cancer Res 1999;5:2344–2348. 44. Wender PA, Baryza JL, Bennett CE, et al. The practical synthesis of a novel and highly potent analogue of bryostatin. J Am Chem Soc 2002;124:13648–13649. 45. Wender PA, Baryza JL. Identification of a tunable site in bryostatin analogs: C20 bryologs through late stage diversification. Org Lett 2005;7:1177–1180. 46. Wender PA, Clarke MO, Horan JC. Role of the A-ring of bryostatin analogues in PKC binding: Synthesis and initial biological evaluation of new A-ring-modified bryologs. Org Lett 2005;7:1995–1998. 47. Yeon SW, Jung MW, Ha MJ, et al. Blockade of PKC epsilon activation attenuates phorbol ester-induced increase of alpha-secretase-derived secreted form of amyloid precursor protein. Biochem Biophys Res Commun 2001;280:782–787. 48. Zhang X, Zhang R, Zhao H, et al. Preclinical pharmacology of the natural product anticancer agent bryostatin-1, an activator of protein kinase C. Cancer Res 56:802–808,1996. 49. Zhao M, Rudek MA, He P, Smith, BD, Baker S.D.. Validation and implementation of a method for determination of bryostatin-1 in human plasma by using liquid chromatography/tandem mass spectrometry. Ann Biochem 337:143–148,2005. 50. Zhu G, Wang D, Lin YH, McMahon T, Koo EH, Messing RO. Protein kinase C epsilon suppresses Abeta production and promotes activation of á-secretase. Biochem Biophys Res Commun 2001;285:997–1006.

CNS Drug Reviews, Vol. 12, No. 1, 2006