1
|
Farella I, D’Amato G, Orellana-Manzano A, Segura Y, Vitale R, Clodoveo ML, Corbo F, Faienza MF. "OMICS" in Human Milk: Focus on Biological Effects on Bone Homeostasis. Nutrients 2024; 16:3921. [PMID: 39599707 PMCID: PMC11597255 DOI: 10.3390/nu16223921] [Citation(s) in RCA: 0] [Impact Index Per Article: 0] [Reference Citation Analysis] [Abstract] [Key Words] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 10/29/2024] [Revised: 11/12/2024] [Accepted: 11/14/2024] [Indexed: 11/29/2024] Open
Abstract
Human milk (HM) is a complex biofluid rich in nutrients and bioactive compounds essential for infant health. Recent advances in omics technologies-such as proteomics, metabolomics, and transcriptomics-have shed light on the influence of HM on bone development and health. This review discusses the impact of various HM components, including proteins, lipids, carbohydrates, and hormones, on bone metabolism and skeletal growth. Proteins like casein and whey promote calcium absorption and osteoblast differentiation, supporting bone mineralization. Long-chain polyunsaturated fatty acids like docosahexaenoic acid (DHA) contribute to bone health by modulating inflammatory pathways and regulating osteoclast activity. Additionally, human milk oligosaccharides (HMOs) act as prebiotics, improving gut health and calcium bioavailability while influencing bone mineralization. Hormones present in HM, such as insulin-like growth factor 1 (IGF-1), leptin, and adiponectin, have been linked to infant growth, body composition, and bone density. Research has shown that higher IGF-1 levels in breast milk are associated with increased weight gain, while leptin and adiponectin influence fat mass and bone metabolism. Emerging studies have also highlighted the role of microRNAs (miRNAs) in regulating key processes like adipogenesis and bone homeostasis. Furthermore, microbiome-focused techniques reveal HM's role in establishing a balanced infant gut microbiota, indirectly influencing bone development by enhancing nutrient absorption. Although current findings are promising, comprehensive longitudinal studies integrating omics approaches are needed to fully understand the intricate relationships among maternal diet, HM composition, and infant bone health. Bridging these gaps could offer novel dietary strategies to optimize skeletal health during infancy, advancing early-life nutrition science.
Collapse
Affiliation(s)
- Ilaria Farella
- Department of Medicine and Surgery, LUM University, 70010 Casamassima, Italy;
| | - Gabriele D’Amato
- Neonatal Intensive Care Unit, Di Venere Hospital, 70012 Bari, Italy;
| | - Andrea Orellana-Manzano
- Escuela Superior Politécnica del Litoral, ESPOL, Laboratorio Para Investigaciones Biomédicas, Facultad de Ciencias de la Vida (FCV), ESPOL Polytechnic University, Campus Gustavo Galindo Km 30.5 vía Perimetral, Guayaquil P.O. Box 09-01-5863, Ecuador; (A.O.-M.); (Y.S.)
| | - Yaritza Segura
- Escuela Superior Politécnica del Litoral, ESPOL, Laboratorio Para Investigaciones Biomédicas, Facultad de Ciencias de la Vida (FCV), ESPOL Polytechnic University, Campus Gustavo Galindo Km 30.5 vía Perimetral, Guayaquil P.O. Box 09-01-5863, Ecuador; (A.O.-M.); (Y.S.)
| | - Rossella Vitale
- Giovanni XXIII Pediatric Hospital, University of Bari “A. Moro”, 70124 Bari, Italy;
| | - Maria Lisa Clodoveo
- Interdisciplinary Department of Medicine, School of Medicine, University of Bari “A. Moro”, 70100 Bari, Italy;
| | - Filomena Corbo
- Department of Pharmacy-Drug Sciences, University of Bari “A. Moro”, 70125 Bari, Italy;
| | - Maria Felicia Faienza
- Pediatric Unit, Department of Precision and Regenerative Medicine and Ionian Area, University of Bari “A.Moro”, 70124 Bari, Italy
| |
Collapse
|
2
|
Sun H, Xie Q, Zhou Y, Liu Y, Pan J, Xu Y, Jiang S, Li K. An investigation of association between human milk mineral patterns and infant growth. Front Nutr 2024; 11:1387956. [PMID: 38962446 PMCID: PMC11220249 DOI: 10.3389/fnut.2024.1387956] [Citation(s) in RCA: 0] [Impact Index Per Article: 0] [Reference Citation Analysis] [Abstract] [Key Words] [Track Full Text] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Received: 02/19/2024] [Accepted: 06/10/2024] [Indexed: 07/05/2024] Open
Abstract
Introduction Human milk is widely acknowledged as the optimal food for infant aged 0 ~ 6 months. While there has been extensive documentation on the mineral and trace element composition of human milk, results on the relationship between mineral content and infant growth remain mixed. This cross-sectional study aims to explore human milk mineral patterns and to investigate associations between human milk mineral patterns, human milk metabolomic profile and infant growth. Methods A total of 200 breast milk samples from seven cities in China was included. Human milk mineral and trace elements was detected by inductively coupled plasma mass spectrometer (ICP-MS). K-means cluster analysis was utilized to derived human milk mineral patterns. Untargeted human milk metabolomic profiles was determined using high performance liquid chromatography-tandem mass spectrometry (HPLC-MS/MS). Differences of infant growth rate and metabolomic profiles were then compared across patterns identified. Results Three human milk mineral patterns were identified. Cluster I was characterized as the highest levels of potassium, magnesium and calcium, while the lowest levels of copper, zinc, manganese and selenium. Cluster II showed the most abundant sodium, iron, zinc, manganese and selenium. Cluster III had the lowest levels of sodium, potassium, magnesium, iron and calcium. Infants of cluster I showed significantly higher length-for-age z score (0.60 ± 2.03, p = 0.03). Compared with other clusters, samples of cluster I showed lower expression of metabolites of arachidonic acid (ARA) and nicotinate and nicotinamide metabolism pathway. Discussion A human milk mineral pattern was identified which is related to increased infant growth rate and altered metabolic signature. Future work is needed to understand these human milk patterns in terms of biologic mechanisms and generalization to other populations.
Collapse
Affiliation(s)
- Han Sun
- Feihe Research Institute, Heilongjiang Feihe Dairy Co., Ltd., Beijing, China
- PKUHSC-China Feihe Joint Research Institute of Nutrition and Healthy Lifespan Development, Beijing, China
| | - Qinggang Xie
- Feihe Research Institute, Heilongjiang Feihe Dairy Co., Ltd., Beijing, China
- PKUHSC-China Feihe Joint Research Institute of Nutrition and Healthy Lifespan Development, Beijing, China
| | - Yalin Zhou
- PKUHSC-China Feihe Joint Research Institute of Nutrition and Healthy Lifespan Development, Beijing, China
- Department of Nutrition and Food Hygiene, School of Public Health, Peking University, Beijing, China
- Beijing Key Laboratory of Toxicological Research and Risk Assessment for Food Safety, Peking University, Beijing, China
| | - Yang Liu
- Feihe Research Institute, Heilongjiang Feihe Dairy Co., Ltd., Beijing, China
- PKUHSC-China Feihe Joint Research Institute of Nutrition and Healthy Lifespan Development, Beijing, China
| | - Jiancun Pan
- Feihe Research Institute, Heilongjiang Feihe Dairy Co., Ltd., Beijing, China
- PKUHSC-China Feihe Joint Research Institute of Nutrition and Healthy Lifespan Development, Beijing, China
| | - Yajun Xu
- PKUHSC-China Feihe Joint Research Institute of Nutrition and Healthy Lifespan Development, Beijing, China
- Department of Nutrition and Food Hygiene, School of Public Health, Peking University, Beijing, China
- Beijing Key Laboratory of Toxicological Research and Risk Assessment for Food Safety, Peking University, Beijing, China
| | - Shilong Jiang
- Feihe Research Institute, Heilongjiang Feihe Dairy Co., Ltd., Beijing, China
- PKUHSC-China Feihe Joint Research Institute of Nutrition and Healthy Lifespan Development, Beijing, China
| | - Kaifeng Li
- Feihe Research Institute, Heilongjiang Feihe Dairy Co., Ltd., Beijing, China
- PKUHSC-China Feihe Joint Research Institute of Nutrition and Healthy Lifespan Development, Beijing, China
| |
Collapse
|
3
|
Lin Z, McClure MJ, Zhao J, Ramey AN, Asmussen N, Hyzy SL, Schwartz Z, Boyan BD. MicroRNA Contents in Matrix Vesicles Produced by Growth Plate Chondrocytes are Cell Maturation Dependent. Sci Rep 2018; 8:3609. [PMID: 29483516 PMCID: PMC5826934 DOI: 10.1038/s41598-018-21517-4] [Citation(s) in RCA: 25] [Impact Index Per Article: 3.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Received: 05/19/2017] [Accepted: 02/06/2018] [Indexed: 01/01/2023] Open
Abstract
Chondrocytes at different maturation states in the growth plate produce matrix vesicles (MVs), membrane organelles found in the extracellular matrix, with a wide range of contents, such as matrix processing enzymes and receptors for hormones. We have shown that MVs harvested from growth zone (GC) chondrocyte cultures contain abundant small RNAs, including miRNAs. Here, we determined whether RNA also exists in MVs produced by less mature resting zone (RC) chondrocytes and, if so, whether it differs from the RNA in MVs produced by GC cells. Our results showed that RNA, small RNA specifically, was present in RC-MVs, and it was well-protected from RNase by the phospholipid membrane. A group of miRNAs was enriched in RC-MVs compared RC-cells, suggesting that miRNAs are selectively packaged into MVs. High throughput array and RNA sequencing showed that ~39% miRNAs were differentially expressed between RC-MVs and GC-MVs. Individual RT-qPCR also confirmed that miR-122-5p and miR-150-5p were expressed at significantly higher levels in RC-MVs compared to GC-MVs. This study showed that growth plate chondrocytes at different differentiation stages produce different MVs with different miRNA contents, further supporting extracellular vesicle miRNAs play a role as "matrisomes" that mediate the cell-cell communication in cartilage and bone development.
Collapse
Affiliation(s)
- Zhao Lin
- Department of Periodontics, School of Dentistry, Virginia Commonwealth University, Richmond, VA, USA
- Department of Biomedical Engineering, School of Engineering, Virginia Commonwealth University, Richmond, VA, USA
| | - Michael J McClure
- Department of Biomedical Engineering, School of Engineering, Virginia Commonwealth University, Richmond, VA, USA
| | - Junjun Zhao
- Department of Periodontics, School of Dentistry, Virginia Commonwealth University, Richmond, VA, USA
- Department of Biomedical Engineering, School of Engineering, Virginia Commonwealth University, Richmond, VA, USA
- General Dentistry, 9th People's Hospital, College of Stomatology, Shanghai JiaoTong University School of Medicine, Shanghai, China
| | - Allison N Ramey
- Department of Biomedical Engineering, School of Engineering, Virginia Commonwealth University, Richmond, VA, USA
| | - Niels Asmussen
- School of Integrated Life Science, Virginia Commonwealth University, Richmond, VA, USA
| | - Sharon L Hyzy
- Department of Biomedical Engineering, School of Engineering, Virginia Commonwealth University, Richmond, VA, USA
| | - Zvi Schwartz
- Department of Biomedical Engineering, School of Engineering, Virginia Commonwealth University, Richmond, VA, USA
- Department of Periodontics, The University of Texas Health Science Center at San Antonio, San Antonio, TX, USA
| | - Barbara D Boyan
- Department of Biomedical Engineering, School of Engineering, Virginia Commonwealth University, Richmond, VA, USA.
- Wallace H. Coulter Department of Biomedical Engineering, Georgia Institute of Technology, Atlanta, GA, USA.
| |
Collapse
|
4
|
Abstract
Apart from its classical function in bone and calcium metabolism, vitamin D is also involved in immune regulation and has been linked to various cancers, immune disorders and allergic diseases. Within the innate and adaptive immune systems, the vitamin D receptor and enzymes in monocytes, dendritic cells, epithelial cells, T lymphocytes and B lymphocytes mediate the immune modulatory actions of vitamin D. Vitamin D insufficiency/deficiency early in life has been identified as one of the risk factors for food allergy. Several studies have observed an association between increasing latitude and food allergy prevalence, plausibly linked to lower ultraviolet radiation (UVR) exposure and vitamin D synthesis in the skin. Along with mounting epidemiological evidence of a link between vitamin D status and food allergy, mice and human studies have shed light on the modulatory properties of vitamin D on the innate and adaptive immune systems. This review will summarize the literature on the metabolism and immune modulatory properties of vitamin D, with particular reference to food allergy.
Collapse
|
5
|
Doroudi M, Plaisance MC, Boyan BD, Schwartz Z. Membrane actions of 1α,25(OH)2D3 are mediated by Ca(2+)/calmodulin-dependent protein kinase II in bone and cartilage cells. J Steroid Biochem Mol Biol 2015; 145:65-74. [PMID: 25263660 DOI: 10.1016/j.jsbmb.2014.09.019] [Citation(s) in RCA: 16] [Impact Index Per Article: 1.6] [Reference Citation Analysis] [Abstract] [Key Words] [MESH Headings] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Received: 06/24/2014] [Revised: 09/12/2014] [Accepted: 09/16/2014] [Indexed: 01/05/2023]
Abstract
1α,25(OH)2D3 regulates osteoblasts and chondrocytes via its membrane-associated receptor, protein disulfide isomerase A3 (Pdia3) in caveolae. 1α,25(OH)2D3 binding to Pdia3 leads to phospholipase-A2 (PLA2)-activating protein (PLAA) activation, stimulating cytosolic PLA2 and resulting in prostaglandin E2 (PGE2) release and PKCα activation, subsequently stimulating differentiation. However, how PLAA transmits the signal to cPLA2 is unknown. Ca(2+)/calmodulin (CaM)-dependent protein kinase II (CaMKII) activation is required for PLA2 activation in vascular smooth muscle cells, suggesting a similar role in 1α,25(OH)2D3-dependent signaling. The aim of the present study is to evaluate the roles of CaM and CaMKII as mediators of 1α,25(OH)2D3-stimulated PLAA-dependent activation of cPLA2 and PKCα, and downstream biological effects. The results indicated that 1α,25(OH)2D3 and PLAA-peptide increased CaMKII activity within 9 min. Silencing Cav-1, Pdia3 or Plaa in osteoblasts suppressed this effect. Similarly, antibodies against Plaa or Pdia3 blocked 1α,25(OH)2D3-dependent CaMKII. Caveolae disruption abolished activation of CaMKII by 1α,25(OH)2D3 or PLAA. CaMKII-specific and CaM-specific inhibitors reduced cPLA2 and PKC activities, PGE2 release and osteoblast maturation markers in response to 1α,25(OH)2D3. Camk2a-silenced but not Camk2b-silenced osteoblasts showed comparable effects. Immunoprecipitation showed increased interaction of CaM and PLAA in response to 1α,25(OH)2D3. The results indicate that membrane actions of 1α,25(OH)2D3 via Pdia3 triggered the interaction between PLAA and CaM, leading to dissociation of CaM from caveolae, activation of CaMKII, and downstream PLA2 activation, and suggest that CaMKII plays a major role in membrane-mediated actions of 1α,25(OH)2D3.
Collapse
Affiliation(s)
- Maryam Doroudi
- School of Biology, Georgia Institute of Technology, 310 Ferst Drive NW, Atlanta, GA 30332, USA
| | - Marc C Plaisance
- Wallace H. Coulter Department of Biomedical Engineering, Georgia Institute of Technology, 313 Ferst Drive NW, Atlanta, GA 30332, USA
| | - Barbara D Boyan
- School of Biology, Georgia Institute of Technology, 310 Ferst Drive NW, Atlanta, GA 30332, USA; Wallace H. Coulter Department of Biomedical Engineering, Georgia Institute of Technology, 313 Ferst Drive NW, Atlanta, GA 30332, USA; Department of Biomedical Engineering, School of Engineering, Virginia Commonwealth University, 601 West Main Street, Richmond, VA 23284, USA.
| | - Zvi Schwartz
- Department of Biomedical Engineering, School of Engineering, Virginia Commonwealth University, 601 West Main Street, Richmond, VA 23284, USA
| |
Collapse
|
6
|
Chen J, Lobachev KS, Grindel BJ, Farach-Carson MC, Hyzy SL, El-Baradie KB, Olivares-Navarrete R, Doroudi M, Boyan BD, Schwartz Z. Chaperone properties of pdia3 participate in rapid membrane actions of 1α,25-dihydroxyvitamin d3. Mol Endocrinol 2013; 27:1065-77. [PMID: 23660595 DOI: 10.1210/me.2012-1277] [Citation(s) in RCA: 14] [Impact Index Per Article: 1.2] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/19/2022] Open
Abstract
Protein disulfide isomerase family A, member 3 (Pdia3) mediates many of the plasma membrane (PM)-associated rapid responses to 1α,25-dihydroxyvitamin D3 (1α,25[OH]2D3). It is not well understood how Pdia3, which is an endoplasmic reticulum (ER) chaperone, functions as a PM receptor for 1α,25(OH)2D3. We mutated 3 amino acids (K214 and R282 in the calreticulin interaction site and C406 in the isomerase catalytic site), which are important for Pdia3's ER chaperone function, and examined their role in responses to 1α,25(OH)2D3. Pdia3 constructs with and without the ER retention signal KDEL were used to investigate the PM requirement for Pdia3. Finally, we determined whether palmitoylation and/or myristoylation were required for Pdia3-mediated responses to 1α,25(OH)2D3. Overexpressing the Pdia3 R282A mutant in MC3T3-E1 cells increased PM phospholipase A2-activating protein, Rous sarcoma oncogene (c-Src), and caveolin-1 but blocked increases in 1α,25(OH)2D3-stimulated protein kinase C (PKC) seen in cells overexpressing wild-type Pdia3 (Pdia3Ovr cells). Cells overexpressing Pdia3 with K214A and C406S mutations had PKC activity comparable to untreated controls, indicating that the native response to 1α,25(OH)2D3 also was blocked. Overexpressing Pdia3[-KDEL] increased PM localization and augmented baseline PKC, but the stimulatory effect of 1α,25(OH)2D3 was comparable to that seen in wild-type cultures. In contrast, 1α,25(OH)2D3 increased prostaglandin E2 in Pdia3[±KDEL] cells. Although neither palmitoylation nor myristoylation was required for PM association of Pdia3, myristoylation was needed for PKC activation. These data indicate that both the chaperone functional domains and the subcellular location of Pdia3 control rapid membrane responses to 1α,25(OH)2D3.
Collapse
Affiliation(s)
- Jiaxuan Chen
- Wallace H. Coulter Department of Biomedical Engineering, Georgia Institute of Technology, Atlanta, Georgia 30332, USA
| | | | | | | | | | | | | | | | | | | |
Collapse
|
7
|
Abstract
All cells comprising the skeleton-chondrocytes, osteoblasts, and osteoclasts-contain both the vitamin D receptor and the enzyme CYP27B1 required for producing the active metabolite of vitamin D, 1,25 dihydroxyvitamin D. Direct effects of 25 hydroxyvitamin D and 1,25 dihydroxyvitamin D on these bone cells have been demonstrated. However, the major skeletal manifestations of vitamin D deficiency or mutations in the vitamin D receptor and CYP27B1, namely rickets and osteomalacia, can be corrected by increasing the intestinal absorption of calcium and phosphate, indicating the importance of indirect effects. On the other hand, these dietary manipulations do not reverse defects in osteoblast or osteoclast function that lead to osteopenic bone. This review discusses the relative importance of the direct versus indirect actions of vitamin D on bone, and provides guidelines for the clinical use of vitamin D to prevent/treat bone loss and fractures.
Collapse
Affiliation(s)
- Daniel D Bikle
- University of California, San Francisco, San Francisco, CA, USA.
| |
Collapse
|
8
|
Zhong M, Carney DH, Jo H, Boyan BD, Schwartz Z. Inorganic phosphate induces mammalian growth plate chondrocyte apoptosis in a mitochondrial pathway involving nitric oxide and JNK MAP kinase. Calcif Tissue Int 2011; 88:96-108. [PMID: 21104071 DOI: 10.1007/s00223-010-9433-5] [Citation(s) in RCA: 16] [Impact Index Per Article: 1.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Received: 03/05/2010] [Accepted: 09/02/2010] [Indexed: 10/18/2022]
Abstract
Chondrocytes in the hypertrophic zone of the growth plate undergo apoptosis during endochondral bone development via mechanisms that involve inorganic phosphate (Pi) and nitric oxide (NO). Recent evidence suggests that Pi-dependent NO production plays a role in apoptosis of cells in the resting zone as well. This study examined the mechanism by which Pi induces NO production and the signaling pathways by which NO mediates its effects on apoptosis in these cells. Pi decreased the number of viable cells based on MTT activity; the number of TUNEL-positive cells and the level of DNA fragmentation were increased, indicating an increase in apoptosis. Blocking NO production using the NO synthase (NOS) inhibitor L: -NAME or cells from eNOS(-/-) mice blocked Pi-induced chondrocyte apoptosis, indicating that NO production is necessary. NO donors NOC-18 and SNOG both induced chondrocyte apoptosis. SNOG also upregulated p53 expression, the Bax/Bcl-2 expression ratio, and cytochrome c release from mitochondria, as well as caspase-3 activity, indicating that NO induces apoptosis via a mitochondrial pathway. Inhibition of JNK, but not of p38 or ERK1/2, MAP kinase was able to block NO-induced apoptosis, indicating that JNK is necessary in this pathway. Pi elevates NO production via eNOS in resting zone chondrocytes, which leads to a mitochondrial apoptosis pathway dependent on JNK.
Collapse
Affiliation(s)
- M Zhong
- Wallace H. Coulter Department of Biomedical Engineering at Georgia Tech and Emory University, Georgia Institute of Technology, 315 Ferst Drive NW, Atlanta, GA 30332-0363, USA
| | | | | | | | | |
Collapse
|
9
|
Zhong M, Carney DH, Boyan BD, Schwartz Z. 17β-Estradiol regulates rat growth plate chondrocyte apoptosis through a mitochondrial pathway not involving nitric oxide or MAPKs. Endocrinology 2011; 152:82-92. [PMID: 21068162 DOI: 10.1210/en.2010-0509] [Citation(s) in RCA: 18] [Impact Index Per Article: 1.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 11/19/2022]
Abstract
Estrogens cause growth plate closure in both males and females, by decreasing proliferation and inducing apoptosis of postproliferative growth plate chondrocytes. In vitro studies using 17β-estradiol (E(2)) conjugated to bovine serum albumin (E(2)-BSA) show that rat costochondral growth plate resting zone chondrocytes also respond to E(2). Moreover, they are regulated by E(2)-BSA via a protein kinase C and ERK MAPK signaling pathway that is functional only in female cells. To better understand how E(2) regulates apoptosis of growth plate chondrocytes, rat resting zone chondrocytes cells were treated with E(2) or E(2)-BSA. E(2) caused apoptosis in male and female resting zone and growth zone chondrocytes in a dose-dependent manner, based on elevated DNA fragmentation, terminal deoxynucleotidyl transferase dUTP nick end labeling staining and caspase-3 activation. E(2) also up-regulated p53 and Bax protein (Bcl-2-associated X protein) levels and induced release of cytochrome C from the mitochondria, indicating a mitochondrial apoptotic pathway. The apoptotic effect of E(2) did not involve elevated nitric oxide production or MAPKs. It was reduced by ICI 182780, which is an estrogen receptor (ER) antagonist and blocked by antibodies to Erα36, a membrane-associated ER. E(2)-BSA reduced cell viability and increased caspase-3 activity; ICI 182780 had no effect, but anti-ERα36 antibodies blocked the effect. The results indicate that estrogen is able to directly affect the cell population kinetics of growth plate chondrocytes by regulating apoptosis, as well as proliferation and differentiation in both resting zone and growth zone cells. They also have provided further information about the physiological functions of estrogen on longitudinal bone growth.
Collapse
Affiliation(s)
- M Zhong
- Department of Biomedical Engineering, Georgia Institute of Technology, 315 Ferst Drive NW, Atlanta, Georgia 30332-0363, USA
| | | | | | | |
Collapse
|
10
|
Chen J, Olivares-Navarrete R, Wang Y, Herman TR, Boyan BD, Schwartz Z. Protein-disulfide isomerase-associated 3 (Pdia3) mediates the membrane response to 1,25-dihydroxyvitamin D3 in osteoblasts. J Biol Chem 2010; 285:37041-50. [PMID: 20843786 DOI: 10.1074/jbc.m110.157115] [Citation(s) in RCA: 71] [Impact Index Per Article: 4.7] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
Protein-disulfide isomerase-associated 3 (Pdia3) is a multifunctional protein hypothesized to be a membrane receptor for 1,25(OH)(2)D(3). In intestinal epithelium and chondrocytes, 1,25(OH)(2)D(3) stimulates rapid membrane responses that are different from genomic effects via the vitamin D receptor (VDR). In this study, we show that 1,25(OH)(2)D(3) stimulates phospholipase A(2) (PLA(2))-dependent rapid release of prostaglandin E(2) (PGE(2)), activation of protein kinase C (PKC), and regulation of bone-related gene transcription and mineralization in osteoblast-like MC3T3-E1 cells (WT) via a mechanism involving Pdia3. Pdia3 was present in caveolae based on co-localization with lipid rafts and caveolin-1. In Pdia3-silenced (Sh-Pdia3) cells, 1,25(OH)(2)D(3) failed to stimulate PKC and PGE(2) responses; in Pdia3-overexpressing cells (Ov-Pdia3), responses to 1,25(OH)(2)D(3) were augmented. Downstream mediators of Pdia3, PLA(2)-activating protein (PLAA) and arachidonic acid, stimulated similar PKC activation in wild-type, Sh-Pdia3, and Ov-Pdia3 cells supporting the hypothesis that Pdia3 mediates the membrane action of 1,25(OH)(2)D(3). Treatment of MC3T3-E1 cells with 1,25(OH)(2)D(3) for 9 min stimulated rapid phosphorylation of extracellular signal-regulated kinases 1 and 2 (ERK1/2) and increased expression of alkaline phosphatase, MMP-13, and osteopontin but decreased expression of osteocalcin, osteoprotegerin (mRNA and protein), and smad2. These effects were attenuated in Sh-Pdia3 cells. Sh-Pdia3 cells produced higher numbers of von Kossa-positive nodules and alizarin red-positive nodules compared with WT cells with or without 1,25(OH)(2)D(3) treatment whereas Ov-Pdia3 did not show any mineralization. Our data suggest Pdia3 is an important initiator of 1,25(OH)(2)D(3)-stimulated membrane signaling pathways, which have both genomic and non genomic effects during osteoblast maturation.
Collapse
Affiliation(s)
- Jiaxuan Chen
- Wallace H. Coulter Department of Biomedical Engineering, Georgia Institute of Technology, Atlanta, Georgia 30332-0363, USA
| | | | | | | | | | | |
Collapse
|
11
|
The direct role of vitamin D on bone homeostasis. Arch Biochem Biophys 2008; 473:225-30. [PMID: 18424254 DOI: 10.1016/j.abb.2008.03.038] [Citation(s) in RCA: 123] [Impact Index Per Article: 7.2] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 12/21/2007] [Revised: 03/26/2008] [Accepted: 03/27/2008] [Indexed: 12/28/2022]
|
12
|
Böhmer S, Carapito C, Wilzewski B, Leize E, Van Dorsselaer A, Bernhardt R. Analysis of aldosterone-induced differential receptor-independent protein patterns using 2D-electrophoresis and mass spectrometry. Biol Chem 2006; 387:917-29. [PMID: 16913842 DOI: 10.1515/bc.2006.116] [Citation(s) in RCA: 6] [Impact Index Per Article: 0.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/15/2022]
Abstract
In the human body the mineralocorticoid aldosterone is responsible for maintaining water and electrolyte homeostasis and therefore controlling blood pressure. In addition, aldosterone has recently been associated with severe heart failure. Besides receptor-dependent action, the damaging effects of aldosterone may also be partly mediated through non-genomic mechanisms. The present study focuses on the mineralocorticoid receptor-independent action of aldosterone at the protein level. We chose the fission yeast Schizosaccharomyces pombe as a model organism, since this yeast does not contain nuclear steroid receptors, but many genes and regulatory mechanisms that are close to those of mammals. Using 2D-electrophoresis we identified for the first time protein spots affected by aldosterone in a nuclear receptor-free system. Mass spectrometry analysis using MALDI-TOF MS and nanoLC-MS/MS approaches allowed the unambiguous identification of 11 proteins that showed increased or decreased levels, which may represent newly identified players and pathways of aldosterone-induced action. Two proteins with a connection to osmotic regulation (NAD-dependent malic enzyme and glycerol-3-phosphate-dehydrogenase), as well as two proteins involved in the overall organization of the cytoskeleton, vip1 and glyceraldehyde-3-phosphate dehydrogenase, which was also found to be specifically affected by aldosterone in human HCT116 cells, are discussed.
Collapse
Affiliation(s)
- Susanne Böhmer
- Universität des Saarlandes, FR 8.3 Biochemie, Postfach 151150, D-66041 Saarbrücken, Germany
| | | | | | | | | | | |
Collapse
|
13
|
Boyan BD, Wong KL, Wang L, Yao H, Guldberg RE, Drab M, Jo H, Schwartz Z. Regulation of growth plate chondrocytes by 1,25-dihydroxyvitamin D3 requires caveolae and caveolin-1. J Bone Miner Res 2006; 21:1637-47. [PMID: 16995819 DOI: 10.1359/jbmr.060713] [Citation(s) in RCA: 36] [Impact Index Per Article: 1.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 11/18/2022]
Abstract
UNLABELLED We examined the role of caveolae and caveolin-1 in the mechanism of 1alpha,25(OH)(2)D(3) action in growth plate chondrocytes. We found that caveolae are required for rapid 1alpha,25(OH)(2)D(3)-dependent PKC signaling, and caveolin-1 must be present based on studies using chondrocytes from Cav-1(-/-) mice. INTRODUCTION 1,25-dihydroxyvitamin D(3) [1alpha,25(OH)(2)D(3)] regulates endochondral ossification in part through membrane-associated mechanisms, including protein kinase C (PKC) signaling activated by a membrane-associated 1alpha,25(OH)(2)D(3)-binding protein, ERp60. We tested the hypothesis that caveolae are required for 1alpha,25(OH)(2)D(3) action and play an important role in regulating chondrocyte biology and growth plate physiology. MATERIALS AND METHODS Rat costochondral chondrocytes were examined for caveolae by transmission electron microscopy of cultured cells and of cells in situ. Western blots and confocal microscopy were used to detect caveolae proteins including caveolin-1 (Cav-1) and 1alpha,25(OH)(2)D(3) receptors. Caveolae cholesterol was depleted with beta-cyclodextrin (CD) and effects of 1alpha,25(OH)(2)D(3) on PKC, DNA synthesis, alkaline phosphatase, and proteoglycan production determined. Chondrocytes from Cav-1(-/-) and C57BL/6 wildtype mice were also treated with 1alpha,25(OH)(2)D(3). Epiphyses and costochondral junctions of 8-week-old male Cav-1(-/-) and wildtype mice (N = 8) were compared by histomorphometry and microCT. Data were analyzed by ANOVA and Bonferroni for posthoc comparisons. RESULTS Growth zone chondrocytes had caveolae and Cav-1, -2, and -3. Resting zone chondrocytes, which do not exhibit a rapid 1alpha,25(OH)(2)D(3)-dependent increase in PKC activity, also had these caveolins, but caveolae were larger and fewer in number. ERp60 but not VDR co-localized with Cav-1 in plasma membranes and in lipid rafts. CD-treatment blocked 1alpha,25(OH)(2)D(3) effects on all parameters tested. The Cav-1(-/-) cells did not respond to 1alpha,25(OH)(2)D(3), although 1alpha,25(OH)(2)D(3) increased PKC, alkaline phosphatase, and [(35)S]-sulfate incorporation in wildtype C57BL/6 cells. Histology and microCT showed that Cav-1(-/-) growth plates were longer and had more hypertrophic cells in each column. Growth plate changes were reflected in the metaphysis. CONCLUSIONS The membrane-mediated effects of 1alpha,25(OH)(2)D(3) require caveolae and Cav-1, and Cav-1 deficiency results in altered growth plate physiology.
Collapse
Affiliation(s)
- Barbara D Boyan
- Department of Biomedical Engineering, Georgia Institute of Technology, Atlanta, Georgia 30332-0363, USA.
| | | | | | | | | | | | | | | |
Collapse
|
14
|
McMillan J, Fatehi-Sedeh S, Sylvia VL, Bingham V, Zhong M, Boyan BD, Schwartz Z. Sex-specific regulation of growth plate chondrocytes by estrogen is via multiple MAP kinase signaling pathways. BIOCHIMICA ET BIOPHYSICA ACTA-MOLECULAR CELL RESEARCH 2006; 1763:381-92. [PMID: 16713447 DOI: 10.1016/j.bbamcr.2006.02.010] [Citation(s) in RCA: 29] [Impact Index Per Article: 1.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Received: 10/23/2005] [Revised: 02/07/2006] [Accepted: 02/27/2006] [Indexed: 10/24/2022]
Abstract
Both male and female rat growth plate cartilage cells possess estrogen receptors (ERs), but 17beta-estradiol (E(2)) activates protein kinase C (PKC) and PKC-dependent biological responses to E(2) only in cells from female animals. PKC signaling can elicit genomic responses via mitogen activated protein kinase (MAPK) and E(2) has been shown to activate ERK MAPK in many cells, suggesting that MAPK may play a role in growth plate chondrocytes as well. We tested if E(2) increases MAPK activity and if so, whether the response is limited to female cells, if it is PKC-dependent, and if the mechanism involves traditional ER pathways. We also determined the contribution of MAPK to the biological response of growth plate chondrocytes and assessed the relative contributions of ERK, p38 and JNK MAPKs. Female rat costochondral cartilage cells were treated with E(2) and MAPK-specific activity determined in cell layer lysates. The mechanism of MAPK activation was determined by treating the cells with E(2) conjugated to bovine serum albumin (E(2)-BSA) to assess if membrane receptors were involved; stereospecificity was determined using 17alpha-estradiol; PKC and phospholipase C (PLC) dependence was determined using specific inhibitors; and the ER agonist diethylstilbestrol, the ER antagonist ICI 182780, and tamoxifen were used to assess the role of traditional ER pathways. E(2) regulation of ERK1/2 MAPK was assessed and the relative roles of ERK1/2, p38 and JNK MAPKs determined using specific inhibitors. E(2) caused a rapid dose-dependent activation of MAPK that was greatest in cells treated for 9 min with 10(-9) M hormone; activity remained elevated for 3 h. E(2)'s effect on MAPK was stereospecific and comparable to that of E(2)-BSA. It was insensitive to DES and ICI 182780, dependent on PKC and PLC, blocked by tamoxifen and it did not require gene transcription or translation. E(2) had no effect on ERK1 or ERK2 mRNA or protein but it caused a rapid phosphorylation of ERK1/2 at 9 min. Inhibition of ERK1/2 and p38 MAPK reduced the stimulatory effects of E(2) on alkaline phosphatase activity and [(35)S]-sulfate incorporation. These results suggest that E(2) regulates MAPK through a sex-specific membrane-mediated mechanism that does not involve cytosolic ERs in a traditional sense and that ERK1/2 and p38 mediate the downstream biological effects of the hormone.
Collapse
Affiliation(s)
- J McMillan
- Wallace H. Coulter Department of Biomedical Engineering, Georgia Institute of Technology, 315 Ferst Drive NW, Atlanta, 30332, USA
| | | | | | | | | | | | | |
Collapse
|
15
|
Schwartz Z, Graham EJ, Wang L, Lossdörfer S, Gay I, Johnson-Pais TL, Carnes DL, Sylvia VL, Boyan BD. Phospholipase A2 activating protein (PLAA) is required for 1alpha,25(OH)2D3 signaling in growth plate chondrocytes. J Cell Physiol 2005; 203:54-70. [PMID: 15368540 DOI: 10.1002/jcp.20212] [Citation(s) in RCA: 30] [Impact Index Per Article: 1.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/11/2022]
Abstract
Phospholipase A2 (PLA2) is pivotal in the rapid membrane-mediated actions of 1,25-dihydroxyvitamin D3 [1alpha,25(OH)2D3]. Microarray analysis indicated that PLA2 activating protein (PLAA) mRNA is upregulated 6-fold before rat growth plate cells exhibit 1alpha,25(OH)2D3-dependent protein kinase C (PKC) increases, suggesting that it plays an important role in 1alpha,25(OH)2D3's mechanism of action. PLAA mRNA was confirmed in 1alpha,25(OH)2D3-responsive growth zone (prehypertrophic and upper hypertrophic cell zones) chondrocytes by RT-PCR and Northern blot in vitro and by in situ hybridization in vivo. PLAA protein was shown by Western blot and immunohistochemistry. PLAAs role in 1alpha,25(OH)2D3 signaling was evaluated in growth zone cell cultures using PLAA peptide. Arachidonic acid release was increased as was PLA2-specific activity in plasma membranes and matrix vesicles. PKCalpha, but not PKCbeta, PKCepsilon, or PKCzeta, was increased. PLAAs effect was comparable to that of 1alpha,25(OH)2D3 and was additive with 1alpha,25(OH)2D3. PLA2 inhibitors quinacrine and AACOCF3, and cyclooxygenase inhibitor indomethacin blocked the effect of PLAA peptide on PKC, indicating arachidonic acid and its metabolites were involved. This was confirmed using exogenous arachidonic acid. Prostaglandin acted via EP1 based on inhibition by SC19220 and not via EP2 since AH6809 had no effect. Like 1alpha,25(OH)2D3, PLAA peptide also increased activity of phospholipase C-specific activity via beta-1 and beta-3 isoforms, but not delta-1 or gamma-1; the effect of PLAA was via lysophospholipid but not via arachidonic acid. PLAA peptide decreased [3H]-thymidine incorporation to 50% of the decrease caused by 1alpha,25(OH)2D3. In contrast, PLAA peptide increased alkaline phosphatase-specific activity and proteoglycan production in a manner similar to 1alpha,25(OH)2D3. This indicates that PLAA is a specific activator of PLA2 in growth plate chondrocytes, and suggests that it mediates the membrane effect of 1alpha,25(OH)2D3, thereby modulating physiological response.
Collapse
Affiliation(s)
- Z Schwartz
- Georgia Institute of Technology, Atlanta, Georgia 30332, USA
| | | | | | | | | | | | | | | | | |
Collapse
|
16
|
Long F, Wang YX, Liu L, Zhou J, Cui RY, Jiang CL. Rapid nongenomic inhibitory effects of glucocorticoids on phagocytosis and superoxide anion production by macrophages. Steroids 2005; 70:55-61. [PMID: 15610897 DOI: 10.1016/j.steroids.2004.10.004] [Citation(s) in RCA: 73] [Impact Index Per Article: 3.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Received: 07/13/2004] [Revised: 10/04/2004] [Accepted: 10/07/2004] [Indexed: 11/19/2022]
Abstract
Traditionally, steroid hormone effects have been described as a result of the modulation of nuclear transcription, thus triggering genomic events that are responsible for physiological effects. Despite early observations of rapid steroid effects that were incompatible with this theory, nongenomic steroid effects have been widely recognized only recently. However, the nongenomic effect of glucocorticoid (GC) on anti-inflammation and immunosuppression has not been reported. Macrophages play important roles in inflammation and the immune response. The present experiment selected macrophages as experimental cells to explore the nongenomic effects and possible mechanisms of GCs on phagocytosis and superoxide anion production. Phagocytosis by macrophages was detected by the neutral red uptake assay. The superoxide anions were measured by cytochrome C reduction assay. It was found that both 10(-4) and 10(-5) mol/L corticosterone (CORT) rapidly inhibited uptake of neutral red by macrophages in less than 30 min, and the inhibition by the former was stronger than that of the latter. CORT (10(-4) to 10(-10) mol/L) rapidly inhibited superoxide anion production by macrophages in less than 30 min. The above-mentioned effects were insensitive to the GC-receptor antagonist mifepristone (RU486) and the translation inhibitor actidione. CORT coupled to bovine serum albumin (BSA-CORT) was able to mimic the rapid inhibitory effects of CORT. The results indicated that CORT could rapidly inhibit phagocytosis and superoxide anion production by mouse peritoneal macrophages in vitro in less than 30 min by a rapid, nongenomic mechanism, which contributes to the anti-inflammatory and immunosuppressive actions of GCs. These data shed a new light on the clinical application of GCs.
Collapse
Affiliation(s)
- Fei Long
- Department of Nautical Medicine, Second Military Medical University, 800 Xiangyin Road, Shanghai 200433, PR China
| | | | | | | | | | | |
Collapse
|
17
|
Boyan BD, Jennings EG, Wang L, Schwartz Z. Mechanisms regulating differential activation of membrane-mediated signaling by 1alpha,25(OH)2D3 and 24R,25(OH)2D3. J Steroid Biochem Mol Biol 2004; 89-90:309-15. [PMID: 15225791 DOI: 10.1016/j.jsbmb.2004.03.027] [Citation(s) in RCA: 17] [Impact Index Per Article: 0.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 11/30/2022]
Abstract
Vitamin D metabolites 1alpha,25(OH)(2)D(3) and 24R,25(OH)(2)D(3) regulate endochondral ossification in a cell maturation-dependent manner via membrane-mediated mechanisms. 24R,25(OH)(2)D(3) stimulates PKC activity in chondrocytes from the growth plate resting zone, whereas 1alpha,25(OH)(2)D(3) stimulates PKC in growth zone chondrocytes. We used the rat costochondral growth plate cartilage cell model to study how these responses are differentially regulated. 1alpha,25(OH)(2)D(3) acts on PKC, MAP kinase, and downstream physiological responses via phosphatidylinositol-specific PLC-beta; 24R,25(OH)(2)D(3) acts via PLD. In both cases, diacylglycerol (DAG) is increased, activating PKC. Both cell types possess membrane and nuclear receptors for 1alpha,25(OH)(2)D(3), but the mechanisms that render the 1alpha,25(OH)(2)D(3) pathway silent in resting zone cells or the 24R,25(OH)(2)D(3) pathway silent in growth zone cells are unclear. PLA(2) is pivotal in this process. 1alpha,25(OH)(2)D(3) stimulates PLA(2) activity in growth zone cells and 24R,25(OH)(2)D(3) inhibits PLA(2) activity in resting zone cells. Both processes result in PKC activation. To understand how negative regulation of PLA(2) results in increased PKC activity in resting zone cells, we used PLA(2) activating peptide to stimulate PLA(2) activity and examined cell response. PLAP is not expressed in resting zone cells in vivo, supporting the hypothesis that PLA(2) activation is inhibitory to 24R,25(OH)(2)D(3) action in these cells.
Collapse
Affiliation(s)
- B D Boyan
- Wallace H. Coulter Department of Biomedical Engineering, Georgia Institute of Technology and Emory University, 315 Ferst Drive NW, Atlanta, GE 30332, USA.
| | | | | | | |
Collapse
|
18
|
Boyan BD, Sylvia VL, McKinney N, Schwartz Z. Membrane actions of vitamin D metabolites 1?,25(OH)2D3 and 24R,25(OH)2D3 are retained in growth plate cartilage cells from vitamin D receptor knockout mice. J Cell Biochem 2003; 90:1207-23. [PMID: 14635194 DOI: 10.1002/jcb.10716] [Citation(s) in RCA: 63] [Impact Index Per Article: 2.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/08/2022]
Abstract
1alpha,25(OH)(2)D(3) regulates rat growth plate chondrocytes via nuclear vitamin D receptor (1,25-nVDR) and membrane VDR (1,25-mVDR) mechanisms. To assess the relationship between the receptors, we examined the membrane response to 1alpha,25(OH)(2)D(3) in costochondral cartilage cells from wild type VDR(+/+) and VDR(-/-) mice, the latter lacking the 1,25-nVDR and exhibiting type II rickets and alopecia. Methods were developed for isolation and culture of cells from the resting zone (RC) and growth zone (GC, prehypertrophic and upper hypertrophic zones) of the costochondral cartilages from wild type and homozygous knockout mice. 1alpha,25(OH)(2)D(3) had no effect on [(3)H]-thymidine incorporation in VDR(-/-) GC cells, but it increased [(3)H]-thymidine incorporation in VDR(+/+) cells. Proteoglycan production was increased in cultures of both VDR(-/-) and VDR(+/+) cells, based on [(35)S]-sulfate incorporation. These effects were partially blocked by chelerythrine, which is a specific inhibitor of protein kinase C (PKC), indicating that PKC-signaling was involved. 1alpha,25(OH)(2)D(3) caused a 10-fold increase in PKC specific activity in VDR(-/-), and VDR(+/+) GC cells as early as 1 min, supporting this hypothesis. In contrast, 1alpha,25(OH)(2)D(3) had no effect on PKC activity in RC cells isolated from VDR(-/-) or VDR(+/+) mice and neither 1beta,25(OH)(2)D(3) nor 24R,25(OH)(2)D(3) affected PKC in GC cells from these mice. Phospholipase C (PLC) activity was also increased within 1 min in GC chondrocyte cultures treated with 1alpha,25(OH)(2)D(3). As noted previously for rat growth plate chondrocytes, 1alpha,25(OH)(2)D(3) mediated its increases in PKC and PLC activities in the VDR(-/-) GC cells through activation of phospholipase A(2) (PLA(2)). These responses to 1alpha,25(OH)(2)D(3) were blocked by antibodies to 1,25-MARRS, which is a [(3)H]-1,25(OH)(2)D(3) binding protein identified in chick enterocytes. 24R,25(OH)(2)D(3) regulated PKC in VDR(-/-) and VDR(+/+) RC cells. Wild type RC cells responded to 24R,25(OH)(2)D(3) with an increase in PKC, whereas treatment of RC cells from mice lacking a functional 1,25-nVDR caused a time-dependent decrease in PKC between 6 and 9 min. 24R,25(OH)(2)D(3) dependent PKC was mediated by phospholipase D, but not by PLC, as noted previously for rat RC cells treated with 24R,25(OH)(2)D(3). These results provide definitive evidence that there are two distinct receptors to 1alpha,25(OH)(2)D(3). 1alpha,25(OH)(2)D(3)-dependent regulation of DNA synthesis in GC cells requires the 1,25-nVDR, although other physiological responses to the vitamin D metabolite, such as proteoglycan sulfation, involve regulation via the 1,25-mVDR.
Collapse
|
19
|
Boyan BD, Sylvia VL, Dean DD, Del Toro F, Schwartz Z. Differential regulation of growth plate chondrocytes by 1alpha,25-(OH)2D3 and 24R,25-(OH)2D3 involves cell-maturation-specific membrane-receptor-activated phospholipid metabolism. CRITICAL REVIEWS IN ORAL BIOLOGY AND MEDICINE : AN OFFICIAL PUBLICATION OF THE AMERICAN ASSOCIATION OF ORAL BIOLOGISTS 2003; 13:143-54. [PMID: 12097357 DOI: 10.1177/154411130201300205] [Citation(s) in RCA: 61] [Impact Index Per Article: 2.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 11/17/2022]
Abstract
This review discusses the regulation of growth plate chondrocytes by vitamin D(3). Over the past ten years, our understanding of how two vitamin D metabolites, 1alpha,25-(OH)(2)D(3) and 24R,25-(OH)(2)D(3), exert their effects on endochondral ossification has undergone considerable advances through the use of cell biology and signal transduction methodologies. These studies have shown that each metabolite affects a primary target cell within the endochondral developmental lineage. 1alpha,25-(OH)(2)D(3) affects primarily growth zone cells, and 24R,25-(OH)(2)D(3) affects primarily resting zone cells. In addition, 24R,25-(OH)(2)D(3) initiates a differentiation cascade that results in down-regulation of responsiveness to 24R,25-(OH)(2)D(3) and up-regulation of responsiveness to 1alpha,25-(OH)(2)D(3). 1alpha,25-(OH)(2)D(3) regulates growth zone chondrocytes both through the nuclear vitamin D receptor, and through a membrane-associated receptor that mediates its effects via a protein kinase C (PKC) signal transduction pathway. PKCalpha is increased via a phosphatidylinositol-specific phospholipase C (PLC)-dependent mechanism, as well as through the stimulation of phospholipase A(2) (PLA(2)) activity. Arachidonic acid and its downstream metabolite prostaglandin E(2) (PGE(2)) also modulate cell response to 1alpha,25-(OH)(2)D(3). In contrast, 24R,25-(OH)(2)D(3) exerts its effects on resting zone cells through a separate, membrane-associated receptor that also involves PKC pathways. PKCalpha is increased via a phospholipase D (PLD)-mediated mechanism, as well as through inhibition of the PLA(2) pathway. The target-cell-specific effects of each metabolite are also seen in the regulation of matrix vesicles by vitamin D(3). However, the PKC isoform involved is PKCzeta, and its activity is inhibited, providing a mechanism for differential autocrine regulation of the cell and events in the matrix by these two vitamin D(3) metabolites.
Collapse
Affiliation(s)
- B D Boyan
- Departments of Orthopaedics, Periodontics, Biochemistry, and Orthodontics, The University of Texas Health Science Center at San Antonio, 7703 Floyd Curl Drive, MS-7774, San Antonio, TX 78229-3900, USA.
| | | | | | | | | |
Collapse
|
20
|
Schwartz Z, Shaked D, Hardin RR, Gruwell S, Dean DD, Sylvia VL, Boyan BD. 1alpha,25(OH)2D3 causes a rapid increase in phosphatidylinositol-specific PLC-beta activity via phospholipase A2-dependent production of lysophospholipid. Steroids 2003; 68:423-37. [PMID: 12798493 DOI: 10.1016/s0039-128x(03)00044-8] [Citation(s) in RCA: 38] [Impact Index Per Article: 1.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 12/18/2022]
Abstract
1alpha,25(OH)(2)D(3) activates protein kinase C (PKC) in rat growth plate chondrocytes via mechanisms involving phosphatidylinositol-specific phospholipase C (PI-PLC) and phospholipase A(2) (PLA(2)). The purpose of this study was to determine if 1alpha,25(OH)(2)D(3) activates PI-PLC directly or through a PLA(2)-dependent mechanism. We determined which PLC isoforms are present in the growth plate chondrocytes, and determined which isoform(s) of PLC is(are) regulated by 1alpha,25(OH)(2)D(3). Inhibitors and activators of PLA(2) were used to assess the inter-relationship between these two phospholipid-signaling pathways. PI-PLC activity in lysates of prehypertrophic and upper hypertrophic zone (growth zone) cells that were incubated with 1alpha,25(OH)(2)D(3), was increased within 30s with peak activity at 1-3 min. PI-PLC activity in resting zone cells was unaffected by 1alpha,25(OH)(2)D(3). 1beta,25(OH)(2)D(3), 24R,25(OH)(2)D(3), actinomycin D and cycloheximide had no effect on PLC in lysates of growth zone cells. Thus, 1alpha,25(OH)(2)D(3) regulation of PI-PLC enzyme activity is stereospecific, cell maturation-dependent, and nongenomic. PLA(2)-activation (mastoparan or melittin) increased PI-PLC activity to the same extent as 1alpha,25(OH)(2)D(3); PLA(2)-inhibition (quinacrine, oleyloxyethylphosphorylcholine (OEPC), or AACOCF(3)) reduced the effect of 1alpha,25(OH)(2)D(3). Neither arachidonic acid (AA) nor its metabolites affected PI-PLC. In contrast, lysophosphatidylcholine (LPC) and lysophosphatidylethanolamine (LPE) activated PI-PLC (LPE>LPC). 1alpha,25(OH)(2)D(3) stimulated PI-PLC and PKC activities via Gq; GDPbetaS inhibited activity, but pertussis toxin did not. RT-PCR showed that the cells express PLC-beta1a, PLC-beta1b, PLC-beta3 and PLC-gamma1 mRNA. Antibodies to PLC-beta1 and PLC-beta3 blocked the 1alpha,25(OH)(2)D(3) effect; antibodies to PLC-delta and PLC-gamma did not. Thus, 1alpha,25(OH)(2)D(3) regulates PLC-beta through PLA(2)-dependent production of lysophospholipid.
Collapse
Affiliation(s)
- Z Schwartz
- Department of Biomedical Engineering, Georgia Institute of Technology, 315 Ferst Drive, Atlanta, GA 30332, USA
| | | | | | | | | | | | | |
Collapse
|
21
|
Schwartz Z, Ehland H, Sylvia VL, Larsson D, Hardin RR, Bingham V, Lopez D, Dean DD, Boyan BD. 1alpha,25-dihydroxyvitamin D(3) and 24R,25-dihydroxyvitamin D(3) modulate growth plate chondrocyte physiology via protein kinase C-dependent phosphorylation of extracellular signal-regulated kinase 1/2 mitogen-activated protein kinase. Endocrinology 2002; 143:2775-86. [PMID: 12072413 DOI: 10.1210/endo.143.7.8889] [Citation(s) in RCA: 52] [Impact Index Per Article: 2.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 11/19/2022]
Abstract
Membrane-mediated increases in protein kinase C (PKC) activity and PKC-dependent physiological responses of growth plate chondrocytes to vitamin D metabolites depend on the state of endochondral maturation; 1alpha,25-dihydroxyvitamin D(3) [1alpha,25-(OH)(2)D(3)] regulates growth zone (GC) cells, whereas 24R,25-(OH)(2)D(3) regulates resting zone (RC) cells. Different mechanisms, including protein kinase A signaling, mediate the effects of 1alpha,25-(OH)(2)D(3) and 24R,25-(OH)(2)D(3) on PKC, suggesting that different mechanisms may also regulate any MAPK involvement in the physiological responses. This study used confluent cultures of rat costochondral chondrocytes as a model. 1alpha,25-(OH)(2)D(3) stimulated MAPK specific activity in GC in a time- and dose-dependent manner, evident within 9 min. 24R,25-(OH)(2)D(3) stimulated MAPK in RC; increases were dose dependent, occurred after 9 min, and were greatest at 90 min. In both cells the effect was due to ERK1/2 activation (p42 > p44 in GC; p42 = p44 in RC). MAPK activation was dependent on PKC, but not protein kinase A. The effect of 1alpha,25-(OH)(2)D(3) required phospholipase C, and the effect of 24R,25-(OH)(2)D(3) required phospholipase D. Inhibition of cyclooxygenase activity reduced the effect of 1alpha,25-(OH)(2)D(3) on MAPK in GC and enhanced the effect of 24R,25-(OH)(2)D(3) in RC. Based on MAPK inhibition with PD98059, ERK1/2 MAPK mediated the effect of 24R,25-(OH)(2)D(3) on [(3)H]thymidine incorporation and [(35)S]sulfate incorporation by RC, but only partially mediated the effect of 1alpha,25-(OH)(2)D(3) on GC. ERK1/2 was not involved in the regulation of alkaline phosphatase specific activity by either metabolite. This paper supports the hypothesis that 1alpha,25-(OH)(2)D(3) regulates the physiology of GC via rapid membrane-mediated signaling pathways, and some, but not all, of the response to 1alpha,25-(OH)(2)D(3) is via the ERK family of MAPKs. In contrast, 24R,25-(OH)(2)D(3) exerts its effects on RC via PKC-dependent MAPK. Whereas 1alpha,25-(OH)(2)D(3) increases MAPK activity via phospholipase C and increased prostaglandin production, 24R,25-(OH)(2)D(3) increases MAPK via phospholipase D and decreased prostaglandin production. The cell specificity, metabolite stereospecificity, and the dependence on PKC argue for the participation of membrane receptors for 1alpha,25-(OH)(2)D(3) and 24R,25-(OH)(2)D(3) in the regulation of ERK1/2 in the growth plate.
Collapse
Affiliation(s)
- Z Schwartz
- Department of Orthopedics, University of Texas Health Science Center, San Antonio 78229, USA
| | | | | | | | | | | | | | | | | |
Collapse
|
22
|
Boyan BD, Sylvia VL, Dean DD, Schwartz Z. Membrane mediated signaling mechanisms are used differentially by metabolites of vitamin D(3) in musculoskeletal cells. Steroids 2002; 67:421-7. [PMID: 11960617 DOI: 10.1016/s0039-128x(01)00178-7] [Citation(s) in RCA: 18] [Impact Index Per Article: 0.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 11/26/2022]
Abstract
1 alpha,25(OH)(2)D(3) and 24R,25(OH)(2)D(3) mediate their effects on chondrocytes and osteoblasts in part through increased activity of protein kinase C (PKC). For both cell types, 1 alpha,25(OH)(2)D(3) exerts its effects primarily on more mature cells within the lineage, whereas 24R,25(OH)(2)D(3) exerts its effects primarily on relatively immature cells. Studies using the rat costochondral cartilage growth plate as a model indicate that the two metabolites increase PKC activity by different mechanisms. In growth zone cells (prehypertrophic/upper hypertrophic cell zones), 1 alpha,25(OH)(2)D(3) causes a rapid increase in PKC that does not involve new gene expression. 1 alpha,25(OH)(2)D(3) binds its membrane receptor (1,25-mVDR), resulting in activation of phospholipase A(2) and the rapid release of arachidonic acid, as well as activation of phosphatidylinositol-specific phospholipase C, resulting in formation of diacylglycerol and inositol-1,4,5-tris phosphate (IP(3)). IP(3) leads to release of intracellular Ca(2+) from the rough endoplasmic reticulum, and together with diacylglycerol, the increased Ca(2+) activates PKC. PKC is then translocated to the plasma membrane, where it initiates a phosphorylation cascade, ultimately phosphorylating the extracellular signal-regulated kinase-1 and -2 (ERK1/2) family of MAP kinases (MAPK). PKC increases are maximal at 9 min, and MAPK increases are maximal at 90 min in these cells. By contrast, 24R,25(OH)(2)D(3) increases PKC through activation of phospholipase D in resting zone cells. Peak production of diacylglycerol via phospholipase D2 is at 90 min, as are peak increases in PKC. Some of the effect is direct on existing plasma membrane PKC, but most is due to new PKC expression; translocation is not involved. Arachidonic acid and its metabolites also play differential roles in the mechanisms, stimulating PKC in growth zone cells and inhibiting PKC in resting zone cells. 24R,25(OH)(2)D(3) decreases phospholipase A(2) activity and prostaglandin production, thereby overcoming this potential inhibitory component, which may account for the delay in the PKC response. Ultimately, ERK1/2 is phosphorylated. PKC-dependent MAPK activity transduces some, but not all, of the physiological responses of each cell type to its respective vitamin D metabolite, suggesting that the membrane receptor(s) and nuclear receptor(s) may function interdependently to regulate proliferation and differentiation of musculoskeletal cells, but different pathways are involved at different stages of phenotypic maturation.
Collapse
Affiliation(s)
- Barbara D Boyan
- Department of Orthopaedics, MC7774, The University of Texas Health Science Center at San Antonio, 7703 Floyd Curl Drive, San Antonio, TX 78229-3900, USA.
| | | | | | | |
Collapse
|
23
|
Schwartz Z, Sylvia VL, Larsson D, Nemere I, Casasola D, Dean DD, Boyan BD. 1alpha,25(OH)2D3 regulates chondrocyte matrix vesicle protein kinase C (PKC) directly via G-protein-dependent mechanisms and indirectly via incorporation of PKC during matrix vesicle biogenesis. J Biol Chem 2002; 277:11828-37. [PMID: 11805100 DOI: 10.1074/jbc.m110398200] [Citation(s) in RCA: 33] [Impact Index Per Article: 1.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
Matrix vesicles are extracellular organelles involved in mineral formation that are regulated by 1alpha,25(OH)(2)D(3). Prior studies have shown that protein kinase C (PKC) activity is involved in mediating the effects of 1alpha,25(OH)(2)D(3) in both matrix vesicles and plasma membranes. Here, we examined the regulation of matrix vesicle PKC by 1alpha,25(OH)(2)D(3) during biogenesis and after deposition in the matrix. When growth zone costochondral chondrocytes were treated for 9 min with 1alpha,25(OH)(2)D(3), PKCzeta in matrix vesicles was inhibited, while PKCalpha in plasma membranes was increased. In contrast, after treatment for 12 or 24 h, PKCzeta in matrix vesicles was increased, while PKCalpha in plasma membranes was unchanged. The effect of 1alpha,25(OH)(2)D(3) was stereospecific and metabolite-specific. Monensin blocked the increase in matrix vesicle PKC after 24 h, suggesting the secosteroid-regulated packaging of PKC. In addition, the 1alpha,25(OH)(2)D(3) membrane vitamin D receptor (1,25-mVDR) was involved, since a specific antibody blocked the 1alpha,25(OH)(2)D(3)-dependent changes in PKC after both long and short treatment times. In contrast, antibodies to annexin II had no effect, and there was no evidence for the presence of the nuclear VDR on Western blots. To investigate the signaling pathways involved in regulating matrix vesicle PKC activity after biosynthesis, matrix vesicles were isolated and then treated for 9 min with 1alpha,25(OH)(2)D(3) in the presence and absence of specific inhibitors. Inhibition of phosphatidylinositol-phospholipase C, phospholipase D, or G(i)/G(s) had no effect. However, inhibition of G(q) blocked the effect of 1alpha,25(OH)(2)D(3). The rapid effect of 1alpha,25(OH)(2)D(3) also involved the 1,25-mVDR. Moreover, arachidonic acid was found to stimulate PKC when added directly to isolated matrix vesicles. These results indicate that matrix vesicle PKC is regulated by 1alpha,25(OH)(2)D(3) at three levels: 1) during matrix vesicle biogenesis; 2) through direct action on the membrane; and 3) through production of other factors such as arachidonic acid.
Collapse
|
24
|
Schwartz Z, Sylvia VL, Guinee T, Dean DD, Boyan BD. Tamoxifen elicits its anti-estrogen effects in growth plate chondrocytes by inhibiting protein kinase C. J Steroid Biochem Mol Biol 2002; 80:401-10. [PMID: 11983487 DOI: 10.1016/s0960-0760(02)00038-9] [Citation(s) in RCA: 19] [Impact Index Per Article: 0.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 11/23/2022]
Abstract
17 beta-Estradiol (E(2)) regulates growth plate cartilage cells via classical nuclear receptor mechanisms, as well as by direct effects on the chondrocyte membrane. These direct effects are stereospecific, causing a rapid increase in protein kinase C (PKC) specific activity, are only found in cells from female rats and are mimicked by E(2)-bovine serum albumin (BSA), which cannot penetrate the cell membrane. E(2) and E(2)-BSA stimulate alkaline phosphatase specific activity and proteoglycan sulfation in female rat costochondral cartilage cell cultures, but traditional nuclear receptors do not appear to be involved. This study examined the effect of the anti-estrogen tamoxifen on these markers of chondrocyte differentiation; the gender-specificity of tamoxifen's effect on PKC, if tamoxifen has an effect on vitamin D metabolite-stimulated PKC, which is mediated via specific membrane receptors (1,25-mVDR; 24,25-mVDR) and whether the effect of tamoxifen is mediated by nuclear estrogen receptors. Tamoxifen dose-dependently inhibited the effect of E(2)-BSA on PKC, alkaline phosphatase and proteoglycan sulfation in confluent cultures of female resting zone (RC) cells and growth zone (GC) (prehypertrophic/upper hypertrophic zones) cells, suggesting that its action is at the membrane and not cell maturation-dependent. Neither the estrogen receptor (ER) antagonist ICI 182780 nor the ER agonist diethylstilbesterol affected E(2) or E(2)-BSA-stimulated PKC in female chondrocytes. Tamoxifen also inhibited the increase in PKC activity due to 1 alpha,25-(OH)(2)D(3) or 24R,25-(OH)(2)D(3) in growth plate cells derived from either female or male rats. Inhibition of PKC by tamoxifen may be a general property of membrane receptors involved in rapid responses to hormones.
Collapse
Affiliation(s)
- Z Schwartz
- Department of Orthopaedics, The University of Texas Health Science Center at San Antonio, San Antonio, TX 78229-3900, USA
| | | | | | | | | |
Collapse
|
25
|
Sunny F, Lakshmy PS, Oommen OV. Rapid action of cortisol and testosterone on lipogenic enzymes in a fresh water fish Oreochromis mossambicus: short-term in vivo and in vitro study. Comp Biochem Physiol B Biochem Mol Biol 2002; 131:297-304. [PMID: 11959013 DOI: 10.1016/s1096-4959(02)00023-4] [Citation(s) in RCA: 24] [Impact Index Per Article: 1.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/17/2022]
Abstract
Rapid action of steroid hormones on lipid metabolism is not reported so far in any vertebrate. The present study was intended to evaluate the quick actions of cortisol and testosterone on enzymes, namely malic enzyme (ME), glucose-6-phosphate dehydrogenase (G6PDH), and isocitrate dehydrogenase (ICDH) in Oreochromis mossambicus. Cortisol and testosterone produced rapid and opposite effects on the lipogenic enzymes studied. Cortisol significantly decreased the activities of ME, G6PDH, as early as 5 min and ICDH as early as 10 min in vitro (10(-6) M), and 30 min in vivo (0.1 microg/g body wt.) whereas the same doses of testosterone significantly stimulated the activity of all enzymes as early as 5 min in vitro and 30 min in vivo. Actinomycin D treatment did not interfere with the inhibiting effect of cortisol on enzyme activities when measured at 10 min in the in vitro system. The transcriptional inhibitor appeared to partially block the effect of cortisol in vivo. The stimulatory effect of testosterone was insensitive to the action of actinomycin D both in vivo and in vitro. These effects appear to be brought about independently of new protein synthesis because the rapid responses occurred within a latent period of 5-30 min and were insensitive to the action of actinomycin D, suggesting a non-genomic action.
Collapse
Affiliation(s)
- Francis Sunny
- Endocrinology and Biochemistry Division, Department of Zoology, University of Kerala, Kariavattom, Trivandrum 695 581, Kerala, India
| | | | | |
Collapse
|
26
|
Boyan BD, Bonewald LF, Sylvia VL, Nemere I, Larsson D, Norman AW, Rosser J, Dean DD, Schwartz Z. Evidence for distinct membrane receptors for 1 alpha,25-(OH)(2)D(3) and 24R,25-(OH)(2)D(3) in osteoblasts. Steroids 2002; 67:235-46. [PMID: 11856547 DOI: 10.1016/s0039-128x(01)00160-x] [Citation(s) in RCA: 55] [Impact Index Per Article: 2.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 11/29/2022]
Abstract
1 alpha,25-(OH)(2)D(3) exerts its effects on chondrocytes and enterocytes via nuclear receptors (1,25-nVDR) and a separate membrane receptor (1,25-mVDR) that activates protein kinase C (PKC). 24R,25-(OH)(2)D(3) also stimulates PKC in chondrocytes, but through other membrane mechanisms. This study examined the hypothesis that osteoblasts possess distinct membrane receptors for 1 alpha,25-(OH)(2)D(3) and 24R,25-(OH)(2)D(3) that are involved in the activation of PKC and that receptor expression varies as a function of cell maturation state. 1 alpha,25-(OH)(2)D(3) stimulated PKC in well differentiated (UMR-106, MC-3T3-E1) and moderately differentiated (ROS 17/2.8) osteoblast-like cells, and in cultures of fetal rat calvarial (FRC) cells and 2T3 cells treated with rhBMP-2 to promote differentiation. 24R,25-(OH)(2)D(3) stimulated PKC in FRC and 2T3 cultures that had not been treated to induce differentiation, and in ROS 17/2.8 cells. MG63 cells, a relatively undifferentiated osteoblast-like cell line, had no response to either metabolite. Ab99, a polyclonal antibody generated to the chick enterocyte 1,25-mVDR, but not a specific antibody to the 1,25-nVDR, inhibited response to 1 alpha,25-(OH)(2)D(3). 1 alpha,25-(OH)(2)D(3) exhibited specific binding to plasma membrane preparations from cells demonstrating a PKC response to this metabolite that is typical of positive cooperativity. Western blots of these membrane proteins reacted with Ab99, and the Ab99-positive protein had an Mr of 64 kDa. There was no cross-reaction with antibodies to the C- or N-terminus of annexin II. The effect of 24,25-(OH)(2)D(3) on PKC was stereospecific; 24S,25-(OH)(2)D(3) had no effect. These results demonstrate that response to 1 alpha,25-(OH)(2)D(3) and 24R,25-(OH)(2)D(3) depends on osteoblast maturation state and suggest that specific and distinct membrane receptors are involved.
Collapse
Affiliation(s)
- Barbara D Boyan
- Department of Orthopaedics, University of Texas Health Science Center at San Antonio, San Antonio, Texas, USA.
| | | | | | | | | | | | | | | | | |
Collapse
|
27
|
Sylvia VL, Del Toro F, Dean DD, Hardin RR, Schwartz Z, Boyan BD. Effects of 1alpha,25-(OH)(2)D(3) on rat growth zone chondrocytes are mediated via cyclooxygenase-1 and phospholipase A(2). JOURNAL OF CELLULAR BIOCHEMISTRY. SUPPLEMENT 2001; Suppl 36:32-45. [PMID: 11455568 DOI: 10.1002/jcb.1072] [Citation(s) in RCA: 22] [Impact Index Per Article: 0.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 11/08/2022]
Abstract
1alpha,25-(OH)(2)D(3) mediates its effects on growth zone chondrocytes via rapid membrane-associated events as well as through traditional nuclear receptor mechanisms. The membrane-associated signaling pathways include rapid production of diacylglycerol and activation of protein kinase C (PKC), as well as activation of phospholipase A(2) (PLA(2)), increased production of arachidonic acid, and increased production of prostaglandins. This study examined the roles of PLA(2) and cyclooxygenase (Cox) in the mechanism of action of 1alpha,25-(OH)(2)D(3) in these cells to determine whether one or both enzymes catalyze the rate limiting step and whether constitutive or inducible Cox is involved. Cultures were incubated with 1alpha,25-(OH)(2)D(3) for 9 min to measure PKC or for 24 h to measure physiological responses ([(3)H]-thymidine incorporation, alkaline phosphatase specific activity, [(35)S]-sulfate incorporation). Based on RT-PCR and Northern blot analysis, growth zone chondrocytes expressed mRNAs for both Cox-1 and Cox-2 and neither Cox was modulated by 1alpha,25-(OH)(2)D(3). To examine the role of Cox, the cultures were also treated with resveratrol (a specific inhibitor of Cox-1), NS-398 (a specific inhibitor of Cox-2), or indomethacin (a general Cox inhibitor). The results showed that Cox-1 inhibition reduced the 1alpha,25-(OH)(2)D(3)-dependent effects on proliferation, differentiation, and matrix production, whereas inhibition of Cox-2 only had an effect on proliferation. The effects of Cox inhibition were not rate limiting, based on experiments in which PLA(2) was activated with melittin or inhibited with quinacrine. However, at least part of the action of 1alpha,25-(OH)(2)D(3) was regulated by metabolism of arachidonic acid to prostaglandins. This supports the hypothesis that 1alpha,25-(OH)(2)D(3) exerts its effects via more than one signaling pathway and that these pathways are interrelated via the modulation of PLA(2) as a rate-limiting step. PKC regulation may occur at multiple stages in the signal transduction cascade. J. Cell. Biochem. Suppl. 36: 32-45, 2001.
Collapse
Affiliation(s)
- V L Sylvia
- Department of Orthopaedics, The University of Texas Health Science Center at San Antonio, San Antonio, TX 78229, USA
| | | | | | | | | | | |
Collapse
|
28
|
Sylvia VL, Del Toro F, Hardin RR, Dean DD, Boyan BD, Schwartz Z. Characterization of PGE(2) receptors (EP) and their role as mediators of 1alpha,25-(OH)(2)D(3) effects on growth zone chondrocytes. J Steroid Biochem Mol Biol 2001; 78:261-74. [PMID: 11595507 DOI: 10.1016/s0960-0760(01)00099-1] [Citation(s) in RCA: 38] [Impact Index Per Article: 1.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 10/18/2022]
Abstract
Growth plate chondrocyte function is modulated by the vitamin D metabolite 1alpha,25-(OH)(2)D(3) via activation of protein kinase C (PKC). In previous studies with cells derived from prehypertrophic and upper hypertrophic zones of rat costochondral cartilage (growth zone cells), inhibition of prostaglandin production with indomethacin caused a decrease in the stimulation of PKC activity, suggesting that changes in prostaglandin levels mediate the 1alpha,25-(OH)(2)D(3)-dependent response in these cells. Growth zone cells also respond to PGE(2) directly, indicating that prostaglandins act as autocrine or paracrine regulators of chondrocyte metabolism in the growth plate. The aim of the present study was to identify which PGE(2) receptor subtypes (EP) mediate the effects of PGE(2) on growth zone cells. Using primers specific for EP1-EP4, reverse transcription-polymerase chain reaction (RT-PCR) amplified EP1 and EP2 cDNA in a RT-dependent manner. In parallel experiments, we used EP subtype-specific agonists to examine the role of EP receptors in 1alpha,25-(OH)(2)D(3)-mediated cell proliferation and differentiation. 17-Phenyl-trinor-PGE(2) (PTPGE(2)), an EP1 agonist, decreased [3H]-thymidine incorporation in a dose-dependent manner and augmented the 1alpha,25-(OH)(2)D(2)-induced inhibition of [3H]-thymidine incorporation. PTPGE(2) also caused significant increases in proteoglycan production, as measured by [35S]-sulfate incorporation, and alkaline phosphatase specific activity. 1alpha,25-(OH)(2)D(3)-induced alkaline phosphatase activity was only slightly stimulated by PTPGE(2). In contrast, 1alpha,25-(OH)(2)D(3)-induced PKC activity was synergistically increased by PTPGE(2), whereas EP1 antagonists SC-19220 and AH6809 inhibited PKC activity in a dose-dependent manner. The EP2, EP3 and EP4 agonists had no effect on the various cell-induced responses measured. EP1 receptor-induced responses were blocked by the phospholipase C inhibitor U73122, and reduced by PKA inhibitors. EP1 receptor-induced PKC activity was insensitive to pertussis toxin or choleratoxin but blocked by the G-protein inhibitor GDPbetaS, suggesting the involvement of G(q). These results suggest that the EP1 receptor subtype mediates various PGE(2)-induced cellular responses in growth zone chondrocytes leading to decreased proliferation and enhanced differentiation, as well as the effect of 1alpha,25-(OH)(2)D(3) on cellular maturation.
Collapse
MESH Headings
- Alkaline Phosphatase/metabolism
- Animals
- Base Sequence
- Calcitriol/pharmacology
- Cell Differentiation/drug effects
- Cell Division/drug effects
- Cells, Cultured
- Chondrocytes/cytology
- Chondrocytes/drug effects
- Chondrocytes/metabolism
- DNA Primers/genetics
- Dinoprostone/metabolism
- Dinoprostone/pharmacology
- Growth Plate/cytology
- Growth Plate/drug effects
- Growth Plate/metabolism
- Protein Kinase C/metabolism
- Proteoglycans/metabolism
- RNA, Messenger/genetics
- RNA, Messenger/metabolism
- Rats
- Receptors, Prostaglandin E/drug effects
- Receptors, Prostaglandin E/genetics
- Receptors, Prostaglandin E/metabolism
- Receptors, Prostaglandin E, EP1 Subtype
- Receptors, Prostaglandin E, EP2 Subtype
- Thymidine/metabolism
Collapse
Affiliation(s)
- V L Sylvia
- Department of Orthopaedics, The University of Texas Health Science Center at San Antonio, Mail Code 7774, 7703 Floyd Curl Drive, 78229-3900, USA
| | | | | | | | | | | |
Collapse
|
29
|
Schwartz Z, Sylvia VL, Luna MH, DeVeau P, Whetstone R, Dean DD, Boyan BD. The effect of 24R,25-(OH)(2)D(3) on protein kinase C activity in chondrocytes is mediated by phospholipase D whereas the effect of 1alpha,25-(OH)(2)D(3) is mediated by phospholipase C. Steroids 2001; 66:683-94. [PMID: 11546556 DOI: 10.1016/s0039-128x(01)00100-3] [Citation(s) in RCA: 43] [Impact Index Per Article: 1.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 11/22/2022]
Abstract
1alpha,25-(OH)(2)D(3) regulates protein kinase C (PKC) activity in growth zone chondrocytes by stimulating increased phosphatidylinositol-specific phospholipase C (PI-PLC) activity and subsequent production of diacylglycerol (DAG). In contrast, 24R,25-(OH)(2)D(3) regulates PKC activity in resting zone (RC) cells, but PLC does not appear to be involved, suggesting that phospholipase D (PLD) may play a role in DAG production. In the present study, we examined the role of PLD in the physiological response of RC cells to 24R,25-(OH)(2)D(3) and determined the role of phospholipases D, C, and A(2) as well as G-proteins in mediating the effects of vitamin D(3) metabolites on PKC activity in RC and GC cells. Inhibition of PLD with wortmannin or EDS caused a dose-dependent inhibition of basal [3H]-thymidine incorporation by RC cells and further increased the inhibitory effect of 24R,25-(OH)(2)D(3). Wortmannin also inhibited basal alkaline phosphatase activity and [35]-sulfate incorporation and decreased the stimulatory effect of 24R,25-(OH)(2)D(3). This inhibitory effect of wortmannin was not seen in cultures treated with the PI-3-kinase inhibitor LY294002, verifying that wortmannin affected PLD. Wortmannin also inhibited basal PKC activity and partially blocked the stimulatory effect of 24R,25-(OH)(2)D(3) on this enzyme activity. Neither inhibition of PI-PLC with U73122, nor PC-PLC with D609, modulated PKC activity. Wortmannin had no effect on basal PLD in GC cells, nor on 1alpha,25-(OH)(2)D(3)-dependent PKC. Inhibition of PI-PLC blocked the 1alpha,25-(OH)(2)D(3)-dependent increase in PKC activity but inhibition of PC-PLC had no effect. Activation of PLA(2) with melittin inhibited basal and 24R,25-(OH)(2)D(3)-stimulated PKC in RC cells and stimulated basal and 1alpha,25-(OH)(2)D(3)-stimulated PKC in GC cells, but wortmannin had no effect on the melittin-induced changes in either cell type. Pertussis toxin modestly increased the effect of 24R,25-(OH)(2)D(3) on PKC, whereas GDPbetaS had no effect, suggesting that PLD2 is the isoform responsible. This indicates that 1alpha,25-(OH)(2)D(3) regulates PKC in GC cells via PI-PLC and PLA(2), but not PC-PLC or PLD, whereas 24R,25-(OH)(2)D(3) regulates PKC in RC cells via PLD2.
Collapse
Affiliation(s)
- Z Schwartz
- Department of Orthopaedics, The University of Texas Health Science Center at San Antonio, 78284, USA
| | | | | | | | | | | | | |
Collapse
|
30
|
Sylvia VL, Walton J, Lopez D, Dean DD, Boyan BD, Schwartz Z. 17 beta-estradiol-BSA conjugates and 17 beta-estradiol regulate growth plate chondrocytes by common membrane associated mechanisms involving PKC dependent and independent signal transduction. J Cell Biochem 2001; 81:413-29. [PMID: 11255224 DOI: 10.1002/1097-4644(20010601)81:3<413::aid-jcb1055>3.0.co;2-m] [Citation(s) in RCA: 60] [Impact Index Per Article: 2.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/26/2022]
Abstract
Nuclear receptors for 17 beta-estradiol (E(2)) are present in growth plate chondrocytes from both male and female rats and regulation of chondrocytes through these receptors has been studied for many years; however, recent studies indicate that an alternative pathway involving a membrane receptor may also be involved in the cell response. E(2) was found to directly affect the fluidity of chondrocyte membranes derived from female, but not male, rats. In addition, E(2) activates protein kinase C (PKC) in a nongenomic manner in female cells, and chelerythrine, a specific inhibitor of PKC, inhibits E(2)-dependent alkaline phosphatase activity and proteoglycan sulfation in these cells, indicating PKC is involved in the signal transduction mechanism. The aims of the present study were: (1) to examine the effect of a cell membrane-impermeable 17 beta-estradiol-bovine serum albumin conjugate (E(2)-BSA) on chondrocyte proliferation, differentiation, and matrix synthesis; (2) to determine the pathway that mediates the membrane effect of E(2)-BSA on PKC; and (3) to compare the action of E(2)-BSA to that of E(2). Confluent, fourth passage resting zone (RC) and growth zone (GC) chondrocytes from female rat costochondral cartilage were treated with 10(-9) to 10(-7) M E(2) or E(2)-BSA and changes in alkaline phosphatase specific activity, proteoglycan sulfation, and [(3)H]-thymidine incorporation measured. To examine the pathway of PKC activation, chondrocyte cultures were treated with E(2)-BSA in the presence or absence of GDP beta S (inhibitor of G-proteins), GTP gamma S (activator of G-proteins), U73122 or D609 (inhibitors of phospholipase C [PLC]), wortmannin (inhibitor of phospholipase D [PLD]) or LY294002 (inhibitor of phosphatidylinositol 3-kinase). E(2)-BSA mimicked the effects of E(2) on alkaline phosphatase specific activity and proteoglycan sulfation, causing dose-dependent increases in both RC and GC cell cultures. Both forms of estradiol inhibited [(3)H]-thymidine incorporation, and the effect was dose-dependent. E(2)-BSA caused time-dependent increases in PKC in RC and GC cells; effects were observed within three minutes in RC cells and within one minute in GC cells. Response to E(2) was more robust in RC cells, whereas in GC cells, E(2) and E(2)-BSA caused a comparable increase in PKC. GDP beta S inhibited the activation of PKC in E(2)-BSA-stimulated RC and GC cells. GTP gamma S increased PKC in E(2)-BSA-stimulated GC cells, but had no effect in E(2)-BSA-stimulated RC cells. The phosphatidylinositol-specific PLC inhibitor U73122 blocked E(2)-BSA-stimulated PKC activity in both RC and GC cells, whereas the phosphatidylcholine-specific PLC inhibitor D609 had no effect. Neither the PLD inhibitor wortmannin nor the phosphatidylinositol 3-kinase inhibitor LY294022 had any effect on E(2)-BSA-stimulated PKC activity in either RC or GC cells. The classical estrogen receptor antagonist ICI 182780 was unable to block the stimulatory effect of E(2)-BSA on PKC. Moreover, the classical receptor agonist diethylstilbestrol (DES) had no effect on PKC, nor did it alter the stimulatory effect of E(2)-BSA. The specificity of the membrane response to E(2) was also demonstrated by showing that the membrane receptor for 1 alpha,25-(OH)(2)D(3) was not involved. These data indicate that the rapid nongenomic effect of E(2)-BSA on PKC activity in RC and GC cells is dependent on G-protein-coupled PLC and support the hypothesis that many of the effects of E(2) involve membrane-associated mechanisms independent of classical estrogen receptors. (c) 2001 Wiley-Liss, Inc.
Collapse
Affiliation(s)
- V L Sylvia
- Department of Orthopaedics, The University of Texas Health Science Center at San Antonio, San Antonio, Texas 78229-3900, USA
| | | | | | | | | | | |
Collapse
|
31
|
Sela J, Gross UM, Kohavi D, Shani J, Dean DD, Boyan BD, Schwartz Z. Primary mineralization at the surfaces of implants. CRITICAL REVIEWS IN ORAL BIOLOGY AND MEDICINE : AN OFFICIAL PUBLICATION OF THE AMERICAN ASSOCIATION OF ORAL BIOLOGISTS 2001; 11:423-36. [PMID: 11132764 DOI: 10.1177/10454411000110040301] [Citation(s) in RCA: 36] [Impact Index Per Article: 1.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 11/17/2022]
Abstract
Osteogenesis around implants is affected by the physical and chemical characteristics of the biomaterials used. The osteoprogenitor cells must migrate to the implant site and synthesize and secrete a mineralizable extracellular matrix. Because this is neo-bone formation, the mechanism by which the cells calcify their matrix involves extracellular organelles called matrix vesicles in a process termed "primary mineralization". Two different methods for assessing the effects of implant materials on primary mineralization are presented in this report. In the first approach, different implant materials used in dentistry and orthopedic surgery were placed in rat tibial bones after marrow ablation. Two groups of implants were used, bone-bonding and non-bonding materials. We examined the effects of the materials on calcification morphometrically by quantitating changes in matrix vesicle morphology and distribution in endosteal tissue around implants as compared with normal endosteal bone healing. In addition, matrix vesicles were isolated from the endosteal tissue around the implant as well as from the contralateral limb and were examined biochemically. The results demonstrated that bone-bonding materials induced a greater increase in matrix vesicle enzyme activity than did non-bonding materials. However, all materials caused changes in matrix vesicles that were different from those seen in normal endosteal bone formation following injury. The effects of implant materials on biochemical markers of mineralization, including specific activities of matrix vesicle alkaline phosphatase and phospholipase A2 and phosphatidylserine content, demonstrated a high correlation with the morphometric observations with regard to enhancement and/or delay of primary mineralization. In the other approach, we used a radioisotopic method to evaluate the effects of implant materials on primary mineralization. This analysis revealed that implants alter bone healing, as shown by the differential uptake of 99mTc and 32P in different bone compartments. Decreased 32P uptake by the organic phase in the presence of bone-bonding implants suggests that cleavage of 99mTcMD32P into its technetium and methylene diphosphonate moieties was inhibited by the presence of the implants. In summary, these approaches to evaluating the effects of materials on primary mineralization demonstrate that the marrow ablation model can easily distinguish between bone-bonding and non-bonding materials. The use of this model can be valuable in the development of new materials.
Collapse
Affiliation(s)
- J Sela
- Division of Oral Pathology, Biomineralization Laboratory, Hebrew University, Hadassah School of Dental Medicine, Jerusalem, Israel
| | | | | | | | | | | | | |
Collapse
|
32
|
Boyan BD, Sylvia VL, Dean DD, Schwartz Z. 24,25-(OH)(2)D(3) regulates cartilage and bone via autocrine and endocrine mechanisms. Steroids 2001; 66:363-74. [PMID: 11179745 DOI: 10.1016/s0039-128x(00)00162-8] [Citation(s) in RCA: 58] [Impact Index Per Article: 2.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 10/18/2022]
Abstract
The purpose of this paper is to summarize recent advances in our understanding of the physiological role of 24(R),25(OH)(2)D(3) in bone and cartilage and its mechanism of action. With the identification of a target cell, the growth plate resting zone (RC) chondrocyte, we have been able to use cell biology methodology to investigate specific functions of 24(R),25(OH)(2)D(3) and to determine how 24(R),25(OH)(2)D(3) elicits its effects. These studies indicate that there are specific membrane-associated signal transduction pathways that mediate both rapid, nongenomic and genomic responses of RC cells to 24(R),25(OH)(2)D(3). 24(R),25(OH)(2)D(3) binds RC chondrocyte membranes with high specificity, resulting in an increase in protein kinase C (PKC) activity. The effect is stereospecific; 24R,25(OH)(2)D(3), but not 24S,25-(OH)(2)D(3), causes the increase, indicating a receptor-mediated response. Phospholipase D-2 (PLD2) activity is increased, resulting in increased production of diacylglycerol (DAG), which in turn activates PKC. 24(R),25(OH)(2)D(3) does not cause translocation of PKC to the plasma membrane, but activates existing PKCalpha. There is a rapid decrease in Ca(2+) efflux, and influx is stimulated. 24(R),25(OH)(2)D(3) also reduces arachidonic acid release by decreasing phospholipase A(2) (PLA(2)) activity, thereby decreasing available substrate for prostaglandin production via the action of cyclooxygenase-1. PGE(2) that is produced acts on the EP1 and EP2 receptors expressed by RC cells to downregulate PKC via protein kinase A, but the reduction in PGE(2) decreases this negative feedback mechanism. Both pathways converge on MAP kinase, leading to new gene expression. One consequence of this is production of new matrix vesicles containing PKCalpha and PKCzeta and an increase in PKC activity. The chondrocytes also produce 24(R),25(OH)(2)D(3), and the secreted metabolite acts directly on the matrix vesicle membrane. Only PKCzeta is directly affected by 24(R),25(OH)(2)D(3) in the matrix vesicles, and activity of this isoform is inhibited. This effect may be involved in the control of matrix maturation and turnover. 24(R),25(OH)(2)D(3) causes RC cells to mature along the endochondral developmental pathway, where they become responsive to 1alpha,25(OH)(2)D(3) and lose responsiveness to 24(R),25(OH)(2)D(3), a characteristic of more mature growth zone (GC) chondrocytes. 1alpha,25(OH)(2)D(3) elicits its effects on GC through different signal transduction pathways than those used by 24(R),25(OH)(2)D(3). These studies indicate that 24(R),25(OH)(2)D(3) plays an important role in endochondral ossification by regulating less mature chondrocytes and promoting their maturation in the endochondral lineage.
Collapse
Affiliation(s)
- B D Boyan
- University of Texas Health Science Center at San Antonio, 78229-3900, USA.
| | | | | | | |
Collapse
|
33
|
Sylvia VL, Schwartz Z, Del Toro F, DeVeau P, Whetstone R, Hardin RR, Dean DD, Boyan BD. Regulation of phospholipase D (PLD) in growth plate chondrocytes by 24R,25-(OH)2D3 is dependent on cell maturation state (resting zone cells) and is specific to the PLD2 isoform. BIOCHIMICA ET BIOPHYSICA ACTA 2001; 1499:209-21. [PMID: 11341968 DOI: 10.1016/s0167-4889(00)00120-8] [Citation(s) in RCA: 32] [Impact Index Per Article: 1.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 10/18/2022]
Abstract
Many of the effects of 1alpha,25-(OH)2D3 and 24R,25-(OH)2D3 on costochondral chondrocytes are mediated by the protein kinase C (PKC) signal transduction pathway. 1alpha,25-(OH)2D3 activates PKC in costochondral growth zone chondrocytes through a specific membrane receptor (1alpha,25-mVDR), involving rapid increases in diacylglycerol via a phospholipase C (PLC)-dependent mechanism. 24R,25-(OH)2D3 activates PKC in resting zone chondrocytes. Although diacylglycerol is increased by 24R,25-(OH)2D3, PLC is not involved, suggesting a phospholipase D (PLD)-dependent mechanism. Here, we show that resting zone and growth zone cells express mRNAs for PLD1a, PLD1b, and PLD2. Both cell types have PLD activity, but levels are higher in resting zone cells. 24R,25-(OH)2D3, but not 24S,25-(OH)2D3 or 1alpha,25-(OH)2D3, stimulates PLD activity in resting zone cells within 3 min via nongenomic mechanisms. Neither 1alpha,25-(OH)2D3 nor 24R,25-(OH)2D3 affected PLD in growth zone cells. Basal and 24R,25-(OH)2D3-stimulated PLD were inhibited by the PLD inhibitors wortmannin and EDS. Inhibition of phosphatidylinositol 3-kinase (PI 3-kinase), PKC, phosphatidylinositol-specific PLC (PI-PLC), and phosphatidylcholine-specific PLC (PC-PLC) had no effect on PLD activity. Thus, 24R,25-(OH)2D3 stimulates PLD, and PI 3-kinase, PI-PLC and PKC are not involved, whereas PLD is required for stimulation of PKC by 24R,25-(OH)2D3. Pertussis toxin, GDPbetaS, and GTPgammaS had no effect on 24R,25-(OH)2D3-dependent PLD when added to cell cultures, indicating that G-proteins are not involved. These data show that PKC activation in resting zone cells is mediated by PLD and suggest that a functional 24R,25-(OH)2D3-mVDR is required. The results also support the conclusion that the 24R,25-(OH)2D3-responsive PLD is PLD2, since this PLD isoform is G-protein-independent.
Collapse
Affiliation(s)
- V L Sylvia
- Department of Orthopaedics, University of Texas Health Science Center at San Antonio, TX 78229-3900, USA
| | | | | | | | | | | | | | | |
Collapse
|
34
|
Zou J, Landy H, Feun L, Xu R, Lampidis T, Wu CJ, Furst AJ, Savaraj N. Correlation of a unique 220-kDa protein with vitamin D sensitivity in glioma cells. Biochem Pharmacol 2000; 60:1361-5. [PMID: 11008130 DOI: 10.1016/s0006-2952(00)00438-x] [Citation(s) in RCA: 14] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/30/2022]
Abstract
We have investigated the antitumor and apoptotic effects of 1, 25-dihydroxyvitamin D(3) (VD(3)) in glioma cell lines and in primary cultures derived from surgical specimens from patients. Our results showed that certain glioma cells underwent apoptosis, whereas others were resistant. In an attempt to search for parameters that dictate VD(3) sensitivity, we discovered a unique 220-kDa protein in glioma cells that were sensitive to VD(3). This protein was not a classical vitamin D receptor (VDR), but was recognized by two different anti-VDR monoclonal antibodies. Furthermore, the level of the 220-kDa protein was inversely correlated with the IC(50) of VD(3) in these glioma cells. This 220-kDa protein was also present in frozen brain tumor samples, and the level of expression appeared to correlate with their corresponding primary cultures. Thus, our findings suggest that this 220-kDa protein may play an important role in determining VD(3) sensitivity in malignant glioma.
Collapse
Affiliation(s)
- J Zou
- Hematology/Oncology Section, Department of Medicine, VA Medical Center, University of Miami School of Medicine, Miami, FL 33125, USA
| | | | | | | | | | | | | | | |
Collapse
|
35
|
Crescioli C, Maggie M, Vannelli GB, Luconi M, Salerno R, Barni T, Gulisano M, Forti G, Serio M. Effect of a vitamin D3 analogue on keratinocyte growth factor-induced cell proliferation in benign prostate hyperplasia. J Clin Endocrinol Metab 2000; 85:2576-83. [PMID: 10902811 DOI: 10.1210/jcem.85.7.6690] [Citation(s) in RCA: 7] [Impact Index Per Article: 0.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 11/19/2022]
Abstract
Prostate enlargement and function is under the dual control of androgens and intraprostatic growth factors. They regulate, in concert, prostate cell proliferation and apoptosis. An increased signaling of both growth factors and androgens are supposed to underlie benign prostate hyperplasia (BPH), one of the more common disorders of the aging male. Since, in clinical practice, androgen ablation resulted in a rather limited decrease in prostate volume, therapeutic strategies targeting intraprostatic growth factors are emerging. The activated form of vitamin D, vitamin D3, and some of its analogues have been described as potent regulators of cell growth and differentiation. In this study, we report the effects of one of these vitamin D3 analogues, 1,25-dihydroxy-16ene-23yne D3, or analogue (V), on the fate of isolated epithelial cells derived from patients with BPH. We essentially found that analogue (V), as well as vitamin D3, inhibited BPH cell proliferation and counteracted the mitogenic activity of a potent growth factor for BPH cells, such as keratinocyte growth factor (KGF). Moreover, analogue (V) induced bcl-2 protein expression, intracellular calcium mobilization, and apoptosis in both unstimulated and KGF-stimulated BPH cells. Since a short-term (5-min) incubation with analogue (V) reduced the KGF-induced tyrosine phosphorylation of a 120-kDA protein, corresponding to the KGF receptor, a rapid and direct cross-talk between these two molecules is suggested. Such a rapid effect of analogue (V), together with the transient induction of intracellular calcium waves, seems to indicate the partial involvement of a membrane, nongenomic receptor for vitamin D3. In conclusion, we demonstrated the antiproliferative and proapoptotic effect of analogue (V) in BPH cells and speculated on its possible use in the therapy of BPH.
Collapse
Affiliation(s)
- C Crescioli
- Department of Clinical Physiopathology, University of Florence, Italy
| | | | | | | | | | | | | | | | | |
Collapse
|
36
|
Sylvia VL, Boyan BD, Dean DD, Schwartz Z. The membrane effects of 17beta-estradiol on chondrocyte phenotypic expression are mediated by activation of protein kinase C through phospholipase C and G-proteins. J Steroid Biochem Mol Biol 2000; 73:211-24. [PMID: 11070350 DOI: 10.1016/s0960-0760(00)00078-9] [Citation(s) in RCA: 41] [Impact Index Per Article: 1.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 10/18/2022]
Abstract
Growth plate chondrocytes from both male and female rats have nuclear receptors for 17beta-estradiol (E(2)); however, recent studies indicate that an alternative pathway involving a membrane receptor may also be involved in the female cell response. E(2) directly affects the fluidity of chondrocyte membranes derived from female, but not male, rats. In addition, E(2) activates PKC in a nongenomic manner in female cells, and chelerythrine, a specific inhibitor of PKC, inhibits E(2)-dependent alkaline phosphatase activity in these cells, indicating PKC is involved in the signal transduction mechanism. The aims of this study were: (1) to examine if PKC mediates the effect of E(2) on chondrocyte proliferation, differentiation, and matrix synthesis; and (2) to determine the pathway that mediates the membrane effect of E(2) on PKC. Confluent, fourth passage resting zone (RC) and growth zone (GC) chondrocytes from female rat costochondral cartilage were treated with 10(-10) to 10(-7) M E(2) in the presence or absence of the PKC inhibitor chelerythrine, and changes in alkaline phosphatase specific activity, proteoglycan sulfation, and [3H]thymidine incorporation were measured. To examine the pathway of PKC activation, chondrocyte cultures were treated with E(2) in the presence or absence of genistein (an inhibitor of tyrosine kinases), U73122 or D609 (inhibitors of phospholipase C [PLC]), quinacrine (an inhibitor of phospholipase A(2) [PLA(2)]), and melittin (an activator of PLA(2)). Alkaline phosphatase specific activity and proteoglycan sulfation were increased and [3H]thymidine incorporation was decreased by E(2). The effects of E(2) on all parameters were blocked by chelerythrine. Treatment of the cultures with E(2) produced a significant dose-dependent increase in PKC. U73122 dose-dependently inhibited the activation of PKC in E(2)-stimulated female chondrocyte cultures. However, the classical receptor antagonist ICI 182780 was unable to block the stimulatory effect of E(2) on PKC. Moreover, the classical receptor agonist diethylstilbestrol (DES) had no effect on PKC, nor did it alter the stimulatory effect of E(2). Inhibition of tyrosine kinase and PLA(2) had no effect on the activation of PKC by E(2). The PLA(2) activator also had no effect on PKC activation by E(2). E(2) stimulated PKC activity in membranes isolated from the chondrocytes, demonstrating a direct membrane effect for this steroid hormone. These data indicate that the rapid nongenomic effect of E(2) on PKC activity in chondrocytes from female rats is sex-specific and dependent upon a G-protein-coupled phospholipase C.
Collapse
Affiliation(s)
- V L Sylvia
- Department of Orthopaedics, The University of Texas Health Science Center at San Antonio, 7703 Floyd Carl Drive, San Antonio, TX 78284-7774, USA
| | | | | | | |
Collapse
|
37
|
Sylvia VL, Schwartz Z, Dean DD, Boyan BD. Transforming growth factor-beta1 regulation of resting zone chondrocytes is mediated by two separate but interacting pathways. BIOCHIMICA ET BIOPHYSICA ACTA 2000; 1496:311-24. [PMID: 10771099 DOI: 10.1016/s0167-4889(00)00030-6] [Citation(s) in RCA: 16] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 11/23/2022]
Abstract
Previous studies have shown that transforming growth factor-beta1 (TGF-beta1) stimulates protein kinase C (PKC) via a mechanism that is independent of phospholipase C or tyrosine kinase, but involves a pertussis toxin-sensitive G-protein. Maximal activation occurs at 12 h and requires new gene expression. To understand the signaling pathways involved, resting zone chondrocytes were incubated with TGF-beta1 and PKC activity was inhibited with chelerythrine, staurosporine or H-7. [(35)S]Sulfate incorporation was inhibited, indicating that PKC mediates the effects of TGF-beta1 on matrix production. However, there was little, if any, effect on TGF-beta1-dependent increases in [(3)H]thymidine incorporation, and TGF-beta1-stimulated alkaline phosphatase was unaffected, indicating that these responses to the growth factor are not regulated via PKC. TGF-beta1 caused a dose-dependent increase in prostaglandin E(2) (PGE(2)) production which was further increased by PKC inhibition. The increase was regulated by TGF-beta1-dependent effects on phospholipase A(2) (PLA(2)). Activation of PLA(2) inhibited TGF-beta1 effects on PKC, and inhibition of PLA(2) activated TGF-beta1-dependent PKC. Exogenous arachidonic acid also inhibited TGF-beta1-dependent increases in PKC. The effects of TGF-beta1 on PKC involve genomic mechanisms, but not regulation of existing membrane-associated enzyme, since no direct effect of the growth factor on plasma membrane or matrix vesicle PKC was observed. These results support the hypothesis that TGF-beta1 modulates its effects on matrix production through PKC, but its effects on alkaline phosphatase are mediated by production of PGE(2) and protein kinase A (PKA). Inhibition of PKA also decreases TGF-beta1-dependent proliferation. We have previously shown that PGE(2) stimulates alkaline phosphatase through its EP2 receptor, whereas EP1 signaling causes a decrease in PKC. Thus, there is cross-talk between the two pathways.
Collapse
Affiliation(s)
- V L Sylvia
- Department of Orthopaedics, The University of Texas Health Science Center at San Antonio, 7703 Floyd Curl Drive, San Antonio, TX, USA
| | | | | | | |
Collapse
|
38
|
Schwartz Z, Sylvia VL, Del Toro F, Hardin RR, Dean DD, Boyan BD. 24R,25-(OH)(2)D(3) mediates its membrane receptor-dependent effects on protein kinase C and alkaline phosphatase via phospholipase A(2) and cyclooxygenase-1 but not cyclooxygenase-2 in growth plate chondrocytes. J Cell Physiol 2000; 182:390-401. [PMID: 10653606 DOI: 10.1002/(sici)1097-4652(200003)182:3<390::aid-jcp10>3.0.co;2-t] [Citation(s) in RCA: 31] [Impact Index Per Article: 1.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/09/2023]
Abstract
Recent studies have shown that 24R,25-(OH)(2)D(3) mediates its effects on growth plate chondrocytes via membrane receptors. This study examined the roles of phospholipase A(2) (PLA(2)) and cyclooxygenase (Cox) in the mechanism of action of 24R, 25-(OH)(2)D(3) in resting zone chondrocytes in order to determine whether the activity of one or both enzymes provides a regulatory checkpoint in the signaling pathway resulting in increased protein kinase C (PKC) activity. We also determined whether constitutive or inducible Cox is involved. Cultures were incubated with 24R, 25-(OH)(2)D(3) for 90 min to measure PKC or for 24 h to measure physiological responses ([(3)H]-thymidine incorporation, alkaline phosphatase-specific activity, [(35)S]-sulfate incorporation). Based on RT-PCR and Northern blot analysis, resting zone chondrocytes express mRNAs for both Cox-1 and Cox-2. Levels of mRNA for both proteins were unchanged from control levels after a 24-h incubation with 24R,25-(OH)(2)D(3). To examine the role of Cox, the cultures were also treated with resveratrol (a specific inhibitor of Cox-1), NS-398 (a specific inhibitor of Cox-2), or indomethacin (a general Cox inhibitor). Cox-1 inhibition resulted in effects on proliferation, differentiation, and matrix production typical of 24R, 25-(OH)(2)D(3). In contrast, inhibition of Cox-2 had no effect, indicating that 24R,25-(OH)(2)D(3) exerts its effects via Cox-1. Inhibition of Cox-1 also blocked 24R,25-(OH)(2)D(3)-dependent increases in PKC. Activation of PLA(2) with melittin inhibited 24R, 25-(OH)(2)D(3)-dependent stimulation of PKC, and inhibition of PLA(2) with quinacrine stimulated PKC in response to 24R, 25-(OH)(2)D(3). Inclusion of resveratrol reduced the melittin-dependent inhibition of PLA(2) and caused an increase in quinacrine-stimulated PLA(2) activity. Metabolism of arachidonic acid to leukotrienes is not involved in the response to 24R, 25-(OH)(2)D(3) because inhibition of lipoxygenase had no effect. The effect of 24R,25-(OH)(2)D(3) was specific because 24S,25-(OH)(2)D(3), the biologically inactive stereoisomer, failed to elicit a response from the cells. These results support the hypothesis that 24R, 25-(OH)(2)D(3) exerts its effects via more than one signaling pathway and that these pathways are interrelated via the modulation of PLA(2). PKC regulation may occur at multiple stages in the signal transduction cascade.
Collapse
Affiliation(s)
- Z Schwartz
- Department of Orthopaedics, The University of Texas Health Science Center at San Antonio, San Antonio, Texas 78229-3900, USA
| | | | | | | | | | | |
Collapse
|
39
|
Del Toro F, Sylvia VL, Schubkegel SR, Campos R, Dean DD, Boyan BD, Schwartz Z. Characterization of prostaglandin E(2) receptors and their role in 24,25-(OH)(2)D(3)-mediated effects on resting zone chondrocytes. J Cell Physiol 2000; 182:196-208. [PMID: 10623883 DOI: 10.1002/(sici)1097-4652(200002)182:2<196::aid-jcp8>3.0.co;2-e] [Citation(s) in RCA: 34] [Impact Index Per Article: 1.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/08/2022]
Abstract
Resting zone chondrocyte differentiation is modulated by the vitamin D metabolite, 24,25-(OH)(2)D(3), via activation of protein kinase C (PKC). In previous studies, inhibition of prostaglandin production with indomethacin caused an increase in PKC activity, suggesting that changes in prostaglandin levels may mediate the 24, 25-(OH)(2)D(3)-dependent response and act as autocrine or paracrine regulators of chondrocyte metabolism. Supporting this hypothesis is the fact that resting zone cells respond directly to prostaglandin E(2) (PGE(2)). The aim of the present study was to identify which PGE(2) receptor subtypes (EP) mediate the effects of PGE(2) on resting zone cells. Using primers specific for EP1-EP4, reverse transcription-polymerase chain reaction (RT-PCR) amplified EP1 and EP2 cDNA in a RT-dependent manner. A variant form of the EP1 cDNA, EPlv, was also amplified in an RT-dependent manner. In parallel experiments, we used EP subtype-specific agonists to examine the role of EP receptors in 24,25-(OH)(2)D(3)-mediated cell proliferation and differentiation. 17-phenyl-trinor-PGE(2) (PTPGE(2)), an EP1 agonist, increased [(3)H]-thymidine incorporation in a dose-dependent manner and reversed the 24, 25-(OH)(2)D(2)-induced inhibition of [(3)H]-thymidine incorporation. SC-19220, an EP1 antagonist, caused a further dose-dependent decrease in 24,25-(OH)(2)D(3)-induced inhibition of [(3)H]-thymidine incorporation. PTPGE(2) also caused a biphasic increase in [(35)S]-sulfate incorporation and increased alkaline phosphatase enzyme activity at high concentrations (10(-8) M). 24, 25-(OH)(2)D(3)-induced alkaline phosphatase activity was synergistically stimulated in a dose-dependent manner by PTPGE(2). In contrast, 24,25-(OH)(2)D(3)-induced PKC activity was inhibited in a dose-dependent manner by PTPGE(2) and SC-19220, the EP1 antagonist, elevated PKC activity at high concentrations (10(-8) M). The EP2 agonist, misoprostol, only affected [(35)S]-sulfate incorporation, but in a dose-dependent manner. The EP3 and EP4 agonists had no effect on cell response. These results suggest that the EP1 receptor subtype mediates some of the PGE(2)-induced cellular responses in resting zone cells that lead to both increased proliferation and differentiation. Because 24,25-(OH)(2)D(3) inhibits PGE(2) synthesis in these cells, EP1-mediated induction of proliferation is blocked, encouraging cellular maturation and activation of PKC activity.
Collapse
Affiliation(s)
- F Del Toro
- Department of Orthodontics, The University of Texas Health Science Center at San Antonio, San Antonio, Texas
| | | | | | | | | | | | | |
Collapse
|
40
|
Lohmann CH, Sagun R, Sylvia VL, Cochran DL, Dean DD, Boyan BD, Schwartz Z. Surface roughness modulates the response of MG63 osteoblast-like cells to 1,25-(OH)(2)D(3) through regulation of phospholipase A(2) activity and activation of protein kinase A. JOURNAL OF BIOMEDICAL MATERIALS RESEARCH 1999; 47:139-51. [PMID: 10449625 DOI: 10.1002/(sici)1097-4636(199911)47:2<139::aid-jbm4>3.0.co;2-2] [Citation(s) in RCA: 68] [Impact Index Per Article: 2.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 01/15/2023]
Abstract
Implant surface roughness influences osteoblast proliferation, differentiation, and local factor production. Moreover, the responsiveness of osteoblasts to systemic hormones such as 1, 25-(OH)(2)D(3) is altered by the effects of surface roughness; on the roughest Ti surfaces the effects of roughness and 1, 25-(OH)(2)D(3) are synergistic. Prostaglandin E(2) (PGE(2)) appears to be involved in mediating the effects of surface roughness on the cells, as well as in the response to 1,25-(OH)(2)D(3). However, it is not yet known through which signaling pathways surface roughness exerts its effects on the response of osteoblasts to 1, 25-(OH)(2)D(3). The present study examined the potential role of protein kinase A (PKA), phospholipase A(2)(PLA(2)), and protein kinase C (PKC) in this process. MG63 osteoblast-like human osteosarcoma cells were cultured on cpTi disks with R(a) values of 0. 54 microm (PT), 4.14 microm (SLA), or 4.92 microm (TPS). PKA was inhibited by adding H8 to the cultures; similarly, PLA(2) was inhibited with quinacrine or activated with melittin, and PKC was inhibited with chelerythrine. Inhibitors or activators were included in the culture media through the entire culture period or for the last 24 h of culture. In addition, cultures were treated for 24 h with inhibitors or activators in the presence of 1,25-(OH)(2)D(3). The effects on cell number and alkaline phosphatase specific activity were determined after 24 h; PKC activity was determined after 9 min and at 24 h. Cell number was reduced on rough surfaces, and alkaline phosphatase activity was increased. 1,25-(OH)(2)D(3) had a synergistic effect with surface roughness on alkaline phosphatase. However, neither surface roughness nor 1,25-(OH)(2)D(3) had an effect on PKC. H8 treatment for 24 h inhibited cell number and alkaline phosphatase on all surfaces; however, when it was present throughout the culture period, the PKA inhibitor had no effect on cell number, but decreased alkaline phosphatase-specific activity. H8 reduced the 1,25-(OH)(2)D(3)-mediated effect on cell number and alkaline phosphatase. Quinacrine inhibited cell proliferation and alkaline phosphatase on all surfaces and further reduced the 1,25-(OH)(2)D(3)-dependent decreases in both parameters. Melittin had no effect when applied for 24 h and did not modify the 1,25-(OH)(2)D(3) effect; however, when present throughout the culture period, it caused a decrease in proliferation and an increase in enzyme activity. Chelerythrine, the PKC inhibitor, only inhibited cell proliferation when it was present throughout the entire culture period. However, it decreased alkaline phosphatase in cultures treated for 24 h, but increased enzyme activity when it was present for the entire culture period. The results indicate that surface roughness and 1,25-(OH)(2)D(3) both mediate their effects through PLA(2) which catalyzes the rate-limiting step in PGE(2) production. Further downstream, PGE(2) activates PKA. Surface roughness-dependent effects are also mediated through PKC, but only after the cells have reached confluence and are undergoing phenotypic maturation. The effect of surface roughness on responsiveness to 1,25-(OH)(2)D(3) is mediated through PLA(2)/PKA and not through PKC.
Collapse
Affiliation(s)
- C H Lohmann
- Department of Orthopaedics, University of Texas Health Science Center, San Antonio, Texas 78284-7774, USA
| | | | | | | | | | | | | |
Collapse
|
41
|
|
42
|
Pedrozo HA, Schwartz Z, Rimes S, Sylvia VL, Nemere I, Posner GH, Dean DD, Boyan BD. Physiological importance of the 1,25(OH)2D3 membrane receptor and evidence for a membrane receptor specific for 24,25(OH)2D3. J Bone Miner Res 1999; 14:856-67. [PMID: 10352093 DOI: 10.1359/jbmr.1999.14.6.856] [Citation(s) in RCA: 76] [Impact Index Per Article: 2.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 11/18/2022]
Abstract
We have recently identified a membrane vitamin D receptor (mVDR) specific for 1,25-dihydroxyvitamin D3 (1,25(OH)2D3) and shown that it mediates the rapid activation of protein kinase C (PKC) in growth zone chondrocytes (GCs). In this study, we examine the role of the 1, 25(OH)2D3-mVDR in chondrocyte physiology and provide evidence for the existence of a specific membrane receptor for 24, 25-dihydroxyvitamin D3 (24,25(OH)2D3-mVDR). Fourth-passage cultures of growth plate chondrocytes at two distinct stages of endochondral development, resting zone (RC) and growth zone (GC) cells, were used to assess the role of the mVDR in cell proliferation, PKC activation, and proteoglycan sulfation. To preclude the involvement of the nuclear vitamin D receptor (nVDR), we used hybrid analogs of 1, 25(OH)2D3 with <0.1% affinity for the nVDR (2a, 1alpha-CH2OH-3beta-25D3; 3a, 1alpha-CH2OH-3beta-20-epi-22-oxa-25D3; and 3b, 1beta-CH2OH-3alpha-20-epi-22-oxa-25D3). To determine the involvement of the mVDR, we used an antibody generated against the highly purified 1,25(OH)2D3 binding protein from chick intestinal basolateral membranes (Ab99). Analog binding to the mVDR was demonstrated by competition with [3H]1,25(OH)2D3 using matrix vesicles (MVs) isolated from cultures of RC and GC cells. Specific recognition sites for 24,25(OH)2D3 in RC MVs were demonstrated by saturation binding analysis. Specific binding of 24,25(OH)2D3 was also investigated in plasma membranes (PMs) from RC and GC cells and GC MVs. In addition, we examined the ability of Ab99 to block the stimulation of PKC by analog 2a in isolated RC PMs as well as the inhibition of PKC by analog 2a in GC MVs. Like 1,25(OH)2D3, analogs 2a, 3a, and 3b inhibit RC and GC cell proliferation. The effect was dose dependent and could be blocked by Ab99. In GC cells, PKC activity was stimulated maximally by analogs 2a and 3a and very modestly by 3b. The effect of 2a and 3a was similar to that of 1, 25(OH)2D3 and was blocked by Ab99, whereas the effect of 3b was unaffected by antibody. In contrast, 2a was the only analog that increased PKC activity in RC cells, and this effect was unaffected by Ab99. Analog 2a had no effect on proteoglycan sulfation in RC cells, whereas analogs 3a and 3b stimulated it and this was not blocked by Ab99. Binding of [3H]1,25(OH)2D3 to GC MVs was displaced completely with 1,25(OH)2D3 and analogs 2a, 3a, and 3b, but 24, 25(OH)2D3 only displaced 51% of the bound ligand. 24,25(OH)2D3 displaced 50% of [3H]1,25(OH)2D3 bound to RC MVs, but 2a, 3a, and 3b displaced <50%. Scatchard analysis indicated specific binding of 24, 25(OH)2D3 to recognition sites in RC MVs with a Kd of 69.2 fmol/ml and a Bmax of 52.6 fmol/mg of protein. Specific binding for 24, 25(OH)2D3 was also found in RC and GC PMs and GC MVs. GC membranes exhibited lower specific binding than RC membranes; MVs had greater specific binding than PMs in both cell types. 2a caused a dose-dependent increase in PKC activity of RC PMs that was unaffected by Ab99; it inhibited PKC activity in GC MVs, and this effect was blocked by Ab99. The results indicate that the 1, 25(OH)2D3 mVDR mediates the antiproliferative effect of 1,25(OH)2D3 on chondrocytes. It also mediates the 1,25(OH)2D3-dependent stimulation of PKC in GC cells, but not the 2a-dependent increase in RC PKC activity, indicating that 24,25(OH)2D3 mediates its effects through a separate receptor. This is supported by the failure of Ab99 to block 2a-dependent stimulation of PKC in isolated PMs. The data demonstrate for the first time the presence of a specific 24, 25(OH)2D3 mVDR in endochondral chondrocytes and show that, although both cell types express mVDRs for 1,25(OH)2D3 and 24,25(OH)2D3, their relative distribution is cell maturation-dependent.
Collapse
Affiliation(s)
- H A Pedrozo
- Department of Orthopaedics, The University of Texas Health Science Center, San Antonio, Texas 78284-7774, USA
| | | | | | | | | | | | | | | |
Collapse
|
43
|
Schwartz Z, Gilley RM, Sylvia VL, Dean DD, Boyan BD. Prostaglandins mediate the effects of 1,25-(OH)2D3 and 24,25-(OH)2D3 on growth plate chondrocytes in a metabolite-specific and cell maturation-dependent manner. Bone 1999; 24:475-84. [PMID: 10321907 DOI: 10.1016/s8756-3282(99)00014-9] [Citation(s) in RCA: 27] [Impact Index Per Article: 1.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 10/17/2022]
Abstract
Prior studies have shown that 1,25-(OH)2D3 stimulates alkaline phosphatase, phospholipase A2 (PLA2), and protein kinase C (PKC)-specific activities, and production of prostaglandin E2 (PGE2) in growth zone chondrocytes. In contrast, 24,25-(OH)2D3 stimulates alkaline phosphatase and PKC-specific activities but inhibits PLA2-specific activity and PGE2 production in resting zone cells. This indicates that different mechanisms are involved in the action of 1,25-(OH)2D3 and 24,25-(OH)2D3 on their respective target cells. In this study, we examined the hypothesis that differential regulation of prostaglandin production modulates the activity of PKC and alkaline phosphatase. To do this, we examined the effect of the cyclooxygenase inhibitor indomethacin (Indo) on alkaline phosphatase, PLA2, and PKC-specific activities in growth plate chondrocytes treated with these two vitamin D metabolites. In addition, we examined whether inhibition of PKC altered PGE2 production. In growth zone cells, Indo inhibited basal alkaline phosphatase and blocked the 1,25-(OH)2D3-dependent increase in alkaline phosphatase. This effect was due to inhibition of both plasma membrane and matrix vesicle alkaline phosphatase. In resting zone cells, Indo increased basal alkaline phosphatase activity in a dose-dependent manner, but it did not further enhance the 24,25-(OH)2D3-dependent stimulation of this enzyme. The effect of Indo was found in both plasma membranes and matrix vesicles. These data indicate that 1,25-(OH)2D3-dependent increases in alkaline phosphatase-specific activity in growth zone cells are mediated through increased prostaglandin production, whereas 24,25-(OH)2D3-mediated changes in enzyme activity in resting zone cells are mediated through decreased prostaglandin production. Regulation of PLA2 by either 1,25-(OH)2D3 or 24,25-(OH)2D3 in their target cells was unaffected by Indo, indicating that the effect of the vitamin D metabolites on this enzyme is not dependent on changes in PGE2 production. The rapid increase in 1,25-(OH)2D3-dependent PKC-specific activity in growth zone cells was inhibited by Indo, whereas there was a potentiation of the effect of 24,25-(OH)2D3 on PKC activity in resting zone cells. In addition, inhibition of PKC blocked the 1,25-(OH)2D3-dependent increase in PGE2 production in growth zone cells and the 24,25-(OH)2D3-dependent decrease in PGE2 production by resting zone cells. These data indicate that prostaglandins are involved in mediating the rapid effects of 1,25-(OH)2D3 on growth zone cells, and contribute to the effects of 24,25-(OH)2D3 on resting zone cells; in both instances, the vitamin D metabolites exert their effects on PKC through changes in arachidonic acid via the action of PLA2. In addition, PKC by itself may mediate the production of PGE2.
Collapse
Affiliation(s)
- Z Schwartz
- Department of Periodontics, University of Texas Health Science Center at San Antonio, 78284-7774, USA
| | | | | | | | | |
Collapse
|
44
|
Grisaru D, Lev-Lehman E, Shapira M, Chaikin E, Lessing JB, Eldor A, Eckstein F, Soreq H. Human osteogenesis involves differentiation-dependent increases in the morphogenically active 3' alternative splicing variant of acetylcholinesterase. Mol Cell Biol 1999; 19:788-95. [PMID: 9858601 PMCID: PMC83935 DOI: 10.1128/mcb.19.1.788] [Citation(s) in RCA: 56] [Impact Index Per Article: 2.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022] Open
Abstract
The extended human acetylcholinesterase (AChE) promoter contains many binding sites for osteogenic factors, including 1,25-(OH)2 vitamin D3 and 17beta-estradiol. In differentiating osteosarcoma Saos-2 cells, both of these factors enhanced transcription of the AChE mRNA variant 3' terminated with exon 6 (E6-AChE mRNA), which encodes the catalytically and morphogenically active E6-AChE isoform. In contrast, antisense oligodeoxynucleotide suppression of E6-AChE mRNA expression increased Saos-2 proliferation in a dose- and sequence-dependent manner. The antisense mechanism of action was most likely mediated by mRNA destruction or translational arrest, as cytochemical staining revealed reduction in AChE gene expression. In vivo, we found that E6-AChE mRNA levels rose following midgestation in normally differentiating, postproliferative fetal chondrocytes but not in the osteogenically impaired chondrocytes of dwarf fetuses with thanatophoric dysplasia. Taken together, these findings suggest morphogenic involvement of E6-AChE in the proliferation-differentiation balance characteristic of human osteogenesis.
Collapse
Affiliation(s)
- D Grisaru
- Department of Biological Chemistry, Life Sciences Institute, The Hebrew University of Jerusalem, Jerusalem 91904, Israel
| | | | | | | | | | | | | | | |
Collapse
|
45
|
Kobayashi T, Pittelkow MR, Warner GM, Squillace KA, Kumar R. Regulation of a novel immediate early response gene, IEX-1, in keratinocytes by 1alpha,25-dihydroxyvitamin D3. Biochem Biophys Res Commun 1998; 251:868-73. [PMID: 9791001 DOI: 10.1006/bbrc.1998.9556] [Citation(s) in RCA: 36] [Impact Index Per Article: 1.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/22/2022]
Abstract
1alpha,25-Dihydroxyvitamin D3 [1alpha,25(OH)2D3] regulates cellular growth and differentiation. We show that in keratinocytes, 1alpha, 25(OH)2D3 reduces concentrations of the messenger RNA of IEX-1, the product of which blocks Fas- or tumor necrosis factor type alpha-induced apoptosis in various cells. In sub-confluent keratinocyte cultures, the addition of 1alpha,25(OH)2D3, in amounts that induce growth arrest, reduces IEX-1 mRNA concentrations. In confluent cells, 1alpha,25(OH)2D3 initially reduces and then increases IEX-1 mRNA concentrations. IEX-1 protein is localized in the nucleus and perinuclear region of keratinocytes. In sub-confluent cells, 1alpha,25(OH)2D3 translocates IEX-1 protein from the nucleus to the perinuclear region and cytoplasm. Since IEX-1 has recently been shown to regulate cell survival and number, we suggest that IEX-1 may play a role in keratinocyte growth and differentiation and that 1alpha,25(OH)2D3 may reduce keratinocyte growth via a reduction in IEX-1 mRNA and a change in the intracellular distribution of IEX-1 protein.
Collapse
Affiliation(s)
- T Kobayashi
- Departments of Medicine, Molecular Biology, Mayo Clinic and Foundation, Rochester, Minnesota, 55905, USA
| | | | | | | | | |
Collapse
|
46
|
Boyan BD, Sylvia VL, Curry D, Chang Z, Dean DD, Schwartz Z. Arachidonic acid is an autocoid mediator of the differential action of 1,25-(OH)2D3 and 24,25-(OH)2D3 on growth plate chondrocytes. J Cell Physiol 1998; 176:516-24. [PMID: 9699504 DOI: 10.1002/(sici)1097-4652(199809)176:3<516::aid-jcp8>3.0.co;2-r] [Citation(s) in RCA: 35] [Impact Index Per Article: 1.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/11/2022]
Abstract
Prior studies have shown that 24,25-(OH)2D3 and 1,25-(OH)2D3 regulate protein kinase C (PKC) in costochondral chondrocytes in a cell maturation-dependent manner, with 1,25-(OH)2D3 affecting primarily growth zone (GC) cells and 24,25-(OH)2D3 affecting primarily resting zone (RC) cells. In addition, 1,25-(OH)2D3 has been shown to increase phospholipase A2 activity in GC, while 24,25-(OH)2D3 has been shown to decrease phospholipase A2 activity in RC. Stimulation of phospholipase A2 in GC caused an increase in PKC, whereas inhibition of phospholipase A2 activity in RC cultures increased both basal and 24,25-(OH)2D3-induced PKC activity, suggesting that phospholipase A2 may play a central role in mediating the effects of the vitamin D metabolites on PKC. To test this hypothesis, RC and GC cells were cultured in the presence and absence of phospholipase A2 inhibitors (quinacrine and oleyloxyethylphosphorylcholine [OEPC]), phospholipase A2 activators (melittin and mastoparan), or arachidonic acid alone or in the presence of the target cell-specific vitamin D metabolite. PKC specific activity in the cell layer was determined as a function of time. Phospholipase A2 inhibitors decreased both basal and 1,25-(OH)2D3-induced PKC activity in GC. When phospholipase A2 activity was activated by inclusion of melittin or mastoparan in the cultures, basal PKC activity in RC was reduced, while that in GC was increased. Similarly, melittin and mastoparan decreased 24,25-(OH)2D3-induced PKC activity in RC and increased 1,25-(OH)2D3-induced PKC activity in GC. For both cell types, the addition of arachidonic acid to the culture media produced an effect on PKC activity that was similar to that observed when phospholipase A2 activators were added to the cells. These results demonstrate that vitamin D metabolite-induced changes in phospholipase A2 activity are directly related to changes in PKC activity. Similarly, exogenous arachidonic acid affects PKC in a manner consistent with activation of phospholipase A2. These effects are cell maturation- and time-dependent and metabolite-specific.
Collapse
Affiliation(s)
- B D Boyan
- Department of Orthopaedics, The University of Texas Health Science Center at San Antonio, 78284-7774, USA.
| | | | | | | | | | | |
Collapse
|
47
|
Thanh XD, Massicot F, Do B, Breget R, Durand D, Warnet JM, Claude JR, Taniguchi N, Clot JP. Effects of 24R,25-dihydroxyvitamin D3 on alkaline phosphatase activity in pig renal epithelial LLC-PK1 cells in culture. GENERAL PHARMACOLOGY 1998; 31:459-62. [PMID: 9703220 DOI: 10.1016/s0306-3623(98)00011-1] [Citation(s) in RCA: 0] [Impact Index Per Article: 0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 11/18/2022]
Abstract
1. The effects of 24R,25-dihydroxyvitamin D3 [24,25(OH)2D3] on alkaline phosphatase activity (ALP) were evaluated in pig kidney LLC-PK1 cells in culture. 2. The vitamin D3 metabolite increased ALP activity in these cells, whereas no effect of the hormone was observed on gamma-glutamyltranspeptidase and acid phosphatase activities. 3. ALP activity was stimulated after 3- to 12-hr incubation in the presence of 10(-9) mol/l 24,25(OH)2D3 with a maximum after 6 hr. 4. The hormonal induction of ALP activity was prevented by pretreatment of cells by actinomycin D. 5. It is proposed that 24,25(OH)2D3 could increase ALP activity by de novo protein synthesis.
Collapse
Affiliation(s)
- X D Thanh
- Laboratoire de Pharmacologie/CHU-UFR de Pharmacie, Université d'Angers, France
| | | | | | | | | | | | | | | | | |
Collapse
|
48
|
Nemere I, Schwartz Z, Pedrozo H, Sylvia VL, Dean DD, Boyan BD. Identification of a membrane receptor for 1,25-dihydroxyvitamin D3 which mediates rapid activation of protein kinase C. J Bone Miner Res 1998; 13:1353-9. [PMID: 9738506 DOI: 10.1359/jbmr.1998.13.9.1353] [Citation(s) in RCA: 132] [Impact Index Per Article: 4.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 11/18/2022]
Abstract
This paper is the first definitive report demonstrating a unique membrane receptor for 1,25-dihydroxyvitamin D3 (1,25(OH)2D3) which mediates the rapid and nongenomic regulation of protein kinase C (PKC). Previous studies have shown that 1,25(OH)2D3 exerts rapid effects on chondrocyte membranes which are cell maturation-specific, do not require new gene expression, and do not appear to act via the traditional vitamin D receptor. We used antiserum generated to a [3H]1,25(OH)2D3 binding protein isolated from the basal lateral membrane of chick intestinal epithelium (Ab99) to determine if rat costochondral resting zone (RC) or growth zone (GC) cartilage cells contain a similar protein and if cell maturation-dependent differences exist. Immunohistochemistry demonstrated that both RC and GC cells express the protein, but levels are highest in GC. The binding protein is present in both plasma membranes and matrix vesicles and has a molecular weight of 66,000 Da. The 66 kDa protein in GC matrix vesicles has a Kd of 17.2 fmol/ml and Bmax of 124 fmol/mg of protein for [3H]1,25(OH)2D3. In contrast, the 66 kDa protein in RC matrix vesicles has a Kd of 27.7 fmol/ml and a Bmax of 100 fmol/mg of protein. Ab99 blocks the 1,25(OH)2D3-dependent increase in PKC activity in GC chondrocytes, indicating that the 1,25(OH)2D3-binding protein is indeed a receptor, linking ligand recognition to biologic function.
Collapse
Affiliation(s)
- I Nemere
- Utah State University, Logan, USA
| | | | | | | | | | | |
Collapse
|
49
|
Sylvia VL, Hughes T, Dean DD, Boyan BD, Schwartz Z. 17beta-estradiol regulation of protein kinase C activity in chondrocytes is sex-dependent and involves nongenomic mechanisms. J Cell Physiol 1998; 176:435-44. [PMID: 9648931 DOI: 10.1002/(sici)1097-4652(199808)176:2<435::aid-jcp22>3.0.co;2-0] [Citation(s) in RCA: 40] [Impact Index Per Article: 1.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/14/2022]
Abstract
17Beta-estradiol (E2) regulates growth plate chondrocyte differentiation in both a sex- and cell maturation-dependent manner, and the sex-specific effects of E2 appear to be mediated in part by membrane events. In this study, we examined whether E2 regulates protein kinase C (PKC) in a cell-maturation and sex-specific manner and whether E2 uses a nongenomic mechanism in regulating this enzyme. In addition, we determined if PKC mediates the E2-dependent stimulation of alkaline phosphatase activity seen in chondrocytes. Confluent, fourth passage resting zone (RC) and growth zone (GC) chondrocytes from male and female rat costochondral cartilage were treated with 10(-10) to 10(-7) M E2. E2 caused a dose-dependent increase in PKC in RC and GC cells from female rats. Peak stimulation was at 90 min. Increased PKC was evident by 3 min in both RC and GC and was still evident in RC cells at 720 min, but in GC cells activity returned to baseline by 270 min. Actinomycin D had no effect at 9, 90, 270, or 720 min, but there was a small decrease in E2-stimulated PKC in RC treated with cycloheximide at 90 and 270 min and in GC treated for 90 min. E2 increased cytosolic and membrane PKC at 9 min and by 90 min promoted translocation of PKC activity from the cytosol to the membranous compartment of female RC cells. Antibodies specific for the alpha, beta, delta, epsilon, and zeta isoforms of PKC revealed that PKCalpha in female GC and RC cells is activated by E2. There was a small, but statistically significant, increase in PKC in male RC cells in response to E2, but it was not dose-dependent, and no effect of E2 was noted in male GC cells. 17Alpha-estradiol, an inactive isomer of E2, did not affect PKC specific activity in RC or GC cells from either female or male rats. Chelerythrine, a specific inhibitor of PKC, inhibited E2-dependent alkaline phosphatase activity, indicating that E2 mediates its rapid effects on alkaline phosphatase via PKC.
Collapse
Affiliation(s)
- V L Sylvia
- Department of Orthopaedics, The University of Texas Health Science Center at San Antonio, 78284-7774, USA
| | | | | | | | | |
Collapse
|
50
|
Affiliation(s)
- J E Ellingsen
- Department of Prosthetic Dentistry and Stomatognathic Physiology, Faculty of Dentistry, University of Oslo, Norway
| |
Collapse
|