Feeds:
Posts
Comments

Archive for the ‘Mg++’ Category

“Minerals in Medicine” –  40 Minerals that are crucial to Human Health and Biomedicine: Exhibit by NIH Clinical Center and The Smithsonian Institution National Museum of Natural History

Reporter: Aviva Lev-Ari, PhD, RN

 

Friday, September 9, 2016

NIH Clinical Center and The Smithsonian Institution partner to launch Minerals in Medicine Exhibition

What

The National Institutes of Health Clinical Center, in partnership with The Smithsonian Institution National Museum of Natural History, will open a special exhibition of more than 40 minerals that are crucial to human health and biomedicine. “Minerals in Medicine” is designed to enthrall and enlighten NIH Clinical Center’s patients, their loved ones, and the NIH community. Media are invited into America’s Research Hospital, the NIH Clinical Center, to experience this unique exhibition during a ribbon cutting ceremony on Monday September 12 at 4pm.

Beyond taking in the minerals’ arresting beauty, spectators can learn about their important role in keeping the human body healthy, and in enabling the creation of life-saving medicines and cutting edge medical equipment that is used in the NIH Clinical Center and healthcare facilities worldwide. The exhibition, which is on an eighteen-month loan from the National Museum of Natural History, includes specimens that were handpicked from the museum’s vast collection by NIH physicians in partnership with Smithsonian Institution geologists. Some of the minerals on display were obtained regionally as they are part of the Maryland and Virginia landscape.

Who

  • John I. Gallin, M.D., Director of the NIH Clinical Center
  • Jeffrey E. Post, Ph.D., Smithsonian Institution National Museum of Natural History, Chair of the Department of Mineral Sciences and Curator of the National Gem and Mineral Collection

When

Monday, September 12, 2016, 4:00 – 5:00 p.m.

Where

NIH Clinical Center (Building 10), 10 Center Drive, Bethesda, MD, 20892; 1st Floor near Admissions

How

RSVP encouraged, but not required, to attend in person. NIH Visitors Map: http://www.ors.od.nih.gov/maps/Pages/NIH-Visitor-Map.aspx

About the NIH Clinical Center: The NIH Clinical Center is the clinical research hospital for the National Institutes of Health. Through clinical research, clinician-investigators translate laboratory discoveries into better treatments, therapies and interventions to improve the nation’s health. More information: http://clinicalcenter.nih.gov.

About the National Institutes of Health (NIH): NIH, the nation’s medical research agency, includes 27 Institutes and Centers and is a component of the U.S. Department of Health and Human Services. NIH is the primary federal agency conducting and supporting basic, clinical, and translational medical research, and is investigating the causes, treatments, and cures for both common and rare diseases. For more information about NIH and its programs, visit www.nih.gov.

SOURCE

https://www.nih.gov/news-events/news-releases/nih-clinical-center-smithsonian-institution-partner-launch-minerals-medicine-exhibition

Read Full Post »

Metabolic Genomics and Pharmaceutics, Vol. 1 of BioMed Series D available on Amazon Kindle

Metabolic Genomics and Pharmaceutics, Vol. 1 of BioMed Series D available on Amazon Kindle

Reporter: Stephen S Williams, PhD

Article ID #180: Metabolic Genomics and Pharmaceutics, Vol. 1 of BioMed Series D available on Amazon Kindle. Published on 8/15/2015

WordCloud Image Produced by Adam Tubman

Leaders in Pharmaceutical Business Intelligence would like to announce the First volume of their BioMedical E-Book Series D:

Metabolic Genomics & Pharmaceutics, Vol. I

SACHS FLYER 2014 Metabolomics SeriesDindividualred-page2

which is now available on Amazon Kindle at

http://www.amazon.com/dp/B012BB0ZF0.

This e-Book is a comprehensive review of recent Original Research on  METABOLOMICS and related opportunities for Targeted Therapy written by Experts, Authors, Writers. This is the first volume of the Series D: e-Books on BioMedicine – Metabolomics, Immunology, Infectious Diseases.  It is written for comprehension at the third year medical student level, or as a reference for licensing board exams, but it is also written for the education of a first time baccalaureate degree reader in the biological sciences.  Hopefully, it can be read with great interest by the undergraduate student who is undecided in the choice of a career. The results of Original Research are gaining value added for the e-Reader by the Methodology of Curation. The e-Book’s articles have been published on the Open Access Online Scientific Journal, since April 2012.  All new articles on this subject, will continue to be incorporated, as published with periodical updates.

We invite e-Readers to write an Article Reviews on Amazon for this e-Book on Amazon.

All forthcoming BioMed e-Book Titles can be viewed at:

http://pharmaceuticalintelligence.com/biomed-e-books/

Leaders in Pharmaceutical Business Intelligence, launched in April 2012 an Open Access Online Scientific Journal is a scientific, medical and business multi expert authoring environment in several domains of  life sciences, pharmaceutical, healthcare & medicine industries. The venture operates as an online scientific intellectual exchange at their website http://pharmaceuticalintelligence.com and for curation and reporting on frontiers in biomedical, biological sciences, healthcare economics, pharmacology, pharmaceuticals & medicine. In addition the venture publishes a Medical E-book Series available on Amazon’s Kindle platform.

Analyzing and sharing the vast and rapidly expanding volume of scientific knowledge has never been so crucial to innovation in the medical field. WE are addressing need of overcoming this scientific information overload by:

  • delivering curation and summary interpretations of latest findings and innovations on an open-access, Web 2.0 platform with future goals of providing primarily concept-driven search in the near future
  • providing a social platform for scientists and clinicians to enter into discussion using social media
  • compiling recent discoveries and issues in yearly-updated Medical E-book Series on Amazon’s mobile Kindle platform

This curation offers better organization and visibility to the critical information useful for the next innovations in academic, clinical, and industrial research by providing these hybrid networks.

Table of Contents for Metabolic Genomics & Pharmaceutics, Vol. I

Chapter 1: Metabolic Pathways

Chapter 2: Lipid Metabolism

Chapter 3: Cell Signaling

Chapter 4: Protein Synthesis and Degradation

Chapter 5: Sub-cellular Structure

Chapter 6: Proteomics

Chapter 7: Metabolomics

Chapter 8:  Impairments in Pathological States: Endocrine Disorders; Stress

                   Hypermetabolism and Cancer

Chapter 9: Genomic Expression in Health and Disease 

 

Summary 

Epilogue

 

 

Read Full Post »

Functional Correlates of Signaling Pathways

Author and Curator: Larry H. Bernstein, MD, FCAP

 

We here move on to a number of specific, key published work on signaling, and look at the possible therapeutic applications to disease states.

Scripps Research Professor Wolfram Ruf and colleagues have identified a key connection between

  • the signaling pathways and the immune system spiraling out of control involving
  • the coagulation system and vascular endothelium that,
  • if disrupted may be a target for sepsis. (Science Daily, Feb 29, 2008).

It may be caused by a bacterial infection that enters the bloodstream, but

  • we now recognize the same cascade not triggered by bacterial invasion.

The acute respiratory distress syndrome (ARDS) has been defined as

  • a severe form of acute lung injury featuring
  • pulmonary inflammation and increased capillary leak.

ARDS is associated with a high mortality rate and accounts for 100,000 deaths annually in the United States. ARDS may arise in a number of clinical situations, especially in patients with sepsis. A well-described pathophysiological model of ARDS is one form of

  • the acute lung inflammation mediated by
  1. neutrophils,
  2. cytokines, and
  3. oxidant stress.

Neutrophils are major effect cells at the frontier of

  • innate immune responses, and they play
  • a critical role in host defense against invading microorganisms.

The tissue injury appears to be related to

  • proteases and toxic reactive oxygen radicals
  • released from activated neutrophils.

In addition, neutrophils can produce cytokines and chemokines that

enhance the acute inflammatory response.

Neutrophil accumulation in the lung plays a pivotal role in the pathogenesis of acute lung injury during sepsis. Directed movement of neutrophils is

  • mediated by a group of chemoattractants,
  • especially CXC chemokines.

Local lung production of CXC chemokines is intensified during experimental sepsis induced by cecal ligation and puncture (CLP).

Sepsis, Multi-organ Dysfunction Syndrome, and Septic Shock: A Conundrum of Signaling Pathways Cascading Out of Control

http://pharmaceuticalintelligence.com/2012/10/13/sepsis-multi-organ-dysfunction-syndrome-and-septic-shock-a-conundrum-of-signaling-pathways-cascading-out-of-control/

Integrins and extracellular matrix in mechanotransduction

ligand binding of integrins

ligand binding of integrins

Integrins are a family of cell surface receptors which

mediate cell–matrix and cell–cell adhesions.

Among other functions they provide an important

mechanical link between the cells external and intracellular environments while

the adhesions that they form also have critical roles in cellular signal-transduction.

Cell–matrix contacts occur at zones in the cell surface where

adhesion receptors cluster and when activated

the receptors bind to ligands in the extracellular matrix.

The extracellular matrix surrounds the cells of tissues and forms the

structural support of tissue which is particularly important in connective tissues.

Cells attach to the extracellular matrix through

specific cell-surface receptors and molecules

including integrins and transmembrane proteoglycans.

The integrin family of αβ heterodimeric receptors act as

cell adhesion molecules

connecting the ECM to the actin cytoskeleton.

The actin cytoskeleton is involved in the regulation of

1.cell motility,

2.cell polarity,

3.cell growth, and

4.cell survival.

The combination of αβ subunits determines

binding specificity and

signaling properties.

Both α and β integrin subunits contain two separate tails, which

penetrate the plasma membrane and possess small cytoplasmic domains which facilitate

the signaling functions of the receptor.

There is some evidence that the β subunit is the principal site for

binding of cytoskeletal and signaling molecules,

whereas the α subunit has a regulatory role. The integrin tails

link the ECM to the actin cytoskeleton within the cell and with cytoplasmic proteins,

such as talin, tensin, and filamin. The extracellular domains of integrin receptors bind the ECM ligands.

binding of integrins depends on ECM divalent cations ch19

binding of integrins depends on ECM divalent cations ch19

integrin coupled to F-actin via linker

integrin coupled to F-actin via linker

http://www.nature.com/nrm/journal/vaop/ncurrent/images/nrm3896-f4.jpg

Schematic of the ‘focal adhesion clutch’ on stiff (a) versus soft (b) extracellular matrix (ECM). In all cases, integrins are coupled to F-actin via linker proteins (for example, talin and vinculin). The linker proteins move backwards (as indicated by the small arrows) as F-actin also moves backwards, under pushing forces from actin polymerization and/or pulling forces from myosin II activity. This mechanism transfers force from actin to integrins, which pull on the ECM. A stiff ECM (a) resists this force so that the bound integrins remain immobile. A compliant matrix (b) deforms under this force (as indicated by the compressed ECM labelled as deformed matrix) so that the bound integrins can also move backwards. Their movement reduces the net loading rate on all the force-bearing elements, which results in altered cellular responses

The ECM is a complex mixture of matrix molecules, including –

  • glycoproteins, collagens, laminins, glycosaminoglycans, proteoglycans,
  • and nonmatrix proteins, – including growth factors

The integrin receptor formed from the binding of α and β subunits is

  • shaped like a globular head supported by two rod-like legs (Figure 1).

Most of the contact between the two subunits occurs in the head region, with

  • the intracellular tails of the subunits forming the legs of the receptor.

Integrin recognition of ligands is not constitutive but

  • is regulated by alteration of integrin affinity for ligand binding.

For integrin binding to ligands to occur

  • the integrin must be primed and activated, both of which involve
  • conformational changes to the receptor.

Linking integrin conformation to function

Figure  Integrin binding to extracellular matrix (ECM). Conformational changes to integrin structure and clustering of subunits which allow enhanced function of the receptor.

Integrins work alongside other proteins such as

cadherins,

immunoglobulin superfamily

cell adhesion molecules,

selectins, and

syndecans

to mediate

cell–cell and

cell–matrix interactions and communication.

Activation of adhesion receptors triggers the formation of matrix contacts in which

bound matrix components,

adhesion receptors,

and associated intracellular cytoskeletal and signaling molecules

form large functional, localized multiprotein complexes.

Cell–matrix contacts are important in a variety of different cell and

tissue properties including

1.embryonic development,

2.inflammatory responses,

3.wound healing,

4.and adult tissue homeostasis.

Integrin extracellular binding activity is regulated from inside the cell and binding to the ECM induces signals that are transmitted into the cell. This bidirectional signaling requires

dynamic,

spatially, and

temporally regulated formation and

disassembly of multiprotein complexes that

form around the short cytoplasmic tails of integrins.

Ligand binding to integrin family members leads to clustering of integrin molecules in the plasma membrane and recruitment of actin filaments and intracellular signaling molecules to the cytoplasmic domain of the integrins. This forms focal adhesion complexes which are able to maintain

not only adhesion to the ECM

but are involved in complex signaling pathways

which include establishing

1.cell polarity,

2.directed cell migration, and

3.maintaining cell growth and survival.

Initial activation through integrin adhesion to matrix recruits up to around 50 diverse signaling molecules

to assemble the focal adhesion complex

which is capable of responding to environmental stimuli efficiently.

Mapping of the integrin

adhesome binding and signaling interactions

a network of 156 components linked together which can be modified by 690 interactions.

Genetic programming occurs with the binding of integrins to the ECM

Signal transduction pathway activation arising from integrin-ECM binding results in

  • changes in gene expression of cells and
  • leads to alterations in cell and tissue function.

Various different effects can arise depending on the

1.cell type,

2.matrix composition, and

3.integrins activated

It has been suggested that integrin-type I collagen interaction is necessary for

  • the phosphorylation and activation of osteoblast-specific transcription factors
  • present in committed osteoprogenitor cells.

During mechanical loading/stimulation of chondrocytes there is an

  1. influx of ions across the cell membrane resulting from
  2. activation of mechanosensitive ion channels
  3. which can be inhibited by subunit-specific anti-integrin blocking antibodies or RGD peptides.

Using these strategies it was identified that

  • α5β1 integrin is a major mechanoreceptor in articular chondrocyte
  • responses to mechanical loading/stimulation.

Osteoarthritic chondrocytes show a depolarization response to 0.33 Hz stimulation

  • in contrast to the hyperpolarization response of normal chondrocytes.

The mechanotransduction pathway in chondrocytes derived from normal and osteoarthritic cartilage

  • both involve recognition of the mechanical stimulus
  • by integrin receptors resulting in
  • the activation of integrin signaling pathways
  • leading to the generation of a cytokine loop.

Normal and osteoarthritic chondrocytes show differences

  • at multiple stages of the mechanotransduction cascade.
Signaling pathways activated in chondrocytes

Signaling pathways activated in chondrocytes

http://dx.doi.org/10.1016/j.matbio.2014.08.007

Chondrocyte integrins are important mediators of cell–matrix interactions in cartilage

  • by regulating the response of the cells to signals from the ECM that
  1. control cell proliferation,
  2. survival,
  3. differentiation,
  4. matrix remodeling.

Integrins participate in development and maintenance of the tissue but also

  • in pathological processes related to matrix destruction, where
  • they likely play a role in the progression of OA.

Cellular adaptation to mechanical stress: role of integrins, Rho, cytoskeletal tension and mechanosensitive ion channels

Cells exhibited four types of mechanical responses:

(1) an immediate viscoelastic response;

(2) early adaptive behavior characterized by pulse-to-pulse attenuation in response to oscillatory forces;

(3) later adaptive cell stiffening with sustained (>15 second) static stresses; and

(4) a large-scale repositioning response with prolonged (>1 minute) stress.

Importantly, these adaptation responses differed biochemically.

The immediate and early responses were affected by

chemically dissipating cytoskeletal prestress (isometric tension), whereas

the later adaptive response was not.

The repositioning response was prevented by

inhibiting tension through interference with Rho signaling,

similar to the case of the immediate and early responses, but it was also prevented by

blocking mechanosensitive ion channels or

by inhibiting Src tyrosine kinases.

All adaptive responses were suppressed by cooling cells to 4°C to slow biochemical remodeling. Thus, cells use multiple mechanisms to sense and respond to static and dynamic changes in the level of mechanical stress applied to integrins.

Microtubule-Stimulated ADP Release, ATP Binding, and Force Generation In Transport Kinesins

All three classes of molecular motor proteins are now known to be

  • large protein families with diverse cellular functions.

Both the kinesin family and the myosin family have been defined and their proteins grouped into subfamilies. Finally, the elusive cytoplasmic version of dynein was identified and a multigene family of flagellar and cytoplasmic dyneins defined. Members of a given motor protein family share

  • significant homology in their motor domains with the defining member,
  • kinesin, dynein or myosin; but they also contain
  • unique protein domains that are specialized for interaction with different cargoes.

This large number of motor proteins may reflect

  • the number of cellular functions that require force generation or movement,
  • ranging from mitosis to morphogenesis to transport of vesicles.

Kinesins are a large family of microtubule (MT)-based motors that play important roles in many cellular activities including

mitosis,

motility, and

intracellular transport

Their involvement in a range of pathological processes

  • also highlights their significance as therapeutic targets and
  • the importance of understanding the molecular basis of their function

They are defined by their motor domains that contain both

  • the microtubule (MT) and
  • ATP binding sites.

Three ATP binding motifs—

  1. the P-loop,
  2. switch I,
  3. switch II–

are highly conserved among

  1. kinesins,
  2. myosin motors, and
  • small GTPases.

They share a conserved mode of MT binding such that

  • MT binding,
  • ATP binding, and
  • hydrolysis

are functionally coupled for efficient MT-based work.

The interior of a cell is a hive of activity, filled with

  • proteins and other items moving from one location to another.

A network of filaments called microtubules forms tracks

  • along which so-called motor proteins carry these items.

Kinesins are one group of motor proteins, and a typical kinesin protein has

  • one end (called the ‘motor domain’) that can attach itself to the microtubules.

The other end links to the cargo being carried, and a ‘neck’ connects the two. When two of these proteins work together,

  • flexible regions of the neck allow the two motor domains to move past one another,
  • which enable the kinesin to essentially walk along a microtubule in a stepwise manner.

Although the two kinesins have been thought to move along the microtubule tracks in different ways, Atherton et al. find that the core mechanism used by their motor domains is the same.

When a motor domain binds to the microtubule, its shape changes,

  • first stimulating release of the breakdown products of ATP from the previous cycle.

This release makes room for a new ATP molecule to bind. The structural changes caused by ATP binding

  • produce larger changes in the flexible neck region that
  • enable individual motor domains within a kinesin pair to
  • co-ordinate their movement and move in a consistent direction.

The major and largely invariant point of contact between kinesin motor domains and the MT is helix-α4,

  • which lies at the tubulin intradimer interface.

The conformational changes in functionally important regions of each motor domain are described,

  • starting with the nucleotide-binding site,
  • from which all other conformational changes emanate.

The nucleotide-binding site (Figure 2) has three major elements:

(1) the P-loop (brown) is visible in all our reconstructions;

(2) loop9 (yellow, contains switch I) undergoes major conformational changes through the ATPase cycle; and

(3) loop11 (red, contains switch II) that connects strand-β7 to helix-α4, the conformation and flexibility of which is

  • determined by MT binding and motor nucleotide state.

Movement and extension of helix-α6 controls neck linker docking

the N-terminus of helix-α6 is closely associated with elements of the nucleotide binding site suggesting that

  • its conformation alters in response to different nucleotide states.

Further,

  • because the orientation of helix-α6 with respect to helix-α4 controls neck linker docking and
  • because helix-α4 is held against the MT during the ATPase cycle,
    • conformational changes in helix-α6 control movement of the neck linker.

Mechanical amplification and force generation involves conformational changes across the motor domain

A key conformational change in the motor domain following Mg-ATP binding is

  • peeling of the central β-sheet from the C-terminus of helix-α4 increasing their separation;
  • this is required to accommodate rotation of helix-α6 and consequent neck linker docking

ATP binding draws loop11 and loop9 closer together; causing

(1) tilting of most of the motor domain not contacting the MT towards the nucleotide-binding site,

(2) rotation, translation, and extension of helix-α6 which we propose contributes to force generation, and

(3) allows neck linker docking and biases movement of the 2nd head towards the MT plus end.

In both motors, microtubule binding promotes

ordered conformations of conserved loops that

stimulate ADP release,

enhance microtubule affinity and

prime the catalytic site for ATP binding.

ATP binding causes only small shifts of these nucleotide-coordinating loops but induces

large conformational changes elsewhere that

allow force generation and

neck linker docking towards the microtubule plus end.

The study presents evidence provide evidence for a conserved ATP-driven

mechanism for kinesins and

reveals the critical mechanistic contribution of the microtubule interface.

Phosphorylation at endothelial cell–cell junctions: Implications for VE-cadherin function

This review summarizes the role of VE-cadherin phosphorylation in the regulation of endothelial cell–cell junctions and highlights how this affects vascular permeability and leukocyte extravasation.

The vascular endothelium is the inner lining of blood vessels and

forms a physical barrier between the vessel lumen and surrounding tissue;

controlling the extravasation of fluids,

plasma proteins and leukocytes.

Changes in the permeability of the endothelium are tightly regulated. Under basal physiological conditions, there is a continuous transfer of substances across the capillary beds. In addition the endothelium can mediate inducible,

transient hyperpermeability

in response to stimulation with inflammatory mediators,

which takes place primarily in post-capillary venules

However, when severe, inflammation may result in dysfunction of the endothelial barrier

  • in various parts of the vascular tree, including large veins, arterioles and capillaries.

Dysregulated permeability is observed in various pathological conditions, such as

  • tumor-induced angiogenesis,
  • cerebrovascular accident and
  • atherosclerosis.

Two fundamentally different pathways regulate endothelial permeability,

  1. the transcellular and
  2. paracellular pathways.

Solutes and cells can pass through the body of endothelial cells via the transcellular pathway, which includes

  • vesicular transport systems,
  • fenestrae, and
  • biochemical transporters.

The paracellular route is controlled by

  • the coordinated opening and closing of endothelial junctions and
  • thereby regulates traffic across the intercellular spaces between endothelial cells.

Endothelial cells are connected by

tight, gap and

adherens junctions,

of which the latter, and particularly the adherens junction component,

vascular endothelial (VE)-cadherin,

are of central importance for the initiation and stabilization of cell–cell contacts.

Although multiple adhesion molecules are localized at endothelial junctions,

  • blocking the adhesive function of VE-cadherin using antibodies
  • is sufficient to disrupt endothelial junctions and
  • to increase endothelial monolayer permeability both in vitro and in vivo.

Like other cadherins, VE-cadherin mediates adhesion via

  • homophilic, calcium-dependent interactions.

This cell–cell adhesion

is strengthened by binding of cytoplasmic proteins, the catenins,

to the C-terminus of VE-cadherin.

VE-cadherin can directly bind

  • β-catenin and plakoglobin, which
  • both associate with the actin binding protein α-catenin.

Initially, α-catenin was thought to directly anchor cadherins to the actin cytoskeleton, but recently it became clear that

  • α-catenin cannot bind to both β-catenin and actin simultaneously.

Numerous lines of evidence indicate that p120-catenin

  • promotes VE-cadherin surface expression and stability at the plasma membrane.

Different models are proposed that describe how

  • p120-catenin regulates cadherin membrane dynamics, including the hypothesis
  • that p120-catenin functions as a ‘cap’ that prevents the interaction of VE-cadherin
  • with the endocytic membrane trafficking machinery.

In addition, p120-catenin might regulate VE-cadherin internalization

  • through interactions with small GTPases.

Cytoplasmic p120-catenin, which is not bound to VE-cadherin, has been shown to

decrease RhoA activity,

elevate active Rac1 and Cdc42, and thereby is thought

to regulate actin cytoskeleton organization and membrane trafficking.

The intact cadherin-catenin complex is required for proper functioning of the adherens junction.

Several mechanisms may be involved in the

  • regulation of the organization and function of the cadherin–catenin complex, including
  1. endocytosis of the complex,
  2. VE-cadherin cleavage and
  3. actin cytoskeleton reorganization.

The remainder of this review primarily focuses on the

role of tyrosine phosphorylation in the control of VE-cadherin-mediated cell–cell adhesion.

Regulation of the adhesive function of VE-cadherin by tyrosine phosphorylation

It is a widely accepted concept that tyrosine phosphorylation of

  • components of the VE–cadherin-catenin complex
  • Correlates with the weakening of cell–cell adhesion.

A general idea has emerged that

tyrosine phosphorylation of the VE-cadherin complex

leads to the uncoupling of VE-cadherin from the actin cytoskeleton

through dissociation of catenins from the cadherin.

However, tyrosine phosphorylation of VE-cadherin

  • is required for efficient transmigration of leukocytes.

This suggests that VE-cadherin-mediated cell–cell contacts

1.are not just pushed open by the migrating leukocytes, but play

2.a more active role in the transmigration process.

A schematic overview of leukocyte adhesion-induced signals leading to VE-cadherin phosphorylation

Regulation of the integrity of endothelial cell–cell contacts by phosphorylation of VE-cadherin.

Regulation of the integrity of endothelial cell–cell contacts by phosphorylation of VE-cadherin

Regulation of the integrity of endothelial cell–cell contacts by phosphorylation of VE-cadherin

N-glycosylation status of E-cadherin controls cytoskeletal dynamics through the organization of distinct β-catenin- and γ-catenin-containing AJs

N-glycosylation of E-cadherin has been shown to inhibit cell–cell adhesion.

Specifically, our recent studies have provided evidence that

  • the reduction of E-cadherin N-glycosylation
  • promoted the recruitment of stabilizing components,
  • vinculin and serine/ threonine protein phosphatase 2A (PP2A), to adherens junctions (AJs)
  • and enhanced the association of AJs with the actin cytoskeleton.

Here, we examined the details of how

N-glycosylation of E-cadherin affected the molecular organization of AJs and their cytoskeletal interactions.

Using the hypoglycosylated E-cadherin variant, V13, we show that

V13/β-catenin complexes preferentially interacted with PP2A and with the microtubule motor protein dynein.

This correlated with dephosphorylation of the microtubule-associated protein tau, suggesting that

increased association of PP2A with V13-containing AJs promoted their tethering to microtubules.

These studies provide the first mechanistic insights into how N-glycosylation of E-cadherin drives changes in AJ composition through

  • the assembly of distinct β-catenin- and γ-catenin-containing scaffolds that impact the interaction with different cytoskeletal components

Cytoskeletal Basis of Ion Channel Function in Cardiac Muscle

MacKinnon. Fig 1  Ion channels exhibit three basic properties

MacKinnon. Fig 1 Ion channels exhibit three basic properties

In order to contract and accommodate the repetitive morphological changes induced by the cardiac cycle, cardiomyocytes

depend on their highly evolved and specialized cytoskeletal apparatus.

Defects in components of the cytoskeleton, in the long term,

affect the ability of the cell to compensate at both functional and structural levels.

In addition to the structural remodeling,

the myocardium becomes increasingly susceptible to altered electrical activity leading to arrhythmogenesis.

The development of arrhythmias secondary to structural remodeling defects has been noted, although the detailed molecular mechanisms are still elusive.

subjects with severe left ventricular chamber dilation such as in DCM can have left bundle branch block (LBBB), while right bundle branch block (RBBB) is more characteristic of right ventricular failure.  LBBB and RBBB have both been repeatedly associated with AV block in heart failure.

The impact of volume overload on structural and electro-cardiographic alterations has been noted in cardiomyopathy patients treated with left ventricular assist device (LVAD) therapy, which puts the heart at mechanical rest.

In LVAD-treated subjects,

QRS- and both QT- and QTc duration decreased,

suggesting that QRS- and QT-duration are significantly influenced by mechanical load and

that the shortening of the action potential duration contributes to the improved contractile performance after LVAD support.

An early postoperative period study after cardiac unloading therapy in 17 HF patients showed that in the first two weeks after LVAD implantation,

HF was associated with a relatively high incidence of ventricular arrhythmias associated with QTc interval prolongation.

In addition, a recent retrospective study of 100 adult patients with advanced HF, treated with an axial-flow HeartMate LVAD suggested that

  • the rate of new-onset monomorphic ventricular tachycardia (MVT) was increased in LVAD treated patients compared to patients given only medical treatment,

The myocardium is exposed to severe and continuous biomechanical stress during each contraction-relaxation cycle. When fiber tension remains uncompensated or simply unbalanced,

it may represent a trigger for arrhythmogenesis caused by cytoskeletal stretching,

which ultimately leads to altered ion channel localization, and subsequent action potential and conduction alterations.

Cytoskeletal proteins not only provide the backbone of the cellular structure, but they also

maintain the shape and flexibility of the different sub-cellular compartments, including the

1.plasma membrane,

2.the double lipid layer, which defines the boundaries of the cell and where

ion channels are mainly localized.

The interaction between the sarcomere, which is the basic for the passive force during diastole and for the restoring force during systole.

Sarcomeric Proteins and Ion Channels

besides fiber stretch associated with mechanical and hemodynamic impairment, cytoskeletal alterations due to primary genetic defects or indirectly to alterations in response to cellular injury can potentially

1.affect ion channel anchoring, and trafficking, as well as

2.functional regulation by second messenger pathways,

3.causing an imbalance in cardiac ionic homeostasis that will trigger arrhythmogenesis.

Intense investigation of

the sarcomeric actin network,

the Z-line structure, and

chaperone molecules docking in the plasma membrane,

has shed new light on the molecular basis of

  • cytoskeletal interactions in regulating ion channels

Actin disruption using cytochalasin D, an agent that interferes with actin polymerization, increased Na+ channel activity in 90% of excised patches tested within 2 min, which indicated that

the integrity of the filamentous actin (F-actin) network was essential for the maintenance of normal Na+ channel function

These data were the first to support a role for the cytoskeleton in cardiac arrhythmias.

Molecular interactions between the cytoskeleton and ion channels

The figure illustrates the interactions between the ion channels on the sarcolemma, and the sarcomere in cardiac myocytes. Note that the Z-line is connected to the cardiac T-tubules. The diagram illustrates the complex protein-protein interactions that occur between structural components of the cytoskeleton and ion channels. The cytoskeleton is involved in regulating the metabolism of ion channels, modifying their expression, localization, and electrical properties.

sarcomere structure

sarcomere structure

It is important to be aware of the enormous variety of clinical presentations that derive from distinct variants in the same pool of genetic factors. Knowledge of these variants could facilitate tailoring the therapy of choice for each patient. In particular,

the recent findings of structural and functional links between

the cytoskeleton and ion channels

could expand the therapeutic interventions in

arrhythmia management in structurally abnormal myocardium, where aberrant binding

between cytoskeletal proteins can directly or indirectly alter ion channel function.

Read Full Post »

Summary of Signaling and Signaling Pathways

Summary of Signaling and Signaling Pathways

Author and Curator: Larry H Bernstein, MD, FCAP

In the imtroduction to this series of discussions I pointed out JEDS Rosalino’s observation about the construction of a complex molecule of acetyl coenzyme A, and the amount of genetic coding that had to go into it.  Furthermore, he observes –  Millions of years later, or as soon as, the information of interaction leading to activity and regulation could be found in RNA, proteins like reverse transcriptase move this information to a more stable form (DNA). In this way it is easier to understand the use of CoA to make two carbon molecules more reactive.

acetylCoA

acetylCoA

In the tutorial that follows we find support for the view that mechanisms and examples from the current literature, which give insight into the developments in cell metabolism, are achieving a separation from inconsistent views introduced by the classical model of molecular biology and genomics, toward a more functional cellular dynamics that is not dependent on the classic view.  The classical view fits a rigid framework that is to genomics and metabolomics as Mendelian genetics if to multidimentional, multifactorial genetics.  The inherent difficulty lies in two places:

  1. Interactions between differently weighted determinants
  2. A large part of the genome is concerned with regulatory function, not expression of the code

The goal of the tutorial was to achieve an understanding of how cell signaling occurs in a cell.  Completion of the tutorial would provide

  1. a basic understanding signal transduction and
  2. the role of phosphorylation in signal transduction.
Regulation of the integrity of endothelial cell–cell contacts by phosphorylation of VE-cadherin

Regulation of the integrity of endothelial cell–cell contacts by phosphorylation of VE-cadherin

In addition – detailed knowledge of –

  1. the role of Tyrosine kinases and
  2. G protein-coupled receptors in cell signaling.
serine

serine

threonine

threonine

protein kinase

protein kinase

We are constantly receiving and interpreting signals from our environment, which can come

  • in the form of light, heat, odors, touch or sound.

The cells of our bodies are also

  • constantly receiving signals from other cells.

These signals are important to

  • keep cells alive and functioning as well as
  • to stimulate important events such as
  • cell division and differentiation.

Signals are most often chemicals that can be found

  • in the extracellular fluid around cells.

These chemicals can come

  • from distant locations in the body (endocrine signaling by hormones), from
  • nearby cells (paracrine signaling) or can even
  • be secreted by the same cell (autocrine signaling).

Notch-mediated juxtacrine signal between adjacent cells. 220px-Notchccr

Signaling molecules may trigger any number of cellular responses, including

  • changing the metabolism of the cell receiving the signal or
  • result in a change in gene expression (transcription) within the nucleus of the cell or both.
controlling the output of ribosomes.

controlling the output of ribosomes.

To which I would now add..

  • result in either an inhibitory or a stimulatory effect

The three stages of cell signaling are:

Cell signaling can be divided into 3 stages:

Reception: A cell detects a signaling molecule from the outside of the cell.

Transduction: When the signaling molecule binds the receptor it changes the receptor protein in some way. This change initiates the process of transduction. Signal transduction is usually a pathway of several steps. Each relay molecule in the signal transduction pathway changes the next molecule in the pathway.

Response: Finally, the signal triggers a specific cellular response.

signal transduction

signal transduction

http://www.hartnell.edu/tutorials/biology/images/signaltransduction_simple.jpg

The initiation is depicted as follows:

Signal Transduction – ligand binds to surface receptor

Membrane receptors function by binding the signal molecule (ligand) and causing the production of a second signal (also known as a second messenger) that then causes a cellular response. These types of receptors transmit information from the extracellular environment to the inside of the cell.

  • by changing shape or
  • by joining with another protein
  • once a specific ligand binds to it.

Examples of membrane receptors include

  • G Protein-Coupled Receptors and
Understanding these receptors and identifying their ligands and the resulting signal transduction pathways represent a major conceptual advance.

Understanding these receptors and identifying their ligands and the resulting signal transduction pathways represent a major conceptual advance.

  • Receptor Tyrosine Kinases.
intracellular signaling

intracellular signaling

http://www.hartnell.edu/tutorials/biology/images/membrane_receptor_tk.jpg

Intracellular receptors are found inside the cell, either in the cytopolasm or in the nucleus of the target cell (the cell receiving the signal).

Note that though change in gene expression is stated, the change in gene expression does not here imply a change in the genetic information – such as – mutation.  That does not have to be the case in the normal homeostatic case.

This point is the differentiating case between what JEDS Roselino has referred as

  1. a fast, adaptive reaction, that is the feature of protein molecules, and distinguishes this interaction from
  2. a one-to-one transcription of the genetic code.

The rate of transcription can be controlled, or it can be blocked.  This is in large part in response to the metabolites in the immediate interstitium.

This might only be

  • a change in the rate of a transcription or a suppression of expression through RNA.
  • Or through a conformational change in an enzyme
 Swinging domains in HECT E3 enzymes

Swinging domains in HECT E3 enzymes

Since signaling systems need to be

  • responsive to small concentrations of chemical signals and act quickly,
  • cells often use a multi-step pathway that transmits the signal quickly,
  • while amplifying the signal to numerous molecules at each step.

Signal transduction pathways are shown (simplified):

Signal Transduction

Signal Transduction

Signal transduction occurs when an

  1. extracellular signaling molecule activates a specific receptor located on the cell surface or inside the cell.
  2. In turn, this receptor triggers a biochemical chain of events inside the cell, creating a response.
  3. Depending on the cell, the response alters the cell’s metabolism, shape, gene expression, or ability to divide.
  4. The signal can be amplified at any step. Thus, one signaling molecule can cause many responses.

In 1970, Martin Rodbell examined the effects of glucagon on a rat’s liver cell membrane receptor. He noted that guanosine triphosphate disassociated glucagon from this receptor and stimulated the G-protein, which strongly influenced the cell’s metabolism. Thus, he deduced that the G-protein is a transducer that accepts glucagon molecules and affects the cell. For this, he shared the 1994 Nobel Prize in Physiology or Medicine with Alfred G. Gilman.

Guanosine monophosphate structure

Guanosine monophosphate structure

In 2007, a total of 48,377 scientific papers—including 11,211 e-review papers—were published on the subject. The term first appeared in a paper’s title in 1979. Widespread use of the term has been traced to a 1980 review article by Rodbell: Research papers focusing on signal transduction first appeared in large numbers in the late 1980s and early 1990s.

Signal transduction involves the binding of extracellular signaling molecules and ligands to cell-surface receptors that trigger events inside the cell. The combination of messenger with receptor causes a change in the conformation of the receptor, known as receptor activation.

This activation is always the initial step (the cause) leading to the cell’s ultimate responses (effect) to the messenger. Despite the myriad of these ultimate responses, they are all directly due to changes in particular cell proteins. Intracellular signaling cascades can be started through cell-substratum interactions; examples are the integrin that binds ligands in the extracellular matrix and steroids.

Integrin

Integrin

Most steroid hormones have receptors within the cytoplasm and act by stimulating the binding of their receptors to the promoter region of steroid-responsive genes.

steroid hormone receptor

steroid hormone receptor

Various environmental stimuli exist that initiate signal transmission processes in multicellular organisms; examples include photons hitting cells in the retina of the eye, and odorants binding to odorant receptors in the nasal epithelium. Certain microbial molecules, such as viral nucleotides and protein antigens, can elicit an immune system response against invading pathogens mediated by signal transduction processes. This may occur independent of signal transduction stimulation by other molecules, as is the case for the toll-like receptor. It may occur with help from stimulatory molecules located at the cell surface of other cells, as with T-cell receptor signaling. Receptors can be roughly divided into two major classes: intracellular receptors and extracellular receptors.

Signal transduction cascades amplify the signal output

Signal transduction cascades amplify the signal output

Signal transduction cascades amplify the signal output

G protein-coupled receptors (GPCRs) are a family of integral transmembrane proteins that possess seven transmembrane domains and are linked to a heterotrimeric G protein. Many receptors are in this family, including adrenergic receptors and chemokine receptors.

Arrestin binding to active GPCR kinase (GRK)-phosphorylated GPCRs blocks G protein coupling

signal transduction pathways

signal transduction pathways

Arrestin binding to active GPCR kinase (GRK)-phosphorylated GPCRs blocks G protein coupling

Arrestin binding to active GPCR kinase (GRK)-phosphorylated GPCRs blocks G protein coupling

Signal transduction by a GPCR begins with an inactive G protein coupled to the receptor; it exists as a heterotrimer consisting of Gα, Gβ, and Gγ. Once the GPCR recognizes a ligand, the conformation of the receptor changes to activate the G protein, causing Gα to bind a molecule of GTP and dissociate from the other two G-protein subunits.

The dissociation exposes sites on the subunits that can interact with other molecules. The activated G protein subunits detach from the receptor and initiate signaling from many downstream effector proteins such as phospholipases and ion channels, the latter permitting the release of second messenger molecules.

Receptor tyrosine kinases (RTKs) are transmembrane proteins with an intracellular kinase domain and an extracellular domain that binds ligands; examples include growth factor receptors such as the insulin receptor.

 insulin receptor and and insulin receptor signaling pathway (IRS)

insulin receptor and and insulin receptor signaling pathway (IRS)

To perform signal transduction, RTKs need to form dimers in the plasma membrane; the dimer is stabilized by ligands binding to the receptor.

RTKs

RTKs

The interaction between the cytoplasmic domains stimulates the autophosphorylation of tyrosines within the domains of the RTKs, causing conformational changes.

Allosteric_Regulation.svg

Subsequent to this, the receptors’ kinase domains are activated, initiating phosphorylation signaling cascades of downstream cytoplasmic molecules that facilitate various cellular processes such as cell differentiation and metabolism.

Signal-Transduction-Pathway

Signal-Transduction-Pathway

As is the case with GPCRs, proteins that bind GTP play a major role in signal transduction from the activated RTK into the cell. In this case, the G proteins are

  • members of the Ras, Rho, and Raf families, referred to collectively as small G proteins.

They act as molecular switches usually

  • tethered to membranes by isoprenyl groups linked to their carboxyl ends.

Upon activation, they assign proteins to specific membrane subdomains where they participate in signaling. Activated RTKs in turn activate

  • small G proteins that activate guanine nucleotide exchange factors such as SOS1.

Once activated, these exchange factors can activate more small G proteins, thus

  • amplifying the receptor’s initial signal.

The mutation of certain RTK genes, as with that of GPCRs, can result in the expression of receptors that exist in a constitutively activate state; such mutated genes may act as oncogenes.

Integrin

 

Integrin

Integrin

Integrin-mediated signal transduction

An overview of integrin-mediated signal transduction, adapted from Hehlgens et al. (2007).

Integrins are produced by a wide variety of cells; they play a role in

  • cell attachment to other cells and the extracellular matrix and
  • in the transduction of signals from extracellular matrix components such as fibronectin and collagen.

Ligand binding to the extracellular domain of integrins

  • changes the protein’s conformation,
  • clustering it at the cell membrane to
  • initiate signal transduction.

Integrins lack kinase activity; hence, integrin-mediated signal transduction is achieved through a variety of intracellular protein kinases and adaptor molecules, the main coordinator being integrin-linked kinase.

As shown in the picture, cooperative integrin-RTK signaling determines the

  1. timing of cellular survival,
  2. apoptosis,
  3. proliferation, and
  4. differentiation.
integrin-mediated signal transduction

integrin-mediated signal transduction

Integrin signaling

Integrin signaling

ion channel

A ligand-gated ion channel, upon binding with a ligand, changes conformation

  • to open a channel in the cell membrane
  • through which ions relaying signals can pass.

An example of this mechanism is found in the receiving cell of a neural synapse. The influx of ions that occurs in response to the opening of these channels

  1. induces action potentials, such as those that travel along nerves,
  2. by depolarizing the membrane of post-synaptic cells,
  3. resulting in the opening of voltage-gated ion channels.
RyR and Ca+ release from SR

RyR and Ca+ release from SR

An example of an ion allowed into the cell during a ligand-gated ion channel opening is Ca2+;

  • it acts as a second messenger
  • initiating signal transduction cascades and
  • altering the physiology of the responding cell.

This results in amplification of the synapse response between synaptic cells

  • by remodelling the dendritic spines involved in the synapse.

In eukaryotic cells, most intracellular proteins activated by a ligand/receptor interaction possess an enzymatic activity; examples include tyrosine kinase and phosphatases. Some of them create second messengers such as cyclic AMP and IP3,

cAMP

cAMP

Inositol_1,4,5-trisphosphate.svg

Inositol_1,4,5-trisphosphate.svg

  • the latter controlling the release of intracellular calcium stores into the cytoplasm.

Many adaptor proteins and enzymes activated as part of signal transduction possess specialized protein domains that bind to specific secondary messenger molecules. For example,

  • calcium ions bind to the EF hand domains of calmodulin,
  • allowing it to bind and activate calmodulin-dependent kinase.
calcium movement and RyR2 receptor

calcium movement and RyR2 receptor

PIP3 and other phosphoinositides do the same thing to the Pleckstrin homology domains of proteins such as the kinase protein AKT.

Signals can be generated within organelles, such as chloroplasts and mitochondria, modulating the nuclear
gene expression in a process called retrograde signaling.

Recently, integrative genomics approaches, in which correlation analysis has been applied on transcript and metabolite profiling data of Arabidopsis thaliana, revealed the identification of metabolites which are putatively acting as mediators of nuclear gene expression.

http://fpls.com/unraveling_retrograde_signaling_pathways:_finding_candidate_signaling_molecules_via_metabolomics_and_systems_biology_driven_approaches

Related articles

  1. Systems Biology Approach Reveals Genome to Phenome Correlation in Type 2 Diabetes (plosone.org)
  2. Gene Expression and Thiopurine Metabolite Profiling in Inflammatory Bowel Disease – Novel Clues to Drug Targets and Disease Mechanisms? (plosone.org)
  3. Activation of the Jasmonic Acid Plant Defence Pathway Alters the Composition of Rhizosphere

Nutrients 2014, 6, 3245-3258; http://dx.doi.org:/10.3390/nu6083245

Omega-3 (ω-3) fatty acids are one of the two main families of long chain polyunsaturated fatty acids (PUFA). The main omega-3 fatty acids in the mammalian body are

  • α-linolenic acid (ALA), docosahexaenoic acid (DHA) and eicosapentaenoic acid (EPA).

Central nervous tissues of vertebrates are characterized by a high concentration of omega-3 fatty acids. Moreover, in the human brain,

  • DHA is considered as the main structural omega-3 fatty acid, which comprises about 40% of the PUFAs in total.

DHA deficiency may be the cause of many disorders such as depression, inability to concentrate, excessive mood swings, anxiety, cardiovascular disease, type 2 diabetes, dry skin and so on.

On the other hand,

  • zinc is the most abundant trace metal in the human brain.

There are many scientific studies linking zinc, especially

  • excess amounts of free zinc, to cellular death.

Neurodegenerative diseases, such as Alzheimer’s disease, are characterized by altered zinc metabolism. Both animal model studies and human cell culture studies have shown a possible link between

  • omega-3 fatty acids, zinc transporter levels and
  • free zinc availability at cellular levels.

Many other studies have also suggested a possible

  • omega-3 and zinc effect on neurodegeneration and cellular death.

Therefore, in this review, we will examine

  • the effect of omega-3 fatty acids on zinc transporters and
  • the importance of free zinc for human neuronal cells.

Moreover, we will evaluate the collective understanding of

  • mechanism(s) for the interaction of these elements in neuronal research and their
  • significance for the diagnosis and treatment of neurodegeneration.

Epidemiological studies have linked high intake of fish and shellfish as part of the daily diet to

  • reduction of the incidence and/or severity of Alzheimer’s disease (AD) and senile mental decline in

Omega-3 fatty acids are one of the two main families of a broader group of fatty acids referred to as polyunsaturated fatty acids (PUFAs). The other main family of PUFAs encompasses the omega-6 fatty acids. In general, PUFAs are essential in many biochemical events, especially in early post-natal development processes such as

  • cellular differentiation,
  • photoreceptor membrane biogenesis and
  • active synaptogenesis.

Despite the significance of these

two families, mammals cannot synthesize PUFA de novo, so they must be ingested from dietary sources. Though belonging to the same family, both

  • omega-3 and omega-6 fatty acids are metabolically and functionally distinct and have
  • opposing physiological effects. In the human body,
  • high concentrations of omega-6 fatty acids are known to increase the formation of prostaglandins and
  • thereby increase inflammatory processes [10].

the reverse process can be seen with increased omega-3 fatty acids in the body.

Many other factors, such as

  1. thromboxane A2 (TXA2),
  2. leukotriene
  3. B4 (LTB4),
  4. IL-1,
  5. IL-6,
  6. tumor necrosis factor (TNF) and
  7. C-reactive protein,

which are implicated in various health conditions, have been shown to be increased with high omega-6 fatty acids but decreased with omega-3 fatty acids in the human body.

Dietary fatty acids have been identified as protective factors in coronary heart disease, and PUFA levels are known to play a critical role in

  • immune responses,
  • gene expression and
  • intercellular communications.

omega-3 fatty acids are known to be vital in

  • the prevention of fatal ventricular arrhythmias, and
  • are also known to reduce thrombus formation propensity by decreasing platelet aggregation, blood viscosity and fibrinogen levels

.Since omega-3 fatty acids are prevalent in the nervous system, it seems logical that a deficiency may result in neuronal problems, and this is indeed what has been identified and reported.

The main

In another study conducted with individuals of 65 years of age or older (n = 6158), it was found that

  • only high fish consumption, but
  • not dietary omega-3 acid intake,
  • had a protective effect on cognitive decline

In 2005, based on a meta-analysis of the available epidemiology and preclinical studies, clinical trials were conducted to assess the effects of omega-3 fatty acids on cognitive protection. Four of the trials completed have shown

a protective effect of omega-3 fatty acids only among those with mild cognitive impairment conditions.

A  trial of subjects with mild memory complaints demonstrated

  • an improvement with 900 mg of DHA.

We review key findings on

  • the effect of the omega-3 fatty acid DHA on zinc transporters and the
  • importance of free zinc to human neuronal cells.

DHA is the most abundant fatty acid in neural membranes, imparting appropriate

  • fluidity and other properties,

and is thus considered as the most important fatty acid in neuronal studies. DHA is well conserved throughout the mammalian species despite their dietary differences. It is mainly concentrated

  • in membrane phospholipids at synapses and
  • in retinal photoreceptors and
  • also in the testis and sperm.

In adult rats’ brain, DHA comprises approximately

  • 17% of the total fatty acid weight, and
  • in the retina it is as high as 33%.

DHA is believed to have played a major role in the evolution of the modern human –

  • in particular the well-developed brain.

Premature babies fed on DHA-rich formula show improvements in vocabulary and motor performance.

Analysis of human cadaver brains have shown that

  • people with AD have less DHA in their frontal lobe
  • and hippocampus compared with unaffected individuals

Furthermore, studies in mice have increased support for the

  • protective role of omega-3 fatty acids.

Mice administrated with a dietary intake of DHA showed

  • an increase in DHA levels in the hippocampus.

Errors in memory were decreased in these mice and they demonstrated

  • reduced peroxide and free radical levels,
  • suggesting a role in antioxidant defense.

Another study conducted with a Tg2576 mouse model of AD demonstrated that dietary

  • DHA supplementation had a protective effect against reduction in
  • drebrin (actin associated protein), elevated oxidation, and to some extent, apoptosis via
  • decreased caspase activity.

 

Zinc

Zinc is a trace element, which is indispensable for life, and it is the second most abundant trace element in the body. It is known to be related to

  • growth,
  • development,
  • differentiation,
  • immune response,
  • receptor activity,
  • DNA synthesis,
  • gene expression,
  • neuro-transmission,
  • enzymatic catalysis,
  • hormonal storage and release,
  • tissue repair,
  • memory,
  • the visual process

and many other cellular functions. Moreover, the indispensability of zinc to the body can be discussed in many other aspects,  as

  • a component of over 300 different enzymes
  • an integral component of a metallothioneins
  • a gene regulatory protein.

Approximately 3% of all proteins contain

  • zinc binding motifs .

The broad biological functionality of zinc is thought to be due to its stable chemical and physical properties. Zinc is considered to have three different functions in enzymes;

  1. catalytic,
  2. coactive and

Indeed, it is the only metal found in all six different subclasses

of enzymes. The essential nature of zinc to the human body can be clearly displayed by studying the wide range of pathological effects of zinc deficiency. Anorexia, embryonic and post-natal growth retardation, alopecia, skin lesions, difficulties in wound healing, increased hemorrhage tendency and severe reproductive abnormalities, emotional instability, irritability and depression are just some of the detrimental effects of zinc deficiency.

Proper development and function of the central nervous system (CNS) is highly dependent on zinc levels. In the mammalian organs, zinc is mainly concentrated in the brain at around 150 μm. However, free zinc in the mammalian brain is calculated to be around 10 to 20 nm and the rest exists in either protein-, enzyme- or nucleotide bound form. The brain and zinc relationship is thought to be mediated

  • through glutamate receptors, and
  • it inhibits excitatory and inhibitory receptors.

Vesicular localization of zinc in pre-synaptic terminals is a characteristic feature of brain-localized zinc, and

  • its release is dependent on neural activity.

Retardation of the growth and development of CNS tissues have been linked to low zinc levels. Peripheral neuropathy, spina bifida, hydrocephalus, anencephalus, epilepsy and Pick’s disease have been linked to zinc deficiency. However, the body cannot tolerate excessive amounts of zinc.

The relationship between zinc and neurodegeneration, specifically AD, has been interpreted in several ways. One study has proposed that β-amyloid has a greater propensity to

  • form insoluble amyloid in the presence of
  • high physiological levels of zinc.

Insoluble amyloid is thought to

  • aggregate to form plaques,

which is a main pathological feature of AD. Further studies have shown that

  • chelation of zinc ions can deform and disaggregate plaques.

In AD, the most prominent injuries are found in

  • hippocampal pyramidal neurons, acetylcholine-containing neurons in the basal forebrain, and in
  • somatostatin-containing neurons in the forebrain.

All of these neurons are known to favor

  • rapid and direct entry of zinc in high concentration
  • leaving neurons frequently exposed to high dosages of zinc.

This is thought to promote neuronal cell damage through oxidative stress and mitochondrial dysfunction. Excessive levels of zinc are also capable of

  • inhibiting Ca2+ and Na+ voltage gated channels
  • and up-regulating the cellular levels of reactive oxygen species (ROS).

High levels of zinc are found in Alzheimer’s brains indicating a possible zinc related neurodegeneration. A study conducted with mouse neuronal cells has shown that even a 24-h exposure to high levels of zinc (40 μm) is sufficient to degenerate cells.

If the human diet is deficient in zinc, the body

  • efficiently conserves zinc at the tissue level by compensating other cellular mechanisms

to delay the dietary deficiency effects of zinc. These include reduction of cellular growth rate and zinc excretion levels, and

  • redistribution of available zinc to more zinc dependent cells or organs.

A novel method of measuring metallothionein (MT) levels was introduced as a biomarker for the

  • assessment of the zinc status of individuals and populations.

In humans, erythrocyte metallothionein (E-MT) levels may be considered as an indicator of zinc depletion and repletion, as E-MT levels are sensitive to dietary zinc intake. It should be noted here that MT plays an important role in zinc homeostasis by acting

  • as a target for zinc ion binding and thus
  • assisting in the trafficking of zinc ions through the cell,
  • which may be similar to that of zinc transporters

Zinc Transporters

Deficient or excess amounts of zinc in the body can be catastrophic to the integrity of cellular biochemical and biological systems. The gastrointestinal system controls the absorption, excretion and the distribution of zinc, although the hydrophilic and high-charge molecular characteristics of zinc are not favorable for passive diffusion across the cell membranes. Zinc movement is known to occur

  • via intermembrane proteins and zinc transporter (ZnT) proteins

These transporters are mainly categorized under two metal transporter families; Zip (ZRT, IRT like proteins) and CDF/ZnT (Cation Diffusion Facilitator), also known as SLC (Solute Linked Carrier) gene families: Zip (SLC-39) and ZnT (SLC-30). More than 20 zinc transporters have been identified and characterized over the last two decades (14 Zips and 8 ZnTs).

Members of the SLC39 family have been identified as the putative facilitators of zinc influx into the cytosol, either from the extracellular environment or from intracellular compartments (Figure 1).

The identification of this transporter family was a result of gene sequencing of known Zip1 protein transporters in plants, yeast and human cells. In contrast to the SLC39 family, the SLC30 family facilitates the opposite process, namely zinc efflux from the cytosol to the extracellular environment or into luminal compartments such as secretory granules, endosomes and synaptic vesicles; thus decreasing intracellular zinc availability (Figure 1). ZnT3 is the most important in the brain where

  • it is responsible for the transport of zinc into the synaptic vesicles of
  • glutamatergic neurons in the hippocampus and neocortex,

Figure 1: Subcellular localization and direction of transport of the zinc transporter families, ZnT and ZIP. Arrows show the direction of zinc mobilization for the ZnT (green) and ZIP (red) proteins. A net gain in cytosolic zinc is achieved by the transportation of zinc from the extracellular region and organelles such as the endoplasmic reticulum (ER) and Golgi apparatus by the ZIP transporters. Cytosolic zinc is mobilized into early secretory compartments such as the ER and Golgi apparatus by the ZnT transporters. Figures were produced using Servier Medical Art, http://www.servier.com/.   http://www.hindawi.com/journals/jnme/2012/173712.fig.001.jpg

Figure 2: Early zinc signaling (EZS) and late zinc signaling (LZS). EZS involves transcription-independent mechanisms where an extracellular stimulus directly induces an increase in zinc levels within several minutes by releasing zinc from intracellular stores (e.g., endoplasmic reticulum). LSZ is induced several hours after an external stimulus and is dependent on transcriptional changes in zinc transporter expression. Components of this figure were produced using Servier Medical Art, http://www.servier.com/ and adapted from Fukada et al. [30].

omega-3 fatty acids in the mammalian body are

  1. α-linolenic acid (ALA),
  2. docosahexenoic acid (DHA) and
  3. eicosapentaenoic acid (EPA).

In general, seafood is rich in omega-3 fatty acids, more specifically DHA and EPA (Table 1). Thus far, there are nine separate epidemiological studies that suggest a possible link between

  • increased fish consumption and reduced risk of AD
  • and eight out of ten studies have reported a link between higher blood omega-3 levels

DHA and Zinc Homeostasis

Many studies have identified possible associations between DHA levels, zinc homeostasis, neuroprotection and neurodegeneration. Dietary DHA deficiency resulted in

  • increased zinc levels in the hippocampus and
  • elevated expression of the putative zinc transporter, ZnT3, in the rat brain.

Altered zinc metabolism in neuronal cells has been linked to neurodegenerative conditions such as AD. A study conducted with transgenic mice has shown a significant link between ZnT3 transporter levels and cerebral amyloid plaque pathology. When the ZnT3 transporter was silenced in transgenic mice expressing cerebral amyloid plaque pathology,

  • a significant reduction in plaque load
  • and the presence of insoluble amyloid were observed.

In addition to the decrease in plaque load, ZnT3 silenced mice also exhibited a significant

  • reduction in free zinc availability in the hippocampus
  • and cerebral cortex.

Collectively, the findings from this study are very interesting and indicate a clear connection between

  • zinc availability and amyloid plaque formation,

thus indicating a possible link to AD.

DHA supplementation has also been reported to limit the following:

  1. amyloid presence,
  2. synaptic marker loss,
  3. hyper-phosphorylation of Tau,
  4. oxidative damage and
  5. cognitive deficits in transgenic mouse model of AD.

In addition, studies by Stoltenberg, Flinn and colleagues report on the modulation of zinc and the effect in transgenic mouse models of AD. Given that all of these are classic pathological features of AD, and considering the limiting nature of DHA in these processes, it can be argued that DHA is a key candidate in preventing or even curing this debilitating disease.

In order to better understand the possible links and pathways of zinc and DHA with neurodegeneration, we designed a study that incorporates all three of these aspects, to study their effects at the cellular level. In this study, we were able to demonstrate a possible link between omega-3 fatty acid (DHA) concentration, zinc availability and zinc transporter expression levels in cultured human neuronal cells.

When treated with DHA over 48 h, ZnT3 levels were markedly reduced in the human neuroblastoma M17 cell line. Moreover, in the same study, we were able to propose a possible

  • neuroprotective mechanism of DHA,

which we believe is exerted through

  • a reduction in cellular zinc levels (through altering zinc transporter expression levels)
  • that in turn inhibits apoptosis.

DHA supplemented M17 cells also showed a marked depletion of zinc uptake (up to 30%), and

  • free zinc levels in the cytosol were significantly low compared to the control

This reduction in free zinc availability was specific to DHA; cells treated with EPA had no significant change in free zinc levels (unpublished data). Moreover, DHA-repleted cells had

  • low levels of active caspase-3 and
  • high Bcl-2 levels compared to the control treatment.

These findings are consistent with previous published data and further strengthen the possible

  • correlation between zinc, DHA and neurodegeneration.

On the other hand, recent studies using ZnT3 knockout (ZnT3KO) mice have shown the importance of

  • ZnT3 in memory and AD pathology.

For example, Sindreu and colleagues have used ZnT3KO mice to establish the important role of

  • ZnT3 in zinc homeostasis that modulates presynaptic MAPK signaling
  • required for hippocampus-dependent memory

Results from these studies indicate a possible zinc-transporter-expression-level-dependent mechanism for DHA neuroprotection.

Read Full Post »

Integrins, Cadherins, Signaling and the Cytoskeleton

Curator: Larry H. Bernstein, MD, FCAP 

 

We have reviewed the cytoskeleton, cytoskeleton pores and ionic translocation under lipids. We shall now look at this again, with specific attention to proteins, transporters and signaling.

Integrins and extracellular matrix in mechanotransduction

Lindsay Ramage
Queen’s Medical Research Institute, University of Edinburgh,

Edinburgh, UK
Cell Health and Cytoskeleton 2012; 4: 1–9

https://s3.amazonaws.com/academia.edu.documents/37116869/CHC-21829-integrins-and-extracellular-matrix-in-mechanotransduction_122311.pdf?response-content-disposition=inline%3B%20filename%3DCell_Health_and_Cytoskeleton_Integrins_a.pdf&X-Amz-Algorithm=AWS4-HMAC-SHA256&X-Amz-Credential=AKIAIWOWYYGZ2Y53UL3A%2F20191231%2Fus-east-1%2Fs3%2Faws4_request&X-Amz-Date=20191231T021009Z&X-Amz-Expires=3600&X-Amz-SignedHeaders=host&X-Amz-Signature=b376084e0e1c31c399ee7fe96eb81b1b65d3346d647192e9ebeff96f577e117d

Integrins are a family of cell surface receptors which

  • mediate cell–matrix and cell–cell adhesions.

Among other functions they provide an important

  • mechanical link between the cells external and intracellular environments while
  • the adhesions that they form also have critical roles in cellular signal-transduction.

Cell–matrix contacts occur at zones in the cell surface where

  • adhesion receptors cluster and when activated
  • the receptors bind to ligands in the extracellular matrix.

The extracellular matrix surrounds the cells of tissues and forms the

  • structural support of tissue which is particularly important in connective tissues.

Cells attach to the extracellular matrix through

  • specific cell-surface receptors and molecules
  • including integrins and transmembrane proteoglycans.

Integrins work alongside other proteins such as

  • cadherins,
  • immunoglobulin superfamily
  • cell adhesion molecules,
  • selectins, and
  • syndecans

to mediate

  • cell–cell and
  • cell–matrix interactions and communication.

Activation of adhesion receptors triggers the formation of matrix contacts in which

  • bound matrix components,
  • adhesion receptors,
  • and associated intracellular cytoskeletal and signaling molecules

form large functional, localized multiprotein complexes.

Cell–matrix contacts are important in a variety of different cell and

tissue properties including

  1. embryonic development,
  2. inflammatory responses,
  3. wound healing,
  4. and adult tissue homeostasis.

This review summarizes the roles and functions of integrins and extracellular matrix proteins in mechanotransduction.

Integrins are a family of αβ heterodimeric receptors which act as

  • cell adhesion molecules
  • connecting the ECM to the actin cytoskeleton.

The actin cytoskeleton is involved in the regulation of

  1. cell motility,
  2. cell polarity,
  3. cell growth, and
  4. cell survival.

The integrin family consists of around 25 members which are composed of differing

  • combinations of α and β subunits.

The combination of αβ subunits determines

  • binding specificity and
  • signaling properties.

In mammals around 19 α and eight β subunits have been characterized.

Both α and β integrin subunits contain two separate tails, which

  • penetrate the plasma membrane and possess small cytoplasmic domains which facilitate
  • the signaling functions of the receptor.

There is some evidence that the β subunit is the principal

site for

  • binding of cytoskeletal and signaling molecules,

whereas the α subunit has a regulatory role. The integrin

tails

  • link the ECM to the actin cytoskeleton within the cell and with cytoplasmic proteins,

such as talin, tensin, and filamin. The extracellular domains of integrin receptors bind the ECM ligands.

The ECM is a complex mixture of matrix molecules, including -glycoproteins, collagens, laminins, glycosaminoglycans, proteoglycans,
and nonmatrix proteins, – including growth factors.
These can be categorized as insoluble molecules within the ECM, soluble molecules, and/or matrix-associated biochemicals, such as systemic hormones or growth factors and cytokines that act locally.

The integrin receptor formed from the binding of α and β subunits is shaped like a globular head supported by two rod-like legs (Figure 1). Most of the contact between the two subunits occurs in the head region, with the intracellular tails of the subunits forming the legs of the receptor.6 Integrin recognition of ligands is not constitutive but is regulated by alteration of integrin affinity for ligand binding. For integrin binding to ligands to occur the integrin must be primed and activated, both of which involve conformational changes to the receptor.

The integrins are composed of well-defined domains used for protein–protein interactions. The α-I domains of α integrin subunits comprise the ligand binding sites. X-ray crystallography has identified an α-I domain within the β subunit and a β propeller domain within the α subunit which complex to form the ligand-binding head of the integrin.

The use of activating and conformation-specific antibodies also suggests that the β chain is extended in the active integrin. It has since been identified that the hybrid domain in the β chain is critical for integrin activation, and a swing-out movement of this leg activates integrins.

http://www.ks.uiuc.edu/Publications/Stories/tcbg_ytt/pdfs/dbp6.pdf

DBP6: Integrin

Integrin

Integrin

Integrin.large

Integrin.large

Linking integrin conformation to function

Figure  Integrin binding to extracellular matrix (ECM). Conformational changes to integrin structure and clustering of subunits which allow enhanced function of the receptor.

integrin coupled to F-actin via linker

integrin coupled to F-actin via linker

http://dx.dio.org:/integrin-coupled-to-f-actin-via-linker-nrm3896-f4.jpg

Integrin extracellular binding activity is regulated from inside the cell and binding to the ECM induces signals that are transmitted into the cell.15 This bidirectional signaling requires

  • dynamic,
  • spatially, and
  • temporally regulated formation and
  • disassembly of multiprotein complexes that
    form around the short cytoplasmic tails of integrins.

Ligand binding to integrin family members leads to clustering of integrin molecules in the plasma membrane and recruitment of actin filaments and intracellular signaling molecules to the cytoplasmic domain of the integrins. This forms focal adhesion complexes which are able to maintain

  • not only adhesion to the ECM
  • but are involved in complex signaling pathways

which include establishing

  1. cell polarity,
  2. directed cell migration, and
  3. maintaining cell growth and survival.

Initial activation through integrin adhesion to matrix recruits up to around 50 diverse signaling molecules

  • to assemble the focal adhesion complex
  • which is capable of responding to environmental stimuli efficiently.

Mapping of the integrin

  • adhesome binding and signaling interactions

identified a network of 156 components linked together which can be modified by 690 interactions.

The binding of the adaptor protein talin to the β subunit cytoplasmic tail is known to have a key role in integrin activation. This is thought to occur through the disruption of

  • inhibitory interactions between α and β subunit cytoplasmic tails.

Talin also binds

  • to actin and to cytoskeletal and signaling proteins.

This allows talin to directly link activated integrins

to signaling events and the cytoskeleton.

 

Genetic programming occurs with the binding of integrins to the ECM

Signal transduction pathway activation arising from integrin-

ECM binding results in changes in gene expression of cells

and leads to alterations in cell and tissue function. Various

different effects can arise depending on the

  1. cell type,
  2. matrix composition, and
  3. integrins activated.

One way in which integrin expression is important in genetic programming is in the fate and differentiation of stem cells.
Osteoblast differentiation occurs through ECM interactions

with specific integrins

  • to initiate intracellular signaling pathways leading to osteoblast-specific gene expression
  • disruption of interactions between integrins and collagen;
  • fibronectin blocks osteoblast differentiation and

Disruption of α2 integrin prevents osteoblast differentiation, and activation of the transcription factor

  • osteoblast-specific factor 2/core-binding factor α1.

It was found that the ECM-integrin interaction induces osteoblast-specific factor 2/core-binding factor α1 to

  • increase its activity as a transcriptional enhancer
  • rather than increasing protein levels.

It was also found that modification of α2 integrin alters

  • induction of the osteocalcin promoter;
  • inhibition of α2 prevents activation of the osteocalcin promoter,
  • overexpression enhanced osteocalcin promoter activity.

It has been suggested that integrin-type I collagen interaction is necessary for the phosphorylation and activation of osteoblast-specific transcription factors present in committed osteoprogenitor cells.

A variety of growth factors and cytokines have been shown to be important in the regulation of integrin expression and function in chondrocytes. Mechanotransduction in chondrocytes occurs through several different receptors and ion channels including integrins. During osteoarthritis the expression of integrins by chondrocytes is altered, resulting in different cellular transduction pathways which contribute to tissue pathology.

In normal adult cartilage, chondrocytes express α1β1, α10β1 (collagen receptors), α5β1, and αvβ5 (fibronectin) receptors. During mechanical loading/stimulation of chondrocytes there is an influx of ions across the cell membrane resulting from activation of mechanosensitive ion channels which can be inhibited by subunit-specific anti-integrin blocking antibodies or RGD peptides. Using these strategies it was identified that α5β1 integrin is a major mechanoreceptor in articular chondrocyte responses to mechanical loading/stimulation.

Osteoarthritic chondrocytes show a depolarization response to 0.33 Hz stimulation in contrast to the hyperpolarization response of normal chondrocytes. The mechanotransduction pathway in chondrocytes derived from normal and osteoarthritic cartilage both involve recognition of the mechanical stimulus by integrin receptors resulting in the activation of integrin signaling pathways leading to the generation of a cytokine loop. Normal and osteoarthritic chondrocytes show differences at multiple stages of the mechanotransduction cascade (Figure 3). Early events are similar; α5β1 integrin and stretch activated ion channels are activated and result in rapid tyrosine phosphorylation events. The actin cytoskeleton is required for the integrin-dependent Mechanotransduction leading to changes in membrane potential in normal but not osteoarthritic chondrocytes.

Cell–matrix interactions are essential for maintaining the integrity of tissues. An intact matrix is essential for cell survival and proliferation and to allow efficient mechanotransduction and tissue homeostasis. Cell–matrix interactions have been extensively studied in many tissues and this knowledge is being used to develop strategies to treat pathology. This is particularly important in tissues subject to abnormal mechanical loading, such as musculoskeletal tissues. Integrin-ECM interactions are being used to enhance tissue repair mechanisms in these tissues through differentiation of progenitor cells for in vitro and in vivo use. Knowledge of how signaling cascades are differentially regulated in response to physiological and pathological external stimuli (including ECM availability and mechanical loading/stimulation) will enable future strategies to be developed to prevent and treat the progression of pathology associated with integrin-ECM interactions.

Cellular adaptation to mechanical stress: role of integrins, Rho, cytoskeletal tension and mechanosensitive ion channels

  1. Matthews, DR. Overby, R Mannix and DE. Ingber
    1Vascular Biology Program, Departments of Pathology and Surgery, Children’s Hospital, and 2Department of Pediatrics, Massachusetts General Hospital, Harvard Medical School, Boston, MA J Cell Sci 2006; 119: 508-518. http://dx.doi.org:/10.1242/jcs.02760

To understand how cells sense and adapt to mechanical stress, we applied tensional forces to magnetic microbeads bound to cell-surface integrin receptors and measured changes in bead isplacement with sub-micrometer resolution using optical microscopy. Cells exhibited four types of mechanical responses: (1) an immediate viscoelastic response;

(2) early adaptive behavior characterized by pulse-to-pulse attenuation in response to oscillatory forces;

(3) later adaptive cell stiffening with sustained (>15 second) static stresses; and

(4) a large-scale repositioning response with prolonged (>1 minute) stress.

Importantly, these adaptation responses differed biochemically. The immediate and early responses were affected by

  • chemically dissipating cytoskeletal prestress (isometric tension), whereas
  • the later adaptive response was not.

The repositioning response was prevented by

  • inhibiting tension through interference with Rho signaling,

similar to the case of the immediate and early responses, but it was also prevented by

  • blocking mechanosensitive ion channels or
  • by inhibiting Src tyrosine kinases.

All adaptive responses were suppressed by cooling cells to 4°C to slow biochemical remodeling. Thus, cells use multiple mechanisms to sense and respond to static and dynamic changes in the level of mechanical stress applied to integrins.

Microtubule-Stimulated ADP Release, ATP Binding, and Force Generation In Transport Kinesins

J Atherton, I Farabella, I-Mei Yu, SS Rosenfeld, A Houdusse, M Topf, CA Moores

1Institute of Structural and Molecular Biology, Department of Biological Sciences, Birkbeck College, University of London, London, United Kingdom; 2Structural Motility, Institut Curie, Centre National de la Recherche Scientifique, Paris, France; 3Department of Cancer Biology, Lerner Research Institute, Cleveland Clinic, Cleveland, United States
eLife 2014;3:e03680. http://dx.doi.org:/10.7554/eLife.03680

Kinesins are a large family of microtubule (MT)-based motors that play important roles in many cellular activities including

  • mitosis,
  • motility, and
  • intracellular transport

Their involvement in a range of pathological processes also highlights their significance as therapeutic targets and the importance of understanding the molecular basis of their function They are defined by their motor domains that contain both the microtubule (MT) and ATP binding sites. Three ATP binding motifs—the P-loop, switch I, switch II–are highly conserved among kinesins, myosin motors, and small GTPases. They share a conserved mode of MT binding such that MT binding, ATP binding, and hydrolysis are functionally coupled for efficient MT-based work.

The interior of a cell is a hive of activity, filled with proteins and other items moving from one location to another. A network of filaments called microtubules forms tracks along which so-called motor proteins carry these items. Kinesins are one group of motor proteins, and a typical kinesin protein has one end (called the ‘motor domain’) that can attach itself to the microtubules.

The other end links to the cargo being carried, and a ‘neck’ connects the two. When two of these proteins work together, flexible regions of the neck allow the two motor domains to move past one another, which enable the kinesin to essentially walk along a microtubule in a stepwise manner.

Atherton et al. use a technique called cryo-electron microscopy to study—in more detail than previously seen—the structure of the motor domains of two types of kinesin called kinesin-1 and kinesin-3. Images were taken at different stages of the cycle used by the motor domains to extract the energy from ATP molecules. Although the two kinesins have been thought to move along the microtubule tracks in different ways, Atherton et al. find that the core mechanism used by their motor domains is the same.

When a motor domain binds to the microtubule, its shape changes, first stimulating release of the breakdown products of ATP from the previous cycle. This release makes room for a new ATP molecule to bind. The structural changes caused by ATP binding are relatively small but produce larger changes in the flexible neck region that enable individual motor domains within a kinesin pair to co-ordinate their movement and move in a consistent direction. This mechanism involves tight coupling between track binding and fuel usage and makes kinesins highly efficient motors.

A number of kinesins drive long distance transport of cellular cargo with dimerisation allowing them to take multiple 8 nm ATP-driven steps toward MT plus ends. Their processivity depends on communication between the two motor domains, which is achieved via the neck linker that connects each motor domain to the dimer-forming coiled-coil

Kinesins are a superfamily of microtubule-based

  • ATP-powered motors, important for multiple, essential cellular functions.

How microtubule binding stimulates their ATPase and controls force generation is not understood. To address this fundamental question, we visualized microtubule-bound kinesin-1 and kinesin-3 motor domains at multiple steps in their ATPase cycles—including their nucleotide-free states—at ∼7 Å resolution using cryo-electron microscopy.

All our reconstructions have, as their asymmetric unit, a triangle-shaped motor domain bound to an αβ-tubulin dimer within the MT lattice (Figure 1). The structural comparisons below are made with respect to the MT surface, which, at the resolution of our structures (∼7 Å, Table 1), is the same (CCC > 0.98 for all). As is well established across the superfamily, the major and largely invariant point of contact between kinesin motor domains and the MT is helix-α4, which lies at the tubulin intradimer interface (Figure 1C, Kikkawa et al., 2001).

However, multiple conformational changes are seen throughout the rest of each domain in response to bound nucleotide (Figure 1D). Below, we describe the conformational changes in functionally important regions of each motor domain starting with the nucleotide-binding site, from which all other conformational changes emanate.

The nucleotide-binding site (Figure 2) has three major elements: (1) the P-loop (brown) is visible in all our reconstructions;

(2) loop9 (yellow, contains switch I) undergoes major conformational changes through the ATPase cycle; and

(3) loop11 (red, contains switch II) that connects strand-β7 to helix-α4,

the conformation and flexibility of which is determined by MT binding and motor nucleotide state.

Movement and extension of helix-α6 controls neck linker docking

the N-terminus of helix-α6 is closely associated with elements of the nucleotide binding site suggesting that its conformation alters in response to different nucleotide states. In addition, because the orientation of helix-α6 with respect to helix-α4 controls neck linker docking and because helix-α4 is held against the MT during the ATPase cycle,

  • conformational changes in helix-α6 control movement of the neck linker.

Mechanical amplification and force generation involves conformational changes across the motor domain

A key conformational change in the motor domain following Mg-ATP binding is peeling of the central β-sheet from the C-terminus of helix-α4 increasing their separation (Figure 3—figure supplement 2); this is required to accommodate rotation of helix-α6 and consequent neck linker docking (Figure 3B–E).

Peeling of the central β-sheet has previously been proposed to arise from tilting of the entire motor domain relative to static MT contacts, pivoting around helix-α4 (the so-called ‘seesaw’ model; Sindelar, 2011). Specifically, this model predicts that the major difference in the motor before and after Mg-ATP binding would be the orientation of the motor domain with respect to helix-α4.

Kinesin mechanochemistry and the extent of mechanistic conservation within the motor superfamily are open questions, critical to explain how MT binding, and ATP binding and hydrolysis drive motor activity. Our structural characterisation of two transport motors now allows us to propose a model that describes the roles of mechanochemical elements that together drive conserved MT-based motor function.

Model of conserved MT-bound kinesin mechanochemistry. Loop11/N-terminus of helix-α4 is flexible in ADP-bound kinesin in solution, the neck linker is also flexible while loop9 chelates ADP. MT binding is sensed by loop11/helix-α4 N-terminus, biasing them towards more ordered conformations.

We propose that this favours crosstalk between loop11 and loop9, stimulating ADP release. In the NN conformation, both loop11 and loop9 are well ordered and primed to favour ATP binding, while helix-α6—which is required for mechanical amplification–is closely associated with the MT on the other side of the motor domain. ATP binding draws loop11 and loop9 closer together; causing

(1) tilting of most of the motor domain not contacting the MT towards the nucleotide-binding site,

(2) rotation, translation, and extension of helix-α6 which we propose contributes to force generation, and

(3) allows neck linker docking and biases movement of the 2nd head towards the MT plus end.

In both motors, microtubule binding promotes

  • ordered conformations of conserved loops that
  • stimulate ADP release,
  • enhance microtubule affinity and
  • prime the catalytic site for ATP binding.

ATP binding causes only small shifts of these nucleotide-coordinating loops but induces

  • large conformational changes elsewhere that
  • allow force generation and
  • neck linker docking towards the microtubule plus end.

Family-specific differences across the kinesin–microtubule interface account for the

  • distinctive properties of each motor.

Our data thus provide evidence for a

conserved ATP-driven

  • mechanism for kinesins and
  • reveal the critical mechanistic contribution of the microtubule interface.

Phosphorylation at endothelial cell–cell junctions: Implications for VE-cadherin function

I Timmerman, PL Hordijk, JD van Buul

Cell Health and Cytoskeleton 2010; 2: 23–31
Endothelial cell–cell junctions are strictly regulated in order to

  • control the barrier function of endothelium.

Vascular endothelial (VE)-cadherin is one of the proteins that is crucial in this process. It has been reported that

  • phosphorylation events control the function of VE-cadherin.

This review summarizes the role of VE-cadherin phosphorylation in the regulation of endothelial cell–cell junctions and highlights how this affects vascular permeability and leukocyte extravasation.

The vascular endothelium is the inner lining of blood vessels and

  • forms a physical barrier between the vessel lumen and surrounding tissue;
  • controlling the extravasation of fluids,
  • plasma proteins and leukocytes.

Changes in the permeability of the endothelium are tightly regulated. Under basal physiological conditions, there is a continuous transfer of substances across the capillary beds. In addition the endothelium can mediate inducible,

  • transient hyperpermeability
  • in response to stimulation with inflammatory mediators,
  • which takes place primarily in postcapillary venules.

However, when severe, inflammation may result in dysfunction of the endothelial barrier in various parts of the vascular tree, including large veins, arterioles and capillaries. Dysregulated permeability is observed in various pathological conditions, such as tumor-induced angiogenesis, cerebrovascular accident and atherosclerosis.

Two fundamentally different pathways regulate endothelial permeability,

  • the transcellular and paracellular pathways.

Solutes and cells can pass through the body of endothelial cells via the transcellular pathway, which includes

  • vesicular transport systems, fenestrae, and biochemical transporters.

The paracellular route is controlled by

  • the coordinated opening and closing of endothelial junctions and
  • thereby regulates traffic across the intercellular spaces between endothelial cells.

Endothelial cells are connected by

  • tight, gap and
  • adherens junctions,

of which the latter, and particularly the adherens junction component,

  • vascular endothelial (VE)-cadherin,
  • are of central importance for the initiation and stabilization of cell–cell contacts.

Although multiple adhesion molecules are localized at endothelial junctions, blocking the adhesive function of VE-cadherin using antibodies is sufficient to disrupt endothelial junctions and to increase endothelial monolayer permeability both in vitro and in vivo. Like other cadherins, VE-cadherin mediates adhesion via homophilic, calcium-dependent interactions.

This cell–cell adhesion

  • is strengthened by binding of cytoplasmic proteins, the catenins,
  • to the C-terminus of VE-cadherin.

VE-cadherin can directly bind β-catenin and plakoglobin, which

  • both associate with the actin binding protein α-catenin.

Initially, α-catenin was thought to directly anchor cadherins to the actin cytoskeleton, but recently it became clear that

  • α-catenin cannot bind to both β-catenin and actin simultaneously.

Data using purified proteins show that

  • monomeric α-catenin binds strongly to cadherin-bound β-catenin;
  • in contrast to the dimer which has a higher affinity for actin filaments,
  • indicating that α-catenin might function as a molecular switch regulating cadherin-mediated cell–cell adhesion and actin assembly.

Thus, interactions between the cadherin complex and the actin cytoskeleton are more complex than previously thought. Recently, Takeichi and colleagues reported that

  • the actin binding protein EPLIN (epithelial protein lost in neoplasm)
  • can associate with α-catenin and thereby
  • link the E-cadherin–catenin complex to the actin cytoskeleton.

Although this study was performed in epithelial cells,

  • an EPLIN-like molecule might serve as
  • a bridge between the cadherin–catenin complex and
  • the actin cytoskeleton in endothelial cells.

Next to β-catenin and plakoglobin, p120-catenin also binds directly to the intracellular tail of VE-cadherin.

Numerous lines of evidence indicate that

  • p120-catenin promotes VE-cadherin surface expression and stability at the plasma membrane.

Different models are proposed that describe how p120-catenin regulates cadherin membrane dynamics, including the hypothesis

  • that p120-catenin functions as a ‘cap’ that prevents the interaction of VE-cadherin
  • with the endocytic membrane trafficking machinery.

In addition, p120-catenin might regulate VE-cadherin internalization through interactions with small GTPases. Cytoplasmic p120-catenin, which is not bound to VE-cadherin, has been shown to

  • decrease RhoA activity,
  • elevate active Rac1 and Cdc42, and thereby is thought
  • to regulate actin cytoskeleton organization and membrane trafficking.

The intact cadherin-catenin complex is required for proper functioning of the adherens junction. Mutant forms of VE-cadherin which

  • lack either the β-catenin, plakoglobin or p120 binding regions reduce the strength of cell–cell adhesion.

Moreover, our own results showed that

  • interfering with the interaction between α-catenin and β-catenin,
  • using a cell-permeable peptide which encodes the binding site in α-catenin for β-catenin,
  • resulted in an increased permeability of the endothelial monolayer.

Several mechanisms may be involved in the regulation of the organization and function of the cadherin–catenin complex, including endocytosis of the complex, VE-cadherin cleavage and actin cytoskeleton reorganization. The remainder of this review primarily focuses on the

  • role of tyrosine phosphorylation in the control of VE-cadherin-mediated cell–cell adhesion.

Regulation of the adhesive function of VE-cadherin by tyrosine phosphorylation

It is a widely accepted concept that tyrosine phosphorylation of components of the VE–cadherin-catenin complex

  • Correlates with the weakening of cell–cell adhesion.

One of the first reports that supported this idea showed that the level of phosphorylation of VE-cadherin was

  • high in loosely confluent endothelial cells, but
  • low in tightly confluent monolayers,

when intercellular junctions are stabilized.

In addition, several conditions that induce tyrosine phosphorylation

of adherens junction components, like

  • v-Src transformation
  • and inhibition of phosphatase activity by pervanadate,

have been shown to shift cell–cell adhesion from a strong to a weak state. More physiologically relevant;

permeability-increasing agents such as

  • histamine,
  • tumor necrosis factor-α (TNF-α),
  • thrombin,
  • platelet-activating factor (PAF) and
  • vascular endothelial growth factor (VEGF)

increase tyrosine phosphorylation of various components of the cadherin–catenin complex.

A general idea has emerged that

  • tyrosine phosphorylation of the VE-cadherin complex
  • leads to the uncoupling of VE-cadherin from the actin cytoskeleton
  • through dissociation of catenins from the cadherin.

However, tyrosine phosphorylation of VE-cadherin is required for efficient transmigration of leukocytes.

This suggests that VE-cadherin-mediated cell–cell contacts

  1. are not just pushed open by the migrating leukocytes, but play
  2. a more active role in the transmigration process.

A schematic overview of leukocyte adhesion-induced signals leading to VE-cadherin phosphorylation

Regulation of the integrity of endothelial cell–cell contacts by phosphorylation of VE-cadherin

Regulation of the integrity of endothelial cell–cell contacts by phosphorylation of VE-cadherin

Regulation of the integrity of endothelial cell–cell contacts by phosphorylation of VE-cadherin.

Notes: A) Permeability-inducing agents such as thrombin, histamine and VEGF, induce tyrosine phosphorylation (pY) of VE-cadherin and the associated catenins. Although the specific consequences of catenin tyrosine phosphorylation in endothelial cells are still unknown, VE-cadherin tyrosine phosphorylation results in opening of the cell–cell junctions (indicated by arrows) and enhanced vascular permeability. How tyrosine phosphorylation affects VE-cadherin adhesiveness is not yet well understood; disrupted binding of catenins, which link the cadherin to the actin cytoskeleton, may be involved. VEGF induces phosphorylation of VE-cadherin at specific residues, Y658 and Y731, which have been reported to regulate p120-catenin and β-catenin binding, respectively. Moreover, VEGF stimulation results in serine phosphorylation (pSer) of VE-cadherin, specifically at residue S665, which leads to its endocytosis. B) Adhesion of leukocytes to endothelial cells via ICAM-1 increases endothelial permeability by inducing phosphorylation of VE-cadherin on tyrosine residues. Essential mediators, such as the kinases Pyk2 and Src, and signaling routes involving reactive oxygen species (ROS) and Rho, have been shown to act downstream of ICAM-1. Different tyrosine residues within the cytoplasmic domain of VE-cadherin are involved in the extravasation of neutrophils and lymphocytes, including Y658 and Y731. (β: β-catenin, α: α-catenin, γ: γ-catenin/plakoglobin).

N-glycosylation status of E-cadherin controls cytoskeletal dynamics through the organization of distinct β-catenin- and γ-catenin-containing AJs

BT Jamal, MN Nita-Lazar, Z Gao, B Amin, J Walker, MA Kukuruzinska
Cell Health and Cytoskeleton 2009; 1: 67–80

N-glycosylation of E-cadherin has been shown to inhibit cell–cell adhesion. Specifically, our recent studies have provided evidence that the reduction of E-cadherin N-glycosylation promoted the recruitment of stabilizing components, vinculin and serine/ threonine protein phosphatase 2A (PP2A), to adherens junctions (AJs) and enhanced the association of AJs with the actin cytoskeleton. Here, we examined the details of how

  • N-glycosylation of E-cadherin affected the molecular organization of AJs and their cytoskeletal interactions.

Using the hypoglycosylated E-cadherin variant, V13, we show that

  • V13/β-catenin complexes preferentially interacted with PP2A and with the microtubule motor protein dynein.

This correlated with dephosphorylation of the microtubule-associated protein tau, suggesting that

  • increased association of PP2A with V13-containing AJs promoted their tethering to microtubules.

On the other hand, V13/γ-catenin complexes associated more with vinculin, suggesting that they

  • mediated the interaction of AJs with the actin cytoskeleton.
  • N-glycosylation driven changes in the molecular organization of AJs were physiologically significant because transfection of V13 into A253 cancer cells, lacking both mature AJs and tight junctions (TJs), promoted the formation of stable AJs and enhanced the function of TJs to a greater extent than wild-type E-cadherin.

These studies provide the first mechanistic insights into how N-glycosylation of E-cadherin drives changes in AJ composition through

  • the assembly of distinct β-catenin- and γ-catenin-containing scaffolds that impact the interaction with different cytoskeletal components.

Cytoskeletal Basis of Ion Channel Function in Cardiac Muscle

Matteo Vatta, and Georgine Faulkner,

1 Departments of Pediatrics (Cardiology), Baylor College of Medicine, Houston, TX 2 Department of Reproductive and Developmental Sciences, University of Trieste, Trieste, Italy
3 Muscular Molecular Biology Unit, International Centre for Genetic Engineering and Biotechnology, Padriciano, Trieste, Italy

Future Cardiol. 2006 July 1; 2(4): 467–476. http://dx.doi.org:/10.2217/14796678.2.4.467

The heart is a force-generating organ that responds to

  • self-generated electrical stimuli from specialized cardiomyocytes.

This function is modulated

  • by sympathetic and parasympathetic activity.

In order to contract and accommodate the repetitive morphological changes induced by the cardiac cycle, cardiomyocytes

  • depend on their highly evolved and specialized cytoskeletal apparatus.

Defects in components of the cytoskeleton, in the long term,

  • affect the ability of the cell to compensate at both functional and structural levels.

In addition to the structural remodeling,

  • the myocardium becomes increasingly susceptible to altered electrical activity leading to arrhythmogenesis.

The development of arrhythmias secondary to structural remodeling defects has been noted, although the detailed molecular mechanisms are still elusive. Here I will review

  • the current knowledge of the molecular and functional relationships between the cytoskeleton and ion channels

and, I will discuss the future impact of new data on molecular cardiology research and clinical practice.

Myocardial dysfunction in the end-stage failing heart is very often associated with increasing

  • susceptibility to ventricular tachycardia (VT) and ventricular fibrillation (VF),

both of which are common causes of sudden cardiac death (SCD).

Among the various forms of HF,

myocardial remodeling due to ischemic cardiomyopathy (ICM) or dilated cardiomyopathy (DCM)

  • is characterized by alterations in baseline ECG,

which includes the

  • prolongation of the QT interval,
  • as well as QT dispersion,
  • ST-segment elevation, and
  • T-wave abnormalities,

especially during exercise. In particular, subjects with

severe left ventricular chamber dilation such as in DCM can have left bundle branch block (LBBB), while right bundle branch block (RBBB) is more characteristic of right ventricular failure.  LBBB and RBBB have both been repeatedly associated with AV block in heart failure.

The impact of volume overload on structural and electro-cardiographic alterations has been noted in cardiomyopathy patients treated with left ventricular assist device (LVAD) therapy, which puts the heart at mechanical rest. In LVAD-treated subjects,

  • QRS- and both QT- and QTc duration decreased,
  • suggesting that QRS- and QT-duration are significantly influenced by mechanical load and
  • that the shortening of the action potential duration contributes to the improved contractile performance after LVAD support.

Despite the increasing use of LVAD supporting either continuous or pulsatile blood flow in patients with severe HF, the benefit of this treatment in dealing with the risk of arrhythmias is still controversial.

Large epidemiological studies, such as the REMATCH study, demonstrated that the

  • employment of LVAD significantly improved survival rate and the quality of life, in comparison to optimal medical management.

An early postoperative period study after cardiac unloading therapy in 17 HF patients showed that in the first two weeks after LVAD implantation,

  • HF was associated with a relatively high incidence of ventricular arrhythmias associated with QTc interval prolongation.

In addition, a recent retrospective study of 100 adult patients with advanced HF, treated with an axial-flow HeartMate LVAD suggested that

  • the rate of new-onset monomorphic ventricular tachycardia (MVT) was increased in LVAD treated patients compared to patients given only medical treatment,

while no effect was observed on the development of polymorphic ventricular tachycardia (PVT)/ventricular fibrillation (VF).

The sarcomere

The myocardium is exposed to severe and continuous biomechanical stress during each contraction-relaxation cycle. When fiber tension remains uncompensated or simply unbalanced,

  • it may represent a trigger for arrhythmogenesis caused by cytoskeletal stretching,
  • which ultimately leads to altered ion channel localization, and subsequent action potential and conduction alterations.

Cytoskeletal proteins not only provide the backbone of the cellular structure, but they also

  • maintain the shape and flexibility of the different sub-cellular compartments, including the
  1. plasma membrane,
  2. the double lipid layer, which defines the boundaries of the cell and where
  • ion channels are mainly localized.

The interaction between the sarcomere, which is the basic for the passive force during diastole and for the restoring force during systole. Titin connects

  • the Z-line to the M-line of the sarcomeric structure
    (Figure 1).

In addition to the strategic

  • localization and mechanical spring function,
  • titin is a length-dependent sensor during
  • stretch and promotes actin-myosin interaction

Titin is stabilized by the cross-linking protein

  • telethonin (T-Cap), which localizes at the Z-line and is also part of titin sensor machinery (Figure 1).

The complex protein interactions in the sarcomere entwine telethonin to other

  • Z-line components through the family of the telethonin-binding proteins of the Z-disc, FATZ, also known as calsarcin and myozenin.

FATZ binds to

  1. calcineurin,
  2. γ-filamin as well as the
  3. spectrin-like repeats (R3–R4) of α-actinin-2,

the major component of the Z-line and a pivotal

  • F-actin cross-linker (Figure 1).contractile unit of striated muscles, and
  • the sarcolemma,

the plasma membrane surrendering the muscle fibers in skeletal muscle and the muscle cell of the cardiomyocyte,

  • determines the mechanical plasticity of the cell, enabling it to complete and re-initiate each contraction-relaxation cycle.

At the level of the sarcomere,

  • actin (thin) and myosin (thick) filaments generate the contractile force,

while other components such as titin, the largest protein known to date, are responsible for

  • the passive force during diastole and for the restoring force during systole, and (titin).
  • the Z-line to the M-line of the sarcomeric structure
    (Figure 1).

In addition to the strategic

  • localization and mechanical spring function,
  • it acts as a length-dependent sensor during stretch and
  • promotes actin-myosin interaction.

Stabilized by the cross-linking protein telethonin (T-Cap),

  • titin localizes at the Z-line and is
  • part of titin sensor machinery

Another cross-linker of α-actinin-2 in the complex Z-line scaffold is

  • the Z-band alternatively spliced PDZ motif protein (ZASP),
  • which has an important role in maintaining Z-disc stability

in skeletal and cardiac muscle (Figure 1).

ZASP contains a PDZ motif at its N-terminus,

  • which interacts with C-terminus of α-actinin-2,
  • and a conserved sequence called the ZASP like motif (ZM)
  • found in the alternatively spliced exons 4 and 6.

It has also been reported

  • to bind to the FATZ (calsarcin) family of Z-disc proteins (Figure 1).

The complex protein interactions in the sarcomere entwine telethonin to other Z-line components through the family of the telethonin-binding proteins of the

  1. Z-disc,
  2. FATZ, also known as calsarcin and
  3. myozenin

FATZ binds to calcineurin,

  1. γ-filamin as well as the
  2. spectrin-like repeats (R3–R4) of α-actinin-2, the major component of the Z-line and a pivotal F-actin cross-linker (Figure 1).
sarcomere structure

sarcomere structure

Figure 1. Sarcomere structure

The diagram illustrates the sarcomeric structure. The Z-line determines the boundaries of the contractile unit, while Titin connects the Z-line to the M-line and acts as a functional spring during contraction/relaxation cycles.

Sarcomeric Proteins and Ion Channels

In addition to systolic dysfunction characteristic of dilated cardiomyopathy (DCM) and diastolic dysfunction featuring hypertrophic cardiomyopathy (HCM), the clinical phenotype of patients with severe cardiomyopathy is very often associated with a high incidence of cardiac arrhythmias. Therefore, besides fiber stretch associated with mechanical and hemodynamic impairment, cytoskeletal alterations due to primary genetic defects or indirectly to alterations in response to cellular injury can potentially

  1. affect ion channel anchoring, and trafficking, as well as
  2. functional regulation by second messenger pathways,
  3. causing an imbalance in cardiac ionic homeostasis that will trigger arrhythmogenesis.

Intense investigation of

  • the sarcomeric actin network,
  • the Z-line structure, and
  • chaperone molecules docking in the plasma membrane,

has shed new light on the molecular basis of

  • cytoskeletal interactions in regulating ion channels.

In 1991, Cantiello et al., demonstrated that

  • although the epithelial sodium channel and F-actin are in close proximity,
  • they do not co-localize.

Actin disruption using cytochalasin D, an agent that interferes with actin polymerization, increased Na+ channel activity in 90% of excised patches tested within 2 min, which indicated that

  • the integrity of the filamentous actin (F-actin) network was essential
  • for the maintenance of normal Na+ channel function.

Later, the group of Dr. Jonathan Makielski demonstrated that

  • actin disruption induced a dramatic reduction in Na+ peak current and
  • slowed current decay without affecting steady-state voltage-dependent availability or recovery from inactivation.

These data were the first to support a role for the cytoskeleton in cardiac arrhythmias.

F-actin is intertwined in a multi-protein complex that includes

  • the composite Z-line structure.

Further, there is a direct binding between

  • the major protein of the Z-line, α-actinin-2 and
  • the voltage-gated K+ channel 1.5 (Kv1.5), (Figure 2).

The latter is expressed in human cardiomyocytes and localizes to

  • the intercalated disk of the cardiomyocyte
  • in association with connexin and N-cadherin.

Maruoka et al. treated HEK293 cells stably expressing Kv1.5 with cytochalasin D, which led to

  • a massive increase in ionic and gating IK+ currents.

This was prevented by pre-incubation with phalloidin, an F-actin stabilizing agent. In addition, the Z-line protein telethonin binds to the cytoplasmic domain of minK, the beta subunit of the potassium channel KCNQ1 (Figure 2).

Molecular interactions between the cytoskeleton and ion channels

Molecular interactions between the cytoskeleton and ion channels

Figure 2. Molecular interactions between the cytoskeleton and ion channels

The figure illustrates the interactions between the ion channels on the sarcolemma, and the sarcomere in cardiac myocytes. Note that the Z-line is connected to the cardiac T-tubules. The diagram illustrates the complex protein-protein interactions that occur between structural components of the cytoskeleton and ion channels. The cytoskeleton is involved in regulating the metabolism of ion channels, modifying their expression, localization, and electrical properties. The cardiac sodium channel Nav1.5 associates with the DGC, while potassium channels such as Kv1.5, associate with the Z-line.

Ion Channel Subunits and Trafficking

Correct localization is essential for ion channel function and this is dependent upon the ability of auxiliary proteins to

  • shuttle ion channels from the cytoplasm to their final destination such as
  • the plasma membrane or other sub-cellular compartments.

In this regard, Kvβ-subunits are

  • cytoplasmic components known to assemble with the α-subunits of voltage-dependent K+ (Kv) channels
  • at their N-terminus to form stable Kvα/β hetero-oligomeric channels.

When Kvβ is co-expressed with Kv1.4 or Kv1.5, it enhances Kv1.x channel trafficking to the cell membrane without changing the overall protein channel content. The regulatory Kvβ subunits, which are also expressed in cardiomyocytes, directly decrease K+ current by

  • accelerating Kv1.x channel inactivation.

Therefore, altered expression or mutations in Kvβ subunits could cause abnormal ion channel transport to the cell surface, thereby increasing the risk of cardiac arrhythmias.

Ion Channel Protein Motifs and Trafficking

Cell membrane trafficking in the Kv1.x family may occur in a Kvβ subunit-independent manner through specific motifs in their C-terminus. Mutagenesis of the final asparagine (N) in the Kv1.2 motif restores the leucine (L) of the Kv1.4 motif

  • re-establishing high expression levels at the plasma membrane in a Kvβ-independent manner

Cytoskeletal Proteins and Ion Channel Trafficking

Until recently, primary arrhythmias such as LQTS have been almost exclusively regarded as ion channelopathies. Other mutations have been identified with regard to channelopathies. However, the conviction that primary mutations in ion channels were solely responsible for

  • the electrical defects associated with arrhythmias

has been shaken by the identification of mutations in the

  • ANK2 gene encoding the cytoskeletal protein ankyrin-B

that is associated with LQTS in animal models and humans.

Ankyrin-B acts as a chaperone protein, which shuttles the cardiac sodium channel from the cytoplasm to the membrane. Immunohistochemical analysis has localized ankyrin-B to the Zlines/T-tubules on the plasma membrane in the myocardium. Mutations in ankyrin-B associated with LQTS

  • alter sodium channel trafficking due to loss of ankyrin-B localization at the Z-line/transverse (T)-tubules.

Reduced levels of ankyrin-B at cardiac Z-lines/T-tubules were associated with the deficiency of ankyrin-B-associated proteins such as Na/K-ATPase, Na/Ca exchanger (NCX) and inositol-1, 4, 5-trisphosphate receptors (InsP3R).

Dystrophin component of the Dystrophin Glycoprotien Complex (DGC)

Synchronized contraction is essential for cardiomyocytes, which are connected to each other via the extracellular matrix (ECM) through the DGC. The N-terminus domain of dystrophin

  • binds F-actin, and connects it to the sarcomere, while
  • the cysteine-rich (CR) C-terminus domain ensures its connection to the sarcolemma (Figure 2).

The central portion of dystrophin, the rod domain, is composed of

  • rigid spectrin-like repeats and four hinge portions (H1–H4) that determine the flexibility of the protein.

Dystrophin possesses another F-actin binding domain in the Rod domain region, between the basic repeats 11- 17 (DysN-R17).

Dystrophin, originally identified as the gene responsible for Duchenne and Becker muscular dystrophies (DMD/BMD), and later for the X-linked form of dilated cardiomyopathy (XLCM), exerts a major function in physical force transmission in striated muscle. In addition to its structural significance, dystrophin and other DGC proteins such as syntrophins are required for the

  • correct localization,
  • clustering and
  • regulation of ion channel function.

Syntrophins have been implicated in ion channel regulation.  Syntrophins contain two pleckstrin homology (PH) domains, a PDZ domain, and a syntrophin-unique (SU) C-terminal region. The interaction between syntrophins and dystrophin occurs at the PH domain distal to the syntrophin N-terminus and through the highly conserved SU domain. Conversely, the PH domain proximal to the N-terminal portion of the protein and the PDZ domain interact with other membrane components such as

  1. phosphatidyl inositol-4, 5-bisphosphate,
  2. neuronal NOS (nNOS),
  3. aquaporin-4,
  4. stress-activated protein kinase-3, and
  5. 5,

thereby linking all these molecules to the dystrophin complex (Figure 2).

Among the five known isoforms of syntrophin, the 59 KDa α1-syntrophin isoform is the most highly represented in human heart, whereas in skeletal muscle it is only present on the

  • sarcolemma of fast type II fibers.

In addition, the skeletal muscle γ2-syntrophin was found at high levels only at the

  • postsynaptic membrane of the neuromuscular junctions.

In addition to syntrophin, other scaffolding proteins such as caveolin-3 (CAV3), which is present in the caveolae, flask-shaped plasma membrane microdomains, are involved

  • in signal transduction and vesicle trafficking in myocytes,
  • modulating cardiac remodeling during heart failure.

CAV3 and α1-syntrophin, localizes at the T-tubule and are part of the DGC. In addition, α1-syntrophin binds Nav1.5, while

  • caveolin-3 binds the Na+/Ca2+ exchanger, Nav1.5 and the L-type Ca2+ channel as well as nNOS and the DGC (Figure 2).

Although ankyrin-B is the only protein found mutated in patients with primary arrhythmias, other proteins such as caveolin-3 and the syntrophins if mutated may alter ion channel function.

Conclusions

It is important to be aware of the enormous variety of clinical presentations that derive from distinct variants in the same pool of genetic factors. Knowledge of these variants could facilitate tailoring the therapy of choice for each patient. In particular, the recent findings of structural and functional links between

  • the cytoskeleton and ion channels

could expand the therapeutic interventions in

  • arrhythmia management in structurally abnormal myocardium, where aberrant binding
  • between cytoskeletal proteins can directly or indirectly alter ion channel function.

Executive Summary

Arrhythmogenesis and myocardial structure

  • Rhythm alterations can develop as a secondary consequence of myocardial structural abnormalities or as a result of a primary defect in the cardiac electric machinery.
  • Until recently, no molecular mechanism has been able to fully explain the occurrence of arrhythmogenesis in heart failure, however genetic defects that are found almost exclusively in ion channel genes account for the majority of primary arrhythmias such as long QT syndromes and Brugada syndrome. The contractile apparatus is linked to ion channels
  • The sarcomere, which represents the contractile unit of the myocardium not only generates the mechanical force necessary to exert the pump function, but also provides localization and anchorage to ion channels.
  • Alpha-actinin-2, and telethonin, two members of the Z-line scaffolding protein complex in the striated muscle associate with the potassium voltage-gated channel alpha subunit Kv1.5 and the beta subunit KCNE1 respectively.
  • Mutations in KCNE1 have previously been associated with the development of arrhythmias in LQTS subjects.
  • Mutations in both alpha-actinin-2, and telethonin were identified in individuals with cardiomyopathy. The primary defect is structural leading to ventricular dysfunction, but the secondary consequence is arrhythmia.

Ion channel trafficking and sub-cellular compartments

  • Ion channel trafficking from the endoplasmic reticulum (ER) to the Golgi complex is an important check-point for regulating the functional channel molecules on the plasma membrane. Several molecules acting as chaperones bind to and shuttle the channel proteins to their final localization on the cell surface
  • Ion channel subunits such as Kvβ enhance Kv1.x ion channel presentation on the sarcolemma. The α subunits of the Kv1.x potassium channels can be shuttled in a Kvβ-independent manner through specific sequence motif at Kv1.x protein level.
  • In addition, cytoskeletal proteins such as ankyrin-G bind Nav1.5 and are involved in the sodium channel trafficking. Another member of the ankyrin family, ankyrin-B was found mutated in patients with LQTS but the pathological mechanism of ankyrin-B mutations is still obscure, although the sodium current intensity is dramatically reduced.

The sarcolemma and ion channels

  • The sarcolemma contains a wide range of ion channels, which are responsible for the electrical propagating force in the myocardium.
  • The DGC is a protein complex, which forms a scaffold for cytoskeletal components and ion channels.
  • Dystrophin is the major component of the DGC and mutations in dystrophin and DGC cause muscular dystrophies and X-linked cardiomyopathies (XLCM) in humans. Cardiomyopathies are associated with arrhythmias
  • Caveolin-3 and syntrophins associate with Nav1.5, and are part of the DGC. Syntrophins can directly modulate Nav1.5 channel function.

Conclusions

  • The role of the cytoskeleton in ion channel function has been hypothesized in the past, but only recently the mechanism underlying the development of arrhythmias in structurally impaired myocardium has become clearer.
  • The recently acknowledged role of the cytoskeleton in ion channel function suggests that genes encoding cytoskeletal proteins should be regarded as potential candidates for variants involved in the susceptibility to arrhythmias, as well as the primary target of genetic mutations in patients with arrhythmogenic syndromes such as LQTS and Brugada syndrome.
  • Studies of genotype-phenotype correlation and and patient risk stratification for mutations in cytoskeletal proteins will help to tailor the therapy and management of patients with arrhythmias.

Read Full Post »

Pentose Shunt, Electron Transfer, Galactose, more Lipids in brief

Pentose Shunt, Electron Transfer, Galactose, more Lipids in brief

Reviewer and Curator: Larry H. Bernstein, MD, FCAP

Pentose Shunt, Electron Transfer, Galactose, and other Lipids in brief

This is a continuation of the series of articles that spans the horizon of the genetic
code and the progression in complexity from genomics to proteomics, which must
be completed before proceeding to metabolomics and multi-omics.  At this point
we have covered genomics, transcriptomics, signaling, and carbohydrate metabolism
with considerable detail.In carbohydrates. There are two topics that need some attention –
(1) pentose phosphate shunt;
(2) H+ transfer
(3) galactose.
(4) more lipids
Then we are to move on to proteins and proteomics.

Summary of this series:

The outline of what I am presenting in series is as follows:

  1. Signaling and Signaling Pathways
    http://pharmaceuticalintelligence.com/2014/08/12/signaling-and-signaling-pathways/
  2. Signaling transduction tutorial.
    http://pharmaceuticalintelligence.com/2014/08/12/signaling-transduction-tutorial/
  3. Carbohydrate metabolism
    http://pharmaceuticalintelligence.com/2014/08/13/carbohydrate-metabolism/

Selected References to Signaling and Metabolic Pathways published in this Open Access Online Scientific Journal, include the following: 

http://pharmaceuticalintelligence.com/2014/08/14/selected-references-to-signaling-
and-metabolic-pathways-in-leaders-in-pharmaceutical-intelligence/

  1. Lipid metabolism

4.1  Studies of respiration lead to Acetyl CoA
http://pharmaceuticalintelligence.com/2014/08/18/studies-of-respiration-lead-to-acetyl-coa/

4.2 The multi-step transfer of phosphate bond and hydrogen exchange energy
http://pharmaceuticalintelligence.com/2014/08/19/the-multi-step-transfer-of-phosphate-
bond-and-hydrogen-exchange-energy/

5.Pentose shunt, electron transfers, galactose, and other lipids in brief

6. Protein synthesis and degradation

7.  Subcellular structure

8. Impairments in pathological states: endocrine disorders; stress
hypermetabolism; cancer.

Section I. Pentose Shunt

Bernard L. Horecker’s Contributions to Elucidating the Pentose Phosphate Pathway

Nicole Kresge,     Robert D. Simoni and     Robert L. Hill

The Enzymatic Conversion of 6-Phosphogluconate to Ribulose-5-Phosphate
and Ribose-5-Phosphate (Horecker, B. L., Smyrniotis, P. Z., and Seegmiller,
J. E.      J. Biol. Chem. 1951; 193: 383–396

Bernard Horecker

Bernard Leonard Horecker (1914) began his training in enzymology in 1936 as a
graduate student at the University of Chicago in the laboratory of T. R. Hogness.
His initial project involved studying succinic dehydrogenase from beef heart using
the Warburg manometric apparatus. However, when Erwin Hass arrived from Otto
Warburg’s laboratory he asked Horecker to join him in the search for an enzyme
that would catalyze the reduction of cytochrome c by reduced NADP. This marked
the beginning of Horecker’s lifelong involvement with the pentose phosphate pathway.

During World War II, Horecker left Chicago and got a job at the National Institutes of
Health (NIH) in Frederick S. Brackett’s laboratory in the Division of Industrial Hygiene.
As part of the wartime effort, Horecker was assigned the task of developing a method
to determine the carbon monoxide hemoglobin content of the blood of Navy pilots
returning from combat missions. When the war ended, Horecker returned to research
in enzymology and began studying the reduction of cytochrome c by the succinic
dehydrogenase system.

Shortly after he began these investigation changes, Horecker was approached by
future Nobel laureate Arthur Kornberg, who was convinced that enzymes were the
key to understanding intracellular biochemical processes
. Kornberg suggested
they collaborate, and the two began to study the effect of cyanide on the succinic
dehydrogenase system. Cyanide had previously been found to inhibit enzymes
containing a heme group, with the exception of cytochrome c. However, Horecker
and Kornberg found that

  • cyanide did in fact react with cytochrome c and concluded that
  • previous groups had failed to perceive this interaction because
    • the shift in the absorption maximum was too small to be detected by
      visual examination.

Two years later, Kornberg invited Horecker and Leon Heppel to join him in setting up
a new Section on Enzymes in the Laboratory of Physiology at the NIH. Their Section on Enzymes eventually became part of the new Experimental Biology and Medicine
Institute and was later renamed the National Institute of Arthritis and Metabolic
Diseases.

Horecker and Kornberg continued to collaborate, this time on

  • the isolation of DPN and TPN.

By 1948 they had amassed a huge supply of the coenzymes and were able to
present Otto Warburg, the discoverer of TPN, with a gift of 25 mg of the enzyme
when he came to visit. Horecker also collaborated with Heppel on 

  • the isolation of cytochrome c reductase from yeast and 
  • eventually accomplished the first isolation of the flavoprotein from
    mammalian liver.

Along with his lab technician Pauline Smyrniotis, Horecker began to study

  • the enzymes involved in the oxidation of 6-phosphogluconate and the
    metabolic intermediates formed in the pentose phosphate pathway.

Joined by Horecker’s first postdoctoral student, J. E. Seegmiller, they worked
out a new method for the preparation of glucose 6-phosphate and 6-phosphogluconate, 
both of which were not yet commercially available.
As reported in the Journal of Biological Chemistry (JBC) Classic reprinted here, they

  • purified 6-phosphogluconate dehydrogenase from brewer’s yeast (1), and 
  • by coupling the reduction of TPN to its reoxidation by pyruvate in
    the presence of lactic dehydrogenase
    ,
  • they were able to show that the first product of 6-phosphogluconate oxidation,
  • in addition to carbon dioxide, was ribulose 5-phosphte.
  • This pentose ester was then converted to ribose 5-phosphate by a
    pentose-phosphate isomerase.

They were able to separate ribulose 5-phosphate from ribose 5- phosphate and demonstrate their interconversion using a recently developed nucleotide separation
technique called ion-exchange chromatography. Horecker and Seegmiller later
showed that 6-phosphogluconate metabolism by enzymes from mammalian
tissues also produced the same products
.8

Bernard Horecker

Bernard Horecker

http://www.jbc.org/content/280/29/e26/F1.small.gif

Over the next several years, Horecker played a key role in elucidating the

  • remaining steps of the pentose phosphate pathway.

His total contributions included the discovery of three new sugar phosphate esters,
ribulose 5-phosphate, sedoheptulose 7-phosphate, and erythrose 4-phosphate, and
three new enzymes, transketolase, transaldolase, and pentose-phosphate 3-epimerase.
The outline of the complete pentose phosphate cycle was published in 1955
(2). Horecker’s personal account of his work on the pentose phosphate pathway can
be found in his JBC Reflection (3).1

Horecker’s contributions to science were recognized with many awards and honors
including the Washington Academy of Sciences Award for Scientific Achievement in
Biological Sciences (1954) and his election to the National Academy of Sciences in
1961. Horecker also served as president of the American Society of Biological
Chemists (now the American Society for Biochemistry and Molecular Biology) in 1968.

Footnotes

  • 1 All biographical information on Bernard L. Horecker was taken from Ref. 3.
  • The American Society for Biochemistry and Molecular Biology, Inc.

References

  1. ↵Horecker, B. L., and Smyrniotis, P. Z. (1951) Phosphogluconic acid dehydrogenase
    from yeast. J. Biol. Chem. 193, 371–381FREE Full Text
  2. Gunsalus, I. C., Horecker, B. L., and Wood, W. A. (1955) Pathways of carbohydrate
    metabolism in microorganisms. Bacteriol. Rev. 19, 79–128  FREE Full Text
  3. Horecker, B. L. (2002) The pentose phosphate pathway. J. Biol. Chem. 277, 47965–
    47971 FREE Full Text

The Pentose Phosphate Pathway (also called Phosphogluconate Pathway, or Hexose
Monophosphate Shunt) is depicted with structures of intermediates in Fig. 23-25
p. 863 of Biochemistry, by Voet & Voet, 3rd Edition. The linear portion of the pathway
carries out oxidation and decarboxylation of glucose-6-phosphate, producing the
5-C sugar ribulose-5-phosphate.

Glucose-6-phosphate Dehydrogenase catalyzes oxidation of the aldehyde
(hemiacetal), at C1 of glucose-6-phosphate, to a carboxylic acid in ester linkage
(lactone). NADPserves as electron acceptor.

6-Phosphogluconolactonase catalyzes hydrolysis of the ester linkage (lactone)
resulting in ring opening. The product is 6-phosphogluconate. Although ring opening
occurs in the absence of a catalyst, 6-Phosphogluconolactonase speeds up the
reaction, decreasing the lifetime of the highly reactive, and thus potentially
toxic, 6-phosphogluconolactone.

Phosphogluconate Dehydrogenase catalyzes oxidative decarboxylation of
6-phosphogluconate, to yield the 5-C ketose ribulose-5-phosphate. The
hydroxyl at C(C2 of the product) is oxidized to a ketone. This promotes loss
of the carboxyl at C1 as CO2.  NADP+ again serves as oxidant (electron acceptor).

pglucose hd

pglucose hd

https://www.rpi.edu/dept/bcbp/molbiochem/MBWeb/mb2/part1/images/pglucd.gif

Reduction of NADP+ (as with NAD+) involves transfer of 2e- plus 1H+ to the
nicotinamide moiety.

nadp

NADPH, a product of the Pentose Phosphate Pathway, functions as a reductant in
various synthetic (anabolic) pathways, including fatty acid synthesis.

NAD+ serves as electron acceptor in catabolic pathways in which metabolites are
oxidized. The resultant NADH is reoxidized by the respiratory chain, producing ATP.

nadnadp

https://www.rpi.edu/dept/bcbp/molbiochem/MBWeb/mb2/part1/images/nadnadp.gif

Regulation: 
Glucose-6-phosphate Dehydrogenase is the committed step of the Pentose
Phosphate Pathway. This enzyme is regulated by availability of the substrate NADP+.
As NADPH is utilized in reductive synthetic pathways, the increasing concentration of
NADP+ stimulates the Pentose Phosphate Pathway, to replenish NADPH.

The remainder of the Pentose Phosphate Pathway accomplishes conversion of the
5-C ribulose-5-phosphate to the 5-C product ribose-5-phosphate, or to the 3-C
glyceraldehyde -3-phosphate and the 6-C fructose-6-phosphate (reactions 4 to 8
p. 863).

Transketolase utilizes as prosthetic group thiamine pyrophosphate (TPP), a
derivative of vitamin B1.

tpp

tpp

https://www.rpi.edu/dept/bcbp/molbiochem/MBWeb/mb2/part1/images/tpp.gif

Thiamine pyrophosphate binds at the active sites of enzymes in a “V” conformation.The amino group of the aminopyrimidine moiety is close to the dissociable proton,
and serves as the proton acceptor. This proton transfer is promoted by a glutamate
residue adjacent to the pyrimidine ring.

The positively charged N in the thiazole ring acts as an electron sink, promoting
C-C bond cleavage. The 3-C aldose glyceraldehyde-3-phosphate is released.
2-C fragment remains on TPP.

FASEB J. 1996 Mar;10(4):461-70.   http://www.ncbi.nlm.nih.gov/pubmed/8647345

Reviewer

The importance of this pathway can easily be underestimated.  The main source for
energy in respiration was considered to be tied to the

  • high energy phosphate bond in phosphorylation and utilizes NADPH, converting it to NADP+.

glycolysis n skeletal muscle in short term, dependent on muscle glycogen conversion
to glucose, and there is a buildup of lactic acid – used as fuel by the heart.  This
pathway accounts for roughly 5% of metabolic needs, varying between tissues,
depending on there priority for synthetic functions, such as endocrine or nucleic
acid production.

The mature erythrocyte and the ocular lens both are enucleate.  85% of their
metabolic energy needs are by anaerobic glycolysis.  Consider the erythrocyte
somewhat different than the lens because it has iron-based hemoglobin, which
exchanges O2 and CO2 in the pulmonary alveoli, and in that role, is a rapid
regulator of H+ and pH in the circulation (carbonic anhydrase reaction), and also to
a lesser extent in the kidney cortex, where H+ is removed  from the circulation to
the urine, making the blood less acidic, except when there is a reciprocal loss of K+.
This is how we need a nomogram to determine respiratory vs renal acidosis or
alkalosis.  In the case of chronic renal disease, there is substantial loss of
functioning nephrons, loss of countercurrent multiplier, and a reduced capacity to
remove H+.  So there is both a metabolic acidosis and a hyperkalemia, with increased
serum creatinine, but the creatinine is only from muscle mass – not accurately
reflecting total body mass, which includes visceral organs.  The only accurate
measure of lean body mass would be in the linear relationship between circulating
hepatic produced transthyretin (TTR).

The pentose phosphate shunt is essential for

  • the generation of nucleic acids, in regeneration of red cells and lens – requiring NADPH.

Insofar as the red blood cell is engaged in O2 exchange, the lactic dehydrogenase
isoenzyme composition is the same as the heart. What about the lens of and cornea the eye, and platelets?  The explanation does appear to be more complex than
has been proposed and is not discussed here.

Section II. Mitochondrial NADH – NADP+ Transhydrogenase Reaction

There is also another consideration for the balance of di- and tri- phospopyridine
nucleotides in their oxidized and reduced forms.  I have brought this into the
discussion because of the centrality of hydride tranfer to mitochondrial oxidative
phosphorylation and the energetics – for catabolism and synthesis.

The role of transhydrogenase in the energy-linked reduction of TPN 

Fritz HommesRonald W. Estabrook∗∗

The Wenner-Gren Institute, University of Stockholm
Stockholm, Sweden
Biochemical and Biophysical Research Communications 11, (1), 2 Apr 1963, Pp 1–6
http://dx.doi.org:/10.1016/0006-291X(63)90017-2

In 1959, Klingenberg and Slenczka (1) made the important observation that incubation of isolated

  • liver mitochondria with DPN-specific substrates or succinate in the absence of phosphate
    acceptor resulted in a rapid and almost complete reduction of the intramitochondrial TPN.

These and related findings led Klingenberg and co-workers (1-3) to postulate

  • the occurrence of an ATP-controlled transhydrogenase reaction catalyzing the reduction of
    mitochondrial TPN by DPNH. A similar conclusion was reached by Estabrook and Nissley (4).

The present paper describes the demonstration and some properties of an

  • energy-dependent reduction of TPN by DPNH, catalyzed by submitochondrial particles.

Preliminary reports of some of these results have already appeared (5, 6 ) , and a
complete account is being published elsewhere (7).We have studied the energy- dependent reduction of TPN by PNH with submitochondrial particles from both
rat liver and beef heart. Rat liver particles were prepared essentially according to
the method of Kielley and Bronk (8), and beef heart particles by the method of
Low and Vallin (9).

PYRIDINE NUCLEOTIDE TRANSHYDROGENASE  II. DIRECT EVIDENCE FOR
AND MECHANISM OF THE
 TRANSHYDROGENASE REACTION*

BY  NATHAN 0. KAPLAN, SIDNEY P. COLOWICK, AND ELIZABETH F. NEUFELD
(From the McCollum-Pratt Institute, The Johns Hopkins University, Baltimore,
Maryland)  J. Biol. Chem. 1952, 195:107-119.
http://www.jbc.org/content/195/1/107.citation

NO Kaplan

NO Kaplan

Sidney Colowick

Sidney Colowick

Elizabeth Neufeld

Elizabeth Neufeld

Kaplan studied carbohydrate metabolism in the liver under David M. Greenberg at the
University of California, Berkeley medical school. He earned his Ph.D. in 1943. From
1942 to 1944, Kaplan participated in the Manhattan Project. From 1945 to 1949,
Kaplan worked with Fritz Lipmann at Massachusetts General Hospital to study
coenzyme A. He worked at the McCollum-Pratt Institute of Johns Hopkins University
from 1950 to 957. In 1957, he was recruited to head a new graduate program in
biochemistry at Brandeis University. In 1968, Kaplan moved to the University of
California, San Diego
, where he studied the role of lactate dehydrogenase in cancer. He also founded a colony of nude mice, a strain of laboratory mice useful in the study
of cancer and other diseases. [1] He was a member of the National Academy of
Sciences.One of Kaplan’s students at the University of California was genomic
researcher Craig Venter.[2]3]  He was, with Sidney Colowick, a founding editor of the scientific book series Methods
in Enzymology
.[1]

http://books.nap.edu/books/0309049768/xhtml/images/img00009.jpg

Colowick became Carl Cori’s first graduate student and earned his Ph.D. at
Washington University St. Louis in 1942, continuing to work with the Coris (Nobel
Prize jointly) for 10 years. At the age of 21, he published his first paper on the
classical studies of glucose 1-phosphate (2), and a year later he was the sole author on a paper on the synthesis of mannose 1-phosphate and galactose 1-phosphate (3). Both papers were published in the JBC. During his time in the Cori lab,

Colowick was involved in many projects. Along with Herman Kalckar he discovered
myokinase (distinguished from adenylate kinase from liver), which is now known as
adenyl kinase. This discovery proved to be important in understanding transphos-phorylation reactions in yeast and animal cells. Colowick’s interest then turned to
the conversion of glucose to polysaccharides, and he and Earl Sutherland (who
will be featured in an upcoming JBC Classic) published an important paper on the
formation of glycogen from glucose using purified enzymes (4). In 1951, Colowick
and Nathan Kaplan were approached by Kurt Jacoby of Academic Press to do a
series comparable to Methodem der Ferment Forschung. Colowick and Kaplan
planned and edited the first 6 volumes of Methods in Enzymology, launching in 1955
what became a series of well known and useful handbooks. He continued as
Editor of the series until his death in 1985.

http://bioenergetics.jbc.org/highwire/filestream/9/field_highwire_fragment_image_s/0/F1.small.gif

The Structure of NADH: the Work of Sidney P. Colowick

Nicole KresgeRobert D. Simoni and Robert L. Hill

On the Structure of Reduced Diphosphopyridine Nucleotide

(Pullman, M. E., San Pietro, A., and Colowick, S. P. (1954)

J. Biol. Chem. 206, 129–141)

Elizabeth Neufeld
·  Born: September 27, 1928 (age 85), Paris, France
·  EducationQueens College, City University of New YorkUniversity of California,
Berkeley

http://fdb5.ctrl.ucla.edu/biological-chemistry/institution/photo?personnel%5fid=45290&max_width=155&max_height=225

In Paper I (l), indirect evidence was presented for the following transhydrogenase
reaction, catalyzed by an enzyme present in extracts of Pseudomonas
fluorescens:

TPNHz + DPN -+ TPN + DPNHz

The evidence was obtained by coupling TPN-specific dehydrogenases with the
transhydrogenase and observing the reduction of large amounts of diphosphopyridine nucleotide (DPN) in the presence of catalytic amounts of triphosphopyridine
nucleotide (TPN).

In this paper, data will be reported showing the direct

  • interaction between TPNHz and DPN, in thepresence of transhydrogenase alone,
  • to yield products having the propertiesof TPN and DPNHZ.

Information will be given indicating that the reaction involves

  • a transfer of electrons (or hydrogen) rather than a phosphate 

Experiments dealing with the kinetics and reversibility of the reaction, and with the
nature of the products, suggest that the reaction is a complex one, not fully described
by the above formulation.

Materials and Methods [edited]

The TPN and DPN used in these studies were preparations of approximately 75
percent purity and were prepared from sheep liver by the chromatographic procedure
of Kornberg and Horecker (unpublished). Reduced DPN was prepared enzymatically with alcohol dehydrogenase as described elsewhere (2). Reduced TPN was prepared by treating TPN with hydrosulfite. This treated mixture contained 2 pM of TPNHz per ml.
The preparations of desamino DPN and reduced desamino DPN have been
described previously (2, 3). Phosphogluconate was a barium salt which was kindly
supplied by Dr. B. F. Horecker. Cytochrome c was obtained from the Sigma Chemical Company.

Transhydrogenase preparations with an activity of 250 to 7000 units per mg. were
used in these studies. The DPNase was a purified enzyme, which was obtained
from zinc-deficient Neurospora and had an activity of 5500 units per mg. (4). The
alcohol dehydrogenase was a crystalline preparation isolated from yeast according to the procedure of Racker (5).

Phosphogluconate dehydrogenase from yeast and a 10 per cent pure preparation of the TPN-specific cytochrome c reductase from liver (6) were gifts of Dr. B. F.
Horecker.

DPN was assayed with alcohol and crystalline yeast alcohol dehydrogenase. TPN was determined By the specific phosphogluconic acid dehydrogenase from yeast and also by the specific isocitric dehydrogenase from pig heart. Reduced DPN was
determined by the use of acetaldehyde and the yeast alcohol dehydrogenase.
All of the above assays were based on the measurement of optical density changes
at 340 rnp. TPNHz was determined with the TPN-specific cytochrome c reductase system. The assay of the reaction followed increase in optical density at 550 rnp  as a measure of the reduction of the cytochrome c after cytochrome c
reductase was added to initiate the reaction. The changes at 550 rnp are plotted for different concentrations of TPNHz in Fig. 3, a. The method is an extremely sensitive and accurate assay for reduced TPN.

Results
[No Figures or Table shown]

Formation of DPNHz from TPNHz and DPN-Fig. 1, a illustrates the direct reaction between TPNHz and DPN to form DPNHZ. The reaction was carried out by incubating TPNHz with DPN in the presence of the
transhydrogenase, yeast alcohol dehydrogenase, and acetaldehyde. Since the yeast dehydrogenase is specific for DPN,

  • a decrease in absorption at340 rnp can only be due to the formation of reduced DPN. It can
    be seen from the curves in Fig. 1, a that a decrease in optical density occurs only in the
    presence of the complete system.

The Pseudomonas enzyme is essential for the formation of DPNH2. It is noteworthy
that, under the conditions of reaction in Fig. 1, a,

  • approximately 40 per cent of theTPNH, reacted with the DPN.

Fig. 1, a also indicates that magnesium is not required for transhydrogenase activity.  The reaction between TPNHz and DPN takes place in the absence of alcohol
dehydrogenase and acetaldehyde
. This can be demonstrated by incubating the
two pyridine nucleotides with the transhydrogenase for 4 8 12 16 20 24 28 32 36
minutes

FIG. 1. Evidence for enzymatic reaction of TPNHt with DPN.

  • Rate offormation of DPNH2.

(b) DPN disappearance and TPN formation.

(c) Identification of desamino DPNHz as product of reaction of TPNHz with desamino DPN.  (assaying for reduced DPN by the yeast alcohol dehydrogenase technique.

Table I (Experiment 1) summarizes the results of such experiments in which TPNHz was added with varying amounts of DPN.

  • In the absence of DPN, no DPNHz was formed. This eliminates the possibility that TPNH 2 is
    converted to DPNHz
  • by removal ofthe monoester phosphate grouping.

The data also show that the extent of the reaction is

  • dependent on the concentration of DPN.

Even with a large excess of DPN, only approximately 40 per cent of the TPNHzreacts to form reduced DPN. It is of importance to emphasize that in the above
experiments, which were carried out in phosphate buffer, the extent of  the reaction

  • is the same in the presence or absence of acetaldehyde andalcohol dehydrogenase.

With an excess of DPN and different  levels of TPNHZ,

  • the amount of reduced DPN which is formed is
  • dependent on the concentration of TPNHz(Table I, Experiment 2).
  • In all cases, the amount of DPNHz formed is approximately
    40 per cent of the added reduced TPN.

Formation of TPN-The reaction between TPNHz and DPN should yield TPN as well as DPNHz.
The formation of TPN is demonstrated in Table 1. in Fig. 1, b. In this experiment,
TPNHz was allowed to react with DPN in the presence of the transhydrogenase
(PS.), and then alcohol and alcohol dehydrogenase were added . This
would result in reduction of the residual DPN, and the sample incubated with the
transhydrogenase contained less DPN. After the completion of the alcohol
dehydrogenase reaction, phosphogluconate and phosphogluconic dehydrogenase (PGAD) were added to reduce the TPN. The addition of this TPN-specific
dehydrogenase results in an

  • increase inoptical density in the enzymatically treated sample.
  • This change represents the amount of TPN formed.

It is of interest to point out that, after addition of both dehydrogenases,

  • the total optical density change is the same in both

Therefore it is evident that

  • for every mole of DPN disappearing  a mole of TPN appears.

Balance of All Components of Reaction

Table II (Experiment 1) shows that,

  • if measurements for all components of the reaction are made, one can demonstrate
    that there is
  • a mole for mole disappearance of TPNH, and DPN, and
  • a stoichiometric appearance of TPN and DPNH2.
  1. The oxidized forms of the nucleotides were assayed as described
  2. the reduced form of TPN was determined by the TPNHz-specific cytochrome c reductase,
  3. the DPNHz by means of yeast alcohol dehydrogenase plus

This stoichiometric balance is true, however,

  • only when the analyses for the oxidized forms are determined directly on the reaction

When analyses are made after acidification of the incubated reaction mixture,

  • the values found forDPN and TPN are much lower than those obtained by direct analysis.

This discrepancy in the balance when analyses for the oxidized nucleotides are
carried out in acid is indicated in Table II (Experiment 2). The results, when
compared with the findings in Experiment 1, are quite striking.

Reaction of TPNHz with Desamino DPN

Desamino DPN

  • reacts with the transhydrogenase system at the same rate as does DPN (2).

This was of value in establishing the fact that

  • the transhydrogenase catalyzesa transfer of hydrogen rather than a phosphate transfer reaction.

The reaction between desamino DPN and TPNHz can be written in two ways.

TPN f desamino DPNHz

TPNH, + desamino DPN

DPNH2 + desamino TPN

If the reaction involved an electron transfer,

  • desamino DPNHz would be
  • Phosphate transfer would result in the production of reduced

Desamino DPNHz can be distinguished from DPNHz by its

  • slowerrate of reaction with yeast alcohol dehydrogenase (2, 3).

Fig. 1, c illustrates that, when desamino DPN reacts with TPNH2, 

  • the product of the reaction is desamino DPNHZ.

This is indicated by the slow rate of oxidation of the product by yeast alcohol
dehydrogenase and acetaldehyde.

From the above evidence phosphate transfer 

  • has been ruled out as a possible mechanism for the transhydrogenase reaction.

Inhibition by TPN

As mentioned in Paper I and as will be discussed later in this paper,

  • the transhydrogenase reaction does not appear to be readily reversible.

This is surprising, particularly since only approximately 

  • 40 per cent of the TPNHz undergoes reaction with DPN
    under the conditions described above. It was therefore thought that
  • the TPN formed might inhibit further transfer of electrons from TPNH2.

Table III summarizes data showing the

  • strong inhibitory effect of TPN on thereaction between TPNHz and DPN.

It is evident from the data that

  • TPN concentration is a factor in determining the extent of the reaction.

Effect of Removal of TPN on Extent of Reaction

A purified DPNase from Neurospora has been found

  • to cleave the nicotinamide riboside linkagesof the oxidized forms of both TPN and DPN
  • without acting on thereduced forms of both nucleotides (4).

It has been found, however, that

  • the DPNase hydrolyzes desamino DPN at a very slow rate (3).

In the reaction between TPNHz and desamino DPN, TPN and desamino DPNH:,

  • TPNis the only component of this reaction attacked by the Neurospora enzyme
    at an appreciable rate

It was  thought that addition of the DPNase to the TPNHZ-desamino DPN trans-
hydrogenase reaction mixture

  • would split the TPN formed andpermit the reaction to go to completion.

This, indeed, proved to be the case, as indicated in Table IV, where addition of
the DPNase with desamino DPN results in almost

  • a stoichiometric formation of desamino DPNHz
  • and a complete disappearance of TPNH2.

Extent of Reaction in Buffers Other Than Phosphate

All the reactions described above were carried out in phosphate buffer of pH 7.5.
If the transhydrogenase reaction between TPNHz and DPN is run at the same pH
in tris(hydroxymethyl)aminomethane buffer (TRIS buffer)

  • with acetaldehydeand alcohol dehydrogenase present,
  • the reaction proceeds muchfurther toward completion 
  • than is the case under the same conditions ina phosphate medium (Fig. 2, a).

The importance of phosphate concentration in governing the extent of the reaction
is illustrated in Fig. 2, b.

In the presence of TRIS the transfer reaction

  • seems to go further toward completion in the presence of acetaldehyde
    and 
    alcohol dehydrogenase
  • than when these two components are absent.

This is not true of the reaction in phosphate,

  • in which the extent is independent of the alcoholdehydrogenase system.

Removal of one of the products of the reaction (DPNHp) in TRIS thus

  • appears to permit the reaction to approach completion,whereas
  • in phosphate this removal is without effect on the finalcourse of the reaction.

The extent of the reaction in TRIS in the absence of alcohol dehydrogenase
and acetaldehyde
 is

  • somewhat greater than when the reaction is run in phosphate.

TPN also inhibits the reaction of TPNHz with DPN in TRIS medium, but the inhibition

  • is not as marked as when the reaction is carried out in phosphate buffer.

Reversibility of Transhydrogenase Reaction;

Reaction between DPNHz and TPN

In Paper I, it was mentioned that no reversal of the reaction could be achieved in a system containing alcohol, alcohol dehydrogenase, TPN, and catalytic amounts of
DPN.

When DPNH, and TPN are incubated with the purified transhydrogenase, there is
also

  • no evidence for reversibility.

This is indicated in Table V which shows that

  • there is no disappearance of DPNHz in such a system.

It was thought that removal of the TPNHz, which might be formed in the reaction,
could promote the reversal of the reaction. Hence,

  • by using the TPNHe-specific cytochrome c reductase, one could
  1. not only accomplishthe removal of any reduced TPN,
  2. but also follow the course of the reaction.

A system containing DPNH2, TPN, the transhydrogenase, the cytochrome c
reductase, and cytochrome c, however, gives

  • no reduction of the cytochrome

This is true for either TRIS or phosphate buffers.2

Some positive evidence for the reversibility has been obtained by using a system
containing

  • DPNH2, TPNH2, cytochrome c, and the cytochrome creductase in TRIS buffer.

In this case, there is, of course, reduction of cytochrome c by TPNHZ, but,

  • when the transhydrogenase is present.,there is
  • additional reduction over and above that due to the added TPNH2.

This additional reduction suggests that some reversibility of the reaction occurred
under these conditions. Fig. 3, b shows

  • the necessity of DPNHzfor this additional reduction.

Interaction of DPNHz with Desamino DPN-

If desamino DPN and DPNHz are incubated with the purified Pseudomonas enzyme,
there appears

  • to be a transfer of electrons to form desamino DPNHz.

This is illustrated in Fig. 4, a, which shows the

  • decreased rate of oxidation by thealcohol dehydrogenase system
  • after incubation with the transhydrogenase.
  • Incubation of desamino DPNHz with DPN results in the formation of DPNH2,
  • which is detected by the faster rate of oxidation by the alcohol dehydrogenase system
  • after reaction of the pyridine nucleotides with thetranshydrogenase (Fig. 4, b).

It is evident from the above experiments that

the transhydrogenase catalyzes an exchange of hydrogens between

  • the adenylic and inosinic pyridine nucleotides.

However, it is difficult to obtain any quantitative information on the rate or extent of
the reaction by the method used, because

  • desamino DPNHz also reacts with the alcohol dehydrogenase system,
  • although at a much slower rate than does DPNH2.

DISCUSSION

The results of the balance experiments seem to offer convincing evidence that
the transhydrogenase catalyzes the following reaction.

TPNHz + DPN -+ DPNHz + TPN

Since desamino DPNHz is formed from TPNHz and desamino DPN,

  • thereaction appears to involve an electron (or hydrogen) transfer
  • rather thana transfer of the monoester phosphate grouping of TPN.

A number of the findings reported in this paper are not readily understandable in
terms of the above simple formulation of the reaction. It is difficult to understand
the greater extent of the reaction in TRIS than in phosphate when acetaldehyde
and alcohol dehydrogenase are present.

One possibility is that an intermediate may be involved which is more easily converted
to reduced DPN in the TRIS medium. The existence of such an intermediate is also
suggested by the discrepancies noted in balance experiments, in which

  • analyses of the oxidized nucleotides after acidification showed
  • much lower values than those found by direct analysis.

These findings suggest that the reaction may involve

  • a 1 electron ratherthan a 2 electron transfer with
  • the formation of acid-labile free radicals as intermediates.

The transfer of hydrogens from DPNHz to desamino DPN

  • to yield desamino DPNHz and DPN and the reversal of this transfer
  • indicate the unique role of the transhydrogenase
  • in promoting electron exchange between the pyridine nucleotides.

In this connection, it is of interest that alcohol dehydrogenase and lactic
dehydrogenase cannot duplicate this exchange  between the DPN and
the desamino systems.3  If one assumes that desamino DPN behaves
like DPN,

  • one might predict that the transhydrogenase would catalyze an
    exchange of electrons (or hydrogen) 3.

Since alcohol dehydrogenase alone

  • does not catalyze an exchange of electrons between the adenylic
    and inosinic pyridine nucleotides, this rules out the possibility
  • that the dehydrogenase is converted to a reduced intermediate
  • during electron between DPNHz and added DPN.

It is hoped to investigate this possibility with isotopically labeled DPN.
Experiments to test the interaction between TPN and desamino TPN are
also now in progress.

It seems likely that the transhydrogenase will prove capable of

  • catalyzingan exchange between TPN and TPNH2, as well as between DPN and

The observed inhibition by TPN of the reaction between TPNHz and DPN may
therefore

  • be due to a competition between DPN and TPNfor the TPNH2.

SUMMARY

  1. Direct evidence for the following transhydrogenase reaction. catalyzedby an
    enzyme from Pseudomonas fluorescens, is presented.

TPNHz + DPN -+ TPN + DPNHz

Balance experiments have shown that for every mole of TPNHz disappearing
1 mole of TPN appears and that for each mole of DPNHz generated 1 mole of
DPN disappears. The oxidized nucleotides found at the end of the reaction,
however, show anomalous lability toward acid.

  1. The transhydrogenase also promotes the following reaction.

TPNHz + desamino DPN -+ TPN + desamino DPNH,

This rules out the possibility that the transhydrogenase reaction involves a
phosphate transfer and indicates that the

  • enzyme catalyzes a shift of electrons (or hydrogen atoms).

The reaction of TPNHz with DPN in 0.1 M phosphate buffer is strongly
inhibited by TPN; thus

  • it proceeds only to the extent of about40 per cent or less, even
  • when DPNHz is removed continuously by meansof acetaldehyde
    and alcohol dehydrogenase.
  • In other buffers, in whichTPN is less inhibitory, the reaction proceeds
    much further toward completion under these conditions.
  • The reaction in phosphate buffer proceedsto completion when TPN
    is removed as it is formed.
  1. DPNHz does not react with TPN to form TPNHz and DPN in the presence
    of transhydrogenase. Some evidence, however, has been obtained for
    the reversibility by using the following system:
  • DPNHZ, TPNHZ, cytochromec, the TPNHz-specific cytochrome c reductase,
    and the transhydrogenase.
  1. Evidence is cited for the following reversible reaction, which is catalyzed
    by the transhydrogenase.

DPNHz + desamino DPN fi DPN + desamino DPNHz

  1. The results are discussed with respect to the possibility that the
    transhydrogenase reaction may
  • involve a 1 electron transfer with theformation of free radicals as intermediates.

 

BIBLIOGRAPHY

  1. Coiowick, S. P., Kaplan, N. O., Neufeld, E. F., and Ciotti, M. M., J. Biol. Chem.,196, 95 (1952).
  2. Pullman, 111. E., Colowick, S. P., and Kaplan, N. O., J. Biol. Chem., 194, 593(1952).
  3. Kaplan, N. O., Colowick, S. P., and Ciotti, M. M., J. Biol. Chem., 194, 579 (1952).
  4. Kaplan, N. O., Colowick, S. P., and Nason, A., J. Biol. Chem., 191, 473 (1951).
  5. Racker, E., J. Biol. Chem., 184, 313 (1950).
  6. Horecker, B. F., J. Biol. Chem., 183, 593 (1950).

Section !II. 

Luis_Federico_Leloir_-_young

The Leloir pathway: a mechanistic imperative for three enzymes to change
the stereochemical configuration of a single carbon in galactose.

Frey PA.
FASEB J. 1996 Mar;10(4):461-70.    http://www.fasebj.org/content/10/4/461.full.pdf
PMID:8647345

The biological interconversion of galactose and glucose takes place only by way of
the Leloir pathway and requires the three enzymes galactokinase, galactose-1-P
uridylyltransferase, and UDP-galactose 4-epimerase.
The only biological importance of these enzymes appears to be to

  • provide for the interconversion of galactosyl and glucosyl groups.

Galactose mutarotase also participates by producing the galactokinase substrate
alpha-D-galactose from its beta-anomer. The galacto/gluco configurational change takes place at the level of the nucleotide sugar by an oxidation/reduction
mechanism in the active site of the epimerase NAD+ complex. The nucleotide portion
of UDP-galactose and UDP-glucose participates in the epimerization process in two ways:

1) by serving as a binding anchor that allows epimerization to take place at glycosyl-C-4 through weak binding of the sugar, and

2) by inducing a conformational change in the epimerase that destabilizes NAD+ and
increases its reactivity toward substrates.

Reversible hydride transfer is thereby facilitated between NAD+ and carbon-4
of the weakly bound sugars.

The structure of the enzyme reveals many details of the binding of NAD+ and
inhibitors at the active site
.

The essential roles of the kinase and transferase are to attach the UDP group
to galactose, allowing for its participation in catalysis by the epimerase. The
transferase is a Zn/Fe metalloprotein
, in which the metal ions stabilize the
structure rather than participating in catalysis. The structure is interesting
in that

  • it consists of single beta-sheet with 13 antiparallel strands and 1 parallel strand
    connected by 6 helices.

The mechanism of UMP attachment at the active site of the transferase is a double
displacement
, with the participation of a covalent UMP-His 166-enzyme intermediate
in the Escherichia coli enzyme. The evolution of this mechanism appears to have
been guided by the principle of economy in the evolution of binding sites.

PMID: 8647345 Free full text

Section IV.

More on Lipids – Role of lipids – classification

  • Energy
  • Energy Storage
  • Hormones
  • Vitamins
  • Digestion
  • Insulation
  • Membrane structure: Hydrophobic properties

Lipid types

lipid types

lipid types

nat occuring FAs in mammals

nat occuring FAs in mammals

Read Full Post »

Signaling transduction tutorial

Larry H. Bernstein, MD, FCAP, Reporter and Curator
Leaders in Pharmaceutical Intelligence

http://pharmaceuticalintelligence.com/8-10-2014/Signaling transduction tutorial

This portion of the discussion is a series of articles on signaling and signaling pathways. Many of the protein-protein interactions or protein-membrane interactions and associated regulatory features have been referred to previously, but the focus of the discussion or points made were different.  I considered placing this after the discussion of proteins and how they play out their essential role, but this is quite a suitable place for a progression to what follows.  This is introduced by material taken from Wikipedia, which will be followed by a series of mechanisms and examples from the current literature, which give insight into the developments in cell metabolism, with the later goal of separating views introduced by molecular biology and genomics from functional cellular dynamics that are not dependent on the classic view.  The work is vast, and this discussion does not attempt to cover it in great depth.  It is the first in a series.  This discussion, in particular is a tutorial on signaling transduction that was already available, and relevant.  One may note that all of the slides used herein were also used in the previous blog, but in a different construction.  I shall tweak the contents, as I find helpful.

  1. Signaling and signaling pathways
  2. Signaling transduction tutorial.
  3. Carbohydrate metabolism
  4. Lipid metabolism
  5. Protein synthesis and degradation
  6. Subcellular structure
  7. Impairments in pathological states: endocrine disorders; stress hypermetabolism; cancer.

 

Signal Transduction Tutorial

The goal of this tutorial is for you to gain an understanding of how cell signaling occurs in a cell.  Upon completion of the tutorial,

  • you will have a basic understanding signal transduction and
  • the role of phosphorylation in signal transduction.

You will also have detailed knowledge of

  • the role of Tyrosine kinases and
  • G protein-coupled receptors in cell signaling.
  1. Description of Signal Transduction

As living organisms

  • we are constantly receiving and interpreting signals from our environment.

These signals can come

  • in the form of light, heat, odors, touch or sound.

The cells of our bodies are also

  • constantly receiving signals from other cells.

These signals are important to

  • keep cells alive and functioning as well as
  • to stimulate important events such as
  • cell division and differentiation.

Signals are most often chemicals that can be found

  • in the extracellular fluid around cells.

These chemicals can come

  • from distant locations in the body (endocrine signaling by hormones), from
  • nearby cells (paracrine signaling) or can even
  • be secreted by the same cell (autocrine signaling).
intercellular signaling

intercellular signaling

http://www.hartnell.edu/tutorials/biology/images/intercellularsignaling.jpg

Signaling molecules may trigger any number of cellular responses, including

  • changing the metabolism of the cell receiving the signal or
  • result in a change in gene expression (transcription) within the nucleus of the cell or both.

Overview of Cell Signaling

Cell signaling can be divided into 3 stages.

  1. Reception: A cell detects a signaling molecule from the outside of the cell. A signal is detected when the chemical signal (also known as a ligand) binds to a receptor protein on the surface of the cell or inside the cell.
  2. Transduction: When the signaling molecule binds the receptor it changes the receptor protein in some way. This change initiates the process of transduction. Signal transduction is usually a pathway of several steps. Each relay molecule in the signal transduction pathway changes the next molecule in the pathway.
  3. Response: Finally, the signal triggers a specific cellular response.
signal transduction_simple

signal transduction_simple

http://www.hartnell.edu/tutorials/biology/images/signaltransduction_simple.jpg

Reception

Signal Transduction - ligand binds to surface receptor

Signal Transduction – ligand binds to surface receptor

 

 

Membrane receptors function by binding the signal molecule (ligand) and causing the production of a second signal (also known as a second messenger) that then causes a cellular response. These types of receptors transmit information from the extracellular environment to the inside of the cell

  • by changing shape or
conformational-rearrangements

conformational-rearrangements

Enzyme_Model  allosterism

Enzyme_Model allosterism

  • by joining with another protein
  • once a specific ligand binds to it.

Examples of membrane receptors include

  • G Protein-Coupled Receptors and
membrane_receptor_g protein

membrane_receptor_g protein

 

 

 

 

  • Receptor Tyrosine Kinases.
activation of receptor Tyrosine Kinase

activation of receptor Tyrosine Kinase

http://www.hartnell.edu/tutorials/biology/images/membrane_receptor_tk.jpg

Intracellular receptors are found inside the cell, either in the cytopolasm or in the nucleus of the target cell (the cell receiving the signal).

Chemical messengers that are hydrophobic or very small (steroid hormones for example) can

  • pass through the plasma membrane without assistance and
  • bind these intracellular receptors.

Once bound and activated by the signal molecule,

  • the activated receptor can initiate a cellular response, such as a
  • change in gene expression.

Note that this is the first time that change in gene expression is stated.  Is the change in gene expression implication of a change in the genetic information – such as – mutation?  That does not have to be the case in the normal homeostatic case.  This might only be

  • a change in the rate of a transcription or a suppression of expression through RNA.
intracellular_receptor_steroid

intracellular_receptor_steroid

http://www.hartnell.edu/tutorials/biology/images/intracellular_receptor_steroid.jpg

Transduction

Since signaling systems need to be

  • responsive to small concentrations of chemical signals and act quickly,
  • cells often use a multi-step pathway that transmits the signal quickly,
  • while amplifying the signal to numerous molecules at each step.
Signal transduction cascades amplify the signal output

Signal transduction cascades amplify the signal output

Steps in the signal transduction pathway often involve

  • the addition or removal of phosphate groups which results in the activation of proteins.
  • Enzymes that transfer phosphate groups from ATP to a protein are called protein kinases.

Many of the relay molecules in a signal transduction pathway are protein kinases and

  • often act on other protein kinases in the pathway. Often
  • this creates a phosphorylation cascade, where
  • one enzyme phosphorylates another, which then phosphorylates another protein, causing a chain reaction.
phosphorylation-cascade

phosphorylation-cascade

Also important to the phosphorylation cascade are

  • a group of proteins known as protein phosphatases.

Protein phosphatases are enzymes that can rapidly remove phosphate groups from proteins (dephosphorylation) and thus inactivate protein kinases. Protein phosphatases are

  • the “off switch” in the signal transduction pathway.

Phosphorylation Dephosphorylation

 

Turning the signal transduction pathway off when the signal is no longer present is important

  • to ensure that the cellular response is regulated appropriately.

Dephosphorylation also makes protein kinases

  • available for reuse and
  • enables the cell to respond again when another signal is received.

Kinases are not the only tools used by cells in signal transduction. Small, nonprotein, water-soluble molecules or ions called second messengers (the ligand that binds the receptor is the first messenger) can also

  • relay signals received by receptors on the cell surface
  • to target molecules in the cytoplasm or the nucleus.
membrane protein receptor binds with hormone

membrane protein receptor binds with hormone

 

insulin receptor and and insulin receptor signaling pathway (IRS)

insulin receptor and and insulin receptor signaling pathway (IRS)

 

 

binding-proteins-and-bioavailable-25-hydroxyvitamin-d

binding-proteins-and-bioavailable-25-hydroxyvitamin-d

 

 

Examples of second messengers include cyclic AMP (cAMP) and calcium ions.

membrane_receptor_g protein

membrane_receptor_g protein

http://www.hartnell.edu/tutorials/biology/images/membrane_receptor_gprotein.jpg

Response

Cell signaling ultimately leads to the regulation of one or more cellular activities. Regulation of gene expression (turning transcription of specific genes on or off) is a common outcome of cell signaling. A signaling pathway may also

  • regulate the activity of a protein, for example
ion-transporters-and-channels-in-mammalian-choroidal-epithelium

ion-transporters-and-channels-in-mammalian-choroidal-epithelium

Ca(2+) and contraction

Ca(2+) and contraction

 

transepithelial-electrogenic-ion-transport

transepithelial-electrogenic-ion-transport

 

calcium release flux

calcium release flux

 

coupled receptors

 

 

 

 

  1. opening or closing an ion channel in the plasma membrane or
  2. promoting a change in cell metabolism such as catalyzing the breakdown of glycogen.

Signaling pathways can also lead to important cellular events such as

  • cell division or apoptosis (programmed cell death).
ubiquitylation-is-a-multistep-reaction.

ubiquitylation-is-a-multistep-reaction.

 

Involvement of VSMCs apoptosis in fibrous plaque rupture.

Involvement of VSMCs apoptosis in fibrous plaque rupture.

 

 

 

 

 

 

 

 

 

 

 

 

G- Protein-Coupled Receptor

 

membrane_receptor_g protein

membrane_receptor_g protein

Arrestin binding to active GPCR kinase (GRK)-phosphorylated GPCRs blocks G protein coupling

Arrestin binding to active GPCR kinase (GRK)-phosphorylated GPCRs blocks G protein coupling

Signal Transduction Tutorial bDr. Katherine Harris is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 3.0 Unported License.

Funded by the U.S. Department of Education, College Cost Reduction and Access (CCRAA) grant award # P031C080096.

http://creativecommons.org/licenses/by-nc-sa/3.0/

  • NonCommercial — You may not use the material for commercial purposes.
  • ShareAlike — If you remix, transform, or build upon the material, you must distribute your contributions under the same license as the original.

Adapt — remix, transform, and build upon the material

hormone + receptor signaling

http://home.earthlink.net/~dayvdanls/SignalTrans.gif

Signal-Transduction-Pathway

http://pi-silico.hkbu.edu.hk/wp-content/uploads/2012/12/Signal-Transduction-Pathway.png

http://upload.wikimedia.org/wikipedia/commons/a/a4/1Signal_Transduction_Pathways_Model.jpg

Akt mTOR pathway

Akt mTOR pathway

http://cc.scu.edu.cn/G2S/eWebEditor/uploadfile/20120810155043970.jpg

Quia – 9AP Chapter 11 – Cell Commun

http://www.quia.com/files/quia/users/lmcgee/membranetransport/cell_communication/reception_transduction_resp.gif

http://cc.scu.edu.cn/G2S/eWebEditor/uploadfile/20120810155043970.jpg

HER2 in Breast Cancer–What Does it Mean?

http://img.medscape.com/fullsize/migrated/editorial/clinupdates/2000/681/tu02.fig2.jpg

Protease signalling: the cutting edge

http://emboj.embopress.org/content/embojnl/31/7/1630/F5.large.jpg

Quia – 9AP Chapter 11 – Cell Commun

http://www.quia.com/files/quia/users/lmcgee/membranetransport/cell_communication/phosphorylation-cascade.gif

 

Signal Transduction in Autism

http://www.mun.ca/biology/desmid/brian/BIOL2060/BIOL2060-14/1403.jpg

The multiple protein-dependent steps in signal transduction

http://www.nature.com/nrm/journal/v1/n2/images/nrm1100_145a_i2.gif

CONVERSING AT THE CELLULAR LEVEL: AN INTRODUCTION TO SIGNAL …

  1. scq.ubc.ca

 

http://www.scq.ubc.ca/wp-content/uploads/2006/07/transduction.gif

 

Biology 1710 > Davis > Flashcards > exam 1 | StudyBlue

  1. studyblue.com

 

http://classconnection.s3.amazonaws.com/602/flashcards/1005602/png/bio101332955375817.png

 

Read Full Post »

Effect of Dietary Magnesium Intake on Insulin Resistance

Reporter: Larry H. Bernstein, MD, FCAP

Dietary Magnesium Intake Improves Insulin Resistance among Non-Diabetic Individuals with Metabolic Syndrome Participating in a Dietary Trial

J Wang1,2,†, G Persuitte3,†, BC Olendzki2, NM Wedick2, …, and Yunsheng Ma 2,*
1 Department of Preventive Medicine, Medical School of Yangzhou University, Yangzhou 225001, China
2 Division of Preventive and Behavioral Medicine, Department of Medicine, University of Massachusetts Medical School, Worcester, MA 01655, USA
3 Division of Biostatistics and Health Services Research, Department of Quantitative Health Science, University of Massachusetts Medical School, Worcester, MA 01655, USA †

Nutrients 27 Sep 2013; 5(10):3910-3919; http://dx.doi.org/10.3390/nu5103910

Many cross-sectional studies show

  1. an inverse association between dietary magnesium and insulin resistance, but
  2. few longitudinal studies examine the ability to meet the Recommended Dietary Allowance (RDA)
  • for magnesium intake through food and
  • its effect on insulin resistance among participants with metabolic syndrome (MetS).

The dietary intervention study examined this question in 234 individuals with MetS. Magnesium intake was assessed using 24-h dietary recalls at baseline, 6, and 12 months.

  1. Fasting glucose and insulin levels were collected at each time point; and
  2. insulin resistance was estimated by the homeostasis model assessment (HOMA-IR).

The relation between magnesium intake and HOMA-IR was assessed using linear mixed models adjusted for covariates.

  • Baseline magnesium intake was 287 ± 93 mg/day (mean ± standard deviation), and
  • HOMA-IR, fasting glucose and fasting insulin were 3.7 ± 3.5, 99 ± 13 mg/dL, and 15 ± 13 μU/mL, respectively.

At baseline, 6-, and 12-months, 23.5%, 30.4%, and 27.7% met the RDA for magnesium. After multivariate adjustment,

    • magnesium intake was inversely associated with metabolic biomarkers of insulin resistance (P < 0.01).

Further, the likelihood of elevated HOMA-IR (>3.6) over time was 71% lower [odds ratio (OR): 0.29; 95% confidence interval (CI): 0.12, 0.72] in participants

  • in the highest quartile of magnesium intake than those in the lowest quartile.

For individuals meeting the RDA for magnesium,

  • the multivariate-adjusted OR for high HOMA-IR over time was 0.37 (95% CI: 0.18, 0.77).

These findings indicate that dietary magnesium intake is inadequate among non-diabetic individuals with MetS and suggest that

    • increasing dietary magnesium to meet the RDA has a protective effect on insulin resistance.

Keywords: magnesium; insulin resistance; metabolic syndrome; epidemiology

Cite This Article

Wang J, Persuitte G, Olendzki BC, Wedick NM, Zhang Z, Merriam PA, Fang H, Carmody J, Olendzki G-F, Ma Y. Dietary Magnesium Intake Improves Insulin Resistance among Non-Diabetic Individuals with Metabolic Syndrome Participating in a Dietary Trial. Nutrients. 2013; 5(10):3910-3919.

Read Full Post »

The Role of Tight Junction Proteins in Water and Electrolyte Transport

 

Reviewer and Curator: Larry H. Bernstein, MD, FCAP 

 

This article is Part II of a series that explores the physiology, genomics, and the proteomics of water and electrolytes in human and mammalian function in health and disease.  In this portion of curation, we examine the role of special proteins at the tight junctions of cells, including the claudins.  Consistent with the exploration of cation homeostasis, the last featured article is one of the altered handling of calcium (Ca2+) in CHF, and the closely regulated calcium efflux by the sodium-calcium exchanger (NCX).

The Role of Aquaporin and Tight Junction Proteins in the Regulation of Water Movement in Larval Zebrafish (Danio rerio).

Kwong RW, Kumai Y, Perry SF.
Department of Biology, University of Ottawa, Ottawa, Ontario, Canada.
PLoS One. 2013 Aug 14;8(8):e70764.
http://dx.doi.org/10.1371/journal.pone.0070764   eCollection 2013.

Teleost fish living in freshwater are challenged by passive water influx; however the molecular mechanisms regulating water influx in fish are not well understood. The potential involvement of aquaporins (AQP) and epithelial tight junction proteins in the regulation of transcellular and paracellular water movement was investigated in larval zebrafish (Danio rerio).

We observed that the half-time for saturation of water influx (K u) was 4.3±0.9 min, and reached equilibrium at approximately 30 min. These findings suggest a high turnover rate of water between the fish and the environment. Water influx was reduced by the putative AQP inhibitor phloretin (100 or 500 μM). Immunohistochemistry and confocal microscopy revealed that AQP1a1 protein was expressed in cells on the yolk sac epithelium. A substantial number of these AQP1a1-positive cells were identified as ionocytes, either H(+)-ATPase-rich cells or Na(+)/K(+)-ATPase-rich cells. AQP1a1 appeared to be expressed predominantly on the basolateral membranes of ionocytes, suggesting its potential involvement in regulating ionocyte volume and/or water flux into the circulation.

Additionally, translational gene knockdown of AQP1a1 protein reduced water influx by approximately 30%, further indicating a role for AQP1a1 in facilitating transcellular water uptake. On the other hand, incubation with the Ca(2+)-chelator EDTA or knockdown of the epithelial tight junction protein claudin-b significantly increased water influx. These findings indicate that the epithelial tight junctions normally act to restrict paracellular water influx. Together, the results of the present study provide direct in vivo evidence that water movement can occur through transcellular routes (via AQP); the paracellular routes may become significant when the paracellular permeability is increased.

PMID:  23967101  PMCID: PMC3743848    http://www.ncbi.nlm.nih.gov/pubmed/23967101

The tight junction protein claudin-b regulates epithelial permeability and sodium handling in larval zebrafish, Danio rerio.

Kwong RW, Perry SF.
Department of Biology, University of Ottawa, Ottawa, Ontario, Canada. wkwong@uottawa.ca
Am J Physiol Regul Integr Comp Physiol. Apr 1, 2013; 304(7):R504-13. http://dx.doi.org/10.1152/ajpregu.00385.2012  Epub 2013 Jan 30.

The functional role of the tight junction protein claudin-b in larval zebrafish (Danio rerio) was investigated. We showed that claudin-b protein is expressed at epithelial cell-cell contacts on the skin. Translational gene knockdown of claudin-b protein expression caused developmental defects, including edema in the pericardial cavity and yolk sac.

Claudin-b morphants exhibited an increase in epithelial permeability to the paracellular marker polyethylene glycol (PEG-4000) and fluorescein isothiocyanate-dextran (FD-4). Accumulation of FD-4 was confined mainly to the yolk sac and pericardial cavity in the claudin-b morphants, suggesting these regions became particularly leaky in the absence of claudin-b expression.

Additionally, Na(+) efflux was substantially increased in the claudin-b morphants, which contributed to a significant reduction in whole-body Na(+) levels. These results indicate that claudin-b normally acts as a paracellular barrier to Na(+). Nevertheless, the elevated loss of Na(+) in the morphants was compensated by an increase in Na(+) uptake.

Notably, we observed that the increased Na(+) uptake in the morphants was attenuated in the presence of the selective Na(+)/Cl(-)-cotransporter (NCC) inhibitor metolazone, or during exposure to Cl(-)-free water. These results suggested that the increased Na(+) uptake in the morphants was, at least in part, mediated by NCC. Furthermore, treatment with an H(+)-ATPase inhibitor bafilomycin A1 was found to reduce Na(+) uptake in the morphants, suggesting that H(+)-ATPase activity was essential to provide a driving force for Na(+) uptake. Overall, the results suggest that claudin-b plays an important role in regulating epithelial permeability and Na(+) handling in zebrafish.
PMID: 23364531   http://www.ncbi.nlm.nih.gov/pubmed/23364531

Evidence for a role of tight junctions in regulating sodium permeability in zebrafish (Danio rerio) acclimated to ion-poor water.

Kwong RW, Kumai Y, Perry SF.
Department of Biology, University of Ottawa, Ottawa, ON, Canada. wkwong@uottawa.ca
J Comp Physiol B. Feb 2013 ;183(2):203-13.
http://dx.doi.org/10.1007/s00360-012-0700-9  Epub 2012 Jul 29.

Freshwater teleosts are challenged by diffusive ion loss across permeable epithelia including gills and skin. Although the mechanisms regulating ion loss are poorly understood, a significant component is thought to involve paracellular efflux through pathways formed via tight junction proteins. The mammalian orthologue (claudin-4) of zebrafish (Danio rerio) tight junction protein, claudin-b, has been proposed to form a cation-selective barrier regulating the paracellular loss of Na(+).

The present study investigated the cellular localization and regulation of claudin-b, as well as its potential contribution to Na(+) homeostasis in adult zebrafish acclimated to ion-poor water. Using a green fluorescent protein-expressing line of transgenic zebrafish, we found that claudin-b was expressed along the lamellar epithelium as well as on the filament in the inter-lamellar regions. Co-localization of claudin-b and Na(+)/K(+)-ATPase was observed, suggesting its interaction with mitochondrion-rich cells. Claudin-b also appeared to be associated with other cell types, including the pavement cells. In the kidney, claudin-b was expressed predominantly in the collecting tubules. In addition,

exposure to ion-poor water caused a significant increase in claudin-b abundance as well as a decrease in Na(+) efflux, suggesting a possible role for claudin-b in regulating paracellular Na(+) loss. Interestingly, the whole-body uptake of a paracellular permeability marker, polyethylene glycol-400, increased significantly after prolonged exposure to ion-poor water, indicating that an increase in epithelial permeability is not necessarily coupled with an increase in passive Na(+) loss. Overall, our study suggests that in ion-poor conditions, claudin-b may contribute to a selective reduction in passive Na(+) loss in zebrafish.
PMID: 22843140   http://www.ncbi.nlm.nih.gov/pubmed/22843140

Claudin-16 and claudin-19 function in the thick ascending limb.

Hou J, Goodenough DA.
Washington University School of Medicine, Div Renal Diseases, St Louis, Missouri
Curr Opin Nephrol Hypertens. Sep 2010; 19(5):483-8. http://dx.doi.org/10.1097/MNH.0b013e32833b7125.

The thick ascending limb (TAL) of the loop of Henle is responsible for reabsorbing 25–40% of filtered Na+, 50–60% of filtered Mg2+ and 30–35% of filtered Ca2+. The dissociation of salt and water reabsorption in the TAL serves both to dilute the urine and to establish the corticomedullary osmolality gradient. Active transcellular salt reabsorption results in a lumen-positive transepithelial voltage that drives passive paracellular reabsorption of divalent cations. Claudins are the key components of the paracellular channel. The paracellular channels in the tight junction have properties of ion selectivity, pH dependence and anomalous mole fraction effects, similar to conventional transmembrane channels. Genetic mutations in claudin-16 and claudin-19 cause an inherited human renal disorder, familial hypomagnesemia with hypercalciuria and nephrocalcinosis (FHHNC).

In the TAL of Henle’s loop, the epithelial cells form a water-impermeable barrier, actively transport Na+ and Cl− via the transcellular route, and provide a paracellular pathway for the selective absorption of cations. Na+ K+ and Cl− enter the cell through the Na-K-2Cl cotransporter (NKCC2) in the luminal membrane. Na+ exits the cell through the Na+/K+-ATPase, in exchange for K+ entry. K+ is secreted into the lumen through the renal outer medullary potassium channel. Cl− leaves the cell through the basolateral Cl− channel, made up of two subunits, ClCKb and barttin. The polarized distribution of luminal K+ versus basolateral Cl− conductance generates a spontaneous voltage source (Vsp) of +7−8mV , depending on active transcellular NaCl reabsorption. With continuous NaCl reabsorption along the axis of the TAL segment, the luminal fluid is diluted to 30–60mmol/l  and a large NaCl transepithelial chemical gradient develops at the end of the TAL. Because the paracellular permeability of the TAL is cation-selective (with a PNa/PCl value between 2 and 4), the diffusion voltage (Vdi) is superimposed onto the active transport voltage (Vsp) and becomes the major source of the transepithelial voltage (Vte), which now increases up to +30mV.

Early in-vivo micropuncture studies have shown that approximately 50–60% of the filtered Mg2+ is reabsorbed in the TAL. The flux–voltage relationship indicates that Mg2+ is passively reabsorbed from the lumen to the peritubular space through the paracellular pathway in this segment, driven by a lumen positive Vte.  Vte is made of the sum of Vsp and Vdi. There are two prerequisites required for the paracellular Mg2+ reabsorption in the TAL: the lumen-positive Vte as the driving force and the paracellular permeability for the divalent cation Mg2+.

Claudin-16 and claudin-19 underlie familial hypercalciuric hypomagnesemia with nephrocalcinosis

Claudin-16 and claudin-19 play a major role in the regulation of magnesium reabsorption in the thick ascending limb (TAL). This review describes recent findings of the physiological function of claudin-16 and claudin-19 underlying normal transport function for magnesium reabsorption in the TAL. Mutations in the genes encoding the tight junction proteins claudin-16 and claudin-19 cause the inherited human renal disorder familial hypomagnesemia with hypercalciuria and nephrocalcinosis. FHHNC, OMIM #248250, is a rare autosomal recessive tubular disorder. As a consequence of excessive renal Mg2+ and Ca2+ wasting, patients develop the characteristic triad of hypomagnesemia, hypercalciuria and nephrocalcinosis. Recurrent urinary tract infections and polyuria/polydipsia are frequent initial symptoms. Other clinical symptoms include nephrolithiasis, abdominal pain, convulsions, muscular tetany, and failure to thrive. Additional laboratory findings include elevated serum parathyroid hormone levels before the onset of chronic renal failure, incomplete distal tubular acidosis, hypocitraturia, and hyperuricemia. In contrast to hypomagnesemia and secondary hypocalcemia (HSH, OMIM #602014), FHHNC is generally complicated by end-stage renal failure in early childhood or adolescence.

Simon et al. used the positional cloning strategy and identified claudin-16 (formerly known as paracellin-1), which is mutated in patients with FHHNC. Most mutations reported to date in claudin-16 are missense mutations clustering in the first extracellular loop composing the putative ion selectivity filter. Konrad et al. have found mutations in another tight junction gene encoding claudin-19 from new cohorts of FHHNC patients (OMIM #248190). The renal tubular phenotypes are indistinguishable between patients with mutations in claudin-16 and those with mutations in claudin-19. Although claudin-16 and claudin-19 underlie FHHNC and paracellular Mg2+ reabsorption in the TAL, the transient receptor potential channel melastatin 6 (TRPM6) regulates the apical entry of Mg2+ into the distal convoluted tubule epithelia. Mutations in TRPM6 cause the HSH syndrome.

These above data suggested the hypothesis that claudin-16 and/or claudin-19 forms a selective paracellular Mg2+/Ca2+ channel, which was tested in a number of in-vitro studies. Ikari et al. transfected low-resistance Madin-Darby canine kidney (MDCK) cells with claudin-16 and reported that the Ca2+ flux in these cells was increased in the apical to basolateral direction, whereas the Ca2+ flux in the opposite direction remained unchanged. The Mg2+ flux was without any noticeable change. Kausalya et al.  transfected the high-resistance MDCK-C7 cell line and found that claudin-16 only moderately increased Mg2+ permeability without any directional preference. The effects of claudin-16 on Mg2+/Ca2+ permeation appeared so small (<50%) that the Mg2+/Ca2+ channel theory incompletely explains the dramatic effect of mutations in claudin-16 on Mg2+ and Ca2+ homoeostasis in FHHNC patients. However, , Hou et al.  transfected the anion-selective LLC-PK1 cell line with claudin-16 and found a large increase in Na+ permeability (PNa) accompanied by a moderately enhanced Mg2+ permeability (PMg). The permeability of claudin-16 to other alkali metal cations was found to be: K+ > Rb+ > Na+.  Yu et al. emphasized that these residue replacements can influence protein structures that may have impacts on ion permeability independent of amino acid charge.

The cation selectivity of the tight junction is vital for generating the lumen positive transepithelial potential in the TAL, which drives paracellular absorption of magnesium. Claudin-16 and claudin-19 require each other for assembly into tight junctions in the TAL. Heteromeric claudin-16 and claudin-19 interaction forms a cation selective tight junction paracellular channel. Loss of either claudin-16 or claudin-19 in the mouse kidney abolishes the cation selectivity for the TAL paracellular pathway, leading to excessive renal wasting of magnesium.

Claudins interact with each other both intracellularly and intercellularly: they copolymerize linearly within the plasma membrane of the cell, together with the integral protein occludin, to form the classical intramembrane fibrils or strands visible in freeze-fracture replicas. These intramembrane interactions (side-to-side) can involve one claudin protein (homomeric or homopolymeric) or different claudins (heteromeric or heteropolymeric). In the formation of the intercellular junction, claudins may interact head-to-head with claudins in an adjacent cell, generating both homotypic and heterotypic claudin–claudin interactions. Using the split-ubiquitin yeast 2-hybrid assay, Hou et al. found strong claudin-16 and claudin-19 heteromeric interaction. The point mutations in claudin-16 (L145P, L151F, G191R, A209T, and F232C) or claudin-19 (L90P and G123R) that are known to cause human FHHNC disrupted the claudin-16 and claudin-19 heteromeric interaction. In mammalian cells such as the human embryonic kidney 293 cells, claudin-16 can be coimmunoprecipitated with claudin-19. Freeze-fracture replicas revealed the assembly of tight junction strands in L cells coexpressing claudin-16 and claudin-19, supporting their heteromeric interaction.

Coexpression of claudin-16 and claudin-19 in LLC-PK1 cells resulted in a dramatic upregulation of PNa and down-regulation of PCl, generating a highly cation-selective paracellular pathway. Certain FHHNC mutations in claudin-16 (L145P, L151F, G191R, A209T, and F232C) or claudin-19 (L90P and G123R) that disrupted their heteromeric interaction abolished this physiological change. As claudin-16 colocalizes with claudin-19 in the TAL epithelia of the kidney, claudin-16 and claudin-19 association through heteromeric interactions confers cation selectivity to the tight junction in the TAL. Human FHHNC mutations in claudin-16 or claudin-19 that abolish the cation selectivity diminish the lumen-positive Vdi as the driving force for Mg2+ and Ca2+ reabsorption, readily explaining the devastating phenotypes in FHHNC patients.  Hou et al. generated claudin-16 deficient mouse models using lentiviral transgenesis of siRNA to knock down claudin-16 expression by more than 99% in mouse kidneys. Claudin-16 knockdown mice show significantly reduced plasma Mg2+ levels and excessive urinary excretions (approximately four-fold) of Mg2+ and Ca2+. Calcium deposits are observed in the basement membranes of the medullary tubules and the interstitium in the kidney of claudin-16 knockdown mice. These phenotypes of claudin-16 knockdown mice recapitulate the symptoms in human FHHNC patients.

The paracellular reabsorption of Mg2+ and Ca2+ is driven by a lumen-positive Vte made up of two components: Vsp and Vdi. When isolated TAL segments were perfused ex vivo with symmetrical NaCl solutions, there was no difference in Vsp between claudin-16 knockdown and wild-type mice, indicating Vsp was normal in claudin-16 knockdown. Blocking the NKCC2 channel with furosemide (thus dissipating VSP), the cation selectivity (PNa/PCl) was significantly reduced from3.1 ± 0.3 in wild type to 1.5 ± 0.1 in claudin-16 knockdown, resulting in the loss of Vdi. When perfused with a NaCl gradient of 145mmol/l (bath) versus 30mmol/l (lumen), the resulting Vdi was +18mV in wild type, but only +6.6mV in claudin-16 knockdown. Thus, the reduction in Vdi accounted for a substantive loss of the driving force for Mg2+ and Ca2+ reabsorption.

Renal handling of Na+ in claudin-16 knockdown mice is more complex. In the early TAL segment, the transcellular and paracellular pathways form a current loop in which the currents traversing the two pathways are of equal size but opposite direction. Net luminal K+ secretion and basolateral Cl− absorption polarize the TAL epithelium and generate Vsp. As the paracellular pathway is cation selective (PNa/PCl=2–4 , the majority of the current driven by Vsp through the paracellular pathway is carried by Na+ moving from the lumen to the interstitium. Hebert et al. estimated that, for each Na+ absorbed through the trans-cellular pathway, one Na+ is absorbed through the paracellular pathway. With the loss of claudin-16 and the concomitant loss of paracellular cation selectivity, Na+ absorption through the paracellular pathway is reduced.  In the late TAL segment, dilution of NaCl in the luminal space creates an increasing chemical transepithelial gradient; back diffusion of Na+ through the cation-selective tight junction generates a lumen-positive Vdi across the epithelium. The paracellular absorption of Na+ will be diminished when Vdi equals Vsp, and reversed when Vdi exceeds Vsp. As an equilibrium potential, Vdi blocks further Na+ backleak into the lumen. Without claudin-16, Vdi will be markedly reduced well below normal, providing a driving force for substantial Na+ secretion. Indeed, claudin-16 knockdown mice had increased fractional excretion of Na+ (FENa) and developed hypotension and secondary hyperaldosteronism. The observed Na+ and volume loss are consistent with human FHHNC phenotypes. For example, polyuria and polydipsia are the most frequently reported symptoms from FHHNC patients.

Epithelial paracellular channels are increasingly understood to be formed from claudin oligomeric complexes. In the mouse TAL, claudin-16 and claudin-19 cooperate to form cation-selective paracellular channels required for normal levels of magnesium reabsorption. Different subsets of the claudin family of tight junction proteins are found distributed throughout the nephron, and understanding their roles in paracellular ion transport will be fundamental to understanding renal ion homeostasis.

Keywords: claudin; hypomagnesemia; thick ascending limb; tight junction; transepithelial voltage.     PMID: 20616717  PMCID: PMC3378375  http://www.ncbi.nlm.nih.gov/pubmed/20616717

Function and regulation of claudins in the thick ascending limb of Henle.

Günzel D, Yu AS.
Depart Clin Physiol, Charité, Campus Benjamin Franklin, Berlin, Germany.
Pflugers Arch. May 2009; 458(1):77-88.
http://dx.doi.org/10.1007/s00424-008-0589-z  Epub 2008 Sep 16.

The thick ascending limb (TAL) of Henle mediates transcellular reabsorption of NaCl while generating a lumen-positive voltage that drives passive paracellular reabsorption of divalent cations. Disturbance of paracellular reabsorption leads to Ca(2+) and Mg(2+) wasting in patients with the rare inherited disorder of familial hypercalciuric hypomagnesemia with nephrocalcinosis (FHHNC). Recent work has shown that the claudin family of tight junction proteins form paracellular pores and determine the ion selectivity of paracellular permeability. Importantly, FHHNC has been found to be caused by mutations in two of these genes, claudin-16 and claudin-19, and mice with knockdown of claudin-16 reproduce many of the features of FHHNC. Here, we review the physiology of TAL ion transport, present the current view of the role and mechanism of claudins in determining paracellular permeability, and discuss the possible pathogenic mechanisms responsible for FHHNC.

Tight junctions form the paracellular barrier in epithelia. Claudins are ~22 kDa proteins that were first identified by Mikio Furuse in the laboratory of the late Shoichiro Tsukita as proteins that copurified in a tight junction fraction from the chicken liver [23]. The observation that they were transmembrane proteins with 4 predicted membrane domains and 2 extracellular domains raised early on the possibility that they could play a key role in intercellular adhesion and formation of the paracellular barrier. In 1999, Richard Lifton’s group identified mutations in a novel gene, which they called paracellin, as the cause of familial hypercalciuric hypomagnesemia, an inherited disorder believed to be due to failure of paracellular reabsorption of divalent cations in the thick ascending limb of the renal tubule. Paracellin turned out to be a claudin family member (claudin-16). This suggested that claudins in general might be directly involved in regulating paracellular transport in all epithelia. This is now supported by numerous studies demonstrating that overexpressing or ablating expression of various claudin isoforms in cultured cell lines or in mice affects both the degree of paracellular permeability and its selectivity (vide infra). Furthermore, in mammals alone there are ~24 claudin genes and each exhibits a distinct tissue-specific, pattern of expression. Thus, the specific claudin isoform(s) expressed in each tissue might explain its paracellular permeability properties.

Each nephron segment expresses a unique set of multiple claudin isoforms, and each isoform is expressed in multiple segments, thus making a complicated picture which even varies between different species. The role of combinations of claudins in determining paracellular permeability properties has hardly been studied yet. In mouse, rabbit and cattle, the thick ascending limb of Henle’s loop is thought to express claudins 3, 10, 11, 16  in adulthood, and, at least in mouse, additionally claudin-6 during development. In addition, claudin-4 has been found in cattle and claudin-8 in rabbit. To date, the distribution of Claudin-19 has been investigated in mouse, rat, and man where its presence in the TAL was demonstrated.

The thick ascending limb (TAL) of Henle’s loop, working as “diluting segment” of the nephron, is characterized by two major properties: high transepithelial, resorptive transport of electrolytes and low permeability to water. Major players to achieve electrolyte transport are the apical Na+-K+-2Cl−symporter (NKCC2), the apical K+ channel ROMK, the basolateral Cl− channel (CLC-Kb) together with its subunit barttin and the basolateral Na+/K+-ATPase . The combined actions of these transport systems have been extensively reviewed and are therefore only briefly summarized here. Na+ and Cl− are resorbed by entering the cells apically through NKCC2 and leaving the cells basolaterally through the Na+/K+-ATPase and CLC-Kb, respectively. In contrast, K+ is either recycled across the apical membrane as it is entering through NKCC2 and leaving through ROMK, or even secreted, as it is also entering the cells basolaterally through the Na+/K+-ATPase. Taking these ion movements together, there is a net movement of positive charge from the basolateral to the apical side of the epithelium, giving rise to a lumen positive voltage (3 – 9 mV [11]; about 5 – 7 mV [30,31]; 7 – 8 mV [57]). Over the length of the TAL, luminal NaCl concentration decreases gradually to concentrations of 30 – 60 mM at the transition to the distal tubule, depending on the flow rate within the tubule (low flow rates resulting in low concentrations).

To keep up such a high gradient, the TAL epithelium has to be tight to water and various studies summarized by Burg and Good report water permeability values from 28 µm/s down to values indistinguishable from zero. Tight junctions of the TAL are, however, highly permeable to cations with PNa being about 2 – 2.7 fold, 2.5 fold or even up to 6 fold that of PCl. Amongst the monovalent cations, a permeability sequence of PK > PNa > PRb = PLi > PCs > Porganic cation was observed which is similar to Eisenman sequence VIII or IX, indicating a strong interaction between the permeating ion and the paracellular pore that enables at least partial removal of the hydration shell (see below). As reviewed by Burg and Good, the transepithelial sodium and chloride permeabilities, estimated from radioisotope fluxes, are high (in the range of 10 – 63·10−6 cm/s) and the transepithelial electrical resistance is correspondingly low (21 – 25 W cm2 ; 30 – 40 W cm2 ; 11 – 34 W cm2 . Blocking active transport by the application of furosemide or ouabain increases transepithelial resistance only slightly, indicating that the low values are primarily due to a very high paracellular permeability. Due to these properties of TAL epithelial cells, Na+ ions leak back into the lumen of the tubule, creating a diffusion (dilution) potential that adds another 10 – 15 mV to the lumen positive potential, so that considerable potential differences (25 mV ; 30 mV; cTAL 23 mV, mTAL 17 mV ) may be reached at very slow flow rates.

Considerable proportions of the initially filtrated Mg2+ (50 – 60%; 50 – 70%; 65 – 75%) and Ca2+ (20%; 30 – 35% are resorbed in the TAL. The transepithelial potential is considered to provide the driving force for the predominantly paracellular resorption of Mg2+ and Ca2+ as in many studies, transport of both divalent ions in the TAL has been found to be strictly voltage dependent (resorbtive at lumen positive potentials, zero at 0 mV and secretory at lumen negative potentials) and permeability considerable (PCa 7.7·10−6 cm/s, i.e. approximately 25% of PNa). There is however, some conflicting evidence, e.g. by Suki et al. and Friedman. Both studies used cTAL (cortical TAL) and found that decreasing the transepithelial potential by applying furosemide did either not alter the unidirectional lumen to bath Ca2+ flux (rabbit) or left a substantial net Ca2+ resorption (mouse). Similarly, Rocha et al. found that bath application of ouabain almost abolished the transepithelial potential, but hardly affected net Ca2+ resorption and conclude that (a) all segments of Henle’s loop are relatively impermeable to calcium and (b) net calcium resorption occurs in the thick ascending limb which cannot be explained by passive mechanisms.  Mandon et al. conclude that both Mg2+ and Ca2+ are transported in the cTAL but not in the mTAL (medullary TAL) of rat and mouse, although transepithelial potential differences were similar in both segments, and even if the transepithelial potential was experimentally elevated to values above 20 mV. Wittner et al. even found evidence that in mouse mTAL the passive permeability to divalent cations is very low and that Ca2+ and Mg2+ can be secreted into the luminal fluid under conditions which elicit large lumen-positive transepithelial potential differences. They conclude that this Ca2+ and Mg2+ transport is most probably of cellular origin. In contrast, in rabbit, both ions are transported along the whole length of the TAL.

Both, Mg2+ and Ca2+ resorption are modulated through the action of the basolateral Ca (and Mg) sensing receptor (CaSR) which is found along the entire nephron but especially in the loop of Henle, distal convoluted tubule (DCT) and the inner medullary collecting duct. Different modes of action on Ca2+ and Mg2+ homeostasis exist, such as an indirect action through the modulation of PTH secretion or direct effects on the cells expressing CaSR. In the TAL the latter model is based on the assumptions depicted above, i.e. that Mg2+ and Ca2+ are resorbed paracellularly, driven by the lumen positive potential, so that a reduction in NaCl resorption causes a reduction in driving force for Mg2+ and Ca2+ resorption. As reviewed by Hebert and Ward, CaSR is activated through an increase in basolateral Ca2+ and/or Mg2+ concentration which triggers an increase in the intracellular Ca2+ concentration. This reduces the activity of the adenylate cyclase which, in turn, inhibits transcellular transport of Na+ and Cl−. In addition, the increase in intracellular Ca2+ activates phospholipase A2 (PLA2) and thus increases the intracellular concentration of arachidonic acid and its derivative, 20-HETE. 20-HETE inhibits NKCC2, ROMK and the Na+/K+-ATPase and by this Mg2+ and Ca2+ resorption. In keeping with this hypothesis, mutations in CaSR affect Ca/Mg resorption. Inactivating mutations cause hypercalcemia, hypocalciuria, hypomagnesiuria and, in some patients hypermagnesemia. Conversely, activating mutations (gain of function mutations) lead to hypocalcemia, hypercalciuria, hypermagnesiuria and in up to 50% of the patients mild hypomagnesemia.

Bartter syndrome type I (mutations in NKCC2), and type II (mutations in ROMK) lead to hypercalciuria and thus cause nephrocalcinosis, but no hypomagnesemia is observed. Reports on hypermagnesiuria are conflicting: while Kleta and Bockenhauer link it to nephrocalcinosis seen in these patients, Rodriguez-Soriano states that patients with neonatal Bartter syndrome (i.e. type I or II) show a lack of hypermagnesiuria that may be explained by compensation in the DCT. Hypomagnesemia is occationally present in Bartter type III (CLC-Kb). However, here it is believed to be mainly due to effects on DCT, where CLC-Kb shows highest expression. Patients with Bartter syndrome IV (CLC-Kbsubunit barttin) may or may not present nephrocalcinosis, while Mg2+ homeostasis appears undisturbed. Interestingly, the largest effects on Mg2+-homeostasis are observed in Gitelman syndrome, a defect in the Na+/Cl− symport (NCC) predominantly found in the DCT, where Mg2+ is transported along the transcellular route. Affected patients suffer from hypomagnesemia, hypermagnesuria and hypocalciuria. The effect on Mg2+ in Gitelman syndrome is still poorly understood and possibly due to a concomitant down-regulation of TRPM6, the apical Mg2+ uptake channel in DCT.

More than 30 different claudin-16 mutations have now been reported in families with FHHNC. Because of the large number of unique mutations, it has not been possible to identify any clear qualitative correlation between the phenotype and individual mutations, although certain mutations are associated with greater severity of disease. In 2006, a second locus was identified, CLDN19, which encodes claudin-19. In the initial report, the phenotype appeared similar to that due to claudin-16 mutations, with the exception that there was a high prevalence of ocular abnormalities, including macular colobomata, nystagmus and myopia. Claudin-19 is normally expressed at high levels in the retina, but why it causes these ocular disorders is unknown.

In vitro studies of claudin function comprise inducible or non-inducible transfection of various cells lines with cDNA for claudins that are not endogenously expressed by the cell line used. Alternatively, cells can be transfected with siRNA directed against an endogenous claudin. In both cases, cells are then grown to confluence on permeable filter supports that allow measurement of transepithelial permeabilities. Before the results of permeability studies can be interpreted, however, several parameters have to be controlled.

First, special care has to be taken to make sure that the exogenous claudin is correctly inserted into the tight junction. This can be achieved e.g. by confocal laser scanning microscopy, colocalizing the claudin of interest  with a tight junction marker protein such as occludin.

Second, it has to be ensured that endogenous claudin expression remains unaffected, as permeability changes always result from the combined effects of alterations in endogenous and exogenous claudins.

Third, it has to be kept in mind that, typically, epithelia express several different claudins that act together to produce tissue specific permeability properties. Thus, ideally, a cell line should be chosen that provides a claudin background resembling that usually experienced by the claudin investigated. The latter two points may be the reason for contradicting results obtained in permeability studies expressing a specific claudin in different cell lines.

Studies of paracellular permeabilities can be divided into two groups, those employing electrophysiological measurements (e.g. determination of diffusion potentials), and those measuring ion or solute flux, using either radioactive isotopes or various analytical methods to determine the amount transported.

Although transepithelial conductances depend on paracellular permeabilities of the predominant ions in the bath solution, conductance changes alone cannot be used to predict ion permeabilities.

Firstly, conductances always depend on both ion and counter-ion, not on one ion species alone.

Secondly, transepithelial conductances are the sum of the conductances of the transcellular and paracellular pathways.

Thus, they only reflect paracellular permeability, if paracellular conductance dominates transepithelial conductance and if transcellular conductance remains constant throughout the experiment. This, however, is often not the case, as concentration changes of the ions investigated may affect transcellular conductance, e.g. by activating ion transporters or by inhibiting ion channels. Thirdly, the specific conductance of each solution employed may differ and has therefore to be assessed and taken into account. Thus, comparison of results from diffusion potential measurements or flux studies and conductance measurements may even yield contradicting results. For the same reasons, other methods based on pure conductance/resistance measurements, including the more sophisticated conductance scanning method or one-path impedance spectroscopy are not ion specific and do not allow the measurement of paracellular permeabilities to single ions.

In contrast to electrophysiological measurements, flux measurements are not limited to ions but can also be extended to uncharged molecules.  Flux measurements per se do not distinguish between transcellular or paracellular transport. Therefore, to estimate paracellular permeabilities, inhibition or at least estimation of the transcellular flux is necessary. Assuming that transcellular flux for energetic reasons is not easily reversible while paracellular flux is passive and thus generally assumed to be symmetric, the transcellular proportion is often estimated by calculating the difference between apical to basolateral and basolateral to apical fluxes. All flux measurements are very sensitive to the development of diffusion zones (“unstirred layers”) near the cells. These layers are depleted/enriched in the compound transported and thus alter the driving forces acting on these compounds, if bath solutions are not continually circulated. If ionic fluxes are investigated, transepithelial potentials may develop that diminish or completely inhibit the flux investigated.

All the techniques described above have been employed to investigate the function of claudin-16 and -19, especially with respect to their ability to increase paracellular permeability to divalent cations. The hypothesis, that claudin-16 (then called paracellin-1) may be a paracellular Mg2+ and Ca2+ pore was originally expressed by Simon et al. It was based on the findings that mutations in claudin-16 were the cause of the severe disturbance in Mg2+ and Ca2+ homeostasis in FHHNC patients together with the observations that claudin-16 is a tight junction protein located in the TAL, i.e. the nephron segment responsible for bulk Mg2+ resoption along the paracellular pathway. When, recently, it was found that claudin-19 mutations were underlying hitherto unexplained cases of FHHNC and that claudin-19 co-localized with claudin-16, the hypothesis was extended to claudin-19.

  1. Cole DEC, Quamme GA. Inherited disorders of renal magnesium handling. J Am Soc Nephrol 2000;11:1937–1947. [PubMed: 11004227]
  2. de Rouffignac C, Quamme G. Renal magnesium handling and its hormonal control. Physiol Rev 1994;74:305–322. [PubMed: 8171116]  
  3. Meij IC, van den Heuvel LP, Knoers NV. Genetic disorders of magnesium homeostasis. BioMetals 2002;15:297–307. [PubMed: 12206395]
  4. Satoh J, Romero MF. Mg2+ transport in the kidney. BioMetals 2002;15:285–295. [PubMed: 12206394]  
  5. Schlingmann KP, Konrad M, Seyberth HW. Genetics of hereditary disorders of magnesium homeostasis. Pediatr Nephrol 2004;19:13–25. [PubMed: 14634861]  
  6. Simon DB, Lu Y, Choate KA, et al. Paracellin-1, a renal tight junction protein required for paracellular Mg2+ resorption. Science 1999;285:103–106. [PubMed: 10390358]

PMID: 18795318  PMCID:  PMC2666100  http://www.ncbi.nlm.nih.gov/pubmed/18795318

Deletion of claudin-10 (Cldn10) in the thick ascending limb impairs paracellular sodium permeability and leads to hypermagnesemia and nephrocalcinosis.

Breiderhoff T, Himmerkus N, Stuiver M, Mutig K, Will C, Meij IC et al.
Max Delbrück Center for Molec Med, Berlin, Germany. t.breiderhoff@mdc-berlin.de

Erratum in Proc Natl Acad Sci. 2012 Sep 11;109(37):15072.
Proc Natl Acad Sci. Aug 28, 2012; 109(35):14241-6.   http://dx.doi.org/10.1073/pnas.1203834109. Epub 2012 Aug 13.

In the kidney, tight junction proteins contribute to segment specific selectivity and permeability of paracellular ion transport. In the thick ascending limb (TAL) of Henle’s loop, chloride is reabsorbed transcellularly, whereas sodium reabsorption takes transcellular and paracellular routes. TAL salt transport maintains the concentrating ability of the kidney and generates a transepithelial voltage that drives the reabsorption of calcium and magnesium. Thus, functionality of TAL ion transport depends strongly on the properties of the paracellular pathway. To elucidate the role of the tight junction protein claudin-10 in TAL function, we generated mice with a deletion of Cldn10 in this segment. We show that claudin-10 determines paracellular sodium permeability, and that its loss leads to hypermagnesemia and nephrocalcinosis. In isolated perfused TAL tubules of claudin-10-deficient mice, paracellular permeability of sodium is decreased, and the relative permeability of calcium and magnesium is increased. Moreover, furosemide-inhibitable transepithelial voltage is increased, leading to a shift from paracellular sodium transport to paracellular hyperabsorption of calcium and magnesium. These data identify claudin-10 as a key factor in control of cation selectivity and transport in the TAL, and deficiency in this pathway as a cause of nephrocalcinosis.

Whereas regulation of transporters and channels involved in trans-cellular ion transport has been characterized in much detail, the functional and molecular determinants of paracellular ion trans­port in the kidney remain incompletely understood. In the thick ascending limb (TAL) of Henle’s loop, both trans-cellular and paracellular ion transport pathways contribute to reabsorption of Na+, Cl, Mg2+, and Ca2+. Na+ and Clare reabsorbed mostly transcellularly by the concerted action of chan­nels and transporters. Mutations in five of the genes involved lead to Bartter syndrome, a disorder characterized by salt wasting and polyuria. Whereas Clis transported exclusively transcellularly, 50% of the Na+ load, as well as Ca2+ and Mg2+, are reabsorbed via paracellular pathways. In the TAL, this paracellular route is highly cation-selective. The paracellular passage is largely controlled by the tight junction (TJ), a supramolecular structure of membrane-spanning proteins, their intracellular adapters, and scaffolding proteins. Claudins, a family comprising 27 members, are the main components of the TJ defining the permeability properties. They interact via their extracellular loops with corre­sponding claudins of the neighboring cell to allow or restrict pas­sage of specific solutes (5, 6). In the kidney, their expression pattern is closely related to the corresponding segment-specific solute reabsorption profile. Several claudins are expressed in the TAL, including claudin-16, -19, -10, -3, and -18 The importance of claudin-16 and -19 in this tissue is documented by mutations in CLDN16 and CLDN19, which cause familial hypomagnesemia, hypercalciuria, and nephrocalcinosis, an autosomal recessive dis­order that leads to end-stage renal disease. The relevance of CLDN16 for paracellular reabsorption of Mg2+ and Ca2+ was confirmed in mouse models with targeted gene disruption. In addition, claudin-14, expressed in the TAL of mice on a high-calcium diet, was identified as negative regulator of claudin-16 function (15), and sequence variants in CLDN14 have been asso­ciated with human kidney stone disease. The functional significance of claudin-10, which is also ex­pressed in the TAL, remains unclear. This TJ protein is expressed in two isoforms, claudin-10a and claudin-10b, which differ in their first extracellular loop. In cultured epithelial cells, heter-ologous expression of claudin-10a increases paracellular anion transport, whereas claudin-10b expression increases paracellular cation transport. Both isoforms are expressed differentially along the nephron, with claudin-10a found predominantly in cortical segments, whereas claudin-10b is enriched in the medullary region.  In the present study we generated a mouse model with a TAL-specific Cldn10 gene defect to query the role of this protein in renal paracellular in transport in vivo. We found that claudin-10 is crucial to paracellular Na+ handling in the TAL, and that its absence leads to a shift from paracellular sodium transport to paracellular hyperreabsorption of Ca2+ and Mg2+.

Analysis of claudin-10 expression in the kidney. (A) Western blot analysis of kidney membrane extracts from control (ctr) and cKO mice. A dramatic reduction in claudin-10 protein can be seen in kidneys of cKO mice. Levels of the TJ marker occludin are unchanged. (B) Gene expression analysis of Cldn10 variants on cDNA from isolated segments of the nephron. (C) Immunohistological detection of claudin-10 and markers for PCT (NHE3) and TAL (NKCC2) on sections from control mice (ctr) and cKO mice demonstrates no difference in the signal for claudin-10 in the PCT between WT and cKO. Claudin-10 is expressed in TAL tubules positive for NKCC2. No specific clau-din-10 staining is evident in the TAL of cKO mice. Claudin-10 is detected in TJs positive for ZO-1. This signal is absent in cKO mice, whereas ZO-1 staining is unchanged. (Scale bar: 25 μm.)

In control animals, claudin-10 is located mainly in the TAL, as documented by coimmunostaining with the Na+K+2Clcotrans-porter (NKCC2). In this segment, a large portion of the claudin-10 immunofluorescence signal is located outside of the TJ; however, claudin-10 is present in the TJ, as demonstrated by colocalization with the TJ protein ZO-1. PCTs positive for the sodium-proton exchanger NHE-3 showed a considerably weaker signal restricted to the TJ area. Claudin-10 immunoreactivity was virtually absent in NKCC2-positive tubules of cKO mice, in line with the activity of Cre recombinase in this cell type. The immu-noreactivity of claudin-10 in PCTs of cKOs remained unchanged, however. ZO-1 staining in TAL sections of cKOs was unchanged compared with controls, indicating no unspecific effect on TJ structures. The TJ localization of claudin-16 and claudin-19 in medullary rays was similar in cKOs and controls.   To investigate the phenotypic consequences of renal claudin-10 deficiency, we per­formed a histological examination of the kidneys of 10-wk-old cKO mice and their respective controls. Kidneys from cKO mice contained extensive medullary calcium deposits, as revealed by von Kossa and alizarin red S staining. The deposits were found along the outer stripe of the outer medulla. The detection of extensive calcification suggests alterations in renal ion homeostasis in mice deficient for claudin-10.  Serum Na+ and Cllevels and their renal FE excretion rates were not different be­tween genotypes. In addition, serum creatinine and glomerular filtration rate were not altered compared with controls. Taken together, these findings indicate that calcium deposition does not nonpecifically affect overall glomerular or tubular function.

Fig 4. Gene expression analysis of renal claudins (A) and representative renal ion transporters and channels (B) by real-time PCR. Cldn10 deficiency results in differential gene expression of several genes. Values from cKO animals are shown relative to control mice (mean ± SEM). Wnk1, Wnk1-KS, Kcnj1, and Trpm6, n = 5/4; all other genes, n = 10/10. *P < 0.05; **P < 0.01; ***P < 0.001.    The thiazide-sensitive NaCl cotransporter NCC (Slc12a3), the protein involved in NaCl absorption in the DCT, and the respective inhibitory, kidney-specific kinase-defective KS-WNK1 were expressed at lower levels in the cKO mice. Taken together, these data suggest specific compensatory alterations in components of both paracellular and transcellular renal ion transport mechanisms in mice deficient in claudin-10 in the TAL.

Urinalysis demonstrated that the inhibition of TAL tubular transport by furosemide resulted in a completely differ­ent pattern of tubular Ca2+ and Mg2+ handling that identifies the TAL as the major nephron segment affected by claudin-10 deficiency.  Interestingly, the different effects on plasma Mg2+ and Ca2+ levels reflect the different major reabsorption sites of these ions. Some 60% of the filtered Mg2+ is reabsorbed in the TAL, compared with only 20% of the filtered Ca2+ load (20). Ca2+ hyperreabsorption in TAL seems to be balanced by reduced (proximal and) distal tubular Ca2+ transport. The hyperreabsorption of divalent cations in mice deficient in claudin-10 is in opposition to the loss of divalent cations seen in mouse models of claudin-16 deficiency and in human patients with mutation in CLDN16 or CLDN19. This finding indicates that claudins in the TAL have functions that differentially affect paracellular cation transport in this segment. Mice deficient for claudin-10b in the TAL exhibit decreased permeability for Na+ and increased permeability for Ca2+ and Mg2+, whereas in mice with claudin-16 or claudin-19 deficiency, decreased sodium per­meability in the TAL is paralleled by decreased reabsorption of Ca2+ and Mg2+.

1. Greger R (1981) Cation selectivity of the isolated perfused cortical thick ascending limb of Henle’s loop of rabbit kidney. Pflugers Arch 390:30–37.
2. Furuse M (2010) Molecular basis of the core structure of tight junctions. Cold Spring Harb Perspect Biol 2:a002907.
3. Konrad M, et al. (2006) Mutations in the tight-junction gene claudin 19 (CLDN19) are associated with renal magnesium wasting, renal failure, and severe ocular in­volvement. Am J Hum Genet 79:949–957.
4.  Simon DB, et al. (1999) Paracellin-1, a renal tight junction protein required for par-acellular Mg2+ resorption. Science 285:103–106.
5. Hou J, et al. (2007) Transgenic RNAi depletion of claudin-16 and the renal handling of magnesium. J Biol Chem 282:17114–17122.
6. Himmerkus N, et al. (2008) Salt and acid-base metabolism in claudin-16 knockdown mice: Impact for the pathophysiology of FHHNC patients. Am J Physiol Renal Physiol 295:F1641–F1647.

 PMID: 22891322  PMCID: PMC3435183   http://www.ncbi.nlm.nih.gov/pubmed/22891322

Paracellin-1 is critical for magnesium and calcium reabsorption in the human thick ascending limb of Henle.

Blanchard A, Jeunemaitre X, Coudol P, Dechaux M, Froissart M, et al.
Université Pierre et Marie Curie, INSERM and Laboratoire de Génétique Moléculaire, Hôpital Universitaire Européen Georges Pompidou, Paris, France. blanch@ccr.jussieu.fr
Kidney Int. 2001 Jun; 59(6):2206-15.

A new protein, named paracellin 1 (PCLN-1), expressed in human thick ascending limb (TAL) tight junctions, possibly plays a critical role in the control of magnesium and calcium reabsorption, since mutations of PCLN-1 are present in the hypomagnesemia hypercalciuria syndrome (HHS).
No functional experiments have demonstrated that TAL magnesium and calcium reabsorption were actually impaired in patients with HHS.
Genetic studies were performed in the kindred of two unrelated patients with HHS.

We found two yet undescribed mutations of PCLN-1 (Gly 162 Val, Ala 139 Val). In patients with HHS, renal magnesium and calcium reabsorptions were impaired as expected; NaCl renal conservation during NaCl deprivation and NaCl tubular reabsorption in diluting segment were intact. Furosemide infusion in CS markedly increased NaCl, Mg, and Ca urinary excretion rates. In HHS patients, furosemide similarly increased NaCl excretion, but failed to increase Mg and Ca excretion. Acute MgCl(2) infusion in CS and ERH patient provoked a dramatic increase in urinary calcium excretion without change in NaCl excretion. When combined with MgCl(2) infusion, furosemide infusion remained able to induce normal natriuretic response, but was unable to increase urinary magnesium and calcium excretion further. In HHS patients, calciuric response to MgCl(2) infusion was blunted.

In patients with HHS, levels of circulating renin and aldosterone were normal, suggesting normal blood and extracellular volume. In addition, HHS patient 2 was normally able to lower her sodium excretion below 10 mmol/day during sodium deprivation, and in HHS pa­tient 1, sodium reabsorption in the diluting segment was normal as assessed by hypotonic saline infusion.  After oral NH4Cl load: Minimal urinary pH was 5.8 (normal value <5.4), and maximal net acid excretion reached only 24 pmol/min (normal value >80). Both subjects had hypocitraturia. The latter data suggested in the two probands distal defect of urinary acidification, probably related to nephrocalcinosis.

Because the filtered load of calcium but not the filtered load of magnesium remains unchanged during acute magnesium infusion in humans, the increase in calcium excretion is a better index of the inhibitory effect of peritubular magnesium on renal tubular divalent cation transport.  Urinary sodium excretion remained almost constant in both subjects during MgCl2 infusion (data not shown). Accordingly, the FECa/FENa ratio, which should remain constant if sodium reabsorption was primarily affected, increased in the CS and EHR patient.  Before the furosemide infusion, serum ultrafilterable (UF) Ca concentrations were similar in patients with HHS and the controls. However, Ca excretion markedly differed and was approximately five times higher in HHS patients than in controls.

In the two patients with homozygous mutations in the PCLN-1 gene, an impairment in renal tubular magne­sium and calcium reabsorption with normal NaCl recla­recla­mation was demonstrated. Accordingly, comparative studies performed under baseline condition in one pa­tient with ERH and in HHS patients demonstrated that the magnesium and calcium excretion in HHS patients were inappropriately high when compared with serum magnesium and calcium concentrations. However, renal NaCl reabsorption in HHS patients was intact. There was no clinical evidence of extracellular fluid volume  contraction. Furthermore, basal circulating renin and aldosterone concentrations were normal and adapted to the normal Na intake. Finally, abnormal NaCl reclama­tion in the diluting segment of the nephron was excluded in one patient, while the other was able to adapt normally to a sodium deprived diet.

This study is the first to our knowledge to demonstrate that homozygous mutations of PCLN-1 result in a selective defect in paracellular Mg and Ca reabsorption in the TAL, with intact NaCl reabsorption ability at this site. In addition, the study supports a selective physiological effect of basolateral Mg(2+) and Ca(2+) concentration on TAL divalent cation paracellular permeability, that is, PCLN-1 activity.   PMID: 11380823   http://www.ncbi.nlm.nih.gov/pubmed/11380823

Development of a Novel Sodium-Hydrogen Exchanger Inhibitor for Heart Failure

Elizabeth Juneman*, Reza Arsanjani, Hoang M Thai, Jordan Lancaster, Jeffrey B Madwed, Steven Goldman
Citation: Elizabeth Juneman, et al. (2013) Development of a Novel Sodium-Hydrogen Exchanger Inhibitor for Heart Failure. J Cardio Vasc Med 1: 1-6

This study was designed to determine the potential therapeutic effects of a new sodium-hydrogen exchanger (NHE-1) inhibitor in the rat coronary artery ligation model of chronic heart failure. After the induction of acute myocardial infarction, rats were entered randomly dose dranging from 0.3 mg/kg, 1.0 mg/kg, and 3.0 mg/kg. Solid state micrometer hemodynamics, echocardiographic, and pressure-volume relationships were measured after 6 weeks of treatment. Treatment with this NHE- 1 inhibitor at 3 mg/kg increased (P< 0.05) ejection fraction from 23±3% (N=6) to 33±2% (N=13) while the 1 mg/kg dose decreased (P< 0.05) the infarct size in CHF rats from 21.7±1.4% (N=7) to 15.9±0.7% (N=3) and prevented (P< 0.05) dilatation of the left ventricle in CHF rats in diastole (1.0±0.1 cm, N=6) to 0.9±0.1 cm, N=10) and in systole (0.9±0.1 cm, N=6) to 0.8±0.1, N=10). These study results suggest that this new NHE-1 inhibitor may be potentially useful in treating CHF with an improvement in maladaptive left ventricule remodeling. Because the mechanism of action of this agent is entirely different than the currently applied approach in treating CHF that focuses on aggressive neurohormonal blockade and because this agent does not adversely affect important hemodynamic variables, further investigations with this agent may be warranted.

Keywords: Congestive heart failure; Sodium/hydrogen exchange; Cardiovascular disease; Cardiovascular drugs; CHF: Chronic Heart Failure; NHE-1: Sodium-Hydrogen Exchanger; NCX: Sodium-Calcium Exchanger; Ca2+: Calcium; Na+: Sodium; Na+-K+ATPase: Sodium-Potassium ATPase; NKCC: Sodium-Potassium-Chloride co-transporter; MI: Myocardial Infarction; BI: Boehringer Ingelheim; LV: left Ventricle; EF: Ejection Fraction; LVD: left ventricular dysfunction; PV: Pressure-Volume; SE: Standard Error; ARB: Angiotensin Receptor Blocker; ACE: Angiotensin Converting Enzyme

Without reviewing the pathophysiology of CHF here, altered calcium (Ca2+) handling is a hallmark of CHF. Intracellular Ca2+ concentration is closely regulated by sodium-calcium exchanger (NCX) and Ca2+ efflux is dependent on the intracellular sodium (Na+) concentration and trans-sarcolemmal Na gradient. Multiple channels including sodium-potassium ATPase (Na+-K+ ATPase), sodium-hydrogen exporter (NHE), sodium-bicarbonate co-transporter, sodium-potassium-chloride co-transporter (NKCC), and sodium-magnesium exchanger are responsible for regulation of intracellular sodium in cardiac myocytes. The intracellular concentration of Na+ is significantly increased in heart failure, primarily due to influx of Na+. The NHE plays an integral part in rise of intracellular Na+ concentration and development of hypertrophy in heart failure. Because of its multifaceted role in myocardial function, there has been interest in examining the effects of NHE-1 inhibitors in heart failure.

In this study we report the physiologic responses of a new NHE-1 inhibitor, in a rodent model of heart failure. Previous evaluation of the pharmacokinetic properties of this agent in rat and dog revealed low clearance and robust oral bioavailability, suggesting a potential for once daily oral administration. This new compound was found to be potentially effective in preventing ischemic injury in isolated cells systems and in ischemic injury in isolated cells systems and in a Langendorff isolated heart preparation. Based on these encouraging a pharmacokinetic data, and the established preclinical roof of principle, the next step in new drug development was to test this inhibitor in an appropriate disease-relevant animal model. For this, we chose the rat coronary ligation model of CHF, which is the established model of chronic ischemic heart failure and well performed in our laboratory. The model with permanent occlusion of the left coronary artery is important because this model a similar to the clinical syndrome of CHF. This rat coronary artery model of CHF is the same model used in the classic study defining the beneficial use of angiotensin converting enzyme inhibition with captopril in the treatment of CHF. Thus results in this model have the potential to be predictive of the clinical response seen in patients.

Results

In vivohemodynamic effect of NHE1: As noted previously by our laboratory, rats with severe CHF compared to Sham had changes (P< 0.05) in right ventricular weight, mean arterial pressure, tau, the time constant of LV relaxation, LV systolic pressure, LV end-diastolic pressure, +LVdP/dt, -LVdP/dt, dead volume and peak developed pressure. In this study, treatment resulted in no changes in body weight, chamber weight or hemodynamics. Because we stopped the lowest dose (0.3 mg/kg) there are only hemodynamic data with this dose in rats with CHF.

Echocardiographic changes in LV function and Dimensions with NHE1: Rats with CHF have decreases in EF accompanied by increases in LV systolic and diastolic dimensions. There was no change in anterior wallsystolic displacement. These data are consistent with other reports in this model showing that at 6 weeks after left coronary artery ligation, rats with large MIs have dilated left ventricles with LV remodeling and poor LV function (14,15). Treatment with the highest dose of 3 mg/kg increased (P< 0.05) ejection fraction from 23±3% (N=6) to 33±2% (N=13). Treatment with 1 mg/kg prevented maladaptive LV remodeling, it prevented (P< 0.05) dilatation of the LV in CHF rats in diastole (1.0±0.1 cm, N=6) to 0.9±0.1 cm, N=10) and in systole (0.9±0.1 cm, N=6) to 0.8±0.1, N=10) with no change anterior wall thickening.

Pressure-Volume relationships:  Although there are no significant changes in the PV relationships for either the Sham or CHF rats, there is a trend for the PV loop in CHF to be shifted toward the pressure axis with treatment. These data are consistent with the trend toward decreases in LV dimensions seen with treatment in CHF rats.

Discussion

This study can be viewed as a corollary of a pilot Phase II clinical trial to look for a signal of a beneficial physiologic effect of this new NHE-1 inhibitor in CHF. In terms of drug development, this is an appropriate approach, i.e., take an agent with a therapeutic focus, with an acceptable toxicology profile, alter its pharmacokinetics to improve its oral delivery and bioavailability and then study the drug in an appropriate animal model. The administration of this agent to rats with CHF after left coronary artery ligation resulted in a therapeutic benefit with an increase in EF and a decrease in infarct size in rats with the largest infarcts. There is a suggestion of the prevention of LV remodeling with decreases in LV end-diastolic and end-systolic dimensions accompanied by a similar trend in the PV loop with a shift toward the pressure axis. There were no changes in hemodynamics.

Importantly, the decrease in infarct size with no changes in hemodynamicswould positively affect LV remodeling by minimizing LV dilatation without changes in LV afterload. From a therapeutic perspective, an agent like this may be advantageous in the treatment of heart failure after MI. The lack of hemodynamic changes is not a clinical problem because we already have agents that decrease afterload and lower LV end-diastolic pressure such as angiotensin converting enzyme (ACE) inhibitors and angiotens in receptor blockers (ARBs). In treating CHF, we also have diuretics to control blood volume, which in turn reduces LV end-diastolic pressure. The other potential advantage of an NHE-1 inhibitor is that as opposed to our current use of aggressive neurohormonal blockade, this represents a different approach to treating CHF. This is attractive because we essentially have exhausted or maximized our effects of neurohumoral blockade and with no real new treatments for CHF introduced in the last 10-15 years, need to look for other approaches to treat CHF.

Drug development is obviously a complicated and expensive undertaking. In exploring this agent, we would proposea stepwise approach. In this case with an agent whose analogs have been studied extensively, our thought would be to perform a larger dose ranging study in CHF rats to define dose response curves for systolic function as well as obtain more information on pharmacokinetics, as well as diastolic function and structural changes.

An attractive aspect of this work is that we are examining an agent with a different mechanisms of action that current treatments for heart failure. The stimulus to study the agent in an animal model of heart failure was based on multifaceted roles of sodium-hydrogen exchangers on myocardial function. Nine isoforms of NHE have currently been identified, with NHE- 1 being the predominant isoform in the plasma membrane of the myocardium [3,24]. Because NHE is activated by intracellular acidosis, angiotensin II, and catecholamines, its activity is expectedly increased in heart failure. Inhibition of NHE-1 has previously been associated with decreased fibrosis, apoptosis, preserved contractility, and attenuation of hypertrophy and development of heart failure.

1. Baartscheer A, van Borren MMGJ (2008) Sodium Ion Transporters as New Therapeutic Targets in Heart Failure. Cardiovasc Hematol Agents Med Chem 6: 229-236.
2. Murphy E, Eisner DA (2009) Regulation of Intracellular and Mitochondrial Sodium in Health and Disease. Circ Res 104: 292-303
3. Despa S, Islam MA, Weber CR, Pogwizd SM, Bers DM (2002) Intracellular Na(+) Concentration is Elevated in Heart Failure but Na/K Pump Function is Unchanged. Circulation 105: 2543-2548.
4. Baartscheer A, Schumacher CA, van Borren MMGJ, Belterman CNW, Coronel R, et al. (2003) Increased Na+/H+-Exchange Activity is the Cause of Increased [Na+]i and Underlies Disturbed Calcium Handling in the Rabbit Pressure and Volume Overload Heart Failure Model. Cardiovasc Res 57: 1015-1024.
5.  Pieske B, Houser SR (2003) [Na+]i Handling in the Failing Human Heart. Cardiovasc Res 57: 874-886.
6.  Engelhardt S, Hein L, Keller U, Klämbt K, Lohse MJ (2002) Inhibition of Na(+)-H(+) Exchange Prevents Hypertrophy, Fibrosis, and Heart Failure in Beta(1)- Adrenergic Receptor Transgenic Mice. Circ Res 90: 814-819.
7.  Chen L, Chen CX, Gan XT, Beier N, Scholz W, et al. (2004) Inhibition and Reversal of Myocardial Infarction-Induced Hypertrophy and Heart Failure by NHE-1 Inhibition. Am J Physiol Heart Circ Physiol 286: 381-387.
8.  Marano G, Vergari A, Catalano L, Gaudi S, Palazzesi S, et al. (2004) Na+/ H+ Exchange Inhibition Attenuates Left Ventricular Remodeling and Preserves Systolic Function in Pressure-Overloaded Hearts. Br J Pharmacol 141: 526-532.
9. Goldman S, Raya TE (1995) Rat Infarct Model of Myocardial Infarction and Heart Failure. J Card Fail 1: 169-177.
10. Gaballa MA, Goldman S (2002) Ventricular Remodeling in Heart Failure. J Card Fail. 8: 476-485.
11. Pfeffer MA, Pfeffer JM, Steinberg C, Finn P (1985) Survival After Experimental Myocardial Infarction: Beneficial Effects of Long-Term Therapy with Captopril. Circulation 72: 406-412.
12. Raya TE, Gay RG, Aguirre M, Goldman S (1989) Importance of Venodilatation in Prevention of Left Ventricular Dilatation after Chronic Large Myocardial Infarction in Rats: A Comparison of Captopril and Hydralazine. Circ Res 64: 330-337.

Related articles in Pharmaceutical Intelligence:

Part I: Identification of Biomarkers that are Related to the Actin Cytoskeleton

Larry H Bernstein, MD, FCAP

http://pharmaceuticalintelligence.com/2012/12/10/identification-of-biomarkers-that-are-related-to-the-actin-cytoskeleton/

Part II: Role of Calcium, the Actin Skeleton, and Lipid Structures in Signaling and Cell Motility

Larry H. Bernstein, MD, FCAP, Stephen Williams, PhD and Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/08/26/role-of-calcium-the-actin-skeleton-and-lipid-structures-in-signaling-and-cell-motility/

Part III: Renal Distal Tubular Ca2+ Exchange Mechanism in Health and Disease

Larry H. Bernstein, MD, FCAP, Stephen J. Williams, PhD
 and Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/09/02/renal-distal-tubular-ca2-exchange-mechanism-in-health-and-disease/

Part IV: The Centrality of Ca(2+) Signaling and Cytoskeleton Involving Calmodulin Kinases and Ryanodine Receptors in Cardiac Failure, Arterial Smooth Muscle, Post-ischemic Arrhythmia, Similarities and Differences, and Pharmaceutical Targets

Larry H Bernstein, MD, FCAP, Justin Pearlman, MD, PhD, FACC and Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/09/08/the-centrality-of-ca2-signaling-and-cytoskeleton-involving-calmodulin-kinases-and-ryanodine-receptors-in-cardiac-failure-arterial-smooth-muscle-post-ischemic-arrhythmia-similarities-and-differen/

Part V: Heart, Vascular Smooth Muscle, Excitation-Contraction Coupling (E-CC), Cytoskeleton, Cellular Dynamics and Ca2 Signaling

Larry H Bernstein, MD, FCAP, Justin Pearlman, MD, PhD, FACC and Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/08/26/heart-smooth-muscle-excitation-contraction-coupling-cytoskeleton-cellular-dynamics-and-ca2-signaling/

Part VI: Calcium Cycling (ATPase Pump) in Cardiac Gene Therapy: Inhalable Gene Therapy for Pulmonary Arterial Hypertension and Percutaneous Intra-coronary Artery Infusion for Heart Failure: Contributions by Roger J. Hajjar, MD

Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/08/01/calcium-molecule-in-cardiac-gene-therapy-inhalable-gene-therapy-for-pulmonary-arterial-hypertension-and-percutaneous-intra-coronary-artery-infusion-for-heart-failure-contributions-by-roger-j-hajjar/

Part VII: Cardiac Contractility & Myocardium Performance: Ventricular Arrhythmiasand Non-ischemic Heart Failure – Therapeutic Implications for Cardiomyocyte Ryanopathy (Calcium Release-related Contractile Dysfunction) and Catecholamine Responses

Justin Pearlman, MD, PhD, FACC, Larry H Bernstein, MD, FCAP and Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/08/28/cardiac-contractility-myocardium-performance-ventricular-arrhythmias-and-non-ischemic-heart-failure-therapeutic-implications-for-cardiomyocyte-ryanopathy-calcium-release-related-contractile/

Part VIII: Disruption of Calcium Homeostasis: Cardiomyocytes and Vascular Smooth Muscle Cells: The Cardiac and Cardiovascular Calcium Signaling Mechanism

Justin Pearlman, MD, PhD, FACC, Larry H Bernstein, MD, FCAP and Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/09/12/disruption-of-calcium-homeostasis-cardiomyocytes-and-vascular-smooth-muscle-cells-the-cardiac-and-cardiovascular-calcium-signaling-mechanism/

Part IX: Calcium-Channel Blockers, Calcium Release-related Contractile Dysfunction (Ryanopathy) and Calcium as Neurotransmitter Sensor

Justin Pearlman, MD, PhD, FACC, Larry H Bernstein, MD, FCAP and Aviva Lev-Ari, PhD, RN

Part X: Synaptotagmin functions as a Calcium Sensor: How Calcium Ions Regulate the fusion of vesicles with cell membranes during Neurotransmission

Larry H Bernstein, MD, FCAP and Aviva Lev-Ari, PhD, RN

68-76.hwp

pone.0070764.g006  Morpholino knockdown of aquaporin-1a1 reduces water influx.       NIHMS262281.html

nihms81087f1  Localization of claudin proteins in mammalian kidney.      F1.medium  intracellular Mg2+ in normal and Mg2+ depleted immortalized mouse distal convoluted tubule (MDCT) cells

F2.small  membrane voltage influences Mg2+ uptake in MDCT cells    pnas.1203834109fig04  Gene expression analysis of renal claudins

Read Full Post »

Translational Research on the Mechanism of Water and Electrolyte Movements into the Cell

Reviewer and Curator: Larry H. Bernstein, MD, FACP

Introduction

This article is the first in a three part curation covering work that has great importance to our understanding of hydration and possibly the effects of dehydration in cell physiology, and studied effects on renal function and brain, with possible implications for heart failure, myocardial contraction, heart rate, and arrhythmiagenesis.  The discovery of aquaporins and the elucidation of potassium channels and selective ion conduction was jointly awarded the Nobel Prize in Chemistry in 2003 to Peter Agre, at the Johns Hopkins School of Medicine, Baltimore, and Roderick Mac Kinnon, at the Howard Hughes Medical Institute, Laboratory of Molecular Neurobiology and Biophysics, Rockefeller University, New York, NY.  The transport of water, it was assumed, is associated with the movements of Na(+), K(+), Ca(2+), Mg(2+).  The calmodulin kinase, rhyanodine, and calcium sparks in the Ca(2+) release from sarcolemma is covered elsewhere in cardiac contraction, skeletal muscle, smooth muscle, and neural stimulation of muscle and adrenergic release.  The sodium/potassium exchange is depicted in diagrams, but not discussed.  In traditional chemistry we would think in terms of a cationic and anionic balance that has to be maintained in charge equivalents on both sides of a membrane.  However, the intricacies of membrane structure as well as active transporters has been delineated and has been a transformative factor in our understanding of organ function in health and disease.

Aquaporin Water Channels

AQUAPORIN WATER CHANNELS: Nobel Lecture, Dec 8, 2003, by Peter Agre. http://www.nobelprize.org/nobel_prizes/chemistry/laureates/2003/agre-lecture.pdfagre-lecture Fig1 Membrane orientation of AQP1

We have studied the aquaporin water channels for several years, and we now understand that they explain how water crosses biological membranes. Our bodies are 70% water, and all other vertebrates, invertebrates, microbes, and plants are also primarily water. The organization of water within biological compartments is fundamental to life, and the aquaporins serve as the plumbing systems for cells. Aquaporins explain how our
brains secrete and absorb spinal fluid, how we can generate aqueous humor within our eyes, how we can secrete tears, saliva, sweat, and bile, and how our kidneys can concentrate urine so effectively. These proteins are fundamental to mammalian physiology, but they are also very important in the lives of microorganisms and plants.
It was correctly proposed  in the 1920’s that water could move through the cell membrane simply by diffusing through the lipid bilayer. The current view is that the lipid bilayer has a finite permeability for water, but, in addition, a set of proteins exists that we now refer to as “aquaporins.” Their existence was suggested by a group of pioneers in the water transport field who preceded us by decades – people including Arthur K. Solomon in Boston, Alan Finkelstein in New York, Robert Macey in Berkeley, Gheorghe Benga in Romania, Guillermo Whittembury in Venezuela, Mario Parisi in Argentina – who by biophysical methods predicted that water channels must exist in certain cell types with high water permeability such as renal tubules, salivary glands, and red cells (reviewed by Finkelstein, 1987).
The difference between diffusional and channel-mediated water perme-ability is fairly distinct. Diffusion is a low capacity, bidirectional movement of water that occurs in all cell membranes, whereas the membranes of a subset of cells with aquaporin proteins have very high capacity for permeation by water.
This permeability is selective, since water (H O) crosses through the membranes with almost no resistance, while acid, the hydronium ion (H O ) does not permeate the proteins. This distinction is essential to life. The movement of water is directed by osmotic gradients, so aquaporins are not pumps or exchangers. They form a simple pore that allows water to rapidly pass through membranes by osmosis. There are also other differences between diffusion and channel-mediated water transport. No inhibitors are known for simple diffusion. In contrast, mercurials were discovered by Robert Macey to inhibit water transport in red cells but water permeability was restored by treatment with reducing agents (Macey and Farmer, 1970). These observations predicted that water channels must be proteins with sulfhydryls and characteristically low Arrhenius activation energy.
A number of investigators using ver y logical approaches attempted to identify the water channel molecule; identification proved a very difficult prolem. Isotopic mercurials labeled several membrane proteins – the anion exchanger (band 3). Solomon and a group of several proteins (band 4.5) by Benga. None of the proteins were isolated, reconstituted, and shown  to transport watter (reviewed by Agre et al., 1993a).

DISCOVERY OF AQP1

The field was essentially stuck, but following the well known scientific approach known as “sheer blind luck,” we stumbled upon the protein. Looking through our notebooks for the earliest studies that showed there was such a protein water channel. We were at that time attempting to raise antibodies in rabbits to the denatured partially purified Rh polypeptide.  The rabbits vigorously produced antibodies, but we failed to recognize initially that our antibody did not react with the core Rh polypeptide that migrated at 32 kDa, seen clearly by silver staining of sodium dodecyl sulfate polyacryamide electrophoresis gels (SDS-PAGE). Instead, our antibodies reacted only with a 28 kDa polypeptide. The 28 kDa was an unrelated protein.  Silver staining of SDS-PAGE migration of the isolated protein revealed a discrete band of 28 kDa in detergent insoluble extracts (it failed to stain with the conventional protein stains such as Coomassie blue). The protein was then purified in large amounts from human red cell membranes (Denker et al., 1988; Smith and Agre, 1991).  The 28 kDa protein was strikingly abundant. With approximately  copies per red cell, it was one of the major proteins in the membrane. The protein had features suggesting that it was a tetrameric membrane-spanning protein – suggesting that it was a channel, but a channel for what? The purified protein also provided us the N-terminal amino acid sequence that we used for cDNA cloning. Using our antibody, we looked at several other tissues and found the protein is also strikingly abundant in human kidney. We observed staining over the apical and basolateral membranes of proximal renal tubules and the descending thin limb of the loops of Henle, but we were still frustrated by our failure to recognize what the protein’s function might be.  My clinical mentor, John C. Parker (1935–1993) at the University of North Carolina at Chapel Hill, was the first to suggest to me that red cells and renal tubules were exceedingly permeable to water. He recommended that we consider a role in membrane water transport. While John did not live to see our later studies, he did live to see our initial discovery and we celebrated together.
Postdoctoral fellow Gregor y Preston cloned the cDNA from an erythroid brary (Preston and Agre, 1991). The coding region corresponded to a 269 amino acid polypeptide, predicted by hydropathy analysis to have six bilayer-spanning domains. Interestingly, the amino terminal half (repeat-1) and the carboxy terminal half of the molecule (repeat-2) were genetically related – about 20% identical. Two loops B and E were more highly related to each other, and each contained the signature motif – asparagine, proline, alanine (NPA) [Fig. 1]. Examining the genetics database, we recognized several sequence-related DNAs from diverse sources: lens of cow eyes, brains of fruit flies, bacteria, and plants. Nevertheless, none was functionally defined.
Figure 1. Membrane orientation of AQP1 predicted from primary amino acid sequence. Two tandem repeats each have three bilayer-spanning domains; the repeats are oriented 180˚ with respect to each other. The loops B and E each contain the conser ved motif, Asn- Pro-Ala (NPA)
These clues heightened our suspicion that the 28 kDa protein was a transporter, so we tested for possible water transport function with our colleague Bill Guggino at Johns Hopkins. We used oocytes the frog Xenopus laevis, a useful model, since frog oocytes have very low water permeability. Control oocytes were injected with water alone; oocytes were injected with 2 ng of cRNA encoding our protein. After days of protein synthesis, the oocytes appeared essentially identical. Then we stressed the oocytes by transferring them to distilled water, and an amazing difference was immediately apparent. Having exceedingly low water perme-
ability, the control oocytes failed to swell. In contrast, the test oocytes were highly permeable to water and exploded like popcorn [Fig. 2] (Preston et al., 1992).  The protein was christened “aquaporin” and is now officially designated “AQP1,” the first functionally defined water channel protein (Agre et al., 1993b).
Figure 2. Functional expression of AQP1 water channels in Xenopus laevis oocytes. Control oocyte (left) was injected with water; AQP1 oocyte (right) was injected with cRNA. The oocytes were transferred to hypotonic buffer. After 30 seconds (top) the AQP1 oocyte has begun to swell; after 3 minutes (bottom), the AQP1 oocyte has exploded. Modified and reprinted from Science with permission (Preston et al., 1992).
We  confirmed the function of this protein by studying the purified AQP1 reconstituted into synthetic lipid vesicles of ~0.1 micron diameter prepared by our colleague Suresh Ambudkar at Johns Hopkins (Zeidel et al., 1992). These simple membrane vesicles were examined by freeze fracture electron microscopy by our colleague Arvid Maunsbach, from the University of Aarhus. When lipid was reconstituted without protein, the membrane surfaces were smooth; however, membranes reconstituted with AQP1 contained many intramembraneous particles 0.01 micron diameter (Zeidel et al., 1994). We tested the membranes for water permeability in collaboration with Mark Zeidel at Har vard Medical School. Using stopped flow transfer to hypertonic buffer, the simple liposomes shrank, reaching equilibrium in about one half
second; this is believed to represent the baseline water permeability. When membranes reconstituted with AQP1 were examined, the shrinking occurred much more rapidly, reaching equilibrium in about 20 milliseconds. The channel-mediated flow of water was confirmed, since it was inhibited with mercurials. We calculated the Arrhenius activation energy (<5 kcal/mol), and we determined the unit permeability to be ~3×10 water molecules per subunit per second. Importantly, we attempted to measure proton permeation of AQP1, but despite massive water permeability, acid permeation was not detected. These studies verified that we had, in fact, isolated the long-sought water channel protein.

STRUCTURE OF AQP1

Subsequent efforts were devoted to identifying the mercurial inhibitory site predicted by the studies of Macey. Mercurials react with free sulfhydryls in the amino acid cysteine. Four cysteines are found in the AQP1 polypeptide, but only the residue in loop E (Cys-189 proximal to the second NPA motif) is inhibited by mercurials. We altered the AQP1 sequence by site-directed mutagenesis and expressed the recombinants in oocytes for water permeability studies. Mutation of this residue to serine (Cys-189-Ser) resulted in full water permeability without mercurial inhibition. When we then replaced the alanine in the corresponding position of loop B with a cysteine (Ala-73-Cys), the protein exhibited mercurial sensitive water permeability (Preston et al., 1993). Substitutions elsewhere in the AQP1 failed to produce this behavior. This suggested to us that loops B and E in opposite parts of the molecule must somehow form the aqueous pore. The model that we concocted turned out to be schematically correct and was termed “the hourglass.” The ancient timepiece allows sand to run from upper chamber to lower chamber; if inverted, the sand will flow in the opposite direction. Six bilayer spanning domains were predicted to surround a central domain containing loop B, dipping into the membrane from the cytoplasmic surface, and loop E, dipping into the membrane from the extracellular surface [Fig. 3 left and right].
Figure 3. Hourglass model for membrane topology of AQP1 subunit.
Left panel – Schematic folding of loops B and E overlap within the lipid bilayer to form a single aqueous pathway.
Right panel – Ribbon model of three dimensional structure of AQP1 subunit confirms hourglass with single aqueous pathway. Modified and reprinted with permission from Jour-
nal of Biological Chemistr y (Jung et al., 1994b) and Journal of Clinical Investigation (Kozono et al., 2002).
The overlap of loops B and E was predicted to form a single aqueous pore through the center of the molecule with the NPA motifs juxtaposed and mercurial inhibitory site alongside (Jung et al., 1994b). The AQP1 protein tetrameric with a central pore in each subunit. Thus, AQP1 is structurally like ion channel proteins where four subunits surround a single central as discussed by Rod MacKinnon in his lecture.
We  then sought to establish the high resolution structure of AQP1 in collaboration with Andreas Engel and his group at the Biozentrum in Basel. We were later joined with Yoshinori Fujiyoshi and his group at Kyoto University. Human red cell AQP1 protein was purified by Barb Smith in our lab; Andreas’s student Tom Walz reconstituted it into synthetic membranes at very high protein concentrations. Under these conditions, the AQP1 protein forms remarkably symmetrical arrays referred to as membrane crystals. By measuring the water permeability, we confirmed that the function was 100% retained, giving us confidence that the structure we deduced would be the biologically relevant structure (Walz et al., 1994).
Figure 4. Functional representation for selective water flow through AQP1 subunit and residues involved in human disease.
Left panel – Schematic of sagittal cross-section of AQP1 reveals bulk water in extracellu- lar and intracellular vestibules of hourglass. These are separated by a 20Å span where water passes in single file with transient interactions with pore-lining residues that prevent hy- drogen bonding between water molecules (bold colors). Two structures are believed to pre- vent permeation by protons (H O ): electrostatic repulsion is created by a fixed positive
charge from pore-lining arginine (R195) at a 2.8Å narrowing in the channel; water dipole reorientation occurs from simultaneous hydrogen bonding of water molecule with side chains of two asparagines residues in NPA motifs (N192 and N76). Two partial positive charges at the center of the channel result from orientation of two non-membrane span- ning alpha helices distal to the NPA motifs

THE AQUAPORIN AND AQUAGLYCEROPORIN PROTEIN FAMILY

While we were pursuing studies of AQP1, several other research groups from around the world became interested in what is now known to be a large family of related proteins. The combined efforts of these labs have led to the molecular identification of 12 mammalian aquaporin homologs, and several hundred related proteins have been recognized in other vertebrates as well as invertebrates, plants, and unicellular micro-organisms. The mammalian homologs may be loosely clustered into two subsets [Fig. 5]. The first is referred to as “classical aquaporins”, since they were initially considered to be exclusive water pores. The second is referred to as “aquaglyceroporins”, since they are permeated by water plus glycerol. Interestingly, E. coli has one member of
each – AqpZ (Calamita et al., 1995), and GlpF, isolated by other investigators much earlier. Together, the mammalian aquaporins and aquaglyceroporins are now known to contribute to multiple physiological processes that occur during our daily lives.
Figure 5. Human aquaporin gene family contains two subsets. Homologs freely permeated by water (classical aquaporins, blue) or water and glycerol (aquaglyceroporins, yellow) are represented. E. coli has one aquaporin (AqpZ) and one aquaglyceroporin (GlpF). Reprinted with permission from Journal of Physiology (Agre et al., 2002)
 The remainder of the Nobel Lecture (2003) can be found at the Nobel Prize site.  This portion is sufficient to cover the genesis and advancement of the water transport discovery.

Urinary Excretion of Aquaporin-2 Water Channel Differentiates Psychogenic Polydipsia from Central Diabetes Insipidus

T Saito, San-e Ishikawa, T Ito, H Oda, F Ando, … and T Saito Division of Endocrinology and Metabolism (Ta.S., S.I., F.A., Mi.H., S.N., To.S.), Department of Medicine, Jichi Medical School, Tochigi 329-0498; and Departments of Medicine and Psychiatry (T.I., H.O., Ma.H.), Tokyo Metropolitan Matsuzawa Hospital, Tokyo, Jp 
correspondence to: San-e Ishikawa, M.D., Division of Endocrinology and Metabolism, Department of Medicine, Jichi Medical School, Tochigi 329-0498, Japan. E-mail: saneiskw@jichi.ac.jp. http://jcem.endojournals.org/full/84/6/2235
The present study was undertaken to determine whether urinary excretion of aquaporin-2 (AQP-2) water channel under ad libitum water intake is of value to differentiate polyuria caused by psychogenic polydipsia from central diabetes insipidus. A 30-min urine collection was made at 0900 h in 3 groups of: 11 patients with central diabetes insipidus (22–68 yr old), 10 patients with psychogenic polydipsia (28–60 yr old), and 15 normal subjects (21–38 yr old). In the patients with central diabetes insipidus, the plasma arginine vasopressin level was low despite hyperosmolality, resulting in hypotonic urine. Urinary excretion of AQP-2 was 37 ± 15 fmol/mg creatinine, a value one-fifth less than that in the normal subjects. In the patients with psychogenic polydipsia, plasma arginine vasopressin and urinary osmolality were as low as those in the patients with central diabetes insipidus. However, urinary excretion of AQP-2 of 187 ± 45 fmol/mg creatinine was not decreased, and its excretion was equal to that in the normal subjects. These results indicate that urinary excretion of AQP-2, under ad libitum water drinking, participates in the differentiation of psychogenic polydipsia from central diabetes insipidus. 
PSYCHOGENIC polydipsia causes a marked polyuria with hypotonic urine (1, 2). Arginine vasopressin (AVP) secretion is suppressed by hypoosmolality caused by excess intake of water. Suppression of AVP release obliges us to differentiate psychogenic polydipsia from central diabetes insipidus. Osmotic stimulation tests have been carried out to determine the reserve function of the posterior pituitary gland. Plasma AVP levels increase in response to an increase in plasma osmolality (Posm) in patients with psychogenic polydipsia but not in those with central diabetes insipidus.
In response to AVP, concentrated urine is produced by water reabsorption across the renal collecting duct (3, 4). Aquaporin-2 (AQP-2) is an AVP-regulated water channel of the collecting duct; it is translocated from the cytoplasmic vesicles to the apical plasma membranes by shuttle trafficking when the cells are stimulated by AVP (5, 6, 7), and it is again redistributed into the cytoplasmic vesicles after removal of AVP stimulation (8). Also, AQP-2 is, in part, excreted into the urine (9, 10). We demonstrated that urinary excretion of AQP-2 is of great value in diagnosing central diabetes insipidus in the hypertonic saline infusion test and impaired water excretion in the acute oral water load test (11, 12).   The present study was undertaken to determine whether urinary excretion of AQP-2, under ad libitum water intake, is a useful tool for diagnosing psychogenic polydipsia.

Subjects and study design

Three groups of subjects were examined in the present study.
[1]  11 patients who had been diagnosed as having idiopathic central diabetes insipidus. They had taken 1-deamino-8-D-AVP (DDAVP) intranasally, twice a day, and discontinued the DDAVP therapy 24 h before the study.
[2] 10 patients were diagnosed as having psychogenic polydipsia. They had been treated for psychiatric disorders, including schizophrenia, atypical psychiatric disorder, and chronic alcoholism.
[3] 15 normal volunteers, with ages ranging from 21–38 yr. (the age range of [1] and [2] reached 60)
All the subjects drank water ad libitum, and 30-min urine collection was made and blood drawn at 0900 h. Urine samples were subjected to measurements of urinary osmolality (Uosm) and urinary excretion of creatinine and AQP-2. Blood samples were used to measure Posm and plasma AVP levels. Uosm and Posm were measured by freezing-point depression (Model 3W2, Advanced Instruments, Needham Height, MA). Urinary creatinine was measured with an automatic clinical analyzer (Model 736, Hitachi Co., Tokyo, Jp). Plasma AVP levels were determined by RIA using AVP RIA kits (Mitsubishi Chemistry, Tokyo, Jp) (13). Urinary excretion of AQP-2 was measured as described below.

RIA of AQP-2

The RIA of urinary AQP-2 was performed by the method described in our previous reports (11, 12). Urinary AQP-2-like immunoreactivity was measured by a specific RIA that used the polyclonal antibody against a synthetic portion (Tyr0-AQP-2[ V257-A271]) of the C-terminal of human AQP-2 raised in rabbits. A synthetic peptide [Tyr0-AQP-2 (V257-A271)] was radioiodinated with iodine-125 (New England Nuclear, Boston, MA) by the chloramine-T method.  All samples were analyzed in duplicate. The intra- and interassay coefficients of variation were less than 10%. The minimal detectable quantity of AQP-2 was 0.86 pmol/tube, and an amount equivalent to 6.9 pmol/tube caused 50% inhibition of binding of the radiolabeled ligand.

Results

In the patients with central diabetes insipidus, the plasma AVP level was low despite hyperosmolality of 297.8 ± 3.4 mosmol/kg H2O, resulting in hypotonic urine (Fig. 1⇓). Urinary excretion of AQP-2 was one-fifth less in the patients with central diabetes insipidus than in the normal subjects. AQP-2 is the AVP-dependent water channel of collecting duct cells and is recycling between the cytoplasmic vesicles and the apical plasma membranes in the cells (5, 6, 7, 8). AQP-2 is partly excreted into the urine, which is approximately 3% of AQP-2 in the collecting duct cells (14). In normal subjects, urinary excretion of AQP-2 is changeable in a wide range in physiological conditions (11). Because urinary excretion of AQP-2 has a positive correlation with plasma AVP levels in normal subjects (11), the reduced urinary excretion of AQP-2 was in concert with the impaired secretion of AVP in central diabetes insipidus.
Figure 1.
Posm, plasma AVP (Pavp), Uosm, and urinary excretion of AQP-2 (UAQP-2), under ad libitum water drinking, in 15 normal subjects (NL, •), 11 patients with central diabetes insipidus (CDI, ○) and 10 patients with psychogenic polydipsia (PP, □). *, P < 0.01; **, P < 0.05 vs. the normal subjects. Value are means ± sem.
In the patients with psychogenic polydipsia, Uosm was as low as that in the patients with central diabetes insipidus (Fig. 1⇑). The plasma AVP level was low because of the reduced Posm, which was derived from an exaggerated intake of water. Urinary excretion of AQP-2, however, was not decreased; and rather, its excretion kept the normal range. The relationship between plasma AVP levels and urinary excretion of AQP-2 is shown in Fig. 2⇓. The urinary excretion of AQP-2 in the patients with psychogenic polydipsia was dissociated from the positive correlation between plasma AVP and urinary excretion of AQP-2 in the normal subjects.
Figure 2.
Relationship between plasma AVP levels and UAQP-2. •, Normal subjects (n = 15); ○, patients with central diabetes insipidus (n = 11); □, patients with psychogenic polydipsia (n = 10). Values are means ± sem.

Discussion

The present study demonstrated the clinical tool, of urinary excretion of AQP-2, in differentiating psychogenic polydipsia from central diabetes insipidus. What is involved in the marked difference in urinary excretion of AQP-2 in these two disorders? There is a possibility that, as patients with psychogenic polydipsia reduce water intake during sleep, antidiuresis may occur periodically at night and the production of AQP-2 be somewhat restored. Because approximately 3% of AQP-2 in collecting duct cells is excreted into the urine, urinary excretion of AQP-2 may keep relatively high, despite hypotonic urine. The difference may come from the periodicity of water intake in a day, in the patients with psychogenic polydipsia. As a whole, these changes may disrupt the positive relationship between urinary excretion of AQP-2 and plasma AVP levels. At the present time, however, other factors involved in urinary excretion of AQP-2 remain undetermined.
In conclusion, urinary excretion of AQP-2, under ad libitum water drinking, participates in the differentiation of polyuria caused by psychogenic polydipsia from central diabetes insipidus.

References

Jose CI, Perez-Cruet J. 1979 Incidence and morbidity of self-induced water intoxication in state mental hospital patients. Am J Psychiatry. 136:221–222.  Medline
Goldman MB, Luchins DJ, Robertson GL. 1988 Mechanisms of altered water metabolism in psychiatric patients with polydipsia and hyponatremia. N Engl J Med. 318:397–403.  Medline
Ishikawa S. 1993 Cellular action of arginine vasopressin in the kidney. Endocr J. 40:373–386.  Medline
Fushimi K, Uchida S, Hara Y, Hirata Y, Marumo F, Sasaki S. 1993 Cloning and expression of apical membrane water channel of rat kidney collecting tubule. Nature. 361:549–552.  CrossRefMedline
Sasaki S, Fushimi K, Saito H, et al. 1994 Cloning, characterization and chromosomal mapping of human aquaporin of collecting duct. J Clin Invest. 93:1250–1256.
Nielsen S, DiGiovanni SR, Christensen EI, Knepper MA, Harris HW. 1993 Cellular and subcellular immunolocalization of vasopressin-regulated water channel in rat kidney. Proc Natl Acad Sci USA. 90:11663–11667.  Abstract/FREE Full Text
Saito T, Ishikawa S, Sasaki S, et al. 1997 Alteration in water channel AQP-2 by removal of AVP stimulation in collecting duct cells of dehydrated rats. Am J Physiol. 272:F183–F191.
Kanno K, Sasaki S, Ishikawa S, et al. 1995 Urinary excretion of aquaporin-2 in patients with diabetes insipidus. N Engl J Med. 332:1540–1545. CrossRefMedline
Elliot S, Goldsmith P, Knepper MA, Haughey M, Olson B. 1996 Urinary excretion of aquaporin-2 in humans: a potential marker of collecting duct responsiveness to vasopressin. J Am Soc Nephrol. 7:403–409.  Abstract
Saito T, Ishikawa S, Sasaki S, et al. 1997 Urinary excretion of aquaporin-2 in the diagnosis of central diabetes insipidus. J Clin Endocrinol Metab. 82:1823–1827.   Abstract/FREE Full Text
Saito T, Ishikawa S, Ando F, et al. 1998 Exaggerated urinary excretion of aquaporin-2 in the pathological state of impaired water excretion dependent upon arginine vasopressin. J Clin Endocrinol Metab. 83:4034–4040.  Abstract/FREE Full Text
Terris J, Ecelbarger CA, Nielsen S, Knepper MA. 1996 Long-term regulation of four renal aquaporins in rats. Am J Physiol. 271:F414–F422. Clin Endocrinology & Metabolism  1999; 84(6):2235-2237   http://dx.doi.org/10.1210/jc.84.6.2235

Comparison of cardiovascular aquaporin-1 changes during water restriction between 25- and 50-day-old rats.

Netti VA, Vatrella MC, Chamorro MF, Rosón MI, Zotta E, Fellet AL, Balaszczuk AM.
Cátedra de Fisiología, Facultad de Farmacia y Bioquímica, Universidad de Buenos Aires, IQUIMEFA, CONICET, Junín 956, C1113AAD, Buenos Aires, Argentina, vnetti@conicet.gov.ar.
Eur J Nutr. Apr 27, 2013
Aquaporin-1 (AQP1) is the predominant water channel in the heart, linked to cardiovascular homeostasis. Our aim was to study cardiovascular AQP1 distribution and protein levels during osmotic stress and subsequent hydration during postnatal growth.
Rats aged 25 and 50 days were divided in: 3d-WR: water restriction 3 days; 3d-WAL: water ad libitum 3 days; 6d-WR+ORS: water restriction 3 days + oral rehydration solution (ORS) 3 days; and 6d-WAL: water ad libitum 6 days. AQP1 was evaluated by immunohistochemistry and western blot in left ventricle, right atrium and thoracic aorta.
Water restriction induced a hypohydration state in both age groups (40 and 25 % loss of body weight in 25- and 50-day-old rats, respectively), reversible with ORS therapy. Cardiac AQP1 was localized in the endocardium and endothelium in both age groups, being evident in cardiomyocytes membrane only in 50-day-old 3d-WR group, which presented increased protein levels of AQP1; no changes were observed in the ventricle of pups. In vascular tissue, AQP1 was present in the smooth muscle of pups; in the oldest group, it was found in the endothelium, increasing after rehydration in smooth muscle. No differences were observed between control groups 3d-WAL and 6d-WAL of both ages.
Our findings suggest that cardiovascular AQP1 can be differentially regulated in response to hydration status in vivo, being this response dependent on postnatal growth. The lack of adaptive mechanisms of mature animals in young pups may indicate an important role of this water channel in maintaining fluid balance during hypovolemic state.

 Clinical application of aquaporin research: aquaporin-1 in the peritoneal membrane

Nishino T, Devuyst O.
Division of Renal Care Unit, Second Department of Internal Medicine, Nagasaki University School of Medicine, Nagasaki, Jp
Peritoneal dialysis (PD) is an established mode of renal replacement therapy based on the exchange of fluid and solutes between blood and a dialysate that has been instilled in the peritoneal cavity. The dialysis process involves osmosis, as well as diffusive and convective transports through the highly vascularized peritoneal membrane. The membrane contains ultrasmall pores responsible for the selective transport of water across the capillary endothelium. The distribution of the water channel aquaporin-1 (AQP1), as well as its molecular structure ensuring an exquisite selectivity for water, fit with the characteristics of the ultrasmall pore. Peritoneal transport studies using AQP1 knockout mice demonstrated that the osmotic water flux across the peritoneal membrane is mediated by AQP1. This water transport accounts for 50% of the ultrafiltration during PD. Treatment with high-dose corticosteroids upregulates the expression of AQP1 in peritoneal capillaries, resulting in increased water transport and ultrafiltration in rats. These data illustrate the potential of the peritoneal membrane as an experimental model in the investigation of the role of AQP1 in the endothelium. They emphasize the critical role of AQP1 during PD and suggest that manipulating AQP1 expression could be clinically useful in PD patients.

Corticosteroids induce expression of aquaporin-1 and increase transcellular water transport in rat peritoneum

Stoenoiu MS, Ni J, Verkaeren C, Debaix H, Jonas JC, Lameire N, Verbavatz JM, Devuyst O.
Division of Nephrology and ENDO Unit, Université Catholique de Louvain Medical School, Brussels, Belgium
J Am Soc Nephrol. Mar 2003; 14(3):555-565.
The water channel aquaporin-1 (AQP1) is the molecular counterpart of the ultrasmall pore responsible for transcellular water permeability during peritoneal dialysis (PD). This water permeability accounts for up to 50% of ultrafiltration (UF) during a hypertonic dwell, and its loss can be a major clinical problem for PD patients. By analogy with the lung, the hypothesis was tested that corticosteroids may increase AQP1 expression in the peritoneal membrane (PM) and improve water permeability and UF in rats. First, the expression and distribution of the glucocorticoid receptor (GR) in the PM and capillary endothelium was documented. Time-course and dose-response analyses showed that a daily IM injection of dexamethasone (1 or 4 mg/kg) for 5 d induced an approximately twofold increase in the expression of AQP1 at the mRNA and protein levels. The GR antagonist RU-486 completely inhibited the dexamethasone effect. The functional counterpart of the increased AQP1 expression was a significant increase in sodium sieving and net UF across the PM, contrasting with a lack of effect on the osmotic gradient and permeability for small solutes. The latter observation reflected the lack of effect of corticosteroids on nitric oxide synthase (NOS) activity and endothelial NOS isoform expression in the PM. In conclusion, corticosteroids induce AQP1 expression in the capillary endothelium of the PM, which is reflected by increased transcellular water permeability and UF. These data emphasize the critical role of AQP1 during PD and suggest that pharmacologic regulation of AQP1 may provide a target for manipulating water permeability across the PM.

Aquaporins: relevance to cerebrospinal fluid physiology and therapeutic potential in hydrocephalus

Owler BK, Pitham T, Wang D.
Kids Neurosurgical Research Unit, Children’s Hospital at Westmead, Westmead NSW 2145, Australia. brian@sydneyneurosurgeon.com.au.
Cerebrospinal Fluid Res.  Sep 22, 2010; 7:15.  http://dx.doi.org/10.1186/1743-8454-7-15.
The discovery of a family of membrane water channel proteins called aquaporins, and the finding that aquaporin 1 was located in the choroid plexus, has prompted interest in the role of aquaporins in cerebrospinal fluid (CSF) production and consequently hydrocephalus. While the role of aquaporin 1 in choroidal CSF production has been demonstrated, the relevance of aquaporin 1 to the pathophysiology of hydrocephalus remains debated. This has been further hampered by the lack of a non-toxic specific pharmacological blocking agent for aquaporin 1. In recent times aquaporin 4, the most abundant aquaporin within the brain itself, which has also been shown to have a role in brain water physiology and relevance to brain oedema in trauma and tumours, has become an alternative focus of attention for hydrocephalus research. This review summarises current knowledge and concepts in relation to aquaporins, specifically aquaporin 1 and 4, and hydrocephalus. It also examines the relevance of aquaporins as potential therapeutic targets in hydrocephalus and other CSF circulation disorders.
PMID: 20860832  PMCID:  PMC2949735

Pathophysiology of the aquaporin water channels

King LS, Agre P.
Department of Medicine, Johns Hopkins University School of Medicine, Baltimore, Maryland 21205, USA.
Annu Rev Physiol. 1996; 58:619-48.
Discovery of aquaporin water channel proteins has provided insight into the molecular mechanism of membrane water permeability. The distribution of known mammalian aquaporins predicts roles in physiology and disease.
Aquaporin-1 mediates proximal tubule fluid reabsorption, secretion of aqueous humor and cerebrospinal fluid, and lung water homeostasis.
Aquaporin-2 mediates vasopressin-dependent renal collecting duct water permeability; mutations or downregulation can cause nephrogenic diabetes insipidus.
Aquaporin-3 in the basolateral membrane of the collecting duct provides an exit pathway for reabsorbed water.
Aquaporin-4 is abundant in brain and probably participates in reabsorption of cerebrospinal fluid, osmoregulation, and regulation of brain edema.
Aquaporin-5 mediates fluid secretion in salivary and lacrimal glands and is abundant in alveolar epithelium of the lung.
Specific regulation of membrane water permeability will likely prove important to understanding edema formation and fluid balance in both normal physiology and disease.

Discovery of aquaporins: a breakthrough in research on renal water transport

van Lieburg AF, Knoers NV, Deen PM.
Department of Pediatrics, University of Nijmegen, The Netherlands.
Pediatr Nephrol. Apr 1995; 9(2):228-34.
Several membranes of the kidney are highly water permeable, thereby enabling this organ to retain large quantities of water. Recently, the molecular identification of water channels responsible for this high water permeability has finally been accomplished. At present, four distinct renal water channels have been identified, all members of the family of major intrinsic proteins.
Aquaporin 1 (AQP1), aquaporin 2 (AQP2) and the mercury-insensitive water channel (MIWC) are water-selective channel proteins, whereas the fourth,
Aquaporin 3 (AQP3), permits transport of urea and glycerol as well. Furthermore, a putative renal water channel (WCH3) has been found.
AQP1 is expressed in apical and basolateral membranes of proximal tubules and descending limbs of Henle,
AQP2 predominantly in apical membranes of principal and inner medullary collecting duct cells and
AQP3 in basolateral membranes of kidney collecting duct cells.
MIWC is expressed in the inner medulla of the kidney and has been suggested to be localised in the vasa recta.
The human genes encoding AQP1 and AQP2 have been cloned, permitting deduction of their amino acid sequence, prediction of their two-dimensional structure by hydropathy analysis, speculations on their way of functioning and DNA analysis in patients with diseases possibly caused by mutant aquaporins. Mutations in the AQP1 gene were recently detected in clinically normal individuals, a finding which contradicts the presumed vital importance of this protein. Mutations in the AQP2 gene were shown to cause autosomal recessive nephrogenic diabetes insipidus. The renal unresponsiveness to arginine vasopressin, which characterises this disease, is in accordance with the assumption that AQP2 is the effector protein of the renal vasopressin pathway.(ABSTRACT TRUNCATED AT 250 WORDS)

Selectivity of the renal collecting duct water channel aquaporin-3

Echevarría M, Windhager EE, Frindt G.
Depart Physiol Biophys, Cornell University Medical College, New York, NY
J Biol Chem. Oct 11, 1996; 271(41):25079-82.
Aquaporin-3 (AQP3) is a water channel found in the basolateral cell membrane of principal cells of the renal collecting tubule as well as in other epithelia. To examine the selectivity of AQP3, the permeability to water (Pf), urea (Pur), and glycerol (Pgly) of Xenopus oocytes injected with cRNA encoding AQP3 was measured. Oocytes injected with cRNA encoding either human or rat aquaporin-1 (AQP1) were used as controls. Although both aquaporins permit water flow across the cell membrane, only AQP3 was permeable to glycerol and urea (Pgly > Pur). The uptake of glycerol into oocytes expressing AQP3 was linear up to 165 mM. For AQP3 the Arrhenius energy of activation for Pf was 3 kcal/mol, whereas for Pgly and Pur it was >12 kcal/mol. The sulfhydryl reagent p-chloromercuriphenylsulfonate (1 mM) abolished Pf of AQP3, whereas it did not affect Pgly. In addition, phloretin (0.1 mM) inhibited Pf of AQP3 by 35%, whereas it did not alter Pgly or Pur. We conclude that water does not share the same pathway with glycerol or urea in AQP3 and that this aquaporin, therefore, forms a water-selective channel.

The aquaporin family of water channels in kidney

Agre P, Nielsen S.
Depart of Med, Johns Hopkins University School of Medicine, Baltimore, MD
Nephrologie. 1996;17(7):409-15.
The longstanding puzzle of membrane water-permeability was advanced by discovery of a new class of proteins known as the “aquaporins” (AQPs). First identified in red blood cells, AQP1 was shown to function as a water channel when expressed in Xenopus oocytes or when pure AQP1 protein was reconstituted into synthetic membranes. Analysis of the primary sequence revealed that the two halves of the AQP1 polypeptide are tandem repeats; site directed mutagenesis studies indicate that the repeats may fold into an obversely symmetric structure which resembles an hourglass. Electron crystallography elucidated the tetrameric organization of AQP1, and functional studies suggest that each tetramer contains multiple functionally independent aqueous pores.
AQP1 is abundant in the apical and basolateral membranes of renal proximal tubules and descending thin limbs, and is also present in multiple extra renal tissues.
AQP2 is expressed only in the principal cells of renal collecting duct where it is the predominant vasopressin (ADH, antidiuretic hormone) regulated water channel. AQP2 is localized in the apical membrane and in intracellular vesicles which are targeted to the apical plasma membranes when stimulated by ADH. Humans with mutations in genes encoding AQP1 and AQP2 exhibit contrasting clinical phenotypes.
AQP3 resides in the basolateral membranes of renal collecting duct principal cells providing an exit pathway for water;
AQP4 is abundant in brain where it may function as the hypothalamic osmoreceptor responsible for secretion of ADH. Continued analysis of the aquaporins is providing detailed molecular insight into the fundamental physiological problems of water balance and disorders of water balance.

Aquaporins in the kidney: from molecules to medicine

Nielsen S, Frøkiaer J, Marples D, Kwon TH, Agre P, Knepper MA.
The Water and Salt Res Center, Anatomy and Exper Clin Res Institutes, University of Aarhus, Aarhus, Denmark. sn@ana.au.dk
Physiol Rev. Jan 2002; 82(1):205-44.  http://dx.doi.org/10.1152/physrev.00024.2001

The molecular identity of membrane water channels long-standing biophysical question of how water crosses long remained elusive until the pioneering discovery of biological membranes specifically, and provided insight, at the molecular level, of AQP1 by Agre and colleagues around 1989 –1991,  The discovery of aquaporin-1 (AQP1) answered the long-standing biophysical question of how water specifically crosses biological membranes. In the kidney, at least seven aquaporins are expressed at distinct sites. AQP1 is extremely abundant in the proximal tubule and descending thin limb and is essential for urinary concentration. AQP2 is exclusively expressed in the principal cells of the connecting tubule and collecting duct and is the predominant vasopressin-regulated water channel. AQP3 and AQP4 are both present in the basolateral plasma membrane of collecting duct principal cells and represent exit pathways for water reabsorbed apically via AQP2. Studies in patients and transgenic mice have demonstrated that both AQP2 and AQP3 are essential for urinary concentration.

Since the discovery of aquaporins, major efforts have been aimed at elucidating their structural organization. Hydropathy analysis of the deduced amino acid sequence of AQP1 led to the prediction that the protein resides primarily within the lipid bilayer (191), consistent with the initial studies of AQP1 in red cell membranes (46). AQP1 contains an internal repeat with the NH – and the first provided a molecular answer to the long-standing COOH-terminal halves being sequence related and each
containing the signature motif Asn-Pro-Ala (NPA) (181,252). This is consistent with earlier observations on the homologous major intrinsic protein from lens, (MIP, nowreferred to as AQP0). When evaluated by hydropathy analysis, six bilayer-spanning domains are apparent (Fig.1); however, the apparent interhelical loops B and E also exhibit significant hydrophobicity. Critical to the topology is the location of loop C which connects the two halves of the molecule. Preston et al. (194) demonstrated that loop C resides at the extracellular surface of the oocytes, confirming the obverse sym-metry of the NH – and COOH-terminal halves of the mol-lar surface of the oocytes, confirming the obverse symmetry of the NH – and COOH-terminal halves of the mol-ecule.The structural organization of other aquaporins such as bacterial aquaporin-Z and plant aquaporins have also been deduced. How can water channels avoid passage of protons (H O )? As predicted, loops B and E are associated by Van der Waals interactions between the two NPA motifs. Free hydrogen bonding occurs in the column of water within the pore, except at the very center where a single water molecule transiently reorients to bond with the two asparagines residues of the NPA motif. This results in minimum resistance to the flow of water, thus permitting kidneys to perform their important physiological roles of reabsorbing water while excreting acid.

FIG. 1. A: schematic representation of the structural organization of aquaporin-1 (AQP1) monomers in the membrane (top and bottom). Aquaporins have six membrane-spanning regions, both intracellular NH and COOH termini, and internal tandem repeats that, presumably, are due to an ancient gene duplication (top). The topology is consistent with an obverse symmetry for the two similar NH – and COOH- 2 terminal halves (bottom). The tandem repeat structure with two asparagine-proline-alanine (NPA) sequences has been proposed to form tight turn structures that interact in the membrane to form the pathway for translocation of water across the plasma membrane. Of the five loops in AQP1, the B and E loops dip into the lipid bilayer, and it has been proposed that they form “hemichannels” that connect between the leaflets to form a single aqueous pathway within a symmetric structure that resembles an “hourglass.” B: AQP1 is a multisubunit oligomer that is organized as a tetrameric assembly of four identical polypeptide subunits with a large glycan attached to only one.

Discovery and Biophysical Characterization of the First Molecular Water Channel AQP1 Expression of AQP1 in X. laevis oocytes by Preston et al. (192) demonstrated that AQP1-expressing oocytes exhibited remarkably high osmotic water permeability (P
cm/s), causing the cells to swell rapidly and explode in hypotonic buffer. The osmotically induced swelling of oocytes expressing AQP1 occurs with a low activation energy and is reversibly inhibited by HgCl or other mercurials. Only inward water flow (swelling) was examined, but it was predicted that the direction of water flow through AQP1 is determined by the orientation of the osmotic gradient. Consistent with this, it was later demonstrated that AQP1-expressing oocytes swell in hyposmolar buffers but shrink in hyperosmolar buffers (160).  Swelling of oocytes expressing AQP1 occurs with a low activation energy and is reversibly inhibited by HgCl or other mercurials. Only inward water flow (swelling) was examined, but it was predicted that the direction of water flow through AQP1 is determined by the orientation of the osmotic gradient. Consistent with this, it was later demonstrated that AQP1-expressing oocytes swell in hyposmolar buffers but shrink in hyperosmolar buffers (160).

Over the past 4 years a series of studies have explored the issues of selectivity and polytransport function of aquaporins. This has led to a division of aquaporins (4) into a group that transports water relatively selectively (the “orthodox” set or “aquaporins”) and a group of water channels that also conduct glycerol and other small solutes in addition to water (the “cocktail” set or aquaglyceroporins). This appears to represent an ancient phylogenetic divergence between glycerol transporters and pure water channels (185). Recently, it has become clear that transport properties are even more diverse, since AQP6 has been demonstrated to conduct anions as well (263), and it has also been demonstrated that aquaporins can be regulated by gating, as discussed below.

The signal transduction pathways have been de­scribed thoroughly in previous reviews. cAMP levels in collecting duct principal cells are in­creased by binding of vasopressin to V2 receptors. The synthesis of cAMP by adenylate cyclase is stim­ulated by a V2 receptor-coupled heterotrimeric GTP-bind-ing protein, Gs. Gs interconverts between an inactive responses to vasopressin. In this study it was demon­strated that changes in AQP2 labeling density of the apical plasma membrane correlated closely with the water per­meability in the same tubules, while there were reciprocal changes in the intracellular labeling for AQP2. In vivo studies using normal rats or vasopressin-deficient Brattleboro rats also showed a marked increase in apical plasma membrane labeling of AQP2 in response to vasopressin or dDAVP treatment.  The acute treatment of rats with vasopressin V2-receptor antagonist or acute water loading (to reduce endogenous vasopressin levels, both re­ducing vasopressin action, resulted in a prominent inter­nalization of AQP2 from the apical plasma membrane to small intracellular vesicles further underscoring the role of AQP2 trafficking in the regulation of collecting duct water permeability.

PGE2 inhibits vasopressin-induced water permeabil­ity by reducing cAMP levels. In preliminary studies, Zelenina et al. investigated the effect of PGE2 on PKA phosphorylation of AQP2 in kidney papilla, and the results suggest that the action of prostaglandins is associated with retrieval of AQP2 from the plasma membrane, but that this appears to be independent of AQP2 phosphorylation by PKA.  Phosphorylation of AQP2 by other kinases, e.g., pro­tein kinase C or casein kinase II, may potentially partici­pate in regulation of AQP2 trafficking (Fig. 9C). Phosphorylation of other cytoplasmic or vesicular regulatory proteins may also be involved. These issues remain to be investigated directly.

Since the fundamentals of the shuttle hypothesis have been confirmed, interest has turned to the cellular mechanisms mediating the vasopressin-induced transfer of AQP2 to the apical plasma membrane. The shuttle hypothesis has a number of features whose molecular basis remains poorly understood. First, AQP2 is delivered in a relatively rapid and coordinated fashion, and vesicles move from a distribution throughout the cell to the apical region of the cell in response to vasopressin stimulation. Furthermore, AQP2 is delivered specifically to the apical plasma membrane. Finally, AQP2-bearing vesicles fuse with the apical plasma membrane in response to vasopressin, but not to a significant degree in the absence of stimulation (e.g., in vasopressin-deficient Brattleboro rats where < 5% of total AQP2 is present in the apical plasma membrane. Thus there must be some kind of a “clamp” preventing fusion in the unstimulated state and/or a “trigger” when activation occurs.

The coordinated delivery of AQP2-bearing vesicles to the apical part of the cell appears to depend on the translocation of the vesicles along the cytoskeletal ele­ments. In particular, the microtubular network has been implicated in this process, since chemical disruption of microtubules inhibits the increase in permeability both in the toad bladder and in the mammalian collecting duct. Because microtubule-disruptive agents inhibit the development of the hydrosmotic response to vaso-pressin, but have no effect on the maintenance of an established response, and because they have been re­ported to slow the development of the response without affecting the final permeability in toad bladders , it has been deduced that microtubules appear to be involved in the coordinated delivery of water channels, without being involved in the actual insertion process.

In addition to increasing cAMP levels in collecting duct principal cells, vasopressin acting through the V2 receptor has also been demonstrated to transiently in­crease intracellular Ca2+. The increase occurs in the absence of activation of the phosphoinositide signaling pathway and has recently been dem­onstrated to be due to activation of ryanodine-sensitive calcium release channels in the collecting duct cells. Buffering intracellular calcium with BAPTA or inhibition of calmodulin completely blocked the water permeability response to vasopressin in isolated perfused inner med­ullary collecting ducts, suggesting a critical role for cal­cium at some step in the process of AQP2 vesicle traffick­ing.

In addition to the acute regulation of collecting duct water permeability brought about by the trafficking of AQP2 described above, it is now clear that there are longer term adaptational changes that modulate this acute response. These occur during prolonged changes in body hydration status and form an appropriate physiolog­ical response to such challenges. However, similar long ­term changes also appear to be important in a wide variety of pathological conditions,  and an understanding of the mechanisms involved in these adaptational responses may provide the basis both for a better understanding of, and for potential therapeutic ap­proaches to, pathological disorders of water balance.  Microtubules are polar structures, arising from microtubule organizing centers (MTOCs), at which their minus ends are anchored, and with the plus ends growing away “into” the cell. In fibroblastic cells, there is a single MTOC in the perinuclear region, and the plus ends project to the periphery of the cell. However, there is increasing evidence that in polarized epithelia microtubules arise from multiple MTOCs in the apical region, with their plus ends projecting down toward the basolateral membrane. If this is the case in collecting duct cells, and there is some evidence that it is , then a minus end-directed motor protein such as dynein would be expected to be involved in the movement of vesicles toward the apical plasma membrane.  Recently, it has been shown that dynein is present in the kidney of several mammalian spe­cies and that both dynein and dynactin, a protein complex believed to mediate the interaction of dynein with vesicles, associate with AQP2-bearing vesicles. It seems likely that dynein may drive the microtubule-dependent delivery of AQP2-bearing vesicles toward the apical plasma mem­brane.

The apical part of the collecting duct principal cells contains a prominent terminal web made up of actin filaments. These also appear to be involved in the hydrosmotic response, since disruption of microfilaments with cytochalasins inhibits the response in the toad bladder. Cytochalasins can also inhibit an estab­lished response, and even the offset of the response. From this it has been concluded that microfilaments are probably involved in the final movement of vesicles through the terminal web, their fusion with the plasma membrane, and the subsequent endocytic retrieval of the water channels. Interestingly, vasopressin itself causes actin depolymerization, suggesting that reor­ganization of the terminal web is an important part of the cellular response to vasopressin, a conclusion reached on morphological grounds by DiBona.

The problem of delivering vesicles to a particular domain and allowing them to fuse when, and only when, a signal arrives is conceptually very similar to the situa­tion in the neuronal synapse. It therefore seemed possible that a molecular apparatus similar to the SNAP/SNARE system described there might be present in the collecting duct principal cells.  There are specific proteins on the vesicles (vSNAREs) and the target plasma membrane (tSNAREs) that interact with components of a fusion complex to induce fusion of the vesicles only with the required target membrane. The process is thought to be regulated by other protein com­ponents that sense the signal for fusion (i.e., increased calcium in the synapse). Several groups have now shown that vSNAREs such as VAMP-2 are present in the collect­ing duct principal cells and colocalize with AQP2 in the same vesicles .

A putative tSNARE, SNAP23, has been found in collecting duct principal cells both in the apical plasma membrane and in AQP2-bearing vesicles. Some soluble components of the fusion complex, including NEM-sensitive factor (NSF) and a-soluble NSF-associated protein (SNAP), have also been identified in these cells. Thus it seems likely that the exocytic insertion of AQP2 is indeed controlled by a set of proteins similar to those involved in synaptic transmission, al­though considerable work remains to be done in isolating and characterizing the components, their regulation, and prime physiological function.

 Body water balance is tightly regulated by vasopressin, and multiple studies now have underscored the essential roles of AQP2 in this.
Vasopressin regulates acutely the water permeability of the kidney collecting duct by trafficking of AQP2 from intracellular vesicles to the apical plasma membrane.
The long-term adaptational changes in body water balance are controlled in part by regulated changes in AQP2 and AQP3 expression levels. Lack of functional AQP2 is seen in primary forms of diabetes insipidus, and reduced expression and targeting are seen in several diseases associated with urinary concentrating defects such as acquired nephrogenic diabetes insipidus, postobstructive polyuria, as well as acute and chronic renal failure. In contrast, in conditions with water retention such as severe congestive heart failure, pregnancy, and syndrome of inappropriate antidiuretic hormone secretion, both AQP2 expression levels and apical plasma membrane targetting are increased, suggesting a role for AQP2 in the development of water retention. Continued analysis of the aquaporins is providing detailed molecular insight into the fundamental physiology and pathophysiology of water balance and water balance disorders.
Three additional aquaporins are present in the kidney. AQP6 is present in intracellular vesicles in collecting duct intercalated cells, and AQP8 is present intracellularly at low abundance in proximal tubules and collecting duct principal cells, but the physiological function of these two channels remains undefined. AQP7 is abundant in the brush border of proximal tubule cells and is likely to be involved in proximal tubule water reabsorption.

Fluid transport across leaky epithelia: central role of the tight junction and supporting role of aquaporins.

Fischbarg J.
Institute of Cardiology Research , A. C. Taquini, University of Buenos Aires and National Council for Scientific and Technical Investigations, Buenos Aires, Argentina. jf20@columbia.edu
Physiol Rev. Oct 2010; 90(4):1271-90. http://dx.doi.org/10.1152/physrev.00025.2009.
The mechanism of epithelial fluid transport remains unsolved, which is partly due to inherent experimental difficulties. However, a preparation with which our laboratory works, the corneal endothelium, is a simple leaky secretory epithelium in which we have made some experimental and theoretical headway. As we have reported, transendothelial fluid movements can be generated by electrical currents as long as there is tight junction integrity. The direction of the fluid movement can be reversed by current reversal or by changing junctional electrical charges by polylysine. Residual endothelial fluid transport persists even when no anions (hence no salt) are being transported by the tissue and is only eliminated when all local recirculating electrical currents are.   The notion that transepithelial movement of water depends on the movement of electrolytes arises from a finding by Peter Curran and Arthur K. Solomon that transintestinal water flow (“solvent” flow) depended on the transport of NaCl (“solute” flux) by that layer. That gave birth to the question of how the flow of solute (or “salt”) is linked to the movement of solvent (or “fluid”), or in the short jargon of the field, how solute-solvent cou­pling arises. 
To be noted, gradientless flow is different from transepithelial osmosis a` la Dutrochet. In this last one, in the presence of an osmotic gradient across an epithelial layer, water obligingly traverses the layer. This is well exempli­fied by the kidney collecting duct, a tight epithelium for which we accept nowadays that the water goes across both cell plasma membranes in series, traversing their aquaporins.  There is also the special case of the anuran skin epithelia, whose intercellular junctions are tight, and which water also appears to traverse through cell membrane aquaporins. As a rule, epithelia specialized to transport fluid do so in the absence of any external osmotic gradient across their layers; that is, fluid is transported between compartments of similar osmolarity.  That gave birth to the question of how the flow of solute (or “salt”) is linked to the movement of solvent (or “fluid”), or in the short jargon of the field, how solute-solvent cou­pling arises.
The progression of the ideas on fluid transport is linked to those in a parallel field, that of water channels.  After early advances in their characterization and isolation, they were molecularly identified by Peter Agre and co-workers in the early 1990s, who termed them aquaporins (AQPs). It was subsequently de­termined that AQPs were present in many fluid transport­ing epithelia  and were also present in water-perme­able kidney segments while absent in relatively water-impermeable ones . By then, the measurements of osmotic permeabilities of epithelial cell membranes had been refined using video microscopy techniques. The lab­oratories of Kenneth Spring (working on gallbladders)  and of the Welling brothers (working on kidney proximal tubule) found rather high osmotic perme­ability (or “filtration” permeability, Pf) values (Persson and Spring: 550 and 1,200 pm/s for the apical and baso-lateral membranes, respectively; Welling: -300 pm/s). Both laboratories suggested that, given such high Pf values, a few milliosmoles of osmotic pressure difference across the cell boundaries would suffice to drive the transported fluids through the cells.

There had been all along experimental evidence for the diverging view that fluid transport across leaky epi­thelia took place via paracellular, transjunctional water flow. That contrary evidence came from the laboratories of Adrian Hill using gallbladder, John Pappenheimer and his fellow James Madara using intestine, and Guillermo Whittembury and Gerhard Malnic using kidney proximal tubule. The contrary view of paracellular flow had remained a minority opinion. Still and all, these “rebels” stood their ground, led by an utterly unconvinced Adrian Hill. Con­sidering the divergent views, Kenneth Spring and col­leagues decided to take the bull by the horns and use confocal microscopy to look for evidence for or against transjunctional water flow in epithelia.
Paracellular, transjunctional fluid flow in an absorbing epithelium would lead to significant dilution of a paracellular fluorescent marker trapped in the inter­cellular spaces, which in turn would be detectable by the optical sectioning methods they mastered; all very ele­gant, for sure.

And so we come to the paper Spring and colleagues published in May of 1998  reporting that they had found no transjunctional water flow in cultured Madin-Darby canine kidney (MDCK) cell layers. Understandably, their statement had a very large impact. And yet, only some months afterwards, this notion had to be revised as it became clear that the preparation they had chosen presumably transported little if any water. By Spring’s own admission in October of the same 1998, “ . . . the fluid transport rate of MDCK cells is only about 1% of that of the renal proximal tubule… ”  To spell out the obvious, little or no fluid transport means no transjunctional (or trans-cellular) water flow either, so in perspective, the findings of Spring and colleagues (“absence of junctional flow”) bring no surprise and have no bearing on the issue of the route of fluid flow in general.

After the demise of the 1998 paper above, doubts about local osmosis continued to be fueled. Adrian Hill had been joined in his criticism of it by Thomas Zeuthen and Ernest Wright. In particular, Zeuthen and co-workers had developed an alternative model for transcellular water transfer based on molecular cotransport through transporters. Predictably, Hill’s views were newly sought out. In a thorough review written with his wife and colleague Bruria Shachar-Hill, they restated the evidence from theirs and collaborating laboratories for junctional flow for Necturus and rabbit gallbladder, Necturus intestine, Rhodnius Malpighian tubule, and rat and rabbit salivary gland. In addition, they gave a convincing account of the evidence consistent with junctional water flow for renal proximal tubule, exocrine gland (salivary, lacrimal), and small intestine. Here we will simply call attention to those arguments and will concentrate on other arguments plus additional evidence of our own.

By the end of the 1990s, Alan Verkman’s laboratory had been investigating the physiological effects of knock­ing out AQPs in mice.  The dele­tion of AQPs resulted in drastic decreases of cell mem­brane osmotic permeability, but only in rather mild decreases in rates of fluid transport, and this last to boot only in tissues that transported fluid at high rates. Verkman and colleagues generally discuss those results in a guarded manner, underlining the role of aquaporins as routes for cell water permeability without making pro­nouncements on the mechanism of transtissue fluid trans­port. Yet, paraphrasing the comments by Hill and col­leagues in another cogent review, the effects seen in the AQP knockouts are sometimes difficult to explain, and not commensurate with the deletion of what would be hypothetically a major route for transcellular transtissue water transfer.

Perhaps the existence and the location of electrogenic transporters and channels are telling us something very fundamental about the function of these layers. There does not seem to be an explanation of why epithelia in general, and specifically leaky epithelia, would have evolved to have an electrical potential difference across the layer. In principle, salts could simply be transported neutrally. In a similar vein, apical Na channels that allow Na to leak back into the cell would not make sense if the task of an epithelial cell would be to transport salt from the serosal (basal) to the luminal (apical) side. However, both of these apparent incongruencies suddenly make sense if the raison d’être of these epithelia is to perform tasks such as electro-osmosis. The electrical potential might not be an evolutionary leftover but a central fea­ture. The Na channel would not be apical by accident but to help build up the local current meant for electro-osmosis. As mentioned above, aside from the corneal endothelium , there is evidence for electro-osmosis in small intestine, kidney proximal tubule, and frog skin glands. Hence, it would be desirable if the presence of electro-osmosis would be explored in other fluid-transporting epithelia.

Electro-osmotic coupling would result in somewhat (perhaps 30%) hypotonic emerging fluid. This entails that the fluid left behind at the intercellular spaces might be correspondingly hypertonic. Such osmolarity difference in turn might be sensed by the cell and trigger mechanisms that would affect sites for regulation at basolateral and apical sites for HCO3  and Na transports, and perhaps also at the junction so as to modify the characteristics of the coupling. It is conceivable that such regulation might take place with some degree of period­icity. There may be a role for AQP1 in this regulation, which would explain the mild effects seen on fluid trans­port in this and other preparations in experiments done with AQP1 null cells. This would explain what has been noted by Verkman and colleagues, namely, that effects of AQP deletion are more pronounced in epithelia that gen­erate higher rates of fluid transport. Thus AQP deletion reduced near-isosmolar fluid transport in kidney proximal tubule and salivary gland, where fluid transport is rapid, but not in lung, lacrimal gland, sweat gland, or corneal endothelium where fluid trans­port is relatively slow.

Aquaporin (AQP) 1 is the only AQP present in these cells, and its deletion in AQP1 null mice significantly affects cell osmotic permeability (by > 40%) but fluid transport much less ( > 20%), which militates against the presence of sizable water movements across the cell. In contrast, AQP1 null mice cells have reduced regulatory volume decrease (only 60% of control), which suggests a possible involvement of AQP1 in either the function or the expression of volume-sensitive membrane channels/transporters. A mathematical model of corneal endothelium we have developed correctly predicts experimental results only when paracellular electro-osmosis is assumed rather than transcellular local osmosis. Our evidence therefore suggests that the fluid is transported across this layer via the paracellular route by a mechanism that we attribute to electro-osmotic coupling at the junctions. From our findings we have developed a novel paradigm for this preparation that includes

1) paracellular fluid flow;
2) a crucial role for the junctions;
3) hypotonicity of the primary secretion; and
4) an AQP role in regulation rather than as a significant water pathway.
These elements are remarkably similar to those proposed by the laboratory of Adrian Hill for fluid transport across other leaky epithelia.

Related articles in Pharmaceutical Intelligence:

Part I: Identification of Biomarkers that are Related to the Actin Cytoskeleton

Larry H Bernstein, MD, FCAP

http://pharmaceuticalintelligence.com/2012/12/10/identification-of-biomarkers-that-are-related-to-the-actin-cytoskeleton/

Part II: Role of Calcium, the Actin Skeleton, and Lipid Structures in Signaling and Cell Motility

Larry H. Bernstein, MD, FCAP, Stephen Williams, PhD and Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/08/26/role-of-calcium-the-actin-skeleton-and-lipid-structures-in-signaling-and-cell-motility/

Part III: Renal Distal Tubular Ca2+ Exchange Mechanism in Health and Disease

Larry H. Bernstein, MD, FCAP, Stephen J. Williams, PhD
 and Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/09/02/renal-distal-tubular-ca2-exchange-mechanism-in-health-and-disease/

Part IV: The Centrality of Ca(2+) Signaling and Cytoskeleton Involving Calmodulin Kinases and Ryanodine Receptors in Cardiac Failure, Arterial Smooth Muscle, Post-ischemic Arrhythmia, Similarities and Differences, and Pharmaceutical Targets

Larry H Bernstein, MD, FCAP, Justin Pearlman, MD, PhD, FACC and Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/09/08/the-centrality-of-ca2-signaling-and-cytoskeleton-involving-calmodulin-kinases-and-ryanodine-receptors-in-cardiac-failure-arterial-smooth-muscle-post-ischemic-arrhythmia-similarities-and-differen/

Part V: Heart, Vascular Smooth Muscle, Excitation-Contraction Coupling (E-CC), Cytoskeleton, Cellular Dynamics and Ca2 Signaling

Larry H Bernstein, MD, FCAP, Justin Pearlman, MD, PhD, FACC and Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/08/26/heart-smooth-muscle-excitation-contraction-coupling-cytoskeleton-cellular-dynamics-and-ca2-signaling/

Part VI: Calcium Cycling (ATPase Pump) in Cardiac Gene Therapy: Inhalable Gene Therapy for Pulmonary Arterial Hypertension and Percutaneous Intra-coronary Artery Infusion for Heart Failure: Contributions by Roger J. Hajjar, MD

Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/08/01/calcium-molecule-in-cardiac-gene-therapy-inhalable-gene-therapy-for-pulmonary-arterial-hypertension-and-percutaneous-intra-coronary-artery-infusion-for-heart-failure-contributions-by-roger-j-hajjar/

Part VII: Cardiac Contractility & Myocardium Performance: Ventricular Arrhythmiasand Non-ischemic Heart Failure – Therapeutic Implications for Cardiomyocyte Ryanopathy (Calcium Release-related Contractile Dysfunction) and Catecholamine Responses

Justin Pearlman, MD, PhD, FACC, Larry H Bernstein, MD, FCAP and Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/08/28/cardiac-contractility-myocardium-performance-ventricular-arrhythmias-and-non-ischemic-heart-failure-therapeutic-implications-for-cardiomyocyte-ryanopathy-calcium-release-related-contractile/

Part VIII: Disruption of Calcium Homeostasis: Cardiomyocytes and Vascular Smooth Muscle Cells: The Cardiac and Cardiovascular Calcium Signaling Mechanism

Justin Pearlman, MD, PhD, FACC, Larry H Bernstein, MD, FCAP and Aviva Lev-Ari, PhD, RN

http://pharmaceuticalintelligence.com/2013/09/12/disruption-of-calcium-homeostasis-cardiomyocytes-and-vascular-smooth-muscle-cells-the-cardiac-and-cardiovascular-calcium-signaling-mechanism/

Part IX: Calcium-Channel Blockers, Calcium Release-related Contractile Dysfunction (Ryanopathy) and Calcium as Neurotransmitter Sensor

Justin Pearlman, MD, PhD, FACC, Larry H Bernstein, MD, FCAP and Aviva Lev-Ari, PhD, RN

Part X: Synaptotagmin functions as a Calcium Sensor: How Calcium Ions Regulate the fusion of vesicles with cell membranes during Neurotransmission

Larry H Bernstein, MD, FCAP and Aviva Lev-Ari, PhD, RN

 1743-8454-7-15-1  Distribution in brain of aquaporin-1 (AQP1, blue) and AQP4 (orange), schematically illustrated on a sagittal section of a human brain
centralpore-small  Tetrameric Pore                     AQP-highlight
Created with The GIMP                           Gating of aquaporins
AQP-thumbnail  Gas Molecules Commute into Cell      aqpz-glpf  water channels
GlpF-ABF  Molecular Obstacle Course              nihms365271f1   Roles of water-selective aquaporins (AQPs, shown in purple).
building_a_model-02-full     nihms365271f2  Roles of water-glycerol-transporting aquaporins (aquaglyceroporins).

Read Full Post »