Designing Heart Performance by Gene Transfer

Size: px
Start display at page:

Download "Designing Heart Performance by Gene Transfer"

Transcription

1 Physiol Rev 88: , 2008; doi: /physrev Designing Heart Performance by Gene Transfer JENNIFER DAVIS, MARGARET V. WESTFALL, DEWAYNE TOWNSEND, MICHAEL BLANKINSHIP, TODD J. HERRON, GUADALUPE GUERRERO-SERNA, WANG WANG, ERIC DEVANEY, AND JOSEPH M. METZGER Department of Integrative Biology and Physiology, University of Minnesota Medical School, Minneapolis, Minnesota I. Introduction 1568 A. Perspective 1568 B. Scope of review 1568 II. Cardiac Gene Transfer Tools and Principles 1569 A. Viral vectors 1569 B. Nonviral vectors 1574 C. In vivo vector delivery techniques 1575 D. In vitro gene transfer/acute genetic engineering 1578 III. Ca 2 Handling Proteins 1579 A. Regulators of Ca 2 release from the SR 1581 B. Regulators of cytoplasmic Ca 2 removal 1584 C. Ca 2 binding proteins that modulate cardiac performance 1589 IV. Sarcomeric Targets and Templates 1592 A. Protein turnover and stoichiometry 1593 B. Thin filament proteins, isoforms, mutants, and chimeras 1595 C. Thick filament proteins 1601 V. Cytoskeletal Proteins 1605 A. Dystrophin and dystrophin-associated proteins 1605 B. Intermediate filaments (desmin) 1608 C. Microtubules 1609 VI. Cardiac Signaling Pathways 1609 A. Gene transfer influencing the -adrenergic signaling pathway 1610 B. Gene transfer of Ca 2 /calmodulin kinase 1613 C. Gene transfer and PKC signaling 1614 D. Gene transfer of protein phosphatases 1620 E. Gene transfer and MAPK signaling 1620 F. Myocardial nitric oxide synthase signaling and contractile function 1621 G. Other signaling proteins of interest 1623 VII. Future Directions 1624 Davis J, Westfall MV, Townsend D, Blankinship M, Herron TJ, Guerrero-Serna G, Wang W, Devaney E, Metzger JM. Designing Heart Performance by Gene Transfer. Physiol Rev 88: , 2008; doi: / physrev The birth of molecular cardiology can be traced to the development and implementation of high-fidelity genetic approaches for manipulating the heart. Recombinant viral vector-based technology offers a highly effective approach to genetically engineer cardiac muscle in vitro and in vivo. This review highlights discoveries made in cardiac muscle physiology through the use of targeted viral-mediated genetic modification. Here the history of cardiac gene transfer technology and the strengths and limitations of viral and nonviral vectors for gene delivery are reviewed. A comprehensive account is given of the application of gene transfer technology for studying key cardiac muscle targets including Ca 2 handling, the sarcomere, the cytoskeleton, and signaling molecules and their posttranslational modifications. The primary objective of this review is to provide a thorough analysis of gene transfer studies for understanding cardiac physiology in health and disease. By comparing results obtained from gene transfer with those obtained from transgenesis and biophysical and biochemical methodologies, this review provides a global view of cardiac structure-function with an eye towards future areas of research. The data presented here serve as a basis for discovery of new therapeutic targets for remediation of acquired and inherited cardiac diseases /08 $18.00 Copyright 2008 the American Physiological Society 1567

2 1568 DAVIS ET AL. I. INTRODUCTION A. Perspective Heart disease is the leading cause of combined morbidity and mortality in the western world. Globally, it is estimated that over the course of the next three decades heart disease will be the leading cause of death worldwide, including both high- and low-income countries (970). The past years have seen tremendous insights into the molecular underpinnings of cardiac performance leading to clinically relevant therapeutics to treat heart disease. Nonetheless, the growing burden of cardiovascular disease in this country and throughout the world necessitates continued vigor directed at the mechanistic basis of heart disease with the goal of identifying new therapeutic targets and implementing effective treatment modalities. The elucidation of the human genome, together with a growing appreciation of the complexities of cardiac gene expression and proteome, lends hope that new discoveries will be forthcoming in treating acquired and inherited diseases of the heart. Cardiac gene transfer presents a unique strategy to design cardiac performance by tailoring specific physiological outcomes in the heart. B. Scope of Review This review is focused on the application of recombinant viral vector systems as gene delivery vehicles to the normal and diseased heart. One could argue that the birth of molecular cardiology was ushered in during the early 1990s by the development and implementation of genetic strategies for targeted engineering of gene expression in cardiac muscle. Since that time, there have been a number of excellent reviews focused on the application of transgenesis and ES cell gene targeting in mammals relating to the heart (151, 400, 517, 746, 999). This review concentrates on another avenue of gene-based engineering by featuring emergent viral vector technologies. First, the origins and applications of gene transfer technologies for the heart are reviewed, with an emphasis on the strengths and limitations of recombinant viral vector and nonviral systems for cardiac gene delivery. Next, cardiac muscle targets and the applications of vector technology to the heart are highlighted by discussing gene transfer of key elements of cardiac excitation-contraction (EC) coupling, with an emphasis on intracellular Ca 2 handling and contractile/regulatory proteins of the sarcomere (Fig. 1). In concert, new discoveries in the cytoskeletal matrix as applied to inherited and acquired cardiac disease are discussed with an eye towards new targets for acute and long-term genetic engineering in the heart. Finally, we FIG. 1. Cardiac muscle subcellular organization. Key elements of cardiac muscle structure and function: from the organ to the molecular level with emphasis on intracellular Ca 2 handling and the constituents of the contractile apparatus, the cardiac sarcomere (circular inset). The adult myocardium is comprised primarily of striated muscle cells (cardiac myocytes) organized as a functional syncytium such that a single stimulus causes the entire myocardium to synchronously depolarize and contract. Within a cardiac myocyte, sarcomeres are arranged in series and in parallel providing optimal contractile architecture. Surrounding each myofibril is a highly organized sarcoplasmic reticulum (SR, green) and transverse tubule (T-tubules, blue) network that contains the elements responsible for the electrochemical coupling from action potential to Ca 2 to force generation. The dynamic interplay of these elements forms the basis of excitation-contraction (EC coupling) in cardiac myocytes.

3 CARDIAC MUSCLE GENE TRANSFER 1569 review vector-based approaches for modifying essential components of the cell signaling network in the heart. Recent advances in gene transfer of cardiac membrane channels and biological pacemakers are addressed elsewhere (17, 184, 185, 524). Collectively, the goal of this review is to provide the reader with a comprehensive analysis of the state of the art in cardiac gene transfer with the aim of identifying and evaluating new targets and opportunities directed at genebased remediation of cardiac disease in vivo. II. CARDIAC GENE TRANSFER TOOLS AND PRINCIPLES With the continued advancement in gene transfer technologies, genetic manipulation of cardiac muscle both in vivo and in vitro has gained tremendous momentum as an important experimental and therapeutic reagent. Gene transfer, either through permanent modification of the mammalian genome or expression of a transgene in somatic cells, is a powerful experimental tool for resolving basic science questions as well as discovering the primary etiologies and mechanistic basis for disease pathogenesis. Transgenic animal models have been essential for understanding the effects of gene expression on organ function in a physiologically relevant and integrative context. The confounding influence, however, of complex compensatory (mal)adaptations may make data interpretation from transgenic animal models difficult. Specifically, it is challenging to ascertain whether a functional defect is a direct or adaptive manifestation of a given gene product. Thus results can represent a combination of primary and secondary outcomes. Acute gene transfer has the potential to alleviate these issues. For instance, in vitro gene transfer to isolated adult cardiac myocytes utilizes a stably differentiated primary cardiac muscle cell and offers an experimental system devoid of complex systemic and environmental interactions. In addition to experimental advantages, direct gene transfer technologies constitute a viable therapeutic modality for remediation of acquired and inherited cardiac diseases. The field of gene therapy and gene transfer in general has utilized an array of unique vector systems. As a thorough review of all of these systems could easily fill a textbook, this section only examines commonly used gene transfer vectors (Table 1) for myocardial and isolated myocyte transduction and briefly describes their major advantages and drawbacks. The field of immunology, as it pertains to delivered transgenes, could also constitute an entire review by itself. As such, discussions of immunology here are focused only on responses to a particular gene transfer vector. Gene transfer vectors are generally classified as viral or nonviral based, and this distinction provides the framework for the following discussion. A. Viral Vectors 1. Retroviral vectors Retroviral vectors are derived from a variety of wildtype retroviruses. Moloney murine leukemia virus (MLV) and lentivirus are two common examples, but employing less common retroviruses, such as foamy virus, for gene transfer is also gaining in popularity (39, 555). Despite many differences, these vectors all share common traits. One is the storage of genetic information in the form of single-stranded RNA that is reverse-transcribed into double-stranded DNA (provirus) once the virus enters the host cell (636). A vital characteristic of retroviral infection is its integration as double-stranded DNA into the host s genome (636). Unlike some retroviruses, lentivirus can efficiently transduce nondividing cells such as cardiac myocytes (633, 1009), and therefore, this section will focus on the use of lentiviral vectors (503). As lentivirus integrates into the host genome, lifetime transduction could potentially be achieved following a single transduction event. This is not without risk, however, as any integrating vector holds the potential for serious insertional mutagenesis events. Like many integrating vectors, lentivirus shows preference for inserting into active chromatin (123, 162, 472). TABLE 1. Viral vectors for cardiac gene transfer Viral Vector Capsid Type Packaging Size, kb Delivery Method Toxicity Adeno-associated Proteinaceous 4.5 Direct injection, local vascular isolation, virus (AAV) systemic vascular delivery Adenovirus Proteinaceous 8 10 Direct injection, local vascular isolation Gutted adenovirus Proteinaceous 36 Direct injection, local vascular isolation Lentivirus Enveloped proteinaceous 8 10 Direct injection Summary of four commonly used cardiac gene transfer vectors reviewed in this article. Capsid structure, average vector packaging size, delivery methods, and relative toxicity are described. Toxicity is rated in a relative scale, where the number of signs represents increasing toxicity. Adenovirus refers to first- and second-generation vectors. The packaging size of lentivirus vectors is not an absolute limit, though larger inserts package with progressively less efficiency.

4 1570 DAVIS ET AL. Lentivirus is a retrovirus related to the HIV virus (443). It contains a proteinaceous capsid surrounded by an envelope derived from the host plasma membrane. Production of the vector is an ever-evolving field, with improvements being made in production efficiency, purity, and safety (481). Production generally involves either a stably transduced cell line, which buds off lentivirus vectors into the supernatant and is collected and concentrated, or a plasmid cotransfection system that introduces a genome coding plasmid and helper virus into cells (103, 418, 515, 529). The plasmid cotransfection system is widely used to produce lentiviral vectors and has undergone a number of generations of development. Each generation was aimed at increasing growth efficiency while reducing the chance of generating replication-competent viruses. Currently, lentiviral vectors cannot be grown to titers on par with adenoviral or adeno-associated viral vectors (AAV) (294, 529, 807). Also, unlike vectors such as AAV, highly purified, large-scale, lentiviral vector production is extremely challenging. Lentivirus vectors are mostly concentrated but with some concomitant impurities. However, technologies such as high-performance liquid chromatography (HPLC) offer the possibility of truly purifying vector stocks (796). A final disadvantage is that lentiviral vectors are less stable than other vectors and are more difficult to manipulate due to this lability (529). Currently, relatively few studies have been reported using lentivirus vectors to transduce the myocardium in vivo (71, 238). Although these vectors are capable of transducing % of cardiac myocytes in vitro (71, 588, 779, 780), in vivo efficiencies rarely achieve a transduction efficiency of above 30% (71, 238). For cardiac expression, direct injection is by far the most efficient means of delivery, with little expression seen after vascular delivery (996). However, in cardiac transplant rejection studies, this relatively low level of expression has yielded significant results (1008). The immunology/toxicology of lentivirus vectors when used for cardiac delivery is not yet well understood, but data will undoubtedly be available in the future. Lentiviral vectors represent an appealing method for cardiac transduction. They offer the ability to stably transduce nondividing cardiac myocytes, a potentially useful experimental/therapeutic characteristic. As the use of lentiviral vectors to transduce the myocardium is a relatively young field, many questions remain unanswered. Future studies will have to focus on increasing transduction efficiencies, determining optimal administration methods, and elucidating the host immune response following vector administration/transduction. Advances in production and purification of these vectors will also expand the potential applications and ease of use of this vector. 2. Adenoviral vectors Adenovirus-based vectors (Ad) have been a workhorse for a variety of gene transfer studies spanning several decades. Despite its limitations, this experimental approach will likely continue to be a valuable tool in the future. Along with AAV vectors, Ad vectors represent one of the most efficient means of both in vivo and in vitro cardiac transduction. Adenovirus has been an invaluable reagent for cardiac muscle gene transfer (170, 245, 294, 439, 813, 858, 966). Adenovirus is a member of the Adenoviridae class of viruses. It is a double-stranded DNA virus (dsdna) with a 36-kb genome capped with inverted terminal repeats (ITRs). The ITRs function as origins of viral genome replication. The genome encodes dozens of protein products divided into early and late transcriptional events using a variety of space-efficient internal promoters and splice sites. Adenovirus has a nonenveloped icosahedron capsid (60 90 nm in diameter) surrounding the dsdna virion (119, 267). Its proteinaceous capsid consists of many different proteins with functions ranging from providing structure to docking and infection. The 12 vertices each contain a penton base and fibrous spike that is used to attach the virus to coxsackievirus (CAR) or adenovirus receptors on the host cell membrane (Fig. 2) (46). Once attached, the adenovirus is endocytosed and transported to the endosome where the virion is thought to escape through a ph-dependent endoplasmolysis (291, 971). These events occur over an estimated min time span (291, 971), and subsequently the dsdna migrates to the nucleus for transcription (466). There are more than 50 infectious adenoviral serotypes, with some being causative agents of human diseases like conjunctivitis and the common cold. Engineered replication-deficient serotypes 2 and 5 are the most commonly used biomedical reagents. Adenoviral vectors are produced in several varieties. The early generation vectors (sometimes called first- or second-generation Ad vectors) contain deletions in key areas of their genome. These deletions render the adenovirus replication deficient in nonpermissible cells and provide a substantial cloning capacity ( 7 kb, Table 1) for exogenous expression cassettes. Vector DNA is transfected into permissive packaging cells (commonly HEK 293), which are eventually lysed as a result of viral production. The viral lysate is plaque purified, and a pure clone is used to seed more packaging cells. The viral lysate/cell debris is collected, and the Ad vector is purified/concentrated using a variety of methods such as centrifugation or chromatography. Gutted adenovirus has large portions of the adenoviral genome deleted, yielding vectors with a much larger cloning capacity (30 40 kb, Table 1) relative to Ad and AAV (35, 196, 264, 265, 807). Gutted adenovirus can be

5 CARDIAC MUSCLE GENE TRANSFER 1571 grown concurrently with a helper virus (which resembles first-generation Ad vectors) to provide critical functions in trans. After production, the gutted vector is separated from the helper virus by density using equilibrium centrifugation. Commonly, the packaging signal of the helper virus is flanked by recombinase targets resulting in excision of the packaging signal in the presence of the proper recombinase. The recombinase is expressed in the packaging cells, thus limiting the amount of packaged helper that needs to be removed by differential centrifugation. Adenovirus represents some of the largest viral vectors available for cardiac gene transfer. Most gene transfer to adult cardiac myocytes has been performed with secondand third-generation serotype 5 adenovirus (439, 965). Adenoviral vectors are extremely efficient at transducing the myocardium. This is true with a variety of injection methods and experimental animal models. After direct injection into the myocardium (Fig. 3, Table 2), 60 80% of the exposed cardiac cells can be transduced (170, 307, 439, 568, 858). Peak expression is generally seen within 3 days and is typically very strong. In addition to its use as an experimental tool, adenoviral delivery has been used to improve cardiac function in a number of studies and as such makes this vector a potential candidate for clinical applications (583, 801, 858, 968). Injection of the vector directly into the intravascular space results in poor transduction of the myocardium, and large amounts of Ad vector are acutely toxic, especially to the liver (720). However, when the cardiac circulation is isolated, either during transplantation studies or during heart-lung bypass, global transduction efficiencies of 30 50% can be achieved (83, 84, 187, 188, 801, 845). These techniques frequently make use of elevated hydrostatic pressure and permeabilizing adjuvants (187, 188). In vitro adenoviral transduction is even more efficient, and a preferred gene delivery vector for cultured cardiac myocytes will be discussed in subsequent sections. Reporter assays in which adenoviral delivery of chloramphenicol acetyltransferase (CAT), green fluorescent protein (GFP), or LacZ( -galactosidase) genes to isolated adult cardiac myocytes have shown a transduction efficiency approaching 100%, which occurs rapidly between days 1 2 in culture (132, 424, 439, 490, 765, 965). Despite its many advantages, Ad vectors have some serious limitations. Adenovirus elicits a potent cellular immune response (416, 991). Cells transduced with early generation Ad vectors are cleared rapidly from the body. This potent immune response is against both capsid and viral proteins that are encoded by the residual adenoviral DNA present in the vector. Gutted adenoviral vectors were developed to circumvent this problem as these vectors are devoid of most viral genes except for the ITRs and packaging signal (35, 172, 333, 807). Although these gutted-ad vectors demonstrate lower immunogenicity, they still trigger a cellular immune response presumably due to residual helper virus contamination or the vector capsid itself (172, 333, 334). However, even with this immune response, expression can be detected longer than 18 mo after administration. The various serotypes or chimeras of Ad vectors continue to be extremely useful cardiac gene transfer reagents, but their inherent immunogenicity/ toxicity will likely limit them to experimental/ therapeutic protocols that do not require long-term expression and can tolerate a potential immune response. 3. AAV FIG. 2. Adenoviral-mediated gene transfer to cardiac myocytes. Adenovirons (serotype 2 and 5) for gene transfer have a double-stranded DNA genome that is rendered replication incompetent by deleting several viral transcriptional elements required for replication (E1, E3 or E2, E4). The 36-kb adenoviral genome is packaged in an icosahedral protein capsid that contains a penton base and fiber knob at each vertex. The fiber knob of the adenovirus binds to coxsackievirus/adenovirus receptors (CAR) permitting entry into the cardiac myocyte by endocytosis. Once internalized, the virion DNA leaves the endosome and translocates to the nucleus where the double-stranded DNA can be transcribed and translated by the myocyte s own machinery into the recombinant protein of interest (circular inset). Recombinant AAVs are generating considerable interest in the field of cardiac gene transfer (294, 390, 668, 893, 997). AAV is a Dependovirus member of Parvoviridae (63, 821) with a particle size of 20 nm. As a Dependovirus, AAV is incapable of replicating in host cells under most physiological circumstances and requires coinfection with a helper virus for replication. Adenovirus or herpesvirus is the most frequently used helper virus, but others including human papillimo virus have also

6 1572 DAVIS ET AL. FIG. 3. In vivo viral vector delivery techniques. A: systemic delivery involves injection of the viral vectors into the venous space for delivery to the heart and throughout the circulation, including the coronary circulation. The vector (green) should gain access to most cardiac myocytes via the heart s capillary network. B: coronary delivery involves open-chest surgery in which the heart circulation (inflow, outflow, or both) is isolated with clamping (cross-clamp) or balloon angioplasty. The vector is introduced in this case under high pressure and allowed to dwell in the coronary circulation. C: direct injection involves injection of the vector by a syringe or similar device directly into the heart musculature. The vector gains access to the myocardial cells through the interstitial space and enters by vector-specific mechanisms and is more local in terms of transduction (shown in green). Downloaded from been effective helper viruses (63). The AAV capsid is a nonenveloped proteinaceous capsid made of three proteins termed VP1, VP2, and VP3, and the capsid seems to be devoid of most posttranslational modifications. AAV has a single-stranded DNA (ssdna) genome where both Watson and Crick strands appear to be packaged equally. The AAV genome is relatively small ( 4.5 kb, Table 1). Several promoters and alternate splicing control the expression of the capsid and Rep proteins. Rep proteins are critical elements involved in genome replication, integration, and packaging. The AAV genome also contains two ITRs, which are important for genome packaging, replication, and stability (47, 516, 836, 842, 984). Several distinct serotypes of AAV have been identified (Table 3), with serotype 2 being the most commonly used for gene transfer vectors. (For reviews on AAV, see Ref. 443.) AAV has several features that make it an appealing cardiac gene transfer vector. AAV vectors have a favor- by on June 29, 2017 TABLE 2. Recombinant viral vector delivery methods for the heart in vivo Delivery Method Advantages Disadvantages Vectors Direct injection Systemic Local vascular Restricted transduction, increased safety Global cardiac transduction, transduction of noncardiac tissue, ease of administration Global cardiac transduction, low chance of noncardiac tissue transduction Difficult to achieve global cardiac transduction, requires physical disruption of the heart, may require visualization/access to the heart (thoracotomy) Requires large concentration of vector, chance of increased toxicity, potential transduction of noncardiac tissue Can require toxic adjuvants, requires greater technical expertise AAV, adenovirus, gutted adenovirus, lentivirus AAV AAV, adenovirus, gutted adenovirus A summary of the three primary viral vector delivery methods reviewed in this article and their respective advantages/disadvantages as it pertains to myocardial gene transfer in vivo. Included are the common viral vectors used with each delivery method. AAV, adeno-associated virus.

7 CARDIAC MUSCLE GENE TRANSFER 1573 TABLE 3. Tissue tropism of identified adeno-associated virus serotypes Tissue AAV Serotype Reference Nos. Cardiac muscle AAV9, AAV8, AAV6, AAV1, AAV7, AAV5, AAV2 390, 431, 672, 893, 944 Skeletal muscle AAV1, AAV7, AAV6, AAV8, AAV9 62, 108, 254, 255, 982 Vascular endothelium AAV1, AAV2, AAV5 112, 298 Vascular smooth muscle AAV2 112, 645 CNS AAV5, AAV1, AAV4, AAV2 12, 91, 157, 248, 938 Midbrain AAV2 157 Ependyma / astrocytes AAV4 157, 501 Eye Retina AAV5, AAV4 512, 949, 979 Photoreceptors AAV5 949 Cochlear hair cells AAV3 502 Lung AAV9, AAV6, AAV5 75, 254, 321, 1000 Liver AAV9, AAV8, AAV6, AAV2, AAV1 254, 255, 627, 885 Pancreas AAV8, AAV1, AAV2, AAV6 114, 507, 936, 945 Kidney AAV2, AAV9 (neonatal) 75, 866 Summary of the transduction of tissues by various adeno-associated virus (AAV) serotypes. Several tissues can be transduced by multiple AAV serotypes albeit with varying efficiencies. Tissue transduction efficiency is not only dependent on AAV serotype but on delivery method, vector titers, and promoters. able safety profile relative to other vectors (Table 1, Refs. 21, 294, 522) as no AAV serotype or Dependovirus has been implicated as a causative agent in human disease (59). Early work suggested that AAV vectors may be able to circumvent an immune reaction, but recent studies have now demonstrated that AAV vectors can elicit both a humoral and cellular immune response in a variety of animal models (116, 294, 432, 522, 579, 604). This immune response is generally less robust than those triggered by other viral vectors like adenovirus, making AAV a preferred vector for in vivo gene transfer. While data from human clinical trials demonstrated a clear immune response against AAV-mediated gene transfer (432, 522, 604), the response was mild, and investigations into transient immune suppression to overcome this difficulty are underway. Integration of gene transfer vectors into the host genome is a safety concern that can be seen as both a benefit and a potential limitation. Integrating gene transfer vectors offer the possibility of permanent transduction avoiding the need for repeat gene delivery. Integrating vectors, aside from AAV, have occasionally caused human disease through insertional mutagenesis (309, 310, 800). For AAV vectors, this phenomenon must be interpreted in light of several facts. Classically, wild-type AAV integrates preferentially into two specific sites in the human genome, perhaps guided by the homology between the ITRs and genomic sequence. This integration is mediated by Rep proteins (785, 935). As the Rep proteins are deleted in AAV vectors, the rate of integration is much lower than that of the wild-type virus. Also, much of the site specificity of integration appears to be lost in recombinant AAV vectors (700). Recombinant AAV vectors typically have a low rate and random pattern of integration and appear to favor integration into transcriptionally active and open chromatin ( ). Even in the absence of integration, expression from AAV vectors can persist in striated musculature for many years (21). AAV vectors can be produced to high purity and titer (63). Early methods used adenovirus as a helper virus for AAV vector production, but adenoviral contamination was often found in these preparations. Current protocols now employ a helper-free plasmid cotransfection system to avoid an Ad contamination (299, 300, 983). Here a plasmid containing the AAV genome is transfected into cells, frequently HEK293 cells, along with a plasmid containing helper genes to function in trans. Vector purification and concentration are dependent on the specific capsid serotype. Capsids from some serotypes, such as serotypes 2 and 6, bind heparin avidly enough to be purified over a heparin sulfate-sepharose column (124, 321, 716). Vector production of all AAV serotypes can use serial centrifugation methods. Of note, the capsids of most serotypes can cross-package the ITRs from genomes of other virus serotypes (commonly serotype 2) in a process termed pseudotyping (63, 716). This phenomenon permits easy manipulation of the vector capsid, with altered production and tissue tropism characteristics (Table 3) without having to clone new vector genomes. In the last several years methods using baculovirus expression to produce AAV vectors have also been described (831, 910, 911). One of the main disadvantages of AAV vectors is their relatively limited cloning capacity (Table 1). The entire expression cassette of interest, including the open reading frame and transcriptional regulatory sequence, must not exceed 4.5 kb, allowing for 300 bp of ITR sequence (836). This cloning capacity severely limits the potential for gene transfer cassettes. Strategies to overcome this limitation include cotransduction with two vectors containing one-half of an expression cassette. The two genomes then rearrange in vivo to yield a larger, intact

8 1574 DAVIS ET AL. expression cassette (261, 262, 469). Also, many large genes have been modified into micro versions that still retain functionality, yet meet the size requirements for AAV vectors (329, 893, 997). Initial reports using this approach have been very encouraging, although a broad range of transduction efficiencies have been reported. Another concern is that the general human population has an antibody titer to AAV. This titer generally has a reasonable amount of cross-reactivity with several other serotypes (116, 322), raising the possibility that a neutralizing titer may limit AAV s therapeutic potential in vivo (522). AAV vectors have been widely used for cardiac gene transfer. Serotypes 1, 6, 7, 8, and 9 (Table 3) appear to be the most efficient for transduction, although many of the newer less described serotypes may also efficiently transduce myocardial tissue (293, 294, 672, 844). Depending on the exact capsid serotype and delivery method, AAV vectors have transduced nearly 100% of cardiac tissues (294, 893). Various studies utilizing intravenous delivery demonstrate global cardiac transduction, an exciting finding in terms of therapeutic strategies as well as experimental manipulations (294, 390, 893). More focused delivery to the heart by direct injection results in strong local expression of the gene product (997). Using AAV vectors, expression levels do not peak as fast as with adenovirus and generally take between 10 days and 2 mo to reach their peak. Once present, AAV expression can last for years (21, 63). AAV vectors are currently one of the most promising vectors for genetic manipulation of the heart in vivo. Studies using AAV for cardiac transduction have been very successful in delivering a variety of genes to several mammalian species including mice, rats, rabbits, dogs, pigs, and nonhuman primates (21, 422, 668, 672, 893). Additionally, several clinical trials have demonstrated clear human transduction with AAV vectors (432, 522, 604). Studies in heart kinetics and energetics, transplant rejection, and structural abnormalities have all been successfully performed using AAV vectors (113, 293, 778). These include several very promising therapeutic reports where a dramatic reduction of disease morbidity and mortality was demonstrated in several animal models of muscular dystrophy (293, 294, 893). The future use of AAV for clinical application will likely require further investigation of the immune response to AAV, development of smaller regulatory sequences for use in AAV vectors, and elucidation of the intracellular handling of the AAV capsid and genome, a complicated and poorly understood topic not covered in this review. Finally, other viral-based vectors such as Epstein- Barr, foamy, and simian virus 40 (SV40) (555, 843, 890) have also been reported for use in gene transfer studies but have not been included in this discussion for reasons of space, frequency of use, and a lack of a rich literature in cardiac transduction. This does not preclude these vectors, however, as viable options for cardiac gene transfer in future studies. B. Nonviral Vectors 1. Naked DNA Pure DNA carrying an expression cassette is perhaps one of the most basic transfer vectors. In this situation, clonal DNA that is generally derived from a bacterial plasmid (pdna) is introduced to the tissue of interest. Through means not completely understood, the cells take up the pdna, transport it to the nucleus, and express the exogenous gene. This type of striated musculature transduction has been known for approximately two decades (974, 975). pdna vectors have many potential advantages. pdna can be readily produced in large amounts to very high purity through a variety of methods available commercially and for the laboratory environment. Common laboratory methods for purifying pdna include alkaline lysis and phenol-chloroform extraction and DEA-dextran (diethylaminoethyl) binding columns. Huge fermentors are commercially available to produce and HPLC purify gram quantities of extremely pure plasmid under good manufacturing practice (GMP) conditions. Because there are no true infection particles with pdna, the resultant material can also be stored easily for long periods of time without loss of potency. Additionally, recombinant manipulation of the pdna is much easier. Unlike recombinant viral genomes, expression cassette size is much less of a concern. Although smaller genomes are generally propagated more efficiently, bacterial strains can easily maintain plasmids of 30 kb and above. This allows for the cloning of extremely large expression cassettes with larger portions of native regulatory sequence compared with viral vectors. This is advantageous in terms of having more natural transcriptional activity and possible tissue-specific expression patterns, a challenge with viral vectors such as AAV. pdna vectors are also attractive gene transfer reagents because they lack a viral proteinaceous or membranous component. This contributes to both the stability and small immune response associated with pdna vectors. Because organisms have evolved ways of neutralizing viral transduction events, virally delivered genes can elicit very potent humoral and cellular immune responses against the genetic vectors. Overall pdna does not tend to elicit a potent immune reaction, but bacterial DNA sequences are recognized by the body and can spark a small immune reaction. In a potentially related protective mechanism, a DNA sequence that is covalently attached to bacterial sequences can be silenced in vivo. In a series of experiments, delivery of circular DNA with excised

9 CARDIAC MUSCLE GENE TRANSFER 1575 bacterial sequence resulted in prolonged expression of the transgene. There are many potential benefits of using pdna vectors to transduce the myocardium, but at least two major obstacles remain: low efficiencies of transduction and persistence of transduction. Despite attempts with a variety of administration techniques, pdna vectors have very low cardiac transduction efficiency compared with their viral counterparts (975). Nonetheless, pdna vectors are still being explored for cardiac transduction due to their many advantageous qualities, and they are still considered useful experimental and therapeutic reagents for studies that do not require global and persistent expression. Continued investigation into pdna vectors may result in gains in both transduction efficiencies and persistence of expression. C. In Vivo Vector Delivery Techniques Delivery of foreign expression cassettes to the myocardium has a range of possible experimental and therapeutic applications. Regardless of the experimental or therapeutic objective, these genetic manipulations will likely rely on efficient transduction of the myocardium. Transduction efficiency requires the marriage of safe, efficient, and producible gene transfer vectors with a delivery system that is technically efficient, feasible, and well tolerated. The following two sections describe the current state of vector delivery technologies used for in vivo transduction of the myocardium. In general, the ideal gene delivery method should be technically simple, inexpensive, safe, and only transduce specified regions of the targeted tissue. At present, delivery techniques meeting all of these criteria do not exist. Current delivery methods can be classified into two broad categories: direct delivery to the myocardium and intravascular systemic delivery (Fig. 3, Table 2). As so many specific methods have been developed for vector delivery within each category, the following discussion will highlight the basic permutations of each delivery method. While this section is separated into vectors and delivery methods for the purpose of discussion, the pairing of vector and delivery system can have specific advantages and disadvantages as they pertain to transduction efficiency. For instance, intravascular delivery is often combined with AAV vectors (Table 2), while AAV is rarely if ever delivered by electroporation. 1. Direct myocardial injection Direct injection of the gene transfer vector to the myocardium is perhaps conceptually the most obvious approach. This method has been extensively used with most available gene transfer vectors and involves the introduction of an injectant directly into the heart musculature by a syringe or similar device (858) (Fig. 3). The vector gains direct access to the myocardial cells through the interstitial space and enters via the specific vector s entry mechanisms. This method has been utilized in studies addressing a variety of disease models including ischemic heart disease, heart failure, and muscular dystrophies. Additionally, several different animal models ranging from common laboratory animals (rodents and rabbits) to larger mammals (canines and pigs) are amenable to this type of manipulation. The exact method of injection varies by the experimental system and objectives. Overall, direct injection has been associated with excellent survival rates. Injections can be done blind by targeting the heart through the chest wall or trans-diaphragm from the abdominal cavity. While this method may seem rather unreliable, it has been used with effectiveness by a number of investigators, especially in small animals (858). The injection can also be done through the chest wall, under the guidance of ultrasound technology. Ultrasound guidance improves the accuracy of direct injection by allowing some visualization of the heart itself. Despite the technical simplicity of the direct injection method, it is difficult to reproducibly transduce the same area of the heart to similar magnitudes between animals and studies, even with ultrasound guidance. Direct injection can also be done following surgeries that expose the heart, permitting direct visualization for delivering injectant to the target tissue (858). This technique has been used both on beating hearts and after a transient cardiac arrest with similar effectiveness. Either ultrasound guidance or injection based on coronary vascular anatomy permits reasonably comprehensive transduction of the vascular walls of rodents (320). Although this necessitates a surgery and all of the associated complications, it heightens investigator confidence in the site of injection(s). With either blind or visualized direct injection technique, one major limitation is the targeted musculature must be physically large enough and easily accessible for the procedure. Thus, in most rodent studies, the sites of injection to nonseptal walls of the ventricles are limited. It is possible that in larger animals with thicker atrial walls the technique could also be applied to atrial tissue. Another limitation of this delivery route is the poor accessibility of the septal wall, trabeculae, and papillary muscles, but ultrasound guidance has been used to improve gene transfer to the septal wall by direct injection. Transduction efficiency not only depends on the method of direct injection but also appears to be highly dependent on the vector and vector dosage. The use of adenovirus and AAV vectors with direct injection has resulted in strong transduction of much of the ventricular tissue (858). In contrast, injection with retrovirus-based vectors, such as lentivirus, results in comparatively lower expression and amount of transduced cardiac tissue (71,

10 1576 DAVIS ET AL. 779, 780, 1009). The delivery of naked or complexed pdna with this technique results in patchy and lower intensity expression throughout the injected wall relative to the transduction efficiency of viral-based vectors. One permutation of the direct injection method is to inject the vector into the pericardial space rather than the myocardium itself. The injection of adenoviral vectors into the pericardial space resulted in moderate to weak expression in 30 40% of the rodent heart 6 wk after injection (247). As might be expected, the pattern of expression appeared to have a graded intensity with the strongest areas of expression neighboring the pericardial space. In this case, transduction was observed after including enzymes such as collagenase in the injectant to disrupt the extracellular matrix and to allow for more diffuse dissemination of the vector. The vast majority of studies have introduced the injectant with a syringe of some type; however, vectors have also been introduced using gene gun technology (537, 645, 907, 908). In this instance, the vector consists of gold particles complexed with pdna. Transduction involves surgically exposing the heart of the animal, commonly a rat, and bombarding the cardiac musculature with the gold particles under gas pressure with a device such as the Helios Gene Gun. Although this technique is not dependent on biologically active particles as when using virally derived vectors, it has several limitations. These studies have reported a reasonably high death rate from the gene gun procedure that seems related mostly to complications from the thoracotomy rather than the use of the gene gun itself. Heart expression levels with this technique are not robust (537, 645, 907, 908). The transduced cells are limited to a rather superficial layer of the myocardium, and expression within this layer is not strong. Given the nature of the technique, it is unlikely that many deeper cells could be transduced without killing the more superficial muscle. Presently, this approach is not significantly advantageous relative to direct injection with viral or nonviral vectors. Most injectants contain the vector and a physiological salt solution, but some studies have reported the addition of enzymes like proteases or hyaluronidase to degrade the extracellular matrix (ECM) (247, 462, 694, 695). The idea behind adding enzymes to the injectant is to facilitate vector diffusion through the tissue by degrading the ECM, thus yielding greater transduction efficiency. In fact, such techniques have improved the transduction efficiency of direct injections targeted for tumors and the pericardial space. Whether these compounds have a place in direct injection of the cardiac musculature is unclear as degradation of the cardiac ECM is fraught with potential complications and needs careful control. Adjuvants have a clearer role in intravascular delivery to the heart, which will be described shortly. Direct injection of vectors to the myocardium is a useful technique due to its relative simplicity and application to a variety of vectors, animal models, and experimental systems. It is, however, limited by transduction patterns and the requirement for directly visualizing the heart through surgery to achieve significant levels of transduction. For experimental and therapeutic manipulations that only require transduction of a limited amount of the myocardium (e.g., focal revascularization), direct injection remains an attractive technique. It is likely that any experiment or therapy requiring global transduction of the myocardium will be unable to use direct injection methods. 2. Intravascular delivery methods The vasculature offers an attractive portal for delivering gene transfer vectors as most cells, including those in the myocardium, lie in close proximity to capillary beds. Vectors placed in the venous space travel to the heart and throughout the circulation, including the coronary circulation (Fig. 3, Table 2). By using the heart s vast capillary network in this way, gene tranfer vectors gain access to a majority of the heart s myocytes. However, a number of imposing roadblocks complicate this scenario. The blood itself may contain neutralizing antibodies to the vector of choice, especially when they are viral-based. The blood also contains proteins, like albumin and platelets, which may absorb or inactivate the vector. There are also physical barriers to transduction including the endocardial cells and the capillary endothelium. Once out of the vascular lumen, the vector must also cross the ECMfilled interstitial space to transduce the target cell. The lung also tends to act as a sponge for many gene transfer vectors ( ), and any vectors placed in the venous space have to pass through the lung before reaching the left heart for ejection into the systemic circulation. Finally, there is the response of the host itself. Large intravascular amounts of a foreign material, especially from viral delivery, can lead to organ toxicity and/or an allergic reaction (720). Despite early difficulties, many groups have developed several vascular delivery methods that can result in high-efficiency transduction. Historically, direct injection of pdna and adenovirus vectors into the bloodstream has resulted in poor transduction of the myocardium, although some tissues such as the liver transduced well (383, 743, 884). With intravenous injection of pdna, transduction of the heart was sparse. Viral vectors, such as adenovirus, are not tolerated well when intravenously injected, and they have inefficient expression. These results likely reflect the difficulty in overcoming the physical barriers to transduction at levels of vector that are readily prepared and tolerated by the experimental animal. Recently this approach has been revisited using a high-pressure/high-volume tech-

11 CARDIAC MUSCLE GENE TRANSFER 1577 nique sometimes called hydroporation (359, 974, 1001). In this technique a viral or nonviral vector is delivered in a very high volume of injectant (approximately blood volume in some cases) resulting in high vascular pressures. Frequently, this is restricted to a specific limb by occlusion of the local vasculature but has been applied bodywide to rodents. Large volumes and pressures of injectant are thought to physically disrupt barriers formed by the endothelium and ECM, which in turn permit greater escape of the vector from the vascular lumen and wider dissemination throughout the tissue. This may result in a mild and temporary disruption of the plasma membranes themselves, which would further aid transduction. This high pressure/volume technique can result in moderatehigh transduction of the skeletal but not the cardiac musculature (292). Also, it is unclear how this technique would be tolerated by various animals and how easily it can be adapted to the heart. In light of the initial disappointing cardiac transduction obtained after the administration of vectors to the vascular space, many groups began developing intravascular delivery techniques aimed at overcoming these barriers to cardiac transduction. These techniques mainly focused on increasing local cardiac vector dose and/or dwell time and increasing the permeability of the cardiac microvasculature. For instance, percutaneous catheters have been used to deliver vectors directly to the heart (346, 373) as this technique should increase the local vector dosage in the heart, thereby increasing transduction efficiency. Catheters have been used to deliver a variety of vectors (pdna, adenovirus, and AAV) to several areas of the heart including the right atrium and in the root of the aorta just above the sinuses for access to the coronary arteries. A guiding modality such as ultrasound or fluoroscopy is often used (377). The technique has also proven effective in a range of animal models including rodents, canines, and sheep (84, 377, 658). Some studies have used balloon catheters to occlude the vascular outflow of the heart in an attempt to further increase dwell time and pressure of the injectant (346). Adjuvants such as nitroprusside, substance P, adenosine, and histamine have been added to the injectant to increase microvascular permeability to aid in vector extravasation (186, 188, 294, 504, 505). Overall, these techniques have resulted in impressive levels of myocardial transduction. Additionally techniques that use permeabilizing agents and viral vectors tend to be highly efficient. Recently, this technique has been combined with others such as sonoporation to increase the efficiency of complexed DNA transduction. This technique is mildly invasive, requiring vascular catheters that are associated with risks such as bleeding and infection. Some of the adjuvants used are toxic at high levels and can expose multiple tissues to the gene transfer vector, resulting in nonspecific transduction. Expertise in catheter manipulation and access to appropriate equipment are also required. Despite these drawbacks, percutaneous catheters represent an attractive and efficient means of cardiac vector gene delivery. A method with similar goals to the percutaneous catheters has been termed cardiac isolation or crossclamping (83, 84, 292, 318). This is a widely varied method with many individual techniques. In general, the animal is subjected to open-chest surgery where the heart circulation (inflow, outflow, or both) is isolated with clamping or balloon angioplasty. The animal is placed on heart-lung bypass with induced cardioplegia. Frequently, the blood is washed out and replaced with a buffer containing permeabilizing agents such as adenosine, histamine, and papaverine. The vector is then introduced under pressure and commonly allowed to have a dwell time of up to 15 min. Vectors are most frequently an adenovirus, though AAV and plasmid vectors have been used. The cardioplegia is reversed, and the animal is taken off heart-lung bypass (83, 84). In another variation, the heart is excised and manipulated in situ before transplant (662). This technique is comparable to the catheter-based method, it is highly invasive and has surgical risks, that can include infection or aortic dissection, commensurate with any open chest manipulation of the heart. Obviously, this technique also requires tremendous surgical expertise and an operating suite capable of supporting the procedures. One potential advantage of this technique is that it can limit exposure of the vector and potentially toxic adjuvants, such as papaverine, to the heart only (83, 84). Thus systemic toxicity and transduction of noncardiac tissues may be avoided. Recently, some promising results were reported with AAV vectors. AAV vectors pseudotyped with capsid proteins from some of the less commonly used serotypes, such as 1, 6, 8, and 9 among others, are known to transduce muscle cells much more efficiently than the commonly used serotype 2 capsid (Table 3). Several reports have been published demonstrating global cardiac transduction, with both marker and therapeutic genes, in mice after a single injection of AAV into the tail vein (Fig. 3) (262, 293, 294, 390, 893, 944, 1014). Surprisingly, tail vein injection does not require permeabilizing adjuvants, but at suboptimal vector doses the permeabilizing agent vascular endothelial growth factor (VEGF) has been advantageous (294). Potential drawbacks of tail vein injection include the requirement of high vector doses and the possibility of vector expression in noncardiac tissues. This technique is capable of transducing other tissues, such as skeletal muscle, although the development of cardiac specific expression cassettes may be able to overcome this possible problem (783). Nonetheless, cardiac restricted expression has been achieved (893). Despite the use of high AAV doses, tail vein injection was tolerated in mice and did not result in early morbidity, obvious toxicity, or immune responses directed at the vector cap-

12 1578 DAVIS ET AL. sid itself (294). At present, tail vein injection represents the only method available to transduce virtually every cardiac myocyte in small rodents, such as mice and hamsters. Future studies are needed to demonstrate if this delivery method can be efficiently translated to larger animals. One canine has been manipulated by systemic injection with a suboptimal vector dose (based on mouse data). Encouragingly, this canine demonstrated 60% cardiac transduction, similar to the mouse model, with very little expression in other tissues (61). Another limitation is that systemic delivery currently achieves high expression levels only with AAV vectors, which have a limited packaging capacity (Table 1). Similar experiments with other vectors, such as adenovirus, fail to transduce the myocardium to the same extent. These intravascular delivery methods represent the best current technology for achieving global cardiac transduction. D. In Vitro Gene Transfer/Acute Genetic Engineering A complementary approach to in vivo gene transfer and transgenic animal models is gene transfer to isolated cardiac myocytes in vitro (Figs. 2 and 4). Acute gene transfer in vitro offers a powerful experimental approach for understanding the direct effects of a genetic manipulation on cardiac myocyte structure and function. Traditional transfection methods have been partially successful in understanding the mechanisms of cardiac growth and differentiation in neonatal and fetal cardiac myocytes as these myocytes are easily cultured and are amenable to transfection by foreign DNA (306, 674, 676, 886, 900). Extrapolation of these data to the adult cardiac myocyte has been difficult due to several key differences between neonatal and adult cardiac myocytes that include features like morphological and sarcomeric organization, contractile and Ca 2 handling protein isoform expression, signaling molecule and transcription factor expression, and the dynamic versus quiescient nature of the culturing system (397, 624, 803). Compared with neonatal myocytes, acute genetic modification of adult cardiac myocytes is more challenging, because differentiated cardiac myocytes are more difficult to maintain in primary culture, and they are not amenable to traditional transfection techniques (DEAdextran, electroporation, Ca 2 phosphate, and lipofection; Refs. 424, 439, 765). Alternative methods, like direct injection of foreign DNA into the myocardium, are equally inefficient ( 0.02%) as the foreign DNA tends to localize to the injection site (93, 94, 494). Thus the aforementioned approaches are unsuitable for comprehensively manipulating cardiac gene expression in a controlled environment. The development and implementation of viral vectors and improved serum-free culturing methods were critical for overcoming these limitations. To date, the greatest success for in vitro gene transfer has come from recombinant adenoviral vectors, which will be one focus of this and subsequent sections. Adenoviral vectors have some distinct advantages over other viral vectors for genetically engineering postmitotic and terminally differentiated cell types like the adult cardiac myocyte. In contrast to other viral vectors, adenovirus can uniformly and efficiently transduce nonreplicating cells. Additionally, adenoviral DNA remains episomal, thereby eliminating any confounding influences of integration site and mutagenesis within the host cell s genome (11, 466). Importantly, adenovirus can be grown to high titers ( pfu/ml) and can transduce myocytes in vitro with little to no toxicity (11, 466). The first successful reports of adenoviral gene transfer to isolated adult cardiac myocytes were published in the early 1990s (424, 439). Isolated adult rodent cardiac myocytes cultured in serum-free conditions were adenovirally transduced with various reporter gene constructs driven by powerful, constitutively active promoters. In both studies nearly 100% of the adult cardiac myocytes were uniformly transduced, which is in stark contrast to the low transduction efficiency seen with pdna transfection to neonatal cells (424, 439). Kass-Eisler et al. (424) also performed direct injection of the adenoviral reporter gene construct into the myocardium and found a 5,000- fold increase in transduction efficiency relative to direct pdna injection. Viral-based reporter gene activity was first measured 4 h posttransduction (424), and there was a dosedependent increase in reporter activity which was assessed across time in culture (424, 439). These studies showed great promise for adenoviral gene transfer as a tool for understanding cardiac function, but questions still remained about transgene stability, the effects of adenovirus on myocyte morphology, contractile protein expression, myocyte contractile function, and the stability of cardiac myocytes in serum-free primary culture. Rust et al. (765) answered several of these questions by reporting culturing methods in which adult cardiac myocytes were stable and retained their differentiated state for 1 wk in serum-free culture conditions (765). In cardiac myocyte primary culture, adenovirus-mediated gene expression achieved nearly 100% transduction efficiency (765). Furthermore, adenoviral transduction did not affect the normal rod-shaped adult cardiac myocyte morphology, contractile protein isoform expression, or the isometric tension-pca relationship for the duration of the culturing period providing evidence that the adult cardiac myocyte phenotype is truly retained (765). 1. Cardiac myocyte culturing systems Isolating adult cardiac myocytes for study in shortterm primary culture has a reputation for being challeng-

13 CARDIAC MUSCLE GENE TRANSFER 1579 ing as they can be unstable in the presence of physiological extracellular Ca 2 and they readily dedifferentiate in the presence of serum (928). Nonetheless, cell culture experiments are important for uncovering the primary molecular and cellular effects of an experimental manipulation in a controlled environment. It is debatable whether there are any true immortal cardiac cell lines that retain both the genetic and protein profile, and morphological and contractile characteristics of bonafide adult cardiac myocytes. For instance, the SV40 large T transformed atrial cells (125) and ventricular tumor cells (770) possess some adult cardiac myocyte features but are largely inadequate both structurally and functionally compared with isolated adult cardiac myocytes. For this reason, many laboratories have become quite adept at cardiac myocyte isolation for primary culture and gene transfer (231, 317, 319, 347, 606, , 1012). Historically, the rat has been the preferred species for cardiac myocyte culture mainly because of availability and size (582). Isolation and culturing methods have now been expanded to include mouse and larger mammals such as rabbit, cat, dog, and human (109, 156, 160, 169, 355, 361). For contractile structure-function studies, it is imperative that cultured cardiac myocytes meet the following criteria: 1) be tolerant of physiologic Ca 2 ( mm), 2) retain a functional metabolic system, 3) create a homogeneous population of myocytes absent of contaminating and proliferating cells, 4) maintain their ultrastructure, cellular morphology and Ca 2 handling systems, and 5) remain quiescent and stable in culture. These criteria ensure that myocyte preparations are repeatable and yield a standard for healthy and physiologically relevant myocytes. Early reports of Ca 2 -tolerant isolated cardiac myocytes were published more than 30 years ago (397). There were two main cardiac myocyte culturing approaches, one requiring serum and the other termed the cell reattached method which used serum-free conditions. Serum-based cardiac myocyte culturing protocols were successful and could maintain cells for weeks and even months (125, 384, 396, 397), but the serum contained enough growth and miscellaneous factors that within days these cultured myocytes dedifferentiated and lost the adult myocyte phenotype. In contrast, serum-free culture conditions permitted adult cardiac myocytes to retain their highly differentiated rod-shaped morphology, myofilament and metabolic ultrastructure, and intact Ca 2 handling and transverse-tubule density (581, 693, 765). Many of the current protocols have been derived from the original work on serum-free rat myocyte culturing methods of Haworth et al. (343, 344) and Jacobson and Piper (397). Although isolation and culturing procedures differ slightly between laboratories and across species, they generally require certain fundamental elements described below. Once the heart is excised or tissue fragments obtained (as in human heart isolation) and cannulated, the hearts/myocardial tissue undergoes retrograde perfusion on a modified Langendorff apparatus (Fig. 4) such that a low Ca 2 enzymatic solution travels from the aorta through the coronaries for global exposure of the myocardium to the enzyme mixture. Most enzymatic solutions contain collagenase, hyaluronidase, protease, or a combination of these enzymes. Enzymatic perfusion is followed by gentle mechanical digestion and a slow titration of Ca 2 to bring the concentration back up to physiological levels. For some species, such as mouse or human hearts, Ca 2 titration is performed in the presence of an EC coupling inhibitor like 2,3-butanedione monoxime (BDM) as these myocytes tend to be ultrasensitive to extracellular Ca 2 after isolation. There are numerous publications documenting the successful isolation and use of adenoviral gene transfer to cardiac myocytes from failing and nonfailing rodents (158, 319, 355, 378, 564, 764, ), rabbits (109, 156, 355, 810, 878), canines (361), felines (166, 527), and humans (160, 168, 169). In all cases, adenoviral gene transfer is reported to be highly efficient and efficacious. Adenoviral gene transfer to myocytes isolated from a variety of species is complementary to making multiple transgenic lines in both rodent and larger mammals without the added cost, time, or larger mammal limitations to transgenesis. In addition, the availability of adenoviral gene transfer is tremendously important as intact Ca 2 handling is a critical component of both physiological and pathophysiological processes. Rodent EC coupling is quite different from that of larger mammals (rabbit, dog, and human; Ref. 49), which can impact the extrapolation of results from rodent models to the human. Acute genetic engineering has become a valuable experimental approach for elucidating the physiological role of normal and disease-related Ca 2 handling, myofilament, cytoskeletal, and signaling proteins in cardiac muscle. The following sections report the physiological insights gained from a rich and growing body of literature involving vector-mediated gene transfer to cardiac myocytes in vitro and in vivo. III. Ca 2 HANDLING PROTEINS Proper intracellular Ca 2 handling is essential for the normal beat-to-beat function of cardiac muscle. The cardiac myocyte is designed for highly orchestrated changes in cytosolic [Ca 2 ] during EC coupling (Fig. 5) in the heart (50). In the healthy heart, a transient increase in intracellular Ca 2 concentration ([Ca 2 ] i ) is the initial driving force for mechanical contraction, and Ca 2 removal initiates relaxation. The intracellular Ca 2 movement is tightly controlled by proteins associated with the sarcoplasmic reticulum (SR) and sarcolemmal membrane (Fig. 5). At the molecular level, EC coupling is a process

14 1580 DAVIS ET AL. whereby a small amount of Ca 2 enters through voltagegated dihydropyridine receptor (DHPR) or L-type Ca 2 channels to trigger large-scale Ca 2 release from the ryanodine receptor (RyR), the Ca 2 release channel located on the SR membrane. This process is known as Ca 2 - induced Ca 2 release (CICR, Fig. 5). The released Ca 2, which rises from 100 nm at diastole to 500 nm to 1 M during systole binds to troponin C and induces myofilament activation initiating cross-bridge cycling. The removal of Ca 2 from the cytoplasm during relaxation is carried out by the ATP-dependent Ca 2 pump, SERCA2a, and the sarcolemmal sodium-calcium exchanger (NCX) (50) (Fig. 5). Impaired EC coupling is a prominent feature of the diseased and failing heart. Inherited mutations in key Ca 2 handling proteins have been associated with cardiomyopathy (602). In some cases, altered Ca 2 cycling precedes the observed depression of mechanical performance. In acquired heart failure (HF), SR Ca 2 content is typically decreased due to the following HF-related maladaptations: upregulation of NCX, reduction in SERCA2a content/activity, decreased SERCA2a/PLN ratio, and increased RyR open probability which produces spontaneous Ca 2 leak (450, 517, 577, 681, 682). Decreased SR Ca 2 content contributes to reduced EC coupling gain and diminished Ca 2 transients, which negatively affect cardiac contractility. Reduced SERCA2a activity and insufficient energy supply impair Ca 2 reuptake resulting in Ca 2 transient remodeling. Furthermore, delayed Ca 2 recycling can result in elevated diastolic Ca 2 concentration that contributes to heightened arrhythmogenic potential. Additionally, elevated cytosolic Ca 2 concentration has been implicated as a stimulus for maladaptive growth and other morphological changes associated with HF (57, 337, 690). High-resolution biophysical and genetic approaches including transgenesis and acute gene transfer have been used to elucidate the mechanistic role of important Ca 2 cycling modulators in the heart under normal and diseased conditions (49, 55, 121, 140, 451, 517, 952). Acute genetic engineering provides a means for directly studying gene dosage effects. Titration of the transgene expression to obtain physiological versus pharmacological levels of expression is an important consideration when interpreting the functional outcomes of genetically manipulating Ca 2 handling proteins. Acute gene transfer studies are not limited to overexpression strategies as FIG. 4. Enzymatic isolation and acute gene transfer of adult cardiac myocytes in vitro. Hearts are excised and retrograde perfused with enzymatic solution containing a combination of collagenase, hyaluronidase, and/or protease on a modified Langendorff apparatus. Once adult cardiac myocytes are isolated, they are plated on laminin-coated coverslips and transduced in serum-free conditions with viral vectors for subsequent studies of myocyte structure and function.

15 CARDIAC MUSCLE GENE TRANSFER 1581 FIG. 5. Cardiac muscle excitation-contraction coupling. Illustration of the dynamic interplay between an action potential, intracellular Ca 2 fluxes, Ca 2 handling proteins, and the sarcomere in a process known as excitation-contraction coupling. A single action potential initiates extracellular Ca 2 entry through voltage-gated Ca 2 channels (DHPR). This Ca 2 entry triggers the release of a large amount of Ca 2 from the intracellular organelle, the sarcoplasmic reticulum (SR), through the ryanodine receptor (RyR). The rise in cytosolic Ca 2 reaches a [Ca 2 ] i sufficient for Ca 2 to bind to the myofilament regulatory protein, troponin, which initiates cross-bridge cycling and force generation (systole). Myocyte relaxation (diastole) requires the removal of cytosolic Ca 2 via two primary mechanisms: sequestration of Ca 2 into the SR by the Ca 2 -ATPase pump (SERCA2a, pink) or Ca 2 extrusion by the sodium-calcium exchanger (NCX) or sarcolemma Ca 2 -ATPase pump. SERCA2a activity is tightly regulated by the phospho-protein phospholamban (PLN). dominant negative mutants or delivery of knockdown molecules offer alternative approaches for understanding the direct role of Ca 2 handling proteins in adult cardiac myocyte function. This section focuses on acute genetic engineering approaches to modulate Ca 2 cycling and cardiac function. This will be accomplished by highlighting the following three components of EC coupling: 1) molecules involved in Ca 2 release from the SR, 2) molecules involved in Ca 2 sequestration, and 3) additional Ca 2 binding proteins that alter cardiac performance. A. Regulators of Ca 2 Release From the SR (610). The interactions between these proteins contribute to the CICR mechanism of EC coupling (Fig. 5). RyR structure-function has not yet been studied using acute genetic engineering techniques. The large open reading frame of RyR raises challenges for viral-mediated gene transfer strategies (Table 1). Additional challenges to studying the physiological role of RyR through gene transfer technology includes finding physiologically relevant doses of RyR subunit overexpression and the incorporation of exogenous RyR subunits into the SR membrane. Nonetheless, several laboratories have addressed the molecular mechanisms of CICR and RyR function through modulating the RyR s regulatory proteins (FKBP12.6, CSQ, junctin, and triadin) as reviewed in the following sections. RyR (RYR2 isoform) is the Ca 2 release channel in cardiac muscle SR. RyR is a homotetramer composed of four 565-kDa monomeric subunits (667). RyR is extensively regulated via several associated proteins including protein kinase A (PKA), anchoring protein (makap), protein phosphatases (PP1 and PP2A), sorcin, calmodulin, S100 proteins, and FKBP12.6 (Calstabin2) (531). RyR is also a substrate for posttranslational modification, and its functionality is altered by several small molecules including Ca 2, ATP, and Mg 2 (551). In the junctional SR, RyR interacts with calsequestrin (CSQ), triadin, and junctin 1. FKBP12.6/calstabin2 FKBP12.6 is a cis-trans isomerase protein that binds tightly to the cardiac ryanodine receptor (RYR2). It is a 12.6-kDa protein that is thought to bind to each subunit of the channel in a 1:1 ratio on the cytosolic side of RyR (478, 951). FKBP12.6 is considered a RyR channel stabilizer in the closed state during diastole, and it appears to faciltitate the functional coupling of RyR channels (534, 981). Elegant biophysical, transgenesis, and acute gene transfer strategies have yielded important insights into the physi-

16 1582 DAVIS ET AL. ological role of FKBP12.6 (478, 531, 952). Adenoviralmediated gene transfer of FKBP12.6 in both rat and rabbit cardiac myocytes revealed consistent gains in cardiac myocyte function. A five- to sixfold overexpression of FKBP12.6 in rat and rabbit myocytes directly reduced SR Ca 2 leak and Ca 2 spark frequency and amplitude (280, 513, 710), corroborating biophysical evidence for FKBP12.6 s role in stabilizing channel gating during diastole (534, 981). The reduction in SR Ca 2 leak in turn elevated SR Ca 2 load and increased myocyte fractional shortening (280, 710). Unique to the rat myocyte was the finding that FKBP12.6 overexpression hastened Ca 2 transient decay (280, 710). FKBP12.6 gene transfer and overexpression appears to synchronize RyR channel opening during systole, which in turn stabilizes the RyR and reduces random RyR openings and Ca 2 leak during diastole. It should be noted that the effects of RyR phosphorylation on FKBP12.6-RyR interactions and the subsequent functional outcome still requires further clarification as there is some discrepancy in the literature (376, 410, 534, 727, 837, 951). Nonetheless, FKBP12.6 represents an excellent candidate molecule for further gene transfer studies that could be used to explore the role of FKBP12.6 in heart failure, in posttranslational modification of the RyR, and as a potential therapeutic agent. A related FKBP isoform, FKBP12.0, is also expressed in the heart, and in some mammals the concentration of FKBP12.0 is higher than FKBP12.6 (409). Both FKBP12.0 and 12.6 have highly homologous structure and function, yet their affinities for the cardiac RyR (RYR2) differ significantly (409, 710). FKBP12.0 has a high affinity for the skeletal muscle RyR (RYR1) and can interact with RYR2, but FKBP12.6 has a much higher affinity and seemingly preferential interaction with the cardiac RYR2 (409, 710). Interestingly, knocking out FKBP12.0 in a transgenic mouse is lethal early in development due to severe congenital cardiac defects in the absence of skeletal muscle pathology (819). This mouse model implicates a key role for FKBP12.0 in the early functioning myocardium, but the physiological function of FKBP12.0 versus FKBP12.6 in the adult heart remains unclear. Recently, FKBP12.0 was acutely overexpressed in isolated adult rabbit cardiac myocytes by adenoviral gene transfer (809). A threefold FKBP12.0 overexpression increased SR Ca 2 content similar to FKBP12.6 (809). This acute gene transfer model also uncovered several functional differences between FKBP isoforms. Overexpression of FKBP12.0 (809) had opposing effects on Ca 2 spark amplitude, duration, and frequency relative to acutely transduced FKBP12.6 myocytes (513, 710). Additionally, FKBP12.0 reduced the sensitivity of the cardiac RyR to the CICR mechanism causing a decrease in EC coupling gain (809), a result not seen in FKBP12.6 transduced rabbit myocytes (513, 710). These acute gene transfer studies suggest that FKBP12.0 has distinct and possibly reciprocal effects on RyR function relative to FKBP12.6 in adult cardiac myocytes, although ultimately both FKBP proteins caused an increase in SR Ca 2 content. 2. Junctional SR proteins: CSQ, triadin, junctin, and histidine-rich Ca 2 -binding protein RyR has several binding partners located on the luminal side of the channel within the SR (Fig. 5). Calsequestrin (CSQ, 46 kda) is the most abundant SR Ca 2 binding protein. It binds 20 Ca 2 /molecule (cardiac isoform) with moderate affinity (K d M), and it actively participates in Ca 2 cycling by regulating the SR s luminal Ca 2 concentration (825). CSQ s binding partners triadin (isoform 1, 35 kda) and junctin ( 26 kda) act together to tether CSQ to the RyR complex, permitting a physical coupling of CSQ to RyR (1003). The interaction between these proteins is thought to play a vital role in CICR. In addition, histidine-rich Ca 2 -binding protein (HRC; 170 kda) is a moderate-affinity, high-capacity SR Ca 2 binding protein that interacts with SR luminal proteins (e.g., triadin) and is considered a secondary intra-sr Ca 2 storage source other than CSQ. A) CSQ. Physiological insights into CSQ s role in cardiac muscle function have been obtained from transgenic mouse (414, 792) and acute genetic manipulation studies (571, 877), but the focus here will be placed primarily on results obtained from acute gene transfer. Acute two- to fourfold overexpression of CSQ in adult rat and rabbit cardiac myocytes directly increased SR Ca 2 content as assessed by rapid caffeine application (571, 877). Rat cardiac myocytes acutely overexpressing CSQ had increased amplitude and duration of Ca 2 sparks and waves in the absence of changes in the frequency of these Ca 2 release events (458, 877), providing evidence that CSQ overexpression delays the closure of RyR. Imperatoxin A, a pharmacological activator of local RyR-mediated Ca 2 release events, was also used to assess the direct effects of CSQ overexpression on RyR refractory period (877). Results from this experiment demonstrated that imperatoxin A-induced Ca 2 spark frequency was reduced with CSQ overexpression in rat cardiac myocytes, suggesting that CSQ influences the RyR s Ca 2 -dependent refractory period through its buffering of luminal Ca 2 in the SR (877). Targeted knockdown of CSQ by adenoviral delivery of antisense CSQ to isolated adult rat cardiac myocytes caused the opposite effects of CSQ overexpression (458, 877) in which SR Ca 2 content and Ca 2 current (I Ca )-triggered Ca 2 transient amplitude were significantly reduced. CSQ knockdown also reduced the periodicity of Ca 2 sparks as well as increased the probability of Ca 2 wave propagation (458, 877). The combined approach of acute CSQ overexpression and knockdown provides evidence that CSQ plays a vital role in determining SR

17 CARDIAC MUSCLE GENE TRANSFER 1583 Ca 2 storage capacity and in modulating RyR function through its influence on SR luminal [Ca 2 ] (458, 877). The effects of acute CSQ overexpression on EC coupling appeared species dependent, as differences in CICR and Ca 2 wave and spark generation varied between rodents and larger mammals. In contrast to rat myocytes overexpressing CSQ, Ca 2 transient amplitude, when triggered by I Ca, was reduced in rabbit myocytes overexpressing CSQ (571, 877). Additional species-dependent differences were seen with Ca 2 spark measurements in which CSQ overexpression did not alter the amplitude, duration, or frequency of Ca 2 sparks in rabbit myocytes. These discrepancies on the direct effects of CSQ on EC coupling may be attributed to several factors including 1) species-dependent differences in I Ca (49), 2) gene dosage effects in which rat had 4-fold overexpression versus rabbit myocyte transduction which had 1.5-fold overexpression, or 3) the difference in CSQ backbone used for mutagenesis as the rat studies utilized the canine CSQ cdna while the rabbit studies used rabbit CSQ sequence. In comparing complementary genetic models of acute in vitro and chronic in vivo overexpression of CSQ, data from both models suggest that CSQ is a vital mediator of Ca 2 storage in the SR. Functional findings in transgenic mouse models (414, 792), however, differed from acute gene transfer studies as 10- to 20-fold overexpression of CSQ caused a reduction in SR Ca 2 release, which attenuated CICR mechanisms. These transgenic mice also showed signs of cardiac hypertrophy (414) and a transition to the fetal gene program (792). Interpreting the results from CSQ overexpression in transgenic mouse models compared with those of adenoviral mediated acute gene transfer is challenging as the transgenic mouse models had adaptive responses to the transgene that resulted in hypertrophy, altered cellular morphology, and changes in the expression of Ca 2 release and Ca 2 reuptake proteins. All of these alterations could contribute to the difference in functional outcomes measured in acute versus long-term gene transfer models. Additionally, the pharmacological levels of CSQ overexpression obtained in the transgenic models relative to the lower level of CSQ overexpression obtained with acute adenoviral gene transfer likely contribute to the differential findings and should be considered when performing a comparative analysis between models. CSQ dysregulation has been associated with arrhythmic disorders. Seven different allelic variants in the CSQ locus have been linked to inherited forms of catecholaminergic polymorphic ventricular tachycardia (CPVT), a disorder that is associated with stress-induced sudden cardiac death ( inherited arrhythmias database). To date, only two of the mutant CSQ alleles, D307H and R33Q, have been studied using acute genetic engineering (876, 925). Acute, fourfold overexpression of the CSQ mutant, D307H, in adult rat cardiac myocytes had a dominant negative effect to decrease SR Ca 2 content and I Ca -induced Ca 2 transient amplitude (925). These results demonstrated that the D307H mutant directly decreases the sensitivity of the SR to CICR mechanisms. Additionally, isoproterenol and escalations in pacing frequency induced extra nonrhythmic Ca 2 transients, a cellular mimetic of delayed afterdepolarizations (DAD) that are characteristic of CPVT. In contrast, threefold overexpression of the R33Q CSQ mutant in isolated rat cardiac myocytes did not affect SR Ca 2 content and had a dominant effect to increase Ca 2 transient amplitude in response to a triggering I Ca (876). The R33Q mutant also increased spontaneous Ca 2 spark and wave frequency, suggesting this mutant enhances Ca 2 leak and thus RyR activity (925). The physiological consequences of expressing D307H versus R33Q mutant CSQ were clearly different, yet both resulted in CPVT. In the case of the D307H mutant, the arrhythmic phenotype was ascribed to the inability of this mutant to bind Ca 2, thus causing a misregulation of Ca 2 release (876) similar to the results from acute CSQ knockdown (877). The R33Q CSQ mutant, however, had Ca 2 binding properties similar to wild type, but the heightened Ca 2 leak was attributed to altered CSQ-RyR interactions that caused overactivation of RyR (876). B) JUNCTIN. Junctin has been identified as an important tethering component in CSQ s interaction with RyR. To date, there is only one report of acute gene manipulation of junctin in the heart (260). Canine junctin was adenovirally delivered to isolated adult cardiac myocytes. An acute twofold overexpression of junctin directly decreased SR Ca 2 content as well as decreased the Ca 2 transient amplitude but did not alter cellular fractional shortening (260). Overexpression of junctin also increased myocyte contractility and accelerated relaxation kinetics (260). The mechanistic basis for this disconnect between contractile function and Ca 2 transient parameters in junctin overexpressing myocytes remains unclear. It is possible that junctin may affect additional aspects of EC coupling beyond that of Ca 2 release. Interestingly, complementary transgenic mouse models in which pharmacological doses of junctin were achieved also demonstrated reduced SR Ca 2 content, but the Ca 2 transient amplitude was preserved (369, 1002). The differences between acute and long-term genetic engineering in this case were likely due to the compensatory changes in Ca 2 handling proteins that were detected in the transgenic mouse models (369, 1002). Additionally, junctin overexpressing mouse models demonstrated remodeling of the SR and t-tubule system with 10-fold junctin overexpression (1002) and hypertrophy and histopathology with 30-fold overexpression (369).

18 1584 DAVIS ET AL. C) TRIADIN. There are three cardiac triadin isoforms, triadins 1, 2 and 3, with the dominant isoform being triadin 1 (304). Triadin and junctin are the products of different genes but have highly homologous sequences and structures that are postulated to play redundant roles in regulating Ca 2 release through RyR. The anchoring role of triadin, like junctin, has been fairly well defined, but to date, triadin s functional role in the Ca 2 release process has remained elusive. To address these issues, adenoviral-mediated acute overexpression of triadin in cultured adult rat ventricular myocytes revealed a direct ability of triadin to activate RyR and promote Ca 2 release (875). With threefold triadin overexpression, Ca 2 release from RyR was enhanced as Ca 2 spark frequency increased, but spark amplitude was lowered. Triadin overexpression also increased RyR open probability during single-channel recording and increased Ca 2 transient amplitude at smaller trigger I Ca (875). Consequently, SR Ca 2 content was decreased due to heightened spontaneous Ca 2 release at rest. Transduced myocytes were also arrhythmogenic when stimulated in the presence of isoproterenol (875). Triadin s direct activation of RyR may be through its interaction with CSQ and RyR, since a truncated triadin mutation lacking the domain important for CSQ interaction showed no effect on RyR Ca 2 release (875). An acute genetic approach showed consistent results supporting a direct regulatory role of triadin on RyR. The discrepancy between results from acute versus longterm triadin overexpression in transgenic mice (438) was postulated to involve compensatory adaptations by other key Ca 2 handling proteins, culture-related changes of myocyte structure and function, the level of overexpression, and/or location of these overexpressed triadin molecules (600). The transgenic mouse models have also shown that pharmacological doses of triadin can negatively affect EC coupling and lead to maladaptive cardiac hypertrophy (438). D) HRC. HRC is a luminal SR binding protein that has a histidine-rich repeat region located at the center of the molecule and is responsible for both Ca 2 binding and interactions with other junctional proteins. The Ca 2 binding ability of HRC and its interaction with junctional proteins suggest a potentially important physiological role of HRC in cardiac function. Adenoviral-mediated acute gene transfer of HRC demonstrated a direct and significant inhibitory effect on the Ca 2 transient and myocyte contractility when overexpressed at low levels of 1.7-fold (223). In this study, HRC overexpression slowed Ca 2 transient decay and as a consequence myocyte relaxation (223). Additionally, HRC overexpression increased SR Ca 2 load but blunted Ca 2 transient amplitude and fractional shortening, suggesting that HRC not only affects SR Ca 2 storage but also modulates Ca 2 release (223). Unexpectedly, acute HRC overexpression induced an upregulation of triadin and junctin in the absence of changes in other SR Ca 2 release (CSQ and RyR) and Ca 2 reuptake (SERCA2a, PLN) proteins (223). The increased protein levels of triadin and junctin contribute to the difficulty of understanding HRC s direct role in regulating both Ca 2 storage and release. It is likely that the combined effects of the changes in these luminal SR proteins are affecting myocyte physiology reported here. Interestingly, acute gene transfer with triadin and junctin, when expressed at levels similar to those measured in acutely engineered HRC myocytes, had opposite effects on SR Ca 2 load and Ca 2 release (260, 875). These findings further underscore the complex interactions between HRC, triadin, and junctin and their effects on myocyte physiology. While HRC affects SR Ca 2 content, similar to that demonstrated with the acute expression of CSQ (877), HRC has divergent effects on SR Ca 2 release, suggesting that CSQ and HRC are modulating RyR Ca 2 release via different mechanisms. To date, the direct effects of HRC still remain unresolved, possibly necessitating the use of other acute genetic engineering approaches including adenoviral-mediated HRC downregulation or dominant negative mutagenesis strategies. Interestingly, HRC null and overexpression transgenic mouse models have been generated and showed an increase in triadin expression with no change in SR load (297, 398). These findings confound our understanding of HRC s role in cardiac muscle function as both downregulation and overexpression of HRC produced similar effects on SR Ca 2 storage. Results from the HRC overexpression mouse model (297) show important differences compared with results from acute gene transfer (223). In contrast to the acute HRC overexpression in rat myocytes, transgenic mice had slow Ca 2 reuptake by SERCA2a and slow Ca 2 extrusion by NCX. This in turn slowed the Ca 2 transient decay rate but not cellular contractile function (297). Chronic HRC overexpression also caused the development of cardiac hypertrophy and an age-dependent transition to congestive heart failure (297). Remodeling of the heart was not apparent in transgenic lines with less than fourfold overexpression, suggesting that high levels of HRC are not tolerated at the organismal level (297). Perhaps gene dosage is partially responsible for the disparity between the complementary overexpression models as are the compensatory changes in other Ca 2 handling proteins including SERCA and NCX that occurred with chronic overexpression of HRC (297). B. Regulators of Cytoplasmic Ca 2 Removal 1. SERCA2A and phospholamban Cardiac cytosolic Ca 2 content is highly regulated via transport proteins that either sequester Ca 2 into the SR or extrude Ca 2 across the sarcolemma. The mechanisms

19 CARDIAC MUSCLE GENE TRANSFER 1585 for extruding Ca 2 from the cytosol in mammalian cardiac muscle include the following: SERCA2a, sarcolemmal NCX, mitochondrial Ca 2 uniporter, and sarcolemmal Ca 2 -ATPase (50). SERCA2a sequesters cytosolic Ca 2 into the SR by transporting two Ca 2 per molecule of hydrolyzed ATP against a steep Ca 2 gradient. There are five SERCA isoforms that are encoded by three genes, with SERCA2a being the predominant isoform expressed in the cardiac muscle (263). SERCA2a is regulated by the closely associated phosphoprotein, phospholamban (PLN), although sarcolipin, described below, can also contribute to SERCA regulation (55, 539). PLN, a small pentameric protein complex comprised of 6-kDa monomers (55), dynamically regulates SERCA2a function on a beat-to-beat basis by its phosphorylation state. It is the interplay between kinases and phosphatases that determines PLN s phosphorylation status (539). In the unphosphorylated state, PLN lowers the affinity of SERCA2a for Ca 2, thereby inhibiting Ca 2 transport (25, 451). Phosphorylation of PLN by PKA at Ser-16 or by CAMKII at Thr-17 reverses the PLN-mediated inhibition of SERCA2a (see also sect. VI) (539). PLN phosphorylation is an important contributor to the hastening of myocardial relaxation during -adrenergic stimulation as this PLN modification increases Ca 2 reuptake into the SR (950). PLN protein levels are generally unchanged in failing human hearts; however, a reduction in the extent of PKA-mediated Ser-16 phosphorylation of PLN and/or an increase in PLN to SERCA2a ratio are often noted in heart failure. Consequently, SERCA2a function is decreased in heart failure (517). Thus both SERCA2a and PLN are considered attractive gene therapy candidates for improving Ca 2 handling in heart failure patients. PLN phosphorylation status is contingent on the activity of PP1 (106, 539), the phosphatase involved in counteracting PKA-mediated phosphorylation of PLN. PP1 dephosphorylates PLN which in turn inhibits SERCA2a activity. An additional level of regulation is due to the actions of inhibitor 1 (I-1) protein (539). When activated by PKA, I-1 negatively controls the phosphatase activity of PP1 (539). The resulting I-1-dependent inhibition of PP1 maximizes the phosphorylation state of PLN, thereby increasing SERCA2a function. I-1 plays a vital role in the positive inotropic effects of -adrenergic stimulation as it assists in maximizing PKA activity in a cardiac myocyte (678, 850). The roles of PP1 and I-1 have been studied using acute gene transfer and are fully reviewed in section VI. Important insights into the physiological role of SERCA2a and PLN in cardiac muscle have come from transgenic mouse models and are reviewed elsewhere (517, 558). This section highlights experimental results obtained through the use of acute genetic engineering. Several of these gene transfer studies have been performed with the eventual goal of restoring SR Ca 2 uptake and SERCA2a activity (263, 337, 478) which become diminished in the failing heart. The reduction in Ca 2 reuptake in the failing heart is attributed to reduced expression of SERCA2a or a decreased pump activity, which can also be associated with an increase in the PLN/ SERCA2a ratio (18, 559). Acute genetic engineering strategies have been used to increase SERCA2a expression and/or activity to restore cardiac function in models of heart failure. Studies employing adenoviral gene transfer of SERCA2a have shown that overexpression of SERCA2a can significantly enhance Ca 2 release, hasten relaxation, and decrease diastolic [Ca 2 ] (167, 169, 268, 317, 319, 559, 583, 801). Acute overexpression of PLN increases the PLN/SERCA2a ratio, which functionally results in decreased Ca 2 transient amplitude, prolonged Ca 2 transient decay time, and increased diastolic [Ca 2 ], characteristics that are similar to myocytes from failing hearts (320). Overexpressing SERCA2a can rescue the failing myocyte phenotype that is created when the PLN/ SERCA2a ratio is increased (168, 169, 317, 319, 583, 801). Importantly, aberrant Ca 2 cycling and contractile deficits characteristic of failing human myocytes can be corrected by restoring the level of SERCA2a expression. This functional correction of the failing myocyte by SERCA2a gene transfer is manifest in increased shortening and relaxation velocity, heightened peak systolic Ca 2, and lower diastolic [Ca 2 ], in addition to a corrected cell shortening-frequency response (169). Hirsch et al. (361) demonstrated the effects of adenoviral-mediated overexpression of SERCA2a in cardiac myocytes isolated from a canine model of diastolic heart failure. This model was generated by a descending thoracic aortic coarctation resulting in left ventricular (LV) pressure overload over a year. Cardiac myocytes isolated from canines with diastolic dysfunction were transduced with SERCA2a Ad5 vectors. Acute expression of SERCA2a enhanced relaxation in the failing canine myocytes. In this model, SERCA2a overexpression unexpectedly resulted in a loss of isoproterenol-mediated inotropy during cardiac myocyte contractile measurements in vitro (361). The mechanism of this effect is unknown but is postulated to be due to a diminished PLN/SERCA2a ratio in SERCA2a transduced myocytes. Considering that cardiac reserve is critical to global cardiac performance, a full understanding of the effects of SERCA2a overexpression on -adrenergic molecular inotropy is important to ascertain. In vivo adenoviral gene transfer of SERCA2a has been performed in several animal models of heart failure. In a rat model of heart failure induced by aortic banding, Miyamoto et al. (583) used intracoronary gene delivery of SERCA2a at the time of transition from compensated hypertrophy to heart failure. SERCA2a expression restored systolic and diastolic function in this model (583). In a subsequent study, the same group demonstrated that

20 1586 DAVIS ET AL. in vivo SERCA2a gene transfer normalized SERCA2a expression levels in the failing myocardium which in turn improved cardiac function, energetics, and survivability (170). Appropriate titration of SERCA2a expression in failing cardiac muscle can restore the normal stoichiometry between PLN and SERCA, which prevents cytosolic Ca 2 overload and left ventricular dysfunction. Taken together, these studies implicate SERCA2a gene transfer as a potential treatment for contractile dysfunction in failing hearts. PLN gene knockdown and targeted mutant gene delivery have also been used successfully to modulate SERCA2a activity and consequently myocardial physiology (206, 227, 347, 1018). Acute expression of a dominant negative PLN construct, K3E/R14E, significantly increased fractional shortening and hastened Ca 2 transient decay and relaxation times in isolated rabbit ventricular myocytes (347). Gene transfer of PLN K3E/R14E to myocytes isolated from a rabbit HF model directly increased SR Ca 2 content, which in turn corrected the contractile dysfunction in the failing rabbit myocytes (1018). A V49A PLN mutant also acted as a dominant negative form of PLN to enhance myocyte contractility and relaxation (577). Both sets of dominant negative PLN variants identified critical sites within PLN that have dominant functional consequences over native PLN to reverse its normal inhibition of SERCA2a activity. An alternative gene transfer strategy to enhance SERCA2a activity involves using a constitutively phosphorylated PLN mimetic developed by substituting the serine residue at codon 16 with glutamic acid (PLN S16E) (372). Acute adenoviral delivery of this PLN phosphomimetic to neonatal rat cardiac myocytes increased contractility and had positive lusitropic effects in the absence of any -adrenergic stimulation (372), suggesting that the PLN S16E enhances Ca 2 cycling and thus contractile function relative to baseline. The efficacy of PLN S16E to halt heart failure progression was also examined using in vivo transcoronary delivery of recombinant AAV serotype 2 to BIO14.6 cardiomyopathic hamsters (372). Transcoronary delivery of an AAV2 reporter construct had 79% transduction efficiency, and the transgene was stable for at least 7 mo (372). In this model, PLN S16E increased SR Ca 2 cycling and slowed the loss of systolic and diastolic function characteristic of the cardiomyopathic hamster (372). Additionally, PLN S16E prevented thinning of the posterior LV wall (372). In a complementary rodent heart failure model, AAV in vivo delivery of PLN S16E blocked the transition to ventricular dilation observed in the nontransduced infarcted rodent model (394). Hemodynamically, PLN S16E improved LV contractility and diastolic function as well as lowered end diastolic pressures and prevented the transition to LV dilation in this HF rat model (394). Together, these studies highlight the potential for in vivo AAV-mediated delivery of PLN S16E as a therapy for progressive dilated cardiomyopathy (DCM) and associated heart failure. Several acute PLN gene knock-down strategies have also been used to alter SERCA2a function. Adenoviral gene transfer of antisense RNA directed towards PLN in neonatal adult rat (206, 347) and human failing cardiac myocytes (168) demonstrated decreased PLN expression and consequently increased rates of Ca 2 transient decay and enhanced myocyte contractility. In another knockdown approach, AAV-mediated gene transfer of a short hairpin RNA (shrna) to neonatal rat myocytes silenced PLN but did not significantly influence the levels of other Ca 2 cycling or cellular PKA targets (227). This silencing of PLN significantly increased basal SR Ca 2 uptake, but reduced PKA-directed Ca 2 uptake for up to 7 days after gene transfer. While delivery of antisense PLN significantly decreased PLN levels in neonatal rat myocytes, it failed to decrease PLN levels in adult cardiac myocytes (347), causing some disagreement about using an antisense PLN approach to knockdown PLN. In a separate study, the expression of PLN was significantly decreased after adenoviral delivery of antisense PLN to failing human myocytes (168). Viral-mediated gene delivery of an antibody targeted to the cytoplasmic portion of PLN has also been used to acutely alter SERCA2a activity (557). Expression of this antibody did not affect the expression or localization of PLN and affected the Ca 2 transient similar to phosphorylated PLN when expressed in adult mouse myocytes (557). In vivo gene delivery of this antibody also acutely improved basal contractility and relaxation in mice with diabetic cardiomyopathy. However, -adrenergic modulation of contractile function was muted in myocytes expressing this antibody, which may ultimately be detrimental for long-term survival. Although the increases in basal Ca 2 uptake observed with knockdown strategies are beneficial in terms of improved relaxation rates, the loss of adrenergic modulation may also be important for contractile function in failing hearts. In a subsequent study, silencing PLN by adenovirally delivering an antibody directed towards PLN s cytoplasmic domain restored contractile function and Ca 2 handling both at the myocyte and organ level in cardiomyopathic hamsters (178). Gene transfer of PLN antibodies represents a new approach for modifying PLN-SERCA interactions and provides a novel therapeutic strategy for improving failing cardiac myocyte contractile function. 2. Sarcolipin Sarcolipin (SLN) is a 31-amino acid SR membrane protein (23, 25, 660) with close structural similarity to PLN (660). The SLN and PLN sequences diverge significantly at the NH 2 and COOH termini (55). SLN does not contain the Ser-16 or Thr-17 phosphorylation sites but has

21 CARDIAC MUSCLE GENE TRANSFER 1587 the potential for phosphorylation at Thr-5 (55). The sequence variation in the COOH terminus may be responsible for the differential regulatory function of the proteins at high [Ca 2 ] (24, 921). SLN is predominantly expressed in the atria and fast skeletal muscle, and it can be found in the ventricle but at much lower concentrations than PLN (25, 55). Similar to PLN, SLN inhibits SERCA activity in the basal state, but the mechanism of reversal of SLNmediated inhibition by -adrenergic activation is unclear (23, 24). SLN expression is highly species dependent. In rodents, both SLN mrna and protein are expressed in the atria but are almost negligible in the ventricle. In contrast, in larger mammals including humans, SLN mrna is highly expressed in fast-twitch skeletal muscle compared with atria and ventricle (920). SLN expression is developmentally regulated and has a pathophysiological component as patients with atrial arrhythmias have low levels of SLN transcript (578, 920). Adenoviral (31) or transgenic (25, 30) overexpression of SLN has been performed to identify the role of SLN in cardiac physiology. Isolated adult rat cardiac myocytes acutely engineered with SLN had decreased myocyte shortening in the absence of any changes in Ca 2 transient amplitude relative to control myocytes (31). Acute SLN overexpression also slowed myocyte relaxation and Ca 2 transient decay time, suggesting that SLN depresses SERCA2a activity in ventricular myocytes (31). Immunofluorescence assays showed an overlap of vector-derived SLN with the native SERCA2a and PLN providing evidence of SLN colocalization with SERCA2a and PLN (31). This was further supported in PLN pull-down assays with SLN (31). In two different complementary transgenic mouse models (25, 30), chronic SLN overexpression in the ventricle caused a loss of function similar to the acute gene transfer model in which a reduction in rate of shortening and relaxation was observed at the organ (25) and myocyte (30) level. One SLN transgenic mouse model revealed diminished Ca 2 transient and myocyte shortening amplitude (30), which was ascribed to SLNdependent changes in SERCA2a s Ca 2 affinity. Stimulating -adrenergic pathways corrected the loss in contractility reported in SLN overexpressing mice (25, 30) and in PLN knockouts with SLN overexpression (289). These findings suggest that SLN inhibition of SERCA may be reversed by -adrenergic signaling, with the caveat that compensatory changes in SLN mouse models may also be contributing to the isoproterenoldependent enhancement of contractility and Ca 2 dynamics. These studies provide evidence that SLN is another potential therapeutic target, although additional work is required to identify the in vivo mechanism of SLN-mediated SERCA2a inhibition. 3. NCX The NCX transports 3 Na for 1 Ca 2 and is an important regulator of Ca 2 homeostasis and contractility in cardiac myocytes, especially in larger mammals. NCX can operate in both Ca 2 efflux (forward) and Ca 2 influx (reverse) modes, depending on the internal and external concentration of both Na and Ca 2 and membrane potential (53). During diastole, SERCA2a and NCX both contribute to cytosolic Ca 2 removal. The competition between SERCA2a and NCX for Ca 2 consequently determines the Ca 2 content of the SR per cardiac cycle (49). NCX has also been postulated as an important component in heart failure as several animal models of hypertrophy and heart failure have shown elevated NCX expression and/or activity (53, 824), a characteristic that has also been reported in tissue from failing human hearts (53, 824). There are, however, conflicting reports of heart failure-induced decreases or unchanged NCX content. Hassenfuss and Pieske (337) found that in failing human heart samples the ratio of NCX to SERCA protein expression consistently increased two- to fourfold relative to nonfailing hearts (824). Insights into NCX function under normal and diseased conditions have been gained through acute gene transfer studies. Adenoviral-mediated gene transfer of NCX to isolated cardiac myocytes has shown somewhat conflicting results. Several studies have demonstrated that gene transfer of NCX causes contractile dysfunction in rabbit ventricular myocytes due to a reduced SR Ca 2 load (719, 798, 799). Bölck et al. (69) reported that NCX gene transfer enhanced both systolic Ca 2 amplitude and myocyte fractional shortening in rat cardiac myocytes at low stimulation rates, but these functional parameters became reduced at higher pacing frequencies. In a recent study, Münch et al. (611) investigated the functional consequences of NCX overexpression using both in vitro and in vivo adenoviral gene transfer to nonfailing and failing rabbit hearts. With in vitro adenoviral gene transfer, NCX overexpression in isolated nonfailing cardiac myocytes depressed contractility at all pacing frequencies, while NCX overexpression in failing myocytes had an even greater frequency-independent reduction in contractility. In contrast, in vivo NCX overexpression in failing rabbit hearts improved contractility and contractile reserve as the hearts had increased responsiveness to -adrenergic stimulation and, therefore, enhanced inotropic and lusitropic responses (611). Long-term NCX overexpression in nonfailing rabbits in vivo had only minor effects on myocardial contractility and led to myocardial hypertrophy, in accordance with previous findings in transgenic mice (8, 736, 879, 992). The mixed results obtained from rat versus rabbit NCX gene transfer models might be attributed to the higher intracellular Na concentration in rodent cardiac myocytes, which would favor reverse-mode NCX

22 1588 DAVIS ET AL. action (51). Elevated intracellular Na concentrations have been found in human (692) and rabbit (177) failing hearts, in which case overexpression of NCX may be beneficial for heart failure by improving SR Ca 2 load; however, this benefit may be offset by heightened arrhythmogenicity. As NCX is the primary mechanism of Ca 2 efflux from cardiac myocytes, it has been postulated that loss of this extrusion mechanism should cause significant Ca 2 overload and thus contribute to heart failure (52, 54, 379, 948). Several laboratories have demonstrated that global NCX1 knockout in mice results in embryonic lethality (117, 452, 934). Interestingly, a recently engineered cardiac specific NCX1 knockout mouse survived well into adulthood with no signs of cardiomyopathy (354, , 737, 738). Pharmacological inhibitors of NCX function have been utilized, but the results are more difficult to interpret owing to drug specificity issues (739). To better understand NCX s role in cardiac myocyte physiology in the absence of complex influences from adaptive responses in transgenic mice or the pleiotropic effects of pharmacological inhibitors, transient knockdown strategies have been employed to acutely suppress NCX expression and activity. The initial reports of NCX knockdown used an antisense oligonucleotide strategy in which naked DNA was delivered to embryonic (863) or neonatal rat cardiac myocytes (497, 827). These studies showed that naked DNA delivery of antisense NCX is a viable strategy for knocking down NCX expression in neonatal (497, 827) but not adult rat cardiac myocytes (379, 859). To improve efficiency, Tadros et al. (859) used adenoviral delivery of the antisense NCX and achieved a 30 and 66% knockdown at 3 and 6 days after gene transfer, respectively. Under physiological conditions (37 C, 1.1 mm [Ca 2 ] o, 1 3 Hz pacing frequency), a 30% knockdown of NCX in adult rat cardiac myocytes resulted in heightened diastolic [Ca 2 ] i in the absence of action potential remodeling or changes in myocyte contractility/ca 2 transient amplitude during caffeine contracture. NCX knockdown tempered the contractile and Ca 2 transient response to varying extracellular [Ca 2 ] in cardiac myocytes, showing the importance of extracellular [Ca 2 ] in addition to NCX density on Ca 2 efflux. Higher levels of NCX knockdown have been achieved with adenoviral delivery of interference RNA (RNAi) to cultured neonatal rat myocytes (379). This strategy was highly efficient and elicited 90% knockdown of NCX expression. NCX knockdown in this system depressed the Ca 2 transient amplitude and relaxation rate in addition to action potential remodeling. Taken together, these studies have demonstrated that acute genetic engineering strategies can be used to knockdown NCX protein expression. These studies, however, have not entirely resolved the direct effects of NCX knockdown on adult cardiac myocyte function as only rodent models have been tested at this point. Considering that rodent Ca 2 extrusion relies heavily on SR reuptake whereas NCX plays a greater role in larger mammals (50), further study in larger animal models is required to fully understand the physiological and pathophysiological role of NCX. 4. Phospholemman Phospholemman (PLM) is a 72-amino acid transmembrane protein localized to the sarcolemma of cardiac and skeletal muscle. It is a member of the recently discovered FKYD family that regulates ion transport (854) and is a substrate for - and -adrenergic signaling molecules PKA and protein kinase C (PKC) (496, 709). PLM associates with Na -K -ATPase pump (NKA) and can regulate the pump s affinity for Na (143). Several studies have now demonstrated that PLM acts on NKA in a manner similar to phospholamban s regulation of SERCA2a, such that phosphorylation of PLM by PKA increases NKA pump activity through a mechanism of PLM-NKA disinhibition which concomitantly increases the pump s affinity for Na (176, 251, 822). Sodium homeostasis is critical under both normal cardiac function and during heart failure. There is evidence that PLM expression is decreased relative to NKA and that a higher percentage of PLM is phosphorylated in failing human and rabbit hearts (74). In contrast, PLM expression was elevated in a rat model of heart failure (115, 808, 1006). Together these studies suggested an important role for PLM in both healthy and diseased cardiac myocytes. Song et al. (832) virally transduced isolated adult cardiac myocytes with PLM which resulted in PLM overexpression ( 42% greater than than wild-type) within 3 days after gene transfer. Overexpression of PLM increased maximal myocyte shortening and Ca 2 transient amplitudes at low [Ca 2 ] o (0.6 mm). These parameters were blunted at higher [Ca 2 ] o (5.0 mm) and had minimal affects at physiological [Ca 2 ] (832). This same model was also used to assess the direct effects of phospholemman overexpression on SR Ca 2 content, NCX current, and action potential generation (1006). In this study, 3.5- fold overexpression of PLM at 3 days after gene transfer caused similar contractile changes at high and low [Ca 2 ] o as demonstrated earlier (832). These contractile abnormalities, however, could be corrected by coexpressing the PLM with NCX (1006). Here, SR Ca 2 content in PLM overexpressing myocytes was unchanged, but the Ca 2 decay rate was significantly slower, suggesting that NCX and PLM are functionally interacting. Additionally, the myocyte action potential and resting membrane potential were not significantly altered, but the reversemode NCX current was lower at higher clamped voltages. Taken together, these studies suggest that PLM directly affects cardiac myocyte function, perhaps through NCX. Recent evidence from in vitro pull-down assays

23 CARDIAC MUSCLE GENE TRANSFER 1589 and cotransfection of PLM and NCX into HEK cells suggests there is a direct physical interaction between these proteins (941). There are a few limitations to the acute gene transfer studies presented here that may be confounding our understanding of PLM s role in cardiac myocyte physiology. For example, PLM is a target for - and -adrenergic posttranslational modification which regulates NKA. In these published acute gene transfer studies, there is no control for phosphorylation status of PLM. Acute gene transfer of targeted amino acid substitutions to create phosphomimetics or permanently inactive PLM could yield more insight into the regulatory role of PLM in cardiac function. Furthermore, species differences in Ca 2 and Na homeostasis, particularly with respect to the role of NCX, may result in differential speciesdependent roles for PLM. C. Ca 2 Binding Proteins That Modulate Cardiac Performance 1. Parvalbumin Parvalbumin (Parv) is a cytosolic divalent cation buffer that is a member of the EF-hand Ca 2 binding protein family (199). Parv contains two high-affinity Ca 2 binding sites (K pvca M 1 ) that competitively bind Mg 2 but with much lower Mg 2 affinity (K pvmg M 1 ) (199). Parv is expressed in several tissues including fast skeletal muscle but is not normally found in the mammalian heart (199). In fast skeletal muscle, Parv is thought to act as a delayed Ca 2 buffer that hastens relaxation in a dose-dependent manner (352, 374). Although not naturally occurring in the heart, there is evidence of successful human Parv gene transfer in cardiac myocytes. The effects of ectopic Parv gene transfer on cardiac performance have been pursued both in vitro and in vivo. Parv gene transfer in isolated rodent cardiac myocytes caused a significant increase in Ca 2 sequestration rate and myocyte relaxation performance under basal conditions (931). The same acute Parv gene transfer procedure rescued diastolic function in myocytes obtained from hypothyroid rats, a model of diastolic heart failure (931). Diastolic dysfunction is also a clinical feature of hypertrophic cardiomyopathy (HCM). As such a cellular mimetic of HCM was developed by acutely expressing an HCM-linked mutant -tropomyosin in isolated cardiac myocytes. Notably, cotransduction of Parv corrected the slow relaxation, a characteristic of most HCM myocyte models, by increasing the rate of Ca 2 removal (139). In addition, acute Parv gene transfer successfully reversed the slow relaxation kinetics of senescent rodent cardiac myocytes (378). In vitro gene transfer of Parv or SERCA2a in myocytes from a canine model of diastolic heart failure showed comparative efficacy in hastening myocyte relaxation (361). In considering Parv as a potential therapeutic for the treatment of diastolic heart failure, a notable feature of Parv-transduced myocytes is their retained capacity for responding to -adrenergic stimulation, a result not obtained with SERCA2a overexpression (361). These studies indicate that Parv gene transfer may offer unique potential as a primary treatment for diastolic dysfunction in failing hearts. The effects of Parv in single myocytes have been translated to organ-level function. In vivo gene transfer of Parv via direct intramyocardial injection achieved significant Parv expression (858). The Parv expressing hearts showed accelerated relaxation rates as measured by a working heart preparation as well as by in vivo micromanometry and echocardiography (858). In vivo adenovirusmediated Parv gene transfer in aged rats showed a reduced expression efficiency (568); nonetheless, Parv still redressed the slow relaxation in these aged hearts (568). A model of Parv s role in cardiac myocyte performance has been proposed (140) (Fig. 6). Briefly, Parv binds Mg 2 during late diastole due to the high [Mg 2 ] relative to [Ca 2 ] (Fig. 6A). The voltage-dependent increase in cytosolic [Ca 2 ](1 M) triggers an exchange of bound Mg 2 to Ca 2 at Parv s metal binding sites (Fig. 6B). This process results in delayed Ca 2 binding by Parv through early diastole because the unbinding of Mg 2 is relatively slow. As intracellular [Ca 2 ] declines to near resting levels in mid to late diastole, Ca 2 dissociates from Parv (Fig. 6, C and D). Therefore, significant amounts of Ca 2 can be transiently buffered by Parv during early relaxation, which in turn accelerates relaxation in Parv-transduced cardiac myocytes relative to controls. A limitation of wild-type Parv is that high doses can lead to excessive Ca 2 buffering during systole, which negatively affects fractional shortening and Ca 2 transient amplitude (361). Experimental data as well as mathematical modeling have been carried out to determine the optimal concentration range of Parv for expression in cardiac myocytes. This range was set at mm for wild-type Parv (141). This critical concentration range limits the buffering of Ca 2 to the relaxation phase. Therefore, peak contraction and Ca 2 transient amplitude remain unaffected while diastolic relaxation and Ca 2 transient decay times are accelerated. Recent studies using Parv isoforms suggest that experimentally modified Parv constructs may be developed to limit the deleterious aspects of Parv function (i.e., depressing contractility) while maintaining enhanced relaxation performance (755). 2. Sorcin Sorcin, soluble resistance-related calcium-binding protein, is a highly conserved 22-kDa protein that has five EF-hand Ca 2 binding domains (519). Sorcin is expressed

24 1590 DAVIS ET AL. FIG. 6. Model of parvalbumin s effects on cardiac myocyte function. A: late diastole in which [Ca 2 ] i is low and parvalbumin (PARV) is binding Mg 2 with the myofilaments in a relaxed state. B: systole in which [Ca 2 ] i is rapidly rising. The heightened [Ca 2 ] i causes Mg 2 to slowly dissociate from PARV. During this time, Ca 2 is binding to the myofilament proteins, which initiates cross-bridge cycling and thus muscle contraction. C: late systole and early diastole in which PARV competes with other intracellular buffers for Ca 2.Ca 2 binds to PARV, which facilitates a faster decay of the Ca 2 transient and subsequently accelerates sarcomeric relaxation. D: mid to late diastole in which [Ca 2 ] i is low causing PARV to exchange bound Ca 2 for Mg 2, while Ca 2 is being sequestered into sarcoplasmic reticulum and cross-bridge cycling is sterically blocked. in several tissues, including the heart (561). Sorcin is localized to the t tubules and SR where it can interact with the RyR in cardiac myocytes (508, 561, 563). Increases in intracellular [Ca 2 ] initiate a Ca 2 -dependent movement of Sorcin from the cytosol to the SR (563), where it can interact with specific target proteins including RyR (561), L-type Ca 2 channel (DHPR) (562), and SERCA2a (536). In lipid bilayer preparations, sorcin directly alters the open probability of single RyR units inhibiting local Ca 2 release (225, 508). This effect is abrogated upon sorcin phosphorylation by PKA (508). Sorcin also plays a role in the diseased heart, as sorcin expression is altered in several heart failure models (536, 664) and mislocalized in spontaneously hypertensive rats (560). Both acute viral-based genetic engineering (242, 536, 810, 846) and mouse transgenesis (560) have been used to elucidate the physiological role of sorcin. Gene transfer of sorcin to isolated adult rat (242, 536) and mouse (846) cardiac myocytes caused a gain in contractility with a 3.5-fold increase in sorcin expression. In isolated rodent myocytes, sorcin overexpression increased fractional shortening, Ca 2 transient amplitude, and SR Ca 2 content (242, 536, 846). Sorcin s role as an acute positive inotrope was also confirmed at the organ level using in vivo adenoviral gene delivery to adult rodent hearts. In vivo, sorcin acutely hastened the rate of pressure development and relaxation and significantly enhanced systolic pressure (242, 846). In contrast, acute overexpression of sorcin in vitro (810) had the opposite effect on cellular function relative to in vivo gene transfer to the rodent myocardium. In rabbit myocytes, sorcin significantly blunted shortening and Ca 2 transient amplitudes (810). Additionally, sorcin overexpression in the rabbit model had reduced SR Ca 2 content and Ca 2 spark amplitude and duration, suggesting that sorcin has a negative effect on EC coupling in rabbit myocytes (810). The discrepancy between these acute gene transfer models could be related to the following factors: 1) the different levels of sorcin overexpression in rabbit versus rodent myocytes and 2) the enhanced contribution of NCX to Ca 2 handling in rabbit versus rodent myocytes (50). For instance, in rabbit myocytes overexpressing sorcin, there was an NCX-dependent slowing of Ca 2 decay time (810). In contrast, rodent myocytes that acutely overexpressed sorcin had enhanced Ca 2 transient decay time (536, 846) and SERCA2a activity (536), providing evidence of species-dependent differences in sorcin function. Notably, pharmacological levels of sorcin expression (20-fold increase) in a transgenic mouse model caused significant reductions in cardiac contractility and relaxation properties in the absence of cardiac hypertrophy or morphological remodeling (560), highlighting the importance of sorcin dosage and time of exposure to functional outcome at the organ level.

25 CARDIAC MUSCLE GENE TRANSFER 1591 The functional role of PKA-dependent posttranslational modification of sorcin is still unclear. Valdivia et al. (912) suggest that sorcin itself might be influenced by -adrenergic stimuli, since PKA phosphorylation of sorcin diminishes its inhibitory effect on RyRs (508). In contrast, Frank et al. (242) used -adrenergic agonists and blockade with acute sorcin overexpression in rat cardiac myocytes and found that sorcin-mediated gains in contractile function occurred independent of -adrenergic stimulation. These data suggest that unmodified sorcin may be regulating RyR function. Interestingly, PKA-mediated hyperphosphorylation of sorcin was identified in animal models of heart failure (536). PKA-mediated phosphorylation enhanced sorcin s Ca 2 sensitivity, which in turn may modulate the translocation of this protein to the SR during pathophysiological conditions in the heart. Structural and genetic studies have suggested causality between a naturally occurring sorcin mutant (F112L) and inherited hypertension and hypertrophic cardiomyopathy (552, 595, 912). Currently, the sorcin F112L mutant has not been studied using acute gene transfer, but a transgenic mouse model with 20-fold overexpression of this mutation has a dilated phenotype (131). Surprisingly, isolated cardiac myocytes from sorcin F122L mice have increased SR Ca 2 load, which enhanced both Ca 2 transient and shortening amplitudes (131). These results are opposite to the results obtained with similar levels of wild-type sorcin overexpression (560). The gene dosage effects of sorcin F112L on cardiac function remain unclear and require further experiments to determine the primary dosage effects of sorcin variants on cardiac myocyte physiology. 3. S100 proteins S100 proteins constitute a highly conserved 21 member family of EF-hand Ca 2 -binding proteins (351, 525). S100 proteins are small in size (10 12 kda) and have similar structural features that include a conventional COOH-terminal divalent cation binding domain and a lower affinity NH 2 -terminal unconventional EF-hand metal binding domain (525). The S100 isoforms differ in terms of their metal binding affinities, dimerization capacity, and posttranslational modifications (525). S100 isoforms are differentially expressed in a tissue-dependent manner (351, 525). In cardiac muscle, S100A1 is the predominant isoform, but S100A4, S100A6, and S100B have also been identified in the heart (315, 425). In general, the binding of Ca 2 to an S100 protein exposes a hydrophobic region of the molecule that is available to interact/modulate various intracellular target proteins (525). These Ca 2 -dependent protein-protein interactions are important for various molecular processes including contraction, cell cycle regulation, cell growth, and secretion. Several of the S100 proteins can also interact with target proteins independent of binding Ca 2 (525). Gene transfer studies performed in vitro and in vivo have contributed greatly to understanding the role of S100A1 protein on cardiac function in the healthy and diseased heart (607). Acute adenoviral-mediated gene transfer of S100A1 to healthy isolated rabbit (605) and rat (731) adult ventricular cardiac myocytes and engineered cardiac tissue (605, 732) caused a positive inotropic response in which cell shortening and Ca 2 transient amplitudes were significantly enhanced. A fivefold overexpression of S100A1 directly hastened the time to peak Ca 2 and Ca 2 transient decay velocity, which in turn increased myocyte shortening and relaxation kinetics (605). This gain in inotropy was attributed to heightened Ca 2 reuptake by SERCA2a (605). Although S100A1 is downregulated in heart failure (730), S100A1 knockout (194), and cardiac-specific S100A1 overexpressing (608) transgenic mouse hearts have normal morphology and tissue histology. Similar to the acute gene transfer studies, a fourfold overexpression of S100A1 resulted in increased contractility and Ca 2 transient amplitudes at both the cellular and organ level under basal and -adrenergic stimulated conditions despite the Ca 2 transient kinetics remaining unchanged (608). The gain in inotropy was ascribed to an increased SR Ca 2 load that occurred without detectable changes in the expression of several key Ca 2 handling proteins (608). The increased SR Ca 2 content could be due to several factors including S100A1 s interaction with RyR to decrease Ca 2 leak (927) and the enhanced Ca 2 uptake by SERCA2a when S100A1 is overexpressed (605). Target proteins of S100A1 have been reported to include titin (986), SERCA2a (435, 436), and RyR (435, 436, 608, 895, 927), which are all molecular elements that regulate intracellular Ca 2 homeostasis and/or cardiac myocyte contractility. Interestingly, S100A1 knockout mice appear to have normal cardiac function under basal conditions but have impaired cardiac contractility when challenged with hemodynamic stressors like -adrenergic stimulation (194). Infarcted S100A1 knockout mice also have a hastened transition to heart failure, low survivability, and significant loss in contractile function relative to infarcted wild-type mice and nontransgenic littermates (609). In contrast, S100A1 overexpressing mice have improved mortality, normal cardiac function, and protection from hypertrophic remodeling after infarct (609). Taken together, these complementary genetic models highlight a potential role for S100A1 in the compromised heart and in the contractile responsiveness to -adrenergic signaling. S100A1 is considered an attractive prospect for heart failure gene therapy because of its positive inotropic qualities under baseline conditions. As S100A1 protein levels are substantially reduced during heart failure, S100A1 s capabilities as a therapeutic transgene have been studied

26 1592 DAVIS ET AL. in the failing rodent myocardium both in vitro (606) and in vivo (606, 697, 698) using either adenoviral or adenoassociated (AAV, serotype 6) viral gene delivery. These studies utilized a cryothermic approach to induce myocardial damage leading to progressive heart failure over a 12-wk time span. In vitro adenoviral S100A1 delivery to isolated failing cardiac myocytes returned S100A1 expression to nonfailing myocyte levels and significantly enhanced failing myocyte contractility, Ca 2 transient amplitude, and SR Ca 2 content (606). S100A1 delivery to failing cardiac myocytes in vitro partially restored SERCA2a ATPase activity, decreased diastolic Ca 2 leak, and returned intracellular [Na ] back to nonfailing cardiac myocyte concentrations (606). After infarct, in vivo S100A1 delivery by adenovirus (606, 698) and AAV (697) had comparable rescuing effects on both organ level and myocyte contractility under baseline conditions and during -adrenergic stimulation despite having inhomogeneous expression across the myocardium. Additionally, increased S100A1 expression after infarct reversed changes in gene expression classically observed during heart failure (606, 697, 698). Collectively, these reports suggest that S100A1 gene transfer could correct aberrant Ca 2 cycling and contractility common to heart failure, thereby improving the performance of the diseased heart. S100B expression, which is normally undetectable in healthy myocardium, is induced in the diseased heart (419, 675, 902). Tsoporis and co-workers (901, 902) have shown in vitro and in vivo that the induction of S100B expression is a compensatory response to myocardial stress (901, 902). They demonstrated that the expression of S100B limited norepinephrine-induced cardiac hypertrophy structurally and genetically. It appears that S100B expression could be a component of the myocyte response to trophic stimulation that serves as a negativefeedback mechanism to limit cellular growth and associated alterations in gene expression. It is still unclear, however, whether S100B expression contributes directly to the pathogenesis of the failing heart. Unlike S100B, S100A6 protein (calcyclin) is also normally expressed in cardiac muscle as well as in a wide distribution of tissues (463). S100A6 expression in the heart is upregulated in rodent infarct models and in neonatal myocytes that are exposed to hypertrophic agents (901). The precise function of S100A6 in the adult heart, however, is unclear at present and could be readily studied using gene transfer strategies. IV. SARCOMERIC TARGETS AND TEMPLATES The sarcomere is the primary functional unit of striated muscle (Figs. 1 and 7). It is arranged as a hexagonal array of overlapping thick and thin myofilaments (575). The thin filament (Fig. 7) consists of 1- m-long filaments of polymerized actin monomers, along with troponin (Tn) Downloaded from by on June 29, 2017 FIG. 7. Three-dimensional schematic of the cardiac sarcomere. Thin filaments of polymerized actin monomers and regulatory proteins troponin and tropomyosin form a hexagonal lattice surrounding the myosin containing thick filament. Projections from the myosin rod contain the motor domain which cyclically interacts with actin in an ATP-dependent process of cross-bridge cycling. Inset is a schematic depicting the secondary structure of troponin subunits ctnc (Ca 2 binding subunit, yellow), ctnt (tropomyosin binding subunit, purple), and ctni (inhibitory subunit, red).

27 CARDIAC MUSCLE GENE TRANSFER 1593 and tropomyosin (Tm) (282, 888). The thick filament (Fig. 7) consists of the molecular motor myosin (MyHC), its regulatory and essential light chains, and other accessory proteins (i.e., titin and myosin binding protein C) (725). Cardiac muscle contraction occurs when chemical energy in the form of ATP is directly converted into mechanical work, such that actin and myosin interact to generate force and movement via cross-bridge cycling. The forcegenerating capacity of the sarcomere is ultimately responsible for myocardial performance. This process is regulated by the thin filament regulatory machinery, consisting of Tn and Tm, and the activating ligand Ca 2 (282, 888). The highly orchestrated events of Ca 2 binding to Tn during systole permits cross-bridge cycling and thus cardiac contraction (Fig. 5). As cytosolic Ca 2 is lowered, the thin filament regulatory system sterically blocks cross-bridge cycling, which results in cardiac relaxation (282, 888). Owing to the sarcomere s central role in cardiac performance, it has been an attractive target for genetic engineering by gene transfer technology. A. Protein Turnover and Stoichiometry The human heart faithfully generates 3 billion contractile cycles over an average life span. During this time, the heart must regularly replace old sarcomeric proteins with new ones (Fig. 8). It is extraordinary that mechanisms are in place to accomplish high-fidelity myofilament replacement while preserving sarcomeric stoichiometry and while maintaining full cardiac functionality. The sarcomere is a highly ordered three-dimensional lattice, and in the living adult cardiac myocyte this complex architecture has to be maintained as the myocyte contracts. In this context, sarcomere maintenance reflects a state of dynamic equilibrium in which newly synthesized myofilament proteins incorporate into the contractile apparatus as the old myofilament proteins are replaced (Fig. 8). Insights into the process of myofilament protein dynamics have been achieved using multiple approaches. One approach has been to study myofilament assembly during development (27, 203, 205). These seminal studies uncovered the mechanistic basis for the de novo synthesis of myofibrils or what has been termed myofibrillogenesis. With the use of cultured embryonic cardiac myocytes and high-resolution imaging techniques, the formation of nascent myofibrils could be temporally and spatially visualized. Several studies used isoform-specific antibodies (27, 180, 495, 741, 942), transfection (150), and/or microinjection of recombinant fluorophore-tagged myofilament proteins to observe the time course of nascent myofibril formation and the highly ordered assembly of new myofilament proteins in the developing sarcomere (788, 789, 929). These studies and others also documented the precise regulation of myofilament protein gene expression FIG. 8. Model of myofilament turnover and the principles of sarcomeric gene transfer. A: the sarcomere is a three-dimensional lattice with highly order architecture that is maintained as the myocyte contracts and generates force. Sarcomeric maintenance is modeled as a dynamic equilibrium in which old myofilaments turnover and are replaced by the precise incorporation of newly synthesized myofilament proteins into the contractile apparatus (represented by altered colors). This stoichiometric replacement occurs in the adult cardiac sarcomere within a short time frame. B: principles of adenoviral-mediated troponin gene transfer. Viral gene transfer strategies utilize a strong constitutively active promoter (e.g., CMV) to drive expression of recombinant myofilament protein, in this case troponin. The CMV promoter out-competes the endogenous myofilament protein message, so as troponin turns over in a 3-day time window the old native troponin (blue circles) is degraded and replaced with new vector-derived troponins (red circles) in the absence of changes in myofilament stoichiometry. Bottom panels illustrate that gene transfer of troponin/myofilaments results in a timedependent stoichiometric replacement of old troponin/myofilaments with new ones. during myofibrillogenesis (27, 203, 205, 495, 659, 741, 797, 914). Collectively, these studies demonstrated that during myofibrillogenesis sarcomere assembly is extremely dynamic as it involves orchestrated changes in myofilament gene expression, cyclical patterns of protein degradation and synthesis, and new protein incorporation. De novo synthesis of myofibrils and sarcomere assembly in embryonic striated muscle differs both temporally and spatially from the process of sarcomeric maintenance in the postmitotic adult cardiac myocyte. Sarcomeric maintenance involves the continual replacement of old or damaged sarcomeric proteins in the midst of the unceasing cardiac cycle. There are several reports documenting the turnover of myofilament proteins. Martin

28 1594 DAVIS ET AL. et al. (532) used radiolabeled leucine to estimate the turnover rates of myofilament proteins in the rodent heart in vivo. The half-lives of the myofilament proteins varied from 3 to 10 days, with troponin I (TnI) and troponin T (TnT) days, troponin C (TnC) 5.3 days, Tm 5.5 days, myosin heavy chain (MyHC) 5.5 days, actin 10 days, and the myosin light chains (MLC) 7 9 days (Fig. 8A). Studies using microinjection of recombinant epitope-tagged myofilament proteins in adult cardiac myocytes or permeabilized dedifferentiated muscle fibers soaked in labeled sarcomeric proteins further demonstrated that exogenous myofilament proteins can rapidly (on a minute time scale) incorporate and localize appropriately into the sarcomere (183, 388, 511, 788). More recent studies using gene transfer strategies that rely on the myocyte s native transcriptional/translational machinery have been highly instructive in ascertaining the mechanisms of sarcomeric maintenance in adult cardiac myocytes. Michele et al. (565) used adenoviral gene transfer of recombinant epitope-tagged tropomyosin ( Tm) and cardiac troponin I (ctni) into isolated adult rat cardiac myocytes and high-resolution confocal imaging to examine sarcomere assembly in a physiologically intact adult cardiac myocyte. They found that newly synthesized epitope-tagged thin filament proteins (ctni and Tm) stoichiometrically incorporate into the adult sarcomere in a time-dependent manner (565). The timedependent stoichiometric replacement revealed itself through the increase in epitope-tagged Tm and ctni with a concomitant decrease in the targeted native sarcomeric protein, leaving total thin filament protein expression unchanged as quantified by Western blot and confocal microscopy. This result was interpreted as the native old thin filament protein turning over and being replaced by the newly synthesized Tm and ctni (Fig. 8B). Notably, the rate of incorporation correlated to the previously determined biochemical halflives of the thin filament (532). This study also demonstrated that Tm has an ordered stoichiometric replacement, where newly synthesized Tms have targeted replacement at the pointed end of the thin filament (565). In contrast, TnI replacement was stochastic (565). These results, in conjunction with previous work (532, 533), suggested a model of Tm and Tn maintenance in which newly synthesized Tms and TnIs incorporate into the thin filament in an ordered (Tm) and stochastic (TnI) fashion, directly replacing the old Tms/Tns (Fig. 8B) (565). Transgenic mouse models corroborate this model of stoichiometric myofilament replacement as transgenic animals also maintain myofilament protein stoichiometry despite an overabundance of sarcomeric transgene transcripts (151, 258, 403, 567, 621, 735, 872, 874, 903, 999). The mechanism that governs ordered stoichiometric replacement in the sarcomere is not well understood; however, structural constraints and free energy barriers are likely influential variables. The mechanism of Tn turnover as a unit also remains in question. One hypothesis is that Tn stays bound to actin as newly synthesized subunits replace endogenous components. Alternatively, it is possible that the entire old Tn complex dissociates, reassembles with new Tn constituents (i.e., TnI, TnT, and TnC), and then reattaches to actin-tm. Studies of sarcomeric maintenance with acute gene transfer have been performed primarily in rodent models, but more recently the overexpression of sarcomeric genes and stoichiometric replacement in larger mammals including rabbit (355, 786) and human (160) has now been demonstrated. The conservation of myofilament protein stoichiometry in adult striated muscle highlights sarcomeric proteins as unique targets for genetic manipulation. Sarcomeric proteins manipulated by acute gene transfer include isoforms, mutants, and chimeric Tm, TnT, and TnI (564, 565, 764, ) as well as MyHC (355). A model system of replacement, as demonstrated by several cardiac myofilament proteins, is well suited for elucidating the fundamental contributions of the myofilaments to cardiac muscle physiology using genetic engineering strategies. With the advances in gene transfer technology, genetic manipulation of myofilament proteins has been used to study function both in vitro and in vivo. Two genetic approaches, transgenic animals and gene transfer to isolated adult cardiac myocytes, have served as the predominant methods for studying myofilament function in intact myocytes. Both model systems offer unique and complementary opportunities for advancement in understanding cardiac muscle function. Transgenic animal models have been critically important for understanding the role of sarcomeric proteins in myocardial function under normal and diseased conditions in vivo, and these models and functional outcomes have been extensively reviewed (151, 400, 744, 745, 999). The strategy governing adenoviral gene transfer of myofilament proteins to isolated adult cardiac myocytes involves a competition between endogenous and vector-derived gene expression. Gene transfer utilizes the host cell s own transcriptional/translational machinery to express and incorporate the engineered gene product into the sarcomere (Fig. 2). As the native myofilament proteins turnover, they are replaced one-toone by vector-derived myofilament proteins, resulting in a time-dependent and progressive increase in the percent replacement of native myofilament protein (Fig. 8B). For specific sarcomeric proteins, near full replacement can be achieved by 6 7 days in primary culture, particularly for those myofilament proteins with shorter half-lives.

29 CARDIAC MUSCLE GENE TRANSFER 1595 B. Thin Filament Proteins, Isoforms, Mutants, and Chimeras The use of acute genetic engineering is an attractive approach for studying the role of thin filament proteins on cardiac myocyte structure and function. Several thin filament proteins have been studied via targeted acute gene transfer technology. Mouse transgenesis and biochemical reconstitution protocols have also tremendously improved our understanding of thin filament regulation (278, 279, 871, 999), but this section will focus on key observations and insights gained from acute genetic engineering of the thin filament by gene transfer. 1. Cardiac TnI Cardiac TnI was the first myofilament protein targeted by acute gene transfer technology. The recent crystal structure determination in conjunction with NMR, biochemical, and functional studies has modeled ctni as a molecular switch within the sarcomere (Fig. 7) that regulates Ca 2 -dependent cross-bridge cycling (484, 556, 830, 867). During diastole, myocyte intracellular Ca 2 is low and ctni binds tightly to actin, inhibiting strong actinmyosin interactions. The elevation of intracellular Ca 2 initiates systole by weakening the ctni-actin interaction, causing a conformational change in ctni that promotes strong ctni-ctnc interaction and in turn permits active cross-bridge cycling. Cardiac TnI (ctni) plays a fundamental role in cardiac function under both physiological and pathophysiological conditions. ctni is also a prominent player in the heart s developmental transition from an embryonic stage, where the dominant TnI isoform expressed is slow skeletal troponin I (sstni), to the adult heart where the predominant isoform is ctni (729, 774). During physiological stress, ctni can also be posttranslationally modified in response to -adrenergic signaling and changes in ph (556). Pathophysiologically, ctni has been implicated as a dominant player in both acquired and inherited cardiomyopathies (278, 279, 638), and it has been identified as a site for proteolytic cleavage resulting in an acute loss of function known as myocardial stunning (617). 2. Gene transfer of TnI isoforms Westfall et al. (966) first determined the direct functional effects of two TnI isoforms expressed in the heart using adenoviral gene transfer of sstni into isolated adult cardiac myocytes. With near full replacement of native ctni by the sstni isoform, there was a concomitant gain in myofilament Ca 2 sensitivity of tension and altered cooperativity demonstrating that TnI isoforms directly influence sarcomeric function. Cardiac myocytes transduced with sstni also showed a gain in function during a bout of acute acidosis (ph 6.2) where myofilament Ca 2 sensitivity of tension was preserved in sstni myocytes relative to the marked drop in sensitivity sustained by control ctni myocytes. In confirmation, an sstni transgenic mouse model also demonstrated a gain in myofilament Ca 2 and ph sensitivity that resulted in delayed relaxation and Ca 2 transient decay rates in unloaded isolated myocytes (228). At the organ level, expression of the sstni isoform in the adult heart caused diastolic dysfunction (228). The TnI isoform specific differences in Ca 2 and ph sensitivity suggested there were unique domains within each TnI isoform influencing sarcomeric function. To establish the relative contributions of TnI isoform specific domains on myofilament function, chimeras of the ctni and sstni isoforms were engineered and expressed via adenoviral gene transfer in adult cardiac myocytes (956, 957). The N-slow/card-C chimera was developed by engineering 68 amino acids from the NH 2 terminus of sstni onto the 110 COOH-terminal domain of ctni. The other chimera, N-card/slow-C, was designed by joining the first 100 amino acids of ctni with the last 120 amino acids of sstni. Gene transfer of N-slow/card-C chimera desensitized the myofilaments to Ca 2 in tension-pca assays. In contrast, N-card/slow-C myocytes had a dramatic increase in myofilament Ca 2 sensitivity of tension. With these TnI variants expressed at near full replacement, the following myofilament Ca 2 sensitivity of tension hierarchy was established: N-slow/card-C ctni sstni N-card/slow-C. Chimeric TnI studies revealed the importance of TnI s COOH-terminal domain to myofilament function and unexpectedly showed an additive effect of ctni s unique NH 2 terminal (N-card/slow-C chimera) beyond that of sstni alone. Additionally, ctni s NH 2 -terminal domain proved to be critical to normal ctni function as the N-slow/card-C chimera caused a loss of function. Together, these results showed that the NH 2 -terminal domain of ctni is an additional and important contributor to myofilament Ca 2 sensitivity, a result that was not predicted in earlier biochemical studies (721, 896, 915). Complementary sstni and TnI chimera gene transfer approaches underscore TnI s prominent role in affecting both myofilament Ca 2 and ph sensitivity. Tension-pCa measurements at ph 6.2 showed a loss of Ca 2 -activated tension generation in N-slow/card-C myocytes similar to wild-type cardiac myocytes. In contrast, N-card/slow-C myocytes had a tempered response to acidosis similar to sstni myocytes. These data show that the COOH-terminal region of TnI contains the ph-sensitive domain that induces a TnI isoform-dependent change in myofilament Ca 2 sensitivity during acidosis. Gene transfer of sstni variants with targeted amino acid substitutions to adult cardiac myocytes identified a single amino acid that fully converted the ph sensitivity observed with sstni to the ctni phenotype (160, 959, 961). Acute gene transfer experiments showed that stoi-

30 1596 DAVIS ET AL. chiometric replacement of native ctni with a sstni variant (sstniqae: R125Q, H132A, V134E) converted sstni to the ctni phenotype during Ca 2 -activated tension measurements at neutral and acidic ph in permeabilized myocytes (959, 961, 964). A single codon substitution, sstnih132a, fully converted sstni to the ctni phenotype (160, 959, 961). In subsequent studies using biochemical exchange, Dargis et al. (155) reported results from the converse experiment where His was substituted for an Ala at codon 162 in human ctni sequence. This substitution caused a conversion of the ctni acidic phenotype to that of sstni in ATPase assays. This result was also demonstrated by targeted adenoviral gene transfer of the A164H rat ctni variant in intact adult cardiac myocytes (160). Not detected in solution biochemical studies (155) was the gain in Ca 2 -activated tension at neutral ph that was observed in A164H myocytes from both acute gene transfer and the A164H transgenic mouse model (160). These results highlight the importance of studying function through multiple approaches and in the context of the intact adult cardiac myocyte. Gene transfer of the ctni A164H variant was also shown to correct contractile dysfunction associated with isolated failing human myocytes, raising the prospect that this phenotype conversion may have protective effects in chronic heart failure. A164H transgenic mice showed that a single amino acid substitutions in ctni can dramatically alter function, as these mice achieved better myocardial performance with greater energetic economy during bouts of ischemia and postmyocardial infarction (160). Collectively, these studies show that a unique histidine residue in sstni is responsible for TnI isoform differences in myofilament Ca 2 and ph sensitivity. Complementary gene transfer approaches have aided in understanding isoform-specific functional outcomes, and they highlight the mechanistic influence of sarcomeric molecules on myocardial performance under physiological and pathophysiological conditions (964). Acute gene transfer with sstni, TnI chimeras, and the A164H variant of ctni demonstrate the functional importance of ctni s COOH-terminal region in determining cardiac myofilament Ca 2 sensitivity. The A164H transgenic mouse did not show histopathology or myocyte disarray at the organ level (160, 228). This is interesting given that other Ca 2 -sensitizing TnI molecules can cause disease, including inherited HCM. Several publications have linked HCM alleles to mutant sarcomeric gene products that act as dominant alleles by causing a gain of myofilament Ca 2 sensitivity (258, 278, 279, 403, 811, 871, 960). Mutations in the ctni gene, TNNI3, cosegregated with clinical cases of inherited cardiomyopathy. The HCM-linked ctni mutation R146G increased Ca 2 -activated tension generation in detergent-skinned papillary muscles from the R146G transgenic mice and rabbits (403, 786). Unlike A164H mice, the R146G transgenic animals developed overt histopathology and myocyte disarray (403, 786). Both sstni and R146G mutant ctni caused similar gains in myofilament Ca 2 sensitivity, but at the organ level these two TnI alleles had very different effects on cardiac morphology and life span. The mechanistic basis for these divergent outcomes, in lieu of similar gains in Ca 2 sensitivity, remains unknown. Gene transfer of R146G ctni to adult rat cardiac myocytes showed that this HCM linked mutant ctni behaves quite differently from sstni or A164H in the context of the intact myocyte. R146G mutant ctni had poor incorporation efficiency into the sarcomere relative to wild-type ctni, ctni flag, and sstni (960). R146G ctni also yielded a lower total replacement ( 45%), but it still caused a gain in myofilament Ca 2 sensitivity of tension at neutral ph. When the analogous mutation was engineered into the sstni backbone (R115G), Ca 2 -activated tension was similar to R146G ctni and sstni. Thus the R115G sstni mutation did not additively increase myofilament Ca 2 sensitivity beyond that of sstni alone, suggesting that the R146G and sstni may affect Ca 2 sensitivity via the same mechanism. Tension-pCa relationships were measured during acidosis (ph 6.2) and showed that, unlike sstni, the R146G mutant ctni and R115G sstni responded similarly without having a protective functional effect during acidosis. R115G sstni attenuated the protective ph effect of sstni, making it more similar to R146G ctni mutant. Collectively, these results suggest that a combination of increased Ca 2 sensitivity in conjunction with increased ph sensitivity may be responsible for the overt organ level histopathology associated with some HCM cases but not seen in other Ca 2 -sensitive models. The R146G ctni mutation is only one of a number of cardiomyopathy-linked mutations identified in ctni which now total over 20 (278, 279, 638). Recently, six new mutations in ctni were identified and linked to a highly malignant and clinically distinct disease entity, restrictive cardiomyopathy (RCM) (590). The six identified RCMlinked mutant TNNI3 alleles result in the following amino acid substitutions L145Q, R146W, A172T, K179E, D191G, and R193H, which occupy several of ctni s critical functional domains. Interestingly, mutations at ctni codon R146 can result in HCM (R146G or R146Q) or RCM (R146W). These shared domains suggest that the location of the mutation is important to the functional outcome, but they do not address the phenotypic divergence seen at the organ level. Biochemical reconstitution studies showed that RCM-linked mutations in ctni hypersensitize the myofilaments to Ca 2 relative to the already sensitized HCM counterparts (277, 445, 998). Acute gene transfer experiments demonstrated that RCM-linked mutant R193H ctni myocytes are hypersensitive to Ca 2 (158). Unexpectedly, R193H mutant ctni also disinhibited the thin filament at resting Ca 2 concentrations and caused an acute cellular remodeling from normal rod-shaped

31 CARDIAC MUSCLE GENE TRANSFER 1597 myocyte morphology to a short-squat phenotype. This acute cellular remodeling occurred with the progressive stoichiometric incorporation of RCM-linked mutant ctni into the cardiac sarcomere and could not be explained by altered Ca 2 sensitivity alone. The 193 codon in ctni is not only the site of an RCM-linked de novo point mutation but also a site of proteolytic cleavage during myocardial infarction (MI). MI can cause an acute but reversible disease state called myocardial stunning. The pathogenesis of myocardial stunning still remains elusive. It has been investigated with a transgenic mouse model (618) and biochemical reconstitution studies (239) but with conflicting results. The stunned transgenic mouse model had a desensitization of the myofilaments to Ca 2 -activated tension (618), while reconstitution studies found the opposite effect on ATPase activity (239). These divergent results highlight the importance of delineating the primary versus secondary effects of the stunned ctni variant for understanding the direct effects of truncated ctni on myocyte function. 3. TnI phosphorylation by PKA Studies using gene transfer of TnI isoforms and chimeras constructed from the fetal, slow skeletal isoform and the adult, cardiac isoform provided insight into the functional TnI domains that contribute to the decrease in myofilament Ca 2 sensitivity observed with PKA phosphorylation (967). PKA phosphorylates the Ser-23/24 cluster, which is located within the 32-amino acid extension of ctni, but gene transfer studies demonstrated that it does not phosphorylate sstni (967). In permeabilized myocytes, PKA significantly decreases myofilament Ca 2 sensitivity of tension (967). However, expression of sstni or a TnI chimera lacking the 32-amino acid ctni extension prevented this rightward shift in Ca 2 sensitivity. Phosphorylation of the TnI chimera with the 32-amino acid extension and the COOH terminus of sstni resulted in a shift that was comparable to ctni. Together, these results indicated that the functional domain responsible for the phosphorylation-dependent Ca 2 shift is located within the NH 2 terminus of ctni. This conclusion is further supported by a recent study that used both viral-based gene transfer and transgenic mouse models in which tandem serine codons (23/24) in ctni s NH 2 terminus were converted to aspartic acid residues to mimic PKA-mediated ctni phosphorylation. Myocytes from both genetic models had decreased myofilament Ca 2 sensitivity and significantly faster relaxation times relative to controls, while the addition of isoproterenol only minimally hastened relaxation beyond that of the ctni phosphomimetic (993). This study underscores ctni s key contribution to cardiac myocyte relaxation during -adrenergic stimulation. 4. Cardiac TnT As a component of the troponin complex, TnT holds both a structural and functional role in Ca 2 -mediated regulation of cross-bridge cycling (Fig. 7). The TnT molecule can be separated into two distinct functional domains: the NH 2 terminus, which is responsible for anchoring the entire troponin complex to tropomyosin and mediating the cooperative thin filament transitions, while the COOH terminus plays an important role in the Ca 2 - dependent interactions between TnI and TnC (282, 444). Cardiac TnT (ctnt) was the second thin filament regulatory protein to be manipulated by gene transfer technology. The impetus for studying ctnt with a gene transfer strategy was likely motivated by the percentage ( 15%) of reported inherited cardiomyopathy cases that have been attributed to mutations in the gene that encodes for ctnt, TnnT2 (946). Gene transfer strategies have been critical for understanding ctnt s role both physiologically and during the HCM pathogenic processes (528, 765, 855, 947). The first acute gene transfer study utilized Ca 2 phosphate transfection of embryonic quail skeletal myotubes to express recombinant wild-type ctnt and several HCM-linked mutant ctnts: I79N, R92Q, E160, and a truncated ctnt resulting from a premature splice donor site in intron 15 (855, 947). This model system had high transfection efficiency, but at least 10% of the myotube s sarcomeres were structurally dysfunctional, which could result from the dynamic nature of avian myotubes in culture. Functionally, all ctnt mutants decreased maximum Ca 2 -activated tension and tended to desensitize the myofilaments to Ca 2 in tension-pca assays (855, 947). Adenoviral gene transfer of these ctnt mutants into stable, intact, adult cardiac myocytes was also used to study HCM-linked mutations (528, 764, 765). Marian et al. (528) transduced feline cardiac myocytes with the human ctnt mutant R92Q and showed a dose-dependent loss of contractile function starting 48 h after gene transfer. This group also performed direct injection of these recombinant adenoviruses into adult rabbit hearts. Due to the wide range in transduction efficiency (2 60%), they were unable to show any myofibrillar morphology changes and thus did not examine function. Rust et al. (764) transduced isolated adult rat cardiac myocytes with recombinant adenoviral vectors harboring TnT mutants R92Q and I79N mutations. These HCM-linked mutations caused a significant desensitization of the myofilaments to Ca 2 in the absence of a concomitant change in maximum tension generation. Acute gene transfer results (764) are in apparent conflict with those obtained from quail myotube system (855, 947). Adult cardiac myocyte primary culture and avian embryonic skeletal myotubes are very different cellular systems in terms of the dynamic nature and sarco-

32 1598 DAVIS ET AL. meric components characteristic of skeletal myotubes. Despite conflicting results, these studies have shown that disease-linked mutant ctnts play a direct role in determining the myofilament s Ca 2 sensitivity and contractile function. Mutations in the gene that encodes for slow skeletal muscle TnT, TNNT1, have been linked to nemaline myopathy (276). The pathogenic process associated with nemaline myopathy (described below) is widely unknown and is another potential gene transfer target. Diseaselinked single point mutations in TnT can result in a loss or gain of sarcomeric function showing the relative importance of TnT to the regulation of cross-bridge cycling. Many TnT mutants have been studied using biochemical and reconstitution preparations, which goes beyond the scope of this review but are no less important in understanding TnT s role in thin filament function (276, 279, 871). 5. Cardiac TnC TnC is the Ca 2 sensor of the thin filament regulatory system (Fig. 7). The crystal structure of TnC in both the Ca 2 -bound and unbound state has been elucidated (224, 282, 888), and the atomic structure of Tn s core domain has recently been published (867). The crystal structure in addition to NMR and fluorescence resonance energy transfer (FRET) has identified TnC as a dumbbell-shaped protein that contains a central -helical core with a globular domain on either end. There are two isoforms of TnC, cardiac or slow TnC (ctnc) and fast skeletal TnC (stnc), which are the products of two different genes, TnnC1 and TnnC2. TnC is highly conserved across isoforms and vertebrate species ( 70% similarity). Cardiac/slow TnC (ctnc) is exclusively expressed in cardiac and slow skeletal muscles while stnc is expressed exclusively in fast skeletal muscles. Structurally, ctnc and stnc differ at the low-affinity NH 2 -terminal Ca 2 binding sites (I/II). Three divalent cation binding sites are dispersed throughout the globular domains of ctnc with one low-affinity binding site (II) in the NH 2 terminus and two high-affinity Ca 2 / Mg 2 binding sites (III/IV) in the COOH-terminal region. At rest, when cytosolic [Ca 2 ] is low, Mg 2 binds the high-affinity divalent cation sites III/IV until Mg 2 becomes displaced by Ca 2 during the systolic rise in intracellular Ca 2. It is the binding of Ca 2 to site II, however, that initiates the cascade of structural events permissive of cross-bridge cycling (282, 444, 888). Most of what is known about TnC isoform structure and function has come from solution biochemistry and replacement studies (224, 282, 888). In 1988, the first cloning and mutagenesis of avian stnc cdna was used by Reinach and Karlsson to examine structure function relationships in stnc (728). This opened the door for further mutagenesis studies in both stnc and ctnc from different species to elucidate the mechanism of Ca 2 -induced regulation of contraction (28, 29, 286, 483, 715, 815, 816). Currently, there are no published reports describing the use of acute genetic engineering for understanding TnC s role in cardiac physiology. Work by Edwards et al. (202a) demonstrated that adenoviral-mediated acute gene transfer of TnC variants to isolated adult cardiac myocytes caused direct changes in cardiac myocyte function. This study sets the stage for using gene transfer as an effective method for understanding TnC s physiological role in cardiac muscle. Mutations in ctnc have been linked to both HCM and DCM (56, 364, 591, 711, 802). Critical to understanding the pathogenic process of HCM and DCM is determining the primary effects of disease-linked mutant TnCs on cardiac muscle physiology, an aim that would benefit from acute genetic engineering technology. One potential limitation is the slower turnover rate of ctnc ( 5 days) versus ctni or ctnt ( 3 days) (532). This could constrain experimental outcomes due to limits in the amount of exogenous TnC replacement that can be achieved within the time frame of myocyte viability in serum-free primary culture. Nonetheless, elucidating the role of TnC in an intact myocyte and at the myocyte and whole organ level is still an under researched area that could be pursued through acute genetic engineering strategies. 6. Tropomyosin Tropomyosin (Tm) is an extended coiled-coil peptide that spans seven actin monomers in the thin filament (Fig. 7). Tm interacts with neighboring Tm molecules in a head to tail fashion, providing stability to overlapping tropomyosins and promoting binding to sarcomeric actin. Aside from binding actin, Tm is tethered to Tn via TnT. During diastole, when [Ca 2 ] is low, Tm blocks strong cross-bridge binding sites on actin. As cytosolic [Ca 2 ] increases during systole, Tm moves from a closed to an open position, exposing sites on actin for stereospecific myosin binding. Thus Tm can be modeled as a relay switch that promotes strong cross-bridge binding in response to the Ca 2 signal (282, 283). There are four isoforms of Tm (,,, and ), and Tm is the predominant adult cardiac isoform. There is a developmental transition from 20% Tm expression in the embryonic heart to less than 10% in a newborn, and it becomes almost negligible in the adult myocardium (622). Tropomyosin can exist as either a homo- or heterodimer in the adult heart (,, or but is mostly limited to homodimers). Tropomyosin has been the subject of a range of structure-function studies that utilized elegant biochemical methodologies (282, 283, 353, 360, 362, 550, 888, 889). A better understanding of Tm s role under both physiological and pathophysiological conditions in intact cardiac myocytes has been gained through the use of gene

33 CARDIAC MUSCLE GENE TRANSFER 1599 transfer technology. Over the past 10 years, several studies have successfully used acute genetic engineering to elucidate the direct effects of diseased linked mutant Tm on cardiac myocyte function. Mutations in Tm (TPM1) have been linked to inherited cardiomyopathy and nemaline myopathy (skeletal muscle Tm gene, TPM3). Michele et al. (564) used acute adenoviral gene transfer of HCM-linked Tm mutations D175N, E180G, K70T, and A63V to isolated adult rat cardiac myocytes to directly elucidate the structure-function effects of these mutant Tms. The incorporation of mutant Tm closely followed the turnover of native Tm, yielding 40% stoichiometric replacement. Isometric tension-pca assays revealed that E180G, K70T, and A63V increased myofilament Ca 2 sensitivity and had no effect on maximal tension generation. The D175N Tm mutant had no effect on Ca 2 -activated tension even when myocytes were exposed to higher viral titers. The heightened Ca 2 sensitivity of E180G and A63V Tm mutations slowed relaxation in unloaded intact myocytes (566). The D175N and E180G mutations are in a region thought to be important for Tm-TnT interactions (969), while the NH 2 terminus mutations K70T and A63V are in a region important for actin binding (360). Despite these distinct regional differences, three of the mutations had similar effects on myofilament Ca 2 sensitivity, suggesting that these mutations cause similar changes to Tm structure. K70T, A63V, and E180G are located in regions of the heptad repeat structure of Tm where charge is critical for maintaining salt bridge interactions that stabilize the Tm dimer heptad positions e and g (546, 841). It was postulated that the loss of charge caused by K70T and E180G would destabilize the coiled-coil Tm structure causing an increase in myofilament Ca 2 sensitivity. Biochemical reconstitution preparations in conjunction with circular dichroism analysis have shown a loss of thermal stability in Tm NH 2 terminus with K70T and A63V mutated Tm, confirming the loss of stability created by these mutations (353). The E180G and D175N mutant Tms have been expressed in transgenic mice (567, 623, 706, 707) and more recently in rats (955). The direct effects of the E180G Tm mutation in transgenic mice closely paralleled that of acute gene transfer, but the compensatory changes varied between E180G transgenic mouse models. Compensatory changes in Ca 2 handling proteins in myocytes from the E180G-FVB/N strain mouse model contributed to a marked change in diastolic function (623, 706, 707). Additionally, E180G Tm, when expressed on an FVB/N mouse genetic background, resulted in marked histopathology and hypertrophy closely mimicking HCM (706). In contrast, E180G Tm when expressed on C57BL/6 background did not show altered myocyte morphology or any histopathology (567). These reported compensatory adaptations in E180G transgenic mice underscore the important difference between acute and long-term genetic strategies and the contribution of background strain effects on organ-level phenotype. With transgenic models alone, it would be difficult to differentiate the direct effects of the E180G mutation in Tm as there were changes in myofilament Ca 2 sensitivity, Ca 2 handling function, and hypertrophic remodeling. However, coupling acute gene transfer with a transgenic mouse model pinpoints the primary versus adaptive changes that result from the transgene expression. The D175N transgenic mouse model was also engineered (FVB/N strain) producing several lines with varying replacement. To see any changes in working heart performance or myofilament Ca 2 sensitivity, assays had to be performed on lines with 40% replacement. In higher replacement lines, the D175N Tm mutation increased myofilament Ca 2 sensitivity of tension and slowed myocardium relaxation times (218, 623). Contrasting the D175N transgenic mouse results with those of acute gene transfer illustrates the importance of the amount of replacement achieved by the transgene. Acute gene transfer of D175N mutant Tm yielded 40% replacement with no change in myofilament Ca 2 sensitivity. The ceiling for percent replacement of native Tm with mutant Tm is set by the endogenous turnover of myofilament proteins and the time-dependent viability for myocytes in serum-free primary culture. Therefore, a transgenic mouse model in this case permits higher levels of replacement because it is not constrained by primary culture conditions. The different functional phenotypes reported for the D175N mouse model and acute gene transfer myocytes could be explained in part by the heightened percent Tm replacement reported in the D175N transgenic mouse. This underscores the importance of gene dosage effects on cardiac function, which is particularly relevant in inherited cardiomyopathy cases where replacement in human heterozygotes is estimated to be at or near 50% (811). Notably, the development of E180G and D175N rat transgenic models (955) yielded different results from those described previously in mouse (567, 623) or acute gene transfer studies (564). Extremely low levels of E180G or D175N Tm replacement had no effect on myofilament Ca 2 sensitivity or relaxation function in E180G rat myocytes, while D175N Tm desensitized the myofilaments to Ca 2 and accelerated relaxation rates (955). The divergent transgenic rat phenotype could be due to expressing a human Tm sequence in the rat (955), but Michele et al. (564) used the human Tm cdna as the backbone for mutagenesis in the previously described acute gene transfer studies which closely paralleled the work done in transgenic mouse models. Alternative explanations are the potential differential contributions of rat versus mouse genetic modifiers and/or low levels of replacement obtained in the transgenic rat.

34 1600 DAVIS ET AL. Mutations in Tm alleles are also causal for nemaline myopathy (NM). NM is considered a clinically heterogeneous group of congenital skeletal muscle myopathies with a histopathological marker of nemaline rod formation in skeletal muscle fibers. With few exceptions, the clinical NM phenotype is characterized by generalized muscle weakness and respiratory insufficiency with early morbidity (386, 387, 654, 656, 657). Identified NM mutations include the slow skeletal muscle gene TPM3, which is 90% homologous to the cardiac Tm gene TPM1, as well as slow skeletal TnT and -skeletal actin. It is notable that mutations in genes that are nearly identical in both sequence and protein structure could result in such divergent clinical presentation as seen in HCM versus NM. To understand the divergent pathogenic processes, Michele et al. (566) used adenoviral gene transfer of the NM-linked M9R Tm and HCM-linked A63V mutant Tm to isolated adult cardiac myocytes. The functional effects of the NM M9R mutation were indistinguishable from wild-type myocytes at core body temperature (37 C) (566). The M9R mutant s phenotype became apparent at lower temperatures (30 C) that have been reported in vivo temperatures for limb muscles. At these temperatures the M9R mutant myocytes had increased relaxation speeds relative to wild-type. This is opposite of the slowed relaxation measured in the A63V HCM mutant myocytes (566). Biochemical studies showed that the M9R mutant decreases Tm s affinity for actin with no significant changes in Ca 2 -activated ATPase activity relative to wild-type Tm (362). These results were obtained at room temperature, which may have affected the Ca 2 sensitivity of the myofilaments. In comparison, a transgenic mouse overexpressing the M9R mutant Tm in skeletal muscle resulted in histopathology and muscle weakness similar to the human NM phenotype (137). In situ measurements of force and power from the gastrocnemius muscle of M9R Tm transgenic mice were indistinguishable from wildtype skeletal muscles (161). Transgenic mouse models of NM are very difficult to use for extrapolation to the human condition as few murine muscles are considered purely slow twitch. Furthermore, these studies did not exclusively examine slow-twitch muscles like the soleus or measure mutant protein expression, which adds to the difficulty in extrapolating these results to the human disease condition. 7. Actin Actin is a globular protein that spontaneously polymerizes and forms F-actin, the backbone of the thin filament (Fig. 7). F-actin is visualized as a double-stranded helical filament and is a structurally polar molecule with a barbed end at the Z-line and one pointed end at the middle of the sarcomere (699). F-actin contains strong and weak myosin binding sites that are exposed when Tm slides from a blocked/closed to an open position in response to cyclical [Ca 2 ] i fluctuations (282, 283). The polar arrangement of actin is critical as it directs myosin heads to slide actin filaments in the appropriate direction, resulting in myocyte shortening (913). Much of what is known about actin s functional role in cardiac muscle physiology has come from various biophysical and biochemical techniques including X-ray diffraction, cryoelectron microscopy, and in vitro motility (367, 417, 457, 510, 546, 547, , 724, 814). The physiological role of actin isoforms and disease-linked actin mutants is a field in its infancy. Acute genetic engineering involving actin may be limited by its endogenous half-life of 10 days. Actin is the slowest myofilament protein to turnover, which is the biggest hurdle to overcome within the 1-wk window of viable myocytes in serum-free cell culture. This temporal constraint may not provide ample time for incorporation and a measurable functional phenotype. Additionally, it is not well known whether actin will replace and incorporate in the same manner that has been described for other thin filament proteins given the tight regulation of its length by capping proteins and tropomodulin, as well as its interaction with Z-line proteins. Further understanding of actin s role in cardiac muscle function will likely come from assessing the direct functional effects of diseased-linked mutations in the actin gene. Missense mutations in alpha-cardiac actin gene (ACTC, nine identified to date) and -skeletal muscle actin gene (ACTA1) have been linked to cases of DCM, HCM (589, 592, 665), and NM (386, 387, 654, 656, 657; described above), respectively. The clinical phenotype associated with ACTC mutations ranges from asymptomatic to severe cardiac dysfunction. Mutations identified in actin within Z-line anchoring domains (e.g., R312H and E361G) are associated with DCM. Other actin mutations within myosin binding domains (e.g., A331P and A295S) are associated with HCM. One actin HCM mutation (E99K) has been evaluated in vitro by a series of biochemical assays, which demonstrated a weakened interaction with myosin and reduced force production (73). Several other actin mutations have been expressed in vitro and found to demonstrate variable defects in protein folding which correlated with impaired incorporation into the cytoskeleton of mammalian (noncardiac) cell lines (919). To date, there are no published studies using acute gene transfer of cardiac or skeletal muscle actin variants. Unpulished results have shown that four representative cardiac actin HCM and DCM mutations can be expressed in isolated adult rat cardiac myocytes with efficient sarcomeric incorporation by gene transfer (177a), presenting proof of principle that acute gene transfer of actin is a feasible approach for assessing actin s role in cardiac muscle function.

35 CARDIAC MUSCLE GENE TRANSFER 1601 Myofilament turnover in the adult myocyte is another area that might benefit from the use of acute genetic engineering. Actin thin filament length is highly uniform in muscle sarcomeres ( 1 m) (750). Isolated myofibrils are very stable in vitro, and fluorescently labeled actin monomers can be incorporated only after removal of the capping proteins CapZ or tropomodulin (500). Despite these observations, in vivo thin filament maintenance is a dynamic process. In living cells, fluorescent actin rapidly incorporates into thin filaments, and monomer exchange is relatively quick compared with protein turnover, suggesting that thin filament capping does not restrict monomer addition (499). Injection of fluorescently labeled actin into thin filaments of living myocytes has been reported by several groups (183, 270, 336, 545, 817), but the site (pointed end, barbed end, or both) of actin incorporation remains debatable (499). 8. Capping proteins and molecular rulers Actin capping proteins such as CapZ and tropomodulin (Tmod) contribute to the actin monomer dynamics at the ends of the thin filament (499, 500) and regulate the length uniformity of actin filaments within the sarcomere (241). Specifically, CapZ appears to nucleate the developing actin filament at the barbed end within the Z-line during myocyte fibrillogenesis. The capping of actin by Tmod at the pointed end is a critical regulator of thin filament assembly, length regulation, and function (793). The importance of these capping proteins has been demonstrated by several transgenic knock-out models that had early morbidity (122, 246, 332), and with overexpression models that had a DCM phenotype (853). The giant molecular ruler protein nebulin is another important regulator of actin filament length albeit in skeletal muscle (371, 543). Its role in cardiac muscle is not well understood. Compared with skeletal muscle, in which there are approximately two nebulin molecules per thin filament, nebulin is expressed at significantly substoichiometric levels in cardiac muscle. Despite these low levels, knockdown of nebulin in neonatal rat cardiac myocytes using RNAi transfection caused dramatic elongation of thin filaments from the pointed end without affecting the barbed end (544), suggesting a potential role of nebulin in thin filament assembly. Given the complexity of actin dynamics and observed effects in transgenic mouse models, maintenance of thin filament length clearly requires precise regulation. To date, the precise role of actin capping proteins and molecular rulers in adult cardiac myofibrillar assembly and myocyte function are not fully resolved. Much of what is known about myofibrillar assembly comes from direct injection or plasmid transfection of highly dynamic cell culture systems (i.e., chick embryonic myocytes, neonatal myocytes, and serum-induced differentiating myocytes) (233, 240, 295). Concerns have been raised about microinjection techniques as they result in high levels of physical trauma to cultured myocytes (820). Additionally, most myofibrillar assembly studies have focused on morphology and histological analysis with very little analysis of the resulting functional effects of modulating capping proteins. Studying myofilament assembly and the functional role of capping proteins in the adult cardiac myocyte presents another opportunity for the use of acute genetic engineering. The major limitation of an acute genetic engineering approach will again rely on the turnover time of the capping proteins, which is thought to be fairly long (499). It is well known that the stoichiometry of capping proteins to actin is critically important (241), and studies suggest that, unlike other myofilament proteins, capping proteins can be overexpressed (852). At this time there has only been one report of the use of adenoviral gene transfer for studying the role of capping protein, Tmod, in embryonic cardiac myocytes from chicken (852). Adenoviral delivery of sense or antisense Tmod to this culture system was highly efficient, and significant overexpression/knockdown of the Tmod protein was seen by 24 h after gene transfer (852). In this setting, Tmod overexpression yielded shortened thin filaments, while Tmod downregulation caused longer thin filaments. This study showed that Tmod is important for maintaining actin stability and provides a basis for using genetic engineering as an approach for studying capping protein function. C. Thick Filament Proteins The thick filament is predominantly composed of the molecular motor protein myosin (Fig. 7) that cyclically interacts with actin to produce force and sarcomere shortening. Cardiac myosin is a hexamer consisting of two myosin heavy chains (200 kda each, discussed above), two essential light chains (light chain 1, MLC-1), and two regulatory light chains (light chain 2, MLC-2) (797). The thick filament also contains other myosin binding proteins such as C, H, X, and M proteins and the large elastic protein titin. Regulation of cardiac muscle contraction involves complex interactions between Ca 2 and proteins of the thin and thick filaments. While much is known about the contribution of Ca 2 and the thin filament to regulation of contractile function, much less is known about modulation by thick filament accessory proteins and their posttranslational modification. Acute gene transfer has not been extensively employed for studying thick filament proteins, but it offers an intriguing approach for directly assessing the thick filament s role in cardiac muscle physiology under normal and diseased conditions. Notably, the genes that encode for myosin heavy chain, myosin light chain, and the myosin binding

36 1602 DAVIS ET AL. proteins have identified mutant alleles that are causative for inherited cardiomyopathy. In addition, there are known changes in myosin and myosin light chain isoform expression during heart failure (26, 585, 794). Detailed transgenic and biophysical studies have been performed on thick filament mutant alleles (151, 258, 456, 723, 724, 905, 999); however, understanding the direct effects of these mutant proteins on intact cardiac muscle function is incomplete. A potential limitation of utilizing acute gene transfer for modulating the thick filament is the turnover rate of thick filament proteins. Some of the longest myofilament half-lives are related to constituents of the thick filament. The half-life of myosin heavy chain is estimated at 5.5 days, and the light chains are even longer ( 7 9 days) (533), which could limit the amount of replacement or dose-response assessment garnered through acute gene transfer in serum-free myocyte cultures. 1. Cardiac myosin Cardiac myosin is the most abundant protein in cardiac muscle and is the primary consumer of cellular energy (ATP). Two structurally and functionally distinct isoforms of myosin, - and -myosin heavy chain (MyHC), result from two separate genes (MYH6 and MYH7, respectively) that are differentially expressed in cardiac muscle. The transcription of each gene is independently controlled but coordinately regulated (313). Cardiac myosin isoforms can form homodimers or heterodimers and have been electrophoretically separated leading to the following nomenclature: myosin V 1 ( - homodimer), V 2 ( - heterodimer), and V 3 ( - homodimer) (366, 701). Despite sharing 90% amino acid homology, cardiac myosin isoforms are functionally quite distinct. For example, -MyHC, the slow molecular motor, hydrolyzes ATP three to seven times slower than -MyHC (330, 918). The fast motor, -MyHC, represents 90% of the total myosin expressed in the normal adult rodent heart (235, 356, 553). Expression of the slow -MyHC motor increases relative to -MyHC in rodent models of cardiovascular disease including diabetes (762), hypothyroidism, cardiac hypertrophy (553), and aging (99, 236, 932). Therefore, increased -MyHC motor expression is one biomarker of cardiac disease in rodents. The adult cardiac ventricles of healthy larger mammals have a different myosin isoform profile where 90% of the total myosin content consists of -MyHC (26, 284, 498, 585). In humans, mutations in each cardiac myosin gene can cause familial cardiomyopathy (100, 226, 811), and alterations in myosin isoform expression are associated with heart failure (514, 585, 632). The most commonly affected sarcomeric protein in inherited cardiomyopathy patients is -MyHC. In the heart, mutations in the -MyHC gene MYH7 have been associated with both HCM and DCM (638, 811). Although the human heart expresses predominantly -MyHC, mutations of the -MyHC gene MYH6 have recently been linked to either HCM or DCM (100). The clinical importance of -MyHC mutations is further underscored by the finding that -MyHC mutations have been linked to skeletal muscle myopathies including Laing distal myopathy (554), myosin storage myopathy (861), and hyaline body myopathy (68). The clinical importance of cardiac myosin makes it an attractive target for acute gene transfer strategies; however, only two studies to date have utilized such technology for studying cardiac myosin structure and function. The first HCM-causing myosin mutation was discovered in the motor domain of the myosin molecule (258). This missense mutation results in an arginine to glutamine substitution at amino acid position 403 (R403Q), which is located in the actin-binding interface (723). Since the discovery of this myosin mutation, much effort has been devoted to determining how it affects myosin motor function and triggers cardiomyopathy (60, 146, 147, 671, 856, 905, 987). Extensive studies performed on a transgenic mouse model of the R403Q mutation demonstrated delayed cardiac relaxation and slowed chamber filling with a progressive transition to hypertrophy and myocyte disarray common to HCM (258, 259). In these studies, the mutation was engineered in the -MyHC gene, which differs from the human HCM-linked mutation that occurs in the -MyHC gene (258). The primary effects of the R403Q substitution in -MyHC on myosin motor and cardiac function are therefore important to determine. A recombinant adenovirus has been used to study the structural effects of mutant -MyHC expression on sarcomere assembly and myofibrillar organization (527). Marian et al. (527) used acute gene transfer to study the effects of the HCM-associated R403Q mutation of human MYH7 on cardiac sarcomere structure. This study used acute gene transfer in adult feline cardiac myocytes in vitro, which had an efficiency 95%. In the adult feline cardiac myocytes, expression of mutant and normal -MyHC by viral gene transfer was determined by reverse transcription-pcr. Electron micrographs and fluorescent imaging only showed subtle alterations of sarcomere structure in all experimental groups. The R403Q mutant resulted in sarcomeric disarray in 50% of the myocytes that were examined, while no sarcomere disruption was observed in either control (no virus) or wild-type -MyHC transduced myocytes. This result was found in both electron micrographs and in experiments using indirect immunofluorescence and fluorescent imaging. The physiological consequence of long-term expression R403Q -MyHC mutation was also examined using a transgenic rabbit model (526). R403Q transgenic rabbits with 40% -MyHC replacement had significant septal and posterior wall hypertrophy with myocyte disarray and heightened collagen content without changes in fractional shortening

37 CARDIAC MUSCLE GENE TRANSFER 1603 (526). Additionally, R403Q rabbits had markedly reduced survival rates (526). These results suggest that the primary effect of this mutant myosin is to disrupt sarcomere assembly and myofilament organization and perhaps is a constituent in the pathogenic process leading to HCM. Most studies that have examined the effect of myosin isoform switching on cardiac muscle function have used hyper- or hypothyroid animals (235, 356) and transgenic mice (455, 873) or rabbits (401). It is difficult to determine the direct effect of myosin isoform shifts in these models because the phenotype represents the combined effects of myosin isoform and the compensatory response to the hormonal or genetic manipulation. Recently, a recombinant adenovirus was used to determine the functional consequence of increased relative -MyHC expression in rodent cardiac myocytes (355). -MyHC gene transfer demonstrated nearly 100% efficiency at the level of the isolated adult cardiac myocytes. Functionally, adenoviral gene transfer of -MyHC attenuated sarcomere shortening compared with control myocytes (AdGFP or no viral treatment) while Ca 2 transients were unaltered. These results suggest that -MyHC expression represents a Ca 2 -independent negative inotrope among the cardiac myofilament proteins and demonstrate the feasibility of using recombinant adenoviruses to study structure-function relationships of the heart s molecular motor. 2. The myosin essential (MLC-1/ELC) and regulatory (MLC-2/ RLC) light chains Ventricular muscle myosin is comprised of two unique accessory proteins, the essential (MLC-1) and regulatory (MLC-2) light chains. Both MLCs are bound to the neck region of myosin where they are thought to modulate myosin motor function. MLC-1 is a 17-kDa protein product of the MYL3 gene. It contains six functional domains, an actin binding site, a proline-rich region, and four helix-loop-helix regions and belongs to the superfamily of EF-hand Ca 2 binding proteins (899). The regulatory light chain, MLC-2, is a 22-kDa protein product of a different gene, MYL2. MLC-2 has several important structural regions including a phosphorylation site on serine-15 and four EF-hand domains (149). Like MLC-1, it also belongs to the superfamily of EF-hand Ca 2 binding proteins. Mutations in the myosin light chains have been linked to cardiomyopathy (226, 638). Collectively, these data suggest that both light chains play a critical role in normal cardiac physiology, particularly in cross-bridge kinetics. To date, there are no reports of utilizing acute genetic engineering for understanding the role of myosin light chains in cardiac function. However, three different transgenic mouse models that overexpressed either the ventricular isoform of RLC (RLC2 v ) or ELC (ELC1 v )inthe atria or the atrial isoform of ELC (ELC1 a ) in the ventricle showed that protein stoichiometry is tightly conserved despite overexpressing light-chain transcripts (230, 402, 670). Furthermore, multiple transgenic lines have been generated with varying levels of replacement that ranged from 0 to 95%. Taken together, these data show that like other myofilament proteins, MLCs have tightly regulated stoichiometric replacement and offer titratable gene dosing. Permeabilized fiber studies from these mice showed a change in cardiac fiber function such that the incorporation of either ELC1 v or RLC2 v into atrial fibers caused decreased unloaded shortening velocity (787). Interestingly, myosin light-chain isoform switching, where the ventricular isoforms convert to the atrial light chains, has been documented in various human heart failure cases such as in congestive heart failure or in DCM (26, 601, 794), suggesting important and divergent physiological roles in cardiac muscle for the ventricular versus atrial myosin light chains. The precise role of the different cardiac muscle MLC isoforms in cardiac function is not fully known. A potential limitation to MLC gene transfer is the long turnover time of the light chains. The half-lives of the MLCs have been estimated at 7 9 days, which may limit the amount of replacement one can achieve with recombinant light chains delivered to myocytes in primary culture (532, 533). Modest levels of replacement ( 30%) in the transgenic mouse model with ECL1a overexpression in the ventricle resulted in a change in function (230), suggesting that the high levels of replacement may not be necessary to affect cardiac myocyte physiology. 3. Thick filament accessory proteins: C-protein (MyBP-C) Myosin binding protein C (MyBP-C) is a myofilament protein of the intracellular immunoglobulin/ fibronectin superfamily that is located in the C-zone of the A-band in seven to nine transverse stripes spaced 43 nm apart (142, 661, 757). Distinct genes encode for the three isoforms of MyBP-C that are expressed in adult mammalian striated muscle and include fast skeletal, slow skeletal, and cardiac MyBP-C. MYBP3 is the gene that encodes for the cardiac isoform (cmybp-c) (257), and it is expressed exclusively in cardiac muscle throughout development (256). Cardiac MyBP-C is organized similarly to the skeletal isoforms with 10 globular domains (C 1 -C 10 ) and a highly conserved linker region, the MyBP-C motif (257). Cardiac MyBP-C has several unique features including an extra IgI-like NH 2 -terminal domain (termed C 0 ), phosphorylation sites within the MyBP-C motif, and a 28- amino acid insertion within the C5 domain. All isoforms of MyBP-C interact with the light meromyosin (LMM) region of the myosin rod, the region that forms the thick filament backbone (599). The COOH-terminal C 10 domain contains the LMM binding site of all MyBP-C isoforms (14,

38 1604 DAVIS ET AL. 663). In addition to binding the myosin rod, MyBP-C isoforms also bind to the thick filament region of the massive structural protein titin (252, 449). The interactions between MyBP-C, myosin, and titin likely contribute to the highly ordered structure of the sarcomere. The precise role of cmybp-c in cardiac function still remains unclear, but there is evidence suggesting that MyBP-C contributes to thick filament formation and sarcomere assembly (252, 449) and plays a regulatory role in cardiac muscle contraction. MyBP-C s role in cardiac function is still widely unknown but is suspected to contribute to -adrenergic-mediated gains in function (105, 357, 771, 772, 839, 840), myofilament Ca 2 sensitivity (102, 331, 365, 542), and length-dependent changes in myofilament Ca 2 sensitivity (105, 358). Additionally, MyBP-C must play a critical role in normal cardiac function as 20 30% of inherited HCM cases are attributed to mutations in the gene that encodes for MyBP-C (226, 638). To date, the role of MyBP-C in cardiac muscle function has not been studied using acute genetic engineering. Despite being a fairly sizable gene ( 21 kb) with 35 exons, the MyBP-C transcript is 4.5 kb (257) and can be accommodated by common adenoviral vectors (Table 1). Gene transfer of MyBP-C may be limited by its turnover time, which at present is unknown, and acute gene transfer may be further complicated by its stoichiometry. To date, it is still debatable whether MyBP-C stoichiometry is maintained. This debate is largely fueled by conflicting evidence from both human and transgenic mouse models that have various levels of MyBP-C expression. For instance, a clinical study reported no detectable levels of MyBP-C in HCM patients with MyBP-C mutant alleles (760), suggesting that the mutated MyBP-C is unstable with a pathogenic mechanism leaning towards one of haploinsufficiency. A heterozygous MyBP-C null mouse model also lends some support to the mechanism of haploinsufficiency as a knockout of one MyBP-C allele resulted in a slight but significant decrease in total MyBP-C content (102). In contrast, the development of transgenic mouse models that have two- to eightfold overexpression of either wild-type MyBP-C or an HCM-linked truncated MyBP-C demonstrated that MyBP-C stoichiometry is conserved and that the truncated MyBP-C is indeed stable (990). This result is further supported by ES cellderived transgenic mouse models with various COOHterminal truncations that retained total MyBP-C stoichiometry (542). The cardiac isoform of MyBP-C is the only isoform that contains phosphorylation sites. Phosphorylation of MyBP-C prevents myosin S2 and MyBP-C binding (301). The phosphorylation state of various C-protein sites alters thick filament structure and function. Ser-273, Ser-282, and Ser-302 of cmybp-c are key substrates for PKA (237, 972), Ca 2 /calmodulin-dependent kinase CAM kinase; Ref. 335), and PKC (924). Importantly, in addition to ctni, phospholamban and other Ca 2 handling proteins, cmybp-c is phosphorylated in response to -adrenergic stimulation (e.g., isoproterenol) by PKA, thereby suggesting a role for cmybp-c in the regulation of cardiac muscle contractility. Studies using gene transfer to investigate the contribution of posttranslationally modified MyBP-C on contractile function and/or studies to develop therapeutic strategies utilizing the phosphorylated versions of this protein have not been published. However, MyBP-C knockout and phosphorylation mimetic transgenic mouse models have been used to examine the role of MyBP-C phosphorylation as it pertains to the regulation of cardiac muscle contractility. Knockout models suggest that PKAmediated phosphorylation of MyBP-C contributes to both the magnitude of desensitization of the myofilaments to Ca 2 (102) and enhanced cross-bridge cycling during -adrenergic stimulation (345, 357, 434). A phosphomimetic transgenic mouse was designed using aspartic acid substitutions and showed that constitutive phosphorylation of MyBP-C conferred cardioprotection during ischemia-reperfusion injury (772). The precise mechanism whereby cmybp-c and its phosphorylation status affect cardiac contractility and cardioprotective behavior during ischemia is unclear, but acute genetic engineering of cmybp-c may offer a new tool to further uncover the role of MyBP-C phosphorylation in the fine-tuning of cardiac muscle contractility. 4. Thick filament accessory proteins: titin Titin is the largest mammalian protein currently identified. There are several titin isoforms ranging from 2.9 to 3.7 MDa in size that are all splice variants of the same gene, TTN. The titin gene is made up of 363 exons (33) and has a cdna of 82 kb (465). Elegant biophysical studies have shown titin s structural and functional role in cardiac muscle (290). Titin is largely responsible for the viscoelastic characteristic of striated muscle cells. Structurally, titin plays a key role in the maintenance of thick filament length and construction of Z and M-lines (296). Titin connects the Z-line to the M-line and has been modeled as a molecular spring as its structure contains an extensible region, the Ig-segment (immunoglobulin-like domain), and PEVK region (rich in proline, glutamate, valine, and lysine) that generates passive tension as sarcomeres become stretched. The relationship between passive tension and resting sarcomere length turns hyperbolic beyond that of optimal length. The different titin isoforms confer distinct variations in passive stiffness as demonstrated by the cardiac N2B isoform which elicits higher levels of passive tension per increase in sarcomere length relative to the N2A isoform due to N2B s shorter extensible region (104, 244, 465, 898). The heightened stiffness of the N2B isoform is thought to promote rapid return to diastolic filling during the cardiac cycle. Further-

39 CARDIAC MUSCLE GENE TRANSFER 1605 more, titin can be posttranslationally modified and may directly interact with Ca 2 (290). Titin s large size presents the biggest limitation to overcome when trying to utilize gene transfer for titin manipulation. Perhaps the use of titin fragments or truncations for acute gene transfer may be a way of overcoming the packaging size limitations but with the caveat that truncated titins cannot entirely extrapolate to physiological titin function. Gutted adenoviral vectors may also be a promising candidate for titin gene delivery. Furthermore, titin s turnover time is unknown and may be rather lengthy, offering an additional limitation for using acute gene transfer in vitro. V. CYTOSKELETAL PROTEINS FIG. 9. Schematic of the cardiac cytoskeletal network. The 427-kDa protein dystrophin links submembranous actin networks which connect to the sarcomeres via desmin to the transmembrane dystroglycan complex (DGC). The DGC interacts with the laminin-2 receptor and the extracellular matrix. The sarcoglycan complex and sarcospan are associated with the DGC. Cardiac myocytes have an extensive nonsarcomeric cytoskeleton network (Fig. 1). This noncontractile cytoskeleton consists of microtubules and lamins supporting the nucleus of the cardiac myocyte (773, 823). Intermediate filaments, primarily desmin, radiate from the Z-line and interact with neighboring sarcomeres, mitochondria, the nucleus, and the sarcolemma (Fig. 9). Near the plasma membrane these desmin filaments bind to the submembranous actin network (679). This submembranous actin lattice serves as an anchoring point for the mechanical and signaling complexes that form the costamere (Fig. 9) of cardiac myocytes (784). The costamere is thought to play a critical role in mechanotransduction and force transmission (65, 214, 784). Mediating these functions of the costamere are the proteins associated with the integrin heterodimer and dystrophin (327, 784). Cardiac muscle cytoskeletal proteins are critically important for normal physiology and are often affected in disease states (204, 891). For many of these proteins, only a partial picture of their role in cardiac pathophysiology is currently available. Despite this incomplete understanding of their function, it is clear that this is an important group of molecules in the development of heart disease. The utilization of gene transfer technologies has been an important factor in understanding cardiac diseases caused by defective cytoskeletal proteins. Furthermore, gene transfer has promise as a therapeutic agent for the treatment of cytoskeletal-based cardiac disease. A. Dystrophin and Dystrophin-Associated Proteins Dystrophin and its associated proteins have been linked to many forms of inherited muscular dystrophy, several of which present as cardiomyopathies. The nidus of this protein complex is dystrophin (Fig. 9), a 427-kDa protein that binds to the submembranous actin lattice and bridges the gap to the membrane. At the membrane, dystrophin interacts with the integral membrane protein -dystroglycan, which binds to the laminin receptor -dystroglycan, sarcospan, and the sarcoglycan complex (Fig. 9). This latter group of membrane proteins consists of -, -, -, and -sarcoglycan. In addition to these interactions with membrane-bound proteins, dystrophin binds to a variety of adaptor and signaling proteins. The following sections will discuss how gene transfer has been used to understand and treat diseases resulting from the disruption of the dystrophin glycoprotein complex. 1. Dystrophin The dystrophin protein is responsible for Duchenne and Becker forms of muscular dystrophy (209). In the more severe Duchenne muscular dystrophy, the dystrophin protein is completely absent from all tissues in the body, while in the milder Becker muscular dystrophy, a truncated, but partially functional, dystrophin protein is expressed in striated muscles. The dystrophin gene spans 2.4 million base pairs of the short arm of the X-chromosome. The resulting 14-kb transcript consists of 79 exons and produces a protein with 3,685 amino acids and a molecular mass of 427 kda, making it one of the largest known proteins (363). Dystrophin can be divided into four distinct functional regions (2). The NH 2 -terminal domain contains an actin binding region demonstrating homology with the actin binding proteins -actinin and -spectrin. A second actin binding domain is located in the central rod

40 1606 DAVIS ET AL. domain (766). These actin binding domains do not bind to sarcomeric -actin in vivo, but rather interact with G- actin present just below the membrane (326). It has been proposed that this submembranous actin network is connected to the sarcomeres by an interaction with desmin (97). The bulk of dystrophin consists of a central rod domain that contains 24 spectrin-like repeats, each of which is 100 amino acids in length. Interspersed within these repeats are hinge regions that are believed to provide flexibility to the molecule. Following this rod domain is a cystine-rich domain which interacts with the transmembrane protein, -dystroglycan, through a WW domain. -Dystroglycan forms a heterodimer with the laminin-2 receptor -dystroglycan (Fig. 9) (215, 216). The sarcoglycan complex (see below) and sarcospan are associated with the dystroglycan heterodimer. In skeletal muscle, both the dystroglycan and sarcoglycan complex expression require functional dystrophin. In contrast, most of these proteins are present at relatively normal levels in the dystrophin-deficient heart (316, 893). The fourth domain is the COOH-terminal region containing an -dystrobrevin binding domain. Both this dystrophin domain and -dystrobrevin each contains syntrophin binding sites, and syntrophin contains a PDZ domain that may function to bring signaling molecules (such as nitric oxide synthase) into close proximity of the dystroglycan and sarcoglycan complexes (287). One of dystrophin s proposed functions is as a stabilizer of the plasma membrane through its interactions with the intracellular cytoskeleton and extracellular matrix (153, 686). This function is particularly evident during lengthening contractions, where the extracellular matrix and intracellular cytoskeleton are moving in opposite directions (171). Another important function ascribed in part to dystrophin is the transmission of sarcomeric force to the extracellular matrix at the level of the costamere. Evidence supporting this function has been derived largely from the reduced force generation of both skeletal and cardiac muscle tissues from dystrophin null mice (171, 404, 790, 994). A third proposed function of dystrophin is related to its ability to form the nidus for the interaction and localization of many signaling molecules (287), although in skeletal muscle the contribution of this activity to the overall function of dystrophin appears to be limited (718). Duchenne muscular dystrophy is an X-linked disorder characterized by progressive skeletal muscle weakness (209). In addition to this overt skeletal phenotype, patients are afflicted with an insidious cardiomyopathy (209). Heart disease in Duchenne muscular dystrophy is characterized by both conduction, structural, and contractile abnormalities. End-stage heart disease in this case most frequently results in DCM and significant fibrosis, especially within the posterobasal region (232). An X- linked form of DCM has been linked to alterations within the 5 -end of the dystrophin gene (892). These patients have normal levels of dystrophin expressed in skeletal muscles, but no dystrophin present in the heart (570, 615, 892). Many of these patients have disruptions in the promoter driving dystrophin expression in muscle. In the heart, the inactivation of this promoter results in the absence of dystrophin (37, 38). In contrast, there is an upregulation of alternative dystrophin isoforms in skeletal muscle which functionally replace the muscle dystrophin (614, 631). In Becker muscular dystrophy (437, 460), truncated dystrophin molecules resulting from in-frame mutations are expressed (597). These truncated dystrophin molecules have varying levels of function, which are not associated solely with the size of the deletion (43). Information gained from analyzing genotype-phenotype relationships in Becker muscular dystrophy patients provided critical information for the design of functional truncated dystrophin molecules that are amenable to gene transfer. 2. Dystrophin gene transfer The vast majority of studies utilizing dystrophin gene transfer have been centered on replacing dystrophin in skeletal muscle for the treatment of Duchenne muscular dystrophy. Many of these advances are also critical to implementing dystrophin gene transfer within the heart. The most straightforward gene transfer approach is the direct injection of full-length dystrophin pdna (Fig. 3). Unfortunately, this method results in extremely low levels of dystrophin expression in skeletal muscle (7). Although inclusion of additional transfection agents may improve the efficiency of transduction (897, 988), the levels achieved are likely too low to have physiological/therapeutic significance. The low efficiency of these methods raises the question of what are therapeutic levels of dystrophin. Studies involving female carriers of mutated dystrophin alleles suggested that levels slightly greater than 50% of normal are protective against skeletal muscle disease (127, 538), although other studies of X-linked DCM patients report that levels as low as 30% of normal dystrophin expression can prevent significant skeletal muscle disease (637). The therapeutic threshold in the heart is complicated by the inability to obtain cardiac biopsies to determine the level of expression and by variability in the manifestation of cardiac disease in these patients (288, 370, 580, 649). The disappointing results from direct plasmid injections underscored the need for a more efficient method of gene delivery. Viral-based gene delivery vectors have been shown to be the most effective and efficient methods of introducing exogenous DNA into living cells (see sect. IIA). The only current viral vector capable of delivering full-length dystrophin is gutted adenoviral vectors (Table 1) (314, 447, 459). Direct injection of this vector into

41 CARDIAC MUSCLE GENE TRANSFER 1607 skeletal muscle resulted in an inefficient transfer of fulllength dystrophin to muscle fibers local to the site of injection (172, 264, 535), although even these modest levels of transduction improved muscle function (172, 535). Clearance of transduced cells is a significant problem with adenoviral-based vector systems (see sect. IIA2), but expression can be prolonged by performing injections in neonates, which do not develop a strong immune response to the vector (110, 126, 196). One of the most promising gene therapy vectors for striated muscle is recombinant AAV. Significantly truncated dystrophin molecules that retain functionality were first described in patients with Becker muscular dystrophy (212), and additional studies mapped the critical domains required for dystrophin function (144, 329, 687, 718, 777). These studies culminated in several functional dystrophin constructs small enough for AAV delivery (219, 329, 893, 937). Most of these truncated dystrophin constructs improved muscle pathology and membrane integrity following intramuscular injection into skeletal muscle. Only one of these constructs has been accessed to any degree in the myocardium (293, 893, 997). Dystrophindeficient hearts expressing this truncated micro-dystrophin showed a significant improvement in ventricular geometry (893). The expression of micro-dystrophin protected dystrophin-deficient mice from acute heart pump failure during a dobutamine stress test (893). This same construct also significantly extended the life span of mice deficient in both dystrophin and utrophin, a dystrophin homolog upregulated in the dystrophin-deficient mouse (293). Despite these significant improvements in global cardiac function, detailed physiological assessment of this truncated dystrophin molecule revealed that deficits remain, including a reduced transmission of force by striated muscle expressing micro-dystrophin (293, 893). These studies indicate that there is room for improvement in the design of these micro-dystrophins. While truncated dystrophin molecules are widely expressed in the heart when delivered by a single AAV (293, 294, 893), they are not fully functional. Other approaches are being explored, including the use of two AAVs, each containing half of a fully-functional dystrophin protein with splice sites engineered between them. After transduction of the cell, the AAV genomes concatamerize. Following transcription of these AAV genomes, the resulting RNA is processed at the engineered splice sites. The result is a dystrophin molecule that is larger and more functional than the micro-dystrophin delivered by a single AAV (469). The efficiency of this splicing event, however, appears somewhat less than single vector alone (262, 469). Further studies are necessary to determine the functional effects of this trans-splicing approach. Another promising use of cardiac gene transfer in the treatment of Duchenne muscular dystrophy is the use of AAV to introduce small RNA molecules that modulate splicing activity (174). In many patients with Duchenne muscular dystrophy, skipping a few exons yielded a highly functional dystrophin molecule. This approach has been shown to be effective in the heart and shows therapeutic promise for a significant subset of Duchenne muscular dystrophy patients. 3. Sarcoglycans Shortly after its initial characterization, dystrophin was found to interact with a group of membrane-bound glycoproteins (217). These proteins formed a complex, which functionally links dystrophin to the extracellular matrix by binding to laminin (215, 216, 381). This complex of dystrophin-associated proteins consists of two groups of proteins: the dystroglycan and sarcoglycan complexes (995). The laminin binding component of the dystroglycan complex consists of - and -dystroglycan (Fig. 9), which are cleaved components of a common precursor (381). The sarcoglycan complex (Fig. 9) consists of four subunits: -, -, -, and -sarcoglycan (587, 643). All four of these genes have been implicated in various forms of limb-girdle muscular dystrophy (LGMD), and many of them have significant cardiac phenotypes. 4. -Sarcoglycan -Sarcoglycan ( -SG) is a 387-amino acid protein with a single transmembrane domain and a large extracellular NH 2 -terminal domain. Extensive glycosylation of two conserved asparagine residues yields a final molecular mass of 50 kda (747). Mutations resulting in the loss of -SG cause LGMD-2D (101, 749). Interestingly, the absence of -SG results in the concurrent loss of the other members of the sarcoglycan complex (195). -SG deficiency results in severe muscular dystrophy that is essentially localized to skeletal muscle, although there are instances of cardiac disease (195, 691). Gene transfer of -SG in the skeletal muscle of -SG-deficient mice was shown to fully restore the entire sarcoglycan complex (13, 193). Interestingly, the overexpression of -SG following gene transfer caused significant cellular toxicity. The toxic effects of -SG were independent of an immune response and seemingly secondary to alterations in the assembly of the sarcoglycan complex in the presence of excess -SG (193). 5. -Sarcoglycan -Sarcoglycan ( -SG) has a similar structure to that of -SG, with a single transmembrane domain and the majority of the protein present in the extracellular space. -SG is glycosylated at three conserved asparagine residues, resulting in an increase in molecular mass from 35 kda to a final weight of 43 kda (72, 493). The clinical importance of this new protein became evident as genetic

42 1608 DAVIS ET AL. studies linked -SG to LGMD-2E (72, 493). Clinically, -SG deficiency is characterized by significant skeletal muscle disease (72, 493). Cardiomyopathy has also been reported in patients with LGMD-2E (36). Mice in which -SG has been ablated also mirror the condition in humans, with significant pathology present in both skeletal and cardiac muscle (19, 197). -SG functions as a nidus for the entire sarcoglycan complex, and the absence of -SG causes a complete loss of other sarcoglycan complex members (197, 647). In contrast to -SG, -SG is also required for the formation of the sarcoglycan complex in vascular smooth muscle, which may in part explain the greater cardiac involvement associated with -SG deficiency (197). Gene transfer of -SG was shown to restore the expression of the entire sarcoglycan complex and improve pathology (193, 197). Unlike -SG, overexpression of -SG did not result in significant cellular toxicity (193). 6. -Sarcoglycan -Sarcoglycan ( -SG) is a glycosylated protein with a molecular mass of 35 kda. Mutations within this gene have been implicated in LGMD-2C (549, 646). In contrast to other sarcoglycanopathies, the sarcoglycan complex is not completely lost in the absence of -SG (145, 549). Clinically, LGMD-2C is characterized as a severe form of muscular dystrophy, with significant cardiac disease (44, 45). Consistent with this clinical presentation, mice lacking -SG develop severe skeletal and cardiac pathology (312). Gene transfer of -SG into skeletal muscle was shown to largely reconstitute the expression of -SG and corrected the skeletal muscle pathology (138). Interestingly, similar to -SG, overexpression of -SG resulted in significant pathology, presumably secondary to disruption of normal assembly of the sarcoglycan complex (1016). 7. -Sarcoglycan Shortly after the first description of the dystrophinassociated proteins, it was found that the cardiomyopathic hamster was lacking the sarcoglycan complex (748). The absence of mutations within the -, -, or -SG genes strongly suggested a fourth gene may cause this disease in the hamster (548). To uncover this putative fourth sarcoglycan, a homology screen found a transcript similar to -SG (643). -SG is a 35-kDa glycoprotein with a structure similar to -SG (643). This gene was quickly linked to LGMD-2F (641) and to the cardiomyopathic hamster (642, 777). Similar to the other disorders involving sarcoglycans, patients with mutations in -SG often have severe skeletal muscle disease (643). In contrast to the other sarcoglycanopathies, mutations within -SG have been linked to cardiac disease in the absence of skeletal muscle disease (904). Mice lacking -SG also have significant cardiac and skeletal muscle pathology (136, 311). Intriguingly, the cardiac damage initiated by exercise in the -SG null mice was prevented by pretreatment with vascular relaxing agents, suggesting an important pathophysiological role of vascular smooth muscle (129, 136). Like the other sarcoglycans, -SG gene transfer into skeletal muscle largely corrected the muscular dystrophy present in the -SG-deficient hamster (368). Similarly, direct injection of myocardium with Sendai virus-coated proteoliposomes or AAV containing -SG resulted in reconstitution of the entire sarcoglycan complex ( ). The levels of expression obtained through direct injection sufficiently improved global cardiac function and increased the life span of the cardiomyopathic hamster (428, 429). Global cardiac transduction, through the infusion of gene therapy vectors into the coronary circulation, yielded results similar to that obtained by direct injections (385). Systemic intravascular administration of AAV containing -SG resulted in wide-spread expression of -SG in both skeletal and cardiac muscle (1014) and reconstituted the sarcoglycan complex, correcting pathological changes in striated muscle. -SG gene transfer also improved cardiac function and significantly extended the life span of treated hamsters compared with those receiving no gene therapy (1014). B. Intermediate Filaments (desmin) The most prominent intermediate filament in cardiac muscle is desmin (Fig. 9), which forms coiled dimers that self-assemble into homo- and heteropolymers with a number of proteins including nestin and synemin to form filaments 10 nm in diameter (249). This process is facilitated by the heat shock protein B -crystallin (639). The lattice of desmin filaments extends from the Z-line interacting with a variety of cellular structures including the nucleus and mitochondria. In addition to interactions with these organelles, many desmin filaments connect adjacent Z-lines, functioning to keep neighboring sarcomeres registered. Desmin filaments emanating from Z- lines at the periphery of the myocyte extend towards the sarcolemma. Near the membrane, these intermediate filaments interact with the proteins that make up the costameres. These proteins, including the complex of proteins associated with the integrins and dystrophin, are localized in the region of the Z-line via an interaction with desmin (576). Desmin filaments are also prominent in Purkinje fibers and at the intercalated disc (442, 887). A variety of inherited cardiomyopathies have been defined as desmin-related myopathies and are characterized by the deposition of desmin aggregates within cardiac and/or skeletal muscle. These disorders can be categorized into two groups: primary mutations within the desmin gene and those within desmin-associated proteins (96, 679). The latter group consists of mutations of B -

43 CARDIAC MUSCLE GENE TRANSFER 1609 crystallin (392, 926). These mutations appear to result in defective assembly and subsequent aggregation of the desmin filaments (76). Expression of mutated forms of B -crystallin in adult cardiac myocytes was detrimental to contractile function (521). The myofilaments remained intact in the presence of the mutated B -crystallin, but contraction was attenuated. Further studies found a disruption of mitochondrial structure and function in myocytes expressing a mutated B -crystallin (521). Mutations within desmin can result in pathology in skeletal muscle, cardiac muscle, or both. Most of these mutations occur in the central -helical domain of desmin. Many of these mutated proteins are unable to form filaments in vitro (34), and all result in the formation of desmin aggregates in diseased tissue (272, 285, 613, 826). Desmin mutations have been linked to both dilated (482, 584) and restrictive cardiomyopathies (20, 712, 930). Gene transfer of mutated desmin into neonatal cardiac myocytes disrupted the sarcomeric banding pattern (341). These studies demonstrate that alterations in desmin filament assembly have a dominant effect on cellular sarcomeric assembly, an observation consistent with the dominant mode of inheritance that characterizes desminrelated cardiomyopathies. C. Microtubules Microtubules are hollow filaments consisting of and -heterodimers of tubulin. Both tubulin subtypes are GTP-binding protein; however, only the GTP bound to -tubulin is hydrolyzed during polymerization. Microtubules are dynamic structures that run longitudinally across the myocyte and are concentrated in the perinuclear regions (275). Under normal conditions, microtubules have small effects on the structural properties of cardiac myocytes (135, 644). They appear to function primarily as a means of organizing particles, vesicles, and organelles within the myocyte. Microtubules serve as the tracks for the complementary motor proteins kinesin and dynein. These proteins contribute to properly distributing macromolecules and organelles within the cell. In addition to this role in subcellular organization, microtubules also have critical roles in signal transduction within the myocyte (95). Microtubular accumulation is implicated in the pathophysiology of ischemic cardiomyopathy, depression of function with hypertrophy, and diabetic cardiomyopathy (95). This effect is particularly apparent with the cardiac dysfunction associated with pressure overload hypertrophy. The importance of microtubule accumulation is evidenced by the remediation of the contractile deficit by agents that depolymerize microtubules (134). This accumulation of polymerized microtubules is associated with increased expression of microtubule associated protein 4 (MAP4) and 1 -tubulin, but not 4 -tubulin, which is the predominate isoform in the heart (634). Adenoviral-mediated overexpression of these proteins in normal adult cardiac myocytes revealed that MAP4 is sufficient for these myocytes to develop microtubular structure similar to that observed in myocytes isolated from pressure-overloaded hearts. In contrast, adenoviralmediated expression of either 4 -or 1 -tubulin had no significant effect on microtubule assembly (864). The increased density of microtubules present within pressureoverloaded myocytes resulted in a significant increase in the viscous load placed on the myofilaments. This increased load significantly reduced the efficiency of the contracting cardiac myocyte and clearly contributed to the poor contractile performance of pressure-overloaded hearts (860). Microtubules play an important role in the pathophysiology of several models of heart failure secondary to pressure overload, but the importance of microtubules in other models of heart failure remains unclear (130, 163). The potential of microtubules to be manipulated by gene transfer technologies introduces the possibility of experimental or therapeutic modulation of microtubules to improve our understanding of these cytoskeletal elements in the failing heart. VI. CARDIAC SIGNALING PATHWAYS Gene transfer has been a valuable approach for studying the effects of signaling pathways on myocyte function and for identifying the cellular targets within a signaling cascade. This approach also is helpful for determining whether a signaling pathway acutely influences cardiac function in a setting that is independent of load and possible compensatory adaptations that may develop in transgenic animals. Gene transfer offers the ability to study the temporal and dose-dependent modulation of contractile function by a signaling pathway, which may be difficult and/or expensive to follow in transgenic animals. This approach could be useful for studying the relationship between expression of a specific signaling protein in cardiac pathophysiology and heart failure. The goal of this portion of the review is to focus on information gained about signaling pathways and their influence on contractile function using gene transfer. While gene transfer studies of signaling via transcription factors are important, studies described here focus primarily on the influence of signaling pathways on ion transport, Ca 2 cycling, and myofilament proteins, as well as contractile function. A potential limitation of studies using in vivo adenoviral-mediated gene transfer is the difficulty in achieving long-term, homogeneous expression at the organ level. However, a new generation of vectors appears to be overcoming issues of expression duration and homogeneous expression (954). Heterogeneous and/or transient

44 1610 DAVIS ET AL. expression may be a desirable goal for studies of signaling cascades, which more likely operate under spatial and/or temporal activation. Overexpression of signaling proteins may be successful if targeted expression below pharmacological levels is achieved in the intact heart. Many of the signaling pathways discussed below are mediated via G protein-coupled receptors. The focus of this section is on interactions between signaling proteins involved in modulating Ca 2 cycling and myofilament proteins and their overall influence on cardiac myocyte contractile function. Importantly, signaling pathways can exert both acute and chronic influences on contractile function. The signaling section will begin with reviewing the -adrenergic signaling pathway as it pertains to the contributions to the field gleaned from gene transfer studies. A. Gene Transfer Influencing the -Adrenergic Signaling Pathway 1. -Adrenergic receptors and G proteins One of the most studied and best understood signaling pathways involved in modulating cardiac contractile function is the -adrenergic receptor ( -AR) pathway (48, 506, 696, 883). Catecholamine binding to the 1 -AR activates adenylyl cyclase via G s to increase camp production, which in turn activates PKA (Fig. 10). The PKA catalytic subunit phosphorylates multiple targets, including proteins located within the sarcolemma, SR, and sarcomeres (Fig. 10) to increase cardiac contractility and relaxation rate. This basic signaling cascade is well understood based on a variety of experimental approaches including biochemical, transgenic, and knockout models (506, 883). While decreases in -AR signaling were noted during heart failure in the 1980s (79), -AR agonists paradoxically increased mortality in failing hearts (342, 669, 818). Gene transfer studies have been important for providing new insight into the pathway of -AR cycling during desensitization/downregulation and the role of -AR cycling during heart failure (340, 635, 669). Gene transfer approaches have also provided fundamental knowledge about adenylyl cyclase and A-kinase anchoring proteins (AKAPs) (324, 761). These latter proteins serve as scaffolding proteins that contribute to macromolecular assembly of activated PKA with phosphorylation targets Downloaded from by on June 29, 2017 FIG. 10. Model of protein kinase A (PKA)-mediated -adrenergic signaling in cardiac myocytes. -Adrenergic-1 receptors ( 1 AR) are G protein-coupled receptors that when stimulated initiate adenylyl cyclase (AC)/cAMP-dependent activation of PKA. PKA has several intracellular targets in the cardiac myocyte including voltage-gated Ca 2 channels (DHPR), ryanodine receptor (RyR), phospholamban (PLN), and the myofilament proteins cardiac troponin I (ctni) and myosin binding protein C (MyBP-C). Inhibitory protein 1 (I-1) and phosphodiesterase (PDE) are also substrates of PKA-mediated phosphorylation. -Adrenergic-2 receptors ( 2 AR) are also G protein-coupled receptors that when activated can initiate the phosphatidylinositol 3-kinase (PI3K) and ERK signaling cascades or activate inhibitory protein 1 (I-1), which decreases protein phosphatase 1 (PP1) activity. PP1 is a negative regulator of PKA activity and dephosphorylates PKA targets like PLN.

45 CARDIAC MUSCLE GENE TRANSFER 1611 (761). In addition, this approach has been instrumental in identifying the relative contribution of individual targets to the inotropic and lusitropic effects of -adrenergic signaling in cardiac myocytes (178, 207, 967). Most importantly, studies on the -AR signaling pathway have provided insight into the paradoxical observation that -AR agonists increase mortality by demonstrating that -agonists activate a cascade of pathophysiological events in addition to the downstream increase in contractile function. By addressing the events that directly influence contractile function, viral-mediated delivery systems have begun to provide therapeutic strategies for treating heart failure using the -AR signaling pathway (201, 635, 881). An example of therapeutic strategies designed to improve cardiac function involves viral-mediated gene delivery of 2 -adrenergic receptors (192, 202, 540, 882). While the highest proportion of -ARs in the heart are 1 -receptors, the proportion of 2 -receptors increases with heart failure, as 1 -ARs are downregulated during chronic stimulation (754, 828, 1010). Adenoviral gene transfer of two different naturally occurring 1 -AR polymorphisms to isolated rodent cardiac myocytes had significant chronotropic effects, and one of the 1 -AR variants greatly enhanced the sensitivity of the myocytes to a common -blocker used to treat heart failure (751). Like 1 -ARs, 2 -receptors are also coupled to PKA activation (85). Signaling via 2 -ARs, however, is more complex than 1 (883), because the 2 receptor is coupled to both G s and G i. G i subunits negatively influence contractile function (980) (Fig. 10). In addition, 2 -receptors show significant differences in their spatial activation of PKA compared with 1 -ARs (15, 883). Complementary transgenic models with moderate overexpression of 2 -ARs improved contractile function without significant hypertrophy (753). Gene transfer of -ARs has also been used as a therapy to treat failing hearts (412, 464, 569, 882). Gene transfer of 2 -AR during cardiopulmonary bypass in neonatal pigs increased total -AR density and improved left ventricular function in response to -adrenergic stimulation (412, 413). Increases in these receptors after gene transfer improved ventricular function at baseline and with -adrenergic stimulation in loaded and unloaded rabbit hearts (540, 812) as well as in response to higher preloads (464). The beneficial effects of 2 -AR overexpression are clearly dose-dependent, and high levels of overexpression are not considered beneficial (492). 2 - ARs signal via multiple downstream pathways. Gene transfer of 1 -AR or 2 -AR into double knockout mice indicated that the protective effects of 2 -AR are likely mediated via G i and the downstream activation of phosphoinositide 3-kinase (PI3K) and Akt (477, 1015) or via functionally discrete signaling pools (506, 828). Gene transfer approaches have also been used to understand the contribution of G i to contractile function. Increased expression of G i was observed with heart failure, which may influence signaling via 2 receptors (208). Atrioventricular viral-mediated gene transfer of wild-type G i or constitutively active G i (Q205L) significantly improved heart rate during persistent atrial fibrillation in a pig model (41, 186). Gene transfer of G i did not significantly influence baseline isometric force in rabbit trabeculae or shortening in isolated myocytes (405). However, expression of G i blunted the response to -AR activation by isoproterenol in both the trabecular and myocyte preparations, and this blockade was prevented with pertussis toxin. Overall, these results indicate that signaling via G i acts as a brake on the -adrenergic contractile response, and it may be important for the beneficial effects of 2 receptor gene therapy. Another strategy along these same lines has been to increase other G protein-coupled receptors, including parathyroid hormone related peptide (PTHrp) and vasopressin2 (V2) receptors, using viral-mediated gene transfer (474). Like 2 -ARs, PTHrp receptors are coupled to G s and G i and downstream activation of phospholipase C, while V2 receptors are strongly linked only to G s (58, 474, 806). Neither receptor, however, is expressed endogenously in cardiac myocytes. The increased expression of V2 receptors had no significant effect on basal contractile function but significantly improved the response to agonists. In contrast, gene transfer of PTHrp receptors increased basal contractile velocity in failing and nonfailing myocytes, but there was no further response to agonists. At present, it is unclear whether therapies with these receptors significantly improve contractile function at the cellular and/or organ level under pathophysiological conditions. 2. Cycling of -ARs G protein-coupled receptor kinases (GRKs) are important for the rapid modulation of -AR density on the myocyte cell surface (752). The most prominent of these GRKs, ARK1 (or GRK2), is elevated during pathophysiological conditions (339), such as ischemia (480), heart failure (192, 201, 752, 909), and coronary artery bypass (rabbits) (881). ARK1 is responsible for phosphorylation of agonist-occupied -ARs, which leads to uncoupling of the receptor from downstream signaling pathways (446). Adenoviral delivery of the ARK1 inhibitor ARK-1ct, which is constructed from the COOH terminus of ARK1 and acts as a dominant negative, significantly attenuated the left ventricular dysfunction in hearts undergoing coronary artery bypass and after myocardial ischemia (813, 881, 968). Viral delivery of this inhibitor to myocytes from spontaneously hypertensive rats in heart failure also significantly improved -AR-mediated camp production, peak shortening, and the rates of rat myocyte contraction as well as improved signaling in failing rabbit myocytes (9, 201). While this inhibitor restored ARK1 levels in two

46 1612 DAVIS ET AL. dilated cardiomyopathy models (200, 328), it did not consistently prevent the progression of cardiomyopathy or premature mortality. These results suggest that ARK-1 plays a critical role in modulating -AR density on the sarcolemma, but the animal model, type of cardiac pathophysiology, and timing of gene delivery may be critical for the future development of therapeutic treatments for human heart failure. As described below, events and/or other signaling pathways may also contribute to the final progression to heart failure. More recently, gene transfer of the ARK1ct has been compared with truncated phosducin, another GRK inhibitor (489). Both inhibitors bind to G subunits, yet only ARK-1 contributes to -AR cycling and improves camp accumulation in myocytes from failing rabbit hearts (489). Gene transfer of either inhibitor improved the left ventricular contractile response to isoproterenol, and it improved fractional shortening and LV end-diastolic dimension in rapidly paced, failing hearts. These results suggest that the beneficial influence of GRK inhibitors on failing hearts maybe due, at least in part, to the inhibition of G subunits. Proteins associated with ARK1 in the myocyte have also been targets for gene transfer and treatment of heart failure. PI3K binds to ARK1 in the cytosol, is targeted to -ARs during agonist activation of -ARs, and regulates -AR internalization (625). Overexpression of the kinase domain, PIK, competitively displaces PI3K from ARK1 and prevents -AR internalization and desensitization. Transient PIK expression decreased PI3K localization with -ARs in sarcoma cells, and viral-mediated gene transfer of PIK into myocytes from failing pig hearts restored the isoproterenol-mediated enhancement of peak shortening and contraction and relaxation rates toward the nonfailing phenotype (683). The loss of -ARs, increased norepinephrine (NE) levels, and/or chronic -AR activation that result from heart failure are ultimately maladaptive (640, 683). Future studies may focus on gene therapy utilizing modified proteins associated with the -AR cycling pathway, such as -arrestin (477) and PDE4 (375, 488), which serve as important proteins in scaffolding and/or desensitization of 2 -ARs. In addition to -ARs and associated proteins, studies have also focused on clearance of hormones involved in -AR signaling, especially NE. Accumulation of NE during hyperstimulation of -AR was postulated to be a major cause of structural and functional impairment during heart failure (726). Efforts to improve local NE clearance using gene transfer of uptake-1 into rabbit hearts prior to pacing-induced heart failure acutely improved NE uptake, 1 -AR receptor, and SERCA2 expression, as well as diastolic and systolic contractile function, over a 2-wk time period (612). Gene transfer of uptake-1 had no significant influence on contractile function in nonfailing hearts. It is currently unclear whether the uptake-1 expression develops in cardiac myocytes or in associated neurons after gene transfer, and whether or not this strategy may be appropriate for future therapy to treat heart failure. 3. Downstream -AR signaling A) ADENYLYL CYCLASE. Gene transfer strategies are providing significant strides toward understanding signaling downstream from the -AR. Viral-mediated gene transfer of adenylyl cyclase VI (AC-VI) was first studied in neonatal rat myocytes, which selectively increased camp production in response to -AR activation but not to other agonists coupled to adenylyl cyclase (666). Gene transfer also increased the cardiac response to -adrenergic activation days after recombinant viral delivery in normal, ischemic, and failing hearts (324, 467, 468, 759, 865). Further work demonstrated that AC-VI delivery increased the baseline rate of contraction (dp/dt max ), and -adrenergic responses were maintained after gene transfer into mice (758). Importantly, gene transfer of adenylyl cyclase into pigs experiencing pacing-induced heart failure improved contractile function and decreased ventricular dilatation, as well as improved camp production and reduced indicators of hypertrophy. These findings also agree with results obtained after transgenic expression of AC-VI, which improved mortality and ventricular function in mice with ischemic heart failure (865). Many of the beneficial effects of AC-VI can be explained by the resulting increase in camp production. However, increased camp is associated with arrhythmias, which were not observed in the transgenic model. Thus the improved mortality observed in AC-VI transgenic mice may be due to 1) the close proximity of this isoform to the sarcolemma and 2) camp-dependent effects of AC-VI (324, 479). It remains unclear whether gene delivery of AC-VI is capable of reversing LV morphological and functional remodeling, but this gene is garnering consideration for clinical heart failure therapy (324). B) G S-COUPLED RECEPTORS. Gene transfer has been used for upstream activation of adenylyl cyclase using receptors for arginine vasopressin (AVP). Increased AVP expression is associated with congestive heart failure and is linked to a poor prognosis (274). The vasopressin 1 (V1) receptor is expressed in the heart while V2 receptors are expressed in renal collecting ducts but not in cardiac myocytes. The V1 receptor is coupled to phospholipase C (PLC)- while V2 is coupled to G s and adenylyl cyclase. Viral-mediated gene transfer of V2 in adult rat ventricular myocytes resulted in dose-dependent expression, an increase in camp formation, and an increased amplitude of contraction in isolated myocytes that was blocked by a V2-specific vasopressin antagonist (473). Coronary-based adenoviral gene delivery to the myocardium also improved fractional shortening and the rate of contraction in response to V2 stimulation (953). However, it remains to

47 CARDIAC MUSCLE GENE TRANSFER 1613 be determined whether gene delivery of V2 improves cardiac performance in failing hearts. In addition, the long-term benefits of this approach are controversial, as blockade of both V1 and V2 is also predicted to improve function in individuals with congestive heart failure (273). This general approach has also been utilized to determine whether delivery of other noncardiac G s -coupled receptors to cardiac myocytes acts as a novel therapeutic strategy for bypassing the -AR and boosting contractile function (474). C) AKAPS. Several experiments on downstream signaling within the -AR pathway have focused on AKAPs. AKAPs serve as cellular scaffolding for PKA by tethering the type II regulatory subunit (RII), as well as target proteins, and phosphatase/phosphodiesterases. At least 13 different AKAPs have been identified in myocardium, and the localization and function of these AKAPs was recently reviewed in detail by Reuhr et al. (761). Gene transfer and expression of Ht-31, a peptide derived from human thyroid AKAP with a similar binding affinity as AKAP for the RII domain of PKA, resulted in redistribution of the RII subunits within isolated myocytes (231). This change in PKA localization enhanced -AR-mediated peak shortening as well as the rate of shortening and relengthening compared with controls (231). Paradoxically, the level of myofilament protein phosphorylation (e.g., ctni, MyBP-C) in response to -AR stimulation was significantly reduced in these myocytes compared with controls. Results from these studies provided insight into the temporal and spatial targeting of AKAPS within cardiac myocytes (231, 763). In addition, this approach has provided a better understanding of the role played by each AKAP in myocyte hypertrophy (182). Future studies are needed to investigate the influence of AKAPs on contractile function under pathophysiological conditions, as modified AKAPs could one day be used to treat heart failure. D) END-TARGET PROTEINS. The end targets for phosphorylation have also been studied using gene transfer, and this approach has contributed significantly towards understanding -AR signaling. These studies have focused primarily on Ca 2 cycling and sarcomeric proteins that serve as targets for -AR signaling, with direct influences on contractile function. A high proportion of these studies have focused on phospholamban and sarcomeric troponin I, which have been described above (see sects. III and IV). Below are studies of other downstream targets. 4. Myosin binding protein C The -AR pathway phosphorylates a number of other proteins, including the sarcolemmal L-type Ca 2 channel (DHPR) and sarcomeric MyBP-C. Studies using gene transfer to investigate the contribution of these proteins in contractile function and/or studies to develop therapeutic strategies utilizing the phosphorylated versions of these proteins have not been addressed in detail using viral-mediated gene transfer approaches. In the case of Ca 2 channels, as well as other ion channels or transporters, there are multiple subunits. Thus cellular or organdirected gene transfer and overexpression may not necessarily result in appropriate channel subunit organization. Experiments with modifications of MyBP-C on the PKA-targeted Ser283 (593) also may be difficult to perform, as the turnover of MyBP-C is not clear and may be too long for the typical 1 7 day time frame for myocyte contractile assays in vitro. 5. Heat shock proteins, p20 Proteomic analysis of the -AR response to isoproterenol has revealed phosphorylation of the p20/phspb6 heat shock protein. This small heat shock protein is an -crystallin and is typically present in the cytosol of cardiac myocytes (120, 476). Phosphorylation of p20 results in its colocalization with actin (222), although recent studies suggest that p20 does not bind directly to actin (90). It has been postulated that the p20 phosphorylation state determines the translocation and the localization pattern. Viral-mediated gene transfer of p20 into adult rat myocytes increased peak shortening and Ca 2 transient amplitudes within 2 days after gene transfer without significantly influencing resting length, relaxation, or Ca 2 decay (120). Although the cellular basis for the increase in contractility and Ca 2 transient remains unknown, p20 localization with actin and its ability to bind PP1 (221) would suggest that this small heat shock protein may have direct influences on proteins within the Ca 2 cycling cascade and/or the myofilaments. It should be noted that gene transfer has been used to study the protective effect of other heat shock proteins against stresses such as heat and/or ischemia in cardiac myocytes (471). However, to date, gene transfer of these other heat shock proteins has not focused on the direct influence on contractile function. 6. Protein phosphatase 1 PKA modulates PP1 activity via phosphorylation of inhibitor proteins. The role played by PP1, the inhibitor proteins, and modulation of these proteins by PKA and PKC is discussed in more detail below. B. Gene Transfer of Ca 2 /Calmodulin Kinase The Ca 2 /calmodulin-dependent protein kinase II (CAMKII ) isoform phosphorylates several proteins involved in EC coupling, including the RyR and PLN. Increased CAMKII activity is observed during heart failure, and overexpression of the cytosolic variant CAMKII c in

48 1614 DAVIS ET AL. transgenic animals causes heart failure (50, 518). Viral delivery and overexpression of wild-type CAMKII c in rabbit myocytes altered Ca 2 cycling, but overexpression did not directly alter other proteins such as NCX or SERCA, changes that are commonly associated with heart failure (448). The major influences on Ca 2 cycling proteins included phosphorylation of RyR, without changes in the association of RyR with FKBP12.6. This phosphorylation in rabbit myocytes was associated with increased SR fractional release of Ca 2, increased diastolic leak of Ca 2 and reduced SR Ca 2 content. Despite the decrease in SR Ca 2 content, the cellular Ca 2 transient and shortening amplitude were maintained due to the increase in peak DHPR current and increased SR fractional Ca 2 release. Overexpression also enhanced the frequency-dependent acceleration of relaxation, presumably due to its phosphorylation of PLN at Ser-17 and the resulting increase in Ca 2 uptake by SERCA2. In contrast, gene transfer of wild-type, constitutively active, and dominant negative CAMKII c into isolated rat myocytes produced quite different results (989). Here, increased SR Ca 2 content and reduced Ca 2 release in response to elevated Ca 2 were observed in myocytes overexpressing wild-type or constitutively active CAMKII c, and these results were interpreted to indicate that CAMKII c acts as a negativefeedback modulator of RyR. The differences between the studies with rabbit and rat myocytes have been attributed to species differences in Ca 2 handling between rat and rabbit hearts (989). Future gene transfer studies will be necessary to better understand the role played by CAMKII and may ultimately be critical for defining the direct versus compensatory adaptations that develop in transgenic mice expressing CAMKII. C. Gene Transfer and PKC Signaling Activation of PKC plays a key role in mediating signaling via multiple receptors (e.g., angiotensin II, endothelin, -adrenergic receptors; Fig. 11) (152, 211, 453, 598, 688, 880, 940) coupled to G proteins (e.g., primarily G q,g i ) followed by PLC activation. PLC breaks down phosphoinositide bisphosphate (PIP 2 ) into inositol 1,4,5-trisphosphate (IP 3 ) and diacylglycerol (DAG). IP 3 is translocated to the SR (Fig. 11), and binding to its receptor contributes to 5% of SR Ca 2 release (399, 603). Recently, FRET was used by labeling the binding domain of the IP 3 receptor with cyan fluorescent protein, while target protein sequences were labeled with yellow. This biosensor approach demonstrated the spatiotemporal distribution of IP 3 in neonatal cardiac myocytes (733). SR Ca 2 release increases in response to low levels of IP 3 (433, 655), and future gene transfer approaches may prove valuable for imaging, Ca 2 cycling, and/or contractile function studies in adult cardiac myocytes. A much larger body of literature has focused on direct and indirect modulation of cardiac function by PKC using gene transfer approaches (556, 958). Divergent contractile responses are observed in response to PKC activation, with reports of both increases and decreases in contractile performance. The multiple agonists that activate PKC, the presence of multiple PLC and PKC isoforms, downstream activation of other signaling pathways, and phosphorylation of multiple end-target proteins (10, 152, 426, 453, 651) add to the complexity of this signaling pathway. Gene transfer studies have helped define and map this pathway in cardiac myocytes, which has improved our understanding of its direct influence on contractile function. The remainder of this section focuses on insights gained from gene transfer studies. 1. Signaling upstream from PKC A) RECEPTORS AND PLC. Gene transfer of receptors that activate the PKC pathway has been utilized in several studies (Fig. 11). These studies have focused on the relative contribution of receptor isoforms to hypertrophy mediated via PKC and/or the ability of a specific isoform to counteract the hypertrophic phenotype produced by a dominant receptor isoform (243). With this emphasis, the acute and chronic influence of receptor isoform expression on cardiac contractile function has not yet been as thoroughly studied by gene transfer. For example, two major isoforms of the angiotensin II (ANG II) receptor (AT-1 and AT-2) are expressed in adult myocytes (82, 349). Adenoviral-mediated gene transfer of AT-1 and AT-2 each produced hypertrophy in neonatal rat myocytes, and direct lentiviral-mediated cardiac delivery of AT-2 reduced the hypertrophic response to chronic ANG II delivery (148, 220). The importance of AT receptors in the hypertrophic process is clear from these studies, but analysis of cardiac myocyte and/or cardiac contractile function would add to our understanding of these receptors. Gene transfer of proteins/enzymes involved in hormone metabolism has been used to demonstrate the importance of cardiac PKC-related paracrine and autocrine mediators. One example is the use of angiotensin converting enzyme 2 (ACE2), which catalyzes the production of the potent vasodilator angiotensin 1-7 (ANG1-7) (229). ANG1 7 works via a non-pkc pathway (269) and counteracts the influences of the PKC-linked AT-1 receptor on cardiac hypertrophy (179, 717, 976). Lentiviral-mediated gene transfer of ACE2 produced strong cardiac expression and renal expression, albeit to a lesser extent. Functionally, ACE2 delivery also reduced hypertension and increased left ventricular end-diastolic and end-systolic dimensions in spontaneously hypertensive but not normotensive rats (179). Although the mechanism responsible for these beneficial effects is currently not clear, results from studies like these suggest that approaches designed

49 CARDIAC MUSCLE GENE TRANSFER 1615 FIG. 11. Model of protein kinase C (PKC) signaling directly involved in modulating contractile performance in cardiac myocytes. Several agonists activate this signaling cascade. Signaling via endothelin A receptors (ET A ) and the angiotensin II receptor (ATII R ) are examples of receptors signaling through G proteins and their subsequent activation of phospholipase C (PLC). The next step in this cascade leads to the production of diacylglycerol (DAG) and inositol 1,4,5-trisphosphate (IP 3 ) from the phosphoinositol PIP 2 by phospholipase C (PLC). DAG then activates PKC, as does synthetic phorbol esters. In nonfailing hearts, the three most predominant PKC isoforms are classical PKC- and novel PKC- and -. Heart failure is associated with upregulation of PKC- and -, along with the production of PKC-. Activated PKC is translocated to target proteins, and this activation/translocation process is modulated by receptors for activated C kinase (RACKS), receptors for inhibition of C kinase (RICKS), and substrates that interact with C kinase (STICKS). A comprehensive understanding of the specific physiological, pathophysiological, and pharmacological targets for each PKC isoform, as well as the potential for cross-talk among isoforms, remains an active area for investigation. One example of a PKC- targeting pathway shown here was identified with the help of gene transfer studies. PKC- phosphorylates inhibitor-1 (I-1) and in turn activates protein phosphatase 1 (PP1). A major target for PP1 is phosphorylated phospholamban (PLN). PP1-dependent reduction in PLN phosphorylation slows Ca 2 uptake by SERCA2a and significantly decreases contractile function. Other PKC targets, including the myofilaments, directly influence cardiac contractile function upon PKC modification and are also included in this model. Targets involving more indirect and/or transcriptional effects of PKC signaling are not shown. to reduce the influence of PKC-linked signaling hormones on contractile function could be beneficial during heart failure. Other steps in the PKC signaling pathway (Fig. 11) studied using gene transfer include PLC, heat shock proteins, and DAG kinase. PLC is critical for the production of DAG, which is critical for PKC activation/translocation in response to mediators such as endothelin, ANG II, and -adrenergic agents. Overexpression of PLC isoforms in neonatal cardiac myocytes was demonstrated after viralbased gene transfer (22). Heat shock proteins can also modulate PKC (128), and viral-mediated gene transfer has been used to investigate the influence of Hsp70 and Hsp90 on PKC- and - expression in neonatal cardiac myocytes (128). DAG kinases breakdown DAG, and the breakdown of DAG in response to overexpression of DAG kinase inhibits endothelin-1-induced activation of PKC- and downstream signaling (mitogen-activated protein kinase, MAPK) of hypertrophy in neonatal cardiac myocytes (862). B) PKC ISOFORMS. The PKC family consists of at least 12 different isoforms, and the characteristics of the 3 subclasses have been reviewed in detail (173, 190, 619, 838). Adult cardiac myocytes in most mammalian species express four major isoforms, including classical PKC-, novel class PKC- and -, and to a lesser extent atypical PKC- (213, 713, 714). Several pathophysiological conditions and heart failure in humans are associated with increased expression of PKC- and -, plus the appearance of PKC-, another classical isoform (77, 648). Comparable changes in these PKC isoforms were observed in rat models of pressure overload (42, 191). The gene transfer studies described below have significantly contributed to our understanding of the role played by many of these isoforms in modulating contractile function. I) PKC-. Classical and novel PKC isoforms are activated by PLC-dependent DAG production or addition of phorbol esters (Fig. 11). The activity of classical isoforms, such as PKC-, also depend on increased cytosolic Ca 2 (173). Early studies demonstrated that peptide inhibitors

Choosing Between Lentivirus and Adeno-associated Virus For DNA Delivery

Choosing Between Lentivirus and Adeno-associated Virus For DNA Delivery Choosing Between Lentivirus and Adeno-associated Virus For DNA Delivery Presenter: April 12, 2017 Ed Davis, Ph.D. Senior Application Scientist GeneCopoeia, Inc. Outline Introduction to GeneCopoeia Lentiviral

More information

VIROLOGY. Engineering Viral Genomes: Retrovirus Vectors

VIROLOGY. Engineering Viral Genomes: Retrovirus Vectors VIROLOGY Engineering Viral Genomes: Retrovirus Vectors Viral vectors Retrovirus replicative cycle Most mammalian retroviruses use trna PRO, trna Lys3, trna Lys1,2 The partially unfolded trna is annealed

More information

This training module is required for all personnel listed on an IBC protocol that describes work utilizing viral vectors (both replication competent

This training module is required for all personnel listed on an IBC protocol that describes work utilizing viral vectors (both replication competent This training module is required for all personnel listed on an IBC protocol that describes work utilizing viral vectors (both replication competent and incompetent) regardless of the biosafety level used

More information

Lecture 2: Virology. I. Background

Lecture 2: Virology. I. Background Lecture 2: Virology I. Background A. Properties 1. Simple biological systems a. Aggregates of nucleic acids and protein 2. Non-living a. Cannot reproduce or carry out metabolic activities outside of a

More information

B19, see Parvovirus B19 Bone marrow, gene transfer with parvovirus. Erythrovirus, see Parvovirus B19, Simian parvovirus

B19, see Parvovirus B19 Bone marrow, gene transfer with parvovirus. Erythrovirus, see Parvovirus B19, Simian parvovirus ... Subject Index Adeno-associated virus Cap and genome encapsidation 87 DNA integration homologous recombination 90, 91 latency vs replication 77, 78 mechanism 79 requirements 78, 79 site in human genome

More information

Viral Vectors In The Research Laboratory: Just How Safe Are They? Dawn P. Wooley, Ph.D., SM(NRM), RBP, CBSP

Viral Vectors In The Research Laboratory: Just How Safe Are They? Dawn P. Wooley, Ph.D., SM(NRM), RBP, CBSP Viral Vectors In The Research Laboratory: Just How Safe Are They? Dawn P. Wooley, Ph.D., SM(NRM), RBP, CBSP 1 Learning Objectives Recognize hazards associated with viral vectors in research and animal

More information

Recombinant Protein Expression Retroviral system

Recombinant Protein Expression Retroviral system Recombinant Protein Expression Retroviral system Viruses Contains genome DNA or RNA Genome encased in a protein coat or capsid. Some viruses have membrane covering protein coat enveloped virus Ø Essential

More information

Feb 11, Gene Therapy. Sam K.P. Kung Immunology Rm 417 Apotex Center

Feb 11, Gene Therapy. Sam K.P. Kung Immunology Rm 417 Apotex Center Gene Therapy Sam K.P. Kung Immunology Rm 417 Apotex Center Objectives: The concept of gene therapy, and an introduction of some of the currently used gene therapy vector Undesirable immune responses to

More information

VIRUSES. 1. Describe the structure of a virus by completing the following chart.

VIRUSES. 1. Describe the structure of a virus by completing the following chart. AP BIOLOGY MOLECULAR GENETICS ACTIVITY #3 NAME DATE HOUR VIRUSES 1. Describe the structure of a virus by completing the following chart. Viral Part Description of Part 2. Some viruses have an envelope

More information

Virology Introduction. Definitions. Introduction. Structure of virus. Virus transmission. Classification of virus. DNA Virus. RNA Virus. Treatment.

Virology Introduction. Definitions. Introduction. Structure of virus. Virus transmission. Classification of virus. DNA Virus. RNA Virus. Treatment. DEVH Virology Introduction Definitions. Introduction. Structure of virus. Virus transmission. Classification of virus. DNA Virus. RNA Virus. Treatment. Definitions Virology: The science which study the

More information

Polyomaviridae. Spring

Polyomaviridae. Spring Polyomaviridae Spring 2002 331 Antibody Prevalence for BK & JC Viruses Spring 2002 332 Polyoma Viruses General characteristics Papovaviridae: PA - papilloma; PO - polyoma; VA - vacuolating agent a. 45nm

More information

OCCUPATIONAL HEALTH CONSIDERATIONS FOR WORK WITH VIRAL VECTORS

OCCUPATIONAL HEALTH CONSIDERATIONS FOR WORK WITH VIRAL VECTORS OCCUPATIONAL HEALTH CONSIDERATIONS FOR WORK WITH VIRAL VECTORS GARY R. FUJIMOTO, M.D. PALO ALTO MEDICAL FOUNDATION ADJUNCT ASSOCIATE CLINICAL PROFESSOR OF MEDICINE DIVISION OF INFECTIOUS DISEASES AND GEOGRAPHIC

More information

Module 2 In vivo gene therapy. Lecture 7. In-situ, in-vivo and ex-vivo gene therapy (part I)

Module 2 In vivo gene therapy. Lecture 7. In-situ, in-vivo and ex-vivo gene therapy (part I) Module 2 In vivo gene therapy Lecture 7 In-situ, in-vivo and ex-vivo gene therapy (part I) Somatic cell gene therapy involves the transfer of gene to a diseased somatic cell either within the body or outside

More information

7.012 Problem Set 6 Solutions

7.012 Problem Set 6 Solutions Name Section 7.012 Problem Set 6 Solutions Question 1 The viral family Orthomyxoviridae contains the influenza A, B and C viruses. These viruses have a (-)ss RNA genome surrounded by a capsid composed

More information

Jumpstart your research with ViraPower Lentiviral Expression Systems

Jumpstart your research with ViraPower Lentiviral Expression Systems ViraPower Lentiviral Expression Systems Jumpstart your research with ViraPower Lentiviral Expression Systems With ViraPower Lentiviral Systems you can: Efficiently transduce both dividing and non-dividing

More information

Chapter 6- An Introduction to Viruses*

Chapter 6- An Introduction to Viruses* Chapter 6- An Introduction to Viruses* *Lecture notes are to be used as a study guide only and do not represent the comprehensive information you will need to know for the exams. 6.1 Overview of Viruses

More information

Zachary I. Hodes, M.D., Ph.D., F.A.C.C.

Zachary I. Hodes, M.D., Ph.D., F.A.C.C. Zachary I. Hodes, M.D., Ph.D., F.A.C.C. Disclamer: I personally have no financial relationship with any company mentioned today. The Care Group, LLC does have a contract with Cardium to participate in

More information

Chapter 19: Viruses. 1. Viral Structure & Reproduction. 2. Bacteriophages. 3. Animal Viruses. 4. Viroids & Prions

Chapter 19: Viruses. 1. Viral Structure & Reproduction. 2. Bacteriophages. 3. Animal Viruses. 4. Viroids & Prions Chapter 19: Viruses 1. Viral Structure & Reproduction 2. Bacteriophages 3. Animal Viruses 4. Viroids & Prions 1. Viral Structure & Reproduction Chapter Reading pp. 393-396 What exactly is a Virus? Viruses

More information

~Lentivirus production~

~Lentivirus production~ ~Lentivirus production~ May 30, 2008 RNAi core R&D group member Lentivirus Production Session Lentivirus!!! Is it health threatening to lab technician? What s so good about this RNAi library? How to produce

More information

19/06/2013. Viruses are not organisms (do not belong to any kingdom). Viruses are not made of cells, have no cytoplasm, and no membranes.

19/06/2013. Viruses are not organisms (do not belong to any kingdom). Viruses are not made of cells, have no cytoplasm, and no membranes. VIRUSES Many diseases of plants and animals are caused by bacteria or viruses that invade the body. Bacteria and viruses are NOT similar kinds of micro-organisms. Bacteria are classified as living organisms,

More information

GENE THERAPY: Twenty-First Century Medicine

GENE THERAPY: Twenty-First Century Medicine Annu. Rev. Biochem. 2005. 74:711 38 doi: 10.1146/annurev.biochem.74.050304.091637 Copyright c 2005 by Annual Reviews. All rights reserved First published online as a Review in Advance on March 11, 2005

More information

Chapter 19: Viruses. 1. Viral Structure & Reproduction. What exactly is a Virus? 11/7/ Viral Structure & Reproduction. 2.

Chapter 19: Viruses. 1. Viral Structure & Reproduction. What exactly is a Virus? 11/7/ Viral Structure & Reproduction. 2. Chapter 19: Viruses 1. Viral Structure & Reproduction 2. Bacteriophages 3. Animal Viruses 4. Viroids & Prions 1. Viral Structure & Reproduction Chapter Reading pp. 393-396 What exactly is a Virus? Viruses

More information

Herpesviruses. Virion. Genome. Genes and proteins. Viruses and hosts. Diseases. Distinctive characteristics

Herpesviruses. Virion. Genome. Genes and proteins. Viruses and hosts. Diseases. Distinctive characteristics Herpesviruses Virion Genome Genes and proteins Viruses and hosts Diseases Distinctive characteristics Virion Enveloped icosahedral capsid (T=16), diameter 125 nm Diameter of enveloped virion 200 nm Capsid

More information

11/15/2011. Outline. Structural Features and Characteristics. The Good the Bad and the Ugly. Viral Genomes. Structural Features and Characteristics

11/15/2011. Outline. Structural Features and Characteristics. The Good the Bad and the Ugly. Viral Genomes. Structural Features and Characteristics Chapter 19 - Viruses Outline I. Viruses A. Structure of viruses B. Common Characteristics of Viruses C. Viral replication D. HIV II. Prions The Good the Bad and the Ugly Viruses fit into the bad category

More information

Replication Defective Enterovirus Infections: Implications for Type I Diabetes

Replication Defective Enterovirus Infections: Implications for Type I Diabetes Replication Defective Enterovirus Infections: Implications for Type I Diabetes N. M. Chapman Department of Pathology & Microbiology University of Nebraska Medical Center Enterovirus Genome and 2 Capsid

More information

7.012 Quiz 3 Answers

7.012 Quiz 3 Answers MIT Biology Department 7.012: Introductory Biology - Fall 2004 Instructors: Professor Eric Lander, Professor Robert A. Weinberg, Dr. Claudette Gardel Friday 11/12/04 7.012 Quiz 3 Answers A > 85 B 72-84

More information

Fayth K. Yoshimura, Ph.D. September 7, of 7 HIV - BASIC PROPERTIES

Fayth K. Yoshimura, Ph.D. September 7, of 7 HIV - BASIC PROPERTIES 1 of 7 I. Viral Origin. A. Retrovirus - animal lentiviruses. HIV - BASIC PROPERTIES 1. HIV is a member of the Retrovirus family and more specifically it is a member of the Lentivirus genus of this family.

More information

Viral structure م.م رنا مشعل

Viral structure م.م رنا مشعل Viral structure م.م رنا مشعل Viruses must reproduce (replicate) within cells, because they cannot generate energy or synthesize proteins. Because they can reproduce only within cells, viruses are obligate

More information

Introduction. Cancer Biology. Tumor-suppressor genes. Proto-oncogenes. DNA stability genes. Mechanisms of carcinogenesis.

Introduction. Cancer Biology. Tumor-suppressor genes. Proto-oncogenes. DNA stability genes. Mechanisms of carcinogenesis. Cancer Biology Chapter 18 Eric J. Hall., Amato Giaccia, Radiobiology for the Radiologist Introduction Tissue homeostasis depends on the regulated cell division and self-elimination (programmed cell death)

More information

Dr. Ahmed K. Ali Attachment and entry of viruses into cells

Dr. Ahmed K. Ali Attachment and entry of viruses into cells Lec. 6 Dr. Ahmed K. Ali Attachment and entry of viruses into cells The aim of a virus is to replicate itself, and in order to achieve this aim it needs to enter a host cell, make copies of itself and

More information

Muscular Dystrophy. Biol 405 Molecular Medicine

Muscular Dystrophy. Biol 405 Molecular Medicine Muscular Dystrophy Biol 405 Molecular Medicine Duchenne muscular dystrophy Duchenne muscular dystrophy is a neuromuscular disease that occurs in ~ 1/3,500 male births. The disease causes developmental

More information

Name Section Problem Set 6

Name Section Problem Set 6 Name Section 7.012 Problem Set 6 Question 1 The viral family Orthomyxoviridae contains the influenza A, B and C viruses. These viruses have a (-)ss RNA genome surrounded by a capsid composed of lipids

More information

numbe r Done by Corrected by Doctor

numbe r Done by Corrected by Doctor numbe r 5 Done by Mustafa Khader Corrected by Mahdi Sharawi Doctor Ashraf Khasawneh Viral Replication Mechanisms: (Protein Synthesis) 1. Monocistronic Method: All human cells practice the monocistronic

More information

VIRAL TITER COUNTS. The best methods of measuring infectious lentiviral titer

VIRAL TITER COUNTS. The best methods of measuring infectious lentiviral titer VIRAL TITER COUNTS The best methods of measuring infectious lentiviral titer FLUORESCENCE CYCLES qpcr of Viral RNA SUMMARY Viral vectors are now routinely used for gene transduction in a wide variety of

More information

Current Strategies in HIV-1 Vaccine Development Using Replication-Defective Adenovirus as a Case Study

Current Strategies in HIV-1 Vaccine Development Using Replication-Defective Adenovirus as a Case Study Note: I have added some clarifying comments to the slides -- please click on Comments under View to see them. Current Strategies in HIV-1 Vaccine Development Using Replication-Defective Adenovirus as a

More information

19 Viruses BIOLOGY. Outline. Structural Features and Characteristics. The Good the Bad and the Ugly. Structural Features and Characteristics

19 Viruses BIOLOGY. Outline. Structural Features and Characteristics. The Good the Bad and the Ugly. Structural Features and Characteristics 9 Viruses CAMPBELL BIOLOGY TENTH EDITION Reece Urry Cain Wasserman Minorsky Jackson Outline I. Viruses A. Structure of viruses B. Common Characteristics of Viruses C. Viral replication D. HIV Lecture Presentation

More information

Physiology sheet #2. The heart composed of 3 layers that line its lumen and cover it from out side, these layers are :

Physiology sheet #2. The heart composed of 3 layers that line its lumen and cover it from out side, these layers are : Physiology sheet #2 * We will talk in this lecture about cardiac muscle physiology, the mechanism and the energy sources of their contraction and intracellular calcium homeostasis. # Slide 4 : The heart

More information

Chapter 18. Viral Genetics. AP Biology

Chapter 18. Viral Genetics. AP Biology Chapter 18. Viral Genetics 2003-2004 1 A sense of size Comparing eukaryote bacterium virus 2 What is a virus? Is it alive? DNA or RNA enclosed in a protein coat Viruses are not cells Extremely tiny electron

More information

1. Virus 2. Capsid 3. Envelope

1. Virus 2. Capsid 3. Envelope VIRUSES BIOLOGY II VOCABULARY- VIRUSES (22 Words) 1. Virus 2. Capsid 3. Envelope 4. Provirus 5. Retrovirus 6. Reverse transcriptase 7. Bacteriophage 8. Lytic Cycle 9. Virulent 10. Lysis 11. Lysogenic Cycle

More information

Viruses defined acellular organisms genomes nucleic acid replicate inside host cells host metabolic machinery ribosomes

Viruses defined acellular organisms genomes nucleic acid replicate inside host cells host metabolic machinery ribosomes The Viruses Viruses Viruses may be defined as acellular organisms whose genomes consist of nucleic acid, obligately replicate inside host cells using host metabolic machinery and ribosomes to form a pool

More information

LESSON 1.4 WORKBOOK. Viral sizes and structures. Workbook Lesson 1.4

LESSON 1.4 WORKBOOK. Viral sizes and structures. Workbook Lesson 1.4 Eukaryotes organisms that contain a membrane bound nucleus and organelles. Prokaryotes organisms that lack a nucleus or other membrane-bound organelles. Viruses small, non-cellular (lacking a cell), infectious

More information

Introduction retroposon

Introduction retroposon 17.1 - Introduction A retrovirus is an RNA virus able to convert its sequence into DNA by reverse transcription A retroposon (retrotransposon) is a transposon that mobilizes via an RNA form; the DNA element

More information

CHAPTER 6 2/9/2016. Learning Objectives List the four traits that all muscle types have in common.

CHAPTER 6 2/9/2016. Learning Objectives List the four traits that all muscle types have in common. Learning Objectives List the four traits that all muscle types have in common. CHAPTER 6 The Muscular System Demonstrate and explain the use of antagonistic muscle pairs. Describe the attachment of muscle

More information

Viruses Tomasz Kordula, Ph.D.

Viruses Tomasz Kordula, Ph.D. Viruses Tomasz Kordula, Ph.D. Resources: Alberts et al., Molecular Biology of the Cell, pp. 295, 1330, 1431 1433; Lehninger CD Movie A0002201. Learning Objectives: 1. Understand parasitic life cycle of

More information

LESSON 4.4 WORKBOOK. How viruses make us sick: Viral Replication

LESSON 4.4 WORKBOOK. How viruses make us sick: Viral Replication DEFINITIONS OF TERMS Eukaryotic: Non-bacterial cell type (bacteria are prokaryotes).. LESSON 4.4 WORKBOOK How viruses make us sick: Viral Replication This lesson extends the principles we learned in Unit

More information

AP Biology. Viral diseases Polio. Chapter 18. Smallpox. Influenza: 1918 epidemic. Emerging viruses. A sense of size

AP Biology. Viral diseases Polio. Chapter 18. Smallpox. Influenza: 1918 epidemic. Emerging viruses. A sense of size Hepatitis Viral diseases Polio Chapter 18. Measles Viral Genetics Influenza: 1918 epidemic 30-40 million deaths world-wide Chicken pox Smallpox Eradicated in 1976 vaccinations ceased in 1980 at risk population?

More information

Hepadnaviruses: Variations on the Retrovirus Theme

Hepadnaviruses: Variations on the Retrovirus Theme WBV21 6/27/03 11:34 PM Page 377 Hepadnaviruses: Variations on the Retrovirus Theme 21 CHAPTER The virion and the viral genome The viral replication cycle The pathogenesis of hepatitis B virus A plant hepadnavirus

More information

Supplementary Information. Supplementary Figure 1

Supplementary Information. Supplementary Figure 1 Supplementary Information Supplementary Figure 1 1 Supplementary Figure 1. Functional assay of the hcas9-2a-mcherry construct (a) Gene correction of a mutant EGFP reporter cell line mediated by hcas9 or

More information

Virus and Prokaryotic Gene Regulation - 1

Virus and Prokaryotic Gene Regulation - 1 Virus and Prokaryotic Gene Regulation - 1 We have discussed the molecular structure of DNA and its function in DNA duplication and in transcription and protein synthesis. We now turn to how cells regulate

More information

STRUCTURE, GENERAL CHARACTERISTICS AND REPRODUCTION OF VIRUSES

STRUCTURE, GENERAL CHARACTERISTICS AND REPRODUCTION OF VIRUSES STRUCTURE, GENERAL CHARACTERISTICS AND REPRODUCTION OF VIRUSES Introduction Viruses are noncellular genetic elements that use a living cell for their replication and have an extracellular state. Viruses

More information

Certificate of Analysis

Certificate of Analysis Certificate of Analysis Catalog No. Amount Lot Number 631987 10 μg Specified on product label. Product Information plvx-ef1α-ires-mcherry is a bicistronic lentiviral expression vector that can be used

More information

Julianne Edwards. Retroviruses. Spring 2010

Julianne Edwards. Retroviruses. Spring 2010 Retroviruses Spring 2010 A retrovirus can simply be referred to as an infectious particle which replicates backwards even though there are many different types of retroviruses. More specifically, a retrovirus

More information

Retroviruses. ---The name retrovirus comes from the enzyme, reverse transcriptase.

Retroviruses. ---The name retrovirus comes from the enzyme, reverse transcriptase. Retroviruses ---The name retrovirus comes from the enzyme, reverse transcriptase. ---Reverse transcriptase (RT) converts the RNA genome present in the virus particle into DNA. ---RT discovered in 1970.

More information

Date. Student Name. Prompt: This passage is called Characteristics of Viruses. It is about viruses.

Date. Student Name. Prompt: This passage is called Characteristics of Viruses. It is about viruses. Student Name Characteristics of Viruses--Part I Level High School - Science Date _ Prompt: This passage is called Characteristics of Viruses. It is about viruses. Similarities and Differences Between Viruses

More information

Muscular Tissue. Functions of Muscular Tissue. Types of Muscular Tissue. Skeletal Muscular Tissue. Properties of Muscular Tissue

Muscular Tissue. Functions of Muscular Tissue. Types of Muscular Tissue. Skeletal Muscular Tissue. Properties of Muscular Tissue Muscular Tissue Functions of Muscular Tissue Muscle makes up a large percentage of the body s weight (40-50%) Their main functions are to: Create motion muscles work with nerves, bones, and joints to produce

More information

MedChem 401~ Retroviridae. Retroviridae

MedChem 401~ Retroviridae. Retroviridae MedChem 401~ Retroviridae Retroviruses plus-sense RNA genome (!8-10 kb) protein capsid lipid envelop envelope glycoproteins reverse transcriptase enzyme integrase enzyme protease enzyme Retroviridae The

More information

Reoviruses. Virion. Genome. Genes and proteins. Viruses and hosts. Diseases. Distinctive characteristics

Reoviruses. Virion. Genome. Genes and proteins. Viruses and hosts. Diseases. Distinctive characteristics Reoviruses Virion Genome Genes and proteins Viruses and hosts Diseases Distinctive characteristics Virion Naked icosahedral capsid (T=13), diameter 60-85 nm Capsid consists of two or three concentric protein

More information

Choosing Optimal Viral Vector for T-cell Transduction. Viral vectors for blood cells

Choosing Optimal Viral Vector for T-cell Transduction. Viral vectors for blood cells Choosing Optimal Viral Vector for T-cell Transduction Max Mamonkin, PhD Center for Cell and Gene Therapy Baylor College of Medicine PACT Webinar Nov 08, 2018 Viral for blood cells Short/long term gene

More information

Viral vaccines. Lec. 3 أ.د.فائزة عبد هللا مخلص

Viral vaccines. Lec. 3 أ.د.فائزة عبد هللا مخلص Lec. 3 أ.د.فائزة عبد هللا مخلص Viral vaccines 0bjectives 1-Define active immunity. 2-Describe the methods used for the preparation of attenuated live & killed virus vaccines. 3- Comparison of Characteristics

More information

Dr. Gary Mumaugh. Viruses

Dr. Gary Mumaugh. Viruses Dr. Gary Mumaugh Viruses Viruses in History In 1898, Friedrich Loeffler and Paul Frosch found evidence that the cause of foot-and-mouth disease in livestock was an infectious particle smaller than any

More information

7.L.1.4 Circulatory System Guided Study Notes. Circulation

7.L.1.4 Circulatory System Guided Study Notes. Circulation 1 7.L.1.4 Circulatory System Guided Study Notes Circulation Sect. 1: The Body s Transport System Sect. 2: A Closer Look at Blood Vessels Sect. 3: Blood and Lymph Sect. 4: Cardiovascular Health Sect. 1:

More information

Oxford Expression Technologies Ltd

Oxford Expression Technologies Ltd Oxford Expression Technologies Ltd Founded in 2007 as a spin out from Oxford Brookes University and Natural Environment Research Council Technology based on the insect baculovirus expression vectors (BEVs)

More information

3D Tissue Models. Simple, Low Cost Fabrication. Simple, Robust Protocols

3D Tissue Models. Simple, Low Cost Fabrication. Simple, Robust Protocols 3D Tissue Models SynVivo is a physiological, cell-based microfluidic platform that provides a morphologically and physiologically realistic microenvironment allowing real-time study of cellular behavior,

More information

Pre-made Reporter Lentivirus for JAK-STAT Signaling Pathway

Pre-made Reporter Lentivirus for JAK-STAT Signaling Pathway Pre-made Reporter for JAK-STAT Signaling Pathway Cat# Product Name Amounts LVP937-P or: LVP937-P-PBS ISRE-GFP (Puro) LVP938-P or: LVP938-P-PBS ISRE-RFP (Puro) LVP939-P or: LVP939-P-PBS ISRE-Luc (Puro)

More information

Chapter 14 Part One Biotechnology and Industry: Microbes at Work

Chapter 14 Part One Biotechnology and Industry: Microbes at Work Chapter 14 Part One Biotechnology and Industry: Microbes at Work Objectives: After reading Chapter 14, you should understand How biotechnology has resulted in numerous pharmaceutical products to help lessen

More information

Experimental Therapeutics I

Experimental Therapeutics I Experimental Therapeutics I Mary Hitt 5142 Katz Group Centre mhitt@ualberta.ca; or Mary.Hitt@albertahealthservices.ca 1 Specific Topics for Today Preclinical and clinical testing Gene therapy Nonviral

More information

Stretching Cardiac Myocytes: A Finite Element Model of Cardiac Tissue

Stretching Cardiac Myocytes: A Finite Element Model of Cardiac Tissue Megan McCain ES240 FEM Final Project December 19, 2006 Stretching Cardiac Myocytes: A Finite Element Model of Cardiac Tissue Cardiac myocytes are the cells that constitute the working muscle of the heart.

More information

Chapter13 Characterizing and Classifying Viruses, Viroids, and Prions

Chapter13 Characterizing and Classifying Viruses, Viroids, and Prions Chapter13 Characterizing and Classifying Viruses, Viroids, and Prions 11/20/2017 MDufilho 1 Characteristics of Viruses Viruses Minuscule, acellular, infectious agent having either DNA or RNA Cause infections

More information

Pre-made Reporter Lentivirus for MAPK/ERK Signal Pathway

Pre-made Reporter Lentivirus for MAPK/ERK Signal Pathway Pre-made Reporter for MAPK/ERK Signal Pathway Cat# Product Name Amounts LVP957-P or: LVP957-P-PBS SRE-GFP (Puro) LVP958-P or: LVP958-P-PBS SRE-RFP (Puro) LVP959-P or: LVP959-P-PBS SRE-Luc (Puro) LVP960-P

More information

Gene Vaccine Dr. Sina Soleimani

Gene Vaccine Dr. Sina Soleimani Gene Vaccine Dr. Sina Soleimani Human Viral Vaccines Quality Control Laboratory (HVVQC) Titles 1. A short Introduction of Vaccine History 2. First Lineage of Vaccines 3. Second Lineage of Vaccines 3. New

More information

In the name of GOD. Animal models of cardiovascular diseases: myocardial infarction & hypertension

In the name of GOD. Animal models of cardiovascular diseases: myocardial infarction & hypertension In the name of GOD Animal models of cardiovascular diseases: myocardial infarction & hypertension 44 Presentation outline: Cardiovascular diseases Acute myocardial infarction Animal models for myocardial

More information

LESSON 4.6 WORKBOOK. Designing an antiviral drug The challenge of HIV

LESSON 4.6 WORKBOOK. Designing an antiviral drug The challenge of HIV LESSON 4.6 WORKBOOK Designing an antiviral drug The challenge of HIV In the last two lessons we discussed the how the viral life cycle causes host cell damage. But is there anything we can do to prevent

More information

INTERNATIONAL CONFERENCE ON HARMONISATION OF TECHNICAL REQUIREMENTS FOR REGISTRATION OF PHARMACEUTICALS FOR HUMAN USE. ICH Considerations

INTERNATIONAL CONFERENCE ON HARMONISATION OF TECHNICAL REQUIREMENTS FOR REGISTRATION OF PHARMACEUTICALS FOR HUMAN USE. ICH Considerations INTERNATIONAL CONFERENCE ON HARMONISATION OF TECHNICAL REQUIREMENTS FOR REGISTRATION OF PHARMACEUTICALS FOR HUMAN USE ICH Considerations General Principles to Address Virus and Vector Shedding 1.0 Introduction

More information

Fayth K. Yoshimura, Ph.D. September 7, of 7 RETROVIRUSES. 2. HTLV-II causes hairy T-cell leukemia

Fayth K. Yoshimura, Ph.D. September 7, of 7 RETROVIRUSES. 2. HTLV-II causes hairy T-cell leukemia 1 of 7 I. Diseases Caused by Retroviruses RETROVIRUSES A. Human retroviruses that cause cancers 1. HTLV-I causes adult T-cell leukemia and tropical spastic paraparesis 2. HTLV-II causes hairy T-cell leukemia

More information

Anonymous MIT students MK and NM Natalie Kuldell May 2009

Anonymous MIT students MK and NM Natalie Kuldell May 2009 Anonymous MIT students MK and NM Natalie Kuldell 20.020 13 May 2009 Technical Document: oncocures Problem: Dynamic Metastasis Imaging Our project has been designed with the hope of creating a way to detect

More information

Some living things are made of ONE cell, and are called. Other organisms are composed of many cells, and are called. (SEE PAGE 6)

Some living things are made of ONE cell, and are called. Other organisms are composed of many cells, and are called. (SEE PAGE 6) Section: 1.1 Question of the Day: Name: Review of Old Information: N/A New Information: We tend to only think of animals as living. However, there is a great diversity of organisms that we consider living

More information

Human Immunodeficiency Virus. Acquired Immune Deficiency Syndrome AIDS

Human Immunodeficiency Virus. Acquired Immune Deficiency Syndrome AIDS Human Immunodeficiency Virus Acquired Immune Deficiency Syndrome AIDS Sudden outbreak in USA of opportunistic infections and cancers in young men in 1981 Pneumocystis carinii pneumonia (PCP), Kaposi s

More information

A virus consists of a nucleic acid surrounded by a protein coat. [2]

A virus consists of a nucleic acid surrounded by a protein coat. [2] GUIDED READING - Ch. 19 - VIRUSES NAME: Please print out these pages and HANDWRITE the answers directly on the printouts. Typed work or answers on separate sheets of paper will not be accepted. Importantly,

More information

Chapter 6: Summary

Chapter 6: Summary Chapter 6: Summary The present thesis describes several clinical studies on the topic of intravenous myocardial contrast echocardiography, and an experimental study on the effect of ultrasound contrast

More information

Chapter 9. Body Fluid Compartments. Body Fluid Compartments. Blood Volume. Blood Volume. Viscosity. Circulatory Adaptations to Exercise Part 4

Chapter 9. Body Fluid Compartments. Body Fluid Compartments. Blood Volume. Blood Volume. Viscosity. Circulatory Adaptations to Exercise Part 4 Body Fluid Compartments Chapter 9 Circulatory Adaptations to Exercise Part 4 Total body fluids (40 L) Intracellular fluid (ICF) 25 L Fluid of each cell (75 trillion) Constituents inside cell vary Extracellular

More information

Hepatitis B Antiviral Drug Development Multi-Marker Screening Assay

Hepatitis B Antiviral Drug Development Multi-Marker Screening Assay Hepatitis B Antiviral Drug Development Multi-Marker Screening Assay Background ImQuest BioSciences has developed and qualified a single-plate method to expedite the screening of antiviral agents against

More information

Lahore University of Management Sciences. BIO314 Virology and Microbiology (Spring 2015)

Lahore University of Management Sciences. BIO314 Virology and Microbiology (Spring 2015) BIO314 Virology and Microbiology (Spring 2015) Instructor Room. Office Hours Email Telephone Secretary/TA TA Office Hours Course URL (if any) Shaper Mirza and Sadia Hamera Shaper.Mirza@uth.tmc.edu Course

More information

Cardiac Gene Therapy: Beyond the Mouse. David M Kaye Heart Failure Research Group Baker IDI, Melbourne, AUSTRALIA

Cardiac Gene Therapy: Beyond the Mouse. David M Kaye Heart Failure Research Group Baker IDI, Melbourne, AUSTRALIA Cardiac Gene Therapy: Beyond the Mouse David M Kaye Heart Failure Research Group Baker IDI, Melbourne, AUSTRALIA Presenter Disclosure Information FINANCIAL DISCLOSURE: Equity: Osprey Medical Grants/Research

More information

Bacteriophage Reproduction

Bacteriophage Reproduction Bacteriophage Reproduction Lytic and Lysogenic Cycles The following information is taken from: http://student.ccbcmd.edu/courses/bio141/lecguide/unit3/index.html#charvir Bacteriophage Structure More complex

More information

As a courtesy to your fellow classmates please refrain from talking, beating, or snoring. And Now Our Feature Presentation.

As a courtesy to your fellow classmates please refrain from talking, beating, or snoring. And Now Our Feature Presentation. As a courtesy to your fellow classmates please refrain from talking, beating, or snoring. And Now Our Feature Presentation. Circulation Sect. 1: The Body s Transport System Sect. 2: A Closer Look at Blood

More information

Marah karablieh. Osama khader. Muhammad khatatbeh. 0 P a g e

Marah karablieh. Osama khader. Muhammad khatatbeh. 0 P a g e 15 Marah karablieh Osama khader 0 P a g e Muhammad khatatbeh Cardiac Muscle Physiology Introduction The heart has two ventricles and two atriums. The heart wall is composed primarily of spirally arranged

More information

Foot and Mouth Disease Vaccine Research and Development in India

Foot and Mouth Disease Vaccine Research and Development in India Foot and Mouth Disease Vaccine Research and Development in India R.Venkataramanan Indian Veterinary Research Institute, Hebbal, Bangalore 560 024 Foot and Mouth Disease in India Present Status Large Susceptible

More information

Pre-made Reporter Lentivirus for NF-κB Signal Pathway

Pre-made Reporter Lentivirus for NF-κB Signal Pathway Pre-made Reporter for NF-κB Signal Pathway Cat# Product Name Amounts LVP965-P or: LVP965-P-PBS NFKB-GFP (Puro) LVP966-P or: LVP966-P-PBS NFKB-RFP (Puro) LVP967-P or: LVP967-P-PBS NFKB-Luc (Puro) LVP968-P

More information

Virology. *Viruses can be only observed by electron microscope never by light microscope. The size of the virus: nm in diameter.

Virology. *Viruses can be only observed by electron microscope never by light microscope. The size of the virus: nm in diameter. Virology We are going to start with general introduction about viruses, they are everywhere around us; in food; within the environment; in direct contact to etc.. They may cause viral infection by itself

More information

Chapter 13B: Animal Viruses

Chapter 13B: Animal Viruses Chapter 13B: Animal Viruses 1. Overview of Animal Viruses 2. DNA Viruses 3. RNA Viruses 4. Prions 1. Overview of Animal Viruses Life Cycle of Animal Viruses The basic life cycle stages of animal viruses

More information

Viral reproductive cycle

Viral reproductive cycle Lecture 29: Viruses Lecture outline 11/11/05 Types of viruses Bacteriophage Lytic and lysogenic life cycles viruses viruses Influenza Prions Mad cow disease 0.5 µm Figure 18.4 Viral structure of capsid

More information

LESSON 1.4 WORKBOOK. Viral structures. Just how small are viruses? Workbook Lesson 1.4 1

LESSON 1.4 WORKBOOK. Viral structures. Just how small are viruses? Workbook Lesson 1.4 1 Eukaryotes- organisms that contain a membrane bound nucleus and organelles Prokaryotes- organisms that lack a nucleus or other membrane-bound organelles Viruses-small acellular (lacking a cell) infectious

More information

Skeletal Muscle and the Molecular Basis of Contraction. Lanny Shulman, O.D., Ph.D. University of Houston College of Optometry

Skeletal Muscle and the Molecular Basis of Contraction. Lanny Shulman, O.D., Ph.D. University of Houston College of Optometry Skeletal Muscle and the Molecular Basis of Contraction Lanny Shulman, O.D., Ph.D. University of Houston College of Optometry Like neurons, all muscle cells can be excited chemically, electrically, and

More information

ווקטורים ויראליים בכליאה ביולוגית ד"ר דליה זגר ראש חידת הבטיחות ממונת בטיחות ביולוגית

ווקטורים ויראליים בכליאה ביולוגית דר דליה זגר ראש חידת הבטיחות ממונת בטיחות ביולוגית ווקטורים ויראליים בכליאה ביולוגית ד"ר דליה זגר ראש חידת הבטיחות ממונת בטיחות ביולוגית Retroviruses הגנום שלהם רנ"א חד גדילי לאחר ההדבקה מתועתק הרנ"א הנגיפי לדנ"א ע"י RT משלבים את הגנום שלהם בדנ"א של התא

More information

Ongoing preclinical studies are providing a sound scientific

Ongoing preclinical studies are providing a sound scientific Review Gene Therapy for Heart Failure Lisa Tilemann, Kiyotake Ishikawa, Thomas Weber, Roger J. Hajjar Abstract: Congestive heart failure accounts for half a million deaths per year in the United States.

More information

Hepatitis virus immunity. Mar 9, 2005 Rehermann and Nascimbeni review Crispe review

Hepatitis virus immunity. Mar 9, 2005 Rehermann and Nascimbeni review Crispe review Hepatitis virus immunity Mar 9, 2005 Rehermann and Nascimbeni review Crispe review HBV & HCV infection outcomes Both viruses cause immune-mediated active and chronic hepatitis HBV Vertical transmission

More information

Page 32 AP Biology: 2013 Exam Review CONCEPT 6 REGULATION

Page 32 AP Biology: 2013 Exam Review CONCEPT 6 REGULATION Page 32 AP Biology: 2013 Exam Review CONCEPT 6 REGULATION 1. Feedback a. Negative feedback mechanisms maintain dynamic homeostasis for a particular condition (variable) by regulating physiological processes,

More information

Viruses. Rotavirus (causes stomach flu) HIV virus

Viruses. Rotavirus (causes stomach flu) HIV virus Viruses Rotavirus (causes stomach flu) HIV virus What is a virus? A virus is a microscopic, infectious agent that may infect any type of living cell. Viruses must infect living cells in order to make more

More information

Introduction to viruses. BIO 370 Ramos

Introduction to viruses. BIO 370 Ramos Introduction to viruses BIO 370 Ramos 1 2 General Structure of Viruses Size range most

More information

Lahore University of Management Sciences. BIO 314- Microbiology and Virology (Spring 2018)

Lahore University of Management Sciences. BIO 314- Microbiology and Virology (Spring 2018) BIO 314- Microbiology and Virology (Spring 2018) Instructor Shaper Mirza Room No. 9-318A Office Hours TBA Email Shaper.Mirza@uth.tmc.edu ; shaper.mirza@lums.edu.pk Telephone 8413 Secretary/TA No TA Office

More information