16
PHYSIOLOGICAL REVIEWS Vol. 71, No. 2, April 1991 Printed in U.S.A. Mechanical Properties of Isolated Cardiac Myocytes ALLAN J. BRADY Department of Physiology and Cardiovascular Research Laboratory, University of California, Los Angeles, School of Medicine, Center for Health Sciences, Los Angeles, Califwnia I. Introduction ........................................................................................... 413 II. Development of Adult Myocyte Preparation ......................................................... 414 III. Mechanical Properties of Cardiac Myocytes .......................................................... 414 A. Attachment procedures ............................................................................ 414 B. Passive mechanical properties .................................................................... 415 C. Contractile responses in active cardiac myocytes ................................................. 418 IV. Indexes of Contractility ............................................................................... 420 A. Shortening as an index of contractility ........................................................... 420 B. Shortening related to Frank-Starling relation ................................................... 421 V. Summary and Conclusions ............................................................................ 424 I. INTRODUCTION Viewed from the perspective of organ function, there has been a long-standing interest in the separa- bility of cardiac muscle length or ventricular volume- dependent effects (i.e., the Frank-Starling relation) from intrinsic inotropic mechanisms in the cardiac con- tractile response. Furthermore, at the cellular level there is continued interest in the mechanisms of passive and active force generation in the sarcomere. In fact, the resolution of both levels of interest has been severely limited in studies over the past decades by the complex- ity of multicellular cardiac preparations and our inabil- ity to measure or control a minimal number of the con- tractile-related parameters. Actually, the results of many of these studies strongly question the separability of geometrical and intrinsic factors of cardiac contrac- tion (12, 71, 97, 99), but the issues are by no means closed. For example, attempts to deduce cross-bridge me- chanical kinetics from multicellular preparations (in- cluding geometrically linear preparations such as tra- beculae and papillary muscles) have led to ambiguous results because of the viscoelastic nature of these prepa- rations (93) and because no clear representation of the relation between passive and active elements of the tis- sue is evident (20). On the one hand, for simplicity, rest- ing mechanical properties of these preparations have been assumed to be in parallel with active properties. This arrangement has been represented as a Maxwell mechanical analogue. Then, again, series elasticity in excised preparations has been largely attributed to end compliance of the preparations that is related to the attachment of recording devices so that a Voigt ana- logue is suggested. Finally, tests of these assumptions reveal a consid- erable complexity of stress and strain distributions at the sarcomere level, reflecting significant structural and functional nonuniformity in both passive and ac- tively expressed responses. Because contractile function is strongly sarcomere length and cross-bridge shorten- ing dependent, this inhomogeneity further reduces the reliability of interpretations of cross-bridge function based on these global or whole tissue measurements. Added to these concerns is the growing evidence that some phase of the contractile activation process may be diminished at short sarcomere length or with the shortening necessary to achieve sarcomere lengths below unattached rest length. Hence it becomes even more critical to know or to control sarcomere length at the sarcomere level. On first impression, the isolated heart cell would seem to be ideal for mechanical studies. A preparation free of collagen and containing relatively few myofibrils in a single cohesive structure would allow almost direct access to the measurement of sarcomere length and cross-bridge contractile responses. At the present time, however, our ability to isolate a single heart cell for mechanical studies has simply reduced our problem from a state of megadimensional ambiguity to one of mere multidimensional ambiguity. Also, counterbalanc- ing the structural simplification of the preparation are the inherent challenges of noninjurious cell attachment, the recording of microgram levels of force measure- ment, and individual sarcomere resolution. These chal- lenges have kept our progress in studies of cardiac myo- cyte contraction at a slow pace; however, alternative mechanical measurements, such as shortening of unat- tached cells, auxotonic responses of cells attached to rel- atively compliant force transducers, stiffness measure- ments, steady-state activation of skinned cells, and labeled antibody staining of cytoskeletal and interme- 0031-9333/91 $1.50 Copyright 0 1991 the American Physiological Society 413

Mechanical Properties of Isolated Cardiac Myocytesweb.mit.edu/smik/www/Brady - Mech Properties of Isolated...PHYSIOLOGICAL REVIEWS Vol. 71, No. 2, April 1991 Printed in U.S.A. Mechanical

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

  • PHYSIOLOGICAL REVIEWS

    Vol. 71, No. 2, April 1991 Printed in U.S.A.

    Mechanical Properties of Isolated Cardiac Myocytes

    ALLAN J. BRADY

    Department of Physiology and Cardiovascular Research Laboratory, University of California, Los Angeles, School of Medicine, Center for Health Sciences, Los Angeles, Califwnia

    I. Introduction ........................................................................................... 413 II. Development of Adult Myocyte Preparation ......................................................... 414

    III. Mechanical Properties of Cardiac Myocytes .......................................................... 414 A. Attachment procedures ............................................................................ 414 B. Passive mechanical properties .................................................................... 415 C. Contractile responses in active cardiac myocytes ................................................. 418

    IV. Indexes of Contractility ............................................................................... 420 A. Shortening as an index of contractility ........................................................... 420 B. Shortening related to Frank-Starling relation ................................................... 421

    V. Summary and Conclusions ............................................................................ 424

    I. INTRODUCTION

    Viewed from the perspective of organ function, there has been a long-standing interest in the separa- bility of cardiac muscle length or ventricular volume- dependent effects (i.e., the Frank-Starling relation) from intrinsic inotropic mechanisms in the cardiac con- tractile response. Furthermore, at the cellular level there is continued interest in the mechanisms of passive and active force generation in the sarcomere. In fact, the resolution of both levels of interest has been severely limited in studies over the past decades by the complex- ity of multicellular cardiac preparations and our inabil- ity to measure or control a minimal number of the con- tractile-related parameters. Actually, the results of many of these studies strongly question the separability of geometrical and intrinsic factors of cardiac contrac- tion (12, 71, 97, 99), but the issues are by no means closed.

    For example, attempts to deduce cross-bridge me- chanical kinetics from multicellular preparations (in- cluding geometrically linear preparations such as tra- beculae and papillary muscles) have led to ambiguous results because of the viscoelastic nature of these prepa- rations (93) and because no clear representation of the relation between passive and active elements of the tis- sue is evident (20). On the one hand, for simplicity, rest- ing mechanical properties of these preparations have been assumed to be in parallel with active properties. This arrangement has been represented as a Maxwell mechanical analogue. Then, again, series elasticity in excised preparations has been largely attributed to end compliance of the preparations that is related to the attachment of recording devices so that a Voigt ana- logue is suggested.

    Finally, tests of these assumptions reveal a consid-

    erable complexity of stress and strain distributions at the sarcomere level, reflecting significant structural and functional nonuniformity in both passive and ac- tively expressed responses. Because contractile function is strongly sarcomere length and cross-bridge shorten- ing dependent, this inhomogeneity further reduces the reliability of interpretations of cross-bridge function based on these global or whole tissue measurements.

    Added to these concerns is the growing evidence that some phase of the contractile activation process may be diminished at short sarcomere length or with the shortening necessary to achieve sarcomere lengths below unattached rest length. Hence it becomes even more critical to know or to control sarcomere length at the sarcomere level.

    On first impression, the isolated heart cell would seem to be ideal for mechanical studies. A preparation free of collagen and containing relatively few myofibrils in a single cohesive structure would allow almost direct access to the measurement of sarcomere length and cross-bridge contractile responses. At the present time, however, our ability to isolate a single heart cell for mechanical studies has simply reduced our problem from a state of megadimensional ambiguity to one of mere multidimensional ambiguity. Also, counterbalanc- ing the structural simplification of the preparation are the inherent challenges of noninjurious cell attachment, the recording of microgram levels of force measure- ment, and individual sarcomere resolution. These chal- lenges have kept our progress in studies of cardiac myo- cyte contraction at a slow pace; however, alternative mechanical measurements, such as shortening of unat- tached cells, auxotonic responses of cells attached to rel- atively compliant force transducers, stiffness measure- ments, steady-state activation of skinned cells, and labeled antibody staining of cytoskeletal and interme-

    0031-9333/91 $1.50 Copyright 0 1991 the American Physiological Society 413

  • 414 ALLAN J. BRADY Volume 71

    diate filaments, are leading to more detailed concepts of the contractile process in cardiac muscle. These develop- ments are discussed in some detail in this review.

    II. DEVELOPMENT OF ADULT MYOCYTE PREPARATION

    During the past 20 years numerous attempts have been made, using several experimental approaches, to develop an excised single cardiac myocyte preparation that would tolerate normal Ca” concentrations and that could be used to evaluate electrical, metabolic, and contractile function at the cellular level. The vast litera- ture on the progress of this development has been ex- tensively reviewed by Dow et al. (28, 29), and recent methods of myocyte isolation have been evaluated by Powell (100).

    During this period the isolation procedures have been modified such that myocytes can be isolated that remain quiescent and relaxed in normal Ca2+-contain- ing perfusates and are electrically excitable. As noted by Haworth et al. (47), a primary step in producing Ca2+-tolerant cardiac myocytes was the recognition that a critical level of Ca2+ must be present in the digestion medium ( -20 PM) for the first few minutes of coronary perfusion in order for the enzymes to separate the cells and break up the extracellular collagen. The addition of a variety of substrates to assist the cells in handling the Ca2’ loading that occurs with the isolation procedure also seems to be beneficial (7, 47, 84, 96, 125, 144).

    Cell separation at the intercalated disks appears to be readily effected by collagenase so that the principal injury to the cells probably occurs in the separation of the cells from each other as they are freed from the collagen matrix. Gap junctional areas of the nexes may be particularly sensitive to mechanical damage, but electron micrographs of isolated cells show many intact gap junctions often with some adhering residue of an adjacent cell (38,82,85). The large number of cells that can be isolated from ventricular tissues [total yields of 6579% in 1 mM Ca2’ (143) and up to 95% in 1.8 mM Ca2’ (S)] indicates that cells from which a gap junction has been removed can reseal the injured area (85) and that, as reported by Severs et al. (llS), most gap junction .s are internalized by endocytosis soon after cell isolation.

    These preparations have been the subject of a wide variety of successful studies of cardiac cellular physiol- ogy, except that the cells, largely devoid of collagen, are extremely sensitive to external mechanical stress. These persisting attachment problems have resulted in a increasing number of studies that have utilized short- ening in unattached cells as a measure of contractility.

    III. MECHANICAL PROPERTIES OF CARDIAC MYOCYTES

    One of the major objectives in studying the me- chanical properties of the isolated cardiac myocyte, as in studies of whole cardiac tissue, is to deduce cross-

    bridge kinetics as a means of understanding the con- tractile process. However, our access to the cross bridges is still indirect. For example, to measure myo- cyte force or stiffness we must make some attachment to the cell, and this attachment may produce distorting strains or stresses that influence the contractile re- sponse. Also, to resolve the small forces developed by the isolated myocyte, the compliance of the force-sens- ing system may be appreciable relative to the length of the cell. Furthermore, the myocytes possess passive viscoelastic properties, the orientation of which with re- spect to the cross bridges is still unclear, so that the way in which these passive elements are represented in rela- tion to the active contraction (series or parallel combina- tions) influences our interpretation of cross-bridge ac- tivity. On the other hand, when shortening in the unat- tached myocyte is measured, what is obtained is a measure of the maximum turnover rate of the cross bridges, but internal viscous loading and loading by lon- gitudinal and radial elastic elements as well as external cell orienting or stabilizing forces against which the cy- cling cross bridges are working may not be well defined. In addition, the consequences of shortening on cross- bridge activation are only beginning to be understood. These problems leave our interpretation of cross-bridge dynamics in active cells still rather clouded.

    It should be emphasized, however, that the variety of force-, stiffness-, and length-measuring techniques that have been utilized in isolated myocyte preparations in recent years has contributed substantially to our un- derstanding of the myocardial contractile process. How- ever, not unexpectedly, these studies have also extended the scope of parameters that must be dealt with in achieving this goal. The following discussion of these data attempts to put the definitive and ambiguous aspects of these studies in perspective.

    A. Attachment Procedures

    The major limitation in attachment to intact cells is that the sarcolemma, without the protective collagen, is extremely sensitive to applied stress. A variety of ce- ments have been tried to bond the cell to the styli of mechanical transducers. Also, the cells will adhere to poly-L-lysine-coated surfaces to some extent, but cell damage or separation often occurs before maximum contractile activation occurs. Tarr et al. (131), using iso- lated frog atria1 cells, and Shepherd and Kavaler (121), using single guinea pig ventricular myocytes, have had success with this coating procedure. The studies with frog atria1 cells took advantage of the extreme length of frog atria1 cells (ZOO-300 pm) such that they were able to wrap the ends of these narrow (~5 pm wide) myocytes around poly-L-lysine-coated glass supporting beams. The calibrated deflection of one beam served as the force transducer. Intact cells survived this procedure well and remained quiescent and electrically excitable. The major disadvantage of this method was the neces-

  • April 1991 MECHANICAL PROPERTIES OF ISOLATED CARDIAC MYOCYTES 415

    sity to construct the glass beams such that a displace- ment of -5% of cell length was necessary to obtain adequate resolution of the force responses. In the guinea pig study the cell was simply placed on and thus adhered to the coated surface of the two supporting glass beams. Sarcomere uniformity in these cells has not been re- ported.

    Silicone rubber cements are useful, but a uniform stress distribution with attachment to the cell is diffi- cult. A fibrin attachment procedure was reported to be successful by Copelas et al. (24). The compliance of the fibrin seemed to be tolerable (1 ,um/mg), but the method has not been widely used.

    Brady et al. (19) attached isolated cardiac myocytes of rats to a force transducer with single-barreled suc- tion micropipettes and recorded electrically stimulated twitch tension as a function of cell length, external Ca2+ variations, and stimulus interval variations. They mea- sured peak twitch forces of -4 pN, but the success rate of such recordings was extremely low because of the high stress on the sarcolemma at the tips of the micro- pipettes.

    Among the more promising attachment methods is a concentric double-barreled micropipette in which the inner pipette is recessed such that approximately four to five sarcomeres of each end of the cell can be drawn in against the inner pipette with mild suction. The glass surfaces are treated with a thin coat of a prototype bar- nacle cement (Biopolymers) that provides sufficient ad- hesion to support full Ca2’ -activated force development in detergent-skinned myocytes. This attachment applies more of the supporting stress to the intercalated disk surface where stress transfer normally occurs (17). This technique has not yet solved the sarcolemmal stress problem with intact myocytes perfused with normal lev- els of extracellular Ca2+. However, this method works well with detergent-skinned preparations and will sup- port forces of maximal activation with sarcomere strain distribution only 12% in excess of that of unattached myocytes (K. P. Roos and A. J. Brady, unpublished ob- servations).

    A major problem with attachment to highly sensi- tive force transducers is that in cases where the force detection system is not submersible, some part of the force-sensing system must pass through a liquid-air in- terface. The surface tension of the meniscus around the connecting element through this interface can be of -10 pN, whereas the myocyte forces of interest may be one to two orders of magnitude below this level. Thus ex- treme surface stability and reproducibility of the perfus- ate surface are required to measure myocyte forces with these transducers. Iwazumi (59) developed a flow con- troller system in which the chamber solution surface is monitored electronically and flow adjusted by means of servo-controlled solenoids to maintain a constant and reproducible solution level. This device, in conjunction with a novel electromagnetic force transducer, made possible the stable recording of tension from single myofibrils.

    B. Passive Mechanical Properties

    1. Comparative elastic properties of skeletal and cardiac cells

    In attempts to determine whether the relatively high resting stiffness of cardiac muscle has an intracel- lular component, a number of studies have compared the stiffness of isolated cellular preparations of cardiac and skeletal muscle.

    Thus comparing the elastic properties of isolated cells of the hamster gracilis and ventricular myocytes at sarcomere lengths 2.20 and 2.6 pm, Fish et al. (36) showed that resting stress values were three to four times greater in the cardiac cells at each sarcomere length. Also, an exponential dependence of stress on sarcomere length was found in which the slope of the natural log stress-sarcomere length relation in the car- diac and skeletal muscle cells was 7.48 and 5.77 mN/ mm2, respectively [where strain was measured as the relative displacement from control length, (L - L,)/L,]. These values were significantly different at the 3% level. Thus passive mammalian isolated cardiac muscle cell stress is severalfold greater than skeletal muscle cells at similar lengths.

    These data are supported by the higher cytoskele- ton protein content of cardiac cells; for example, Maruyama et al. (78) report that 18% of the total myofi- brillar protein is connectin in bovine heart, which is threefold greater than in skeletal muscle. Price (101) found 2% desmin, also in bovine heart, compared with a report of 0.35% in skeletal muscle (90). In a study of frog (Rana catesbeiana) glycerinated semitendinosus and atria1 preparations in which the contractile filaments had been extracted in 0.6 M KCl, Matsubara and Maruyama (81) found that further treatment of the preparations with 1% NaOH or 2% sodium dodecyl sul- fate to remove the remaining striated structures re- duced the semitendinosus resting tension to t20%, but the atria1 preparation retained >50% of its resting ten- sion at a sarcomere length of -2.75 pm. They concluded that because the primary structure remaining after these treatments was probably a network of the large cytoskeletal filament connectin, the higher content of connectin in heart muscle may be responsible for the difference in resting tension.

    2. Relation of cellular to tissue elasticity in cardiac muscle

    In studies of the passive mechanical properties of single isolated frog (R. catesbeiana) atria1 fibers in which the cells were attached by wrapping around sup- porting glass rods, Tarr et al. (131) found that these cells had resting stresses 8- to 30-fold less than intact frog atria1 trabeculae and could be readily extended to sarco- mere lengths of 3.45 pm. In contrast, the passive force recordings in isolated mammalian cardiac myocytes (19,

  • 416 ALLAN J. BRADY Volume 71

    32,35) indicate that substantial resting tension is appar- ent at sarcomere lengths of 2.0-2.2 pm, and the myo- cytes become extremely stiff at sarcomere lengths of 2.4-2.6 pm (l&35,36). Nag and Zak (85) report that some collagen fibrils may be attached to enzymatically disso- ciated rat cardiac myocytes and might be responsible for some of the resting tension. However, other studies showed that these enzymatically dissociated cells are nearly free of collagen (18, 104) so that, in contrast to frog atria1 tissue, the high resting tension of mamma- lian heart muscle must be, at least in part, a character- istic of the cardiac cell ultrastructure.

    The relative contribution of passive cellular elastic- ity to total tissue elasticity is of interest both in terms of resistance to stretch and resistance to shortening. In the latter case a restoring force would be established that may be responsible for some of the steepness of the ac- tive length-tension relation and for the rapid restora- tion of initial length in cells shortened below their unat- tached rest length (sarcomere length - 1.85-1.93 pm) (see sect. IvB). With regard to a resistance to stretch, Kentish et al. (62) measured the passive length-tension relation in intact and detergent-skinned rat trabeculae and found a substantial reduction in resting tension after 30 min of perfusion with 1% Triton X-100. In the skinned preparation at sarcomere length of 2.20 pm, resting stress was -40% of that of the intact prepara- tion. These observations suggest that dissolution of the sarcolemma in detergent may isolate the intracellular elastic elements from the extracellular stroma.

    Comparing the length dependence of passive stress and stiffness of intact papillary muscle and trabeculae with excised single cardiac myocytes, Brady et al. (20) and Brady (16) found that at sarcomere lengths of -1.9 pm the passive stress in trabeculae and collagenase-iso- lated myocytes was similar. On the other hand, the slope of the natural log stiffness-length relation in trabeculae and papillary muscles was three to five times greater than in the excised myocytes. Similar conclusions were expressed by Fish et al. (36). In whole muscle prepara- tions, Krueger (65) and Krueger and Tsujioka (70) re- port two phases of the resting tension-sarcomere length relation with a steep component beginning around a sarcomere length of 2.6 pm. They attribute much of the more compliant component below a sarcomere length of 2.69 pm to intrinsic cellular elements and the steeper portion to extracellular elements.

    These results suggest that the major component of passive stress in whole tissue is probably borne by the extracellular collagen at the longer sarcomere lengths (65,105). However, extrapolating these relations to sar- comere lengths < 1.9 pm suggests that at the shorter lengths a compressive stress or stiffness of the myo- cytes may become a significant or even the dominant factor in total tissue elasticity, particularly if the ele- ments of the strutlike structure of the extracellular col- lagen matrix go slack below unstressed muscle length (13). This possibility is suggested by the fact that reex- tension after active shortening in isolated myocytes is as rapid or more rapid than in multicellular prepara-

    tions (64, 66,87,104,111,113). Krueger (66) emphasizes the fact that myocyte relengthening after active short- ening occurs in a fast and a slower phase. The relative contributions of el astic recoil and cross -bridge detach- ment processes to these phases remain unclear, but as Krueger (66) points out, the magnitude of the restoring forces is sufficient to influence ventricular filling in the whole heart by a sucking action during early diastole.

    3. Structural basis of cardiac myocyte resting tension

    The relatively high resting stiffness of the cardiac cell has led to correlative mechanical and ultrastruc- tural studies aimed at the identification of the cellular structures responsible for the high resting tension. Brady (16) and Brady and Farnsworth (18) found that the stiffness at a sarcomere length of 2.0 pm measured with ~-HZ sinusoidal length perturbation [and at 70 Hz (unpublished observations)] was relatively unchanged with membrane skinning in 1% Triton X-100 detergent but fell to -20% of control after nearly complete A- band extraction in 0.47 M KC1 (pH = 6.2). In 0.6 M KI, after I-band filament extraction, stiffness fell to -10% of the intact value. These data indicate that the A-band filaments are probably directly involved in bearing resting stress in cardiac muscle.

    Similar myosin disassembly st.udies in single skele- tal muscle fibers (48, 49, 58) also gave results implicat- ing A-band filaments in the support of resting stress. A recent study in detergent-skinned cardiac myocytes (110) showed that in attached cells stiffness declined as the A band was extracted in high salt, whereas in unat- tached cells, cell length shortened with A-band extrac- tion. The studies with unattached cells indicate that in the intact cell some longitudinal elements of the cell must be normally stressed at rest length and that this stress is relieved as the A-ba nd fi laments are disasso- ciated, i.e., the cell shortens as the A band disassembles.

    Th .e high-salt filament- extraction data strongly in- dicate that longitudinal elements, related to the integ- rity of the A band, are responsible for a major compo- nent of the relatively high resting tension characteristic of cardiac muscle. These data also point to differences in passive cytoskeletal elastic structure rather than mem- branous elements or contractile filament interactions, since all the membranous systems were disrupted in de- tergent and perfusate free Ca2+ was

  • April 1991 MECHANICAL PROPERTIES OF ISOLATED CARDIAC MYOCYTES 417

    tal muscle cell protein is cytoskeletal filaments and that the larger share, 8-10% cell protein, is titin. These data, along with the report by Maruyama et al. (78) that 18% of cardiac cellular protein is connectin, indicate that this filament system is probably present in sufficient quantity to be a prime candidate for a major resting tension-bearing element.

    Additional functional significance of this filament system has been reported by Horowits and Podolsky (54), who show evidence in skeletal muscle that the posi- tion of the A band in the sarcomere is maintained by titin filaments joining the thick filaments to the Z disks and that they do, indeed, bear resting tension and be- come compressed at I-band widths < 0.1 pm (53). Fur- thermore, Wang and Wright (141) indicate that titin is attached to the thick filaments at multiple sites, possi- bly at the C proteins (80), because titin epitope spacing in the A-band regions remains constant with cell exten- sion, whereas epitope spacing in the I-band region, at least in skeletal muscle, varies with resting muscle length. These latter data would explain why cell stiff- ness declines as the thick filaments are extracted in high salt (18, 49).

    The 500-kDa filament nebulin, earlier thought to be in series with titin, is now reported to lie in parallel with the thin filaments of skeletal muscle and is probably attached to the Z line but unattached to titin (142); how- ever, Wang and Wright (142) did not find nebulin in cardiac muscle.

    Price (101) reports that 2% of cardiac muscle pro- tein is desmin compared with 0.35% in skeletal muscle (87). Tokuyasu (137) and Tokuyasu et al. (138) showed ferritin labels on desmin both in the Z line and in strands running between the myofibrils in cardiac mus- cle. Somewhat in contrast, using cryoultramicrotomy and immunolabeling, Thornell et al. (136) found desmin only in the region of the intercalated disks but mainly near the Z lines. In any case, titin and perhaps desmin are major candidates for longitudinal stress-bearing ele- ments in resting cardiac muscle.

    The possible contribution of microtubules to car- diac resting tension (116) is more difficult to assess, since the development of this tubular system appears to parallel sarcomerogenesis but tends to decline with the maturation in adult cells. Cartwright and Goldstein (23) report that microtubules run helically around myofi- brils in rat neonatal soleus muscle. The number of mi- crotubules was maximal (1 microtubule/pm’) at day 5 but declined to 0.36 microtubules/pm2 cross section in adults. Colchicine disrupted microtubules, which corre- lated with the loss of organization of the myofibrils in myoblasts and regenerating skeletal muscle. Goldstein and Entman (40) found 0.2 microtubules/pm2 in dog and guinea pig papillary muscle, but in functional cells of the adult rat heart, Samuel et al. (116) found tubulin to be only 10 pg/lOO mg (0.01% ) of total protein. Further- more, the reorganization of tubulin distribution in re- sponse to L-thyroxine was not blocked by colchicine in the cardiac myocytes (115). These data indicate that al- though the microtubules mav contribute significantlv to

    the passive mechanical properties of cardiac cells dur- ing myofibrillogenesis, they probably play a minor role in this regard in normal adult myocytes.

    Other candidates for resting cell stress in intact cells where shortening increases cell diameter are the radially stressed cytoskeletal filaments. For example, at the Z line there are reports of desmin and vimentin (43, 103, 138), vinculin (91, 119), filamin (63), and spectrin (83) and at the M line, skelemins (102). These elements may be stressed as the cell changes length at constant volume or where the contractile filament lattice spacing is altered. Some evaluation of the relative magnitude of this radial component is indicated in the work of Roos and co-workers (107, 111) in which they report that in unattached myocytes placed in varied osmotic strength media the cells decreased in both length and width in hypertonic media. On the other hand, in hypotonic solu- tions, cell length increased little above its control value (sarcomere length = 1.83-1.93 pm) while most of the volume increase occurred by an increase in diameter.

    Krueger (66) evaluated the distribution of stress in cardiac myocytes by recording the sarcomere patterns in propagating asynchronous contractions. Krueger noted that, in the traveling wave, a section of lo-15 sar- comeres ahead of the traveling wave underwent a pre- lengthening of -0.1 pm. From these patterns and pas- sive stiffness data from hamster (36) and rat (16) myo- cytes, Krueger calculated that the passive longitudinal component of restoring force in the contracted regions was ~0.5 mN/mm2.

    This evidence indicates that both radial- and longi- tudinal-oriented intracellular elastic structures play a role in setting the resting sarcomere length and that the elastic moduli of these radial and longitudinal compo- nents are likely rather different. The distributions of cytoskeletal filaments are the probable bases of these elastic elements, and since these distributions are com- plex, it becomes extremely important to know their functional relation to the contractile filaments to inter- pret active force developments and shortening in terms of cross-bridge dynamics and Ca2+ activation.

    4. Sarcomere uniformity

    I) UNATTACHED CELLS. The distribution of sarco- mere patterns in isolated cardiac myocytes has been as- sessed in intact unattached cells at rest (108,109,112), in varied osmotic and ionic stresses (107), during isotonic shortening in electrically initiated contraction [26, 67, 87,111,113,114 (the latter two at image rates of 50-500/ s)], and under voltage-clamp conditions (75). A remark- able degree of regional sarcomere uniformity is noted in these unattached cells both at rest in modified media and during active isotonic contraction. At rest in unat- tached myocytes the major sarcomere length deviation is noted near the nuclei, resulting in an overall sarco- mere length variation of t6% (108). In the electrically stimulated unattached cells (113,114), both the general

    l F l 1 1 . 1 1. t - m l 1 . 1 unirormity and regional ciinerences were maintained

  • 418 ALLAN J. BRADY Volume 71

    during active shortening but the dispersion became greater during relaxation due to regional variations in the onset of relaxation. Using laser diffraction and video imaging, Krueger et al. (67) reported a similar uniformity of shortening and of relaxation as well. They did note, however, that although no buckling of myofi- brils was evident at short sarcomere lengths (cl.5 pm), the first-order diffraction band did not go to zero as expected at these lengths as the Z bands approached the A band. Using computerized enhancement of charge- coupled device (CCD) linear-array images, Roos et al. (111) also noted clear imaging of sarcomeres at sarco- mere lengths < 1.6 pm. At this point it is unclear why both the laser diffraction and the CCD imaging resolved I bands at a sarcomere length where the Z line would be expected to be firmly against the thick filaments. Krueger and London (68) suggest that the cross bridges at the end of the thick filament might add 0.29 pm of variability to the A-band length by way of being folded back on the thick filament stalk at short sarcomere lengths. For example, in the extended muscle the end cross bridges would appear more similar to I-band den- sities because of the absence of the thick filament stalk. In effect, then, as far as shortening is concerned, the A bands would appear to be only 1.4 pm long rather than 1.6 pm. Roos et al. (111) also reported sarcomeres visible by CCD imaging down to a length of 1.4 pm.

    Another explanation might involve A-band shorten- ing as suggested by Pollack (98) and recently shown in skeletal muscle (92). They report that in active shorten- ing of toe muscles (lumbricalis digitorum IV) of the frog (R. temporaria) the A bands shorten by 1520% at sar- comere lengths less than L,. It is interesting that in their electron micrographs of these shortened fibers both the M line and the H bands become obscured. Whether this loss of filament definition may be related to an alteration in cytoskeletal filament distribution (e.g., titin, connectin) has not been established.

    II) ATTACHED CELLS. In stressed passive cells at- tached with single-barreled micropipettes and in myo- cytes impaled with microtools, stress distribution is highly nonuniform near the points of attachment, often with substantial skewing of the sarcomere pattern across the width of the cell. In CCD imaging of single- barreled micropipette-supported preparations, Roos (personal communication) reports that a stretched cell with a mean sarcomere length of 2.47 pm had a range of sarcomere lengths of 1.7-2.8 pm, a standard deviation (SD) > 8%) and a median skewed toward 2.6 pm (the number of sarcomeres measured was -500). This devia- tion compares to an SD of -6% in the unattached cell. In a cell at a mean sarcomere length = 2.23 pm, SD was >8%, sarcomere length range was 1.7-2.6 pm, and the median was 2.2 pm. In this cell, axial skewing was rela- tively minor, but sarcomere misalignments across the width of the cell (radial skewing) spanned the entire length of the cell.

    In these modes of attachment, uniform radial sup- port is poor so that the isolated cardiac myocyte offers little advantage over multicellular preparations in

    terms of control of sarcomere length. On the other hand, attachment to detergent-skinned myocytes with double- barreled micropipettes (18) distributes the applied stress much more evenly over the face of the interca- lated disk. In preliminary studies in resting myocytes, sarcomere length SD values were *6-7%, with little ra- dial or axial skewing at cell extensions 30% above rest length compared with a SD of *5-6% in unattached cells (l,OOO-2,000 sarcomeres) (K. P. Roos, personal com- munication).

    In any case, these observations indicate that in order to take advantage of the greater structural sim- plicity of single isolated heart cells in mechanical stud- ies, appropriate attachment procedures must be em- ployed to maintain the inherent sarcomere uniformity of the myocytes.

    C. Contractile Responses in Active Cardiac Myocytes

    I. Force measurements in skinned myocytes

    Early studies of Ca2’ activation of isolated cardiac myocytes made use of the fact that skinned myocytes adhere well to glass microtools. For example, Bloom (9, 10) and Bloom et al. (11) excised small clusters of cardiac cells (3-5 cells wide by 2-3 cells long) from the mouse ventricle by mechanical disaggregation and, in the lat- ter work, measured contractile force by means of micro- tools impaling the preparations. These preparations were, in effect, skinned, and they responded with sponta- neous contractions in micromolar levels of EGTA-buf- fered Ca2+ perfusates. The sarcoreticular network ap- peared to be intact as evidenced by a contracture re- sponse to 1 mM caffeine. Spontaneous irregular active contractions occurred in most preparations, and their amplitude increased with perfusate Ca2’ in the range of l-10 PM. Maximal steady-state contracture responses were ~4 pN, which is similar to tonic responses (3-4 pN) in mechanically skinned cells (35). Sarcomere length and uniformity were not reported in these studies.

    De Clerck et al. (25) measured force and shortening in mechanically skinned rat cardiac myocytes also using a microtool attachment. The cell did not tolerate normal Ca2’ levels, but contraction could be initiated by iono- phoretically released Ca2+ from an extracellular pipette. Loads could be electronically clamped at predetermined levels with a servo-controlled system, but a shortening of ~5 pm occurred during the clamping process. The contractile responses were limited by diffusion from the pipette, but “isometric” contractile forces also of ~4 PN and maximum unloaded shortening velocity averaged only 0.35 cell lengths/s. The resulting force-velocity re- lations resembled those of intact cardiac tissue but were too irregular to deduce the basic nature of the relation, possibly due to the nonuniformity of activation by the electrophoretic procedure.

    In studies of contractile responses in mechanically skinned rat ventricular cells. Fabiato and Fabiato (35)

  • April 1991 MECHANICAL PROPERTIES OF ISOLATED CARDIAC M’fOCYTES 419

    found a damped series of contractile oscillations (0.1-l Hz) occurred in partially activated myocytes in a criti- cal range of Ca2+ concentrations (pCa 6.4-5.7) and at sarcomere lengths < 3.10 pm. These oscillations ap- peared on the rising phase of the contracture between 50-100% of the final tension response. They character- ized these oscillations as local stress-equilibrating ad- justments of the force-sarcomere relations of the sarco- meres such that longer sarcomeres stretch shorter or weaker sarcomeres in series and thus increase the weaker activation levels, whereas the stronger sarco- meres reduce their force by shortening. On the basis of a viscoelastic model of these localized oscillations, they calculate a sarcomere maximum velocity of -10 pm/s (4-5 sarcomere lengths/s), which is similar to that of intact cardiac muscle (95) and that later reported in un- attached cells (-9 pm/s) by Krueger et al. (67), as well as 3.42 cell lengths/s reported by Roos et al. (109) and 3.6 pm/s reported in cat myocytes (61). These damped oscil- lations during the rise of tonic activation are distinct from the oscillations that occur at much lower Ca2+ (pCa = 7.4) (33) that are attributable to Ca2+-induced Ca2+ release from the sarcoreticular network.

    These studies show that the tonically activated skinned cardiac myocyte is capable of force development and shortening comparable to twitch responses in intact tissue.

    2. Active force development in intact cells

    In studies of isolated intact frog atria1 myocytes (R. catesbeiana) (127,131), contractile force was determined from the deflection of a force beam measured with a video system and a calibration of the beam compliance. This attachment procedure was possible because the cells are ZOO-300 pm long; however, only about a 50-pm segment of the cell was used for sarcomere length mea- surements. The cells averaged -5 pm wide and thus contain only a few myofibrils in cross section. The cells were electrically excitable so that twitch responses could be recorded along with a video record of sarcomere length. They found that frog intact atria1 cells develop twitch contractile forces of -150-250 nN, depending on the compliance of the supporting beams (132). Force beam compliance ranged from 0.028 to 0.128 pm/nN, re- sulting in sarcomere length changes of 1020% for con- tractions of this magnitude (128, 132). Contractile strength and shortening velocity continuously increased with length even to sarcomere lengths > 3 pm (129). The strength of the twitch and the shortening velocity could also be graded by both the amplitude and the duration of the stimulus current, indicating a graded release of ac- tivating Ca2+ in the cell (130). With the recording of contractile responses after a variety of quick releases, auxotonic force-sarcomere length relations could be measured so that auxotonic force-velocity relations could be deduced (129, 130).

    At an initial length of 3.0-3.2 pm, these cells short- ened at nearly constant velocity to a sarcomere length of

    -1.6 pm. Earlier, Manring et al. (76) and later Nasser et al. (87) found similar shortening characteristics in both trabeculae and rabbit myocytes and referred to heart muscle in this regard as a “shortening machine.” These data were explained as resulting from the balance of the time-dependent rise in contractile activation of the cell and the shortening-dependent deactivation of the con- tractile process (12, 15, 31).

    Using a similar frog atria1 myocyte preparation and a high-velocity perfusion technique, Shepherd and Kavaler (121) found contractile force and shortening to respond maximally to step changes in Na+, Ca2+, and caffeine in 300 ms and concluded that there is no evi- dence for Ca2+-induced Ca2+ release in frog atria1 cells, i.e., that the Ca2+ source must be external and that caf- feine must have a sarcolemmal action because its aug- mentation of contraction is as fast as the contractile effects of Na+ and Ca2+ changes.

    In an isolated guinea pig myocyte preparation (120), a cell was stimulated with a patch-clamp elec- trode. The measured contractile forces at 35°C averaged 5.3 mN/mm2 (1.2 ,uN) at a sarcomere length of 1.81-1.88 ,urn with shortening up to 8% in the central region of the cell. No measurement of sarcomere uniformity was re- ported in these experiments. In the measurements of shortening in unattached cells, using a Sr2+ perfusate (122), they concluded that, in contrast to the frog atria1 myocytes, activator Ca2+ must come from internal stores and that the sarcoreticular network Ca2+ release requires Ca2+ entry through channels.

    If these measurements are put in perspective with respect to multicellular preparations, then the contrac- tile forces reported for the isolated guinea pig myocytes are somewhat disappointing. For example, Yeatman et al. (146) reported a stress of 42 mN/mm2 in papillary muscle at 37°C. They reported a Q10 of -0.9 so that at 35°C the papillary force would be 43 mN/mm2. Thus the guinea pig myocyte (122) generated forces only 12% (5.3 mN/mm2) of those of intact tissue, not compensating for extracellular space. If the extracellular space is as- sumed to be 20% of tissue volume, then the active force is only 10% (54 mN/mm2). In the guinea pig myocyte study, the beam compliance was 0.88 prn/pN, giving a beam deflection of 0.62-2.9 pm or l.O-4.5% of the 65-pm cell length. The highest recorded force was 3.3 PN in contracture (120), which is comparable to the twitch tension recorded in isolated rat myocytes at room tem- perature by Brady et al. (19). Although the cells did not have an appreciable preload (sarcomere length - 1.8 pm), it is doubtful that the small twitch contractions in the guinea pig preparation were entirely due to the short initial length and the high compliance of the sup- porting beams. Another factor may relate to a low Ca2’ load in the short cell. For example, Nichols (88) found that in cat papillary muscle at 3O”C, changes in diastolic length affected subsequent twitch tension. Because stretch enhanced subsequent twitches more than re- lease, Nichols suggested that the diastolic length phe- nomenon may be related to the surface-to-volume ratio of the cells and that ionic exchange sites may be more

  • 420 ALLAN J. BRADY Volume 71

    accessible in the stretched cell. Thus excised cells may be less Ca” loaded than intact preparations, perhaps because they tend to be shorter after being freed from the collagen matrix. Another concern is that the low contractility may be indicative of a major nonunifor- mity of sarcomere dimensions during active contrac- tion.

    These observations indicate that there are still con- siderable reservations concerning the functional state of attached isolated cardiac myocytes.

    3. Displacement recordings in unattached stimulated cells

    The measurement of shortening in electrically stim- ulated isolated cardiac myocytes has been successful, with considerable new information regarding relations between cell function and metabolism. A wide variety of methods have been devised to record myocyte contrac- tions, including I) direct cell transmittance changes (32), 2) video recording of cell motion (32,45,46, 57, 68, 87), 3) laser diffraction (61, 89, 117, 149, 4 phase-lock detectors (26,27,64), 5) the more complex CCD and video imaging systems (109, 112), and 6) edge detectors in which the motion of a contracting isolated cell can be measured in a variety of contractile states (22,37,72-75, 123). The edge-detection methods rely on some adhesion of the isolated cell to a stable surface or a perfusion micropipette so that a constant reference of cell position is obtainable. Because sarcomere length uniformity ap- pears to be well maintained during shortening in unat- tached myocytes, the problem of sarcomere length uni- formity is of less concern than in attached preparations. On the other hand, the load, both internal and external, against which attached cells contract in all of these methods is not known. Thus even though shortening can be graded by the stimulus and environmental factors in these preparations, without a knowledge of the loading of the unattached cells the interpretation of the con- tractile state of these cells in relation to activating Ca2+ may be ambiguous or, at least, may require the measure- ment of a number of parameters to deduce the basis of changes in contractility. These concerns are outlined next.

    IV. INDEXES OF CONTRACTILITY

    A complete description of the mechanical compo- nent of the contractile process with which to relate the underlying biochemical processes requires measure- ments of force, stiffness, and shortening characteristics under a variety of loading and perturbational condi- tions. With the currently available techniques there is a limitation in the extent of these parameters that can be monitored. Problems with attachment of isolated car- diac myocytes to recording and control devices puts a constraint on any characterization of the myocardial contractile process such that the most extensive data at

    hand have come from shortening responses. The limited information that is available regarding force develop- ment indicates that isolated cardiac myocytes retain at least some of the relatively high passive stiffness char- acteristic of cardiac muscle and that they are capable of active stress generation, both in tonic and in twitch re- sponses, comparable to intact preparations. However, the sensitivity of the collagen-free sarcolemmal mem- brane to applied stress and the absence of adequate cell attachment methods that maintain the inherent sarco- mere uniformity during stress development leave little more than a few, more or less, auxotonic force record- ings under very limited experimental conditions. Cer- tainly a sarcomeric force-velocity relation that is neces- sary to relate the contractile process to biochemical ki- netics cannot be characterized. The recording of isometric and afterloaded force responses and the es- tablishment of the functional relation between active and passive forces as a function of sarcomere length remain among the most important experimental chal- lenges.

    In the face of such limitations in contractile force measurement, a great deal of progress has been made in quantifying the shortening responses of the isolated myocytes under a wide variety of conditions. The more opportune challenge, then, is to understand the intrinsic factors that control myocyte shortening in terms of ex- ternally applied parameters, such as Ca2+ levels, stimu- lus patterns, and pharmacological agents. Some of the advances and concerns related to the measurement of shortening in unattached cells are discussed in the next section.

    A. Shortening as an Index of Contractility

    The active shortening response of an unattached myocyte reveals the maximum degree of shortening and the shortening velocity of the minimally loaded cell manifested under the conditions extant in the internal and external environment of the cell. Both myocyte shortening and shortening velocity are readily measur- able with current techniques. Thus, to the extent that these two parameters are indicative of the state of the contractile process, there is a measure of functional pa- rameters that can be modified by experimental condi- tions. The parallels that have been demonstrated be- tween myocyte shortening responses and those of intact tissue confirm that the isolated myocyte is a useful model of cardiac tissue. These features have been re- cently reviewed by Allen and Kentish (1).

    A number of major concerns need emphasis in the use of unattached cell responses as an index of contrac- tility and, in particular, in the use of this index to infer cytosolic and sarcoreticular network Ca2+. In this re- gard, Krueger (66) emphasizes the point that myocyte shortening and relengthening reflect different compo- nents of internal loading and contractile processes so that a measure of maximum shortening alone may not be a sufficient measure of contractility. Krueger indi-

  • April 1991 MECHANICAL PROPERTIES OF ISOLATED CARDIAC MYOCYTES 421

    cates that the ratio of the rates of shortening to re- lengthening provides additional information that may help to differentiate between passive and active compo- nents of the shortening event. For example, interven- tions that increase the elastic loading of the shortening cell, e.g., compression of cross bridges or increased stress in the cytoskeleton, as may occur in anoxic heart tissue (39,124), would slow the shortening rate but accel- erate relengthening. In contrast, isoproterenol in- creases both shortening and relaxation rates in rat and rabbit myocytes (45) as well as increasing the extent of shortening. Thus in assessing the mechanism of an in- tervention the rate of cell length change should be in- cluded along with the measurement of maximum length change.

    length dependence of activation all are factors that may affect the determination of this steep length-tension curve. Therefore it is helpful to consider what is known about the basis of force-length relations in cardiac muscle.

    There are further concerns when the basis of short- ening is examined in terms of the Frank-Starling length-tension relation. The steep ascending limb of the length-tension curve, relative to skeletal muscle, im- plies that the ability of the cell to generate force must be affected by some property of cardiac contraction at short sarcomere length or by the shortening process it- self, which is required to achieve a sarcomere length below rest length (-1851.90 pm) (42). A question is, What characteristics of the contractile process do mea- surements of maximum unloaded shortening and maxi- mum shortening velocity reveal in altered contractile states relative to changes in cross bridges and/or Ca2+ kinetics? To put some of these questions in perspective, the next section reviews some of the issues relative to the functional basis of the ascending limb of the cardiac length-tension relation.

    Specifically, the following are some of the mecha- nisms that have been proposed to influence the ascend- ing limb of muscle: 1) interference with cross-bridge at- tachment from thin filaments passing through the M zone of the sarcomere at short sarcomere lengths (41, 56); 2) internal passive elastic restoring forces that in- crease with cell shortening or with an increase in cell diameter (41); 3) buckling of the ends of thin filaments of the adjacent halves of the sarcomere as they meet in the M region (21); 4) compression of attached cross bridge as sarcomere shortening occurs, thus creating a cross-bridge-dependent restoring force [in skeletal mus- cle (4)]; 5) shortening deactivation that may occur as a result of quick releases during activation or from the necessary shortening required to reach sarcomere lengths less than rest length (sarcomere length = 1.80- 1.85 pm) before the isometric phase of contraction (12, 14, 30); 6) a length-dependent activation process in which Ca2+ bound to troponin C (TnC) is released when a cross bridge becomes detached during active shortening with a consequent reduction in the number of activation sites as shortening occurs (2,3,52); and 7) the influence of diastolic length on intracellular Ca2+ levels (88). These mechanisms are considered in some detail.

    I. Thin jZament interference of cross-bridge attachment

    The early suggestion that the ascending limb of the B. Shortening Related to Frank-Starling Relation

    Because isolated unattached myocyte preparations are not normally preloaded, except by possible adhesion to the supporting surface and by the viscous loading of the perfusion media, initial length factors affecting con- tractility may not be a variable parameter. Exceptions might appear where Ca2+ loading or unloading affects the initial cell length; however, these variations could be monitored during the interventions under study. On the other hand, when a muscle shortens in a twitch from an initial length near the peak of the ascending limb of the length-tension relation, its capacity to manifest exter- nal force declines as the muscle becomes shorter. Conse- quently, when shortening in unattached cardiac myo- cytes is used as an indicator of contractility, the under- lying length-dependent factors may affect the interpretation of the mechanism of an intervention, de- pending on which process or processes are responsible for the ascending limb. There are consistent data indi- cating that the ascending limb of the length-tension re- lation in cardiac muscle is significantly steeper than in skeletal muscle (for reviews see Refs. 1, 60, 133). The higher resting tension, restoring forces, the absence of tetanization, the relatively low level of Ca2+ activation during a twitch in cardiac muscle. and an atmarent

    length-tension relation might be due to the interference of cross-bridge attachment from thin filaments of oppo- site ends of the sarcomere (41, 56) has been discussed extensively by Jewel1 (60), Gordon and Pollack (42), and Allen (5). The conclusion is that in maximally activated muscle fibers the plateau of tension development ex- tends down to nearly 1.6 pm where the thick filaments begin to interact with the Z lines. Indeed Fabiato and Fabiato (33) showed a decline to only 85% maximum force development in tonically activated skinned cardiac myocytes at a sarcomere length of 1.6 pm compared with 40% in phasic Ca2+-induced Ca2’ release responses. Thus this portion of the ascending limb appears to be due more to incomplete activation or to the development of internal forces that resist sarcomere shortening rather than an interference with cross-bridge attach- ment.

    2. Restoring forces

    The fact that neither cardiac muscle nor isolated myocytes will retain a sarcomere length less than -1.8 pm after active shortening indicates that some restor- ing force must be developed with shortening that re-

    ~, -.~~ -~ --~~ --L- ~- - stores the rest length during relaxation. The source and

  • 422 ALLAN J. BRADY Volume 77

    magnitude of this restoring force is of considerable in- terest not only in regard to the ascending limb of the cardiac length-tension relation but also because its char- acter might give some clue to the functional organiza- tion of passive and active elements of the contractile system. However, Jewel1 (60) concluded that the strong Ca2+ dependence of the ascending limb of the length- tension relation indicated that a passive elastic restor- ing force was probably not the primary basis for the fall in developed tension at short sarcomere lengths. Simi- larly, Fabiato and Fabiato (34) measured the force of myocyte extension by pressing a force transducer against a contracted cell during the relaxation process. They measured a force of ~4% active force as an esti- mate of the restoring force of the relaxing cell. Also by applying small loads to a relaxing trabecular prepara- tion, ter Keurs and co-workers (134,135) concluded that the restoring force developed during shortening of the preparation could not be greater than a few percent of maximum isometric tension. Strauer (126) found that in the absence of an external load relengthening becomes slower with greater shortening. Krueger (64) concluded that the force of relengthening arises from within the cells but that in local asynchronous contractions the re- lengthening force (AP) from a sarcomere length of 1.6 pm is probably only a few percent of maximum isomet- ric tension (-0.4 mN/mm2). In synchronous contrac- tions the restoring forces are estimated to be, at most, l-2 mN/mm2 (34, 66). The latter conclusion is based on Krueger’s estimation of the restoring force using data from the resting stiffness (AP/AZ) of isolated myocytes (16,36) and the extent of segmental shortening (AZ) in a portion of an unattached myocyte caused by the asynchronous segmental contraction of the cell. The ar- gument for a low restoring force is also supported by the idea that during shortening at maximal rates relatively few cross bridges are attached so that the restoring force must be less than the contractile force producing the shortening.

    The location of some of the elements responsible for the restoring force is suggested by the observations of Fabiato and Fabiato (35), who report that detergent- treated isolated single myofibrils do not relengthen after active shortening to a sarcomere length of 1.6 pm, whereas similarly activated mechanically skinned myo- cytes do relengthen to 1.9 pm. These data indicate that the major intracellular restoring force probably is re- lated to stress in the interfibrillar structures that are destroyed by the detergent. At present it is unclear whether the longitudinal cytoskeletal filament systems (titin, desmin) are compressible or whether they simply go slack as the sarcomeres shorten. Because these fila- ments appear to survive detergent treatment, i.e., the high resistance to stretch remains (18), the absence of relengthening suggests that they do indeed go slack.

    However, the possibility must not be overlooked that the restoring force is not simply a passive property of the elastic structure of the sarcomere. It is possible that in the intact cell the restoring force is related, in

    part, to the active force generation process such as the compression of attached cross bridges.

    In this regard, it would be helpful to have a mea- surement of stiffness during controlled shortening and relengthening to evaluate the nature of the stressed ele- ments (cross bridges and elastic structures) extant dur- ing the shortening and relengthening states.

    3. Buckling of thin &laments

    Brown et al. (21) observed the appearance of waves in the contractile filaments of passive semitendinosus fibers of R. pipens compressed in gelatin. Krueger et al. (67) saw similar wavy patterns in isolated ventricular myocytes under similar compressed conditions. In the skeletal muscle study, Brown et al. (21) found that rather than forming a double-overlap pattern in the M zone in the compressed fiber, the ends of the thin fila- ments appeared to abut together, forcing the sarcomere to bend laterally and causing the wavy appearance. They suggested that a repulsive force due to the charges on the thick and/or thin filaments in the resting state may resist thin filament entrance into the M-zone re- gion, whereas with activation thin filaments can be drawn past each other. On the other hand, the implica- tion may be that in the passive sarcomere thin filament alignment between the two halves of the sarcomere is sufficiently precise to orient the end of the filaments into a relatively rigid end-to-end encounter. With active shortening, force exceeds the repulsive force or with cross-bridge attachment the thin filaments become misaligned enough to allow them to pass. Evidence for this process might be seen in a careful analysis of the slope of the myocyte shortening record in the vicinity of sarcomere length = 1.6 pm. On the other hand, the var- ied thin filament length in cardiac muscle (106) might smooth this transition such that only a subtle change in slope would occur.

    4. Compressed cross bridges

    From another point of view, the steep ascending limb of the length-tension relation may reflect a compo- nent due to compressed cross bridge during shortening. In active skeletal muscle, Allen and Moss (4) found that the stiffness-to-force ratio increased at short sarcomere lengths, indicating that an increasing number of cross bridges must be in compression as sarcomere length de- creased. If substantial cross-bridge compression does occur in shortening myocytes, then inotropic interven- tions that change the metabolic state of the cell, and thus the cross-bridge cycling rate by way of pH or aden- osinetriphosphatase changes, might be expressed dif- ferently between the isometric and isotonic modes of contraction. The additional measurement of stiffness during shortening would help resolve this issue.

  • April 1991 MECHANICAL PROPERTIES OF ISOLATED CARDIAC MYOCYTES 423

    5. Shortening deactivation

    When active muscle is allowed to shorten, its ability to generate force is diminished even though the initial length is restored. In skinned cells in which activation can be maintained, the original force can be regenerated at the initial length but the redevelopment of force is relatively slow [half time = 410 and 230 ms in 8 and 15 PM Ca’+, respectively (Z)]. This shortening deactivation has been well demonstrated in both intact cardiac and skeletal muscle (12, 14, 30).

    However, shortening in unloaded cells to ~1.6 pm can occur even in low-Ca2+-perfused experiments. The shortening rate is Ca2+ dependent, but the Ca” level does not determine the extent of shortening in skinned cells (62). Therefore the deactivation process is not a simple function of length change or activating Ca2+.

    6. Length-dependent activation

    Along similar lines, there is a growing body of evi- dence that contractile activation is length dependent, particularly in twitch contractions of intact muscle. These data indicate that Ca” binding to TnC is reduced at short sarcomere lengths. Questions that arise are whether there is a common mechanism between shorten- ing deactivation and the length sensitivity of Ca2+ acti- vation and how this deactivation and length-dependent activation relate to the steep ascending limb of the length-tension relation and, with regard to all these fac- tors, what state of contractility is measured by unat- tached cell shortening when contractile states are changed by various inotropic agents.

    From studies in glycerinated cardiac muscle of cows, in detergent-skinned trabeculae preparations of rats, and from the substitution of skeletal muscle TnC for cardiac TnC in skinned trabeculae of hamsters, the following points concerning myocyte shortening and Ca2+ activation are evident. I) Calcium is released from TnC when cross bridges detach with shortening steps (2, 50, 55). 2) In heart muscle the released Ca2+ correlates best with the force change rather than with the magni- tude of the length change as in skeletal muscle (2,51). 3) With relengthening the Ca2+ may rebind to TnC at least in steady-state activations as in skinned or glycerinated preparations (52). 4) Calcium is bound to TnC in rigor states, and the amount bound varies with sarcomere length (50). 5) Substitution of skeletal muscle TnC for cardiac TnC increases force development at short sarco- mere lengths, indicating that the cardiac TnC regula- tion of contractile force is more length dependent than skeletal TnC regulation (6).

    With regard to deactivation, then, these data indi- cate that as unattached myocytes shorten, Ca2+ binding to TnC is likely to be much reduced relative to an isomet- ric state at each length, since in shortening muscle few cross bridges are attached and thus TnC activation would be low. These effects may have several important

    consequences relative to the method of assessment of cardiac contractility. First, cytosolic Ca2+ would be ex- pected to be higher in unattached shortening, as cross bridges detach during sarcomere shortening causing the release of Ca2+ from TnC. Second, the contraction may be of shorter duration, since sequestering by the sarco- reticular network would be faster. Third, if the cell be- comes afterloaded at some point during the shortening process, then the developed tension would be reduced, since some of the Ca2+ released from TnC would be se- questered by the sarcoreticular network. Indeed, the latter process may be the primary basis for the steep ascending limb of the cardiac length-tension relation as measured by contractile force after shortening to a pre- set length or by the final sarcomere length after active shortening with various afterloads from an initial non- slack rest length. Thus the ascending limb of the Frank- Starling relation may be predominantly determined by the loss of activating Ca2+ bound to TnC as shortening occurs (I).

    With regard to the question of what aspect of con- tractility unloaded shortening measures, the relation of unloaded shortening responses to cytosolic Ca2+ during inotropic interventions becomes complex to interpret. For example, Allen and Kentish (2) showed that free Ca” increased with shortening in skinned cardiac myo- cytes, which is consistent with the hypothesis that bind- ing of Ca2+ to TnC is reduced after detachment of cross bridges from an actin sites. However, increased buffer- ing of the free Ca2+ under these conditions did not alter the development of force after shortening. Further- more, Hofmann and Fuchs (52) showed that, indeed, Ca2+ binding to TnC in glycerinated heart fibers was reversibly dependent on sarcomere length, i.e., Ca” binding to TnC decreased with shortening but increased again with relengthening. Thus with alteration of the inotropic state of a cell, the measurement only of unat- tached cell shortening may not accurately reflect either sarcoreticular network Ca2+ or cytoplasmic Ca2+, i.e., considerably more Ca” might be released by the sarco- reticular network and cytosolic Ca” might be greater than indicated by the magnitude of shortening. An inde- pendent measure of cellular Ca” is still necessary to evaluate the activation processes.

    7. Contractility related to diastolic length

    In cat papillary muscle, Nichols (88) found twitch force to vary with time after a diastolic length change. Also, the increment in twitch force was greater in re- sponse to diastolic length increases than the decrement with similar decreases in diastolic length. This pro- tracted dependence of force development on stepwise length changes suggests that contractile activation can be altered by diastolic length by mechanisms other than simple changes in myofilament overlap. In fact, Allen and Kentish (2) measured intracellular Ca2+ transients as a function of diastolic length in ferret ventricular

  • 424 ALLAN J. BRADY Volume 71

    muscle and found that the transient responses under- went a slow decline after a reduction in diastolic length, paralleling the slow decline in tension. They attribute this length dependence to possible stretch-dependent conductance channels that may influence diastolic in- tracellular Ca2+ levels.

    These observations emphasize the extreme lability of Ca2+ levels in the cardiac myocyte relative to both length and time. Hence interpretations of activation ki- netics based on shortening must include these variables.

    V. SUMMARY AND CONCLUSIONS

    sion. Furthermore, this cellular component is related to the cytoskeleton rather than to membranous elements in the cell. The more likely candidates for the longitu- dinal resting stress-bearing element are titin (connec- tin) and desmin. Titin is the most likely candidate, since it is present in both skeletal and cardiac muscle in rela- tive large quantities, and immunolabeling and stiffness data show that titin extends from the M line to the Z band. These data also show that there is a close associa- tion of titin with the thick filaments in both cardiac and skeletal muscle and that passive stiffness falls as myo- sin is extracted and the titin attachment to myosin is disrupted in high salt.

    A wide variety of techniques have been developed to monitor the mechanical responses of isolated cardiac myocytes. The most successful are those that measure shortening in unattached cells. Because of their relative ease of implementation, edge-detector methods of fol- lowing cell displacement have become most widespread. Laser diffraction techniques have been applied to the single heart cells, and sophisticated sarcomere imaging systems capable of 2-ms time resolution of shortening responses have also been developed.

    It remains to be established whether the influence of passive stress applied to the thick filaments by titin has any regulatory effect on active force development.

    Active force has been recorded in intact single cells from frog atria; however, the compliance of the force transducers was relatively high (-5% L,). (There is an obvious trade-off between transducer sensitivity, which affects noise and drift and compliance.) Some success has been reported with the use of intact rat myocytes supported by suction micropipettes and in guinea pig ventricular myocytes adhering to poly-L-lysine-coated glass beams. With the rat preparation, contractile stress was comparable to that of ventricular muscle, but few cells survived the attachment. In guinea pig myo- cytes, contractile stress in electrically induced twitches was only ~10% of the active stress developed by mam- malian trabeculae or papillary muscles at the same tem- perature (35”C), but, as with the frog atria1 transducer, the compliance of the supporting beams was relatively high. Sarcomere uniformity has not been evaluated in these intact preparations.

    It is apparent that the basis of the ascending limb of the Frank-Starling length-tension relation may be a composite of several factors, each of which may influ- ence the magnitude of shortening in unattached cells when contractility is changed. The sensitivity of acti- vating Ca2’ bound to TnC on cross-bridge attachment may play a predominant role in the length-tension rela- tion. The development of restoring forces with cell short- ening appears to be significant, but their effect may be more in restoring diastolic length than in influencing the length-tension relation. The relative contributions of other factors (cytosolic viscosity, cross-bridge com- pression, diastolic Ca2+) remain to be evaluated.

    Obviously, contractility during varied inotropic in- terventions measured by shortening does not measure the same properties of the contractile system as deter- mined by isometric force development. A complete defi- nition of the contractile process requires measurements of both modes of contraction and a determination of their complex interaction. With these data, an under- standing of the basis of the steep ascending limb of the length-tension relation and its relation to inotropically altered Ca2+ activation may be forthcoming.

    For attachment to the relatively short mammalian cardiac myocytes, the more promising methods that better preserve sarcomere uniformity include double- barreled micropipettes coated with a barnacle adhesive; however, for nonsubmersible transducers, a continuing limitation is the problem of solution surface stability. Unfortunately, the more severe limitation to effective attachment to intact cells is still the extreme sensitivity of the sarcolemma to mechanical stress. The challenge remains to develop an attachment to the intercalated disk such that cell stress can be transferred to the sup- porting transducers along the normal stress-bearing cellular interface.

    The isolated cardiac myocyte represents the most promising preparation available at the present time to aid in the understanding of the cardiac contractile pro- cess. The isolation procedures have been sufficiently re- fined to provide an actively contractile unit that seems reasonably representative of intact tissue. The chal- lenge now is to further develop attachment methods, force transducers, and sarcomere monitoring and con- trol devices with which to relate the mechanical parame- ters of contraction to the myofilament components of the cell and biochemical kinetics. Among these chal- lenges is the necessity to understand the functional re- lation between the passive stress-bearing elements of the cytoskeleton and the contractile filament system.

    The ultrastructural and passive mechanical data I thank Leota Green and Dr. Kenneth Roos for assistance strongly indicate that although the extracellular colla- in the preparation of this review. gen limits the extension of cardiac muscle beyond the This work was supported by the National Heart, Lung, peak of the active length-tension relation, there is also a and Blood Institute Grant HL-30828 and by the Laubisch En- substantial cellular component of resistance to exten- dowment.

  • April 1991 MECHANICAL PROPERTIES OF ISOLATED CARDIAC MYOCYTES 425

    REFERENCES

    1. ALLEN, D. G., AND J. C. KENTISH. The cellular basis of the length-tension relation in cardiac muscle. J. lMoZ. CeZZ. C’urdioZ. 310: 821~840,1985.

    2. ALLEN, D. G., AND J. C. KENTISH. Calcium concentration in the myoplasm of skinned ferret ventricular muscle following changes in muscle length. J. Physiol. Lond. 407: 489-503,1988.

    3. ALLEN, D. G., C. G. NICHOLS, AND G. L. SMITH. The effects of changes in muscle length during diastole on the calcium tran- sient in ferret ventricular muscle. J. Physiol. Lond. 406: 359-370, 1988.

    4. ALLEN, J. D., AND R. L. MOSS. Factors influencing the ascend- ing limb of the sarcomere length-tension relationship in rabbit skinned muscle fibres. J. Physiol. Lond. 390: 119-136, 1987.

    5. ALLEN, R. D. New observations on cell architecture and dy- namics by video-enhanced contrast optical microscopy. Annu. Rev. Biophys. Biophys. Chem. 14: 265~290,1985.

    6. ABU, A., E. SONNENBLICK, AND J. GULATI. Molecular basis for the influence of muscle length on myocardial performance. Science Wash. DC 240: 74-76, 1988.

    7. BIHLER, I., T. H. HO, AND P. C. SAWH. Isolation of Ca2+-toler- ant myocytes from adult rat heart. Can. J. PhysioZ. PharmacoZ. 62: 581-588,1984.

    8. BKAILY, G., N. SPERELAKIS, AND J. DOANE. A new method for preparation of isolated single adult myocytes. Am. J Physiol. 247 (Heart Circ. PhysioZ. 16): HlOl8-H1026, 1984.

    9. BLOOM, S. Spontaneous rhythmic contraction of separated heart cells. Science Wash. DC 167: 1727-1729, 1970.

    10. BLOOM, S. Requirements for spontaneous contractility in iso- lated adult mammalian heart muscle cells. Exip. CeZZ Res. 69: 17- 24, 1971.

    11. BLOOM, S., A. J. BRADY, AND G. A. LANGER. Calcium metabo- lism and active tension in mechanically disaggregated heart muscle. J. MoZ. CeZZ. CardioZ. 6: 137-147, 1974.

    12. BODEM, R., AND E. H. SONNENBLICK. Deactivation of contrac- tion by quick releases in the isolated papillary muscle of the cat: effects of lever damping, caffeine and tetanization. Circ. Res. 34: 214-225, 1974.

    13. BORG, T. K., AND J. B. CAULFIELD. The collagen matrix of the heart. Federation Proc. 40: 2037-2041, 1981.

    14. BRADY, A. J. Time and displacement dependence of cardiac con- tractility: problems in defining the active state and force-velo- city relations. Federation Proc. 24: 1410-1420,1965.

    15. BRADY, A. J. Mechanics of isolated papillary muscle. In: Factors Incfluencing Myocardiac Contractility, edited by R. Tanz, F. Ka- valer, and J. Roberts. New York: Acadamic, 1967, p. 53-64.

    16. BRADY, A. J. Passive stiffness of rat cardiac myocytes. J. Bio-

    17.

    18.

    19.

    20.

    21.

    22.

    BRADY, A. J. Passive elastic properties of cardiac myocytes rela- tive to intact cardiac tissue. In: Cardiac Myocyte-Connective Tis- sue Interactions in HeaZth and Disease, edited by T. F. Robinson and R. K. H. Kinne. Basel: Karger, 1990, p. 37-52. (Issues Biomed. Ser.) BRADY, A. J., AND S. P. FARNSWORTH. Cardiac myocyte stiff- ness following extraction with detergent and high salt solutions. Am. J. PhysioZ. 250 (Heart Circ. Physiol. 19): H932-H943, 1986. BRADY, A. J., S. T. TAN, AND N. V. RICCHIUTI. Contractile force measured in unskinned isolated adult rat heart fibres. Na- ture Land. 282: 728-729,1979. BRADY, A. J., S. T. TAN, AND N. V. RICCHIUTI. Perturbation measurements of papillary muscle elasticity. Am. J. Physiol. 241 (Heart Circ. PhysioZ. 10): Hl55-H173, 1981. BROWN, L. M., H. GONZALEZ-SERRATOS, AND A. B. HUX- LEY. Sarcomere and filament lengths in passive muscle fibres with wavy myofibrils. J. MuscZe Res. CeZZ MotiZ. 5: 293-314, 1984. CAPOGROSSI, M. C., M. D. STERN, H. A. SPURGEON, AND E. G. LAKATTA. Spontaneous Ca2+ release from the sarcoplas- mic reticulum limits Ca2+ -dependent twitch potentiation in indi- vidual cardiac myocytes. A mechanism for maximum inotropy in the myocardium. J. Gen. Physiol. 91: 133-155,1988.

    23. CARTWRIGHT, J., JR., AND M. A. GOLDSTEIN. Microtubules in

    mech. Eng. 106: 25-30,1984.

    24.

    25.

    26.

    27.

    28.

    29.

    30.

    31.

    32.

    33.

    34.

    35.

    36.

    37.

    38.

    39.

    40.

    41.

    42.

    43.

    44.

    45.

    soleus muscles of the postnatal and adult rat. J. UZtrastruct. Res. 79: 74-84, 1982. COPELAS, L., M. BRIGGS, W. GROSSMAN, AND J. P. MOR- GAN. A method for recording isometric tension development by isolated cardiac myocytes: transducer attachment with fibrin glue. Pfluegers Arch. 408: 315-317, 1987. DE CLERCK, N. M., V. A. CLAES, AND D. L. BRUTSAERT. Force velocity relations of single cardiac muscle cells. Calcium dependency. J. Gen. Physiol. 69: 221-241,1977. DE CLERCK, N. M., V. A. CLAES, AND D. L. BRUTSAERT. Uni- form sarcomere behavior during twitch of intact single cardiac cells. J. MoZ. CeZZ. CardioZ. 16: 735-745, 1984. DE CLERCK, N. M., V. A. CLAES, E. R. VAN OCKEN, AND D. L. BRUTSAERT. Sarcomere distribution patterns in single cardiac cells. Biophys. J. 35: 237-242, 1981. DOW, J. W., N. G. L. HARDING, AND T. POWELL. Isolated car- diac myocytes. I. Preparation of adult myocytes and their homol- ogy with the intact tissue. Cardiovasc. Res. 15: 483-514,198l. DOW, J. W., N. G. L. HARDING, AND T. POWELL. Isolated car- diac myocytes. II. Functional aspects of mature cells. Cardiovasc. Res. 15: 549-579, 1981. EDMAN, K. A. P. Mechanical deactivation induced by active shortening in isolated muscle fibres of the frog. J. PhysioZ. Lond. 246: 255-275,1975. EDMAN, K. A. P., AND A. KEISSLING. The time course of the active state in relation to sarcomere length and movement stud- ied in single skeletal muscle fibres of the frog. Acta Physiol. Stand. 81: 182-196,197l. FABIATO, A., AND F. FABIATO. Excitation-contraction cou- pling of isolated cardiac fibers with disrupted or closed sarcolem- mas. Circ. Res. 31: 293-307, 1972. FABIATO, A., AND F. FABIATO. Dependence of the contractile activation of skinned cardiac cells on the sarcomere length. Nu- ture Lond. 256: 54-56, 1975. FABIATO, A., AND F. FABIATO. Dependence of calcium release, tension generation and restoring forces on sarcomere length in skinned cardiac cells. Eur. J. Curdiol., SuppZ. 4: 13-27, 1976. FABIATO, A., AND F. FABIATO. Myofilament-generated ten- sion oscillations during partial calcium activation and activation dependence of the sarcomere length-tension relation of skinned cardiac cells. J. Gen. Physiol. 72: 667-699, 1978. FISH, D., J. ORENSTEN, AND S. BLOOM. Passive stiffness of isolated cardiac and skeletal myocytes in the hamster. Circ. Res. 54: 267-276, 1984. FRATICELLI, A. R. J., R. DANZIGER, E. LAKATTA, AND H. SPURGEON. Morphological and contractile characteristics of rat cardiac myocytes from maturation to senescence. Am. J Phys- ioZ. 257 (Heart Circ. PhysioZ. 26): H259-H265, 1989. FRY, D. M., D. SCALES, AND G. INESI. The ultrastructure of membrane alterations of enzymatically dissociated cardiac myo- cytes. J MoZ. CeZZ. CardioZ. 11: 1151-1163, 1979. GANOTE, C. E., AND R. S. VANDERHEIDE. Cytoskeletal le- sions in anoxic myocardial injury. A conventional and high-vol- tage electron-microscopic and immunofluorescence study. Am. J. Pathol. 129: 327-344, 1987. GOLDSTEIN, M. A., AND J. L. ENTMAN. Microtubules in mam- malian heart muscle. J. CeZZ BioZ. 80: 183-195, 1979. GORDON, A. M., A. F. HUXLEY, AND F. J. JULIAN. The varia- tion in isometric tension with sarcomere length in vertebrate muscle fibres. J. Physioh Lond. 184: 170-192, 1966. GORDON, A. M., AND G. H. POLLACK. Effects of calcium on the sarcomere length-tension relation in rat cardiac muscle: implica- tions for the Frank-Starling mechanism. Circ. Res. 47: 610-619, 1980. GRANGER, B. L., AND E. LAZARIDES. The existence of an insol- uble Z disc scaffold in chicken skeletal muscle. CeZZ 15: 1253-1268, 1978. HAINFELD, J. F., J. S. WALL, AND K. WANG. Titin: quantita- tive mass measurements by scanning transmission electron mi- croscopy and structural implications for the sarcomere matrix of skeletal muscle. FEBS Lett. 234: 145-148, 1988. HARDING, S. E., G. VESCOVO, M. KIRBY, S. M. JONES, J. GURDEN, AND P. A. POOLE-WILSON. Contractile responses of

  • 426 ALLAN J. BRADY Volume 71

    isolated adult rat and rabbit cardiac myocytes to isoproterenol and calcium. J. MoZ. Cell. Cardiol. 20: 635-647, 1988.

    46. HARRIS, P. J., D. STEWART, M. C. CULLINAN, L. M. DEL- BRIDGE, L. DALLY, AND P. GRINWALD. Rapid measurement of isolated cardiac muscle cell length using a line-scan camera. IEEE Trans. Biomed. Eng. 34: 463-467,1987.

    47. HAWORTH, R. A., A. B. GOKNUR, T. F. WARNER, AND H. A. BERKOFF. Some determinants of quality and yield in the isola- tion of adult heart cells from rat. Cell Cakium 10: 57-62, 1989.

    48. HIGUCHI, H., AND S. ISHIWATA. Disassembly kinetics of thick filaments in rabbit skeletal muscle fibers. Effects of ionic strength, Ca2+ concentration, pH, temperature. Biophys. J. 47: 267-275,1985.

    49. HIGUCHI, H., AND Y. UMAZUME. Localization of the parallel elastic components in frog skinned muscle fibers studied by the dissociation of the A- and I-bands. Biophys. J 48: 137-147,1985.

    50. HOFMANN, P. A., AND F. FUCHS. Effect of length and cross- bridge attachment on Ca2+ binding to cardiac troponin C. Am. J Physiol. 253 (CeLZ Physiol. 22): C90-C96, 1987.

    51. HOFMANN, P. A., AND F. FUCHS. Evidence for a force-depen- dent component of calcium binding to cardiac troponin C. Am. J Physiol. 253 (Cell Physiol. 22): C541-C546, 1987.

    52. HOFMANN, P. A., AND F. FUCHS. Bound calcium and force de- velopment in skinned cardiac muscle bundles: effects of sarco- mere length. J. MoZ. CeZC Cardiol. 20: 667-677, 1988.

    53. HOROWITS, R., K. MARUYAMA, AND R. J. PODOLSKY. Elas- tic behavior of connectin filaments during thick filament move- ment in activated skeletal muscle. J. CeZL BioZ. 109: 2169-2176, 1989.

    54. HOROWITS, R., AND R. J. PODOLSKY. The positional stability of thick filaments in activated skeletal muscle depends on sarco- mere length: evidence for the role of titin filaments. J. Cell BioZ. 105: 2217-2223,1987.

    55. HOUSMANS, P. K., H. K. LEE, AND J. R. BLINKS. Active short- ening retards the decline of the intracellular calcium transient in mammalian heart muscle. Science Wash. DC 221: 159-161, 1983.

    56. HUXLEY, H. E. The mechanism of muscular contraction. Science Wash. DC 213: 18-27, 1965.

    57. ISENBERG, G. Ca entry and contraction as studied in isolated bovine ventricular myocytes. 2. Naturfbrsch. Teil C Biochem. Bio- phyys. Biol. Viral. 37: 502-512, 1982.

    58. ISHIWATA, S., K. MURAMATSU, AND H. HIGUCHI. Disassem- bly from both ends of thick filaments in rabbit skeletal muscle fibers. An optical diffraction study. Biophys. J. 47: 257-266,1985.

    59. IWAZUMI, T. High-speed untrasensitive instrumentation for myofibril mechanics measurements. Am. J Physiol. 252 (Cell Physiol. 21): C253-C262, 1987.

    60. JEWELL, B. R. A reexamination of the influence of muscle length on myocardial performance. Circ. Res. 40: 221-230, 1977.

    61. KENT, R. L., D. L. MANN, Y. URABE, R. HISANO, K. W. HEWETT, M. LOUGHNANE, AND G. COOPER IV. Contractile function of isolated feline cardiocytes in response to viscous load- ing. Am. J. Physiol. 257 (Heart Circ. Physiol. 26): H1717-H1727, 1989.

    62. KENTISH, J. C., H. E. D. J. TER KEURS, L. RICCIARDI, J. J. J. BUCX, AND M. I. M. NOBLE. Comparison between the sarcomere length-force relations of intact and skinned trabeculae from rat right ventricle. Influence of calcium concentrations on these re- lations. Circ. Res. 58: 755-768, 1986.

    63. KOTELIANSKY, V. E., M. A. GLUKHOVA, G. N. GNEUSHEV, J. L. SAMUEL, AND L. RAPPAPORT. Isolation and localization of filamin in heart muscle. Eur. J. Biochem. 156: 619-623,1986.

    64. KRUEGER, J. W. Measurement and interpretation of contrac- tion in isolated cardiac cells. In: Biology ofIsolated Ad&t Cardiac Myocytes, edited by W. A. Clark, R. S. Decker, and T. K. Borg. Amsterdam: Elsevier, 1988, p. 172-186.

    65. KRUEGER, J. W. Fundamental mechanisms that govern cardiac function: A short review of cardiac sarcomere mechanics. Heart FaiZure 4: 137-153, 1988.

    66. KRUEGER, J. Rapid relengthening in isolated cardiac cells and the origin of diastolic recoil. In: Cardiac Mechanics and Function in the Normal and Diseased Heart, edited by M. Hori, H. Suga, J. Baan, and E. L. Yellin. Tokyo: Springer-Verlag, 1989, p. 23-34.

    67. KRUEGER, J. W., D. FORLETTI, AND B. A. WITTENBERG. Un- iform sarcomere shortening behavior in isolated cardiac muscle cells. J Gen. Physiol. 76: 587-607, 1980.

    68. KRUEGER, J. W., AND B. LONDON. Contraction bands: differ- ences between physiologically vs. maximally activated single heart muscle cells. In: Contractile Mechanisms in Muscle, edited by G. H. Pollack and H. Sugi. New York: Plenum, 1984, p. 119-134.

    69. KRUEGER, J. W., AND G. H. POLLACK. Myocardial sarcomere dynamics during isometric contraction. J. Physiol. Lond. 251: 627-643,1975.

    70. KRUEGER, J. W., AND K. TSUJIOKA. Sarcomere mechanics-to- ward a physical basis for cardiac contraction. In: SomeMathemat- ical Questions in Biology-Muscle Physiology, edited by R. M. Miura. Providence, RI: Am. Math. Sot., 1986, vol. 16, p. 155-122.

    71. LAKATTA, E. G., AND B. R. JEWELL. Length-dependent activa- tion. Its effect on the length-tension relation in cat ventricular muscle. Circ. Res. 40: 251-257, 1977.

    72. LANGER, G. A. The effect of pH on cellular and membrane cal- cium binding and contraction of myocardium. A possible role for sarcolemmal phospholipid in EC coupling. Circ. Res. 57: 374-382, 1985.

    73. LANGER, G. A., J. S. FRANK, T. L. RICH, AND F. B. ORNER. Calcium exchange, structure, and function in cultured adult myo- cardial cells. Am. J Physiol. 252 (Heart Circ. Ph,ysioZ. 21): H314- H324,1987.

    74. LEWARTOWSKI, B., R. G. HANSFORD, G. A. LANGER, AND E. G. LAKATTA. Contraction and sarcoplasmic reticulum Ca2’ content in single myocytes of guinea pig heart: effect of ryano- dine. Am. J. Physiol. 259 (Heart Circ. Physiol. 28): H1222-H1229, 1990.

    75. LONDON, B., AND J. W. KRUEGER. Contraction in voltage- clamped internally perfused single heart cells. J Gen. Physiol. 88: 475-505,1986.

    76. MANRING, A., R. NASSAR, AND E. A. JOHNSON. Light diffrac- tion of cardiac muscle: an analysis of sarcomere shortening and muscle tension. J. Mol. CeU. Cardiol. 9: 441-459, 1977.

    77. MARBAN, E., AND H. KUSUOKA. Maximal Ca2+-activated force and myofilament Ca2+ sensitivity in intact mammalian hearts. J. Gen. Physiol, 90: 609-623, 1987.

    78. MARUYAMA, K., S. KIMURA, M. JURODA, AND S. HANDA. Connectin, an elastic protein of muscle. Its abundance in cardiac muscle. J Biochem. 82: 347-350,1977.

    79. MARUYAMA, K., H. SAWADA, S. KIMURA, K. OHASI, H. HI- GUCHI, AND Y. UMAZUME. Connectin filaments in stretched skinned fibers of frog skeletal muscle. J. CeZZ BioZ. 99: 1391-1397, 1984.

    80. MARUYAMA, K., H. T. YOSHIOKA, H. HIGUCHI, K. OHASHI, S. KIMURA, AND R. NATORI. Connectin filaments link thick filaments and Z lines in frog skeletal muscle as revealed by im- munoelectron microscopy. J. CelZ BioZ. 101: 2167-2172, 1985.

    81. MATSUBARA, S., AND K. MARUYAMA. Role of connectin in the length-tension relation of skeletal and cardiac muscles. Jpn. J. Physiol. 27: 589-600, 1977.

    82. MAZET, F., I. DUNIA, G. VASSORT, AND J. L. MAZET. Ultra- structural changes in gap junctions associated with CO, uncou- pling in frog atria1 fibres. J. CeZZ Sci. 74: 51-63, 1985.

    83. MESSIMA, D. A., AND L. F. LEMANSKI. Immunocytochemical studies of spectrin in hamster cardiac tissue. Cell Motil Cytoskele- ton 12: 139-149,1989.

    84. MONTINI, J., G. J. BAGBY, AND J. J. SPITZER. Importance of exogenous substrates for the energy production of adult rat heart myocytes. J. Mol. Cell Cardiol. 13: 903-911, 1981.

    85. NAG, A. C., AND R. ZAK. Dissociation of adult mammalian heart into single cell suspension: an ultrastructural study. J Anat. 129: 541-559,1979.

    86. NASSAR, R., A. MANRING, AND E. A. JOHNSON. Light diffrac- tion of cardiac muscle: sarcomere motion during contraction. In: The Physiological Basis of Starling’s Law of the Heart. Amster- dam: Elsevier, 1974, p. 57-82. (Ciba Found. Symp. 24)

    87. NASSAR, R., M. C. REEDY, AND P. A. W. ANDERSON. Develop- mental changes in the ultrastructural and sarcomere shortening of the isolated rabbit ventricular myocyte. Circ. Res. 61: 465-483, 1987.

  • 88. NICHOLS, C. G. The influence of ‘diastolic’ length on the contrac- of sarcomere length from isolated cardiac cells. Am. J Physiol. tility of isolated cat papillary muscle. J. Physiol. Lond. 361: 269- 242 (Heart Circ. Physiol. 11): H68-H78, 1982. 279,1985. 112. ROOS, K. P., AND J. M. PARKER. A low cost two-dimensional

    89. NIGGLI, E. A laser diffraction system with improved sensitivity digital image acquisition sub-system for high speed microscopic for long-time measurements of sarcomere dynamics in isolated