Http://www.abbs.info e-mail:[email protected]

ISSN 0582-9879                 ACTA BIOCHIMICA et BIOPHYSICA SINICA 2003, 35(9): 793-800                   CN 31-1300/Q

 

The Characterization of Ca2+-calmodulin Independent Phosphorylation of Myosin Light Chains by a Fragment from Myosin Light Chain Kinase

YANG Jing-Xian, WANG Xiao-Ming, TANG Ze-Yao, CHEN Hua, XU Hong, LIN Yuan*

( Department of Pharmacology, Dalian Medical University, Dalian 116027, China)

 

Abstract        A constitutively active myosin light chain kinase (MLCK) fragment (MLCKF) was found to phosphorylate myosin light chains (MLC20) in a Ca2+-CaM independent way more effectively than the intact MLCK. The MLCKF was prepared by tryptic digestion of MLCK. Western blot was used to demonstrate the homogeneity of trypsin-digested MLCKF and intact MLCK. Phosphorylation of MLC20 was detected by Gly-PAGE and Scoin Image Software, and Mg2+-ATPase activity of myosin was measured with spectrophotometry. Our results indicated that Ca2+-CaM independent phosphorylation of myosin (CIPM) by MLCKF was more efficient than CIPM by MLCK and less efficient than Ca2+-CaM dependent phosphorylation of myosin (CDPM) by MLCK in phosphorylating MLC20 and stimulating myosin Mg2+-ATPase activity; both CIPM by MLCKF and CIPM by MLCK were less influenced by the rise of incubation-temperature, the prolonging of incubation-time, the increase of ionic strength of KCl and less sensitive to MLCK inhibitor ML-9 [1-(5-chloronaphthalene-1-sulfonyl)-1H-hexahydro-1,4-diazepine] than CDPM by MLCK. The differences were statistically significant (P<0.01, or P<0.05). The results may be valuable to further investigating the mechanisms of sustained tension characterized by less energy consumption.

Key words     sproteolysis; myosin light chain kinase fragment (MLCKF); phosphorylation; Mg2+-ATPase activity

 

It is generally accepted that Ca2+-CaM dependent myosin phosphorylation and dephosphorylation is the principal mechanism of regulation of smooth muscle contraction and relaxation. The rise in [Ca2+]i results in rapid binding of Ca2+ to calmodulin (CaM) and leading to activation of myosin light chain kinase (MLCK). The activated MLCK catalyzes phosphorylation of myosin light chains(MLC20) and triggers cross-bridges cycling and the development of contractile force. Return of [Ca2+]i to resting level followed by dissociation of [Ca2+]i from CaM, inactivation of MLCK and dephosphorylation of MLC20 catalyzed by myosin light chain phosphatase(MLCP), results in relaxation of smooth muscle. Therefore, the view was held for many years that contractile force in smooth muscle was proportional to cytosolic [Ca2+]i concentrations[15].

However, this simple on/off system based on Ca2+-CaM dependent phosphorylation of myosin (CDPM) is not sufficient to explain all the aspects of smooth muscle contractile activity. It is known that the rise in [Ca2+]i and CDPM is in a very short time in physiological conditions, while the tension may keep much longer time without change in [Ca2+]i[6]. The mechanism of this phenomenon is still unclear. It’s supposed that other mechanisms different from CDPM by MLCK may be involved in the force remaining[79]. Our previous study suggested that Ca2+-CaM independent phosphorylation of myosin (CIPM) by MLCK might contribute to the sustained tension[10]. Recently, what we found new was that a constitutively active MLCK fragment (MLCKF) was more efficient than intact MLCK to phosphorylate MLC20 in a Ca2+-CaM independent way. The following experiments are carried out to reveal the characterization of CIPM by the MLCKF compared with CIPM and CDPM by MLCK.

 

1    Materials and Methods

1.1   Materials

Trypsin, diisopropyl fluorophosphates (DFP), phenylmethyl sulfonyl fluoride (PMSF), 1-(5-chloronaphthalene-1-sulfonyl)-1H-hexahydro-1, 4-diazepine (ML-9) and dithiothreitol (DTT) were purchased from Sigma. Ethylene glycol bis (2-aminoethyl ether) tetraacetic acid (EGTA) was purchased from Wako. Calmodulin (CaM) was generously provided by Prof. K. Kohama, Gumma University, School of Medicine, Japan. Scp70H Centrifuge was made in Hitachi and UV-120-02 Spectrophotometer was the product of Shimadzu.

1.2   Protein purification

Myosin and MLCK used in the assay were purified from fresh chicken gizzard smooth muscle and the methods were described previously[1115].

1.3   Tryptic hydrolysis of MLCK

The hydrolysis of MLCK was carried out according to the methods of references [3,16,17] with slight modification. Native MLCK was incubated with trypsin [11000 (W/W ) ratio of trypsinMLCK] at 3 °C for 15 min in 1 mmol/L DTT, 0.1 mmol/L PMSF, 20 mmol/L Tris·HCl (pH 7.4), and 0.5 mmol/L EGTA. The digestion was terminated by the addition of DFP to the final concentration of 1 mmol/L. The polypeptic mixture of MLCK digested by trypsin was applied to DEAE-52 chromatographic column and then the MLCKF about 61 kD was collected according to the band showed in SDS-PAGE.

1.4   Western blot analysis

Native MLCK, tryptic fragment of MLCK collected by application of DEAE-52 chromatographic column and the polypeptic mixture of MLCK digested by trypsin were separated by SDS-PAGE, and then these proteins were transblotted onto a nitrocellulose membrane. The membrane blocked with 50 g/L skimmed milk powder was incubated with primary antibody [rabbit anti-MLCK IgG, 1:500 (V/V)] and then reacted with horseradish peroxidase-conjugated secondary antibody [goat anti-rabbit IgG, 11000 (V/V)]. Proteins bound with the anti-MLCK IgG were detected by means of the peroxidase reaction using 3,3-diaminobenzidine (DAB) as a color substrate[3,17].

1.5 Phosphorylation of myosin light chains

CIPM was carried out according to the method of references [18, 19] in a 20 mmol/L Tris·HCl (pH 7.4) buffer containing 1 mmol/L DTT, 5 mmol/L MgCl2, 60 mmol/L KCl, 2 mmol/L EGTA, 2 mmol/L ATP and 4 mmol/L myosin. Various concentrations of MLCKF or MLCK, different incubation-time, different incubation-temperature, different concentrations of KCl and different concentrations of MLCK inhibitor ML-9 for phosphorylation of myosin were described in detail in the corresponding figure legends. The assay condition for CDPM was same as that for CIPM with the exception of adding CaCl2 and CaM to a final concentration of 0.1 mmol/L and 5 mg/L respectively instead of 2 mmol/L EGTA.

1.6 Phosphorylation determination

After phosphorylation of MLC20 in both Ca2+-dependent and Ca2+-independent ways, solid urea and sample solution which contained bromophenol blue and glycerol were added to reaction mixture. 10% Gly-PAGE was used to measure the extent of phosphorylation of MLC20 and Scoin Image Software, a densitometry Software gotten from Scion Co. Ltd. was applied to measure the relative phosphorylation extent of MLC20.

1.7 Measurement of myosin Mg2+-ATPase activity

The myosin Mg2+-ATPase activities of CIPM by MLCKF, CDPM by MLCK, CIPM by MLCK and dephosphorylated myosin were measured according to the method of references[2022]. The assays were carried out with 4 mmol/L myosin, 2 mmol/L MLCKF or MLCK at 25 °C for 20 min.

1.8 Other procedures

A polyclonal antibody against MLCK was obtained by injecting MLCK to rabbit together with complete Freuds adjvant (purchased from Sigma).

Protein concentrations were determined by the method of Bradford[23] using bovine serum albumin as the standard.

The results of experiments are expressed as x±s and Student's t-test was used to evaluate the significance of differences.

 

2    Results

2.1 The hydrolysis of MLCK and Western blot analysis

Fig.1(A) showed that intact MLCK purified from chicken gizzard smooth muscle was about 108 kD (Lane 2); trypsin-digested MLCKF obtained by application of DEAE-52 chromatographic column was approximately 61kD (Lane 3). Tryptic proteolysis of MLCK which was stopped by addition of DFP to the final concentration of 1mmol/L produced polypeptides as follows (Lane 4): big fragment about 61 kD and some small fragments from 17 kD to 31 kD. All of the molecular weights were evaluated compared to molecular weight standards (Lane 1).

the antibody of MLCK recognized not only intact MLCK(Lane 1) but also all of the tryptic MLCKFs (Lane 2, 3). This indicated that these fragments were all the products of digested MLCK.

 

Fig.1       The hydrolysis of MLCK and Western blot analysis

(A) Proteins separated on SDS-PAGE. 1, molecular weight standard; 2, purified intact MLCK; 3, trypsin-digested MLCKF collected by application of DEAE-52 chromatographic column; 4, proteolytic mixture of MLCK digested by trypsin. (B) The proteins described above which were transbloted from SDS-PAGE onto a nitrocellulose membrane. The MLCK antibodies recognized intact MLCK (Lane 1), trypsin-digested MLCKF collected by application of DEAE-52 chromatographic column (Lane 2) and all of the mixed fragments following the tryptic proteolysis of MLCK (Lane 3).

 

2.2   The characteristics of CIPM by MLCKF compared with CDPM by MLCK and CIPM by MLCK

2.2.1       The comparison between CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK in different concentrations    To find out the differences between CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK, 0.00002 mmol/L to 2 mmol/L MLCKF and MLCK were selected for the assay. The results showed that the extent of MLC20 phosphorylation in CIPM by different concentrations of MLCKF [Fig.2(C), (D)] was obviously lower than that of CDPM by MLCK [Fig.2(A), (D)], but was higher than that of CIPM by MLCK [Fig.2(B), (D)]. The differences were statistically significant (**P<0.01, #P<0.05, ##P<0.01). These results indicated that CIPM by MLCKF was less efficient than CDPM by MLCK, but was more efficient than CIPM by MLCK.

2.2.2       The comparison between CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK at different incubation-time   To determine whether the incubation-time influences CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK or not, 20 min to 80 min were selected for examining MLC20 phosphorylation. The results showed that with the increase of incubation-time, CDPM by MLCK [Fig.3(A), (D)] showed an apparently declined extent of MLC20 phosphorylation(**P<0.01), but CIPM by MLCKF [Fig.3(C), (D)] and CIPM by MLCK [Fig.3(B), (D)] showed no tendency of declined MLC20 phosphorylation (#P>0.05). This proved that CIPM by MLCKF and CIPM by MLCK were less influenced by the prolonging of incubation-time than CDPM by MLCK.

 

Fig.2       The comparison between CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK in different concentrations (x±s, n=6)

The assay was carried out with 4 mmol/L myosin and various concentrations of MLCKF and MLCK at 25 °C for 20 min. (A), (B) and (C) represent glycerol electrophoresis results of CDPM by MLCK, CIPM by MLCK and CIPM by MLCKF, respectively. a0, b0 and c0 represent dephosphorylation controls; a1 and b1, a2 and b2, a3 and b3, a4 and b4 were added with 2 mmol/L, 0.02 mmol/L, 0.0002 mmol/L and 0.00002 mmol/L MLCK, respectively. c1, c2, c3 and c4 were added with the same concentrations of MLCKF, respectively. LC20 represents dephosphorylated 20 kD myosin light chains; p-LC20 represents mono-phosphorylated 20 kD myosin light chains; pp-LC20 represents di-phosphorylated 20 kD myosin light chains. LC17 represents 17 kD myosin essential light chains. The extent of MLC20 phosphorylation of CDPM by MLCK (filled rhombus), CIPM by MLCKF (filled squares) and CIPM by MLCK (filled triangles) were plotted against the concentrations of MLCKF and MLCK (D). **P<0.01 vs. CDPM by MLCK, #P<0.05, ##P<0.01 vs CIPM by MLCK.

 

Fig.3 The comparison between CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK at different incubation-time (x±s, n=6)

Samples were incubated with 4 mmol/L myosin, 2 mmol/L MLCKF and MLCK at 25 °C for different incubation-time. (A), (B) and (C) represent glycerol electrophoresis results of CDPM by MLCK, CIPM by MLCK and CIPM by MLCKF, respectively. Four different incubation-time, i.e., a1, b1 and c1=20 min, a2, b2 and c2=40 min, a3, b3 and c3=60 min, a4, b4 and c4=80 min were selected for the assay. a0, b0 and c0 represent dephosphorylation controls; The extent of MLC20 phosphorylation of CDPM by MLCK (filled rhombus), CIPM by MLCKF (filled squares) and CIPM by MLCK (filled triangles) were plotted against the incubation-time (D). **P<0.01, #P>0.05 vs. the corresponding controls at 20 min.

 

2.2.3       The comparison between CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK at different incubation-temperature       To observe the influences of incubation-temperature on CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK, 15 , 25 , 35 , 45 were chosen for determining MLC20 phosphorylation. It was shown that with the increase of incubation-temperature, CDPM by MLCK [Fig.4(A), (D)] showed a significantly declined tendency of MLC20 phosphorylation (**P<0.01); While CIPM by MLCKF [Fig.4(C), (D)] and CIPM by MLCK [Fig.4(B), (D)] didn't show apparent declined extent of MLC20 phosphorylation(#P>0.05). These results suggested that CIPM by MLCKF and CIPM by MLCK were more sustained and less influenced by the change of incubation-temperature than CDPM by MLCK.

 

Fig.4       The comparison between CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK at different incubation temperatures (x±s, n=6)

The assay was carried out with 4 mmol/L myosin, 2 mmol/L MLCKF and MLCK for 20 min at different incubation temperatures. (A), (B) and (C) represent glycerol electrophoresis results of CDPM by MLCK, CIPM by MLCK and CIPM by MLCKF, respectively. Four different incubation temperatures, i.e., a1, b1 and c1=15 , a2, b2 and c2=25 , a3, b3 and c3=35 , a4, b4 and c4=45 were chosen for the assay. a0, b0 and c0 represent dephosphorylation controls. The extent of MLC20 phosphorylation of CDPM by MLCK (filled rhombus), CIPM by MLCKF (filled squares), CIPM by MLCK (filled triangles) were plotted against the incubation temperatures (D). **P<0.01, #P>0.05 vs. the corresponding controls at 25 .

 

2.2.4       The effects of different ionic strength of KCl on CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK   To investigate the influences of different ionic strength of KCl on CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK, 60 mmol/L to 360 mmol/L KCl were selected for the phosphorylation of MLC20. With the increase of KCl, the extent of CDPM by MLCK [Fig.5(A), (D)] showed an apparently decline(**P<0.01); in contrast, no apparent change wasobserved in the extent of CIPM by MLCKF [Fig.5(C), (D)] and CIPM by MLCK [Fig.5(B), (D)] (#P>0.05). This indicated that CIPM by MLCKF and CIPM by MLCK were less influenced by the change of ionic strength of KCl than CDPM by MLCK.

 

Fig.5       The effects of different concentrations of KCl on CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK (x±s, n=6)

The assay was carried out with 4 mmol/L myosin, 2 mmol/L MLCKF and MLCK at 25 for 20 min. (A), (B) and (C) represent glycerol electrophoresis results of CDPM by MLCK, CIPM by MLCK and CIPM by MLCKF, respectively. Four different concentrations of KCl, i.e., a1, b1 and c1 = 60 mmol/L KCl; a2, b2 and c2= 120 mmol/L KCl; a3, b3 and c3=240 mmol/L KCl; a4, b4 and c4 = 360 mmol/L KCl were chosen for the assay. a0, b0 and c0 represent dephosphorylation controls. The extent of MLC20 phosphorylation of CDPM by MLCK (filled rhombus), CIPM by MLCKF (filled squares) and CIPM by MLCK (filled triangles) were plotted against the concentrations of KCl (D). **P<0.01, #P>0.05 vs. the corresponding controls with 60 mmol/L KCl.

 

 

2.2.5       The effects of different concentration of ML-9 on CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK

ML-9 is a MLCK inhibitor. To observe whether the inhibitory effects of ML-9 on CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK were same or not, 0.1 mmol/L and 0.2 mmol/L ML-9 were selected and MLC20 phosphorylation was determined for 20 min and 40 min, respectively. It was observed that, with 0.1 mmol/L ML-9, the effect of MLCK on CDPM [Fig.6(A), (D1), (D2)] was significantly inhibited (*P<0.01); in contrast, no inhibitory effects on CIPM by MLCKF [Fig.6(C), (D1), (D2)] and CIPM by MLCK [Fig.6(B), (D1), (D2)] were observed in the presence of 0.1 mmol/L ML-9 (#P>0.05). With the increase of ML-9 to 0.2 mmol/L, the inhibitory effects of ML-9 on CIPM by MLCKF [Fig.6(C), (D1), (D2)] and CIPM by MLCK [Fig.6(B), (D1), (D2)] appeared (**P<0.01). These results suggested that CIPM by MLCKF and CIPM by MLCK were more stable and less sensitive to ML-9 than CDPM by MLCK.

Fig.6       The effects of different concentrations of ML-9 on CIPM by MLCKF, CDPM by MLCK and CIPM by MLCK (x±s, n=6)

The assay was carried out with 4 mmol/L myosin, 2 mmol/L MLCKF and MLCK at 25 for 20 min. (A), (B) and (C) represent glycerol electrophoresis results of CDPM by MLCK, CIPM by MLCK and CIPM by MLCKF, respectively. Among (A), (B) and (C), different concentrations of ML-9 were selected, i.e., a1, a2, b1, b2, c1 and c2 = 0 mmol/L ML-9, a3, a4, b3, b4, c3 and c4 = 0.1 mmol/L ML-9, a5, a6, b5, b6, c5 and c6 = 0.2 mmol/L ML-9; Two different incubation-time were chosen i.e., a1, a3, a5, b1, b3, b5, c1, c3 and c5 =20 min; a2, a4, a6, b2, b4, b6, c2, c4, and c6=40 min. a0, b0 and c0 represent dephosphorylation controls. The extent of MLC20 phosphorylation of CDPM by MLCK (filled rhombus), CIPM by MLCKF (filled squares), CIPM by MLCK (filled triangles) were plotted against the concentrations of ML-9 (D1, incubation for 20 min; D2, incubation for 40 min). **P<0.01, #P>0.05 vs. the corresponding controls without ML-9.

 

2.2.6       The comparison of myosin Mg2+-ATPase activities between CIPM by MLCKF, CDPM by MLCK, CIPM by MLCK and dephosphorylated myosin

It was shown in Fig.7 that myosin Mg2+-ATPase activitity of CIPM by MLCKF (Column 3) was lower than that of CDPM by MLCK (Column 4)(**P<0.01), but was higher than those of CIPM by MLCK (Column 2) and dephosphorylated myosin (Column 1) (##P<0.01). This indicated that CIPM by MLCKF was less efficient than CDPM by MLCK but was more efficient than CIPM by MLCK and dephosphorylated myosin to stimulate myosin Mg2+-ATPase activity.

 

Fig.7       The relative Mg2+-ATPase activities of CIPM by MLCKF, CDPM by MLCK, CIPM by MLCK and dephosphorylated myosin (x±s, n=6)

1, dephosphorylatd myosin; 2, CIPM by MLCK; 3, CIPM by MLCKF; 4, CDPM by MLCK. Fig.7 represents the relative Mg2+-ATPase activities of dephosphorylated myosin (Column 1), CIPM by MLCK (Column 2), CIPM by MLCKF (Column 3) and CDPM by MLCK (Column 4). The assay was carried out with 4 mmol/L myosin, 2 mmol/L MLCKF and MLCK at 25 for 20 min. It was designed that the Mg2+-ATPase activity of dephosphorylated myosin was 100%. The others were relative value comparing to the Mg2+-ATPase activity of dephosphorylated myosin. **P<0.01 vs. CDPM by MLCK, ##P<0.01 vs. CIPM by MLCK and dephosphorylated myosin.

 

3    Discussion

It was previously thought that contractile activity of smooth muscle is controlled primarily by the reversible CDPM by MLCK. However, this simple view is not sufficient to explain all the aspects of smooth muscle contractile activity, especially as concerns the sustained tension, termed tonic contraction[7,2426]. Coirault et al.[8] and Hai et al.[2729] suggested that dephosphorylated myosin which attend to form a special slow cycling crossbridges, termed "latch bridges", involve in tonic contraction of smooth muscle. Two facts encouraged us to make the further investigation. One was that our previous study suggested that CIPM by MLCK may contribute to the tonic contraction[10]; the other was that Weber et al.[3] described that the MLCKF prepared by tryptic digestion of MLCK had a specific activity to phosphorylate MLC20 both in the presence and in the absence of Ca2+. Though our previous study suggested that in the presence of Ca2+, CDPM by MLCKF was more efficient than CDPM by MLCK(data not shown), but the characterization of CIPM by MLCKF in the absence of Ca2+ still remained unclear. To further explore the characterization of CIPM by MLCKF, we prepared the constitutively active MLCK fragment (MLCKF) according to the methods of Weber et al.[3] and used it in our assay. It was found that CIPM by MLCKF was more efficient than CIPM by MLCK and was less efficient than CDPM by MLCK in phosphorylating MLC20 and stimulating myosin Mg2+-ATPase activity; both CIPM by MLCKF and CIPM by MLCK were less influenced by the rise of incubation-temperature, the prolonging of incubation-time, the increase of ionic strength of KCl and less sensitive to MLCK inhibitor ML-9 than CDPM by MLCK. These results may be useful to further investigate the mechanism of sustained tension characterized by less energy consumption.
As to the enhancement of the activity of MLCKF in phosphorylating MLC20, a reasonable explanation might be as follows. Previous studies on the proteolysis of MLCK showed that the autoinhibitory domain at the C-terminal of MLCK might be removed by tryptic cleavage, but the catalytic domain might be retained[3,17,30]. We digested MLCK with trypsin according to the methods of above researchers and obtained the tryptic MLCKF which had approximately the same molecular weight (about 61 kD) as the tryptic MLCKF reported previously (61 kD). Western blot has demonstrated that the tryptic MLCKF we obtained was homogenous with intact MLCK. Therefore, we could not rule out the possibility that, in our study, the enhancement of the activity of trypsin-digested MLCKF in phosphorylating MLC20 was due to the removal of C-terminal of MLCK which contained autoinhibitory domain. To confirm this possibility, it is worth to make further investigation to identify the amino acid sequence of the MLCKF.
We also found that during the isolation of MLCK in the absence of PMSF, part of MLCK was automatically cleaved into a constitutively active MLCKF (about 61 kD), which had the same specific activity in phosphorylating MLC20 and the same molecular weight as the tryptic MLCKF we obtained (61 kD). Western blot demonstrated that the automatically proteolytic MLCKF and trypsin-digested MLCKF were all homogenous.

 

References

1     Rembold CM. Regulation of contraction and relaxation in arterial smooth muscle. Hypertension, 1992, 20(2): 129137

2     Walsh MP. The Ayerst Award Lecture 1990. Calcium-dependent mechanisms of regulation of smooth muscle contraction. Biochem Cell Biol, 1991, 69(12): 771800

3     Weber LP, Van Lierop JE, Walsh MP. Ca2+-independent phosphorylation of myosin in rat caudal artery and chicken gizzard myofilaments. J Physiol, 1999, 516(Pt 3): 805824

4     Stull JT, Gallagher PJ, Herring BP, Kamm KE. Vascular smooth muscle contractile elements. Cellular regulation. Hypertension, 1991, 17(6 Pt 1): 723732

5     Kamm KE, Stull JT. The function of myosin and myosin light chain kinase phosphorylation in smooth muscle. Annu Rev Pharmacol Toxicol, 1985, 25: 593620

6     Kumagai H, Kohama K. The activation-mechanism of dephosphorylated myosin of vascular smooth muscle and sustained contraction. Vascular Biology & Medicine, 2001, 2(4): 359366

7     Haeberle, JR. Thin-filament linked regulation of smooth muscle myosin. J Muscle Res Cell Motil, 1999, 20(4): 363370

8     Coirault C, Blanc FX, Chemla D, Salmeron S, Lecarpentier Y. Biomechanics and bio-energetics of smooth muscle contraction. Relation to bronchial hyperreactivity. Rev Mal Respir, 2000, 17(2 Pt 2): 549554

9     Amano M, Ito M, Kimura K, Fukata Y, Chihara K, Nakano T, Matsuura Y et al. Phosphorylation and activation of myosin by Rho-associated kinase (Rho-kinase). J Biol Chem, 1996, 271(34): 2024620249

10    Lin Y, Tang ZY, Chen H, Wang XM, Yang JX. Ca2+-independent phosphorylation of smooth muscl myosin by myosin light chain kinase. Science Technology and Engineering, 2002, 2(6): 3739

11    Jiang H, Stephens NL. Calcium and smooth muscle contraction. Mol Cell Biochem, 1994, 135(1): 19

12    Ye LH, Hayakawa K, Kishi H, Imamura M, Nakamura A, Okagaki T, Takagi T et al. The structure and function of the actin-binding domain of myosin light chain kinase of smooth muscle. J Biol Chem, 1997, 272(51): 3218232189

13    Lin Y, Kishi H, Nakamura A, Takagi T, Kohama K. N-terminal myosin-binding fragment of talin. Biochem Biophys Res Commun, 1998, 249(3): 656659

14    Lin Y, Sun HJ, Dai SF, Tang ZY, He X, Chen H. The bi-directional regulation of filamin on the ATPase activity of smooth muscle myosin. Chin Med Sci J, 2000, 15(3): 162164

15    Lin Y, Ishikawa R, Kohama K. Role of myosin in the stimulatory effect of caldesmon on the interaction between actin, myosin, and ATP. J Biochem (Tokyo), 1993, 114(2): 279283

16   Foyt HL, Guerriero V Jr, Means AR. Functional domains of chicken gizzard myosin light chain kinase. J Biol Chem, 1985, 260(12): 77657774

17    Numata T, Katoh T, Yazawa M. Functional role of the C-terminal domain of smooth muscle myosin light chain kinase on the phosphorylation of smooth muscle myosin. J Biochem (Tokyo), 2001, 129(3): 437444

18    Okagaki T, Higashi-Fujime S, Ishikawa R, Takano-Ohmuro H, Kohama K. In vitro movement of actin filaments on gizzard smooth muscle myosin: Requirement of phosphorylation of myosin light chain and effects of tropomyosin and caldesmon. J Biochem (Tokyo), 1991, 109(6): 858866

19    Perrie WT, Perry SV. An electrophoretic study of the low-molecular-weight components of myosin. Biochem J, 1970, 119(1): 3138

20    Lin Y, Ishikawa R, Okagaki T, Ye LH, Kohama K. Stimulation of the ATP-dependent interaction between actin and myosin by a myosin-binding fragment of smooth muscle caldesmon. Cell Motil Cytoskeleton, 1994, 29(3): 250258

21    Ishikawa R, Okagaki T, Higashi-Fujime S, Kohama K. Stimulation of the interaction between actin and myosin by physarum caldesmon-like protein and smooth muscle caldesmon. J Biol Chem, 1991, 266(32): 2178421790

22    Kodama T, Fukui K, Kometani K. The initial phosphate burst in ATP hydrolysis by myosin and subfragment-1 as studied by a modified malachite green method for determination of inorganic phosphate. J Biochem (Tokyo), 1986, 99(5): 14651472

23    Bradford MM. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem, 1976, 72: 248254

24    Paul RJ, Wendt IR, Walker JS, Gibbs CL. Smooth muscle energetics: Testing theories of crossbridge regulation. Prog Clin Biol Res, 1990, 327: 2938

25    Rembold CM. Relaxation, [Ca2+]i, and the latch-bridge hypothesis in swine arterial smooth muscle. Am J Physiol, 1991, 261(1 Pt 1): C41C50

26    Di Blasi P, Van Riper D, Kaiser R, Rembold CM, Murphy RA. Steady-state dependence of stress on cross-bridge phosphorylation in the swine carotid media. Am J Physiol, 1992, 262(6 Pt 1): C1388C1391

27    Hai CM, Murphy RA. Cross-bridge phosphorylation and regulation of latch state in smooth muscle. Am J Physiol, 1988, 254(1 Pt 1): C99C106

28    Paul RJ. Smooth muscle energetics and theories of cross-bridge regulation. Am J Physiol, 1990, 258(2 Pt1): C369C375

29    Hai CM, Murphy RA. Regulation of shortening velocity by cross-bridge phosphorylation in smooth muscle. Am J Physiol, 1988, 255(1 Pt 1): C86C94

30    Ito M, Dabrowska R, Guerriero V Jr, Hartshorne DJ. Identification in turkey gizzard of an acidic protein related to the C-terminal portion of smooth muscle myosin light chain kinase. J Biol Chem, 1989, 264(24): 1397113974

 


Received: February 28, 2003        Accepted: June 3, 2003

This work was supported by a grant from the National Natural Science Foundation of China ( No. 30070203)

*Corresponding author: Tel, 86-411-4720652; Fax, 86-411-4673573; e-mail, [email protected] or [email protected]