©1995 by The American Society for Biochemistry and Molecular Biology, Inc.
Crystal Structure of a Complement Factor D Mutant Expressing Enhanced Catalytic Activity (*)

(Received for publication, April 4, 1995; and in revised form, July 31, 1995)

Sunghee Kim (1) Sthanam V. L. Narayana (2) John E. Volanakis (1)(§)

From the  (1)Division of Clinical Immunology and Rheumatology, Department of Medicine and Department of Microbiology, and the (2)Center for Macromolecular Crystallography and School of Optometry, University of Alabama at Birmingham, Birmingham, Alabama 35294-0006

ABSTRACT
INTRODUCTION
MATERIALS AND METHODS
RESULTS
DISCUSSION
FOOTNOTES
REFERENCES

ABSTRACT

Complement factor D is a serine protease regulated by a novel mechanism that depends on conformational changes rather than cleavage of a zymogen for expression of proteolytic activity. The conformational changes are presumed to be induced by the single natural substrate, C3bB, and to result in reversible reorientation of the catalytic center and of the substrate binding site of factor D, both of which have atypical conformations. Here we report that replacement of Ser, Thr, and Ser of factor D (chymotrypsinogen numbering has been used for comparison purposes) with the corresponding residues of trypsin, Tyr, Ser, and Trp, is sufficient to induce substantially higher catalytic activity associated with a typical serine protease alignment of the catalytic triad residues His, Asp, and Ser. These results provide a partial structural explanation for the low reactivity of ``resting-state'' factor D toward synthetic substrates.


INTRODUCTION

Activation of complement leads to expression of important host defense functions and proceeds via pathways that consist of successive enzymatic amplification steps. In the alternative pathway, the first enzymatic reaction is catalyzed by factor D, a serine protease that cleaves factor B only in the context of a complex with C3b, a fragment of the third component of complement, C3(1) . Factor D is unique among serine proteases in that it requires neither enzymatic cleavage for expression of proteolytic activity nor inactivation by an inhibitor for its control. Regulation of factor D activity in blood is apparently attained by a novel mechanism that depends on conformational changes and allows for reversible expression of proteolytic activity(2, 3) . The putative conformational changes are believed to be induced by the natural substrate, C3bB, and to result in reorientation of the catalytic center and of the substrate binding site of factor D, both of which have atypical structures(4) . Functional support for this hypothesis has been provided by the seemingly parodoxical observation that, while the reactivity of factor D with thioester substrates and active site inhibitors is 3-4 orders of magnitude lower than that of typical serine proteases(5, 6) , its proteolytic activity during complement activation is comparable to that of other complement enzymes. Obviously, esterolytic assays assess the resting-state inactive conformation of the enzyme, while complement activation assays the substrate-induced proteolytically active conformation. Structural support for the inactive resting-state conformation hypothesis has been provided by the recently determined crystal structure of factor D(4) .

The two non-crystallographically related molecules, A and B, present in the triclinic unit cell of factor D have typical serine protease structural folds. However, they display distinctive orientations of the side chains of the catalytic triad residues Asp and His (chymotrypsinogen numbering (^1)has been used throughout this paper), while the orientation of the third member of the triad, Ser, is similar to that of other serine proteases. In all serine proteases of known structure, the spatial relationships of these three residues are constant and essential for the formation of a functional unit responsible for catalytic activity(7, 8, 9, 10) . By contrast, in molecule A of factor D, the carboxylate of Asp is pointed away from His and is freely accessible to the solvent. In molecule B, the imidazolium of His is oriented away from Ser, having assumed the energetically favored trans conformation(4) . Neither of these orientations would allow expression of catalytic activity, indicating the need for a realignment of the catalytic triad residues, probably induced by the single natural substrate, C3b-complexed factor B. A possible structural explanation for the unusual disposition of Asp/His in factor D is provided by the substitution of a Ser for the bulky aromatic Tyr or Trp residue usually present at position 94 of serine proteases and also of a Thr for the invariant Ser (Fig. 1). Both of these residues are among those identified by Blow et al.(7) as being important for shielding Asp from solvent molecules in chymotrypsin. In addition, a Ser residue that substitutes for the conserved aromatic Trp or Phe at position 215 may also be a determinant of the unique conformation of resting-state factor D. Indeed, in molecule B of factor D, Ser is positioned between Asp and Ser, preventing His from assuming the active gauche conformation that is characteristic of serine proteases. Here we report that replacement of Ser, Thr, and Ser of factor D with the corresponding residues of trypsin, Tyr, Ser, and Trp, is sufficient to induce typical serine protease alignment of the catalytic triad residues and substantially higher catalytic activity. These results provide a structural exegesis for the unique conformation of the active center of resting-state factor D and also for its low reactivity with synthetic esters. The data also demonstrate the previously unrecognized importance of residue 94 of serine proteases for catalysis.


Figure 1: Alignment of partial amino acid sequences of selected serine proteases. Residues participating in shielding the catalytic residue Asp from the solvent (7) are shaded. Conserved residues are boxed. Asterisks indicate amino acid residues subjected to site-directed mutagenesis. Numbers at the top are for residues of the chymotrypsinogen sequence, while the numbers at the bottom are for the human factor D residues. HFD, human factor D; TRP, bovine pancreas trypsin; CHT, bovine pancreas chymotrypsin; ELA, porcine pancreas elastase. Sequence data are from (37) , except for the human factor D sequence, which is taken from (11) .




MATERIALS AND METHODS

Construction of Wild-type (wt) and Mutant Factor D Recombinant Plasmids

The human factor D cDNA hg31 (11) was cloned into the unique HindIII site of the eukaryotic expression vector pRc/CMV (Invitrogen, San Diego, CA) as described previously(12) . The S94Y mutant was constructed from the wt (^2)template by using the mutagenic oligonucleotide 5`-TGGTGTCGGGCTGGTAGTCCGGGTGG-3` (mutated nucleotides are boldface and underlined). The combination mutants S94Y/T214S, S94Y/S215W, and S94Y/T214S/S215W were constructed by using the S94Y mutant cDNA as template and the mutagenic oligonucleotides 5`-GCGAGCCCGAGGAGACCACGCCC-3`, 5`-CGAGCCCCAGGTGACCACGC-3`, and 5`-AACGCGCGAGCCCCAGGAGACCACGCCC-3`, respectively. Mutants were constructed according to the method of Kunkel(13) . All mutations were verified by nucleotide sequencing using the chain termination method(14) . Oligonucleotides were synthesized by using a model 394 DNA/RNA synthesizer (Applied Biosystems, Foster City, CA).

Expression and Purification of wt and Mutant Recombinant Factor D

Thirty µg of each recombinant pRc/CMV plasmid containing wt or mutant factor D cDNA were transfected into 4 times 10^6 Chinese hamster ovary (CHO) cells by electroporation at 1500 V and 25 microfarads by using a Gene Pulser apparatus (Bio-Rad). The transfected CHO cells were grown in Dulbecco's modified Eagle's medium/Ham's F-12 (1:1) supplemented with 10% heat-inactivated fetal calf serum, 2 mM glutamine, 100 units/ml penicillin, and 100 µg/ml streptomycin in 8% CO(2) at 37 °C. After 48 h of culture, stably transfected CHO cells were selected by growth for 2 weeks in the above medium supplemented with increasing concentrations (0.4-1 mg/ml) of Geneticin-418 (Life Technologies, Inc.). Expression of factor D was assessed by a solid phase enzyme-linked immunosorbent assay (ELISA) using the anti-factor D monoclonal antibody FD10-1 in the solid-phase and the IgG fraction of a rabbit anti-factor D serum as detection reagent as described previously(12) . For purification of factor D, transfected CHO cells were grown in serum-free CHO medium (Life Technologies, Inc.), supplemented with 10 mM HEPES in 1-liter spinner flasks at 37 °C. Typically, after 5-7 days of culture, supernatants of pooled transfectants contained approximately 1 µg/ml factor D, which was purified by successive chromatography on Bio-Rex 70 (Bio-Rad) and Mono S HR 5/5 (Pharmacia Biotech Inc.) as described(15) . For crystallization of the triple mutant S94Y/T214S/S215W, a single transfectant was isolated by a series of subcloning steps. The stable transfectant secreted approximately 5-10 µg/ml factor D in 5 days.

Hemolytic Assay for Factor D

Sheep erythrocytes carrying human C3b (EC3b) were prepared as described previously(16) . The complement proteins C3(17) , factor B(18) , and properdin (19) were purified by previously described methods. Native factor D was purified from the urine of a patient with Fanconi's syndrome as described(20) . Hemolytic titrations were performed by incubating 7.5 times 10^6 EC3b with properdin (62 ng), factor B (250 ng), and appropriately diluted factor D samples in 200 µl of half-strength Veronal-buffered saline, pH 7.3, containing 2.5% dextrose, 2.5 mM MgCl(2), 10 mM EGTA, and 0.1% gelatin. The mixtures were incubated at 37 °C for 30 min and then 250 µl of guinea pig serum diluted 1/40 in Veronal-buffered saline (150 mM NaCl, 5 mM Veronal, pH 7.3) containing 10 mM EDTA and 0.1% gelatin (EDTA-GVB) was added. The mixtures were further incubated at 37 °C for 30 min. The reaction was stopped by adding 500 µl of ice-cold EDTA-GVB, and the percentage of lysis was calculated from the absorbance of the supernatants at 413 nm and used to calculate hemolytic units/ml.

Thioester Hydrolysis by Factor D

The assay measuring the rates of hydrolysis of the thioester substrate Z-Lys-SBzl by factor D was performed as described by Kam et al.(5) . Z-Lys-SBzl was purchased from Calbiochem (San Diego, CA). Ellman's reagent 5,5`-dithiobis-(2-nitrobenzoic acid) (Aldrich) was used at a concentration of 20 mM as a chromogen of thioester hydrolysis(21) . Assays were conducted in microtiter wells and the rate of hydrolysis was measured as absorbance at 405 nm by using a Vmax(TM) kinetic microplate reader (Molecular Devices, Menlo Park, CA). The kinetic constants were calculated from Lineweaver-Burk plots.

Crystallization and Data Collection

Diffraction quality crystals of the S94Y/T214S/S215W mutant were grown at 22 °C by the vapor diffusion method by using 50 mM MES buffer and polyethylene glycol 6000 as precipitant. Crystals grew to a size of 0.3 mm times 0.3 mm times 0.5 mm in 3 days. Preliminary crystallographic analysis indicated that the crystals belonged to the hexagonal space group P3(1)21 with unit cell dimensions a = b = 45.4 Å, c = 175.2 Å. The calculated solvent content (22) was 43%, indicating 1 molecule/asymmetric unit. Diffraction data were collected on a Xentronics (Siemens/Nicolet, Madison, WI) area detector, using Cu-Kalpha radiation (40 mA, 100 kV) from a rotating anode generator. Under these conditions the crystals diffracted to a resolution of 2.0 Å. The XENGEN program package (23) was used to process the data.

Structure Determination

The structure of the S94Y/T214S/S215W mutant was solved by molecular replacement methods (24) using the atomic coordinates of native factor D (4) as search model. The starting coordinates were modified by the removal of the side chains of the three mutated residues and also of His, Asp, Asp, Ser, Ser, and Arg. Coordinates of solvent atoms were also excluded in molecular replacement studies. When molecule B of the native factor D triclinic cell was used for molecular replacement, the cross-rotation function calculations (8.0 to 4.0 Å resolution) yielded higher and cleaner peaks than when molecule A was used. Molecule B coordinates were therefore used for the molecular replacement studies. The search model was rotated and translated into the unit cell of the mutant by an R-factor search using XPLOR(25) , which gave an R-factor of 38%. The positional and orientational parameters of the model were refined by using the rigid body refinement protocol in XPLOR (8.0 to 4.0 Å resolution), which gave a R-factor of 36%. This model was subjected to least square refinement of individual atoms with the PROLSQ program (26) using 8.0 to 3.0 Å data. A difference map constructed by using calculated phases to 3.0-Å resolution revealed the positions of the missing side chains (except for residues 216, 217, and 218). The model was then refined further by a combination of simulated annealing using XPLOR and restrained least-square refinement using PROLSQ. The current refined model includes residues 16-243 and 59 water molecules. The model has been fit by visual inspection of electron density maps (2F(o) - F(c)) and (F(o) - F(c)) computed using calculated phases(27) . Table 1summarizes the crystal and x-ray diffraction data and the final refinement statistics.




RESULTS

Four mutant factor D cDNAs, S94Y, S94Y/T214S, S94Y/S215W, and S94Y/T214S/S215W, were constructed for these studies. The mutant and wt factor D cDNAs were stably expressed in CHO cells, and the recombinant proteins were purified to homogeneity. The effects of the mutations on the catalytic efficiency of resting-state factor D were evaluated by an esterolytic assay using the thioester substrate Z-Lys-SBz1(12) . Results of multiple experiments are summarized in Table 2. Previously reported results for the T214S, S215W, and T214S/S215W mutants of factor D (12) have been included for comparison. As shown, mutation of Ser to Tyr in the single S94Y mutant resulted in about 6-fold increase of the catalytic rate constant (k). Larger increases in k were observed when the S94Y mutation was combined with mutations of Thr to Ser and of Ser to Trp. The triple mutant S94Y/T214S/S215W had a k more than 16-fold higher than wt factor D and only 3.5-fold lower than trypsin (Table 2). The enhanced k of all S94Y mutants and of the double T214S/S215W mutant resulted in increased overall catalytic efficiency as assessed by the k/K(m) ratio, although the K(m) was affected only slightly if at all.



The proteolytic activity of mutant factor D was assessed by a sensitive hemolytic assay, and the results (Table 3) were in general agreement with those obtained with the esterolytic assay. The S94Y mutant had higher hemolytic activity than wt factor D, and its activity was synergistically increased in combinations with the S215W and the double T214S/S215W mutations. As expected, all mutants cleaved factor B complexed with cobra venom factor, a C3b analogue of cobra venom, but none cleaved uncomplexed factor B (data not shown).



To investigate the structural correlates of the increased catalytic activity of mutant factor D, we determined the crystal structure of the triple mutant S94Y/T214S/S215W. The model has been refined to an R-factor of 19.8% (Table 1) and is in agreement with the most favored region of a Ramachandran plot (28) (Fig. 2). Fig. 3shows the electron density for the mutated residues 94, 214, and 215 in a difference map calculated by using molecular replacement phases (8.0 to 3.0 Å). Plots of mean temperature factors (B) per residue and of real-space density fit per residue (R) are shown in Fig. 4(a and b, respectively). The highest B-factors and worst real-space R-factors are observed in the segment His-Ile. In contrast, in both molecules A and B of native factor D, the B-factors for this segment were lower than the average values(4) . In typical serine proteases residue 172 is a highly conserved aromatic (Tyr or Trp), which is part of a type I beta-turn and has been shown to be a determinant of substrate specificity(29) . Thus, the apparent flexibility of His in the mutant may have an effect on substrate binding. The second highest B-factors were observed for the Asp-Val segment of mutant factor D. This region is flexible in molecule A, but quite rigid in molecule B of native factor D. Compared to chymotrypsin or trypsin, this segment of factor D has 3 or 4 additional residues, respectively.


Figure 2: Ramachandran plot for the refined atomic model of the S94Y/T214S/S215W mutant. Most favored, favored, and additional allowed regions are indicated by progressively lighter shading. Disallowed regions are not shaded. Regions are marked according to the analysis of highly refined structures in the Protein Data Bank. Circles and diamonds represent nonglycine and glycine residues, respectively.




Figure 3: Stereo representation of (F - F) map calculated using 20.0 to 2.5-Å resolution data and molecular replacement phases. Side chains of residues 94, 214, and 215 had been removed in the model used for molecular replacement. Density is clearly seen for mutated residues and for deleted solvent molecules.




Figure 4: Quality of the S94Y/T214S/S215W structure. a, variation of the isotropic B values averaged (Å^2) for the main chain (mc) and side chain (sc) atoms of each residue. b, real-space electron density fit per residue (R) computed using (2F - F) exp(ialpha) map. The map was calculated using 20.0 to 2.0 Å resolution data.



Overall, the tertiary structure of the S94Y/T214S/S215W mutant is very similar to that of native factor D. The root mean square (r.m.s.) deviations between the coordinates of main chain and side chain atoms of the mutant and molecules A and B of factor D are shown in Fig. 5(a and b, respectively). Substantial structural divergence between the mutant and both molecules of native factor D is observed in two segments. The first includes residues 169-175, which in the mutant form a flexible loop thus making the observed r.m.s. shifts an unreliable indicator of actual structural differences. Furthermore, this segment partially overlaps the region (residues 172-177) with the highest B-factors and worst real-space R-factors (Fig. 4, a and b). Excluding residues 172-177, the r.m.s. shifts between the mutant and molecule A (Fig. 5a) are 0.95 Å for main chain atoms and 1.02 Å for side-chain atoms. The corresponding values for the comparison between the mutant and molecule B are 0.86 Å and 1.59 Å. By comparison, the r.m.s. shifts between molecules A and B of native factor D (4) are 0.96 Å for main chain atoms and 1.58 Å for side chain atoms. The second segment with significant r.m.s. deviations between the mutant and both molecules of factor D includes residues 214-224. Residues 214-220 form one of the three walls that line-up the primary specificity pocket of serine proteases(30) . In addition, residues 214-216 are part of the S(1)-S(3) subsite (^3)that forms a beta-pleated sheet with residues P(1)-P(3) of the substrate(31, 32, 33) . Two of these residues, Thr and Ser, have been mutated, which probably accounts at least in part for the observed r.m.s. deviations. In addition, changes in the orientation of the side chains of Arg and Arg apparently contribute to the observed r.m.s. shifts. In native factor D Arg forms a salt bridge with Asp at the bottom of the primary specificity pocket (Fig. 6). This bridge probably contributes to the low reactivity of resting-state factor D with synthetic substrates, as it restricts access of the positively charged P(1) Arg residue to the negative charge of Asp. In the mutant, Arg seems to be flexible and its side chain points away from the carboxylate of Asp and toward the solvent. Therefore, no salt bridge between this residue and Asp can form. Instead, Arg has been reoriented so that its guanidinium group is H-bonded to Asp (Fig. 6). These changes are accompanied by a substantial narrowing of the primary specificity pocket of the mutant, apparently due to a movement of the segment formed by residues 214-218 (Fig. 6).


Figure 5: Coordinate shifts between the S94Y/T214S/S215W mutant and native factor D. The r.m.s. deviation (Å) of main chain (mc) and side chain (sc) atoms between the mutant and molecule A (a) and molecule B (b) of native factor D crystallized in the triclinic unit cell(4) . The two molecules are related by a non-crystallographic 2-fold in space group P1.




Figure 6: Comparison of the primary specificity pockets of the S94Y/T214S/S215W mutant (black) and molecule B of native factor D (red). The catalytic residue Ser is shown at the toprightcorner and Asp, which in trypsin-like serine proteases forms a salt bridge with the side chain of the P1 Arg, is shown on the right side of the pocket. The guanidino group of Arg is salt-bridged to Asp in the native structure, but it is oriented away from Asp in the mutant. This shift is compensated by a reorientation of Arg, the side chain of which is H-bonded to Asp in the mutant. The largest structural differences between the two structures are due to narrowing of the pocket of the mutant.



Superposition of the mutant on molecule A results in substantial r.m.s. shifts around the SerTyr substitution (Fig. 5a). Asp of molecule A is pointed toward the solvent away from His at the active center of factor D. In the mutant, Asp is forced by the phenyl ring of Tyr to assume a position similar to that observed in typical serine proteases (Fig. 6) and the entire loop joining residues Val and Leu has been rearranged in a way that resembles the corresponding region of trypsin. The movement of loop 89-103 is mainly responsible for the observed average shift of 5.0 Å for main chain atoms and 8.2 Å for side chain atoms of this segment (Fig. 5a). In molecule B of factor D, the orientation of Asp and of the entire loop 89-103 is similar to that of trypsin, which accounts for the lack of large r.m.s. shifts between molecule B and the mutant (Fig. 5b). However, the imidazolium of His, which has an atypical trans orientation in molecule B of native factor B, assumes the catalytically active gauche orientation in the mutant. This movement is probably responsible for the r.m.s. shifts observed in this region of the plot shown in Fig. 5b.

Fig. 7summarizes graphically the main structural differences between native and mutant factor D. The catalytic triad residues Asp, His, and Ser of both molecules A and B present in the triclinic unit cell of native factor D have an atypical alignment inconsistent with catalysis. In molecule A, Asp is turned away from His, having gained access to the solvent, while in molecule B His is pointed away from Ser and Ser is occupying the space usually occupied by His. In contrast, in S94Y/T214S/S215W the catalytic triad residues display an orientation very similar to that of trypsin. This typical serine protease conformation of the catalytic triad probably accounts for the increased catalytic activity of the mutant compared to native factor D.


Figure 7: Active site regions of molecule A (MOLA), and molecule B (MOLB) of native factor D, bovine trypsin, and factor D S94Y/T214S/S215W mutant (STS). The models were generated using the Ribbons program(38) . Green is used for carbon, blue for nitrogen, and red for oxygen. The backbone is represented as a tube, and bonds joining the side-chain atoms Cbeta onward are shown as cylinders. Each model shows the orientation of the catalytic triad and of residues in the immediate vicinity. Bovine trypsin coordinates were obtained from the Brookhaven Protein Data Bank.




DISCUSSION

The present data clearly demonstrate that three residues, Ser, Thr, and Ser, are responsible for the atypical orientation of the catalytic triad of native factor D. Replacement of these three residues with those present in the corresponding positions of trypsin, Tyr, Ser, and Trp, resulted in a mutant enzyme with about 20-fold higher reactivity toward the synthetic thioester Z-Lys-SBzl than native factor D. The increased catalytic efficiency could be accounted for by an increase in k and could be attributed to structural changes of the catalytic center. Most importantly, in contrast to native factor D, the side chains of the catalytic residues Asp and His of the S94Y/T214S/S215W mutant have a typical serine protease orientation very similar to that seen in trypsin (Fig. 7).

Possible lattice effects were considered before concluding that the observed trypsin-like conformational changes could be attributed directly to the mutations. The triclinic unit cell of native factor D contains two non-crystallographically related molecules, whereas the triple mutant crystallizes in a different space group with one molecule in the asymmetric unit. Therefore, different lattice forces than those exercised on either molecule A or B of native factor D should be expected to impinge on the mutant, possibly contributing to the observed realignment of key catalytic residues. However, we have recently crystallized native factor D in space group P2(1) with one molecule in the asymmetric unit. The structure of this crystal form of native factor D has been determined by a combination of isomorphous replacement and molecular replacement methods. The model has been refined to an R-factor of 18.8% by using 8.0 to 2.0 Å resolution data. The catalytic triad of this molecule has a conformation similar to that present in molecule B of the triclinic cell. Similar conformations also have been observed in inhibitor complexes of factor D. It thus seems unlikely that lattice forces are major contributors to the observed structural differences between native and triple mutant factor D.

The spatial relationships of the three catalytic residues are constant in all serine proteases of known structure and are stabilized by a network of H bonds(7, 34) . Specifically, H bonds between the N1 of His and the O2 of Asp, the N2 of His and the O of Ser, and the O2 of Asp and O of Ser that are present in typical serine proteases are also present in the mutant factor D. In contrast the orientation of the side chains of these residues in native factor D precludes the formation of these H bonds. The integrity of the H bonds and the hydrophobic nature of the environment surrounding the buried Asp are essential for the formation of a functional unit responsible for bond formation and cleavage during catalysis(7, 8, 9, 10) . Hence, the proposal for a substrate-induced conformational change of the active center of factor D to explain the efficient activation of the alternative complement pathway(2, 3, 5) .

All three mutated residues, Ser, Thr, and Ser, apparently contribute to the unusual conformation of the catalytic triad of native factor D and the resulting low reactivity toward small synthetic thioester substrates. However, Ser seems to be the principal determinant as indicated by the larger effect of the S94Y mutation on the k for Z-Lys-SBzl than those obtained for the single T214S and S215W mutants (Table 2). The introduction of the bulky Tyr residue at position 94 is probably responsible for forcing Asp and His to assume an active conformation. This is suggested by the r.m.s. deviation plot (Fig. 5a), which shows substantial structural differences in the immediate vicinity of the S94Y substitution when the mutant is superposed on molecule A of factor D.

A contribution of Thr is clearly indicated by the synergistic gain in k (Table 2) observed when the T214S mutation is combined with the S94Y and S215W mutations. This effect is probably due to the formation of a H bond between the side chains of Ser and Asp, which helps stabilize Asp. A role for the highly conserved aromatic at position 215 of serine proteases in catalysis has never been probed before. This residue is part of the S(1)-S(3) subsite in trypsin(33) , which forms a beta-pleated sheet with residues P(1)-P(3) of the substrate. In trypsin, Trp is involved in the formation of a hydrophobic cluster that also includes Tyr, a structural determinant of substrate specificity(29) . However, in the S94Y/T214S/S215W mutant loop 172-177 is flexible, precluding a hydrophobic interaction between His and Trp. Instead, Trp is oriented in a way favoring hydrophobic interactions with His (Fig. 7). Regardless of its precise structural contribution, the S215W mutation had a significant effect on catalytic efficiency against Z-Lys-SBzl (Table 2).

An additional interesting structural difference between S94Y/T214S/S215W and native factor D involves Arg. In both molecules of factor D, the guanidinium group of this residue forms a salt bridge with the carboxyl of Asp(4) . In trypsin and presumably also in all trypsin-like serine proteases, the carboxyl of Asp forms a salt bridge with the side chain of the P(1) Arg or Lys residue of the substrate(35) . This ionic interaction plays a major role in positioning the scissile bond of the substrate for hydrolysis. Thus, the Arg-Asp bridge of native D has been considered an important contributor to the low reactivity of resting-state factor D toward small synthetic substrates(4, 12) . Furthermore, it has been suggested that a reorientation of the side chain of Arg is an important component of the putative substrate-induced conformational change that leads to efficient proteolysis of C3b-bound factor B by factor D(4, 12) . Interestingly, mutation of Arg to Gly, which is present in the corresponding position of trypsin, combined with deletion of Val, which is absent from trypsin, resulted in a small increase of esterolytic activity and an almost complete loss of proteolytic activity(12) . Indeed, even extensive mutations involving the substrate binding pocket and associated structural elements, which yielded a mutant factor D exhibiting 100-fold higher esterolytic activity than native factor D were associated with extremely low proteolytic activity attributable to the R218G and V219Delta mutations (36) . These findings were interpreted to indicate a direct interaction between Arg and determinants on the C3bB complex, the natural substrate of factor D.

The present finding that the realignment of the catalytic triad residues induced by the S94Y/T214S/S215W mutations is associated with a reorientation of Arg away from Asp indicates a linkage among the conformations of these elements of the active center. It thus seems possible that the proposed interaction between Arg and negatively charged residue(s) on C3bB not only makes Asp available to the P(1) Arg of factor B, but also contributes to the realignment of the catalytic triad residues. Obviously, additional conformational changes particularly of the substrate binding pocket are necessary for expression of efficient proteolytic activity. These changes were not induced by the mutation of the three residues investigated in the present study, as indicated by the geometry of the primary specificity pocket (Fig. 6) and the high K(m) of the triple mutant (Table 2).


FOOTNOTES

*
This work was supported in part by National Institutes of Health Grants AI21067 and AI32949 and by Grant NAGW-813 from NASA. The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore by hereby marked ``advertisement'' in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

§
To whom correspondence should be addressed: Division of Clinical Immunology and Rheumatology, UAB Station, THT 437, Birmingham, AL 35294-0006. Tel.: 205-934-5067; Fax: 205-934-2126.

(^1)
For comparison purposes factor D residues are numbered similarly to chymotrypsin from 16 to 243. Nine residues present in chymotrypsin (37, 116, 117, 203, 204, 205, 206, 244, and 245) but missing from factor D are absent from the numbering scheme. Conversely, 9 residues of factor D are not present in chymotrypsin (3 residues after position 61 and one each after positions 78, 124, 129, 170, 177, and 223). These are numbered as the preceding residue with an appropriate suffix: A, B, or C.

(^2)
The abbreviations used are: wt, wild type; CHO, Chinese hamster ovary; EC3b, sheep erythrocytes carrying human C3b; Z, benzyloxycarbonyl; SBzl, thiobenzyl; r.m.s., root mean square; MES, 4-morpholineethanesulfonic acid.

(^3)
The nomenclature used for the individual amino acid residues (P(1), P(2), etc.) of a substrate and the corresponding subsites (S(1), S(2), etc.) of the enzyme is that of Schechter and Berger(39) .


REFERENCES

  1. Müller-Eberhard, H. J., and Götze, O. (1972) J. Exp. Med. 135,1003-1008 [Medline] [Order article via Infotrieve]
  2. Lesavre, P. H., and Müller-Eberhard, H. J. (1978) J. Exp. Med. 148,1498-1509 [Abstract]
  3. Niemann, M. A., Bhown, A. S., Bennett, J. C., and Volanakis, J. E. (1984) Biochemistry 23,2482-2486 [Medline] [Order article via Infotrieve]
  4. Narayana, S. V. L., Carson, M., El-Kabbani, O., Kilpatrick, J. M., Moore, D., Chen, X., Bugg, C. E., Volanakis, J. E., and DeLucas, L. J. (1994) J. Mol. Biol. 235,695-708 [CrossRef][Medline] [Order article via Infotrieve]
  5. Kam, C.-M., McRae, B. J., Harper, J. W., Niemann, M. A., Volanakis, J. E., and Powers, J. C. (1987) J. Biol. Chem. 262,3444-3451 [Abstract/Free Full Text]
  6. Kam, C.-M., Oglesby, T. J., Pangburn, M. K., Volanakis, J. E., and Powers, J. C. (1992) J. Immunol. 149,163-168 [Abstract/Free Full Text]
  7. Blow, D. M., Birktoft, J. J., and Hartley, B. S. (1969) Nature 221,337-340 [Medline] [Order article via Infotrieve]
  8. Shotton, D. M., and Watson, H. C. (1970) Nature 225,811-816 [Medline] [Order article via Infotrieve]
  9. Bode, W., and Schwager, P. (1975) J. Mol. Biol. 98,693-717 [Medline] [Order article via Infotrieve]
  10. Sprang, S., Standing, T., Fletterick, R. J., Stroud, R. M., Finer-Moore, J., Xuong, N.-H., Hamlin, R., Rutter, W. J., and Craik, C. S. (1987) Science 237,905-909 [Medline] [Order article via Infotrieve]
  11. White, R. T., Damm, D., Hancock, N., Rosen, B. S., Lowell, B. B., Usher, P., Flier, J. S., and Spiegelman, B. M. (1992) J. Biol. Chem. 267,9210-9213 [Abstract/Free Full Text]
  12. Kim, S., Narayana, S. V. L., and Volanakis, J. E. (1994) Biochemistry 33,14393-14399 [Medline] [Order article via Infotrieve]
  13. Kunkel, T. A. (1985) Proc. Natl. Acad. Sci. U. S. A. 82,488-492 [Abstract]
  14. Sanger, F., Nicklen, S., and Coulson, A. R. (1977) Proc. Natl. Acad. Sci. U. S. A. 74,5463-5467 [Abstract]
  15. Yamauchi, Y., Stevens, J. W., Macon, K. J., and Volanakis, J. E. (1994) J. Immunol. 152,3645-3653 [Abstract/Free Full Text]
  16. Ueda, A., Kearney, J. F., Roux, K. H., and Volanakis, J. E. (1987) J. Immunol. 138,1143-1149 [Abstract/Free Full Text]
  17. Gresham, H. D., Matthews, D. F., and Griffin, F. M., Jr. (1986) Anal. Biochem. 154,454-459 [Medline] [Order article via Infotrieve]
  18. Niemann, M. A., Volanakis, J. E., and Mole, J. E. (1980) Biochemistry 19,1576-1583 [Medline] [Order article via Infotrieve]
  19. Pangburn, M. K. (1989) J. Immunol. 142,202-207 [Abstract/Free Full Text]
  20. Volanakis, J. E., and Macon, K. J. (1987) Anal. Biochem. 163,242-246 [Medline] [Order article via Infotrieve]
  21. Habeeb, A. F. S. A. (1976) Methods Enzymol. 25,457-464
  22. Matthews, B. S. (1968) J. Mol. Biol. 33,491-497 [Medline] [Order article via Infotrieve]
  23. Howard, A. J., Gilliland, G. L., Finzel, B. C., Poulos, T. L., Ohelendorf, D. H., and Salemme, F. R. (1987) J. Appl. Crystallogr. 20,383-387 [CrossRef]
  24. Crowther, R. A. (1972) in The Molecular Replacement Method (Rossman, M. G., ed) pp. 173-185, Gordon & Breach, New York
  25. Brunger, A. T. (1988) J. Mol. Biol. 203,803-816 [Medline] [Order article via Infotrieve]
  26. Hendrickson, W. A. & Konnert, J. (1981) in Biological Structure, Function, Conformation and Evaluation (Srinivasan, R., ed) Vol. 1, pp. 43-47, Pergamon Press, London
  27. Jones, T. A. (1978) J. Appl. Crystallogr. 11,268-272 [CrossRef]
  28. Ramachandran, G. N., Ramakrishnan, C., and Sasisekharan, V. J. (1963) Mol. Biol. 7,95-99
  29. Hedstrom, L., Perona, J. J., and Rutter, W. J. (1994) Biochemistry 33,8757-8763 [Medline] [Order article via Infotrieve]
  30. Cohen, G. H., Silverton, E. W., and Davies, D. R. (1981) J. Mol. Biol. 148,449-479 [Medline] [Order article via Infotrieve]
  31. Segal, D. M., Powers, J. C., Cohen, G. H., Davies, D. R., and Wilcox, P. E. (1971) Biochemistry 10,3728-3738 [Medline] [Order article via Infotrieve]
  32. Robertus, J. D., Alden, R. A., Birktoft, J. J., Kraut, J., Powers, J. C., and Wilcox, P. E. (1972) Biochemistry 11,2439-2449 [Medline] [Order article via Infotrieve]
  33. Rühlmann, A., Kukla, D., Schwager, P., Bartels, K., and Huber, R. (1973) J. Mol. Biol. 77,417-436 [Medline] [Order article via Infotrieve]
  34. Stroud, R. M., Kay, L. M., and Dickerson, R. E. (1972) Cold Spring Harbor Symp. Quant. Biol. 36,125-140 [Medline] [Order article via Infotrieve]
  35. Bode, W., Schwager, R., and Huber, R. (1978) J. Mol. Biol. 118,99-112 [Medline] [Order article via Infotrieve]
  36. Kim, S., Narayana, S. V. L., and Volanakis, J. E. (1995) J. Immunol. 154,6073-6079 [Abstract/Free Full Text]
  37. Greer, J. (1990) Proteins 7,317-334 [Medline] [Order article via Infotrieve]
  38. Carson, M., and Bugg, C. E. (1986) J. Mol. Graph. 4,121-122 [CrossRef]
  39. Schechter, I., and Berger, A. (1967) Biochem. Biophys. Res. Commun. 27,157-162 [Medline] [Order article via Infotrieve]

©1995 by The American Society for Biochemistry and Molecular Biology, Inc.