The lactose permease (lac) of Escherichia coli is a paradigm for membrane transport proteins. Encoded by the lacY gene, the permease has been solubilized, purified to homogeneity, reconstituted into phospholipid vesicles and shown to catalyse the coupled translocation of β-galactosides and H+ with a stoichiometry of unity. Circular dichroism and other spectroscopic approaches demonstrate that the purified permease is about 80% helical. Based on hydropathy analysis of the primary amino-acid sequence, a secondary structure has been proposed in which the protein has 12 hydrophobic domains in α-helical conformation that traverse the membrane in zigzag fashion connected by hydrophilic loops. A variety of other approaches are consistent with the model and demonstrate that both the N and C termini are on the inner surface of the membrane, and studies on an extensive series of lac permease/alkaline phosphatase fusion proteins provide exclusive support for the topological predictions of the 12-helix motif. This presentation concentrates on the use of site-directed fluorescence spectroscopy to study structure–function relationships in the permease.

An unsolved basic biochemical problem of critical importance is the mechanism of energy transduction in biological membranes. Although the driving force for a variety of seemingly unrelated phenomena (e.g. secondary active transport, oxidative phosphorylation, rotation of the bacterial flagellar motor) is a bulk-phase, transmembrane electrochemical ion gradient, the molecular mechanism(s) by which free energy stored in such gradients is transduced into work or into chemical energy remains enigmatic. Nonetheless, gene sequencing and analyses of deduced amino-acid sequences indicate that many biological machines involved in energy transduction, secondary transport proteins in particular (Henderson, 1990; Marger and Saier, 1993), fall into families encompassing proteins from archaebacteria to the mammalian central nervous system, thereby suggesting that the members may have common basic structural features and mechanisms of action. In addition, certain of these proteins have been implicated in human disease (e.g. glucose/galactose malabsorption, certain forms of drug abuse).

As postulated originally by Peter Mitchell (1963, 1968) and demonstrated conclusively in bacterial membrane vesicles (Kaback, 1986, 1989, 1992), accumulation of a wide variety of solutes against a concentration gradient is driven by a proton electrochemical gradient interior negative and/or alkaline). The work discussed here focuses on one specific secondary transport protein, the lactose or lac permease as a paradigm. -galactoside accumulation in Escherichia coli is catalyzed by lac permease, a hydrophobic polytopic cytoplasmic membrane protein that carries out the coupled stoichiometric translocation of four β-galactosides with H+ (i.e. -galactoside/H+ symport or cotransport). Physiologically, the permease utilizes free energy released from downhill translocation of H+ to drive accumulation of -galactosides against a concentration gradient. In the absence of , lac permease catalyzes the converse reaction, utilizing free energy released from downhill translocation of -galactosides to drive uphill translocation of H+ with the generation of a , the polarity of which depends upon the direction of the substrate concentration gradient.

Lac permease is encoded by the lacY gene, the second structural gene in the lac operon, which has been cloned into a recombinant plasmid (Teather et al. 1978) and sequenced (Buchel et al. 1980). By combining overexpression of lacY with the use of a highly specific photoaffinity probe (Kaczorowski et al. 1980) and reconstitution of transport activity in artificial phospholipid vesicles (i.e. proteoliposomes, Newman and Wilson, 1980), the permease was the first symporter to be solubilized from the membrane, purified to homogeneity (Newman et al. 1981; Foster et al. 1982; Viitanen et al. 1986) and shown to catalyze all the translocation reactions typical of the transport system in vivo with similar turnover numbers (Viitanen et al. 1984; Matsushita et al. 1983). The findings demonstrate that the product of the lacY gene is solely responsible for all of the translocation reactions catalyzed by the β-galactoside transport system.

Lac permease contains 12 transmembrane domains in α.-helical conformation

Circular dichroic measurements on purified lac permease demonstrate that the protein is 75–80% helical, an estimate consistent with the hydropathy profile of the permease which suggests that approximately 70% of its 417 amino acid residues are found in hydrophobic domains with a mean length of 24±4 residues. Based on these findings, it was proposed (Foster et al. 1983) that the permease is composed of a hydrophilic N terminus followed by 12 hydrophobic segments in α.-helical conformation that traverse the membrane in zigzag fashion connected by hydrophilic domains (loops) with a 17-residue C-terminal hydrophilic tail (Fig. 1). Support for general features of the model and evidence that the N and C termini, as well as the second and third cytoplasmic loops, are exposed to the cytoplasmic face of the membrane were then obtained from laser Raman (Vogel et al. 1985) and Fourier transform infrared (P. D. Roepe, K. Rothschild and H. R. Kaback, unpublished information) spectroscopy, immunological studies (Carrasco et al. 1982, 1984a,b; Seckler et al. 1983, 1986; Seckler and Wright, 1984; Herzlinger et al. 1984, 1985; Danho et al. 1985), limited proteolysis (Goldkorn et al. 1983; Stochaj et al. 1986) and chemical modification (Page and Rosenbusch, 1988). However, none of these approaches is able to differentiate between the 12-helix motif and other models containing 10 (Vogel et al. 1985) or 13 (Bieseler et al. 1985) transmembrane domains.

Fig. 1.

Secondary structure model of lac permease based on the hydropathy profile of the protein. The single amino-acid code is used, and hydrophobic segments are shown in boxes as transmembrane (α.-helical domains connected by hydrophilic loops). Residues labeled black yield active permease when replaced with Cys; residues labeled red yield inactive permease when replaced with a number of different residues; residues labeled yellow are charge-paired (Asp237–Lys358 and Asp240–Lys3l9); residues labeled light blue have not been mutagenized. Permease with A177C (green) is inactive, but other replacements have not been made; permease with Y336F (green) is inactive, but other replacements have not been made. Black arrowheads indicate active split permease constructs; red arrowheads signify inactive constructs.

Fig. 1.

Secondary structure model of lac permease based on the hydropathy profile of the protein. The single amino-acid code is used, and hydrophobic segments are shown in boxes as transmembrane (α.-helical domains connected by hydrophilic loops). Residues labeled black yield active permease when replaced with Cys; residues labeled red yield inactive permease when replaced with a number of different residues; residues labeled yellow are charge-paired (Asp237–Lys358 and Asp240–Lys3l9); residues labeled light blue have not been mutagenized. Permease with A177C (green) is inactive, but other replacements have not been made; permease with Y336F (green) is inactive, but other replacements have not been made. Black arrowheads indicate active split permease constructs; red arrowheads signify inactive constructs.

Calamia and Manoil (1990) provided elegant, unequivocal support for the topological predictions of the 12-helix model by analyzing an extensive series of lac permease–alkaline phosphatase (lacY–phoA) fusion proteins. In addition, it was shown that the alkaline phosphatase activity of fusion proteins engineered at every third amino-acid residue in putative helices III and V increases abruptly as the fusion junction proceeds from the eighth to the eleventh residue. Thus, approximately half a transmembrane domain is needed to translocate alkaline phosphatase through the membrane to the external surface. When fusions are constructed at each amino-acid residue in putative helices IX and X, the data are in excellent agreement with the model (M. L. Ujwal, E. Bibi, C. Manoil and H. R. Kaback, in preparation).

Lac permease is functional as a monomer

One difficult problem to resolve with hydrophobic membrane proteins is their functional oligomeric state. Notwithstanding strong evidence that lac permease is functional as a monomer (Dornmair et al. 1985; Costello et al. 1987), Bibi and Kaback (1992) demonstrated that certain paired in-frame deletion mutants are able to complement functionally. Although cells expressing the deletions individually do not catalyze active transport, cells simultaneously expressing specific pairs of deletions catalyze transport up to 60% as well as cells expressing wild-type permease, and it is clear that the phenomenon occurs at the level of the protein. Remarkably, complementation is observed only with pairs of permease molecules containing large deletions and not with missense mutations or point deletions. Although the mechanism of complementation is unclear, it is probably related to the phenomenon whereby independently expressed N-and C-terminal fragments of the permease interact to form a functional complex (Bibi and Kaback, 1990; see below). In any case, the observation that certain pairs of deletion mutants can interact rekindled concern regarding the oligomerization state of wild-type permease.

Sahin-Tóth et al. (1994) recently engineered a fusion protein containing two lac permease molecules covalently linked in tandem (permease dimer). Permease dimer is inserted into the membrane in a functional state, and each half of the dimer exhibits equal activity. Thus, point mutations in either half of the lacY tandem repeat lead to 50% inactivation of transport. Furthermore, the activity of a permease dimer composed of wild-type permease and a mutant devoid of Cys is inactivated by approximately by 60% by N-ethylmaleimide (NEM). In order to test the caveat that oligomerization between dimers might occur in the trans position, a permease dimer was constructed that contains two different deletion mutants which complement when expressed as untethered molecules. This construct does not catalyze lactose accumulation to any extent whatsoever, suggesting that permease dimers do not undergo oligomerization in the trans position. The experiments are clearly consistent with the conclusion that wild-type lac permease is functional as a monomer. It is also noteworthy that this approach can be applied to other membrane transport proteins that have not been solubilized and purified.

Site-directed mutagenesis reveals that few amino-acid residues are essential for activity

By using site-directed mutagenesis with wild-type permease or a functional mutant devoid of Cys residues (C-less permease, van Iwaarden et al. 1991b), individual amino-acid residues in the permease that are important mechanistically have been identified (Fig. 1). About 300 of the 417 residues in C-less permease have been replaced with Cys and, remarkably, over 95% of the mutants retain activity (Sahin-Tóth et al. 1994; Dunten et al. 1993a) [in addition to helices I, IX, X and XI, Cys-scanning mutagenesis has been carried out with helices III (M. Sahin-Tóth, S. Frillingos, E. Bibi, A. Gonzalez and H. R. Kaback, in preparation), V (C. Weizmann, M. Sahin-Tóth and H. R. Kaback, unpublished information) and VII (S. Frillingos, M. Sahin-Tóth and H. R. Kaback, in preparation) (see Fig. 1)]. Of the few mutants that do not catalyze active transport, most retain the ability to catalyze partial reactions or bind ligand. More specifically, none of the eight Cys (van Iwaarden et al. 1991a,b; Trumble et al. 1984; Menick et al. 1985, 1987b; Viitanen et al. 1985; Neuhaus et al. 1985; Sarkar et al. 1986; Brooker and Wilson, 1986), six Trp (Menezes et al. 1990) or twelve Pro (Lolkema et al. 1988; Consler et al. 1991) residues in the permease is obligatory for activity. Only one out of four His residues (Padan et al. 1985; Püttner et al. 1986, 1989; Püttner and Kaback, 1988), one of the 14 Tyr residues (Roepe and Kaback, 1989) (although Tyr→Phe replacements indicate that three Tyr residues are important for activity, cys-scanning mutagenesis has revealed that Tyr-26 and Tyr-236 can be replaced with Cys with retention of significant activity) and one out of 36 Gly residues (K. Jung, H. Jung, P. Colachurchio and H. R. Kaback, in preparation) are important for active transport. However, Glu-269 (helix VIII) (Ujwal et al. 1994), Arg-302 (helix IX) (Menick et al. 1987a), His-322 (helix X) (Padan et al. 1985; Püttner et al. 1986, 1989; Püttner and Kaback, 1988) and Glu-325 (helix X) (Carrasco et al. 1986, 1989) are essential for substrate accumulation and/or binding. Moreover, differences in the properties of the mutants suggest that Arg-302, His-322 and Glu-325 may function in an H+ translocation pathway, although it is possible that the residues also form part of a coordination site for H3O+ (see Kaback, 1992, for a discussion). In any event, very few residues are mandatory for binding or transport, and it is unlikely that individual amino-acid replacements cause global conformational changes.

Site-directed fluorescent labeling and mutagenesis yield helix packing in the C-terminal half of the permease

The observations described above highlight a need for static and dynamic information at high resolution in order to define the transport mechanism, and recent studies with permease molecules engineered specifically for spectroscopic approaches (van Iwaarden et al. 1991b; Menezes et al. 1990) are providing exciting new insights. Since none of the individual Cys residues in the permease is required for activity, all eight Cys residues have been replaced simultaneously with Ser or Val to yield a Cys-less permease that retains at least 50% of wild-type activity (van Iwaarden et al. 1991b). Molecules containing one or two Cys replacements at any desired position(s) can then be tagged with thiol-specific biophysical probes after solubilization and purification. By using this approach, Jung et al. (1993) established proximity relationships between transmembrane helices in the C-terminal half of the permease. Pairs of charged amino-acid residues in transmembrane domains were replaced with Cys in a C-less construct containing a biotin-acceptor domain in the middle cytoplasmic loop to facilitate purification by monovalent avidin affinity chromatography (Consler et al. 1993). As a fluorophore, N-(1-pyrenyl)maleimide (PM) was selected, since it can form an excited state dimer (excimer) that exhibits a unique emission maximum if two conjugated ring systems are within about 0.35 nm and correctly oriented (Kinnunen et al. 1993). The findings obtained indicate: (i) that transmembrane domain X is in α.-helical conformation; (ii) helices IX (Arg-302) and X (Glu-325) are in close proximity; and (iii) helix VIII (Glu-269) is in close proximity to helix X (His-322) (Fig. 2).

Fig. 2.

Helical wheel model of putative helices VII–XI in lac permease viewed from the periplasmic surface. In addition to the symbols given on the figure, G268 and G262 (helix VIII) which are sites of second-site suppressors for D240 and E325, respectively, are emboldened, and V315 is highlighted (see Jung et al. 1993).

Fig. 2.

Helical wheel model of putative helices VII–XI in lac permease viewed from the periplasmic surface. In addition to the symbols given on the figure, G268 and G262 (helix VIII) which are sites of second-site suppressors for D240 and E325, respectively, are emboldened, and V315 is highlighted (see Jung et al. 1993).

Second-site suppressor analysis (King et al. 1991) and site-directed mutagenesis combined with chemical modification (Sahin-Tóth et al. 1992; Dunten et al. 1993b; Sahin-Tóth and Kaback, 1993b) demonstrate that Asp-237 (helix VII) interacts with Lys-358 (helix XI), probably forming a salt bridge. Individual replacement of Asp-237 or Lys-358 in C-less permease with Cys or Ala abolishes active lactose transport, whereas simultaneous replacement of both charged residues with Cys and/or Ala or reversal of the residues leads to fully active permease. Remarkably, mutant D237C is restored to full activity by carboxylmethylation, which recreates a negative charge at position 237, and mutant K358C is restored to full activity by treatment with ethylammonium methanethiosulfonate, which recreates a positive charge at position 358. It has also been shown by site-directed mutagenesis (Sahin-Tóth et al. 1992; Sahin-Tóth and Kaback, 1993b) and second-site suppressor analysis (Lee et al. 1992) that Asp-240 (helix VIII) interacts with Lys-3l9 (helix X). Individual replacement of either Asp-240 or Lys-3l9 in C-less permease with neutral amino acid residues inactivates the permease, but double neutral mutants retain significant activity. In contrast to Asp-237/Lys-358, however, the polarity of the interaction between Asp-240 and Lys-3l9 is important, as reversal of the residues inactivates the permease. In any case, the findings indicate that helix VII (Asp-237 and Asp-240) is in close proximity to helices X (Lys-3l9) and XI (Lys-358). It is also apparent that none of these residues plays a direct role in the mechanism. Taken together with the conclusions derived from site-directed pyrene labeling, the observations describe the packing of helices VII–XI (Fig. 2) provide the first tertiary structure information regarding this class of membrane proteins (Jung et al. 1993; Kaback et al. 1993).

Ligand binding induces widespread conformational changes

Jung et al. (1994b) have demonstrated that excimer fluorescence can also be used to study dynamic aspects of permease folding. Excimer fluorescence formed between transmembrane domains is markedly diminished by denaturants, whereas the excimer fluorescence observed within helix X is unaffected, indicating that tertiary interactions are disrupted with little effect on secondary structure. Consistently, interacting helices do not exhibit excimer fluorescence in detergent, but only after reconstitution into membranes. One of the excimers described also exhibits ligand-induced alterations. Excimer fluorescence due to the interaction between helices VII (E269C) and X (H322C) is quenched by Tl+, and the effect is markedly and specifically attenuated by permease substrates. The reactivity of single Cys residues placed in many transmembrane domains is also dramatically altered in the presence of ligand, implying that transport involves widespread changes in tertiary structure. In one Cys replacement mutant (V315C) tested thus far, the effect of ligand is mimicked by imposition of , providing an exciting preliminary suggestion that either ligand binding or may cause the permease to assume the same conformation (Sahin-Tóth and Kaback, 1993a,b). The initial observations made with right-side-out membrane vesicles have been confirmed recently with purified V315C permease (Jung et al. 1994a).

Cys-148 is in a substrate-binding site

Fox and Kennedy (1965) demonstrated that lac permease is irreversibly inactivated by N-ethylmaleimide (NEM) and that protection is afforded by certain substrates. On the basis of these findings, it was postulated that a Cys residue is at or near the substrate-binding site in the permease, and it was shown later by Bayreuther et al. (1981) that the substrate-protectable Cys is at position 148. Since Cys-l48 can be replaced with Gly (Trumble et al. 1984; Viitanen et al. 1985) or Ser (Neuhaus et al. 1985; Sarkar et al. 1986) with little or no effect on activity, it is apparent that this residue is not essential. However, the observations also raise the possibility that substrate protection may be due to a long-range conformational effect. Recently (H. Jung, K. Jung and H. R. Kaback, in preparation), Cys-148 was replaced with hydrophobic (Ala, Val, Ile, Phe), hydrophilic (Ser, Thr) or charged (Asp, Lys) residues, and the properties of the replacement mutants were analyzed. Although Cys-l48 is not essential for transport, the size and polarity of the side chain at this position modify transport activity and substrate specificity. Thus, small hydrophobic side-chains (Ala, Val) generally increase the apparent affinity of the permease for substrate, whereas hydrophilic side-chains (Ser, Thr, Asp) decrease the apparent affinity and bulky or positively charged side-chains (Phe, Lys) virtually abolish activity. In addition, hydrophilic substitutions (Ser, Thr, Asp) alter the activity of the permease towards monosaccharides relative to disaccharides.

Site-directed fluorescence spectroscopy was also used to study Cys-148 and other residues in the vicinity (J. Wu and H. R. Kaback, in preparation). In the absence of ligand, permease with a single Cys residue at position 148 reacts rapidly with 2-(4’-maleimidylanilino)-naphthalene-6-sulfonic acid (MIANS), a fluorophore whose quantum yield increases dramatically upon reaction with a thiol, indicating that this residue is readily accessible to the probe. Various ligands of the permease block the reaction, and the concentration-dependence is commensurate with the affinity of each ligand for the permease (i.e. β-D-galactopyranosyl 1-thio-β-D-galactopyranoside ≪lactose<galactose), but neither sucrose nor glucose has any effect whatsoever. Interestingly, labeling of Cys-145, which is presumed to be one helical turn removed from Cys-l48, displays properties similar to those observed with Cys-l48 permease, but the effects of ligand are far less dramatic. In contrast, permease with a single Cys residue at position 146 or 147 behaves in a completely different manner. Studies with iodide show that MIANS at positions 145 or 148 is accessible to the collisional quencher, indicating that this face of helix V is solvent-exposed, whereas MIANS at positions 146 or 147 is not quenched by iodide in the presence or absence of ligand. Finally, iodide-quenching of MIANS at position 145 is clearly diminished in the presence of ligand. Taken together with the findings discussed above, the results indicate that Cys-l48 is a likely component of a substrate-binding site that interacts hydrophobically with the galactosyl portion of the substrate, but does not play an essential role in transport. In addition, the observations indicate that residue 145 is in close proximity to Cys-148.

A novel approach to crystallization of hydrophobic membrane proteins

A major thrust of this laboratory is aimed at two-dimensional and three-dimensional crystallization. Since hydrophobic membrane proteins such as lac permease are particularly difficult to crystallize in three-dimensional form, Privé et al. (1994) have devised a novel approach in which a fusion is constructed between the permease and a ‘carrier’ protein. The carrier is a soluble, stable protein with its C and N termini close together in space at the surface of the protein, so that it can be introduced into an internal position of the permease without distorting either molecule. McKenna et al. (1992b) demonstrated that all but three of the hydrophilic domains in the permease can be disrupted by the insertion of two or six contiguous His residues without abolishing activity. The carrier is chosen with convenient spectral properties, making the fusion protein easier to characterize than the native molecule. A chimeric protein with E. coli cytochrome b562 fused into the middle cytoplasmic loop of lac permease and six His residues attached to the C terminus has been constructed, expressed and highly purified by nickel chelate chromatography. The chimera exhibits transport activity similar to that of wild-type lac permease and has a visible absorption spectrum and a redox potential that are identical to those of cytochrome b562. The chimera has a higher proportional polar surface area than wild-type permease and should have better possibilities of forming the strong, directional intermolecular contacts required of a crystal lattice.

Insertion of permease into the membrane

Expression of lactose permease in fragments as a probe for membrane-spanning domains

Lac permease also provides a useful system for studying insertion of polytopic proteins into the membrane. Surprisingly, as shown by Bibi and Kaback (1990), when the lacY gene is split into two approximately equal-sized fragments (N6 and C6) (Fig. 1), cells expressing both fragments catalyze significant lactose accumulation, whereas cells expressing either half of the permease individually are devoid of activity. Intact permease is completely absent from the membrane of cells expressing fragments either individually or together. Thus, transport activity must result from association between independently synthesized portions of lac permease. When the gene fragments are expressed individually, the N-terminal portion of the permease is observed sporadically and the C-terminal portion is not observed; when expressed together, the N-and C-terminal moieties of the permease are found in the membrane. Clearly, the N-or C-terminal halves are proteolyzed when synthesized independently, and association between the two complementing polypeptides leads to a more stable, catalytically active complex.

Co-expression of independently cloned fragments of lacY encoding N2 and C10 (Wrubel et al. 1990), N1 and C11 or N7 and C5 (Zen et al. 1994) also yields stable molecules in the membrane which interact to form functional permease (Fig. 1). Thus, lacY gene fragments encoding contiguous, non-overlapping peptides with discontinuities in either cytoplasmic or periplasmic loops are able to complement functionally. In striking contrast, Zen et al. (1994) have demonstrated that peptide fragments with discontinuities in transmembrane domains are unable to form functional complexes, implying that the ‘split permease approach’ may be useful for approximating helical boundaries. Based on this notion, a series of contiguous, non-overlapping permease fragments with discontinuities at various positions in loop 6 (cytoplasmic), putative helix VII and loop 7 (periplasmic) were co-expressed in order to approximate the boundaries of transmembrane domain VII (Zen et al. 1994). Contiguous fragments with a discontinuity between Leu-222 and Trp-223 (loop 6) or Gly-254 and Glu-255 (loop 7) are functional, but fragments with discontinuities between Cys-234 and Thr-235, Gln-241 and Gln-242 or Phe-247 and Met-248 are inactive (Fig. 1). Therefore, it is likely that Leu-222 and Gly-254 are located in hydrophilic loops 6 and 7, respectively, whereas Cys-234, Gln-241 and Phe-247 are probably within transmembrane domain VII. These and other results are consistent with a secondary structure model of lactose permease in which Asp-237 and Asp-240 are contained within domain VII rather than loop 7, as predicted by hydropathy profiling (King et al. 1991; Sahin-Tóth et al. 1992; Dunten et al. 1993b; Sahin-Tóth and Kaback, 1993b; Zen et al. 1994).

Regarding the contention that contiguous permease fragments with discontinuities in transmembrane domains do not exhibit functional complementation, Roepe et al. (1989) and McKenna et al. (1991, 1992a) have presented evidence that the last turn of putative helix XII must be intact for the permease to insert into the membrane in a stable form. Thus, truncation of the 17-residue C-terminal tail of the permease at position 401 has little or no effect on the activity or stability of the protein, whereas truncation of the permease at position 396 leads to complete loss of activity and rapid degradation (Fig. 1). Truncations at positions 397–400 exhibit progressively increasing activity and are progressively more stable in the membrane. Moreover, replacement of residues 397–400 with Leu-Leu-Leu-Leu yields fully functional permease that is completely stable, while replacement with Gly-Pro-Gly-Pro yields unstable permease with minimal activity. It seems likely, therefore, that incomplete transmembrane domains are not inserted into the membrane in a stable form. However, it should be emphasized that proteolytic degradation cannot explain the lack of transport activity observed with permease duplexes containing discontinuities in transmembrane domains. Thus, an immunoreactive C-terminal fragment is observed with N2.5/C9.5. Moreover, permease deleted of the first 22 amino-acid residues (the N terminus and half of the first transmembrane domain) is stable and functional when expressed at a high rate (Bibi et al. 1992). Therefore, it seems reasonable to suggest that the lack of transport activity observed with duplexes split in transmembrane domains may be due to an alteration in the transfer of conformational information from one side of a transmembrane domain to the other.

Membrane insertion may involve multiple mechanisms

The demonstration that polypeptides corresponding to N1 and C11 form a relatively stable, functional complex argues against the notion that the N terminus of the permease inserts into the membrane as a helical hairpin. However, H. K. Zen, T. G. Consler, D. Hardy and H. R. Kaback (in preparation) (see Consler et al. 1993) have shown that insertion of the biotin-acceptor domain into the second or the fourth periplasmic loops of the permease (between helices III and IV or helices VII and VIII, respectively; Fig. 1) blocks insertion of transmembrane helices III and IV or VII and VIII without altering the insertion of the remainder of the protein, suggesting that these portions of the permease may be inserted as a helical hairpin. It is also important that Dunten et al. (1993b) showed that disruption of the salt bridge between Asp-237 and Lys-358 causes the permease to be inserted into the membrane much less efficiently, which raises the possibility that the C-terminal half of the polypeptide may be inserted post-translationally. Finally, the first 22 amino-acid residues in the permease, which represent the N-terminal hydrophilic domain and the first half of putative helix I, are not important for activity, but enhance the efficiency of insertion into the membrane (Bibi et al. 1992).

S.F. is a Fellow of the Human Frontier Science Program Organization (HSFPO), and this agency is acknowledged for providing financial support.

Bayreuther
,
K.
,
Bieseler
,
B.
,
Ehring
,
R.
and
Malier–hill
,
B.
(
1981
).
In Methods in Protein Sequence Analysis
(ed.
M.
Elzina
), pp.
139
148
.
New York
:
Humana
.
Bibi
,
E.
and
Kaback
,
H. R.
(
1990
).
In vivo expression of the lacY gene in two segments leads to functional lac permease
.
Proc. natn. Acad. Sci. U.S.A.
87
,
4325
4329
.
Bibi
,
E.
and
Kaback
,
H. R.
(
1992
).
Functional complementation of internal deletion mutants in the lactose permease of Escherichia coli
.
Proc. natn. Acad. Sci. U.S.A.
89
,
1524
1528
.
Bibi
,
E.
,
Stearns
,
S. M.
and
Kaback
,
H. R.
(
1992
).
The N-terminal twenty-two amino acid residues in the lactose permease of Escherichia coli are not obligatory for membrane insertion or transport activity
.
Proc. natn. Acad. Sci. U.S.A.
89
,
3180
3184
.
Bieseler
,
B.
,
Heinrich
,
P.
and
Bayreuther
,
C.
(
1985
).
Topological studies of lactose permease of Escherichia coli by protein sequence analysis
.
Ann. N.Y. Acad. Sci.
456
,
309
325
.
Brooker
,
R. J.
and
Wilson
,
T. H.
(
1986
).
Site-specific alteration of cysteine 176 and cysteine 234 in the lactose carrier of Escherichia coli
.
J. biol. Chem.
261
,
11765
11771
.
Buchel
,
D. E.
,
Gronenborn
,
B.
and
Müller-Hill
,
B.
(
1980
).
Sequence of the lactose permease gene
.
Nature
283
,
541
545
.
Calamia
,
J.
and
Manoil
,
C.
(
1990
).
Lac permease of Escherichia coli topology and sequence elements promoting membrane insertion
.
Proc. natn. Acad. Sci. U.S.A..
87
,
4937
4941
.
Carrasco
,
N.
,
Antes
,
L. M.
,
Poonian
,
M. S.
and
Kaback
,
H. R.
(
1986
).
Lac permease of Escherichia coli: His-322 and Glu-325 may be components of a charge-relay system
.
Biochemistry, N.Y.
25
,
4486
4488
.
Carrasco
,
N.
,
Herzlinger
,
D.
,
Mitchell
,
R.
,
Dechiara
,
S.
,
Danho
,
W.
,
Gabriel
,
T. F.
and
Kaback
,
H. R.
(
1984a
).
Intramolecular dislocation of the COOH terminus of the lac carrier protein in reconstituted proteoliposomes
.
Proc. natn. Acad. Sci. U.S.A.
81
,
4672
4676
.
Carrasco
,
N.
,
Püttner
,
I.. B.
,
Antes
,
L. M.
,
Lee
,
J. A.
,
Larigan
,
J. D.
,
Lolkema
,
J. S.
and
Kaback
,
H. R.
(
1989
).
Characterization of site-directed mutants in the lac permease of Escherichia coli. II. Glutamate-325 replacements
.
Biochemistry, N.Y.
28
,
2533
2539
.
Carrasco
,
N.
,
Tahara
,
S. M.
,
Patel
,
L.
,
Goldkom
,
T.
and
Kaback
,
H. R.
(
1982
).
Preparation, characterization and properties of monoclonal antibodies against the lac carrier protein from Escherichia coli
.
Proc. natn. Acad Sci. U.S.A.
79
,
6894
6898
.
Carrasco
,
N.
,
Viitanen
,
P.
,
Herzlinger
,
D.
and
Kaback
,
H. R.
(
1984b
).
Monoclonal antibodies against the lac carrier protein from Escherichia coli. I. Functional studies
.
Biochemistry, N.Y.
23
,
3681
3687
.
Consler
,
T. G.
,
Persson
,
B. L.
,
Jung
,
H.
,
Zen
,
K. H.
,
Jung
,
K.
,
Privé
,
G. G.
,
Verner
,
G. E.
and
Kaback
,
H. R.
(
1993
).
Properties and purification of an active biotinylated lactose permease from Escherichia coli
.
Proc. natn. Acad. Sci. U.S.A.
90
,
6934
6938
.
Consler
,
T. G.
,
Tsolas
,
O.
and
Kaback
,
H. R.
(
1991
).
Role of proline residues in the structure and function of a membrane transport protein
.
Biochemistry, N.Y.
30
,
1291
1298
.
Costello
,
M. J.
,
Escaig
,
J.
,
Matsushita
,
K.
,
Viitanen
,
P. V.
,
Menick
,
D. R.
and
Kaback
,
H. R.
(
1987
).
Purified lac permease and cytochrome o oxidase are functional as monomers
.
J. biol. Chem.
262
,
17072
17082
.
Danho
,
W.
,
Makofske
,
R.
,
Humiec
,
F.
,
Gabriel
,
T. F.
,
Carrasco
,
N.
and
Kaback
,
H. R.
(
1985
).
Use of site-directed polyclonal antibodies as immunotopological probes for the lac permease of Escherichia coli
. In
Peptides: Structure and Function
(ed.
C. M.
Deber
,
V. J.
Hruby
and
K. D.
Kopple
), pp.
59
62
. Pierce Chem. Co.
Dornmair
,
K.
,
Corin
,
A. S.
,
Wright
,
J. K.
and
Jahnig
,
F.
(
1985
).
The size of the lactose permease derived from rotational diffusion measurements
.
EMBO J.
4
,
3633
3638
.
Dunten
,
R. L.
,
Sahin-Tóth
,
M.
and
Kaback
,
H. R.
(
1993a
).
Cysteine scanning mutagenesis of putative helix XI in the lactose permease of Escherichia coli
.
Biochemistry, N.Y.
32
,
12644
12650
.
Dunten
,
R. L.
,
Sahin-Tóth
,
M.
and
Kaback
,
H. R.
(
1993b
).
Role of the charge pair formed by aspartic acid 237 and lysine 358 in the lactose permease of Escherichia coli
.
Biochemistry, N.Y.
32
,
3139
3145
.
Foster
,
D. L.
,
Boublik
,
M.
and
Kaback
,
H. R.
(
1983
).
Structure of the lac carrier protein of Escherichia coli
.
J. biol. Chem.
258
,
31
34
.
Foster
,
D. L.
,
Garcia
,
M. L.
,
Newman
,
M. J.
,
Patel
,
L.
and
Kaback
,
H. R.
(
1982
).
Lactose–proton symport by purified lac carrier protein
.
Biochemistry, N.Y.
21
,
5634
5638
.
Fox
,
C. F.
and
Kennedy
,
E. P.
(
1965
).
Specific labeling and partial purification of the M protein, a component of the -galactoside transport system of Escherichia coli
.
Proc. natn. Acad. Sci. U.S.A.
51
,
891
899
.
Goldkorn
,
T.
,
Rimon
,
G.
and
Kaback
,
H. R.
(
1983
).
Topology of the lac carrier protein in the membrane of Escherichia coli
.
Proc. natn. Acad. Sci. U.S.A.
80
,
3322
3326
.
Henderson
,
P. J. F.
(
1990
).
Proton-linked sugar transport systems in bacteria
.
J. Bioenerg. Biomembr.
22
,
525
569
.
Herzlinger
,
D.
,
Carrasco
,
N.
and
Kaback
,
H. R.
(
1985
).
Functional and immunochemical characterization of a mutant Escherichia coli energy-coupled for lactose transport
.
Biochemistry, N.Y.
24
,
221
229
.
Herzlinger
,
D.
,
Viitanen
,
P.
,
Carrasco
,
N.
and
Kaback
,
H. R.
(
1984
).
Monoclonal antibodies against the lac carrier protein from Escherichia coli. II. Binding studies with membrane vesicles and proteoliposomes reconstituted with purified lac carrier protein
.
Biochemistry, N.Y.
23
,
3688
3693
.
Jung
,
H.
,
Jung
,
K.
and
Kaback
,
H. R.
(
1994a
).
A conformational change in the lactose permease of Escherichia coli is induced by ligand binding or membrane potential
.
Prot. Sci. (in press)
.
Jung
,
K.
,
Jung
,
H.
and
Kaback
,
H. R.
(
1994b
).
Dynamics of lactose permease of Escherichia coli determined by site-directed fluorescence labeling
.
Biochemistry, N.Y.
33
,
3980
3985
.
Jung
,
K.
,
Jung
,
H.
,
Wu
,
J.
,
Privé
,
G. G.
and
Kaback
,
H. R.
(
1993
).
Use of site-directed fluorescence labeling to study proximity relationships in the lactose permease of Escherichia coli
.
Biochemistry, N.Y.
32
,
12273
12278
.
Kaback
,
H. R.
(
1986
).
Active transport in Escherichia coli: from membrane to molecule
. In
Physiology of Membrane Disorders
(ed.
T. E.
Andreoli
,
J. F.
Hoffman
,
D. D.
Fanestil
and
S. G.
Schultz
), pp.
387
408
.
New York
:
Plenum Press
.
Kaback
,
H. R.
(
1989
).
Molecular biology of active transport: from membranes to molecules to mechanism
.
The Harvey Lectures Series
83
,
77
103
.
Kaback
,
H. R.
(
1992
).
In and out and up and down with the lactose permease of Escherichia coli. In Bacterial and Glucose Transporters
(ed.
M.
Friedlander
and
M.
Mueckler
), pp.
97
125
.
Int. Rev. Cytol.
137A
.
New York
:
Academic Press
.
Kaback
,
H. R.
,
Jung
,
K.
,
Jung
,
H.
,
Wu
,
J.
,
Privd
,
G. G.
and
Zen
,
K. H.
(
1993
).
What’s new with lactose permease?
J. Bioenerg. Biomembr.
25
,
627
636
.
Kaczorowski
,
G. J.
,
Leblanc
,
G.
and
Kaback
,
H. R.
(
1980
).
Specific labeling of the lac carrier protein in membrane vesicles of Escherichia coli by a photoaffinity reagent
.
Proc. natn. Acad. Sci. U.S.A.
77
,
6319
6323
.
King
,
S. C.
,
Hansen
,
C. L.
and
Wilson
,
T. H.
(
1991
).
The interaction between aspartic acid 237 and lysine 358 in the lactose carrier of Escherichia coli
.
Biochim. biophys. Acta
1062
,
177
186
.
Kinnunen
,
P. K. J.
,
Koiv
,
A.
and
Mustonen
,
P.
(
1993
).
In Fluorescence Spectroscopy
(ed.
O. S.
Wolfbeis
), pp.
159
. Berlin,
Heidelberg, New York
:
Springer Verlag
.
Lee
,
J. I.
,
Hwang
,
P. P.
,
Hansen
,
C.
and
Wilson
,
T. H.
(
1992
).
Possible salt bridges between transmembrane a.-helices of the lactose carrier of Escherichia coli
.
J. biol. Chem.
267
,
20758
20764
.
Lolkema
,
J. S.
,
Püttner
,
I. B.
and
Kaback
,
H. R.
(
1988
).
Site-directed mutagenesis of Pro327 in the lac permease of Escherichia coli
.
Biochemistry, N.Y.
27
,
8307
8310
.
Marger
,
M. D.
and
Saier
,
M. H.
, JR
(
1993
).
A major superfamily of transmembrane facilitators that catalyse uniport, symport and antiport
.
Trends biochem. Sci.
18
,
13
20
.
Matsushita
,
K.
,
Patel
,
L.
,
Gennis
,
R. B.
and
Kaback
,
H. R.
(
1983
).
Reconsitution of active transport in proteoliposomes containing cytochrome o oxidase and lac carrier protein purified from Escherichia coli
.
Proc. natn. Acad. Sci. U.S.A.
80
,
4889
4893
.
McKenna
,
E.
,
Hardy
,
D.
and
Kaback
,
H. R.
(
1992a
).
Evidence that the final turn of the last transmembrane helix in the lactose permease is required for folding J. biol. Chem.
267
,
6471
6474
.
McKenna
,
E.
,
Hardy
,
D.
and
Kaback
,
H. R.
(
1992b
).
Insertional mutagenesis of hydrophilic domains in the lactose permease of Escherichia coli
.
Proc. natn. Acad. Sci. U.S.A.
89
,
11954
11958
.
McKenna
,
E.
,
Hardy
,
D.
,
Pastore
,
J. C.
and
Kaback
,
H. R.
(
1991
).
Sequential truncation of the lactose permease over a three amino-acid sequence near the carboxyl-terminus leads to progressive loss of activity and stability
.
Proc. natn. Acad. Sci. U.S.A.
88
,
2969
2973
.
Menezes
,
M. E.
,
Roepe
,
P. D.
and
Kaback
,
H. R.
(
1990
).
Design of a membrane transport protein for fluorescence spectroscopy
.
Proc. natn. Acad. Sci. U.S.A.
87
,
1638
1642
.
Menick
,
D. R.
,
Carrasco
,
N.
,
Antes
,
L.
,
Patel
,
L.
and
Kaback
,
H. R.
(
1987a
).
Lac permease of Escherichia coli: arginine-302 as a component of the postulated proton relay
.
Biochemistry, N.Y.
26
,
6638
6644
.
Menick
,
D. R.
,
Lee
,
J. A.
,
Brooker
,
R. J.
,
Wilson
,
T. H.
and
Kaback
,
H. R.
(
1987b
).
Role of cysteine residues in the lac permease of Escherichia coli
.
Biochemistry, N.Y.
26
,
1132
1136
.
Menick
,
D. R.
,
Sarkar
,
H. K.
,
Poonian
,
M. S.
and
Kaback
,
H. R.
(
1985
).
Cys154 is important for lac permease activity in Escherichia coli
.
Biochem. biophys. Res. Commun.
132
,
162
170
.
Mitchell
,
P.
(
1963
).
Molecule, group and electron transport through natural membranes
.
Biochem. Soc. Symp
.
22
,
142
168
.
Mitchell
,
P.
(
1968
).
Chemiosmotic Coupling and Energy Transduction. England: Glynn Research Ltd.
Neuhaus
,
J. M.
,
Soppa
,
J.
,
Wright
,
J. K.
,
Reide
,
I.
,
Blocker
,
H.
,
Frank
,
R.
AND
Overath
,
P.
(
1985
).
Properties of a mutant lactose carrier of Escherichia coli with a Cys148????????Ser148 substitution
.
FEBS Lett.
185
,
83
88
.
Newman
,
M. J.
,
Foster
,
D. L.
,
Wilson
,
T. H.
and
Kaback
,
H. R.
(
1981
).
Purification and reconstitution of functional lactose carrier from Escherichia coli
.
J. biol. Chem.
256
,
11804
11808
.
Newman
,
M. J.
and
Wilson
,
T. H.
(
1980
).
Solubilization and reconstitution of the lactose transport system from Escherichia coli
.
J. biol. Chem.
255
,
10583
10586
.
Padan
,
E.
,
Sarkar
,
H. K.
,
Viitanen
,
P. V.
,
Poonian
,
M. S.
and
Kaback
,
H. R.
(
1985
).
Site-specific mutagenesis of histidine residues in the lac permease of Escherichia coli
.
Proc. natn. Acad. Sci. U.S.A.
82
,
6765
6768
.
Page
,
M. G. P.
and
Rosenbusch
,
J. P.
(
1988
).
Topography of lactose permease from Escherichia coli
.
J. biol. Chem.
263
,
15906
15914
.
Privé
,
G.
,
Verner
,
G. E.
,
Weitzman
,
C.
,
Zen
,
K.
,
Eisenberg
,
D.
and
Kaback
,
H. R.
(
1994
).
Fusion proteins as tools for crystallization: The lactose permease of Escherichia coli
.
Acta cryst. (in press)
.
Püttner
,
I. B.
and
Kaback
,
H. R.
(
1988
).
Lac permease of Escherichia coli containing a single histidine residue is fully functional
.
Proc. natn. Acad. Sci. U.S.A.
85
,
1467
1471
.
Püttner
,
I. B.
,
Sarkar
,
H. K.
,
Padan
,
E.
,
Lolkema
,
J. S.
and
Kaback
,
H. R.
(
1989
).
Characterization of site-directed mutants in the lac permease of Escherichia coli
.
Biochemistry, N.Y.
28
,
2525
2533
.
Püttner
,
I. B.
,
Sarkar
,
H. K.
,
Poonian
,
M. S.
and
Kaback
,
H. R.
(
1986
).
Lac permease of Escherichia coli: His-205 and His-322 play different roles in lactose/H+ symport
.
Biochemistry, N.Y.
25
,
4483
4485
.
Roepe
,
P. D.
and
Kaback
,
H. R.
(
1989
).
Site-directed mutagenesis of tyrosine residues in the lac permease of Escherichia coli
.
Biochemistry, N.Y.
28
,
6127
6132
.
Roepe
,
P. D.
,
Zbar
,
R.
,
Sarkar
,
H. K.
and
Kaback
,
H. R.
(
1989
).
A five residue sequence near the carboxyl terminus of the polytopic membrane protein lac permease is required for stability within the membrane
.
Proc. natn. Acad. Sci. U.S.A.
86
,
3992
3996
.
Sahin-Tóth
,
M.
,
Dunten
,
R. L.
,
Gonzalez
,
A.
and
Kaback
,
H. R.
(
1992
).
Functional interactions between putative intramembrane charged residues in the lactose permease of Escherichia coli
.
Proc. natn. Acad. Sci. U.S.A.
89
,
10547
10551
.
Sahin-Tóth
,
M.
and
Kaback
,
H. R.
(
1993a
).
Cysteine scanning mutagenesis of putative transmembrane helices IX and X in the lactose permease of Escherichia coli
.
Protein Sci.
2
,
1024
1033
.
Sahin-Tóth
,
M.
and
Kaback
,
H. R.
(
1993b
).
Properties of interacting aspartic acid and lysine residues in the lactose permease of Escherichia coli
.
Biochemistry, N.Y.
32
,
10027
10035
.
Sahin-Tóth
,
M.
,
Persson
,
B.
,
Schwieger
,
J.
,
Cohan
,
P.
and
Kaback
,
H. R.
(
1994
).
Cysteine scanning mutagenesis of the N-terminal 32 amino acid residues in the lactose permease of Escherichia coli
.
Protein Sci.
3
,
240
247
.
Sahin-Tóth
,
M.
,
Lawrence
,
M. C.
and
Kaback
,
H. R.
(
1994
).
Properties of permease dimer, a fusion protein containing two lactose permease molecules from E. coli
.
Proc. natn. Acad. Sci. U.S.A.
91
,
5421
5425
.
Sarkar
,
H. K.
,
Menick
,
D. R.
,
Viitanen
,
P. V.
,
Poonian
,
M. S.
and
Kaback
,
H. R.
(
1986
).
Site-specific mutagenesis of cysteine 148 to serine in the lac permease of Escherichia coli
.
J. biol. Chem
.
261
,
8914
8918
.
Seckler
,
R.
,
Moroy
,
T.
,
Wright
,
J. K.
and
Overath
,
P.
(
1986
).
Anti-peptide antibodies and proteases as structural probes for the lactose/H+ transporter of Escherichia coli: a loop around amino acid residue 130 faces the cytoplasmic side of the membrane
.
Biochemistry, N.Y.
25
,
2403
2409
.
Seckler
,
R.
and
Wright
,
J. K.
(
1984
).
Sidedness of native membrane vesicles of Escherichia coli and orientation of the reconstituted lactose/H+ carrier
.
Eur. J. Biochem.
142
,
269
279
.
Seckler
,
R.
,
Wright
,
J. K.
and
Overath
,
P.
(
1983
).
Peptide-specific antibody locates the COOH terminus of the lactose carrier of Escherichia coli on the cytoplasmic side of the plasma membrane
.
J. biol. Chem.
258
,
10817
10820
.
Stochaj
,
V.
,
Bieseler
,
B.
and
Ehring
,
R.
(
1986
).
Limited proteolysis of lactose permease from Escherichia coli
.
Eur. J. Biochem.
158
,
423
428
.
Teather
,
R. M.
,
Müller-Hill
,
B.
,
Abrutsch
,
U.
,
Aichele
,
G.
and
Overath
,
P.
(
1978
).
Amplification of the lactose carrier protein in Escherichia coli using a plasmid vector
.
Molec. gen. Genet.
159
,
239
248
.
Trumble
,
W. R.
,
Viitanen
,
P. V.
,
Sarkar
,
H. K.
,
Poonian
,
M. S.
and
Kaback
,
H. R.
(
1984
).
Site-directed mutagenesis of cys148 in the lac carrier protein of Escherichia coli. Biochem. biophys
.
Res. Commun.
119
,
860
867
.
Ujwal
,
M. L.
,
Sahin-Tóth
,
M.
,
Persson
,
B.
and
Kaback
,
H. R.
(
1994
).
Role of glutamate-269 in the lactose permease of Escherichia coli
.
Molec. Membr. Biol.
1
,
9
16
.
van Iwaarden
,
P. R.
,
Driessen
,
A. J.
,
Menick
,
D. R.
,
Kaback
,
H. R.
and
Konings
,
W. N.
(
1991a
).
Characterization of purified, reconstituted site-directed cysteine mutants of the lactose permease of Escherichia coli
.
J. biol. Chem.
266
,
15688
15692
.
van Iwaarden
,
P. R.
,
Pastors
,
J. C.
,
Kanings
,
W. N.
and
Kaback
,
H. R.
(
1991b
).
Construction of a functional lactose permease devoid of cysteine residues
.
Biochemistry, N.Y.
30
,
9595
9600
.
Viitanen
,
P.
,
Garcia
,
M. L.
and
Kaback
,
H. R.
(
1984
).
Purified reconstituted lac carrier from Escherichia coli is fully functional
.
Proc. natn. Acad. Sci. U.S.A.
81
,
1629
1633
.
Viitanen
,
P.
,
Newman
,
M. J.
,
Foster
,
D. L.
,
Wilson
,
T. H.
and
Kaback
,
H. R.
(
1986
).
Purification, reconstitution and characterization of the lac permease of Escherichia coli
.
Meth. Enzymol.
125
,
429
452
.
Viitanen
,
P. V.
,
Menick
,
D. R.
,
Sarkar
,
H. K.
,
Trumble
,
W. R.
and
Kaback
,
H. R.
(
1985
).
Site-directed mutagenesis of cysteine-148 in the lac permase of Escherichia coli: effect on transport, binding, and sulfhydryl inactivation
.
Biochemistry, N.Y.
24
,
7628
7635
.
Vogel
,
H.
,
Wright
,
J. K.
and
Jahnig
,
F.
(
1985
).
The structure of the lactose permease from Raman spectroscopy and prediction methods
.
EMBO J.
4
,
3625
3631
Wrubel
,
W.
,
Stochaj
,
U.
,
Sonnewald
,
U.
,
Theres
,
C.
and
Ehring
,
R.
(
1990
).
Reconstitution of an active lactose carrier in vivo by simultaneous synthesis of two complementary protein fragments
.
J. Bacteriol.
172
,
5374
5381
.
Zen
,
H. K.
,
McKenna
,
E.
,
Bibi
,
E.
and
Kaback
,
H. R.
(
1994
).
Expression of lactose permease in contiguous fragments as a probe for membrane-spanning domains
.
Biochemistry, N.Y
.
33
,
8198
8206
.