Abstract
Defensins are cationic and disulfide-bonded host defense proteins of many animals that target microbial cell membranes. Elucidating the three-dimensional structure, dynamics and topology of these proteins in phospholipid bilayers is important for understanding their mechanisms of action. Using solid-state NMR spectroscopy, we have now determined the conformation, dynamics, oligomeric state and topology of a human α-defensin, HNP-1, in DMPC/DMPG bilayers. 2D correlation spectra show that membrane-bound HNP-1 exhibits a similar conformation to the water-soluble state, except for the turn connecting the β2 and β3 strands, whose sidechains exhibit immobilization and conformational perturbation upon membrane binding. At high protein/lipid ratios, rapid 1H spin diffusion from the lipid chains to the protein was observed, indicating that HNP-1 was well inserted into the hydrocarbon core of the bilayer. Arg Cζ-lipid 31P distances indicate that only one of the four Arg residues forms tight hydrogen-bonded guanidinium-phosphate complexes. The protein is predominantly dimerized at high protein/lipid molar ratios, as shown by 19F spin diffusion experiments. The presence of a small fraction of monomers and the shallower insertion at lower protein concentrations suggest that HNP-1 adopts concentration-dependent oligomerization and membrane-bound structure. These data strongly support a “dimer pore” topology of HNP-1 in which the polar top of the dimer lines an aqueous pore while the hydrophobic bottom faces the lipid chains. In this structure R25 lies closest to the membrane surface among the four Arg residues. The pore does not have large lipid disorder, in contrast to the toroidal pores formed by protegrin-1, a two-stranded β-hairpin antimicrobial peptide. These results provide the first glimpse into the membrane-bound structure and mechanism of action of human α-defensins.
Keywords: Human alpha defensins, solid-state NMR, depth of insertion, guanidinium-phosphate interaction, antimicrobial peptides
Defensins are host-defense antimicrobial proteins present in many animals and plants. These cationic polypeptides contain three intramolecular disulfide bonds, whose linkage patterns distinguish the α-, β-, and θ-defensins (1). Humans have six α-defensins of about 30 residues long (2). The human neutrophil peptides (HNP) 1–4 were found in the azurophilic granules of neutrophils (3–5), while HD-5 and HD-6 are present in intestinal epithelial cells (6, 7). Except for HD-6, the human α-defensins show wide-spectrum antimicrobial activities in the microgram per milliliter range (8, 9). Similar to many other antimicrobial peptides (AMPs), the general mechanism of antimicrobial action is believed to be permeabilization of the microbial cell membrane (10, 11).
High-resolution crystal structures of all six human α-defensins in the absence of membrane-mimetic solvents have been determined (12, 13). The structures are very similar despite significant sequence differences between HNP 1–3 and the other three defensins. The proteins consist of three antiparallel β-strands (β1, β2, β3) connected by turns and a longer loop (Figure 1a). The proteins are dimerized in the crystal through intermolecular H-bonds between the two β2 strands, extending the triple-stranded β-sheet to a six-stranded β-sheet. The dimer has a basket shape and is amphipathic, with a polar top and an apolar base.
Figure 1.
Different structural models of HNPs in lipid bilayers. The basic oligomeric state of the protein is a basket-shaped dimer. (a) Structure of the HNP monomer (12) with the four Arg sidechains indicated. (b) Surface-bound wedge model with the hydrophobic basket bottom inserted into the bilayer and the polar basket top in contact with water. (c) Membrane-inserted wedge model. (d) Dimer pore model, where the hydrophobic basket bottom contacts the lipids while the polar top faces the water pore. (e) General pore model, where the dimers are rotated by 90° from the dimer pore orientation. (f) Multimeric pore model, where six dimers form a pore with a ~25 Å inner diameter. The height of the multimer, measured between two Pro8 residues, is 27 Å. The approximate dimensions of the various oligomeric assemblies are indicated in (b-f).
Based on the crystal structure of HNP-3 (12), several mechanistic models had been proposed to account for the antimicrobial activities of HNPs. The wedge model (12) suggests that the dimer inserts into the bilayer with the hydrophobic bottom contacting the hydrocarbon region of the membrane, thus disrupting lipid packing (Figure 1b, c). The depth of the wedge was unspecified, so the dimer may be partly or completely immersed in the lipid membrane. Alternatively, HNPs might form membrane-spanning pores with the hydrophobic base facing the lipids. In the so-called “dimer pore” model (12) (Figure 1d), the polar tops of two dimers are oriented to favor transmembrane (TM) orientation of a small solvent channel observed at the dimer interface in the crystal structure (12). In the “general pore” model (12), the dimers are rotated by 90° around the horizontal axis (Figure 1e), so that the side view, parallel to the membrane normal, shows the basket shape of the dimer. A multimeric pore model was also proposed based on vesicle leakage and dextran permeability experiments (14). Here 6–8 HNP dimers may form a large pore with an inner diameter of 20–25 Å. The dimers are oriented with the long axis of the basket top at about 45° from the bilayer normal, so that the Arg residues are located in two rings separated by about 16 Å along the bilayer normal, promoting favorable electrostatic interactions with the lipid phosphates (Figure 1f).
Despite many crystal structures of HNPs, no high-resolution structure information for membrane-bound HNPs has been reported. To elucidate the mechanism of action of human defensins, we have initiated a solid-state NMR study of uniformly 13C and 15N labeled recombinant HNP-1. Since membrane-bound proteins typically have broad linewidths than proteins outside the membrane, due to the conformational disorder of the lipids, we found it necessary to first conduct resonance assignment of the protein in the more ordered water-soluble and lipid-free state. Unlike the bactericidal channel-forming colicins (15–17), which exhibit abundant membrane-induced flexibility, HNPs cannot undergo large conformational rearrangements due to their multiple disulfide bonds. We thus first determined the NMR structure of lipid-free microcrystalline HNP-1 using 2D and 3D magic-angle-spinning (MAS) 13C-13C and 13C-15N correlation experiments (18, 19). Together with distance constraints, these data led to the first solid-state NMR structure of a human α-defensin, which confirmed the three-dimensional fold seen in the crystal structures (PDB code: 2KHT). The NMR structure determination yielded the complete 13C and 15N chemical shifts of the protein (19), with which the membrane-bound HNP-1 chemical shifts can now be compared. In this work, we investigate the conformation and dynamics of HNP-1 in the membrane, its oligomeric state (20, 21), and most importantly, its interactions with lipids and water. These results yielded the global topology of HNP-1 in the lipid bilayer, allowing us to rule out most structural models and propose the membrane-disruptive mechanism of HNP-1.
Materials and Methods
HNP-1 expression and synthesis
Recombinant 13C, 15N-labeled HNP-1 (ACYCRIPACIAGERRYGTCIYGQRLWAFC- C (Figure 3a) was obtained as a cleavage product from its precursor protein, proHNP1. The residues are numbered from A2 to C31 due to sequence alignment with other mammalian defensins, many of which contain an additional N-terminal residue before A2 (12). Using this numbering system also facilitates comparison with HNP-3, which was the first structure determined in the HNP family. Briefly, the expression protocol starts with expressing proHNP1 as a GST-fusion protein in E. coli using 13C, 15N-labeled Spectra 9 media (Cambridge Isotope Laboratories) (19, 22). GST-proHNP1 was folded and then cleaved using thrombin, producing proHNP1. ProHNP1 was purified by reversed-phase HPLC, then cleaved using cyanogen bromide to yield correctly folded HNP-1. Antimicrobial assays confirmed the activity of the protein. For example, 100% killing of S. aureus is reached at 64 μg/ml HNP-1 (19, 22). Crude HNP-1 was purified by reversed-phase HPLC. About 1.5 mg HNP-1 (molecular weight: 3634 Da) was purified from 1 L culture. 19F-labeled HNP-1 for CODEX experiments was synthesized using t-Boc chemistry (23), where Tyr4 was replaced by 4-19F-Phg. The use of phenylglycine removes the possibility of sidechain χ1 motion that could complicate the interpretation of the CODEX data (24, 25). The choice of Tyr4 as the 19F labeled site was based on the crystal structures of various HNPs, all showing close inter-subunit contact of this residue (12, 13, 26). Previous studies of 4-19F-Phg mutants of the related β-hairpin antimicrobial peptide, PG-1 (24), suggest that the mutation generally does not perturb the antimicrobial activities of the peptides.
Figure 3.
Resonance assignment of membrane-bound HNP-1 and comparison with the microcrystalline protein. (a) Amino acid sequence of HNP-1. (b) 2D 13C-13C DARR spectrum of membrane-bound HNP-1 at 273 K with a mixing time of 40 ms. (c-e) Regions that show chemical shift changes between the membrane-bound (black) and microcrystalline (red) states. (c) I21 sidechain. (d) R25 sidechain. (e) R25 and R16 sidechains. Note the significant increase of the R25 sidechain intensities in the membrane-bound state in (d) and (e). (f) Cα and Cβ secondary chemical shifts of HNP-1 in the membrane-bound and microcrystalline states, confirming the β-strand-rich nature of the protein and the overall similarity of the protein structure in the two states.
Membrane sample preparation
DMPC and DMPG lipids (Avanti Polar Lipids, Alabaster, AL) were codissolved in chloroform at a molar ratio of 3:1, dried under a stream of nitrogen gas, redissolved in cyclohexane and lyophilized overnight. The dry and homogeneous lipid powder was suspended in a pH 7 phosphate buffer, vortexed and then freeze-thawed five times. The vesicle solution was then extruded through polycarbonate filters with pore sizes of 400 nm and 100 nm to obtain large unilamellar vesicles. HNP-1 was dissolved in 1 mL phosphate buffer, added to the lipid vesicle solution and dialyzed overnight. The mixture was centrifuged at 200,000 g for 3 hours at 4°C to obtain a membrane pellet. The pellet was pipette-transferred into 4 mm MAS rotors for NMR experiments.
Three samples were prepared: two U-13C, 15N-labeled HNP-1 samples at protein : lipid (P/L) molar ratios of 1: 18 and 1: 40, and one 19F-labeled HNP-1 sample at P/L = 1: 18. The 1:18 molar ratio corresponds to a mass ratio of 1 : 3.3, thus there was sufficient lipid to ensure appropriate protein-lipid interactions.
Solid-state NMR spectroscopy
Most 2D resonance assignment experiments were carried out on a Bruker (Karlsruhe, Germany) AVANCE-600 (14.1 Tesla) spectrometer operating at a 13C Larmor frequency of 150.92 MHz. A Bruker DSX-400 (9.4 Tesla) spectrometer was used for 13C-31P REDOR experiments, 1H spin diffusion and 13C-1H dipolar coupling measurements. Triple-resonance MAS probes with 4-mm rotors were used for all experiments. 13C and 31P chemical shifts were referenced externally to the α-glycine 13CO signal at 176.49 ppm on the TMS scale and the hydroxyapatite 31P signal at 2.73 ppm on the phosphoric acid scale, respectively. 15N chemical shifts were referenced to the 15N signal of N-acetyl-valine at 122.0 ppm on the liquid ammonia scale.
2D 13C-13C DARR correlation spectra (27) were measured at 273 K under 5 kHz MAS with a 13C spin diffusion time of 40 ms. 2D N(CO)CX and N(CA)CX correlation spectra (28, 29) were measured at 273 K under 7.5 kHz MAS with a 13C mixing time of 40 ms for NCACX and 60 ms for NCOCX. The 15N-13C SPECIFIC (30) cross-polarization (CP) contact times were 3 ms. In the intra-residue NCACX experiment, 15N-13Cα magnetization transfer was achieved using a 15N spin-lock field of 17 kHz and an on-resonance 13Cα spin-lock field of 25 kHz. In the inter-residue NCOCX experiment, 15N-13CO magnetization transfer was accomplished with a 15N on-resonance spin-lock field of 27 kHz and a tilted 13CO spin-lock field strength of 35 kHz, which was the result of a 30 kHz applied field on resonance with Cα and an 18 kHz offset for CO.
13C-1H dipolar couplings were measured at 303 K using the 2D dipolar chemical-shift (DIPSHIFT) correlation experiment (31). The sample was spun at 4.5 kHz, and the MREV-8 sequence (32) was used for 1H homonuclear decoupling. The t1-curves were fit using a Fortran program, and the fit values were divided by the MREV-8 scaling factor to obtain the true couplings. To account for uncertainties in the MREV-8 scaling factor, we measured the rigid-limit C-H dipolar couplings using the crystalline model peptide formyl-Met-Leu-Phe (f-MLF) (33, 34), where Leu Cα, Met Cβ, and Leu Cδ signals represented CH, CH2, and CH3 groups, respectively. The CH and CH2 peaks gave C-H rigid-limit dipolar couplings of 11.2 kHz (fit values), which corresponded to true couplings of 23.8 kHz when the theoretical scaling factor of 0.47 for MREV-8 was used (35). For the Leu CH3 signal, the fit value was 3.5 kHz. The ratio of the measured HNP-1 couplings with the rigid-limit f-MLF couplings yielded the order parameter SCH.
To determine the depth of insertion of HNP-1, we carried out 2D 13C-detected 1H spin diffusion experiments (36, 37) at 303 K or 308 K under 5 kHz MAS. For the P/L = 1:18 sample, 1H spin diffusion mixing times (tm) of 2.25 – 225 ms were applied to transfer the 1H magnetization of mobile lipids and water to HNP-1. For the P/L=1:40 sample, two mixing times of 100 ms and 225 ms were measured since the sensitivity was too low to measure a complete buildup curve. The 1H magnetization transfer was detected through the protein 13C signals. To ensure that only the mobile lipid and water polarization served as the 1H magnetization source, we suppressed the 1H magnetization of the rigid protein by a 1H T2 filter of 0.4 × 2 ms before the t1 evolution period. For the 2.25 ms spectrum of the 1:18 sample, the water 1H cross section was extracted to analyze the water-proximal residues in the protein.
A frequency-selective rotational-echo double-resonance (REDOR) experiment (38) was used to measure protein 13C distances to lipid 31P. The experiments were performed at 233 K under 4.5 kHz MAS to suppress lipid motions as well as protein sidechain motions. A rotor-synchronized 13C Gaussian 180° pulse of 444 μs was applied in the middle of the REDOR period to suppress 13C-13C J-couplings between the on-resonance 13C and its directly bonded spins. 31P 180° pulses of 9 μs were applied every half a rotor period to recouple the 13C-31P dipolar coupling. The time-dependent REDOR intensities were fit using two-spin simulations, which had been shown (39) to correspond to the vertical distance between the 13C spin and the 31P plane for long distances of greater than ~6 Å.
19F CODEX experiments (20, 40, 41) for determining the oligomeric state of HNP-1 were carried out at 233 K under 7 kHz MAS. The 19F chemical shift anisotropy (CSA) was recoupled by two rotor periods of 180° pulses. During the mixing time, 19F spin diffusion changes the 19F CSA and prevents complete refocusing of a stimulated echo. A z-filter after the second 180°-pulse train allows the correction of 19F T1 relaxation effects by conducting two experiments for each tm, a control experiment (S0) in which the short z-filter period (10 μs) was applied between the two 180°-pulse trains while the long tm occurred after, and a dephasing experiment (S) where the long tm occurs between the two 180°-pulse trains. The intensity ratio, S/S0, at equilibrium yields the number of 19F spins in close proximity. Two mixing times, 500 ms and 1 s, were measured.
Results
Conformation and dynamics of HNP-1 in the lipid membrane
Figure 2a,b shows the 31P spectrum of the DMPC/DMPG (3:1) membrane with bound HNP-1 above (303 K) and below (253 K) the membrane phase transition temperature. At 303 K the 31P spectrum has the classical uniaxial lineshape indicative of lamellar bilayers and is identical to the protein-free lipid spectrum, indicating that HNP-1 does not cause observable orientation disorder to the lipid bilayer. Below the phase transition temperature (23°C) the 31P spectrum exhibited the expected broadening, with a rigid-limit span of 210 ppm at 253 K.
Figure 2.
1D 31P and 13C spectra of membrane-bound HNP-1. (a) 31P static spectra of DMPC/DMPG membranes with HNP-1 (P/L = 1:18) (thick line) and without HNP-1 (thin line), measured at 303 K. The protein-bound spectrum exhibits no isotropic peak and is identical to the non-protein-containing spectrum. (b) 31P spectrum of HNP-1-bound DMPC/DMPG membrane at 253 K. (c, d) 13C MAS spectrum of membrane-bound HNP-1 at 298 K (c) and 273 K (d). The intensities and linewidths are little affected by temperature, indicating that HNP-1 is immobilized in the liquid-crystalline membrane. (e) 13C spectrum of microcrystalline HNP-1 without the lipids at 268 K. The intensity distribution of the membrane-bound HNP-1 is similar to that of the microcrystalline protein, indicating similar conformations.
We examined the membrane-bound conformation of HNP-1 by first comparing the 1D 13C spectra of the membrane-bound and microcrystalline states. Figure 2c,d show the 13C spectra of DMPC/DMPG bound HNP-1 at 298 K and 273 K. The intensity and linewidths were largely unaffected by temperature, suggesting that HNP-1 was immobilized at ambient temperature. Compared to the microcrystalline protein spectrum (Figure 2e), the membrane-bound sample exhibited similar intensity distributions except for the additional lipid peaks. Thus, HNP-1 adopts similar overall conformation in the membrane as in the microcrystalline state, as expected for this disulfide-bonded protein.
To better resolve the resonances, we measured 2D 13C-13C and 15N-13C correlation spectra. The 2D 13C-13C DARR spectrum at 40 ms mixing (Figure 3b) showed mostly intra-residue cross peaks, which were readily assigned by comparison with the microcrystalline spectra (19). 2D N(CO)CX and N(CA)CX spectra (Figure S1) further corroborated the assignments. Most residues exhibited similar chemical shifts in the membrane as in the microcrystalline state, except for the sidechains of I21 and R25. The I21 Cγ and Cε peaks moved downfield from the microcrystalline state: the Cγ chemical shift increased from 24.6 ppm to 25.5 ppm while the Cε peak moved from 11.5 ppm to 12.4 ppm (Figure 3c). In the microcrystalline state, the R25 sidechain resonances were not visible in the 2D 13C-13C spectrum and much weaker than the R6 and R16 signals in the 3D NCC spectra (19). In comparison, in the membrane-bound state, the R25 sidechain resonances became much stronger, with clear Cβ-Cγ and Cδ-Cβ cross peaks (Figure 3d, e). I21 and R25 are both located in the β2-β3 turn at the hydrophobic base of the dimer (Figure 1a). This turn was seen in the HNP-3 crystal structure to deviate significantly from crystallographic symmetry and was thought to exhibit flexing motions (12). Thus, the membrane-induced conformational and dynamical changes of I21 and R25 sidechains are likely significant, and may reflect site-specific interactions of this β2-β3 turn with lipids (see below). Figure 3f shows the Cα and Cβ secondary chemical shifts (42) of HNP-1 for both the membrane-bound and microcrystalline states, confirming the triple-stranded motif of the protein and the similarity of the protein conformation without and with the lipids. The secondary shifts suggest that the Y17-G18 junction in the middle of the β2 strand deviates from the ideal β-sheet structure, as noted before (19), which may reflect the role of G18 as a hinge for the β2-β3 hairpin (13).
To evaluate the mobility of membrane-bound HNP-1 quantitatively, we measured the 13C-1H dipolar couplings using the 2D DIPSHIFT experiment. Figure 4 shows the 13C chemical shift dimension of the 2D spectrum, with partial resolution of the sites. Several dipolar cross sections representing the backbone Cα-Hα, sidechain CH2 and CH3 groups are shown. We obtained rigid-limit SCH’s of 0.95–1.05 for Cα and methylene groups, while the methyl groups of Ile and Ala exhibited SCH values of 0.26–0.28. Since methyl three-site jumps alone give an order parameter of 0.33, the small degree of additional scaling, by a factor of 0.78–0.85, indicates that other torsional degrees of freedom in the aliphatic sidechains of these residues are insignificant. Thus the DIPSHIFT data is consistent with the temperature insensitivity of the 13C spectra and indicate that HNP-1 backbone is immobilized in the liquid-crystalline phase of the bilayer, and only the sidechains exhibit moderate motion. This finding is not only consistent with the conformational rigidity of this disulfide-bonded protein, but more importantly indicates that HNP-1 is not monomeric in the membrane. If the protein were monomeric, the small molecular weight would cause fast whole-body uniaxial diffusion of the protein in the membrane, as observed in many small and medium-sized membrane proteins with molecule weights up to several tens of kDa (43–47). Since all HNPs form dimers in the crystal (12, 13, 48), the immobilization suggests that when bound to the membrane, HNP-1 is at least dimeric and may form higher-order oligomers.
Figure 4.
13C-1H dipolar couplings of DMPC/DMPG-bound HNP-1 at 303 K under 4.5 kHz MAS. (a) 13C chemical shift dimension of the 2D spectrum. (b) C-H dipolar coupling curves of two Cα cross sections and their best-fit, compared with the Leu Cα data of f-MLF. (c) C-H dipolar coupling curves of two CH2 peaks and their best-fit, compared with the Met Cβ data of f-MLF. (d) C-H dipolar coupling curves of two CH3 cross sections and their best-fit simulations at 3.0 and 3.2 kHz, compared to the Leu Cδ signal of f-MLF. All couplings here are fit values, and convert to true couplings after they are divided by the homonuclear decoupling scaling factor.
Depth of insertion of HNP-1 from lipid-protein 1H spin diffusion
To determine the immersion depth of HNP-1, we carried out a 2D 13C-detected 1H spin diffusion experiment (36). The experiment measures inter-proton distances from the mobile water and lipid acyl chains to the rigid protein through distance-dependent 1H magnetization transfer. TM proteins in contact with both the membrane surface and the hydrocarbon core exhibit fast 1H spin diffusion from both lipid acyl chains and water (37), while surface-bound proteins have much slower 1H spin diffusion from the lipid chains.
Figure 5 shows the 1H spin diffusion data of membrane-bound HNP-1 (P/L = 1:18) at 305 K. A representative 2D spectrum (mixing time of 100 ms) shows the characteristic cross peaks between the protein 13C signals (63 – 37 ppm) and the lipid CH2 protons at 1.3 ppm, indicating that HNP-1 is within spin diffusion reach of the acyl chains. Strong water-protein cross peaks were also observed. To quantify the distances, we measured the 2D spectra as a function of the mixing time. Figure 5b shows the integrated 1H cross sections for 13C chemical shifts of 63–37 ppm for mixing times of 9 to 225 ms. After correcting for 1H T1 relaxation of the water and CH2 protons, we plotted the cross-peak intensities as a function of the square root of the mixing time. Figure 5c shows that both the water and lipid cross peaks reached a plateau by 100 ms, qualitatively indicating that HNP-1 is in close contact with both the hydrophobic core of the membrane and water. The equilibration allows both the water and lipid cross peaks to be normalized with respect to their own maximum intensities at 100 ms.
Figure 5.
1H spin diffusion from lipid chains and water to HNP-1 in DMPC/DMPG bilayers. (a) Representative 2D spectrum for P/L = 1:18 at 303 K, with a 1H mixing time of 100 ms. (b) Integrated 1H cross sections from 13C chemical shifts of 63 – 37 ppm as a function of 1H mixing times, compared with the 1D 1H spectrum at the top. (c) 1H spin diffusion buildup curves for lipid CH2 and water at P/L=1:18. (d) 1H spin diffusion intensities for lipid CH2 and water at P/L=1:40. The CH2 buildup is much slower at the lower protein concentration.
We simulated the CH2 buildup curve using a 1D lattice model to obtain the minimum lipid-protein separation. In this model, 1H magnetization (Mi) transfer was modeled as a discrete process along the bilayer normal (49). The transfer rate, Ω = D/a2, depends on the lattice spacing a, fixed at 2 Å, and the diffusion coefficients of the various membrane components. Based on previous calibrations we used diffusion coefficients of 0.012 nm2/ms for lipids (DL), 0.03 nm2/ms for water (DW) and 0.3 nm2/ms for the protein (DP) (24, 36, 50). For the interfacial DLP between the lipid and the protein, we used a value of 0.0025 nm2/ms, which was within the range reported before (24, 36, 50). The simulation yields the minimum lipid-protein or water-protein distances because once 1H magnetization crosses the interface from the soft lipid matrix (or water) to the nearest protein residues, it is rapidly equilibrated within the immobilized protein. The lack of site specificity allowed us to use the integrated intensities for the protein 13C signals to obtain global information about the depth of insertion of HNP-1. Figure 5c shows that the CH2 buildup curve for the P/L=1:18 sample was best fit to a distance of 2 Å, indicating that HNP-1 was well inserted into the hydrophobic region of the DMPC/DMPG bilayer, in immediate contact with the acyl chains. This close proximity rules out the surface-bound wedge model (Figure 1b): if the basket-shaped dimer is exposed to the aqueous phase to any significant extent, for example one third of its 18 Å height, then the membrane-immersed part of the protein would be ~12 Å, which would make the protein mostly outside the hydrophobic region of the bilayer.
Some antimicrobial peptides have been found to exhibit concentration-dependent interactions with the lipid bilayer: they bind to the surface of the lipid bilayer at low concentrations but insert into the membrane at high concentrations (51, 52). To determine whether the depth of insertion of HNP-1 changes with concentration, we measured the 1H spin diffusion spectra at P/L = 1:40. We measured two mixing times, 100 ms and 225 ms, which were sufficiently long to give equilibrium intensities for the 1:18 sample. The lipid-protein cross peaks at 225 ms were clearly higher than at 100 ms after correcting for 1H T1 relaxation, indicating that the CH2 cross peaks had not equilibrated by 225 ms. To obtain the correct normalization, we scaled the CH2 cross-peak intensities with the water cross-peak intensities in the 100 ms spectrum, since the water intensities were already equilibrated at 100 ms (Figure 5d), consistent with all membrane proteins studied so far (53–55). This ratio was further scaled by the equilibrium intensity ratio (0.61) between the CH2 peak and the water peak in the 1D 1H spectrum:
| (1) |
The normalized CH2 intensities at P/L = 1:40 are significantly lower than those of the 1:18 sample, indicating that HNP-1 inserts more shallowly at lower protein concentrations. Thus, HNP-1 has concentration-dependent insertion, similar to several other antimicrobial peptides (51, 52).
Membrane topology of HNP-1 from 13C-31P distances
To further define the membrane topology of HNP-1, we measured 13C-31P distances between the Arg residues and the lipid headgroups. The well resolved Arg Cζ signal at 157 ppm provides a site-specific probe of the interaction of the four Arg’s with the lipid phosphates. Figure 6a shows the normalized Cζ intensities, S/S0, as a function of REDOR mixing time. The intensities decayed quickly to 0.67 in the first 10 ms, then more slowly to ~0.5 by 18 ms. The bimodal decay cannot be fit to a single distance, but to a short distance of 4 Å for the initial fast regime and three longer distances of ~7 Å for the long-time regime. The weighting factor of 1 : 3 reflects the intensity plateau at 0.67 between 8 ms and 14 ms. Thus, the data suggests that one out of four guanidinium ions forms tight H-bonded complexes with the lipid phosphates, while the other three Arg’s are more distant.
Figure 6.
Arg Cζ-31P distances of HNP-1 in DMPC/DMPG bilayers. (a) Cζ 13C-31P REDOR S/S0 values as a function of mixing time and representative S0 and S spectra. Best-fit simulation used a 1:3 combination of a 4 Å distance and a 7 Å distance. (b)-(e) Solid lines represent REDOR simulations using Cζ-P distances based on the vertical distance differences among the four Arg Cζ atoms. Dashed lines represent simulations based on the Arg Cα vertical distance differences. (b) Membrane-immersed wedge model and the best-fit Arg Cζ-P REDOR curve. Simulated curves decay faster than the experimental data. (c) General pore model and the best-fit Cζ-P curves, which decay faster than the measured data. (d-e) Simulation of the Arg Cζ-P REDOR data using the dimer pore model. (d) Dimer pore model based on the crystal structure of HNP-3 and a 35 Å bilayer thickness. The simulated Cζ-P REDOR curves decay slower than the experimental data. (e) Dimer pore model after adjusting the Arg sidechain rotameric states (Table 1) and using a membrane thickness of 30 Å. The calculated Cζ-P curve fits the data very well.
We now consider the compatibility of the various membrane topological models of HNP-1 with the Arg Cζ-P distances. In the fully immersed wedge model (Figure 6b), R6 and R15 are closest to the membrane surface, with their Cζ depths differing by only ~2 Å when the axis of the basket-shaped dimer is oriented parallel to the bilayer normal (Table 1) (12). Thus, if R6 Cζ is 4 Å away from 31P, then R15 Cζ should be about 6 Å away. Using this approach, we estimated all four Cζ-P distances for the immersed wedge model (Table 1) and simulated the resulting REDOR curve. Figure 6b shows that the simulated REDOR curve decays faster than the experimental intensities, especially in the 8 – 14 ms regime.
Table 1.
Arg Cζ distances to lipid 31P (RCP, Å) in various topology models of HNPs in lipid bilayers
| Residue1 | Original HNP-3 structure 2 35 Å bilayer thickness | Arg-modified structure, 30 Å bilayer thickness | ||||
|---|---|---|---|---|---|---|
| Rotamer | RCP, wedge | RCP, general pore | RCP, dimer pore | Rotamer | RCP, dimer pore | |
| R6 | mmt180 3 mtt-85 | 4 | 7 | 15 | tpt180 | 9 |
| R15 | mtt180 | 6 | 4 | 8 | mtt85 | 6.5 |
| R16 | ttt180 | 16 | 6 | 16 | ptt85 | 10 |
| R25 | mtt-85 | 18 | 15 | 4 | ttt180 | 4 |
Each residue in principle has four distances to the membrane surfaces due to the dimer state and two bilayer surfaces. Tabulated here is the shortest distance, which is the one relevant for the 13C-31P distance experiment.
The crystal structure of HNP-3 was used to model the dimer pore (12), since the NMR structure is for the HNP-1 monomer and has lower resolution for the sidechains.
The symbols m, t, p indicate −60°, 180°, and +60°, respectively, for the consecutive χ torsion angles from the backbone (56).
Since Arg residues have multiple rotameric states with similar energetic stabilities (56) and the lipid bilayer is expected to exert a strong influence to Arg sidechain conformation through salt bridge formation (39, 57), the Arg sidechains may adopt different conformations in the membrane than in the crystal. However, any rotamer differences should shorten the Cζ-P distances and speed up the dipolar dephasing, rather than slowing down the dephasing as required to fit the experimental data. When the four Cζ-P distances were estimated based on the relative positions of the four Arg backbone Cα atoms to eliminate sidechain conformational effects, the simulated REDOR curve (dashed line) (Figure 6b) still decayed too fast compared to the observed data.
In the general pore model (Figure 6c), more than one Arg sidechain was close to the 31P atoms. Specifically, R6, R15, and R16, all have similar Cζ depths based on the crystal structure (Table 1) (12). Keeping the least inserted R15 Cζ-P distance at 4 Å, as constrained by the REDOR data, we obtained R16 and R6 Cζ-P distances of ~6 and 7 Å. The resulting REDOR curve again decays faster than the measured data, and cannot be remedied by changing the sidechain conformations.
Finally, we considered the dimer pore model, where the dimer orientation was rotated by 90° from that of the general pore model (Figure 6d). Here the situation is qualitatively different: R25 is now closest to the membrane surface, with the next nearest Arg, R15 in the other monomer, ~4 Å deeper in the membrane. Using the rotamers in the HNP-3 structure, we found the simulated curve to now decay more slowly than the measured data (Figure 6d), thus allowing us to modify the sidechain conformations to minimize the Cζ-P distances. In addition, 13CO-31P and 13Cα-31P REDOR data suggested a slight membrane thinning from 35 Å to 30 Å (Figure 7). Combining these two changes, we obtained Arg rotamers that gave Cζ-P distances of 4, 6.5, 9, and 10 Å (Table 1), which gave excellent fit to the experimental data (Figure 6e). Moreover, R25, the residue with the shortest Cζ-P distance, is also the residue showing significant immobilization upon membrane binding (Figure 3). Thus, the conformational changes also support the dimer pore topology of HNP-1.
Figure 7.
13C-31P REDOR dephasing of backbone CO and Cα of membrane-bound HNP-1. (a) CO peak. (b) Cα peak. Best-fit curves represent average two-spin distances.
Backbone CO and Cα distances to 31P are consistent with the Arg Cζ REDOR data. Both signals decay slowly, with average S/S0 values of ~0.8 at 14 ms. Simulations yielded average distances of 7.6 Å for CO and 7.2 Å for Cα. Since no site resolution was attempted, we added the REDOR dephasing of every CO and Cα atoms in each structural model based on their distances to the membrane surface. Figure S2 shows the distance distribution for the dimer pore model, where the shortest distance to 31P was fixed at 4 Å. The simulated REDOR curves fit the measured dephasing well. In reaching this agreement, it was necessary to thin the membrane to 30 Å, suggesting that HNP-1 insertion slightly reduces the bilayer thickness. Although the backbone 13C-31P REDOR did not exclude the other structural models, it was consistent with the dimer pore structure. In addition, the 13CO-31P REDOR data verified that HNP-1 was inserted into the gel-phase membrane: simulations using a surface-bound wedge model were inconsistent with the experimental result (Figure S3).
Residue-specific water-protein distances
To complement the topology information from the lipid-protein 1H spin diffusion experiment, we examined the water-protein 1H spin diffusion profile at a short mixing time of 2.2 ms, when the 1H magnetization has not equilibrated in the protein, thus giving residue-specific information about water proximities. Figure 8a shows the water-edited 13C spectrum (red) extracted from the water 1H cross section of the 2D spectrum. This water-edited spectrum was superimposed with the 13C CP spectrum that reflects the equilibrium 13C magnetization. The two spectra were scaled such that the water-edited intensities at best equaled but never exceeded the CP intensities. It can be seen that the intensity distribution of the water-edited 13C spectrum clearly differs from that of the CP spectrum, indicating selectivity of the water-protein spin diffusion at this mixing time. For example, the lipid chain signals from 33 ppm to 22 ppm were significantly attenuated, as expected because of the immiscibility of water with the hydrocarbon core. The protein signals in the 53–40 ppm region, which mainly result from Cα and Cβ, were also reduced compared to the Cα intensities in the 60–55 ppm range.
Figure 8.
Water-protein 1H spin diffusion to determine the membrane topology of HNP-1. (a) 13C spectrum after 2.25 ms 1H spin diffusion from water (red), superimposed with the 13C CP spectrum (black). Simulated stick spectrum based on the dimer pore model is shown in blue. (b) 13C CP spectrum and simulation to fix the CP weighting factors. (c) Difference between the experimental and simulated water-edited spectrum from (a). (d) Dimer pore model, where residues close to water are shown in blue, interfacial residues in gray, and residues far from water in red. The respective intensity weighting factors are 1, 0.55 and 0.1.
To simulate the water-edited 13C spectrum, we first reproduced the equilibrium CP spectrum by using appropriate weighting factors for each carbon. These weighting factors accounted for different CP dynamics of CHn (n=1, 2,3) groups and the different mobilities between the backbone and sidechains (Figure 8b). Next, we assigned each carbon a spin-diffusion weighting factor based on the proximity of each site to water on the membrane surface and water in the pore. Both the crystal structure of HNP-3 (12) and the solid-state NMR structure of HNP-1 (19) were considered in estimating the carbon-water distances, and the simulation was insensitive to the exact starting structure (Figure S4). Three categories of spin diffusion weighting factors were used. Carbons close to water were given a weighting factor of 1, while those far from water a weighting factor of 0.1. Residues in the middle were assigned a weighting factor of 0.55. Combining the water-proximity weighting factors with the CP weighting factors, we simulated the water-edited spectrum expected for the dimer pore topology. Figure 8a shows that the simulated stick spectrum agrees well with the measured spectrum. The RMSD difference between the two was 0.22, normalized to the 54-ppm peak intensity (Figure 8c). This RMSD value is comparable to the experimental RMS noise of 0.18, indicating that the measured spectrum is consistent with the dimer pore topology. We did not consider the intensities in the 33 – 22 ppm region due to the significant overlap between the lipid and protein signals.
Oligomeric structure of HNP-1 in the lipid membrane
So far we have assumed that HNP-1 assembles into a basic unit of dimers in the membrane, upon which loose higher-order oligomers may form. Although the immobilization result supports this assumption, we verified this hypothesis using the 19F CODEX experiment, which detects 19F magnetization exchange between orientationally different fluorinated molecules. The oligomeric number n is obtained from the equilibrium CODEX echo intensity of 1/n at long mixing times. Tyr4 in the β1 strand was replaced by 4-19F-Phg, since crystal structures suggest this residue to lie in the dimer interface, with inter-monomer distances of less than 10 Å (PDB codes: 1DFN, 3HJ2). Figure 9 shows the 19F CODEX control (S0) and exchange (S) spectra of HNP-1 in DMPC/DMPG membranes (P/L = 1:18) at a mixing time of 1 s. The S/S0 value was 0.66 ± 0.06 and was the same between 1 s and 500 ms within experimental uncertainty, indicating that intermolecular spin diffusion has equilibrated. The slightly larger than 0.5 final value indicates that the majority (66%) of Tyr4 exists in a dimer state (n = 2) while the rest (33%) has 19F-19F distances greater than ~15 Å. Thus, when the protein concentration is about 5 mol% in the membrane, the majority of HNP-1 is dimerized, while the rest either exists in a monomer state or forms loose dimers with Tyr4 separations greater than the detection limit of this 19F spin diffusion technique.
Figure 9.

19F CODEX control (S0) and dephased (S) spectra of 4-19F-Phg4 HNP-1 in DMPC/DMPG (3:1) membranes at P/L=1:18. The mixing time was 1 s. The S/S0 value was 0.66 ±0.06.
Discussion
Evidence of pore formation by HNPs in lipid membranes from biochemical data
Since the discovery of human α-defensins (3, 4), many antimicrobial assays, lipid vesicle experiments, and high-resolution structures have been reported to understand the mechanisms of action of these host-defense proteins (1, 58). Similar to smaller β-hairpin antimicrobial peptides, bactericidal activity by HNPs followed inner membrane permeabilization (59). Dye release and fluorescence spectroscopy experiments found that HNP-1 caused both fusion and lysis of negatively charged lipid vesicles through electrostatic interactions (60). For rabbit neutrophil defensins, vesicle leakage depends on the membrane composition: it is all-or-none for whole E. coli lipids but graded for POPG vesicles (61). Electron micrographs of human parasite cells Trypanosoma cruzi in the presence of micromolar concentrations of HNP-1 showed distinct 25-nm sized pores in the cellular and flagellar membranes (62), through which HNP-1 appears to enter the trypanosome cells, causing subsequent DNA fragmentation and cell destruction. Vesicle leakage experiments also showed that HNP-induced pores increase with the concentration of the anionic lipid (63). These biochemical studies all indicate pore formation by human α-defensins in anionic lipid membranes. However, the exact structure and topology of HNP-1 at the pore and the type of lipid disorder have remained elusive.
Structural constraints on HNP-1-induced pores in anionic lipid membranes
The 1H spin diffusion, 13C-31P distances, and 19F spin diffusion results gave the following constraints to HNP-1 structure in the anionic lipid membrane. At high protein concentrations (P/L = 1:18), HNP-1 fully spans the membrane and contacts the hydrophobic chains. Among the four Arg’s, one Arg forms hydrogen-bonded complexes with the lipid phosphates. This Arg is most likely R25 located in the β2-β3 turn, since its signal was enhanced in the 2D spectra by lipid-induced immobilization. The DMPC/DMPG bilayer is thinned slightly upon HNP-1 binding. Seen at Tyr4, the majority of the protein is at least dimerized at P/L = 1:18. These observations support the dimer pore model for HNP-1. The fully immersed wedge model (Figure 1c) and the general pore model (Figure 1e) would place R25 to be the furthest Arg from the membrane surface, which is inconsistent with the 2D 13C spectra, because the hydrocarbon core is the most fluid region of the lipid bilayer and should not immobilize R25. The depth distribution of the four Arg sidechains in the fully immersed wedge and general pore models also do not agree with the stoichiometry that only one Arg Cζ is H-bonded to the lipid phosphates. On the other hand, the surface-bound wedge model (Figure 1b) does not agree with the fast 1H spin diffusion from the lipid chains to the protein. In the multimeric pore model (Figure 1f), the Arg backbones lie at roughly the same depths in each lipid leaflet, with the two Cα rings separated by ~16 Å. Thus, the Cζ atoms would be similarly close to the membrane surface 31P, which does not agree with the distance distribution of the Cζ-P REDOR data. At lower protein concentrations (P/L=1:40), HNP-1 is less inserted into the membrane, as manifested by the slower lipid-to-protein 1H spin diffusion. Although sensitivity limitations preclude the determination of the exact topology, we expect the surface-bound wedge model to be the most likely scenario at lower protein concentrations.
In model-specific fitting of the various experimental data, we primarily used the crystal structure of HNP-3 (PDB: 1DFN) because the solid-state NMR structure (19) (PDB: 2KHT) is that of the monomer, with no direct intermolecular constraints for the dimer. Nevertheless, Figure S4 shows that the hypothetical dimer NMR structure would give similar conclusions as the HNP-3 crystal structure for the water-edited 13C spectrum, suggesting that the exact input structure does not affect the conclusion of the global topology of HNP-1.
Taken together, the solid-state NMR data shown here indicate a dimer pore topology of HNP-1, where the β-sheet dimers span the membrane with the R25-containing β2-β3 turn pointing towards the membrane surface (Figure 10). When water on the membrane surface and in the pore is both considered, the dimer pore topology reproduces the observed water-edited protein 13C spectrum. The 19F spin diffusion data verifies the dimerization of the majority of the protein. The small fraction of monomers is likely related to the observed shallower insertion of the protein at lower concentrations, and suggests that concentration-dependent oligomerization may be important for the membrane-disruptive activity of HNP-1.
Figure 10.
Membrane topology of HNP-1 in anionic lipid bilayers. At high protein concentrations HNP-1 is mostly dimerized and spans the membrane, lining a central water pore. The R25 sidechain lies closest to the membrane surface interacting with the phosphate groups. The lipid chains near the pore do not exhibit significant disorder. (a) Side view. (b) 90° rotated side view from (a), showing the pore behind one dimer. (c) Top view. The distances between the two dimers and the total oligomeric number of the assembly are not probed by the experiments here.
Does the dimer pore mechanism of HNP-1 apply to other human α-defensins? Comparisons of the activities and specificities of the six human α-defensins indicate that their interactions with lipid membranes are diverse. Against Gram-positive S. aureus, the relative potencies were HNP-2 > HNP-1 > HNP-3 > HNP-4, while the relative potencies against the Gram-negative E. coli were HNP-4 > HNP-2 > HNP-1 = HNP-3 (8). Among HNP 1, 2, and 3, whose amino acid sequences differ only in their N-terminal residue, HNP-3, which possesses an N-terminal polar residue Asp, has weaker activities than HNP-1 and HNP-2, which have a hydrophobic N-terminal residue (5, 8). Consistently, acetylation and amidation of HNP-2 to remove the terminal charges modulated the protein’s antimicrobial activity and vesicle leakage (64). Thus, the density and distribution of the positive charge have a significant effect on the membrane interaction of HNPs. In comparison, HD5 and HD6, which are found in intestinal epithelial cells, have much lower sequence homologies to HNPs (6, 7). HD6 has nearly two orders of magnitude weaker activities than the HNPs (8), while HD5, while localized on the cell membrane, was suggested to interact with the cells in a receptor-mediated fashion (65). Based on the high sequence homology and functional similarity between HNP-1 and HNP-2, we speculate that the dimer pore mechanism may apply to HNP-2 and should be relevant to HNP-3 as well, but the other defensins in this family may adopt different orientations and insertions in the lipid membrane.
Comparison of HNP-1 with small β-hairpin antimicrobial peptides
Our previous studies of the two-stranded disulfide-bonded β-hairpin antimicrobial peptide, PG-1, indicated that the strong interactions between Arg guanidinium ions and the lipid phosphate groups drove the formation of toroidal pore defects in the membrane, where PG-1 lines the pore as a TM β-barrel (24, 39, 55). When the guanidinium ions were dimethylated, the mutant showed 3-fold weaker antimicrobial activities and no longer formed large β-barrels in the membrane (66). A PG-1 mutant with only half the number of Arg residues inserted only partly into the anionic membrane and exhibited much weaker interactions with the lipid headgroups (50). Guanidinium-phosphate interactions were also observed in two Arg-rich cell-penetrating peptides, penetratin (67) and the HIV TAT peptide (68). In HNP-1, although only one guanidinium ion is within H-bonding distance to the lipid phosphates, the average distance for the other three Arg residues is 7 Å, which is short compared to the bilayer thickness (Figure 6a). Lu and coworkers have shown by mutagenesis that when three of the four Arg’s were converted to Lys, HNP-1 activity was significantly weakened, and the effect was more pronounced against the Gram-positive S. aureus than the Gram-negative E. coli (69).
It is interesting that no isotropic peak was observed in the 31P NMR spectra, indicating that high-curvature defects such as micelles or toroidal pores were absent in the HNP-bound DMPC/DMPG membrane (Figure 2). Thus, in the DMPC/DMPG membrane, the HNP-1 dimer pores exist in regular lamellar bilayers, consistent with the classical barrel-stave model. The retention of the bilayer integrity contrasts with the behavior of PG-1, which caused substantial membrane disorder to phosphocholine (PC)/phosphatidylglycerol (PG) membranes as well as phosphatidylethanolamine (PE)/PG membranes. The HNP-1 interaction with the DMPC/DMPG membrane is more akin to that of tachyplesins (70) and a synthetic antimicrobial arylamide (71), which did not disrupt PC/PG membranes. However, tachyplesin-1 caused clear disruption to the PE/PG membrane (70). The extent of membrane disorder depends on both the distribution of Arg residues in the protein sequence and the composition of the lipid membrane. It is possible that HNP-lipid interactions are sensitive to the membrane composition, similar to tachyplesins, so that while HNP-1 does not disrupt PC/PG membranes, it may disrupt PE/PG membranes. Similarly, an increased percentage of the negatively charged PG lipids may increase the amount of membrane disorder. A study of rabbit neutrophil defensins suggested that the presence of PE and cardiolipin lipids increased the membrane disruption (61). Given the increased complexity of HNPs over two-stranded β-hairpin antimicrobial peptides, structural investigations as a function of membrane composition will be useful to further elucidate the membrane interaction and mechanism of action of this class of defensins.
Finally, the precise antimicrobial mechanism of HNP-1 in vivo may depend on factors other than HNP-phospholipid interaction. It was recently reported that HNP-1 activity correlates with the amount of lipid II, a bacterial cell wall precursor (63): inhibition of lipid II synthesis weakened the HNP-1 antibacterial activity. This result suggests that interaction of cell wall components may be involved in the antimicrobial action of HNPs, and may explain why D-amino-acid analogs of HNPs appeared to have weaker bactericidal activities but similar membrane-disruptive abilities compared to their L-amino acid counterparts. Thus, the mechanism of action of HNPs, similar to a number of other defensins (72–75), may involve multiple targets in the bacteria. The dual mechanisms may also contribute to the lack of significant disorder seen in the 31P NMR spectra.
Supplementary Material
Acknowledgments
Funding information: This work was supported by the National Institutes of Health grant GM066976.
Abbreviations
- HNP-1
human neutrophil peptide 1
- DMPC
1,2-dimyristoyl-sn-glycero-3-phosphatidylcholine
- DMPG
1,2-dimyristoyl-sn-glycero-3-phosphatidylglycerol
- MAS
magic-angle spinning
- CP
cross-polarization
- DARR
dipolar-assisted rotational resonance
- DIPSHIFT
dipolar-chemical-shift correlation
- REDOR
rotational-echo double-resonance
Footnotes
Additional 2D spectra and 13C-31P REDOR analyses are included. This material is available free of charge via the Internet at http://pubs.acs.org.
References
- 1.Ganz T. Defensins: antimicrobial peptides of innate immunity. Nat Rev Immunol. 2003;3:710–720. doi: 10.1038/nri1180. [DOI] [PubMed] [Google Scholar]
- 2.Xie C, Prahl A, Ericksen B, Wu Z, Zeng P, Li X, Lu WY, Lubkowski J, Lu W. Reconstruction of the Conserved Beta-Bulge in Mammalian Defensins Using D-Amino Acids. J Biol Chem. 2005;280:32921–32929. doi: 10.1074/jbc.M503084200. [DOI] [PubMed] [Google Scholar]
- 3.Selsted ME, Harwig SS, Ganz T, Schilling JW, Lehrer RI. Primary structures of three human neutrophil defensins. J Clin Invest. 1985;76:1436–1439. doi: 10.1172/JCI112121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Gabay JE, Scott RW, Campanelli D, Griffith J, Wilde C, Marra MN, Seeger M, Nathan CF. Antibiotic proteins of human polymorphonuclear leukocytes. Proc Natl Acad Sci U S A. 1989;86:5610–5614. doi: 10.1073/pnas.86.14.5610. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Ganz T, Selsted ME, Szklarek D, Harwig SS, Daher K, Bainton DF, Lehrer RI. Defensins. Natural peptide antibiotics of human neutrophils. J Clin Invest. 1985;76:1427–1435. doi: 10.1172/JCI112120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Jones DE, Bevins CL. Defensin-6 mRNA in human Paneth cells: implications for antimicrobial peptides in host defense of the human bowel. FEBS Lett. 1993;315:187–192. doi: 10.1016/0014-5793(93)81160-2. [DOI] [PubMed] [Google Scholar]
- 7.Jones DE, Bevins CL. Paneth cells of the human small intestine express an antimicrobial peptide gene. J Biol Chem. 1992;267:23216–23225. [PubMed] [Google Scholar]
- 8.Ericksen B, Wu Z, Lu W, Lehrer RI. Antibacterial activity and specificity of the six human {alpha}-defensins. Antimicrob Agents Chemother. 2005;49:269–275. doi: 10.1128/AAC.49.1.269-275.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Wu Z, Ericksen B, Tucker K, Lubkowski J, Lu W. Synthesis and characterization of human alpha-defensins 4–6. J Pept Res. 2004;64:118–125. doi: 10.1111/j.1399-3011.2004.00179.x. [DOI] [PubMed] [Google Scholar]
- 10.Lehrer RI, Ganz T. Antimicrobial peptides in mammalian and insect host defence. Curr Opin Immunol. 1999;11:23. doi: 10.1016/s0952-7915(99)80005-3. [DOI] [PubMed] [Google Scholar]
- 11.Kagan BL, Selsted ME, Ganz T, Lehrer RI. Antimicrobial defensin peptides form voltage-dependent ion-permeable channels in planar lipid bilayer membranes. Proc Natl Acad Sci U S A. 1990;87:210–214. doi: 10.1073/pnas.87.1.210. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Hill CP, Yee J, Selsted ME, Eisenberg D. Crystal structure of defensin HNP-3, an amphiphilic dimer: mechanisms of membrane permeabilization. Science. 1991;251:1481–1485. doi: 10.1126/science.2006422. [DOI] [PubMed] [Google Scholar]
- 13.Szyk A, Wu Z, Tucker K, Yang D, Lu W, Lubkowski J. Crystal structures of human alpha-defensins HNP4, HD5, and HD6. Protein Sci. 2006;15:2749–2760. doi: 10.1110/ps.062336606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Wimley WC, Selsted ME, White SH. Interactions between human defensins and lipid bilayers: evidence for formation of multimeric pores. Protein Sci. 1994;3:1362–1373. doi: 10.1002/pro.5560030902. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Luo W, Yao XL, Hong M. Large Structure Rearrangement of Colicin Ia Channel Domain After Membrane Binding from 2D 13C Spin Diffusion NMR. J Am Chem Soc. 2005;127:6402–6408. doi: 10.1021/ja0433121. [DOI] [PubMed] [Google Scholar]
- 16.Huster D, Xiao LS, Hong M. Solid-State NMR Investigation of the dynamics of colicin Ia channel-forming domain. Biochemistry. 2001;40:7662–7674. doi: 10.1021/bi0027231. [DOI] [PubMed] [Google Scholar]
- 17.Huster D, Yao X, Jakes K, Hong M. Conformational changes of colicin Ia channel-forming domain upon membrane binding: a solid-state NMR study. Biochim Biophys Acta. 2002;1561:159–170. doi: 10.1016/s0005-2736(02)00340-1. [DOI] [PubMed] [Google Scholar]
- 18.Li S, Zhang Y, Hong M. 3D 13C-13C-13C correlation NMR for de novo distance determination of solid proteins and application to a human alpha defensin. J Magn Reson. 2010;202:203–210. doi: 10.1016/j.jmr.2009.11.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Zhang Y, Doherty T, Li J, Lu W, Barinka C, Lubkowski J, Hong M. Resonance assignment and three-dimensional structure determination of a human a-defensin, HNP-1, by solid-state NMR. J Mol Biol. 2010;397:408–422. doi: 10.1016/j.jmb.2010.01.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Luo W, Hong M. Determination of the oligomeric number and intermolecular distances of membrane protein assemblies by anisotropic 1H-driven spin diffusion NMR spectroscopy. J Am Chem Soc. 2006;128:7242–7251. doi: 10.1021/ja0603406. [DOI] [PubMed] [Google Scholar]
- 21.Luo W, Mani R, Hong M. Sidechain conformation and gating of the M2 transmembrane peptide proton channel of influenza A virus from solid-state NMR. J Phys Chem. 2007;111:10825–10832. doi: 10.1021/jp073823k. [DOI] [PubMed] [Google Scholar]
- 22.Pazgier M, Lubkowski J. Expression and purification of recombinant human alpha-defensins in Escherichia coli. Protein Expr Purif. 2006;49:1–8. doi: 10.1016/j.pep.2006.05.004. [DOI] [PubMed] [Google Scholar]
- 23.Wu Z, Powell R, Lu W. Productive folding of human neutrophil alpha-defensins in vitro without the pro-peptide. J Am Chem Soc. 2003;125:2402–2403. doi: 10.1021/ja0294257. [DOI] [PubMed] [Google Scholar]
- 24.Mani R, Cady SD, Tang M, Waring AJ, Lehrer RI, Hong M. Membrane-dependent oligomeric structure and pore formation of a b-hairpin antimicrobial peptide in lipid bilayers from solid-state NMR. Proc Natl Acad Sci USA. 2006;103:16242–16247. doi: 10.1073/pnas.0605079103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Afonin S, Glaser RW, Berditchevskaia M, Wadhwani P, Guhrs KH, Mollmann U, Perner A, Ulrich AS. 4-fluorophenylglycine as a label for 19F NMR structure analysis of membrane-associated peptides. Chembiochem. 2003;4:1151–1163. doi: 10.1002/cbic.200300568. [DOI] [PubMed] [Google Scholar]
- 26.Zou G, de Leeuw E, Lubkowski J, Lu W. Molecular determinants for the interaction of human neutrophil alpha defensin 1 with its propeptide. J Mol Biol. 2008;381:1281–1291. doi: 10.1016/j.jmb.2008.06.066. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Takegoshi K, Nakamura S, Terao T. 13C - 1H dipolar-assisted rotational resonance in magic-angle spinning NMR. Chem Phys Lett. 2001;344:631–637. [Google Scholar]
- 28.Rienstra CM, Hohwy M, Hong M, Griffin RG. 2D and 3D 15N-13C-13C NMR chemical shift correlation spectroscopy of solids: assignment of MAS spectra of peptides. Journal of the American Chemical Society. 2000;122:10979–10990. [Google Scholar]
- 29.Hong M. Resonance Assignment of 13C/15N Labeled Proteins by Two- and Three-Dimensional Magic-Angle-Spinning NMR. Journal of Biomolecular NMR. 1999;15:1–14. doi: 10.1023/a:1008334204412. [DOI] [PubMed] [Google Scholar]
- 30.Baldus M, Petkova AT, Herzfeld J, Griffin RG. Cross polarization in the tilted frame: assignment and spectral simplification in heteronuclear spin systems. Molecular Physics. 1998;95:1197–1207. [Google Scholar]
- 31.Munowitz MG, Griffin RG, Bodenhausen G, Huang TH. Two-dimensional rotational spin-echo NMR in solids: correlation of chemical shift and dipolar interactions. J Am Chem Soc. 1981;103:2529–2533. [Google Scholar]
- 32.Rhim WK, Elleman DD, Vaughan RW. Analysis of multiple-pulse NMR in solids. J Chem Phys. 1973;59:3740–3749. [Google Scholar]
- 33.Hong M, Griffin RG. Resonance Assignment for Solid Peptides by Dipolar-Mediated 13C/15N Correlation Solid-State NMR. J Am Chem Soc. 1998;120:7113–7114. [Google Scholar]
- 34.Rienstra C, Tucker-Kellogg L, Jaroniec C, Hohg M, Reif B, McMahon M, Tidor B, Lozano-Pérez T, Griffin R. De novo determination of peptide structure with solid-state magic-angle spinning NMR spectroscopy. Proc Natl Acad Sci U S A. 2002;99:10260–10265. doi: 10.1073/pnas.152346599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Mehring M. High Resolution NMR in Solids. Springer-Verlag; New York: 1983. [Google Scholar]
- 36.Huster D, Yao XL, Hong M. Membrane protein topology probed by 1H spin diffusion from lipids using solid-state NMR spectroscopy. J Am Chem Soc. 2002;124:874–883. doi: 10.1021/ja017001r. [DOI] [PubMed] [Google Scholar]
- 37.Kumashiro KK, Schmidt-Rohr K, Murphy OJ, Ouellette KL, Cramer WA, Thompson LK. A novel tool for probing membrane protein structure: solid-state NMR with proton spin diffusion and X-nucleus detection. J Am Chem Soc. 1998;120:5043–5051. [Google Scholar]
- 38.Jaroniec CP, Tounge BA, Rienstra CM, Herzfeld J, Griffin RG. Measurement of 13C-15N distances in uniformly 13C labeled biomolecules: J-decoupled REDOR. J Am Chem Soc. 1999;121:10237–10238. [Google Scholar]
- 39.Tang M, Waring AJ, Hong M. Phosphate-Mediated Arginine Insertion into Lipid Membranes and Pore Formation by a Cationic Membrane Peptide from Solid-State NMR. J Am Chem Soc. 2007;129:11438–11446. doi: 10.1021/ja072511s. [DOI] [PubMed] [Google Scholar]
- 40.deAzevedo ER, Bonagamba TJ, Hu W, Schmidt-Rohr K. Centerband-only detection of exchange: efficient analysis of dynamics in solids by NMR. J Am Chem Soc. 1999;121:8411–8412. [Google Scholar]
- 41.Buffy JJ, Waring AJ, Hong M. Determination of Peptide Oligomerization in Lipid Membranes with Magic-Angle Spinning Spin Diffusion NMR. Journal of the American Chemical Society. 2005;127:4477–4483. doi: 10.1021/ja043621r. [DOI] [PubMed] [Google Scholar]
- 42.Wang Y, Jardetzky O. Probability-based protein secondary structure identification using combined NMR chemical-shift data. Protein Sci. 2002;11:852–861. doi: 10.1110/ps.3180102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Cady SD, Goodman C, Tatko C, DeGrado WF, Hong M. Determining the orientation of uniaxially rotating membrane proteins using unoriented samples: a 2H, 13C, and 15N solid-state NMR investigation of the dynamics and orientation of a transmembrane helical bundle. J Am Chem Soc. 2007;129:5719–5729. doi: 10.1021/ja070305e. [DOI] [PubMed] [Google Scholar]
- 44.Hong M, Doherty T. Orientation determination of membrane-disruptive proteins using powder samples and rotational diffusion: a simple solid-state NMR approach. Chem Phys Lett. 2006;432:296–300. doi: 10.1016/j.cplett.2006.10.067. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Aisenbrey C, Bechinger B. Investigations of polypeptide rotational diffusion in aligned membranes by 2H and 15N solid-state NMR spectroscopy. J Am Chem Soc. 2004;126:16676–16683. doi: 10.1021/ja0468675. [DOI] [PubMed] [Google Scholar]
- 46.Lewis BA, Harbison GS, Herzfeld J, Griffin RG. NMR structural analysis of a membrane protein: bacteriorhodopsin peptide backbone orientation and motion. Biochemistry. 1985;24:4671–4679. doi: 10.1021/bi00338a029. [DOI] [PubMed] [Google Scholar]
- 47.Fares C, Qian J, Davis JH. Magic angle spinning and static oriented sample NMR studies of the relaxation in the rotating frame of membrane peptides. J Chem Phys. 2005;122:194908. [Google Scholar]
- 48.Hoover D, RK, Blumenthal R, Puri A, Oppenheim J, Chertov O, Lubkowski J. The structure of human beta-defensin-2 shows evidence of higher order oligomerization. J Biol Chem. 2000;275:32911–32918. doi: 10.1074/jbc.M006098200. [DOI] [PubMed] [Google Scholar]
- 49.Schmidt-Rohr K, Spiess HW. Multidimensional Solid-State NMR and Polymers. 1. Academic Press; San Diego: 1994. [Google Scholar]
- 50.Tang M, Waring AJ, Hong M. Effects of arginine density on the membrane-bound structure of a cationic antimicrobial peptide from solid-state NMR. Biochim Biophys Acta. 2009;1788:514–521. doi: 10.1016/j.bbamem.2008.10.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Huang HW. Action of antimicrobial peptides: two-state model. Biochemistry. 2000;39:8347–8352. doi: 10.1021/bi000946l. [DOI] [PubMed] [Google Scholar]
- 52.Glaser RW, Sachse C, Durr UH, Wadhwani P, Afonin S, Strandberg E, Ulrich AS. Concentration-Dependent Realignment of the Antimicrobial Peptide PGLa in Lipid Membranes Observed by Solid-State 19F-NMR. Biophys J. 2005;88:3392–3397. doi: 10.1529/biophysj.104.056424. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Hong M. Structure, topology, and dynamics of membrane peptides and proteins from solid-state NMR spectroscopy. J Phys Chem B. 2007;111:10340–10351. doi: 10.1021/jp073652j. [DOI] [PubMed] [Google Scholar]
- 54.Hong M. Oligomeric structure, dynamics, and orientation of membrane proteins from solid-state NMR. Structure. 2006;14:1731–1740. doi: 10.1016/j.str.2006.10.002. [DOI] [PubMed] [Google Scholar]
- 55.Tang M, Hong M. Structure and mechanism of beta-hairpin antimicrobial peptides in lipid bilayers from solid-state NMR spectroscopy. Mol Biosyst. 2009;5:317–322. doi: 10.1039/b820398a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Lovell SC, Word JM, Richardson JS, Richardson DC. The penultimate rotamer library. Proteins: Structure, Function, and Genetics. 2000;40 [PubMed] [Google Scholar]
- 57.Schug KA, Lindner W. Noncovalent binding between guanidinium and anionic groups: focus on biological- and synthetic-based arginine/guanidinium interactions with phosph[on]ate and sulf[on]ate residues. Chem Rev. 2005;105:67–114. doi: 10.1021/cr040603j. [DOI] [PubMed] [Google Scholar]
- 58.Lehrer RI, Lichtenstein AK, Ganz T. Defensins: antimicrobial and cytotoxic peptides of mammalian cells. Annu Rev Immunol. 1993;11:105–128. doi: 10.1146/annurev.iy.11.040193.000541. [DOI] [PubMed] [Google Scholar]
- 59.Lehrer RI, Barton A, Daher KA, Harwig SS, Ganz T, Selsted ME. Interaction of human defensins with Escherichia coli. Mechanism of bactericidal activity. J Clin Invest. 1989;84:553–561. doi: 10.1172/JCI114198. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Fujii G, Selsted ME, Eisenberg D. Defensins promote fusion and lysis of negatively charged membranes. Prot Sci. 1993;2:1301–1312. doi: 10.1002/pro.5560020813. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Hristova K, Selsted ME, White SH. Critical role of lipid composition in membrane permeabilization by rabbit neutrophil defensins. J Biol Chem. 1997;272:24224–24233. doi: 10.1074/jbc.272.39.24224. [DOI] [PubMed] [Google Scholar]
- 62.Madison MN, Kleshchenko YY, Nde PN, Simmons KJ, Lima MF, Villalta F. Human Defensin alpha-1 Causes Trypanosoma cruzi Membrane Pore Formation and Induces DNA Fragmentation, Which Leads to Trypanosome Destruction. Infect Immun. 2007;75:4780–4791. doi: 10.1128/IAI.00557-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.de Leeuw E, Li C, Zeng P, Li C, Diepeveen-de Buin M, Lu WY, Breukink E, Lu W. Functional interaction of human neutrophil peptide-1 with the cell wall precursor lipid II. FEBS Lett. 2010;584:1543–1548. doi: 10.1016/j.febslet.2010.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Xie C, Zeng P, Ericksen B, Wu Z, Lu WY, Lu W. Effects of the Terminal Charges in Human Neutrophil alpha-defensin 2 on its Bactericidal and Membrane Activity. Peptides. 2005;26:2377–2383. doi: 10.1016/j.peptides.2005.06.002. [DOI] [PubMed] [Google Scholar]
- 65.de Leeuw E, Rajabi M, Zou G, Pazgier M, Lu W. Selective Arginines are Important for the Antibaterial Activity and Host Cell Interaction of Human alpha-defensin 5. FEBS Letters. 2009;583:2507–2512. doi: 10.1016/j.febslet.2009.06.051. [DOI] [PubMed] [Google Scholar]
- 66.Tang M, Waring AJ, Lehrer RI, Hong M. Effects of Guanidinium-Phosphate Hydrogen Bonding on the Membrane-Bound Structure and Activity of an Arginine-Rich Membrane Peptide from Solid-State NMR. Angew Chem Int Ed Engl. 2008;47:3202–3205. doi: 10.1002/anie.200705993. [DOI] [PubMed] [Google Scholar]
- 67.Su Y, Doherty T, Waring AJ, Ruchala P, Hong M. Roles of Arginine and Lysine Residues in the Translocation of a Cell-Penetrating from 13C, 31P and 19F Solid-State NMR. Biochemistry. 2009;48:4587–4595. doi: 10.1021/bi900080d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Su Y, Waring AJ, Ruchala P, Hong M. Membrane-Bound Dynamic Structure of an Arginine-Rich Cell-Penetrating Peptide, the Protein Transduction Domain of HIV TAT, from Solid-State NMR. Biochemistry. 2010;49:6009–6020. doi: 10.1021/bi100642n. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Zou G, de Leeuw E, Li C, Pazgier M, Li C, Zeng P, Lu W, Lubkowski J, Lu W. Toward understanding the cationicity of defensins. Arg and Lys versus their noncoded analogs. J Biol Chem. 2007;282:19653–19665. doi: 10.1074/jbc.M611003200. [DOI] [PubMed] [Google Scholar]
- 70.Doherty T, Waring AJ, Hong M. Peptide-lipid interactions of the beta-hairpin antimicrobial peptide tachyplesin and its linear derivatives from solid-state NMR. Biochim Biophys Acta. 2006;1758:1285–1291. doi: 10.1016/j.bbamem.2006.03.016. [DOI] [PubMed] [Google Scholar]
- 71.Su Y, DeGrado WF, Hong M. Orientation, dynamics, and lipid interaction of an antimicrobial arylamide investigated by 19F and 31P solid-state NMR spectroscopy. J Am Chem Soc. 2010;132:9197–9205. doi: 10.1021/ja103658h. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Christ K, Wiedemann I, Bakowsky U, Sahl HG, Bendas G. The role of lipid II in membrane binding of and pore formation by nisin analyzed by two combined biosensor techniques. Biochim Biophys Acta. 2007;1768:694–704. doi: 10.1016/j.bbamem.2006.12.003. [DOI] [PubMed] [Google Scholar]
- 73.Sass V, Schneider T, Wilmes M, Körner C, Tossi A, Novikova N, Shamova O, Sahl HG. Human beta-defensin 3 inhibits cell wall biosynthesis in Staphylococci. Infect Immun. 2010;78:2793–2800. doi: 10.1128/IAI.00688-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Schmitt P, Wilmes M, Pugnière M, Aumelas A, Bachère E, Sahl HG, Schneider T, Destoumieux-Garzón D. Insight into invertebrate defensin mechanism of action: oyster defensins inhibit peptidoglycan biosynthesis by binding to lipid II. J Biol Chem. 2010;285:29208–29216. doi: 10.1074/jbc.M110.143388. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Schneider T, Kruse T, Wimmer R, Wiedemann I, Sass V, Pag U, Jansen A, Nielsen AK, Mygind PH, Raventós DS, Neve S, Ravn B, Bonvin AM, De Maria L, Andersen AS, Gammelgaard L, Sahl HG, Kristensen HH. Plectasin, a fungal defensin, targets the bacterial cell wall precursor Lipid II. Science. 2010;328:1168–1172. doi: 10.1126/science.1185723. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.









