Background: Therapeutic engineering of insulin analogs is ordinarily limited by a trade-off between pharmacokinetics and stability.
Results: Substitution of TyrB26 in a rapid-acting insulin analog by 3-iodo-TyrB26 enhances its biophysical and pharmaceutical properties.
Conclusion: An unnatural amino acid substitution circumvents insulin pharmacokinetic/stability trade-off.
Significance: Nonstandard mutagenesis can optimize the molecular properties of therapeutic proteins.
Keywords: Diabetes, Hormone, Insulin, Nuclear Magnetic Resonance (NMR), Protein Design, Protein Stability, Nonstandard Mutagenesis, Protein Allostery
Abstract
Insulin provides a model for the therapeutic application of protein engineering. A paradigm in molecular pharmacology was defined by design of rapid-acting insulin analogs for the prandial control of glycemia. Such analogs, a cornerstone of current diabetes regimens, exhibit accelerated subcutaneous absorption due to more rapid disassembly of oligomeric species relative to wild-type insulin. This strategy is limited by a molecular trade-off between accelerated disassembly and enhanced susceptibility to degradation. Here, we demonstrate that this trade-off may be circumvented by nonstandard mutagenesis. Our studies employed LysB28, ProB29-insulin (“lispro”) as a model prandial analog that is less thermodynamically stable and more susceptible to fibrillation than is wild-type insulin. We have discovered that substitution of an invariant tyrosine adjoining the engineered sites in lispro (TyrB26) by 3-iodo-Tyr (i) augments its thermodynamic stability (ΔΔGu 0.5 ±0.2 kcal/mol), (ii) delays onset of fibrillation (lag time on gentle agitation at 37 °C was prolonged by 4-fold), (iii) enhances affinity for the insulin receptor (1.5 ± 0.1-fold), and (iv) preserves biological activity in a rat model of diabetes mellitus. 1H NMR studies suggest that the bulky iodo-substituent packs within a nonpolar interchain crevice. Remarkably, the 3-iodo-TyrB26 modification stabilizes an oligomeric form of insulin pertinent to pharmaceutical formulation (the R6 zinc hexamer) but preserves rapid disassembly of the oligomeric form pertinent to subcutaneous absorption (T6 hexamer). By exploiting this allosteric switch, 3-iodo-TyrB26-lispro thus illustrates how a nonstandard amino acid substitution can mitigate the unfavorable biophysical properties of an engineered protein while retaining its advantages.
Introduction
Insulin is a small globular protein containing two chains, A (21 residues) and B (30 residues) (see Fig. 1A) (1). Although the hormone functions as a Zn2+-free monomer in the bloodstream (2), it is stored in pancreatic β-cells as a Zn2+-stabilized hexamer (3). Such self-assembly is of overarching importance to stable pharmaceutical formulation. Indeed, in the absence of self-assembly solutions of insulin would have a limited shelf life. Protective self-assembly thus delays the hormone's chemical degradation (rearrangement of atoms or chemical bonds in the molecule) and physical degradation (aggregation-coupled misfolding leading to amyloid) (4). Pharmaceutical formulations, recapitulating the storage strategy of the β-cell (3), have enabled the broad distribution, storage, and clinical use of insulin for more than 80 years in the treatment of diabetes mellitus.
FIGURE 1.
Sequence and structure of insulin. A, sequence with sites of modification. A and B chains are shown in white and gray. KP-insulin (lispro) contains an inversion of wild-type residues B28 and B29 (ProB28 → Lys, LysB28 → Pro). B, 3-I-TyrB26-KP-insulin (iodo-KP) contains an iodine substitution (gray circle) at TyrB26 (*, panel A) in the lispro context. C, Rf state monomer of KP-insulin (PDB code 1LPH) shown as a ribbon model depicting position of TyrB26. A and B chain ribbons are white and gray, respectively. D, aromatic side chains at dimer interface: TyrB16, PheB24, PheB25, and TyrB26 and their respective mates. E, structures of T6, T3Rf3, and R6 insulin hexamers. The two axial zinc ions within each hexamer are aligned at the center. For clarity, the A and B chains are shown in white and gray, with residues B1-B8 shown in black to indicate either an extended (T state) or α-helical (Rf) conformation. Bound phenols are shown in a space-filling representation within T3Rf3 and R6 hexamers. The stability of the hexamers follows the ordering T6 < T3Rf3 < R6. Structures were obtained from the Protein Data Bank (entries 1ZNJ, 1LPH, and 1TRZ).
Despite its biophysical advantages in relation to protein stability, insulin self-assembly delays its absorption from a subcutaneous (SQ)3 depot and so undermines the therapeutic efficacy of mealtime injections. Because wild-type insulin hexamers are absorbed more slowly than the smaller constituent dimers and monomers, onset of insulin action was traditionally limited by the rate of hexamer disassembly (5). To circumvent this pharmacokinetic barrier, insulin analogs were designed containing amino acid substitutions at or near the classical dimerization surfaces of the zinc insulin hexamer (6). These substitutions led to accelerated hexamer disassembly while preserving native receptor binding affinity and biological activity (7). The clinical success of such rapid-acting insulin analogs represents a pioneering triumph of rational protein design in molecular pharmacology (5).
The molecular elegance of rapid-acting insulin analogs has come at a biophysical cost; such formulations are more susceptible to physical and chemical degradation (especially on dilution and above room temperature)4 than is wild-type insulin (8, 9). Such degradation reflects a trade-off intrinsic to the logic of their design as the same amino acid substitutions that enable accelerated hexamer disassembly in the subcutaneous depot also undermine the protective effects of these assemblies against degradation in the vial or delivery device. This “molecular Catch-22” has motivated exploration of ancillary technologies to provide an improved combination of formulation stability and pharmacokinetic performance; examples include novel coatings for vials and catheters to reduce surface denaturation of insulin (10) (i.e. enhancing stability in a vial or device without delaying subcutaneous absorption) or local heating of the catheter injection site to enhance local blood flow (11) (i.e. promoting subcutaneous absorption without degrading formulation stability).
In this study we have sought to investigate whether nonstandard protein mutagenesis may be exploited to circumvent the stability-pharmacokinetic trade-off among prandial insulin analogs. Based on classical structure-function relationships as described by D. C. Hodgkin and co-workers (12–14) and Derewenda et al. (15) more than 40 years ago, we hypothesized that the structural opportunities provided by nonstandard functional groups at insulin's dimer interface might confer simultaneous improvements in stability and pharmacokinetic properties. We thus undertook to synthesize and characterize a derivative of a rapid-acting insulin analog in clinical use (LysB28, ProB29-insulin; lispro or KP-insulin, the active component of Humalog® (Eli Lilly and Co.)) (16) in which an invariant tyrosine (TyrB26) was substituted by 3-iodotyrosine (3-I-Tyr). Choice of this modification was motivated by the unique physicochemical properties of iodo-aromatic groups (17) as exploited in the evolution of thyroid hormones (18). Substitution of TyrB26 by 3-I-Tyr was previously shown to preserve the function of insulin (19, 20).
Our results demonstrate that the 3-I-TyrB26 derivative of KP-insulin exhibits a remarkable combination of biophysical properties pertinent to general principles of protein engineering. In particular, differential effects of this modification within the R6 hexamer (a traditional vehicle for pharmaceutical formulation) versus the T6 hexamer (formed in the SQ depot) promise to provide a molecular strategy to augment protein stability while preserving rapid action.
EXPERIMENTAL PROCEDURES
Preparation of Insulin Analogs
Analogs were generated by trypsin-catalyzed semi-synthesis using des-octapeptide[B23-B30]-insulin and a modified octapeptide as described (21). The insulin fragment was generated by tryptic cleavage of human insulin and purified by reverse-phase high performance liquid chromatography (HPLC), whereas the 3-I-Tyr-containing octapeptide (sequence GFF(3-I-Y)TKPT) was prepared by solid-phase synthesis (22). After trypsin-mediated formation of a peptide bond between ArgB22 and the octapeptide (a reaction favored by the reduced water activity of a mixed solvent system containing 1,4-butanediol and dimethylacetamide (23)), the resulting insulin analog was purified by preparative reverse-phase HPLC (C4 10 μm 250 × 20 mm; Higgins Analytical Inc., Proto 300). Purity was assessed by analytical HPLC (C4 5 μm, 250 × 4.6 mm; Higgins Analytical Inc., Proto 300). The predicted molecular mass of the insulin analog was verified using matrix-assisted laser desorption/ionization time-of-flight mass spectrometry using an Applied Biosystems 4700 proteomics analyzer (determined mass, 5933 Da; predicted mass, 5933 Da).
Circular Dichroism
Far-ultraviolet (UV) spectra were obtained on an AVIV spectropolarimeter equipped with an automated syringe-driven titration unit. Insulin or insulin analogs were made 50 μm in 10 mm potassium phosphate (pH 7.4) and 50 mm KCl. Spectra were obtained from 200–250 nm as described (24). Thermodynamic stabilities were probed by guanidine hydrochloride-induced denaturation monitored by CD at helix sensitive wavelength 222 nm. Data were fit by nonlinear least squares to a two-state model (25),
![]() |
where x is the concentration of guanidine hydrochloride and θA and θB represent respective estimates of the base-line ellipticities of the protein in its native and unfolded states as extrapolated to a guanidine concentration of 0. Base-line values were approximated via pre- and post-transition lines represented by equations θA(x) = θAH2O + mAx and θB(x) = θBH2O + mBx. Such simultaneous fitting avoids artifacts of linear plots of ΔG versus concentration of denaturant (26).
Size-exclusion Chromatography and Multiangle Light Scattering (SEC-MALS)
Self-association was monitored by SEC-MALS as described (27). Insulin analogs were made 0.6 mm in a buffer consisting of 25 mm Tris-HCl (pH 7.4), 0.65 mg/ml phenol, 1.6 mg/ml meta-cresol, 16 mg/ml glycerol, and ZnCl2 at a ratio of 2 zinc ions per insulin hexamer. A protein solution (volume 100 μl) was injected onto a Superdex75 column (nominal fractionation range of 3–70 kDa) at a flow rate of 0.5 ml/min. The mobile phase consisted of 10 mm Tris-HCl (pH 7.4), 140 mm NaCl, 0.02% sodium azide with or without 50 mm cyclohexanol (employed as a replacement for phenolic compounds5 often contained in pharmaceutical formulations). Light scattering was monitored with a Wyatt Technology miniDAWN TREOS® 3-angle detector. Protein concentrations were determined using an inline Shimadzu SPD-6AV UV-visible detector. Data were analyzed using the ASTRA® software obtained from Wyatt Technology (Santa Barbara, CA).
Receptor Binding Assays
Affinities of insulin or insulin analogs to the B isoform of the insulin receptor were measured in a competitive-displacement assay. In brief, microtiter strip plates (Nunc Maxisorb) were incubated at 4 °C overnight with a stock solution (100 μl/well) consisting of 40 μg/ml anti-FLAG immunoglobulin G. Detergent-solubilized lysates of 293 PEAK cells transfected with cDNAs encoding the insulin receptor with C-terminal FLAG tags were purified using wheat germ agglutinin chromatography (28). Partially purified receptors were then immobilized in the coated plates. Plates were extensively washed, and competitive binding assays using labeled tracer 125I-TyrA14-insulin and unlabeled insulin analogs were performed as described (29). Data for homologous and heterologous receptor binding were analyzed as described (30).
Assessment of Fibril Formation
Insulin fibrillation was monitored under two conditions, (i) at 60 μm insulin concentration in phosphate-buffered saline (pH 7.4) containing 0.1% sodium azide with gentle rocking at 37 °C and (ii) at 0.15 mm concentration in a buffer consisting of 7 mm Tris-HCl (pH 7.4), 75 μm ZnCl2, 16 mg/ml glycerin, 1.6 mg/ml meta-cresol, 0.65 mg/ml phenol with gentle rocking at 45 °C. In each case the protein solutions were incubated in glass vials containing a liquid/air interface. Aliquots were taken at regular intervals and frozen to enable analysis of thioflavin T fluorescence at the end of the assay, terminated on visual appearance of cloudiness in the solution (31).
Visible Absorption Spectroscopy
Formation and disassembly of the phenol-stabilized R6 Co2+-substituted insulin hexamer were probed by visible absorption spectroscopy (450–700 nm). Proteins were made 0.6 mm in a buffer consisting of 50 mm Tris-HCl (pH 7.4), 50 mm phenol, 0.2 mm CoCl2, and 1 mm NaSCN (32). Before the studies, samples were incubated overnight at room temperature. Spectra were then acquired to monitor tetrahedral cobalt coordination with its signature d-d absorption band at 574 nm (32). Kinetics studies to determine the rate of Co2+ release from the hexamers exploited metal-ion sequestration by a 10-fold molar excess of EDTA as described (33). The assay was initiated at 25 °C by the addition of an aliquot of EDTA (made 50 mm at pH 7.4) to a final concentration of 2 mm. The intensity of the 574-nm absorption band was monitored on a time scale of seconds-hours. Kinetics data were consistent with a mono-exponential decay due to attenuation of this band in an octahedral Co2+-EDTA complex.
NMR Spectroscopy
Spectra were acquired in aqueous D2O solution (pD 7.7, direct meter reading) at 700 MHz at 32 °C; protein concentrations were ∼0.5 mm. Homonuclear two-dimensional nuclear Overhauser effect (NOE) spectroscopy (NOESY) and total correlation spectroscopy (TOCSY) spectra were obtained with mixing times 100 and 52 ms as described (34). 1H NMR chemical shifts were calibrated in parts per million (ppm) relative to trimethylsilyl propionate as an internal standard, assumed to be a 0.00 ppm.
Kinetics of Zinc Extraction
Insulin or insulin analogs were made 0.17 mm in a solution containing 13.2 mm Tris-HCl (pH 7.4) and 30 mm phenol with 1.2 zinc ions per hexamer. A probe of the free concentration of Zn2+ was provided by the indicator 2,2″,6:2 terpyridine (35); the reagent was obtained from Sigma. Release of Zn2+ from insulin hexamers was thus monitored by UV absorbance at 334 nm upon the addition of a final dye concentration of 500 μm (in marked excess relative to total Zn2+ concentration). Data were fit to a mono-exponential equation.
Rodent Assay
Male Lewis rats (mean body mass ∼300 grams) were rendered diabetic by treatment with streptozotocin (36). To test in vivo potency, protein solutions containing KP-insulin or analog were constituted in a buffer containing 3.8 mg/ml sodium phosphate (pH 7.4), 8.6 μm ZnCl2 (3 Zn2+:1 insulin hexamer) with the addition of 16 mg/ml glycerin, 1.6 mg/ml meta-cresol, and 0.65 mg/ml phenol. To ensure uniformity of the formulations, the insulin analogs were each repurified by reverse-phase HPLC, dried to powder, dissolved in the above buffer in the absence of zinc ions at the same maximum protein concentration, and re-quantitated by analytical C4 reverse-phase HPLC; an aliquot of 100 mm ZnCl2 (pH 4 in 0.1 mm HCl) was added to obtain the targeted Zn2+:insulin ratio.
Rats were injected SQ at time t = 0 with 10 μg or 20 μg of KP-insulin or analog in 100 μl of buffer per 300-g mass of rat with the dose being adjusted to each rat's body mass. Dose-response studies of KP-insulin indicated that at the 20-μg dose, a near-maximal rate of glucose disposal during the first hour after injection was achieved with an EC50 of ∼5 μg. Blood was obtained from the clipped tip of the tail at time 0 and every 10 min up to 60 min for the 1st h, every 20 min for the 2nd h, every 30 min for the 3rd h and every hour thereafter. Serial measurements were made using a clinical glucometer (Hypoguard Advance Micro-Draw meter). The efficacy of insulin action was calculated using (a) the change in blood glucose concentration over the first hour and (b) the integrated area between the glucose time dependence and a horizontal line at the starting blood glucose concentration (area over the curve (AOC)). The AOC was calculated for each individual rat (and not based on the mean curve) to avoid artifacts due to rat-to-rat variation in initial values of the blood-glucose concentration; in our rat colony such base-line values exhibited considerable variation, even in the same rat on different days. This variation can have a marked effect on the averaged data obtained at each time point. Assessment of statistical significance was performed using Student's t test. Intravenous injections into tail veins with the same insulin analogs were carried out at a dose of 10 μg per 100 μl of buffer per 300-g mass of rat; data were analyzed as above.
RESULTS
Synthesis of 3-Iodo-TyrB26-LysB28, ProB29-insulin
Trypsin-catalyzed semi-synthesis enabled efficient preparation of 3-I-TyrB26-substituted KP-insulin (Fig. 1, A and B). The KP substitutions (ProB28 → Lys and LysB29 → Pro) provided both a monomeric template (in the absence of Zn2+; Fig. 1C) and a hexameric template wherein TyrB26 packs at an aromatic-rich dimer interface (in the presence of Zn2+ and phenol, meta-cresol, or cyclohexanol; Fig. 1, D and E) (9). The availability of these alternative templates facilitated comparative studies of biophysical properties pertinent to formulation and biological activity, including the TR allosteric transition6 (Fig. 1E).
Receptor Binding and Activity
The affinity of 3-I-TyrB26-KP-insulin for the lectin-purified insulin receptor (isoform B) is almost 2-fold greater than that of KP-insulin or wild-type insulin (Table 1) in accordance with past studies (20, 37); dissociation constants (Kd) are summarized in Table 1. The potency and duration of action of 3-I-TyrB26-KP-insulin in a rat model of diabetes mellitus were nonetheless similar (but not identical) to those of KP-insulin on SQ injection at 10-μg (intermediate) and 20-μg (near-maximal) doses (Fig. 2). Because of variation in the initial blood glucose level between rats, data are shown both in relation to the actual blood glucose concentrations at time 0 (Fig. 2A,C) and after normalization with respect to their initial values (defined as 1.0; Fig. 2, B and D). Initial rates of fall of the blood glucose concentration were not statistically significant (Fig. 3A,C).7 Although 3-I-TyrB26-KP-insulin appeared to be less potent than KP-insulin when mean values were plotted, quantitative analysis of total areas over the curve (AOC) demonstrated that any differences were not of statistical significance (Fig. 3, B and D). We cannot exclude that in a larger study (i.e. with a greater number of animals per group)8 statistically significant differences might be observed as higher hormone-receptor affinity can lead to more rapid clearance from the circulation (5).
TABLE 1.
Properties of insulin analogs
| Protein | ΔGua | Cmid | m | Fibrillation lag timeb |
IR binding | |
|---|---|---|---|---|---|---|
| 45 °Cc | 37 °Cd | |||||
| kcal/mol | m | kcal/mol/m | days | nm | ||
| Wild type (insulin) | 3.4 ± 0.1 | 4.8 ± 0.1 | 0.69 ± 0.02 | 22.3 ± 9.1 | 3.8 ± 1.5 | 0.031 ± 0.004 |
| KP-insuline | 3.0 ± 0.1 | 4.6 ± 0.1 | 0.65 ± 0.02 | 13.6 ± 4.5 | 3.7 ± 2.0 | 0.032 ± 0.005 |
| Iodo-KP-insuline | 3.5 ± 0.1 | 5.1 ± 0.1 | 0.68 ± 0.02 | 21.5 ± 7.4 | 15.7 ± 7.3 | 0.022 ± 0.003 |
a Thermodynamic parameters were inferred from CD-detected guanidine denaturation data by application of a two-state model. Cmid is defined as the molar concentration of guanidine-HCl indicative of 50% protein unfolding. The m value is the slope of a plot of the unfolding free energy versus the molar concentration of guanidine-HCl. This slope is proportional to the protein surface area exposed on unfolding.
b A 2-fold increase over base line in thioflavin T fluorescence provided a criterion for onset of fibrillation.
c Fibrillation lag times pertain to R6 zinc hexamers in formulation-like conditions (0.15 mm protein, 7 mm Tris-HCl (pH 7.4), 16 mg/ml glycerin, 1.6 mg/ml meta-cresol, 0.6 mg/ml phenol, and 0.075 μm ZnCl2). For statistical analysis one outlier in the KP trial was excluded due to it being >2 S.D. greater than the mean.
d Fibrillation lag times pertain to zinc-free wild-type insulin (in a monomer-dimer equilibrium) and analogs (monomeric); each protein was made 60 μm in phosphate-buffered saline (pH 7.4).
e Analogs KP-insulin and iodo-KP-insulin, respectively, designated LysB28, ProB29-insulin and 3-iodo-TyrB26-LysB28, ProB29-insulin.
FIGURE 2.
Fall in blood glucose concentration after SQ injection of KP-insulin (●) or 3-I-TyrB26-KP-insulin (iodo-KP) (○) at two doses, 10 μg/100 μl per 300-g mass of rat (A) and 20 μg/100 μl per 300-g mass of rat (C). Data are represented as the mean of each group (KP, n = 10 for both doses; iodo-KP, n = 10 for the 10-μg dose and n = 11 for the 20-μg dose); error bars represent S.E. B and D, normalized blood glucose profiles for the data in panels A and C, respectively.
FIGURE 3.
A and C, initial rates of blood glucose clearance over the first hour for each of the doses tested. Box plots represent upper and lower quartiles; the horizontal bars delineate the median (black) and mean (green), and the vertical bars represent minimum and maximum values. No statistically significant difference in initial rates of fall was observed between analogs at either dose (p > 0.05). Potencies of insulin analogs at two doses, 10 μg/100 μl per 300-g mass of rat (B) and 20 μg/100 μl per 300-g mass of rat (D) as measured by the integrated AOC from data in Fig. 2. No statistically significant difference was observed between analogs at either dose (p > 0.05).
Because on SQ injection apparent biological activities of insulin analogs can reflect differences in bio-availability and rates of absorption, their intravenous bolus injection permitted assessment of intrinsic activity. Tail-vein injection of KP-insulin and 3-I-TyrB26-KP-insulin (at a dose of 10 μg/300 g rat in a volume of 100 μl; n = 5 animals per group) revealed similar patterns in the resulting decrease and recovery of the blood glucose concentration with identical rates of fall in blood glucose during the first hour after injection (Fig. 4A). Any differences in AOC were also not of statistical significance (Fig. 4B).
FIGURE 4.

Rat intravenous injection assay. A, KP-insulin (●, KP) or 3-I-TyrB26-KP-insulin (○, iodo-KP) in a standard formulation buffer was injected intravenously at a dose of 10 μg/100 μl per 300-g mass of rat. Blood glucose measurements were obtained at the indicated times (n = 5). B, area over the curves in panel A. Data are represented as box plots with rectangles delineating quartiles, and the horizontal bars indicate the median (black) and mean (green) values. Vertical bars represent minimum and maximum values. No statistically significant differences in potency were found as determined by AOC (p > 0.05).
Structure and Stability
Structural effects of the modification were probed by far-UV CD. The similarity of the spectra of KP-insulin and 3-I-TyrB26-KP-insulin in the absence of zinc ions (respective black and red dashed lines in Fig. 5A) suggests that the modification does not grossly perturb secondary structure. The non-negligible differences between these spectra and that of wild-type insulin (green line in Fig. 5A) reflect dimerization of wild-type insulin (which accentuates α-helical CD features; Ref. 38) and non-local effects of the KP substitutions on the structure of the monomer (39).
FIGURE 5.
Biophysical assays of structure and stability. A, far-UV circular dichroism spectra of human insulin (HI, green line), KP-insulin (KP, black line) and 3-I-TyrB26-KP-insulin (iodo-KP, red dashed line) at pH 7.4 and 25 °C. Ellipticity was normalized per residue. B, guanidine (Gu)-unfolding transitions as monitored by ellipticity at 222 nm. Labeling is as in panel A. Thermodynamic stabilities were derived using a two-state model (Table 1). C, one-dimensional spectra in D2O: aromatic and downfield Hα resonances of insulin and insulin analogs. Top, wild-type insulin; middle, KP-insulin; bottom, 3-I-TyrB26-KP-insulin. Asterisk in the bottom trace indicates the downfield C2H ring resonance of 3-I-TyrB26. Brackets indicate groupings of histidine residues (left) and α protons (right). Selected resonances assignments are as labeled: a, HisB10 C2H; b, HisB5 C2H; c, PheB24 C4H; d, PheB24 C2,6H; e, PheB24 CαH; f, CysA20 CαH; g, CysA6 CαH; h, CysB7 CαH; i, CysA7 CαH. Line broadening and minor peaks in the wild-type spectrum reflect coupled self-association equilibria among dimers, trimers, tetramers, hexamers, and higher-order oligomers. D, two-dimensional spectra of KP-insulin (black) and 3-I-TyrB26-KP-insulin (red) in D2O. Bottom, overlaid TOCSY spectra of aromatic spin systems; top, are overlaid NOESY spectra in regions containing cross-peaks between aromatic protons (horizontal axis; ω2) and aliphatic protons, including upfield-shifted methyl groups (vertical axis; ω1). The asterisk indicates the downfield C2H ring resonance of 3-I-TyrB26. Brackets in the top panel denote aromatic and methyl group resonances. Long range NOE assignments: j, 3-I-TyrB26 C2H/δ1-CH3 LeuB15; k, TyrA19 C2,6H/δ1-CH3 LeuB15; l, TyrB16 C2,6H/δ1-CH3 LeuB15; m, TyrB26 C2,6H/δ1-CH3 LeuB15; n, PheB24 C4H/δ1-CH3 LeuB15; o, PheB24 C3,5H/δ1-CH3 LeuB15; p, PheB24 C2,6H/δ1-CH3 LeuB15; q, TyrA19 C2,6H/δ1-CH3 IleA2; r, TyrA19 C2,6H/δ2-CH3 LeuB15; s, TyrA19 C2,6H/γ2-CH3 IleA2; t, 3-I-TyrB26 C2H/γ2-CH3 IleA2; u, TyrB26 C2,6H/γ2-CH3 IleA2; v, 3-I-TyrB26 C2H/γ1-CH3 ValB12; x, TyrB26 C2,6H/γ1-CH3 ValB12. The red arrow indicates the absence of long-range NOE cross-peaks from the C6H proton of 3-I-TyrB26 and aliphatic protons in this spectra region (red dashed line). 1H NMR spectra in panels C and D were acquired at 700 MHz and 32 °C; the proteins were made zinc-free at a concentration of ∼0.5 mm in 99.9% D2O at pD 7.7 (direct meter reading). Respective TOCSY and NOESY mixing times were 52 and 100 ms.
Despite the similar far-UV spectra of 3-I-TyrB26-KP-insulin and KP-insulin, CD-monitored guanidine denaturation studies (Fig. 5B) indicated significant differences in thermodynamic stability. Estimates of free energies of unfolding based on a two-state unfolding model indicated that the stability of 3-I-TyrB26-KP-insulin (ΔGu 3.5 ± 0.1 kcal/mol at 25 °C) was greater than that of KP-insulin (ΔΔGu 0.5 ± 0.2 kcal/mol) and similar to that of wild-type insulin (ΔGu 3.4 ± 0.1 kcal/mol; Table 1). Iodination of TyrB26, therefore, compensates for the loss of stability associated with the KP substitutions. Such stabilization is not due to restored dimerization, as 1H NMR spectra of KP-insulin and 3-I-TyrB26-KP-insulin at higher concentrations (∼500 μm) exhibit similar resonance line widths and secondary chemical shifts (lower traces in Fig. 5C) (40), in each case lacking 1H NMR features of self-assembly evident in the spectrum of wild-type insulin (top trace in Fig. 5C) (41). An 1H NMR signature of an insulin monomer is the conformational broadening of the C2H resonance of HisB5 (within a crevice between the A and B chains) relative to the narrow C2H resonance of the solvent-exposed and flexible side chain of HisB10 (peaks a and b in Fig. 5C) (42). The characteristic downfield aromatic singlet resonance of 3-I-Tyr (due to C2H) is well resolved in the spectrum of 3-I-TyrB26-KP-insulin in D2O (asterisk in the bottom trace in Fig. 5C) (42).
An overall similarity between KP-insulin and 3-I-TyrB26-KP-insulin was observed with respect to the overall pattern of chemical shifts of corresponding aromatic resonances and downfield Hα resonances (assigned to CysA6, CysA7, CysA20, and CysB7). This correspondence was further delineated in two-dimensional TOCSY spectra as illustrated in the respective aromatic spin systems of KP-insulin (black cross-peaks in the bottom panel of Fig. 5D) and 3-I-TyrB26-KP-insulin (red cross-peaks). Whereas, as expected, no TOCSY cross-peaks involve the isolated C2H proton of 3-I-TyrB26 (asterisk in Fig. 5D), J coupling was observed between the C5H and C6H protons opposite to the site of modification. A corresponding set of long range NOEs was observed from these aromatic protons to aliphatic resonances and most prominently involved upfield-shifted methyl groups (Fig. 5D, top). NOESY spectra were obtained with a mixing time sufficiently short (100 ms) such that at this temperature (32 °C) spin diffusion was negligible, thus enabling direct contacts to be observed between specific aromatic ring protons and individual methyl groups.
The most up-field aliphatic 1H NMR resonances (δ1-CH3 resonance at 0.15 ppm in KP-insulin and 0.10 ppm in 3-I-TyrB26-KP-insulin) were assigned to the side chain of LeuB15, which packs beneath the aromatic ring of PheB24 and so experiences a ring-current shift (43). Of particular interest are native-like contacts between the side chains of PheB24 and LeuB15 (cross-peaks n, o, and p in Fig. 5D), TyrA19 and IleA2 (cross-peaks q and s). Maintenance of these framework interresidue contacts and ring-current effects provides evidence that iodination of TyrB26 does not perturb the overall tertiary structure of the protein. The specific NOEs involving the individual aromatic protons of 3-I-TyrB26 indicate that the modified ring participates in an organized substructure involving the same set of aliphatic side chains as are near TyrB26 in the solution structure of KP-insulin (40) and in crystal structures of a wild-type insulin protomer (1). A striking asymmetry was observed between NOEs involving the respective meta protons of the modified B26 aromatic ring, whereas C2H (lying on the same side of the ring as the iodo-substituent) contacts one methyl group from ValB12 and one from LeuB15 (respective cross-peaks v and j in Fig. 5D, reflecting native-like B-chain super-secondary structure), and the γ2-CH3 group of IleA2 (cross-peak t in Fig. 5D, reflecting insertion of the ring into an interchain crevice) in the case of C6H (opposite side of the ring) such NOEs are markedly attenuated or absent (red arrow and dashed line in Fig. 5D). Such asymmetry implies that the iodo-substituent occupies a favored binding pocket in (or adjoining) the hydrophobic core; this environment hinders rotation of the ring about the Cβ-Cγ bond axis. These qualitative features do not exclude local reorganization of the B26-related substructure, determining the details of this structure will require more complete 1H NMR analysis and molecular modeling.
Insulin Self-assembly
Evidence that 3-I-TyrB26-KP-insulin and KP-insulin are each competent to form R6 phenol-stabilized Co2+-insulin hexamers9 was provided by observation of a characteristic d-d visible absorption band in each case. This blue band, which provides an optical signature of a tetrahedral binding site (Fig. 6A), is prominently observed in the spectrum of 3-I-TyrB26-KP-insulin (open circles in Fig. 6B); its retention suggests that the bulky iodo-substituent is readily accommodated at or near the classical dimer interface (1). The slight attenuation of this feature in the spectrum of the KP-insulin hexamer (thin black line in Fig. 6B; 12 ± 1% lower at 574 nm in triplicate measurements) in relation to the wild-type hexamer (thick black line) has previously been described (9). The visible absorption spectrum of 3-I-TyrB26-KP-insulin as a metal-ion-stabilized complex (open circles in Fig. 6B) more closely resembles the spectrum of the KP-insulin R6 hexamer than it does the spectrum of the wild-type insulin hexamer, a pattern similar to that seen in CD studies of the metal ion-free proteins (above).
FIGURE 6.
Optical absorption spectra of cobalt-stabilized hexamers and kinetics of metal-ion release from insulin hexamers in the presence of phenol. A, in the presence of phenolic excipients or high salt concentrations, one or both of the coordinated metal ions within the core of the insulin hexamer shifts from an octahedral to a tetrahedral coordination to form either T3Rf3 or R6 type hexamers. The tetrahedral ligand field around the metal ion is composed of three HisB10 residues and is completed by an anion. B, Co2+ d-d bands of wild-type human insulin (wt), KP-insulin (KP), and 3-I-TyrB26-KP-insulin (I-KP) near 550 nm provide a signature of the R (or Rf) state. The amplitude of iodo-KP (○) resembles that of KP-insulin (line), in each case attenuated by 12% relative to wild-type insulin (bold line). C, sequestration of cobalt ions from insulin analogs by EDTA: wild-type insulin (▴), KP-insulin (●), or 3-I-TyrB26-KP-insulin (○). Half-lives are given in Table 2. abs, absorbance. D, corresponding sequestration of zinc ions by the addition of 2,2′:6′,2″-terpyridine. Symbols are as in panel A. Formation of terpyridine-Zn2+ complexes was monitored at 334 nm.
To extend these studies from cobalt hexamers to zinc hexamers (as in a pharmaceutical formulation), we employed SEC-MALS, a technique that combines size-exclusion chromatography with multiangle static light scattering. Because phenol and meta-cresol (classical pharmaceutical excipients and allosteric effectors of the TR transition6 present in the samples) interfere with UV detection of the eluting proteins, the running buffer contained the alternative UV-transparent ligand cyclohexanol10 (44) (left-hand panels of Fig. 7). Although there were no zinc ions in the running buffer, past studies have indicated that its release from hexameric complexes is dependent on their disassembly (45, 46).
FIGURE 7.
SEC-MALS. Elution profiles of wild-type human insulin (HI) (A and D), KP-insulin (KP) (B and E), and 3-I-TyrB26-KP-insulin (iodo-KP) (C and F) were obtained on fractionation on a Superdex 75 in the presence (left) or absence (right) of 50 mm cyclohexanol as an allosteric ligand. Molecular masses across the peaks (dots) were calculated from the combination of UV detection (solid lines) and multiangle light scattering. Peaks corresponding to hexamers and monomer-dimer equilibria are denoted a and b, respectively. Iodo-KP elutes primarily as a hexamer (∼35 kDa) in the presence of cyclohexanol (C) and as a monomer in absence of cyclohexanol (F). The minor peak (denoted by an arrow) of the iodo-KP chromatogram in the absence of cyclohexanol represents a monomer-dimer equilibrium (∼7 kDa). The elution profile of iodo-KP is similar to that of KP-insulin in the absence of cyclohexanol, whereas it is similar to wild-type insulin in its presence.
Marked differences were observed in this assay between wild-type insulin, KP-insulin, and 3-I-TyrB26-KP-insulin. Whereas wild-type insulin eluted predominantly as a hexamer (peak a in Fig. 7A) with a minor component due to an equilibrium between monomers and dimers (intermediate peak b), KP-insulin eluted predominantly as a broad distribution of species representing progressive disassembly of hexamers into smaller species; the late-eluting peak represents an equilibrium between dimers and monomers (peak b in Fig. 7B). 3-I-TyrB26-KP-insulin exhibited an elution profile intermediate between the latter two patterns. Its greater fraction of intact hexamers (relative to KP-insulin; peak a in Fig. 7C) suggests that the iodo-modification either (or both) enhances the stability of the variant R6 hexamer or (and) introduces a kinetic barrier to its disassembly. Such stabilization or kinetic barrier would be expected to forestall degradation of 3-I-TyrB26-KP-insulin hexamers in a vial or delivery device (47).
SEC-MALS may readily be adapted to probe the transition of a concentrated R6-based insulin formulation (as stabilized by Zn2+ coordination and an allosteric phenolic ligand) to lower molecular weight species as a model of these molecular events in the SQ depot. To this end we sought to mimic the rapid loss of phenol and meta-cresol from the loaded R6 sample (which occurs on the millisecond time scale as measured by 1H NMR (43)) on diffusion of these ligands into cell membranes. This step is followed by slow hexamer disassembly with release of zinc ions and absorption of insulin molecules into capillaries. The initial steps of this process may be modeled in the SEC-MALS assay by use of a running buffer that contains neither zinc ions nor cyclohexanol (27). Under these conditions wild-type insulin eluted with a peak mass of ∼13 kDa, corresponding to a dimer-trimer equilibrium; a trailing edge was also observed representing a dimer-monomer equilibrium (Fig. 7D). KP-insulin was observed in this assay to elute primarily as a single peak with a monomeric mass of ∼5.8 kDa (Fig. 7E). This chromatographic behavior is in accordance with the design of Humalog® as a rapid-acting insulin analog formulation (16).
Despite the stabilizing effect of 3-I-TyrB26 in the first set of SEC-MALS studies (above), the elution profile of 3-I-TyrB26-KP-insulin (Fig. 7F) in the second chromatographic assay resembles that of KP-insulin (Fig. 7E). A slight initial shoulder in its elution profile (arrow in Fig. 7F) corresponds to a trace dimer component not detectable in KP-insulin. We note in passing that 3-I-TyrB26-KP-insulin also exhibited a small delay in its peak elution time (relative to KP-insulin), which is a likely consequence of an increased interaction with the column matrix due to the added hydrophobicity of 3-I-Tyr. These findings suggest that under formulation conditions thermodynamic and kinetic features relevant to the rapid action of KP-insulin on SQ injection would be retained in 3-I-TyrB26-KP-insulin.
Kinetics of Disassembly
Because SEC-MALS suggested kinetic stabilization of the cyclohexanol-induced R6 hexamer by the iodo-modification, we returned to the cobalt assay as an explicit probe for the lifetime of the tetrahedral metal binding site (Fig. 6A). The d-d absorption band of the wild-type and variant insulin R6 hexamers (stabilized by 50 mm phenol) was, therefore, exploited as a kinetic probe (Fig. 6B). At time t = 0 s, an aliquot of a concentrated EDTA (adjusted to pH 7.4) was added that enabled progressive sequestration of Co2+ ions on their release from protein complexes in the course of a self-association equilibrium (Fig. 6C). The rate of release is slow (minutes to hours) relative to rapid EDTA chelation (whose rate is diffusion limited). Hexamer lifetimes may thus be measured via the progressive disappearance of the d-d optical transitions characteristics of the tetrahedral Co2+-coordination site in R6 hexamers.11 Fitting of the time-dependent attenuation of the 574-nm Co2+ band intensity indicated that the respective lifetimes of R6 assemblies containing wild-type insulin and KP-insulin are 13.8 ± 0.4 and 3.1 ± 0.1 min at 25 °C (Table 2, Fig. 6C). Remarkably, the lifetime of R6 hexamers containing 3-I-TyrB26-KP-insulin is 25 ± 1 min under these conditions is even slower than that of wild-type insulin.
TABLE 2.
Metal ion sequestration studies
| Protein | Co2+ extractiona half-life | Zn2+ extractionb half-life |
|---|---|---|
| min | min | |
| Wild-type insulin | 13.8 ± 0.4 | 3.6 ± 0.1 |
| KP-insulin | 3.1 ± 0.1 | 1.0 ± 0.1 |
| Iodo-KP-insulin | 25 ± 0.8 | 8.9 ± 0.1 |
a Kinetic features of Co2+ release from protein hexamers were monitored by following the attenuation in absorbance at 574 nm after the addition of excess EDTA. Protein solutions were made 0.6 mm (nominal concentration of monomers) in 50 mm Tris-HCl (pH 7.4), 50 mm phenol, 1 mm NaSCN, and 0.2 mm CoCl2.
b Sequestration of Zn2+ was monitored by following the change in absorbance at 334 nm after the addition of excess 2,2′:6′,2″-terpyridine to a solution wherein the proteins were made 0.17 mm (nominal concentration of monomers) in a buffer containing 13.2 mm Tris-HCl (pH 7.4) and 30 mm phenol with 1.2 zinc ions per hexamer.
To extend these kinetic studies to Zn2+-containing R6 hexamers, we employed a molecular probe whose molar absorbance at 334 nm is markedly enhanced on binding Zn2+ ions (35). The same trend was observed among the three proteins as obtained in the Co2+-based EDTA sequestration assay (Table 2, Fig. 6D). Thus, accommodation of the bulky iodo-aromatic substituent within the dimer interface of the R6 hexamer imposes a kinetic barrier to its disassembly. The added barrier height (ΔΔEa) is estimated as RTln(k′/k); relative to the rate of disassembly of KP-insulin, this estimate yields a value of 1.2 ± 0.1 kcal/mole at 25 °C.
Insulin Fibrillation
Clinical formulations of insulin and insulin analogs are ordinarily protected from fibrillation through their self-assembly as R6 zinc hexamers. The above findings, therefore, motivated us to test whether the iodo-Tyr modification might augment such protection. Protein fibrillation is characterized by a lag phase of stochastic onset (before detectable increases in thioflavin T fluorescence or optical scattering) and an elongation phase (48). Although variability in the lag time is characteristic of nucleation-propagation reactions (49), a significant difference was observed between the susceptibility of wild-type insulin to fibrillation under diluted formulation conditions12 (22 ± 9 days in n = 9 trials; as assessed at a protein concentration of 0.15 mm on gentle agitation at 45 °C in the presence of an air-liquid interface) and that of the less stable KP-insulin analog (14 ± 5 days in n = 8 trials; Fig. 8A). The arrow in the KP-insulin sample refers to an outlier that was off scale (44 days) and was excluded from analysis. Although these are overlapping distributions, Student's t test yielded a p value of 0.027. Our findings are in accordance with clinical guidelines regarding differences in product shelf lives at room temperature.
FIGURE 8.

Fibrillation lag-time assays. Box and whisker plot: A, time to fibril formation under formulation-like conditions at 45 °C. Wild-type human insulin (HI; n = 9), KP-insulin (KP; n = 8), or 3-I-TyrB26-KP-insulin (iodo-KP; n = 6) plots are shown. The arrow for KP denotes a single outlier that was off-scale (44 days) and was excluded from statistical analysis. Differences between HI and KP and KP and iodo-KP were significant (p < 0.05). B, wild-type insulin (HI; n = 14), KP-insulin (n = 11), and the 3-I-TyrB26-KP-insulin (n = 11) at 37 °C. Box plots delineate quartiles with bars representing the median (black) and mean (green) values; red error bars represent S.D. Solid circles indicate individual measurements. Brackets designate p values (*) < 0.05 or (**) < 0.01. The presence of fibrils was defined as a 2-fold enhancement of thioflavin T fluorescence. The difference between KP-insulin and 3-I-TyrB26-KP-insulin was highly significant (p < 0.01).
Strikingly, the enhanced susceptibility of KP-insulin to fibrillation on temperature stress was mitigated by the iodo modification (22 ± 7 days in n = 6 trials), leading to a set of lag times indistinguishable from those of wild-type insulin. Interestingly, fibrillation studies of the metal ion-free proteins made 60 μm in phosphate-buffered saline (pH 7.4) demonstrated that, as a monomer, 3-I-TyrB26-KP-insulin is less susceptible to fibrillation at 37 °C than either KP-insulin or wild-type insulin (Fig. 8B).
DISCUSSION
In this study a nonstandard amino acid substitution has been exploited to alter the biophysical properties of a protein. The vast chemical space that can in principle be explored with nonstandard functional groups offers an expanded opportunity in protein design (50). The present work focused on a halogenated derivative of an aromatic amino acid. We chose 3-iodo-Tyr because of its unique physicochemical properties (18), evolutionary history (51), and fortuitous tolerance as a modification of insulin (20).
Iodo derivatives of insulin have long been used as a radioactive tracer (e.g. 3-125I-Tyr or 3-127I-Tyr (52), ordinarily located at position A14 extrinsic to the receptor binding surface of the hormone (19, 37). This A14 modification has no effect on affinity for the insulin receptor. In this context nonradioactive mono-iodo derivatives of insulin were first prepared as reverse-phase HPLC standards to guide purification of a mono-component radioactive derivative (20). Of the three remaining sites of modification in human insulin (i.e. TyrA19, TyrB16, and TyrB26), only the B26 derivative did not exhibit a reduction in receptor binding (20, 37). Indeed, 3-I-TyrB26 and (3,5)-I2-TyrB26 enhance binding (20, 37, 53), motivating the present study. Given recent progress in the crystallographic characterization of model insulin-receptor complexes (54), it would be of future interest to investigate the structural basis of such enhanced affinity.
Protein Allostery
Allosteric regulation of proteins resulting from ligand binding at a site distant from the active site is mediated by transmitted changes in structure (55). A classical example is provided by hemoglobin whose conformational equilibria among T (tense) and R (relaxed) states is modulated by bisphosphoglycerate bound within the central axis of an α2β2 heterotetramer (56). Insulin hexamers likewise participate in coupled conformational equilibria among T6, T3Rf3, and R6 families of structures (Fig. 1E), so named by analogy to hemoglobin. Although the functional significance of the T → R transition is not well understood (46, 57), studies of insulin analogs have suggested that such allostery exploits sites of conformational change relevant to induced fit of the hormone on receptor binding (54, 58).
T6, T3Rf3, and R6 families of insulin hexamers have each been employed in pharmaceutical formulations (9, 59). Their conformational equilibria may be shifted in favor of the more stable R family through the binding of small cyclic alcohols, such as phenol or meta-cresol or with lower affinity, as cyclohexanol (60). The T → R transition is remarkable for a 30-Å displacement in the respective positions of PheB1 as a consequence of a change in B-chain secondary structure (1). Whereas the B1-B8 segment in the T state (as in the solution structure of an insulin monomer (40)) forms an extended strand (residues B1-B6) and β-turn (B7-B10), in the R state this segment provide an N-terminal extension of the central B-chain α-helix (61). The phenolic binding site does not preexist in the structure of the T6 zinc hexamer but is well defined within T3Rf3 (with three ligands) and R6 (with six or seven ligands) hexamers (62). Despite the precise spatial organization of this internal binding site as visualized in crystal structures, 1H NMR studies have established that exchange between bound and free phenol in solution occurs on the millisecond time scale (43).
These concepts are pertinent to the present study, as aromatic alcohols (phenol and meta-cresol) have traditionally been employed in pharmaceutical formulations for their antimicrobial properties (63). Although such insulin formulations date to the 1930s, the additional allosteric role of these ligands was first appreciated only in 1989 (61). Fortuitously, the ligand-dependent T→R transition markedly prolongs the lifetime of the hexamer (35) and augments its resistance to chemical and physical degradation (47). Because such degradation occurs primarily via conformational fluctuations in an insulin monomer (4), protein allostery enhances the effectiveness with which self-assembly enables sequestration of the native monomer (35).
The allosteric role of phenol and meta-cresol gained central importance in the course of efforts to formulate “meal-time” insulin analogs (including KP-insulin) within labile zinc hexamers as a strategy to achieve rapid absorption (9). The two leading such analog formulations (Humalog® and Novolog® thus contain R6 zinc hexamers) (33, 64), which presumably convert transiently to the less stable T6 form in the SQ depot on the rapid dissociation and diffusion of the allosteric ligand (5). Such applications highlight the translational importance of fundamental biophysical concepts.
Although current prandial insulin analog formulations are sufficiently stable to meet the regulatory guidelines of the United States Food and Drug Administration, above room temperature they degrade at rates exceeding that of wild-type human insulin (65). Even at room temperature this differential stability has impact on dilution of the formulation as commonly utilized in the treatment of children (the prescribing information for Humalog can be found on the Eli Lilly website; see also Ref. 67). Whereas vials of wild-type insulin, once opened at room temperature, must be discarded after 28 days at any strength, corresponding vials of Humalog® (LysB28, ProB29-insulin; insulin lispro) must be discarded on dilution after 14 days. Furthermore, the diluted prandial formulations are not recommended for use in insulin pumps due to the risk of degradation and catheter occlusion given the exposure of the protein solution to fluctuations in temperature and agitation.13 This restriction is inherent in the biophysical properties of the mutant proteins irrespective of the potential clinical advantages of pump therapy in such children (68). Such accelerated degradation could in principle be mitigated through further modification of the protein, but only if the modification does not also hinder absorption, which would sacrifice the therapeutic goal of its design (Fig. 9). This would appear to pose a molecular dilemma, as the original design strategies of current insulin analog formulations work at cross-purposes to stability.
FIGURE 9.
Schemes to forestall insulin fibrillation. Top, pathway to fibrillation within insulin formulation vials. The insulin hexamer is the predominant conformational state in the presence of phenol and zinc. At equilibrium a small population of free monomers is nonetheless present, which may undergo partial unfolding en route to fibril formation. Promotion of the hexameric state in 3-I-TyrB26-KP-insulin formulation is envisaged to sequester free monomers, thus delaying fibrillation. Bottom, a rapid pharmacokinetic profile of 3-I-TyrB26-KP-insulin is maintained relative to that of KP-insulin. Diffusion of phenol away from the SQ space leads to a rapid R → T transition and in turn disassembly into dimers and monomers; the latter are efficiently absorbed into capillaries.
The present study suggests that the 3-I-TyrB26 modification of LysB28, ProB29-insulin augments the kinetic stability of the R6 zinc hexamer (Fig. 9, upper panel) without trapping the T6 hexamer in the SQ depot (Fig. 9, lower panel). Furthermore, this modification enhances the intrinsic resistance of the isolated monomer as well as that of the formulated zinc hexamer to fibrillation above room temperature, the principal mode of insulin degradation pertinent to patient care. Moreover, the augmented thermodynamic stability of the modified monomer (as evaluated in guanidine denaturation studies) predicts enhanced resistance to chemical degradation (8). We envisage that 3-I-TyrB26 mediates such protective effects either through direct interactions of the halogen atom with neighboring side chains or through indirect modulation of “weakly polar” electrostatic interactions (69). Such an indirect mechanism may involve transmitted changes through successive edge-face interactions among the eight aromatic rings at the classical dimer interface (TyrB16, PheB24, PheB25, 3-I-TyrB26, and their symmetry-related partners). Dissecting such mechanisms will require future crystallographic, quantum-chemical simulations, and NMR analyses (17).
Properties of Iodo-tyrosine
Iodine lies in column 17 of the periodic table with atomic number 53 and mass 126 atomic mass units; as such, it is the largest element that is essential for vertebrate life. Its atomic radius (∼2 Å) relative to other atoms in a polypeptide (oxygen, nitrogen, carbon, and hydrogen), and indeed relative to smaller halogen atoms, confers unique properties derived from the average distance of its electrons from the nucleus. Unlike smaller halogens, the electronegativity of iodine is approximately equivalent to that of carbon (2.66 versus 2.55) because its electron density is spread over a larger volume; this spreading also imparts a higher polarization potential relative to smaller halogens (18). Together, these factors render 3-I-Tyr hydrophobic despite the inductive effect of the halogen leading to an electrostatic dipole moment (17). It is possible that the greater size and hydrophobicity of 3-I-Tyr relative to Tyr might account for its biophysical effects on LysB28, ProB29-insulin, such as an augmented kinetic barrier to dissociation of the nonpolar dimer interface. The hydrophobicity of the iodine atom in an iodo-aromatic group is in accordance with crystal structures of protein complexes containing thyroid hormone (see below) (70–72). In this context the asymmetric long-range NOEs selectively observed from the C2H proton of 3-I-TyrB26 (relative to the opposite C6H proton) suggest that the iodo-substituent packs within a nonpolar pocket between the A and B chains.
Halogen substitution within an aromatic system (such as in the phenolic ring of Tyr) alters the π orbitals. The electron withdrawing effect of fluorine, for example, diminishes the partial positive charges around the edge of the ring as dramatically observed in perfluorinated benzene with its inverted quadrupole moment (73). By contrast, in the case of iodo-aromatic substituents (and to a lesser extent bromo- and chloro-aromatic substituents), the conjugated π system can, as a seeming paradox, pull electron density away from the halogen, giving rise to a σ-hole on its opposing face (74). This electropositive cap can mediate “halogen bonding,” an interaction analogous to hydrogen bonding wherein the electron-deficient σ-face of the larger halogens can interact with a carbonyl oxygen or other nucleophiles (74). It would be of interest to test whether halogen bonding by 3-I-TyrB26 might contribute to the marked effects of this modification on the kinetic or thermodynamic stability of the R6 hexamer or to its stabilization of the KP-insulin as a metal ion-free monomer. Resolving this issue will require high resolution structures of each conformational state.
Evolutionary Antecedents
Our use of an iodo-Tyr modification to enhance the biophysical properties of a protein may recapitulate an ancestral evolutionary innovation. The relative ease with which iodide can be oxidized (the element is found primarily in seawater as soluble iodide (I−) and iodate groups) (75) has allowed its exploitation by enzymes that direct its incorporation into organic molecules (76). Such an evolutionary innovation lies deep in the history of the Metazoan kingdom as indicated by its detection in certain sponges and corals (77).
The most familiar examples of iodo-aromatic compounds in vertebrate biology are provided by thyroid hormones and their biosynthetic precursors, including 3-I-Tyr and (3,5)-I2-Tyr as modified residues in thyroglobulin (78). Critical to the growth, development, and metabolic homeostasis of vertebrates (including humans) (79), thyroid hormone and its various protein complexes have been extensively studied (80). Co-crystal structures of thyroid hormone-protein complexes (employing transthyretin, thyroid hormone receptor, and thyroxine-binding globulin) and related model complexes have revealed hormone binding pockets that enclose the iodine atoms within nonpolar cavities or crevices (70–72). Close contacts between the halogen and neighboring side chains stabilize such complexes as indicated by studies of thyroid hormone derivatives lacking one or more iodo-aromatic substituents (81). We speculate that such general features will be found in structures of 3-I-TyrB26 derivatives of insulin hexamers and rapid-acting insulin analogs, both as hexameric assemblies (1, 61) and on binding to the insulin receptor (54).
Concluding Remarks
The present study of the 3-I-TyrB26 derivative of a prandial insulin analog has shown that a single halogen modification can markedly alter the biophysical properties of an allosteric protein, including its stability and self-assembly, while preserving biological activity and rapid action in an animal model of diabetes mellitus. We envisage that the structural principles underlying such iodo-aromatic modifications will be found to recapitulate an ancestral evolutionary innovation in biophysical chemistry lying near the root of the Metazoan kingdom.
The general question of how an iodo-aromatic substitution in a protein can modulate its structure, dynamics, and interactions will require high resolution structures and computational investigation at the level of quantum chemistry (17). Because of the small size of insulin and its potential tractability by multidimensional NMR, crystallographic, and multiscale simulation methods (82), iodo-insulin derivatives promise to provide a general model for the integrated analysis of physical and chemical principles pertinent to the complex biology of thyroid hormone and its evolutionary history. In themselves, the surprising properties of 3-I-TyrB26-LysB28, ProB29-insulin uncovered in the present study illustrate the translational promise of nonstandard protein engineering in molecular pharmacology.
Acknowledgments
We thank S. Yadav for peptide synthesis, L. Whittaker for receptor binding assays, A. Minton and S. Kenrick for assistance and training with light scattering techniques, and J. Racca for discussion.
This work was supported, in whole or in part, by National Institutes of Health Grants DK04949 and DK079233 (to M. A. W.). This article is a contribution from the Cleveland Center for Membrane and Structural Biology. In relation to unrelated insulin analogs, M. A. W. and F. I.-B. have stock in Thermalin Diabetes, LLC Cleveland, OH, for which N. B. P. and J. W. are consultants. M. A. W. is the Chief Scientific Officer and a Director.
Dilution of rapid-acting insulin analogs by 5-fold shifts the self-association equilibrium; although a predominance of hexamers is maintained, partial disassembly leads to larger fractions of zinc-free dimers and monomers. Whereas such dilutional disassembly in the SQ depot facilitates absorption (5), in a vial or pump reservoir it is associated with enhanced rates of degradation at or above room temperature.
Cyclic alcohols such as phenol and meta-cresol are widely used as antimicrobial excipients in pharmaceutical formulations and in the case of insulin also function as protective allosteric effectors, decreasing rates of protein degradation (47).
Conformational equilibria among zinc-stabilized insulin hexamer types T6, T3Rf3, and R6 (Rf, frayed R state).
At the lower SQ dose (10 μg; n = 10 per group) the p value was 0.35; at the higher SQ dose (20 μg; n = 10 for the lispro group and n = 11 for the iodo-analog group) the p value was 0.43. If the small apparent differences in mean initial rates of fall in blood-glucose concentration should represent a true difference in initial potency, a power calculation suggests that a trial of >40 rats per group would be required to demonstrate statistical significance.
At the lower SQ dose the p value was 0.54; at the higher SQ dose (20 μg) the p value was 0.17. The latter p value was reduced due to the limited glycemic response of 2 of the 11 rats to the higher dose of the iodo-analog; these outliers may have represented injection site errors or physiologic insulin resistance as occasionally observed among outbred male Lewis rats. Exclusion of these outliers would yield a p value of 0.57 in accordance with the lower-dose study. If the small apparent differences in mean AOC in the present data should represent a true difference in potency, a power calculation suggests that a trial of >100 rats per group would be required to demonstrate statistical significance. In standard pharmaceutical formulations any such differences in intrinsic potency would be compensated by redefinition of the formulation strength in relation to international units per mg of protein such that equivalent biological activity per ml is obtained.
Insulin is capable of coordinating a variety of divalent metal ions to form hexamers in which the three HisB10 side chains in each component trimer provide a coordination site (octahedral in T3 sites and tetrahedral in R3 or Rf3 sites) (1, 61).
Cyclohexanol is the non-aromatic isostere of phenol and is UV-transparent. It is capable of binding to insulin hexamers albeit with lower affinity (44).
Tetrahedrally coordinated Co2+ provides an absorption spectrum in the visible range due to d-d transitions in its incomplete d-shell; this band is negligible in an octahedral complex (32).
Diluted insulin formulations are frequently used in the pediatric population to enhance the precision of dosing (67).
Exposure of insulin or insulin analog formulations to hydrophobic interfaces, elevated temperature, and agitation are factors that accelerate the formulation of insoluble precipitates in pump lines and catheters (4). The recommended time of use for diluted Humalog® formulations is 14 days after opening of the vial at room temperature; use in pumps is not approved, but off-label applications have been reported (66).
- SQ
- subcutaneous
- 3-I-Tyr (3-I-Y)
- derivative of tyrosine containing an iodo-substituent at ring position 3
- CD
- circular dichroism
- KP-insulin
- analog containing substitutions ProB28 → Lys and LysB29 → Pro
- SEC-MALS
- size-exclusion chromatography and multiangle light scattering
- TOCSY
- two-dimensional total correlation spectroscopy
- AOC
- area over the curve
- T
- tense
- R
- relaxed.
REFERENCES
- 1. Baker E. N., Blundell T. L., Cutfield J. F., Cutfield S. M., Dodson E. J., Dodson G. G., Hodgkin D. M., Hubbard R. E., Isaacs N. W., Reynolds C. D. (1988) The structure of 2Zn pig insulin crystals at 1.5 Å resolution. Philos. Trans. R. Soc. Lond. B Biol. Sci. 319, 369–456 [DOI] [PubMed] [Google Scholar]
- 2. De Meyts P. (2004) Insulin and its receptor: structure, function and evolution. Bioessays 26, 1351–1362 [DOI] [PubMed] [Google Scholar]
- 3. Dodson G., Steiner D. (1998) The role of assembly in insulin's biosynthesis. Curr. Opin. Struct. Biol. 8, 189–194 [DOI] [PubMed] [Google Scholar]
- 4. Brange J., Andersen L., Laursen E. D., Meyn G., Rasmussen E. (1997) Toward understanding insulin fibrillation. J. Pharm. Sci. 86, 517–525 [DOI] [PubMed] [Google Scholar]
- 5. Brange J., Owens D. R., Kang S., Vølund A. (1990) Monomeric insulins and their experimental and clinical implications. Diabetes Care 13, 923–954 [DOI] [PubMed] [Google Scholar]
- 6. Brange J., Ribel U., Hansen J. F., Dodson G., Hansen M. T., Havelund S., Melberg S. G., Norris F., Norris K., Snel L. (1988) Monomeric insulins obtained by protein engineering and their medical implications. Nature 333, 679–682 [DOI] [PubMed] [Google Scholar]
- 7. Slieker L. J., Brooke G. S., DiMarchi R. D., Flora D. B., Green L. K., Hoffmann J. A., Long H. B., Fan L., Shields J. E., Sundell K. L., Surface P. L., Chance R. E. (1997) Modifications in the B10 and B26–30 regions of the B chain of human insulin alter affinity for the human IGF-I receptor more than for the insulin receptor. Diabetologia 40, S54–S61 [DOI] [PubMed] [Google Scholar]
- 8. Brems D. N., Brown P. L., Bryant C., Chance R. E., Green L. K., Long H. B., Miller A. A., Millican R., Shields J. E., Frank B. H. (1992) Improved insulin stability through amino acid substitution. Protein Eng. 5, 519–525 [DOI] [PubMed] [Google Scholar]
- 9. Bakaysa D. L., Radziuk J., Havel H. A., Brader M. L., Li S., Dodd S. W., Beals J. M., Pekar A. H., Brems D. N. (1996) Physicochemical basis for the rapid time-action of LysB28ProB29-insulin: dissociation of a protein-ligand complex. Protein Sci. 5, 2521–2531 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Thurow H., Geisen K. (1984) Stabilisation of dissolved proteins against denaturation at hydrophobie interfaces. Diabetologia 27, 212–218 [DOI] [PubMed] [Google Scholar]
- 11. Landau Z., Klonoff D., Nayberg I., Feldman D., Levit S. B., Lender D., Mosenzon O., Raz I., Wainstein J. (March 10, 2014) Improved pharmacokinetic and pharmacodynamic profile of insulin analogs using InsuPatch, a local heating device. Diabetes Metab. Res. Rev., 10.1002/dmrr.2536 [DOI] [PubMed] [Google Scholar]
- 12. Adams M. J., Blundell T. L., Dodson E. J., Dodson G. G., Vijayan M., Baker E. N., Hardine M. M., Hodgkin D. C., Rimer B., Sheet S. (1969) Structure of rhombohedral 2 zinc insulin crystals. Nature 224, 491–495 [Google Scholar]
- 13. Blundell T. L., Cutfield J. F., Cutfield S. M., Dodson E. J., Dodson G. G., Hodgkin D. C., Mercola D. A., Vijayan M. (1971) Atomic positions in rhombohedral 2-zinc insulin crystals. Nature 231, 506–511 [DOI] [PubMed] [Google Scholar]
- 14. Bentley G., Dodson E., Dodson G., Hodgkin D., Mercola D. (1976) Structure of insulin in 4-zinc insulin. Nature 261, 166–168 [DOI] [PubMed] [Google Scholar]
- 15. Derewenda U., Derewenda Z., Dodson G. G., Hubbard R. E., Korber F. (1989) Molecular structure of insulin: the insulin monomer and its assembly. Br. Med. Bull. 45, 4–18 [DOI] [PubMed] [Google Scholar]
- 16. Hirsch I. B. (2005) Insulin analogues. N. Engl. J. Med. 352, 174–183 [DOI] [PubMed] [Google Scholar]
- 17. Lu Y., Liu Y., Xu Z., Li H., Liu H., Zhu W. (2012) Halogen bonding for rational drug design and new drug discovery. Expert. Opin. Drug Discov. 7, 375–383 [DOI] [PubMed] [Google Scholar]
- 18. Küpper F. C., Feiters M. C., Olofsson B., Kaiho T., Yanagida S., Zimmermann M. B., Carpenter L. J., Luther G. W., 3rd, Lu Z., Jonsson M., Kloo L. (2011) Commemorating two centuries of iodine research: an interdisciplinary overview of current research. Angew. Chem. Int. Ed. Engl. 50, 11598–11620 [DOI] [PubMed] [Google Scholar]
- 19. Linde S., Sonne O., Hansen B., Gliemann J. (1981) Monoiodoinsulin labelled in tyrosine residue 16 or 26 of the insulin B-chain. Preparation and characterization of some binding properties. Hoppe Seylers Z Physiol. Chem. 362, 573–579 [DOI] [PubMed] [Google Scholar]
- 20. Frank B. H., Peavy D. E., Hooker C. S., Duckworth W. C. (1983) Receptor binding properties of monoiodotyrosyl insulin isomers purified by high performance liquid chromatography. Diabetes 32, 705–711 [DOI] [PubMed] [Google Scholar]
- 21. Inouye K., Watanabe K., Tochino Y., Kobayashi M., Shigeta Y. (1981) Semisynthesis and properties of some insulin analogs. Biopolymers 20, 1845–1858 [DOI] [PubMed] [Google Scholar]
- 22. Barany G., Merrifield R. B. (1980) in The Peptides (Gross E., Meienhofer J. eds.) pp. 273–284, Academic Press, New York [Google Scholar]
- 23. Kubiak T., Cowburn D. (1986) Enzymatic semisynthesis of porcine despentapeptide (B26–30) insulin using unprotected desoctapeptide (B23–30) insulin as a substrate. Int. J. Pept. Protein Res. 27, 514–521 [DOI] [PubMed] [Google Scholar]
- 24. Sreerama N., Woody R. W. (1993) A self-consistent method for the analysis of protein secondary structure from circular dichroism. Anal. Biochem. 209, 32–44 [DOI] [PubMed] [Google Scholar]
- 25. Sosnick T. R., Fang X., Shelton V. M. (2000) Application of circular dichroism to study RNA folding transitions. Methods Enzymol. 317, 393–409 [DOI] [PubMed] [Google Scholar]
- 26. Pace C. N., Shaw K. L. (2000) Linear extrapolation method of analyzing solvent denaturation curves. Proteins 4, 1–7 [DOI] [PubMed] [Google Scholar]
- 27. Havelund S., Plum A., Ribel U., Jonassen I., Vølund A., Markussen J., Kurtzhals P. (2004) The mechanism of protraction of insulin detemir, a long-acting, acylated analog of human insulin. Pharm. Res. 21, 1498–1504 [DOI] [PubMed] [Google Scholar]
- 28. Hedo J. A., Harrison L. C., Roth J. (1981) Binding of insulin receptors to lectins: evidence for common carbohydrate determinants on several membrane receptors. Biochemistry 20, 3385–3393 [DOI] [PubMed] [Google Scholar]
- 29. Whittaker L., Hao C., Fu W., Whittaker J. (2008) High-affinity insulin binding: insulin interacts with two receptor ligand binding sites. Biochemistry 47, 12900–12909 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30. Wang Z. X. (1995) An exact mathematical expression for describing competitive binding of two different ligands to a protein molecule. FEBS Lett. 360, 111–114 [DOI] [PubMed] [Google Scholar]
- 31. Huang K., Dong J., Phillips N. B., Carey P. R., Weiss M. A. (2005) Proinsulin is refractory to protein fibrillation. Topological protection of a precursor protein from cross-β assembly. J. Biol. Chem. 280, 42345–42355 [DOI] [PubMed] [Google Scholar]
- 32. Roy M., Brader M. L., Lee R. W., Kaarsholm N. C., Hansen J. F., Dunn M. F. (1989) Spectroscopic signatures of the T to R conformational transition in the insulin hexamer. J. Biol. Chem. 264, 19081–19085 [PubMed] [Google Scholar]
- 33. Birnbaum D. T., Kilcomons M. A., DeFelippis M. R., Beals J. M. (1997) Assembly and dissociation of human insulin and LysB28ProB29-insulin hexamers: a comparison study. Pharm. Res. 14, 25–36 [DOI] [PubMed] [Google Scholar]
- 34. Hua Q. X., Hu S. Q., Frank B. H., Jia W., Chu Y. C., Wang S. H., Burke G. T., Katsoyannis P. G., Weiss M. A. (1996) Mapping the functional surface of insulin by design: structure and function of a novel A-chain analogue. J. Mol. Biol. 264, 390–403 [DOI] [PubMed] [Google Scholar]
- 35. Rahuel-Clermont S., French C. A., Kaarsholm N. C., Dunn M. F. (1997) Mechanisms of stabilization of the insulin hexamer through allosteric ligand interactions. Biochemistry 36, 5837–5845 [DOI] [PubMed] [Google Scholar]
- 36. Saker F., Ybarra J., Leahy P., Hanson R. W., Kalhan S. C., Ismail-Beigi F. (1998) Glycemia-lowering effect of cobalt chloride in the diabetic rat: role of decreased gluconeogenesis. Am. J. Physiol. 274, E984–E991 [DOI] [PubMed] [Google Scholar]
- 37. Peavy D. E., Abram J. D., Frank B. H., Duckworth W. C. (1984) Receptor binding and biological activity of specifically labeled [125I]- and [127I]monoiodoinsulin isomers in isolated rat adipocytes. Endocrinology 114, 1818–1824 [DOI] [PubMed] [Google Scholar]
- 38. Pocker Y., Biswas S. B. (1980) Conformational dynamics of insulin in solution. Circular dichroic studies. Biochemistry 19, 5043–5049 [DOI] [PubMed] [Google Scholar]
- 39. Shoelson S. E., Lu Z. X., Parlautan L., Lynch C. S., Weiss M. A. (1992) Mutations at the dimer, hexamer, and receptor binding surfaces of insulin independently affect insulin-insulin and insulin-receptor interactions. Biochemistry 31, 1757–1767 [DOI] [PubMed] [Google Scholar]
- 40. Hua Q. X., Jia W., Weiss M. A. (2011) Conformational dynamics of insulin. Front. Endocrinol. (Lausanne) 2, 48. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41. Weiss M. A., Hua Q. X., Lynch C. S., Frank B. H., Shoelson S. E. (1991) Heteronuclear 2D NMR studies of an engineered insulin monomer: assignment and characterization of the receptor-binding surface by selective 2H and 13C labeling with application to protein design. Biochemistry 30, 7373–7389 [DOI] [PubMed] [Google Scholar]
- 42. Weiss M. A., Nguyen D. T., Khait I., Inouye K., Frank B. H., Beckage M., O'Shea E., Shoelson S. E., Karplus M., Neuringer L. J. (1989) Two-dimensional NMR and photo-CIDNP studies of the insulin monomer: assignment of aromatic resonances with application to protein folding, structure, and dynamics. Biochemistry 28, 9855–9873 [DOI] [PubMed] [Google Scholar]
- 43. Jacoby E., Hua Q. X., Stern A. S., Frank B. H., Weiss M. A. (1996) Structure and dynamics of a protein assembly. 1H NMR studies of the 36-kDa R6 insulin hexamer. J. Mol. Biol. 258, 136–157 [DOI] [PubMed] [Google Scholar]
- 44. Brader M. L., Kaarsholm N. C., Lee R. W., Dunn M. F. (1991) Characterization of the R-state insulin hexamer and its derivatives. The hexamer is stabilized by heterotropic ligand binding interactions. Biochemistry 30, 6636–6645 [DOI] [PubMed] [Google Scholar]
- 45. Hassiepen U., Federwisch M., Mülders T., Wollmer A. (1999) The lifetime of insulin hexamers. Biophys. J. 77, 1638–1654 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. Dunn M. F. (2005) Zinc-ligand interactions modulate assembly and stability of the insulin hexamer; a review. Biometals 18, 295–303 [DOI] [PubMed] [Google Scholar]
- 47. Brange J., Langkjaer L. (1992) Chemical stability of insulin. 3. Influence of excipients, formulation, and pH. Acta Pharm. Nord. 4, 149–158 [PubMed] [Google Scholar]
- 48. Nielsen L., Khurana R., Coats A., Frokjaer S., Brange J., Vyas S., Uversky V. N., Fink A. L. (2001) Effect of environmental factors on the kinetics of insulin fibril formation: elucidation of the molecular mechanism. Biochemistry 40, 6036–6046 [DOI] [PubMed] [Google Scholar]
- 49. Librizzi F., Rischel C. (2005) The kinetic behavior of insulin fibrillation is determined by heterogeneous nucleation pathways. Protein Sci. 14, 3129–3134 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50. Liu C. C., Schultz P. G. (2010) Adding new chemistries to the genetic code. Annu. Rev. Biochem. 79, 413–444 [DOI] [PubMed] [Google Scholar]
- 51. Venturi S. (2011) Evolutionary significance of iodine. Curr. Chem. Biol. 5, 155–162 [Google Scholar]
- 52. Freychet P., Roth J., Neville D. M., Jr. (1971) Monoiodoinsulin: demonstration of its biological activity and binding to fat cells and liver membranes. Biochem. Biophys. Res. Commun. 43, 400–408 [DOI] [PubMed] [Google Scholar]
- 53. Mirmira R. G., Nakagawa S. H., Tager H. S. (1991) Importance of the character and configuration of residues B24, B25, and B26 in insulin-receptor interactions. J. Biol. Chem. 266, 1428–1436 [PubMed] [Google Scholar]
- 54. Menting J. G., Whittaker J., Margetts M. B., Whittaker L. J., Kong G. K., Smith B. J., Watson C. J., Záková L., Kletvíková E., Jiráček J., Chan S. J., Steiner D. F., Dodson G. G., Brzozowski A. M., Weiss M. A., Ward C. W., Lawrence M. C. (2013) How insulin engages its primary binding site on the insulin receptor. Nature 493, 241–245 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55. Monod J., Wyman J., Changeux J.-P. (1965) On the nature of allosteric transitions: a plausible model. J. Mol. Biol. 12, 88–118 [DOI] [PubMed] [Google Scholar]
- 56. Benesch R., Benesch R. E. (1967) The effect of organic phosphates from the human erythrocyte on the allosteric properties of hemoglobin. Biochem. Biophys. Res. Commun. 26, 162–167 [DOI] [PubMed] [Google Scholar]
- 57. Wan Z. L., Huang K., Hu S. Q., Whittaker J., Weiss M. A. (2008) The structure of a mutant insulin uncouples receptor binding from protein allostery. An electrostatic block to the TR transition. J. Biol. Chem. 283, 21198–21210 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Nakagawa S. H., Zhao M., Hua Q. X., Hu S. Q., Wan Z. L., Jia W., Weiss M. A. (2005) Chiral mutagenesis of insulin. Foldability and function are inversely regulated by a stereospecific switch in the B chain. Biochemistry 44, 4984–4999 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59. Jonassen I., Havelund S., Hoeg-Jensen T., Steensgaard D. B., Wahlund P. O., Ribel U. (2012) Design of the novel protraction mechanism of insulin degludec, an ultra-long-acting basal insulin. Pharm. Res. 29, 2104–2114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Brader M. L., Dunn M. F. (1991) Insulin hexamers: new conformations and applications. Trends Biochem. Sci. 16, 341–345 [DOI] [PubMed] [Google Scholar]
- 61. Derewenda U., Derewenda Z., Dodson E. J., Dodson G. G., Reynolds C. D., Smith G. D., Sparks C., Swenson D. (1989) Phenol stabilizes more helix in a new symmetrical zinc insulin hexamer. Nature 338, 594–596 [DOI] [PubMed] [Google Scholar]
- 62. Kaarsholm N. C., Ko H. C., Dunn M. F. (1989) Comparison of solution structural flexibility and zinc binding domains for insulin, proinsulin, and miniproinsulin. Biochemistry 28, 4427–4435 [DOI] [PubMed] [Google Scholar]
- 63. Brange J. (ed) (1987) Galenics of Insulin: The Physico-chemical and Pharmaceutical Aspects of Insulin and Insulin Preparations, Vol. 36, pp. 40–41, Springer Verlag, Berlin, Heidelberg [Google Scholar]
- 64. Whittingham J. L., Edwards D. J., Antson A. A., Clarkson J. M., Dodson G. G. (1998) Interactions of phenol and m-cresol in the insulin hexamer, and their effect on the association properties of B28 Pro → Asp insulin analogues. Biochemistry 37, 11516–11523 [DOI] [PubMed] [Google Scholar]
- 65. Weiss M. A. (2013) Design of ultra-stable insulin analogues for the developing world. J. Health Spec. 1, 59–70 [Google Scholar]
- 66. Bharucha T., Brown J., McDonnell C., Gebert R., McDougall P., Cameron F., Werther G., Zacharin M. (2005) Neonatal diabetes mellitus: insulin pump as an alternative management strategy. J. Paediatr. Child Health 41, 522–526 [DOI] [PubMed] [Google Scholar]
- 67. Phillip M., Battelino T., Rodriguez H., Danne T., Kaufman F. (2007) Use of Insulin Pump Therapy in the Pediatric Age-Group Consensus statement from the European Society for Paediatric Endocrinology, the Lawson Wilkins Pediatric Endocrine Society, and the International Society for Pediatric and Adolescent Diabetes, endorsed by the American Diabetes Association and the European Association for the Study of Diabetes. Diabetes Care 30, 1653–1662 [DOI] [PubMed] [Google Scholar]
- 68. Shalitin S., Phillip M. (2008) The use of insulin pump therapy in the pediatric age group. Horm. Res. 70, 14–21 [DOI] [PubMed] [Google Scholar]
- 69. Burley S. K., Petsko G. A. (1988) Weakly polar interaction in proteins. Adv. Protein Chem. 39, 125–189 [DOI] [PubMed] [Google Scholar]
- 70. Wojtczak A., Cody V., Luft J. R., Pangborn W. (1996) Structures of human transthyretin complexed with thyroxine at 2.0 Å resolution and 3′,5′-dinitro-N-acetyl-l-thyronine at 2.2 Å resolution. Acta Crystallogr. D. Biol. Crystallogr. 52, 758–765 [DOI] [PubMed] [Google Scholar]
- 71. Sandler B., Webb P., Apriletti J. W., Huber B. R., Togashi M., Cunha Lima S. T., Juric S., Nilsson S., Wagner R., Fletterick R. J., Baxter J. D. (2004) Thyroxine-thyroid hormone receptor interactions. J. Biol. Chem. 279, 55801–55808 [DOI] [PubMed] [Google Scholar]
- 72. Zhou A., Wei Z., Read R. J., Carrell R. W. (2006) Structural mechanism for the carriage and release of thyroxine in the blood. Proc. Natl. Acad. Sci. U.S.A. 103, 13321–13326 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. West A. P., Mecozzi S., Dougherty D. A. (1997) Theoretical studies of the supramolecular synthon benzene··· hexafluorobenzene. J. Phys. Org. Chem. 10, 347–350 [Google Scholar]
- 74. Politzer P., Lane P., Concha M. C., Ma Y., Murray J. S. (2007) An overview of halogen bonding. J. Mol. Model 13, 305–311 [DOI] [PubMed] [Google Scholar]
- 75. Truesdale V. W., Luther G. W., III, Canosa-Mas C. (1995) Molecular iodine reduction in seawater, an improved rate equation considering organic compounds. Mar. Chem. 48, 143–150 [Google Scholar]
- 76. Blasiak L. C., Drennan C. L. (2009) Structural perspective on enzymatic halogenation. Acc. Chem. Res. 42, 147–155 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77. Wheeler H. L., Mendel L. B. (1909) The iodine complex in sponges (3,5-diiodtyrosine). J. Biol. Chem. 7, 1–9 [Google Scholar]
- 78. Chemburkar S. R., Deming K. C., Reddy R. E. (2010) Chemistry of thyroxine: an historical perspective and recent progress on its synthesis. Tetrahedron 66, 1955–1962 [Google Scholar]
- 79. Zimmermann M. B. (2009) Iodine deficiency. Endocr. Rev. 30, 376–408 [DOI] [PubMed] [Google Scholar]
- 80. Schussler G. C. (2000) The thyroxine-binding proteins. Thyroid 10, 141–149 [DOI] [PubMed] [Google Scholar]
- 81. Choh H. L. (ed) (1978) Hormonal Proteins and Peptides: Thyroid Hormones, Vol. 6, pp. 107–193, Academic Press, Inc., New York [Google Scholar]
- 82. Groenhof G. (2013) Solving chemical problems with a mixture of quantum-mechanical and molecular mechanics calculations: Nobel Prize in chemistry 2013. Angew. Chem. Int. Ed. Engl. 52, 12489–12491 [DOI] [PubMed] [Google Scholar]








