throbber
Article
`
`pubs.acs.org/IC
`
`Lead(II) Complex Formation with L‑Cysteine in Aqueous Solution
`Farideh Jalilehvand,*,† Natalie S. Sisombath,† Adam C. Schell,† and Glenn A. Facey‡
`†
`Department of Chemistry, University of Calgary, 2500 University Drive NW, Calgary, Alberta T2N 1N4, Canada
`‡
`Department of Chemistry, University of Ottawa, 10 Marie Curie Private, Ottawa, Ontario K1N 6N5, Canada
`*S Supporting Information
`
`ABSTRACT: The lead(II) complexes
`formed with the
`multidentate chelator L-cysteine (H2Cys) in an alkaline
`aqueous solution were studied using 207Pb, 13C, and 1H
`NMR, Pb LIII-edge X-ray absorption, and UV−vis spectro-
`scopic techniques, complemented by electrospray ion mass
`spectrometry (ESI-MS). The H2Cys/PbII mole ratios were
`varied from 2.1 to 10.0 for two sets of solutions with CPbII =
`0.01 and 0.1 M, respectively, prepared at pH values (9.1−10.4)
`for which precipitates of lead(II) cysteine dissolved. At low
`H2Cys/PbII mole ratios (2.1−3.0), a mixture of the dithiolate
`[Pb(S,N-Cys)2]2− and [Pb(S,N,O-Cys)(S-HCys)]− complexes
`with average Pb−(N/O) and Pb−S distances of 2.42 ± 0.04
`and 2.64 ± 0.04 Å, respectively, was found to dominate. At
`high concentration of free cysteinate (>0.7 M), a significant amount converts to the trithiolate [Pb(S,N-Cys)(S-HCys)2]2−,
`including a minor amount of a PbS3-coordinated [Pb(S-HCys)3]− complex. The coordination mode was evaluated by fitting
`linear combinations of EXAFS oscillations to the experimental spectra and by examining the 207Pb NMR signals in the chemical
`Pb = 2006−2507 ppm, which became increasingly deshielded with increasing free cysteinate concentration. One-
`shift range δ
`pulse magic-angle-spinning (MAS) 207Pb NMR spectra of crystalline Pb(aet)2 (Haet = 2-aminoethanethiol or cysteamine) with
`
`PbS2N2 coordination were measured for comparison (δiso = 2105 ppm). The UV−vis spectra displayed absorption maxima at
`298−300 nm (S− → PbII charge transfer) for the dithiolate PbS2N(N/O) species; with increasing ligand excess, a shoulder
`appeared at ∼330 nm for the trithiolate PbS3N and PbS3 (minor) complexes. The results provide spectroscopic fingerprints for
`structural models for lead(II) coordination modes to proteins and enzymes.
`
`■ INTRODUCTION
`
`The efficiency of cysteine-rich proteins and peptides, e.g.,
`metallothioneins and phytochelatins,
`in removing harmful
`heavy metals from the cells and tissues1,2 has inspired the
`assessment of cysteine as an ecofriendly agent for extracting
`heavy metals from a contaminated environment. Cysteine
`+)COO−), as well as penicillamine
`(H2Cys = HSCH2CH(NH3
`(H2Pen) and glutathione (GSH), can liberate lead bound in
`contaminated soil,
`in iron/manganese oxides, and in lead
`phosphate/carbonate salts or in mine tailings by increasing its
`solubility and mobilization.3−5 Moreover, the ability of cysteine
`to capture heavy metals including lead from polluted water can
`be important
`in the development of new materials with
`treatment.6 A
`potential use in drainage and wastewater
`cysteine-based nanosized chelating agent
`that
`selectively
`removes PbII
`ions has recently been developed for
`the
`treatment of lead poisoning.7
`In recent years, cysteine has been introduced as an
`environmentally friendly source of
`sulfur
`for preparing
`nanocrystalline PbS, a widely used semiconductor. Such
`nanocrystals can be prepared by mixing Pb(NO3)2 or
`Pb(OAc)2 (OAc− = acetate) with cysteine to form a lead(II)
`cysteine precursor, followed by hydrothermal decomposition to
`PbS. Different morphologies, shapes, and sizes can be obtained
`
`depending on the metal-to-ligand mole ratio, concentration, or
`pH.8−11 It has been suggested that the precursor is polycrystal-
`line HSCH2CH(NH2)COOPbOH8 or has a polymeric
`[−SCH2CH(COOH)NHPb−]n structure,11 in well-aligned
`one-dimensional nanowires.12
`Corrie and co-workers have reported formation constants for
`several mononuclear lead(II) cysteine complexes in aqueous
`4−, Pb-
`2−, Pb(Cys)3
`solution,
`including Pb(Cys), Pb(Cys)2
`(HCys)+, Pb(Cys)(HCys)−, and Pb(Cys)2(OH)3−, however,
`with revised values, e.g., for the Pb(Cys) complex in their later
`reports.13−15 Bizri and co-workers also reported formation
`4− and
`constants for the above complexes, except for Pb(Cys)3
`Pb(Cys)2(OH)3−, but included a Pb(Cys)(OH)− complex; see
`Figure S-1a in the Supporting Information (SI).16 To explain
`the high stability of the Pb(Cys) complex, cysteinate was
`proposed to act as a tridentate ligand, binding simultaneously
`through the thiolate (−S−), carboxylate (−COO−), and amine
`(−NH2) groups.15−17 However, a subsequent study of the
`COO− stretching frequencies indicated that cysteinate in an
`alkaline solution exclusively binds to PbII via the amine and
`thiolate groups.18 Pardo et al. proposed formation constants for
`
`Received: October 22, 2014
`Published: February 19, 2015
`
`© 2015 American Chemical Society
`
`2160
`
`DOI: 10.1021/ic5025668
`Inorg. Chem. 2015, 54, 2160−2170
`
`Downloaded via REPRINTS DESK INC on March 31, 2020 at 13:20:21 (UTC).
`
`See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
`
`Eton Ex. 1081
`1 of 11
`
`

`

`Inorganic Chemistry
`
`the Pb(Cys), Pb(HCys)+, Pb(HCys)2,
`a set consisting of
`2− complexes.19 Recently, Crea
`Pb(Cys)(HCys)−, and Pb(Cys)2
`et al. obtained formation constants for Pb(Cys), Pb(HCys)+,
`Pb(H2Cys)2+, Pb(Cys)(OH)−, and Pb(Cys)2
`2− to describe the
`stoichiometric composition of the lead(II) cysteine complexes
`formed at several ionic strengths (0 < I ≤ 1.0 M NaNO3) and
`temperatures; see Figure S-2a in the SI.20 The highest PbII
`concentration used in all studies above was 0.5 mM.
`In a high-field 1H and 13C NMR study, Kane-Maguire and
`Riley investigated the PbII binding to cysteine at both acidic
`(pD 1.9) and alkaline (pD 12.9) D2O solutions.21 The report
`includes proton-coupling constants for free cysteine (L), as well
`as mole fractions of its three rotamers: trans (t) and gauche (g
`and h), at different pD values in the range 1.80−12.92 (pD =
`pH reading + 0.4).21,22 Each rotamer was ascribed a preferred
`mode of binding: rotamer t to bidentate (S,N), h to tridentate
`(S, N, O), and g to bidentate (S,O). No significant lead(II)
`cysteine complex formation was observed in the acidic solutions
`(pD 1.9) with H2Cys/PbII mole ratios 0.5−6.0, which is
`consistent with the well-known ability of
`lead(II) to form
`nitrate complexes.23 At H2Cys/Pb(NO3)2 mole ratios ≥2.0
`(CPbII = 10 mM), only PbL2 complexes were proposed to form
`in alkaline media (pD 12.9), with cysteine mainly acting as a
`tridentate (S,N,O) or a bidentate (S,N) ligand. It was also
`suggested that when CPb(NO3)2 = CH2Cys = 10 mM (pD 12.9),
`PbL species with 63% Pb(S,N,O-Cys), 30% Pb(S,O-Cys), and
`7% Pb(S,N-Cys) coordination were formed in proportion to
`the mole fractions of h, g, and t
`rotamers,
`respectively.
`However, we could not prepare aqueous solutions with 1:1
`PbII/cysteine composition because the initially formed
`precipitate did not dissolve in alkaline media even at pH
`12.0, and stability constants for lead(II) hydrolysis indicate that
`precipitation of lead(II) hydroxide should start in such highly
`alkaline media;20 see Figures S-1b and S-2b in the SI.
`Reliable structural
`information to allow a better under-
`standing of the nature of the lead(II) complexes formed with
`cysteine is clearly needed. We used a combination of
`including UV−vis, 207Pb, 13C, and
`spectroscopic techniques,
`1H NMR, extended X-ray absorption fine structure (EXAFS),
`and electrospray ion mass spectrometry (ESI-MS) to study the
`coordination and bonding in lead(II) cysteine complexes
`formed in two sets of alkaline solutions with CPbII = 10 and
`100 mM for H2Cys/PbII mole ratios ≥2.1. To obtain such
`concentrations, the pH was raised (9.1−10.4) to dissolve the
`lead(II) cysteine precipitate that forms when adding lead(II) to
`cysteine solutions. Both the −SH and −NH3
`+ groups of
`cysteine deprotonate at about pH 8.5,24 thus increasing its
`ability to coordinate via the thiolate and amine groups.
`■ EXPERIMENTAL SECTION
`Sample Preparation.
`cysteamine (Haet,
`L-Cysteine,
`·3H2O, and sodium hydroxide
`H2NCH2CH2SH), PbO, Pb(ClO4)2
`were used as supplied (Sigma-Aldrich). All syntheses were carried out
`under a stream of argon gas. Deoxygenated water
`for sample
`preparation was prepared by bubbling argon gas through boiled
`distilled water. The pH values of the solutions were monitored with a
`Thermo Scientific Orion Star pH meter.
`Two sets of lead(II) cysteine solutions were prepared with different
`for CPbII ∼ 10 and 100 mM,
`H2Cys/Pb(ClO4)2 mole ratios
`respectively, at an alkaline pH at which the lead(II) cysteine
`precipitate dissolved; see Table 1. Lead(II) cysteine solutions A−G
`(CPbII ≈ 10 mM) and A*−F* (CPbII ≈ 100 mM) were freshly prepared
`·3H2O (0.05 mmol) to cysteine dissolved in
`by adding Pb(ClO4)2
`deoxygenated water (0.105−0.75 mmol, pH 2.0−2.4). For 207Pb NMR
`
`Article
`
`Table 1. Composition of Lead(II) Cysteine Solutions
`
`H2Cys/PbII mole
`ratio
`2.1
`3.0
`4.0
`5.0
`8.0
`10.0
`15.0
`15.0
`
`pH
`10.4
`9.1
`9.1
`9.1
`9.1
`9.1
`9.1
`8.9
`
`solution
`A
`B
`C
`D
`E
`F
`G
`G′
`
`CPbII
`(mM)
`10
`10
`10
`10
`10
`10
`10
`10
`
`solution
`A*
`B*
`C*
`D*
`E*
`F*
`
`CPbII
`(mM)
`100
`100
`100
`100
`100
`100
`
`and UV−vis measurements of solutions A−G, 50 mM stock solutions
`of enriched 207PbO (94.5%) from Cambridge Isotope Laboratories and
`PbO dissolved in 0.15 M HClO4 were prepared, respectively. Upon
`the dropwise addition of 6.0 M NaOH, an off-white precipitate
`formed, which momentarily dissolved at pH ∼7; after a few seconds, a
`cream-colored precipitate appeared. The addition of a 1.0 M sodium
`hydroxide solution continued until the solid dissolved above pH 8.5
`and gave a clear colorless solution. For solutions A (CPbII = 10 mM)
`and A* (CPbII = 100 mM), with the mole ratio H2Cys/PbII = 2.1, the
`solid dissolved completely at pH ∼10.4, and for solutions B and B*
`(H2Cys/PbII mole ratio = 3.0), it dissolved at pH 9.1. For consistency,
`the pH values of solutions with higher H2Cys/PbII mole ratios were
`also set at pH 9.1. The final volume for each solution was adjusted to
`5.0 mL. Solutions A−F were used for ESI-MS and 1H and 13C NMR
`(prepared in 99.9% deoxygenated D2O) measurements. For solutions
`in D2O, the pH-meter reading was 10.4 for solution A (pD = pH
`reading + 0.4)22 and 9.1 for B−F. 207Pb NMR spectra were measured
`for all solutions (10% v/v D2O), while Pb LIII-edge EXAFS spectra
`were measured for solutions A−E and A*−F*.
`Bis(2-aminoethanethiolato)lead(II) Solid, Pb(aet)2. A total of
`0.964 g (12.5 mmol) of Haet dissolved in 10 mL of deoxygenated
`water (pH 9.6) was added to a suspension of PbO (1.116 g, 5 mmol)
`in 50 mL of ethanol at 50 °C and refluxed for 3 h under an argon
`atmosphere, giving a pale-yellow solution, which was then filtered and
`cooled in a refrigerator. Colorless crystals formed after 48 h and were
`filtered, washed with ethanol, and dried under vacuum (turning
`yellow). Eleme anal. Calcd for Pb(SCH2CH2NH2)2: C, 13.36; H, 3.37;
`N, 7.79. Found: C, 13.41; H, 3.43; N, 7.81. The unit cell dimensions of
`the crystal were also verified, matching the literature values.25
`Methods. Details about instrumentation and related procedures for
`ESI-MS (Aglient 6520 Q-Tof), UV−vis (Cary 300), and 1H, 13C, and
`207Pb NMR spectroscopy (Bruker AMX 300 and Avance II 400 MHz),
`as well as EXAFS data collection and data analyses are provided
`elsewhere.26 UV−vis spectra of solutions A−G were measured using
`0.25, 0.5, and 0.75 nm data intervals, with a 1.5 absorbance Agilent
`rear-beam attenuator mesh filter in the reference position. ESI-MS
`spectra for solutions A, B, and F were measured in both positive- and
`negative-ion modes. 207Pb NMR spectra for solutions A−G enriched in
`207Pb were measured at room temperature using a Bruker AMX 300
`equipped with a 10 mm broad-band probe. For these solutions, the
`207Pb NMR chemical shift was externally calibrated relative to 1.0 M
`Pb(NO3)2 in D2O, resonating at −2961.2 ppm relative to Pb(CH3)4
`(δ = 0 ppm).27 Approximately 12800−51200 scans for 207Pb NMR,
`16−32 scans for 1H NMR, and 500−3000 scans for 13C NMR were
`coadded for the solutions. One-pulse magic-angle-spinning (MAS)
`207Pb NMR spectra for crystalline Pb(aet)2 were acquired with high-
`power proton decoupling on an AVANCE III 200 NMR spectrometer
`at room temperature (207Pb, 41.94 MHz). Ground crystals were
`packed in a 7 mm zirconia rotor, spinning at MAS rates of 5.8 and 5.5
`kHz using 800 and 895 scans, respectively, with a 5.0 s recycle delay.
`Chemical shifts were referenced relative to Pb(CH3)4, by setting the
`207Pb NMR peak of solid Pb(NO3)2 spinning at a 1.7 kHz rate at
`−3507.6 ppm (295.8 K).28,29 Static 207Pb NMR powder patterns were
`reconstructed by iteratively fitting the sideband manifold using the
`Solids Analysis package within Bruker’s TOPSPIN 3.2 software.
`
`2161
`
`DOI: 10.1021/ic5025668
`Inorg. Chem. 2015, 54, 2160−2170
`
`Eton Ex. 1081
`2 of 11
`
`

`

`Inorganic Chemistry
`
`Article
`
`Figure 1. ESI-MS spectra measured in positive-ion mode for solutions A (left) and F (right) (CPbII = 10 mM) with H2Cys/PbII mole ratios 2.1 and
`10.0, respectively. The peak at 122.03 amu has 100% intensity. Selected peaks assigned to lead(II) species with distinct isotopic patterns for Pb are
`shown in the inset.
`
`Pb LIII-edge X-ray absorption spectroscopy (XAS) spectra were
`measured at ambient
`temperature at
`the Stanford Synchrotron
`Radiation Lightsource (SSRL) for freshly prepared solutions A−E
`and A* at BL 7-3 (500 mA; equipped with a rhodium-coated harmonic
`rejection mirror) and for solutions B*−F* at BL 2-3 (100 mA). A
`double-crystal monochromator with Si(220) was used at both
`beamlines. To remove higher harmonics, the monochromator was
`detuned to eliminate 50% of the maximum intensity of the incident
`beam (I0) at the end of the scan at BL 2-3. To avoid photoreduction of
`the samples at BL 7-3, the beam size was adjusted to 1 × 1 mm, and
`the intensity of
`the incident beam was reduced to 80% of
`the
`maximum of I0 at 13806 eV. A high beam intensity could result in
`precipitation of a small amount of black particles in the sample holder,
`especially in solutions containing excess
`ligand. For
`solutions
`containing CPbII = 10 mM, 12−13 scans were measured in both
`transmission and fluorescence modes, detecting X-ray fluorescence
`using a 30-channel germanium detector, while for
`the more
`concentrated solutions with CPbII = 100 mM between three and four
`scans were collected in transmission mode. For each sample,
`consecutive scans were averaged after comparison to ensure that no
`radiation damage had occurred. The energy scale was internally
`calibrated by assigning the first inflection point of a lead foil at 13035.0
`eV. The threshold energy E0 in the XAS spectra of the lead(II)
`cysteine solutions varied within a narrow range: 13034.0−13034.9 eV.
`Least-squares curve fitting of the EXAFS spectra was performed for
`solutions A, B, A*, B*, and F* over the k range = 2.7−11.7 Å−1, using
`the D-penicillaminatolead(II) (PbPen) crystal structure30 as the model
`in the FEFF 7.0 program.31,32 For each scattering path, the refined
`structural parameters were the bond distance (R), the Debye−Waller
`parameter (σ2), and in some cases the coordination number (N). The
`2) was fixed at 0.9 (obtained from
`amplitude reduction factor (S0
`EXAFS data analysis of solid PbPen),33 while ΔE0 was refined as a
`common value for all scattering paths. The accuracy of the Pb−(N/O)
`and Pb−S bond distances and the corresponding Debye−Waller
`parameters is within ±0.04 Å and ±0.002 Å2, respectively. Further
`technical details about EXAFS data collection and data analyses were
`provided previously.26
`Principal component analysis (PCA), introduced in the EXAFSPAK
`suite of programs,34 was applied on the raw k3-weighted experimental
`EXAFS spectra for solutions A−E and A*−F* over the k range of 2.7−
`11.7 Å−1. DATFIT, another program in the EXAFSPAK package, was
`the k3-weighted EXAFS spectra of
`lead(II) cysteine
`used to fit
`solutions A−E and A*−F* to a linear combination of EXAFS
`oscillations for species with PbS2N(N/O), PbS3N, and PbS3
`coordination to estimate the amount of such species in each lead(II)
`theoretical EXAFS
`the PbS3N model,
`cysteine solution. For
`oscillations were simulated by stepwise variation of the Pb−S and
`Pb−N parameters: Pb−S 2.67−2.70 Å [using σ2 = 0.0065 Å2 from
`
`EXAFS least-squares refinement of lead(II) glutathione solutions with
`excess ligand],33 Pb−N 2.40−2.43 Å (σ2 = 0.004, 0.006, and 0.008 Å2),
`2 = 0.9. The best fits were obtained for Pb−S = 2.68 Å (σ2 =
`and S0
`0.0065 Å2) and Pb−N = 2.40 Å (σ2 = 0.0080 Å2).
`■ RESULTS
`ESI-MS. ESI-MS spectra were measured in both positive-
`and negative-ion modes for the lead(II) cysteine solutions A, B,
`and F, with CPbII = 10 mM and H2Cys/PbII mole ratios 2.1, 3.0,
`and 10.0, respectively, as shown in Figures 1 and S-3 in the SI,
`to identify possible charged lead(II) cysteine complexes. The
`assignment of the mass ions, presented in Tables 2 and S-1 in
`
`Table 2. Assignment of Mass Ions Observed in ESI-MS
`Spectra (Positive-Ion Mode) for Lead(II) Cysteine Solutions
`A, B, and F (C
`PbII = 10 mM; H2Cys/PbII Mole Ratios 2.1, 3.0,
`and 10.0, Respectively)a
`
`m/z
`(amu)
`122.03
`144.01
`
`165.99
`252.97
`
`294.00
`
`308.99
`
`311.01
`
`assignment
`[H2Cys + H+]+
`[Na+ + H2Cys]+
`[2Na+ + H2Cys − H+]+
`[Pb(HCOO)]+
`[Pb(H2Cys) − H+ −
`H2S]+
`[3Na+ + 2(H2Cys) −
`2H+]+
`[PbC2H5N3O2]+
`
`m/z
`(amu)
`327.99
`349.97
`
`449.00
`451.99
`
`594.99
`
`635.97
`
`654.96
`676.95
`
`assignment
`[Pb(H2Cys) − H+]+
`[Na+ + Pb(H2Cys) −
`2H+]+
`− H+]+
`[Pb(H2Cys)2
`[4Na+ + 3(H2Cys) −
`3H+]+
`[5Na+ + 4(H2Cys) −
`4H+]+
`[3Na+ + Pb(H2Cys)3
`4H+]+
`− 3H+]+
`[Pb2(H2Cys)2
`−
`[Na+ + Pb2(H2Cys)2
`4H+]+
`
`−
`
`aH2Cys (C3H7NO2S); m /z 121.02.
`
`the SI, is facilitated by the distinct isotopic distribution pattern
`for lead(II) species due to the natural abundance of 52.4%
`208Pb, 22.1% 207Pb, 24.1% 206Pb, and 1.4% 204Pb.35 The ESI-MS
`spectra for solutions A and B were nearly identical, showing
`positive-ion mass peaks for species with metal-to-ligand mole
`ratios 1:1, 1:2, and 2:2. Such mass peaks were also previously
`detected for a 1:1 mixture of Pb(NO3)2 and cysteine in 50%
`methanol/water and considered to be independent of
`the
`reaction mixture stoichiometry (1:10 or 10:1).36 We could also
`− 4H+]+ at 635.97
`detect a 1:3 species [3Na+ + Pb(H2Cys)3
`
`2162
`
`DOI: 10.1021/ic5025668
`Inorg. Chem. 2015, 54, 2160−2170
`
`Eton Ex. 1081
`3 of 11
`
`

`

`Inorganic Chemistry
`
`Article
`
`Figure 2. (left) UV−vis spectra of lead(II) cysteine solutions A−G with CPbII = 10 mM and H2Cys/PbII mole ratios 2.1−15.0 compared with that of
`a 10 mM cysteine solution (dots, pH 9.1). (right) UV−vis spectra of 10 mM lead(II) solutions containing H2Cys/PbII mole ratio 15.0 at pH 9.15
`(solution G) and at pH 8.95 (solution G′) and their difference (G′ − G) compared with that of a lead(II) glutathione solution with GSH/PbII mole
`ratio 10.0 (pH 8.5).33 Data interval = 0.5 nm.
`
`Figure 3. 1H and 13C NMR spectra of 0.1 M cysteine in D2O (pH 9.1) and alkaline lead(II) cysteine solutions (99.9% D2O) with CPbII = 10 mM and
`H2Cys/PbII mole ratios 2.1 (A), 3.0 (B), 4.0 (C), 5.0 (D), 8.0 (E), and 10.0 (F). See Table 1.
`
`amu in the spectrum of solution F. In negative-ion mode, only
`one mass peak corresponding to a lead(II) complex was
`− 3H+]− ion (446.99
`observed, assigned to the [Pb(H2Cys)2
`amu).
`Electronic Absorption Spectroscopy. Figure 2 (left)
`displays the UV−vis spectra for the lead(II) cysteine solutions
`A−G (CPbII = 10 mM). The absorption bands have been
`attributed to a combination of S− 3p → PbII 6p ligand-to-metal
`charge-transfer and PbII intraatomic transitions.37−40 The peak
`maximum for solution A, λ
`max = 298 nm (CH2Cys = 21 mM; pH
`10.4) shows a slight red shift to λ
`max = 300 nm as the ligand
`concentration increases in solution B with H2Cys/PbII mole
`ratio 3.0 at pH 9.1 (Figure S-4a in the SI).
`For solutions E−G with high free ligand concentration (50−
`120 mM), a growing shoulder appears around λ ∼ 330 nm,
`the peak at ∼300 nm reduces
`while the intensity of
`significantly. This shoulder
`is blue-shifted relative to the
`maximum absorption recorded at 335 nm for a lead(II)
`glutathione solution (pH 8.5) containing excess ligand (Figure
`2, right).33 The amplitude of this shoulder is pH-dependent, as
`shown in Figure 2 (right) for 10 mM lead(II) solutions
`
`containing H2Cys/PbII mole ratio 15.0 at pH 9.15 (solution G)
`and pH 8.95 (solution G′). The difference of these two spectra
`(G′ − G) shows that when the pH is lowered by 0.2 units, a
`max in the UV−vis
`peak at 335 nm emerges, very similar to λ
`spectrum of the lead(II) glutathione solution.33 There is no
`true isosbestic point around 312 nm, as shown in Figure S-4b in
`the SI by the systematic movement of crossing points of the
`absorption spectra of solutions B−G with that of solution A.
`1H and 13C NMR Spectroscopy. The 1H and 13C NMR
`spectra of a 0.1 M cysteine solution (pH 9.1) and the lead(II)
`cysteine solutions A−F (CPbII = 10 mM) prepared in D2O are
`shown in Figure 3, with the 1H NMR chemical shifts (δ
`H)
`shown in Table S-2 in the SI. For lead(II)-containing solutions,
`
`−Hc and C1−C3
`only one set of signals was observed for the Ha
`atoms in both PbII-bound and free cysteine because of fast
`ligand exchange on the NMR time scale. These average
`resonances were all
`shifted downfield relative to the
`corresponding peaks in free cysteine (see Table S-2 in the
`SI), with the largest shifts (Δδ) observed for solution A
`(H2Cys/PbII mole ratio = 2.1), which contains the least amount
`of free ligand. Satellites originating from 207Pb nuclei were not
`observed in the 13C NMR spectra.
`
`2163
`
`DOI: 10.1021/ic5025668
`Inorg. Chem. 2015, 54, 2160−2170
`
`Eton Ex. 1081
`4 of 11
`
`

`

`Inorganic Chemistry
`
`Article
`
`Figure 4. 207Pb NMR spectra of the alkaline aqueous lead(II) cysteine solutions A−G enriched in 207Pb (CPbII = 10 mM; H2Cys/Pb2+ mole ratios
`2.1−15.0) and A*−F* with 10% D2O (CPbII = 100 mM; H2Cys/Pb2+ mole ratios 2.1−10.0).
`
`207Pb NMR Spectroscopy. The chemical shift of the 207Pb
`nucleus spans over a wide range (∼17000 ppm). It is sensitive
`to the local structure and electronic environment, the nature of
`surrounding donor atoms,
`the bond covalency, and the
`coordination number and is affected by the temperature and
`concentration.27,28,41−43 We measured 207Pb NMR spectra for
`two sets of alkaline aqueous lead(II) cysteine solutions
`(containing 10% D2O), with increasing H2Cys/PbII mole ratios
`(Table 1). Calculated distribution diagrams based on different
`sets of stability constants indicate that the dominating lead(II)
`complexes would be either [Pb(Cys)2]2− (Figure S-2b in the
`SI),20 or a mixture of [Pb(Cys)2]2− and [Pb(Cys)(HCys)]−
`(Figure S-1b in the SI).16 Figure 4 presents the 207Pb NMR
`spectra for solutions A−G (CPbII = 10 mM; enriched in 207Pb)
`and A*−F* (CPbII = 100 mM), all with only an average NMR
`resonance. Solutions A and A*, both with the H2Cys/PbII mole
`ratio = 2.1 at pH 10.4, show sharp signals at 2006 and 2010
`respectively, which are ∼184−200 ppm deshielded
`ppm,
`relative to that of lead(II) penicillamine (3,3′-dimethylcysteine)
`solutions with similar composition (∼1806−1826 ppm).26 The
`sharpness of this signal results from fast ligand exchange (in the
`NMR time scale) between the lead(II) complexes in solution.
`As the ligand concentration increases in solutions B and B*
`(H2Cys/PbII = 3.0) and the pH changes to 9.1, the 207Pb
`resonance becomes broader and shifts ∼106 (B) and ∼209
`(B*) ppm downfield and more for
`the higher
`ligand
`concentration. Broad averaged signals indicate ligand exchange
`between several lead(II) species at intermediate rates. Solution
`F* containing CPbII = 100 mM and CH2Cys = 1.0 M shows the
`most deshielded 207Pb NMR resonance (2507 ppm), which still
`is ∼286 ppm upfield relative to that of the Pb(S-GSH)3
`complex (2793 ppm) with PbS3 coordination.33
`For comparison, we measured one-pulse MAS 207Pb NMR
`spectra of the crystalline bis(2-aminoethanethiolato)lead(II)
`complex, Pb(S,N-aet)2, at two different spin rates, 5.5 and 5.8
`kHz (see Figures 5 and S-5a in the SI), and observed an
`isotropic chemical shift of δ
`iso = 2105 ppm for this complex
`with PbS2N2 coordination.25 Reconstruction of a static 207Pb
`NMR powder pattern for spin rate 5.8 kHz resulted in the
`following principal components: δ
`11 = 3707.98 ppm, δ
`22 =
`33 = −223.12 ppm, leading to δ
`2831.04 ppm, δ
`iso = 1/3(δ
`11 +
`22 + δ

`33) = 2105.3 ppm (see Figure S-5b in the SI). The
`isotropic chemical shift is ∼600 ppm upfield relative to the only
`other 207Pb chemical shift
`that
`is reported for PbS2N2
`
`Figure 5. One-pulse proton-decoupled MAS 207Pb NMR spectra of
`crystalline Pb(S,N-aet)2, measured at two different spin rates (5.5 and
`5.8 kHz) at room temperature. The dashed vertical line shows the
`isotropic chemical shift δ
`iso = 2105 ppm (identified from overlapping
`spectra in Figure S-5a in the SI).
`
`coordination, δ
`iso = 2733 ppm for Pb(2,6-Me2C6H3S)2(py)2
`(py = pyridine). In that lead(II) complex, all
`ligands are
`monodentate (i.e., not forming a chelate ring) and pyridine is
`the N-donor ligand.44
`We also obtained the 207Pb NMR spectrum of an aqueous
`lead(II) cysteamine solution, prepared by dissolving crystalline
`(mononuclear) Pb(aet)2 in a solution containing the same
`number of moles of cysteamine, with a final
`lead(II)/
`cysteamine mole ratio of 1:3 (10% D2O; CPbII ∼ 76 mM; pH
`10.1). A signal at 2212 ppm was observed (Figure S-6 in the
`SI), probably from a mixture of mononuclear Pb(S,N-aet)2
`(PbS2N2 coordination), [Pb(S,N-aet)(S-Haet)(OH/OH2)]n (n
`= 0, 1; PbS2NO), and [Pb(S,N-aet)(S-Haet)2]+ (PbS3N)
`complexes. A minor amount of PbS3N species is a likely
`reason for the ∼100 ppm deshielding of the 207Pb NMR
`resonance for this solution, relative to the isotropic chemical
`shift of crystalline Pb(aet)2 (δ
`iso = 2105 ppm). Moreover,
`multinuclear species such as the [Pb2(aet)3]+ complex may
`form,45 where PbII ions can adopt PbS3N coordination through
`bridging thiolate groups. However, Li and Martell could only
`identify mononuclear [Pb(Haet)]2+, [Pb(aet)]+, and Pb(aet)-
`(OH) complexes in dilute solutions with CPbII = 1.0 mM (CHaet
`= 1.0−2.0 mM; pH 2−8).46
`Pb LIII-Edge XAS. The near-edge features in the XAS
`spectra are nearly identical for the lead(II) cysteine solutions, as
`shown in Figure S-7 in the SI, and are evidently not sensitive to
`the changes in lead(II) speciation as the ligand concentration
`
`2164
`
`DOI: 10.1021/ic5025668
`Inorg. Chem. 2015, 54, 2160−2170
`
`Eton Ex. 1081
`5 of 11
`
`

`

`Inorganic Chemistry
`
`the
`increases. When the extended regions are compared,
`EXAFS spectra and corresponding Fourier transforms are also
`quite similar for solutions A, B, and E in the CPbII = 10 mM
`series (Figure S-8 in the SI), while for solutions A*−F* (CPbII =
`100 mM), the amplitude of the EXAFS oscillation (and the
`Fourier transform) shows a gradual
`increase as the total
`concentration of cysteine increases from 0.21 to 1.0 M (Figure
`S-8 in the SI). The k3-weighted EXAFS spectrum of
`the
`lead(II) cysteine solution A* (CH2Cys = 210 mM; pH 10.4)
`closely overlaps with that of a lead(II) penicillamine aqueous
`solution (CPbII = 100 mM; H2Pen/PbII = 3.0; pH 9.6), in which
`the dithiolate [Pb(S,N,O-Pen)(S-HnPen)]2−n (n = 0−1) species
`dominates;26 see Figure S-9 in the SI.
`PCA of the raw k3-weighted EXAFS spectra of the six
`lead(II) cysteine aqueous solutions A*−F* and of the five
`solutions A−E displayed two major components with clear
`oscillatory patterns (Figure S-10,
`top,
`in the SI). Using
`TARGET in the EXAFSPAK package on the PCA output file
`showed that
`two of
`the PCA components in each series
`matched fairly well with the raw k3-weighted EXAFS
`oscillations obtained experimentally for 0.1 M lead(II) solutions
`containing penicillamine (CH2Pen = 0.3 M; pH 9.6) or N-
`acetylcysteine (CH2NAC = 1.0 M; pH 9.1). These solutions are
`dominated by PbS2NO and PbS3 species, respectively.26,47 In
`the next step, the DATFIT program in the EXAFSPAK suite
`was used to estimate the relative amounts of such PbS2N(N/
`O)- and PbS3-coordinated species in the lead(II) cysteine
`solutions A*−F*. This was achieved by fitting the k3-weighted
`EXAFS spectra of solutions A*−F* to a linear combination of
`EXAFS oscillations for the lead(II) penicillamine/N-acetylcys-
`teine solutions; however, a clear oscillatory residual was
`observed. That residual amplitude increased when fitting the
`EXAFS spectra of solutions with higher H2Cys/PbII mole
`ratios; see Figure S-11a in the SI,
`indicating a substantial
`additional contribution from another
`lead(II) species (or
`scattering path), probably a four-coordinated PbS3N complex.
`Note that coordinated O and N atoms cannot be distinguished
`by EXAFS spectroscopy, and the PbS2N(N/O) and PbS3N
`coordination modes used here are based on the interpretation
`of 207Pb NMR spectra (see the Discussion section).
`EXAFS oscillations were then theoretically simulated for a
`PbS3N model using the WinXAS48 and FEFF 7.0 programs,
`varying the Pb−S and Pb−N parameters stepwise within
`2 = 0.9 (see the Methods
`reasonable ranges and keeping S0
`section). The best fits with considerably smaller residuals were
`obtained when including the simulated EXAFS oscillation for
`this model in the linear combination, considering Pb−S 2.68 Å
`(σ2 = 0.0065 Å2) and Pb−N 2.40 Å (σ2 = 0.0080 Å2); see
`Figure S-11b in the SI. These fittings clearly show that di- and
`trithiolate lead(II) complexes with PbS2N(N/O) and PbS3N
`coordination give major contributions to the EXAFS spectra of
`the lead(II) cysteine solutions A*−F* (Table 3). Including the
`PbS3-coordinated model in the fitting only slightly improved
`the residual (Figure S-11c in the SI), indicating a minor and
`uncertain contribution (Table S-3 in the SI), especially because
`there are only two major oscillatory PCA components (Figure
`S-10 in the SI, top). Although the EXAFS spectra of solutions
`E* and F* nearly overlap (Figure S-8 in the SI),
`the
`percentages of PbS3N/PbS3 from their linear combination
`fitting results vary about ±10%. Similarly, the PbS3 contribution
`in the EXAFS spectra of solutions A−E is also minor and
`uncertain (Figure S-11e and Table S-3 in the SI). The fittings of
`
`Article
`
`Table 3. Results of Fitting the Raw k3-Weighted Pb LIII-Edge
`EXAFS Spectra of Lead(II) Cysteine Solutions A*−F* and
`A−E with Linear Combinations of EXAFS Oscillations for
`PbS2N(N/O) and PbS3N Models (See the Text; Figures S-
`11b and S-11d in the SI)a
`δ(207Pb)
`solution (H2Cys/PbII mole
`PbS3N
`PbS2NO + PbS2N2
`(%)
`ratio)
`(ppm)
`(%)
`11
`89
`A (2.1)
`2006
`20
`80
`B (3.0)
`2112
`26
`74
`D (5.0)
`2229
`27
`73
`E (8.0)
`2310
`A* (2.1)
`9
`91
`2010
`B* (3.0)
`23
`77
`2219
`C* (4.0)
`31
`69
`2316
`D* (5.0)
`37
`63
`2350
`E* (8.0)
`40
`60
`2418
`F* (10.0)
`49
`51
`2507
`aThe raw k3-weighted EXAFS spectrum of a 0.1 M lead(II)
`penicillamine solution (CH2Pen = 0.3 M; pH 9.6) was used for the
`PbS2N(N/O) model. The EXAFS oscillation for PbS3N was
`theoretically simulated (see the text). The estimated error limit of
`the relative amounts is ±10−15%. For the results when including the
`PbS3 contribution in the fitting, see Table S-3 in the SI.
`
`linear combinations of EXAFS oscillations for the PbS2N(N/
`O) and PbS3N models to the EXAFS spectra of the dilute
`lead(II) cysteine solutions A−E (CPbII = 10 mM) are shown in
`Figure S-11d in the SI, with results in Table 3.
`Least-squares curve fitting of the EXAFS spectra obtained for
`lead(II) cysteine solutions only provides an average of the bond
`distances around the PbII ions in the above species. Thus, a
`coordination model comprising the Pb−(N/O) and Pb−S
`paths was used to simulate the theoretical EXAFS oscillation,
`which was fitted to the extracted EXAFS spectra for the lead(II)
`cysteine solutions A, A* and B, B*, with H2Cys/PbII mole
`ratios 2.1 and 3.0, respectively, for which δ(207Pb) ∼ 2000−
`2220 ppm was observed. The curve-fitting results are displayed
`in Figure 6, with corresponding structural parameters in Table
`4. The average Pb-(N/O) and Pb−S bond distances were 2.42
`
`Figure 6. Least-squares curve fitting of k3-weighted Pb LIII-edge
`EXAFS spectra and corresponding Fourier transforms for lead(II)
`cysteine solutions A and B (CPbII = 10 mM) and A*−F* (CPbII = 100
`mM) with different H2Cys/PbII mole ratios (see Table 4).
`
`2165
`
`DOI: 10.1021/ic5025668
`Inorg. Chem. 2015, 54, 2160−2170
`
`Eton Ex. 1081
`6 of 11
`
`

`

`Inorganic Chemistry
`± 0.04 and 2.64 ± 0.04 Å, respectively, with corresponding
`Debye−Waller parameters varying over the ranges σ2 = 0.014−
`0.024 and 0.0043−0.0061 Å2, respectively. Attempts to resolve
`two different Pb−(N/O) distances resulted in one quite short
`(2.33 Å) and another fairly long (2.50 Å) distances; see model
`III in Table S-4 in the SI. The shortest Pb−N distance found in
`crystal
`structures with PbS2N2 coordination is 2.401 Å
`(Cambridge Structural Database code: NOGQOQ), and the
`shortest recorded Pb−O distance in crystalline PbS2O2
`complexes is 2.349 Å (CSD code: ZOXGAU).33,49
`
`Table 4. Struc

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket