Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 Nov 7.
Published in final edited form as: Chemistry. 2016 Oct 10;22(46):16657–16667. doi: 10.1002/chem.201602901

On the relationship between NMR chemical shifts of thermally polarized and hyperpolarized 89Y complexes and their solution structures

Yixun Xing a, Ashish K Jindal b, Martín Regueiro-Figueroa c, Mariane Le Fur d, Nelly Kervarec e, Piyu Zhao a, Zoltan Kovacs b, Laura Valencia f, Paulo Pérez-Lourido f, Raphaël Tripier d, David Esteban-Gómez c, Carlos Platas-Iglesias c,, A Dean Sherry a,b,
PMCID: PMC5308128  NIHMSID: NIHMS835832  PMID: 27723138

Abstract

Recently developed dynamic nuclear polarization (DNP) technology offers the potential of increasing the NMR sensitivity of even rare nuclei for biological imaging applications. Hyperpolarized 89Y is an ideal candidate because of its narrow NMR linewidth, favorable spin quantum number (I=½) and long longitudinal relaxation times (T1). Strong NMR signals were detected in hyperpolarized 89Y samples of a variety of yttrium complexes. A dataset of 89Y NMR data composed of 23 complexes with polyaminocarboxylate ligands was obtained using hyperpolarized 89Y measurements or 1H,89Y-HMQC spectroscopy. These data were used to derive an empirical equation that describes the correlation between the 89Y chemical shift and the chemical structure of the complexes. This empirical correlation serves as a guide for the design of 89Y sensors. Relativistic (DKH2) DFT calculations were found to predict the experimental 89Y chemical shifts to a rather good accuracy.

Keywords: Yttrium, MRI, Density Functional Calculations, NMR spectroscopy, Hyperpolarization

Graphical abstract

graphic file with name nihms835832u1.jpg

The 89Y NMR chemical shifts measured for a wide range of Y3+ complexes can be estimated from a set of empirical contributions associated to the different types of donor atoms coordinated to the metal ion. This correlation will serve as a guide for the design of Magnetic Resonance Imaging sensors based on hyperpolarized 89Y.

Introduction

Magnetic resonance imaging (MRI) is a technique widely used in medical diagnosis that is based on the same fundamental principles as NMR. Traditional MRI scanners detect the NMR signal of water proton nuclei present in the body, image contrast being achieved by using adequate pulse sequences that take advantage of differences in proton densities and longitudinal and transverse relaxation times among tissues.[1] MRI provides 3D images of the body with very high resolution, but is traditionally regarded as a rather insensitive technique.[2] In an external magnetic field B0, the extent to which a nuclear spin becomes polarized is described by:[3]

P=(N+N)(N+N) [1]

Where N+ and N- denote the number of nuclear spins parallel (“spin up”) and anti-parallel (“spin down”) with respect to B0. The polarization at Boltzmann equilibrium (P0) is given by:[4]

P0=tanh(γB02kBT)γB02kBT [2]

Where γ is the gyromagnetic ratio of the nucleus under study, kB is the Boltzman constant and other symbols have their usual meaning. Eq. 2 shows that spin polarization increases with field so the tendency has been for human MRI scanners to move to higher fields over the years to gain sensitivity. Eventually, this becomes cost prohibitive so alternative ways to increase NMR sensitivity are being explored. One method is to use hyperpolarized (HP) materials that exist, at least temporarily, in a non-equilibrium hyperpolarized state where (N+-N-) is increased several orders of magnitude over thermal equilibrium. Dynamic nuclear polarization (DNP), one of the more popular methods to produce hyperpolarized materials, has been used to amplify the sensitivity of 13C and 15N nuclei to a level where common metabolites can be imaged by MRI. For example, 13C polarization levels as high as 50% have been achieved using DNP at low temperatures and high fields.[5] The experimental protocol involves doping the 13C material with a stable free radical such as trityl or nitroxide, freezing the sample in a solvent that forms a glass, and irradiating the sample with continuous microwave irradiation at a frequency close to the corresponding electron resonance frequency.

Recently Sherry et al. reported the use of DNP methods to hyperpolarize a variety of 89Y complexes.[6-9] Yttrium exists in nature 100% as the stable isotope, 89Y. The predominant chemical form of yttrium is the diamagnetic, trivalent ion, Y3+. 89Y has a much smaller nuclear magnetic moment than 13C so its NMR sensitivity is only 0.668 relative to 13C.[10] Consequently, it is quite difficult and time consuming to acquire an 89Y NMR spectrum at Boltzmann spin levels except for highly concentrated samples. However, 89Y3+ has a nuclear spin quantum number of I=½, exhibits narrow NMR linewidths and long longitudinal relaxation times (T1) so it is an attractive nucleus for hyperpolarization. In addition, yttrium is not naturally present in the human body so there is zero background signal. The effective ionic radius of Y3+ (1.019 Å for CN8) is similar to that of Gd3+ (1.053 Å)[11] so any ligand used to complex Gd3+ will also form stable complexes with Y3+.[12,13] Y3+ complexes can be also used as models for investigating the solution behaviour of the corresponding paramagnetic lanthanide complexes.[14]

Only a single 89Y NMR report existed before 1970 likely due to the extremely low NMR sensitivity of this nucleus.[15] In the late 1970s the first T1 values for some 89Y complexes were found to be unusually long.[16] In 2007, three 89Y complexes, Y-DOTP, Y-DOTA, Y-DTPA, plus YCl3 were hyperpolarized using DNP methods and NMR signals were obtained on complexes at concentrations of only 5-15 mM after a single scan.[6] The enhanced signals showed that yttrium complexes could be polarized to levels reasonable for imaging. Furthermore, the various 89Y complexes displayed a wide range of chemical shifts. This observation in combination with the long relaxation times of these complexes makes them potentially attractive as probes for various physiological parameters such as pH or redox environment. This preliminary study triggered our interest in conducting hyperpolarization experiments on a broader range of 89Y complexes. We report here 89Y NMR spectra of both thermally polarized and hyperpolarized samples of the diverse group of Y3+ complexes shown in Scheme 1 and Scheme 2. Subsequently, density functional theory (DFT) calculations were conducted to better understand the factors that affect the observed 89Y NMR chemical shifts and eventually establish empirical correlations between NMR chemical shifts and the number and types of donor atoms in each complex, which might serve as a guide for designing new 89Y complexes as biosensors.

Scheme 1.

Scheme 1

Complexes with macrocyclic ligands investigated in this work.

Scheme 2.

Scheme 2

Complexes with non-macrocyclic ligands investigated in this work.

Results and Discussion

Structure of the Y3+ complexes

Establishing a correlation between the observed 89Y NMR chemical shifts and the coordination environment around the metal ion requires an accurate determination of the structure in solution of the concerned complexes. Given the similar coordination properties of Y3+ and the lanthanide ions (Ln3+) we sought to expand the family of structurally characterized Y3+ complexes by taking advantage of our recent studies on Ln3+ complexes. Thus, we have included in our studies a series of macrocyclic-based ligand complexes, [Y(Me2DO2PA)]+, [Y(CB-TE2PA)]+, [Y(P2C14Et4)]3+, [Y(P2C14Py4)]3+ and [Y(BP18C6)]+ (Scheme 1), where the structures of the corresponding Ln3+ complexes have been established by X-ray diffraction in the solid state and/or by analysis of the Ln3+-induced paramagnetic 1H NMR shifts in aqueous solution.[17-25] The X-ray structure obtained using single crystals with formula [Y(P2C14Et4)](NO3)3·6H2O confirm that the [Y(P2C14Et4)]3+ complex retains the ten-coordinate structure observed for the Ln3+ series along the whole lanthanide series (Figure 1).[20,25] Similarly, the [Y(Me2DO2PA)]+ and [Y(CB-TE2PA)]+ complexes adopt eight-coordinate structures in the solid state very similar to those established for the corresponding complexes with the Ln3+ ions (Figures 2 and 3).[18,19,21] The family of Y3+ complexes with macrocyclic ligands investigated in this work also includes [Y(DOTA)(H2O)]-, [Y(DOTAM)(H2O)]3+ and [Y(DO3A)(H2O)2]. The structure of [Y(DOTA)(H2O)]- in the solid state has been established by X-ray diffraction,[26,27] while the structure of Ln3+ complexes of these ligands have been also investigated both in the solid state and in solution.[28-33] The Gd3+ complex of DO3A was shown to be present in solution as a hydration equilibrium involving an eight-coordinated [Gd(DO3A)(H2O)] complex and a nine-coordinated bis-hydrated [Gd(DO3A)(H2O)2] complex. The speciation in solution was found to be dominated by the bis-hydrated form, which results in an overall hydration number of 1.8-1.9 at 25 °C.[34,35]

Figure 1.

Figure 1

View of one of the complex cations present in crystals of [Y(P2C14Et4)](NO3)3·6H2O. Water molecules, anions and hydrogen atoms (except those of the hydroxyl groups) have been omitted for simplicity. Bond distances (Å): Y1-O1, 2.435(2); Y1-O2, 2.442(2); Y1-N1, 2.562(2); Y1-N2, 2.628(3); Y1-N3, 2.657(3).

Figure 2.

Figure 2

View of one of the complex cations present in crystals of [Y(CB-TE2PA)](PF6)·H2O. Water molecules, anions and hydrogen atoms have been omitted for simplicity. Bond distances (Å): Y1-O1, 2.275(4); Y1-O3, 2.259(4); Y1-N1, 2.590(5); Y1-N2, 2.526(5); Y1-N3, 2.607(5); Y1-N4, 2.520(4); Y1-N5, 2.459(5); Y1-N6, 2.470(5).

Figure 3.

Figure 3

View of one of the complex cations present in crystals of [Y(Me2DO2PA)](PF6)·H2O. Water molecules, anions and hydrogen atoms have been omitted for simplicity. Bond distances (Å): Y1-O1, 2.272(2); Y1-N1, 2.565(2); Y1-N2, 2.569(3); Y1-N3, 2.425(2).The complex presents a crystallographically imposed C2 symmetry.

The chemical formula of all Y3+ complexes with non-macrocyclic ligands included in this study are shown in Scheme 2. The number of inner-sphere water molecules in some of the complexes was assigned on the basis of available crystallographic data and solution studies performed on the Ln3+ analogues ([Y(EGTA)(H2O)]-,[36,37] [Y(DTPA)(H2O)]2-,[38] [Y(TTHA)]3-39]). In other cases the complexes present variable hydration numbers in the solid state so the exact number of water molecules coordinated to Y3+ cannot be predicted in a straightforward manner. For instance, the sodium salt of [Y(EDTA)(H2O)3]- was found to contain three coordinated water molecules, while the structurally related [Y(DCTA)(H2O)2]- complex was reported to be bis-hydrated in the solid state.[40] The number of coordinated water molecules in these complexes was assessed by using DFT calculations, which provided theoretical 89Y NMR shifts using different plausible coordination numbers. The comparison of experimental and calculated shifts allowed an unequivocal assignment of the hydration state (see below). Most ligands included in this study are known to form 1:1 complexes with the trivalent Ln3+ ions, except for NTA3- and NTPA3-, which can form both 1:1 and 1:2 complexes, or HIDA2- and IMDA2-, which form 1:2 and 1:3 complexes, respectively.[10]

89Y NMR chemical shifts

89Y NMR spectra of hyperpolarized samples of Y3+-EDTA and Y3+-EGTA are illustrated in Figure 4. These spectra demonstrate that one can obtain high quality 89Y NMR spectra relatively easily. The 89Y chemical shifts of seven complexes measured on hyperpolarized samples are listed in Table 1. Thermally polarized 89Y spectra have been reported previously for Y3+-(DTPA), Y3+-EGTA, Y3+-(NTA)2, Y3+-EDTA and Y3+-TTHA, and their chemical shifts agree well with those measured from spectra of hyperpolarized samples.[10,41]

Figure 4.

Figure 4

NMR spectra of hyperpolarized (a) Y3+-EDTA (6.1 mM) and (b) Y3+-EGTA (7.7 mM) at pH 7.4 collected at 9.4T using a single 20° pulse.

Table 1.

89Y chemical shifts and T1 values for seven hyperpolarized Y3+ complexes.[a]

Y3+ complex Concentration (mM) Chemical shift (ppm) T1 (s)
EGTA 7.7 68.6 474 ± 13
(NTA)2 7.2 73.0 534 ± 118
DTPA 6.2 76.0 227 ± 5
DO3A 7.2 103.5 247 ± 91
EDTA 6.1 133.5 310 ± 4
DOTA 12.5 111.3 446 ± 7
TTHA 7.7 96.9 245 ± 90
103.2 230 ± 14
a

∼0.2M samples dissolved in H2O/glycerol (3/1) plus 15 mM OX63 were frozen at 1.4K and polarized for 3 h in an Oxford DNP HyperSense system. The samples were quickly dissolved in a buffer and transferred to a 10 mm NMR tube for data collectionusing a 9.4T magnet. 89Y spectra were collected using 20° pulses every 20 s over a period of 300-600 s and the resulting data were fitted to standard decay curve for polarized samples. All chemical shifts are reported relative to YCl3 set to 0 ppm.

Another set of 89Y NMR shifts was obtained using 1H,89Y heteronuclear shift correlation through scalar coupling (1H,89Y-HMQC). This technique allows a much faster acquisition compared with conventional pulse-acquisition 89Y NMR measurements, which are very time-consuming due to the long T1 relaxation times of 89Y (Table 1). 1H,89Y-HMQC spectroscopy allows an indirect 1H detection of 89Y, thereby overcoming this problem.[42] The representative example of the 1H,89Y-HMQC spectrum of [Y(P2C14Et4)]3+ is shown in Figure 5. Several cross-peaks relating the 89Y NMR signal of the complex at 25.7 ppm and aliphatic protons of the ligand situated at a three-bond distance are observed. A weak correlation involving the pyridyl proton H2, situated four bonds away the metal ion, is also evident. The strongest correlations involve H4ax, H5ax, H7eq and H7ax, which according to the X-ray crystal structure present H-C-N-Y or H-C-O-Y dihedral angles of 80.6, 78.4, 104.1 and 136.9°. However, no cross-peaks are observed in the case of protons H4eq ad H5eq, which are characterized by dihedral angles of 162.2 and 163.5°. These results suggest that the intensity of the scalar coupling between 1H and 89Y follows a Karplus-like relationship, as observed for the 1H scalar coupling constants responsible for the contact shifts induced by paramagnetic Ln3+ ions.[19]

Figure 5.

Figure 5

1H,89Y-HMQC NMR spectrum of [Y(P2C14Et4)]3+ recorded in D2O solution(4.7 mM).

The 89Y NMR chemical shift data of the complexes investigated in this work are compiled in Table 2. The yttrium chemical shifts of our data set composed of 23 polyamino-polycarboxylate complexes vary over 270 ppm. Particularly surprising are the shifts of [Y(Me2DO2PA]+ (245.8 ppm) and [Y(CB-TE2PA)]+ (218.5 ppm) and the negative chemical shift of [Y(BP18C6)]+ (-23.6 ppm), which expand the 89Y NMR chemical shift range of polyaminocarboxylate complexes by ∼200 ppm with respect to the window of 89Y NMR shifts reported previously (∼70 ppm).[10,41]

Table 2.

Measured chemical shifts and predicted chemical shifts from (3) of Y3+ complexes with the numbers of coordinated amine nitrogen atoms (nNam), pyridine nitrogen atoms (nNpy), carboxylate oxygen atoms (nOc), ether oxygen atoms (nOe), amide oxygen atoms (nOa) and water oxygen atoms (nOw).

Y3+ complex δexp (ppm) δcalc (ppm) nNam nNpy nOc nOe nOh nOa nOw
[Y(DOTA)(H2O)]- (1) 111.3 107.0 4 0 4 0 0 0 1
[Y(DOTAM)(H2O)]3+ (2) 123.0[a] 125.1 4 0 0 0 0 4 1
[Y(DO3A)(H2O)2] (3) 103.5 93.4 4 0 3 0 0 0 2
[Y(P2C14Et4)]3+ (4) 25.7 29.8 4 2 0 0 4 0 0
[Y(P2C14Py4)]3+ (5) 78.6 76.0 4 6 0 0 0 0 0
[Y(DOTA-(gly)4)(H2O)]- (6) 121.6 125.1 4 0 0 0 0 4 1
[Y(DOTAM-2C)(H2O)]- (7) 120.9 125.1 4 0 0 0 0 4 1
[Y(DOTAM-3C)(H2O)]- (8) 118.7 125.1 4 0 0 0 0 4 1
[Y(DOTAM-5C)(H2O)]- (9) 117.8 125.1 4 0 0 0 0 4 1
[Y(Me2DO2PA)]+ (10) 244.7 231.0 4 0 0 0 0 0 0
[Y(CB-TE2PA)]+ (11) 218.5 231.0 4 0 0 0 0 0 0
[Y(BP18C6)]+ (12) -23.6 -15.3 2 0 0 4 0 0 0
[Y(TTHA)]3- (13) 103.2 120.6 4 0 5 0 0 0 0
[Y(IMDA)3]3- (14) 98.7[b] 94.7 3 0 6 0 0 0 0
[Y(DTPA)(H2O)]2- (15) 76.0, 82.2,[b] 83.0[c] 81.1 3 0 5 0 0 0 1
[Y(EDTA)(H2O)2]- (16) 133.5, 129.6[b] 135.7 2 0 4 0 0 0 2
[Y(DCTA)(H2O)2]- (17) 132.7[b] 135.7 2 0 4 0 0 0 2
[Y(HEDTA)(H2O)2] (18) 78.6[b] 76.0 2 0 3 0 1 0 3
[Y(DPTA)(H2O)2]- (19) 130.0[b] 135.7 2 0 4 0 0 0 2
[Y(EGTA)(H2O)]- (20) 68.6, 69.4,[c]68.5[b] 52.0 2 0 4 2 0 0 1
[Y(HIDA)2]- (21) 57.0[b] 48.6 2 0 4 0 2 0 1
[Y(NTA)2]3- (22) 73.0, 76.5[b] 162.8 2 0 6 0 0 0 0
55.2 2 0 6 0 0 0 1
[Y(NTPA)2]3- (23) 90.5[b] 162.8 2 0 6 0 0 0 0
55.2 2 0 6 0 0 0 1
[Y(NTA)] (24) 62.6[b] 82.5 1 0 3 0 0 0 4
-25.1 1 0 3 0 0 0 5
[Y(NTPA)] (25) 44.0[b] 82.5 1 0 3 0 0 0 4
-25.1 1 0 3 0 0 0 5
[Y(H2O)8]3+ (26) 0.0 2.5 0 0 0 0 0 0 8
SNam SNpy SOc SOe SOh SOa SOw
68.1+6.4 85.7+5.4 94.0+5.8 95.7+4.9 97.3+4.9 89.5+5.4 107.6+6.2
a

Data from reference [45].

b

Data from reference [10].

c

Thermally polarized data (this work).

The sensitivity of the 89Y NMR chemical shift to the ligand field produced by the chelate has been examined before for a series of organometallic complexes,[43] but few studies have attempted to relate 89Y chemical shifts with coordination number and identity of donor atoms in ligands that could potentially serve as reporter molecules in biological systems. An empirical relationship between the number of coordinated oxygen and nitrogen atoms and 139La chemical shifts was reported previously.[44] In that study, each negatively charged oxygen atom contributed 30 ppm to the chemical shift of 139La, while each ligand nitrogen atom contributed 50 ppm. This same empirical relationship does not hold for yttrium(III) complexes because the range of NMR chemical shifts observed for the two trivalent ions is quite different. Nevertheless, if one makes the same assumption that each ligand nitrogen atom contributes equally and each oxygen atom contributes equally to the 89Y chemical shift, the 89Y chemical shifts of some of the yttrium(III) complexes listed in Table 2 are rather poorly predicted. An inspection of the 89Y chemical shift data reported in Table 2 shows that generally ten-coordinated complexes (i. e. [Y(BP18C6)]+ and [Y(P2C14Et4)]3+) give signals at higher fields, while the89Y resonances of eight-coordinate complexes are more deshielded and nine coordinated complexes give signals in an intermediate range. Furthermore, the replacement of inner sphere water molecules in [Y(EDTA)(H2O)2]- for ligands with different oxygen donor atoms as lactate affects drastically the value of the 89Y NMR chemical shift (Figure S8, Supporting Information). This supports the use of different shielding constants for ligands having oxygen donor atoms of water, hydroxyl or carboxylate groups. In the other hand, the 89Y NMR chemical shifts measured for thermally polarized samples of complexes [Y(DOTAM-2C)]-, [Y(DOTAM-3C)]- and [Y(DOTAM-5C)]- are very similar to that of [Y(DOTA-(gly)4)]- (Table 2, Figure S9, Supporting Information). The varying lengths of the 4 side chains of the ligand have an effect on the inversion of the population of the twisted square antiprismatic (TSAP) and square antiprismatic (SAP) diastereomers. Therefore, it turns out that chemical shifts are mainly dependent of contribution from the coordinated atoms rather than other structural factors. Thus, we attempted to fit the experimental 89Y NMR data to an empirical relationship of the form:

δcalc(89Y)=A(SNamnNam+SNpynNpy+SOcnOc+SOhnOh+SOenOe+SOanOa+SOwnOw) [3]

where A is an empirical constant that can be regarded as the chemical shift of Y3+ in the absence of any ligand, SNam, SNpy, SOc, SOh, SOe, SOa and SOw represent the shielding contribution of amine, pyridine, carboxylate, hydroxyl, ether, amide and water donor atoms and nNam, nNpy, nOc, nOh, nOe, nOa and nOw are the number of donor atoms of each class. The use of different shielding constants for similar donor atoms such as amide or carboxylate oxygen atoms is justified by the different chemical shifts observed for [Y(DOTA)(H2O)]- (111 ppm) and [Y(DOTAM)(H2O)]3+ (123 ppm),[45] which indicate that oxygen atoms of carboxylate groups provide a slightly more important contribution to the shielding of the 89Y NMR resonance than amide oxygen atoms. The best fit of the data provided A = 863 ± 48 ppm and the shielding constants reported in Table 2.

The shielding effect of an amine nitrogen atom (SNam = 68.1 ppm) is considerably smaller than that of pyridyl nitrogen atoms (SNpy = 85.7 ppm). Oxygen donor atoms of organic ligands provoke a rather similar shielding effects (94.0-97.3 ppm), while the shielding contribution of water molecules in significantly higher (SOw = 107.6 ppm). Figure 6 shows a plot of experimental versus calculated 89Y NMR shifts obtained with equation (3) for 21 complexes. The shifts of NTA3- and NTPA3- complexes were found to deviate considerably from the predicted values, and were omitted from Figure 6. The reasons for this discrepancy are likely related to the presence of hydration equilibria in solution, as discussed below. Nevertheless, the error between the calculated versus experimental data for the 21 complexes is relatively small, ±6.9 ppm.

Figure 6.

Figure 6

Plot of experimental 89Y chemical shifts versus those predicted by equation (3). The straight line is theidentity line.

Given the good agreement between the observed and calculated 89Y NMR shifts for most systems, the noticeable deviations from the experimental and calculated data of [Y(NTA)2]3- and [Y(NTPA)2]3- are intriguing. In the case of [Y(NTA)2]3-, equation (3) predicts a shift of 55.2 ppm assuming the presence of a coordinated water molecule and 162.8 ppm if the complex lacks inner-sphere water molecules. The experimental shifts (73.0 ppm) is closer to the value calculated for the q = 0 species. In the case of [Y(NTPA)2]3- the donor atoms of the ligands are identical to those of [Y(NTA)2]3-, leading to the same calculated values. However, the experimental value is quite different (90.5 ppm). We interpret these data in terms of hydration equilibria involving a non-hydrated species and a complex species containing one coordinated water molecule. The calculated 89Y NMR shifts suggest that the major species in solution corresponds to the hydrated complex, the population of the q = 0 species being estimated as ∼15 and ∼30% for [Y(NTA)2]3- and [Y(NTPA)2]3-, respectively. Similar hydration equilibria involving complex species containing four and five coordinated water molecules are probably present in solutions of the [Y(NTA)] and [Y(NTPA)] complexes (Table 2).

The [Y(TTHA)]3- complex presents at least two species in solution at about neutral pH, as two signals at 103.2 ppm and 96.9 ppm are observed in the hyperpolarized 89Y NMR spectrum at pH 6.9 (Table 1). The Ln3+ complexes of TTHA6- form different structures in the solid state depending upon the size of the metal ion.[46] The smallest Ln3+ ions form nine-coordinated complexes through binding of four amine nitrogen atoms and five oxygen atoms of carboxylate groups, the remaining carboxylate unit remaining uncoordinated. For the largest Ln3+ ions ten-coordinate species were observed in the solid state thanks to the coordination of the four amine nitrogen atoms and the six carboxylate oxygen atoms. Additionally, dinuclear [Ln2(HTTHA)2]4- species containing nine-coordinated metal ions were also reported, with the metal coordination environment being fulfilled by three amine nitrogen atoms, four carboxylate groups of one of the ligands and two carboxylate groups of a neighboring ligand unit. The chemical shift values calculated with equation (3) for the nine- and ten-coordinated [Y(TTHA)]3- species are 120.6 ppm and 26.6 ppm, respectively. The calculated value for the dinuclear [Ln2(HTTHA)2]4- species is 94.7 ppm. The latter value is in excellent agreement with the NMR signal observed at 96.9 ppm. Thus, most likely the Y3+-TTHA complex is present in solution as a mixture of a mononuclear and a dinuclear species giving 89Y NMR signals at 103.2 and 96.9 ppm, respectively. A minor amount of ten-coordinate [Y(TTHA)]3- species might be responsible for the 17.5 ppm difference between the experimental and calculated shifts of [Y(TTHA)]3-.

DFT calculations

Aiming to have a deeper insight on the origin of the 89Y NMR chemical shifts observed for polyaminocarboxylate complexes, we turned our attention to theory. Prior to calculation of 89Y NMR chemical shifts using DFT methods, we first performed geometry optimizations of 16 Y3+ complexes with polyaminocarboxylates using the TPSSh functional and a quasirelativistic pseudopotential that includes the inner 28 electrons in the core (see computational details below). For those complexes containing coordinated water molecules, we explicitly included two second-sphere water molecules hydrogen-bonded to each coordinated water molecule, as the inclusion of second-sphere water molecules was found to be critical for an accurate calculation of Gd3+-Owaterdistances and 17O hyperfine coupling constants in Gd3+ complexes of polyaminocarboxylates.[47,48] The optimized geometries of all Y3+ complexes together with the calculated bond distances of the Y3+ coordination environments are shown in Figures S10-S11, Supporting Information). The quality of the optimized structures was assessed by comparison of the calculated bond distances of the metal coordination environment with crystallographic data (Tables S1-S3, Supporting Information).

Since Y3+ is a relatively heavy atom, relativistic effects are expected to play an important role on the 89Y NMR shifts.[49] Thus, the absolute shielding constants were obtained using relativistic DFT calculations at the TPSSh/DKH2/TZVP level. A plot of the isotropic nuclear shielding values (σiso) versus the observed 89Y NMR shifts provides a rather good linear correlation (Figure 7). Furthermore, the chemical shifts obtained using the σiso value calculated for [Y(H2O)8]3+ are also in reasonable agreement with the experimental values (Table 3). Besides, these calculations allowed us to assess the hydration number of those complexes for which the number of coordinated water molecules was uncertain. One of such cases is [Y(DO3A)(H2O)q], for which the 89Y NMR shift calculated for the bis-hydrated complex (103.4 ppm, Table 3) is virtually identical to the experimental one (103.5 ppm). On the contrary, the mono-hydrated complex yields a calculated chemical shift of 181.8 ppm, a value that deviates ∼78 ppm from the experimental one. For [Y(EDTA)(H2O)q]-, the DFT calculations predict 89Y NMR shifts of 20.5 and 119.4 ppm for the tris- and bis-hydrated forms, the latter value being in reasonably good agreement with the experimental one (135.7 ppm). However, for [Y(NTA)2(H2O)q]3- the chemical shifts calculated for both the q = 1 (24.3 ppm) and q = 0 (101.8 ppm) species deviate considerably from the experimental value (73.0 ppm), suggesting that a hydration equilibrium exists in solution involving a mono- and a bis-hydrated species. The presence of hydration equilibria in solution was ascertained using absorption spectroscopy for different Eu3+ complexes, and therefore it is likely that some Y3+ complexes will show similar behavior.[50,51]

Figure 7.

Figure 7

Plot of the experimental 89Y NMR shifts versus the isotropic nuclear shielding values calculated in aqueous solution (COSMO) at the TPSSh/DKH2/TZVP level.

Table 3.

Isotropic 89Y nuclear shielding values (σiso), diamagnetic (σd) and paramagnetic (σp) contributions to σiso and 89Y NMR shifts calculated in aqueous solution (COSMO) at the TPSSh/DKH2/TZVP level.

Y3+ complex σiso σd σp δcalc
[Y(DOTA)(H2O)]- (1) 2843.6 4469.5 -1625.8 96.4
[Y(DOTAM)(H2O)]3+ (2) 2827.6 4464.8 -1637.6 112.4
[Y(DO3A)(H2O)2] (3) 2836.6 4464.9 -1628.3 103.4
[Y(P2C14Et4)]3+ (4) 2899.6 4521.6 -1622.0 40.4
[Y(Me2DO2PA)]+ (10) 2727.2 4461.4 -1734.2 212.8
[Y(CB-TE2PA)]+ (11) 2729.9 4494.0 -1764.2 210.1
[Y(BP18C6)]+ (12) 2977.3 4547.7 -1570.4 -37.3
[Y(TTHA)]3- (13) 2871.1 4489.1 -1618.0 68.9
[Y(IMDA)3]3- (14) 2852.3 4378.4 -1526.1 87.7
[Y(DTPA)(H2O)]2- (15) 2894.9 4446.1 -1551.2 45.1
[Y(EDTA)(H2O)2]- (16) 2820.6 4386.6 -1565.9 119.4
[Y(DCTA)(H2O)2]- (17) 2810.8 4440.9 -16300 129.2
[Y(DPTA)(H2O)2]- (19) 2810.9 4407.4 -1596.5 129.1
[Y(EGTA)(H2O)]- (20) 2898.0 4437.5 -1539.5 42.0
[Y(HIDA)2]- (21) 2882.8 4406.6 -1523.8 57.2
[Y(H2O)8]3+ (26) 2940.0 4060.5 -1120.6 0.0

The calculated σiso values provide a rather good linear correlation with the calculated electron density at the Y3+ nucleus (Figure S12, Supporting Information). The σiso values can be broken down into the diamagnetic (σd) and paramagnetic (σp) contributions, which were obtained using each nucleus as the gauche origin for separation (Table 3).[52] The diamagnetic contribution arises from the magnetic field at the nucleus produced by the diamagnetic current density caused in the electrons by the external magnetic field.[53] Excluding the aquated ion, for the series of complexes listed in Table 3 the diamagnetic contribution varies by up to ∼169 ppm. The σd values calculated for the complexes with polyaminocarboxylate ligands are larger than that calculated for [Y(H2O)8]3+, which shows that the diamagnetic contribution provides a shielding effect to the 89Y NMR shifts. The analysis of the natural electron configurations obtained for Y3+ in this series of complexes with Natural Bond Orbital (NBO) analysis provides some insight into the nature of this shielding effect (Table S4, Supporting Information). The natural electron configurations reveal significant changes in the 5s and 4d populations depending upon the metal coordination environment. The 5s populations are relatively small and vary from 0.162 for [Y(EDTA)(H2O)2]- to 0.186 for [Y(IMDA)3]3-, while the calculated 4d populations present values in the range 0.82-1.08. The 5p populations are smaller (< 0.014) than the 5s and 4d populations, as would be expected. Plots of the diamagnetic shielding contribution versus the calculated 4d, 5s and 5p populations (Figure S13, Supporting Information) reveal linear trends that indicate increasing shielding effects upon increasing the population of the Y3+ valence orbitals.

The paramagnetic contribution accounts for the mixing of certain excited states with the electronic ground state in the presence of a magnetic field. The calculated σp values for complexes with polyaminocarboxylate ligands are more negative than that obtained for [Y(H2O)8]3+, indicating that the paramagnetic contribution results in a deshielding effect in the family of Y3+ complexes investigated here. The σp term generally becomes more important as the HOMO-LUMO gap is reduced (Figure S14, Supporting Information). This is expected as smaller HOMO-LUMO gaps tend to decrease the energy differences between all occupied and unoccupied orbitals.[54]

Conclusions

This study showed that Y3+ complexes with polyaminocarboxylate ligands provide a rather wide range of 89Y NMR chemical shifts that expands ∼270 ppm, which represents an expansion of about 200 ppm with the chemical shift range observed previously. The 89Y NMR chemical shifts can be obtained either using hyperpolarized samples or 1H,89Y-HMQC NMR experiments to overcome the long acquisition times required for traditional 89Y NMR experiments due to the long relaxation times. The 89Y NMR chemical shift values were found to change considerably depending upon the coordination number of the complex and the nature of the donor atoms of the ligand. A relativistic DFT study was conducted to rationalize the origin of the 89Y NMR chemical shifts. These calculations can predict the observed chemical shifts to a rather good accuracy. The breakdown of the calculated isotropic shielding constants revealed that the calculated 89Y NMR chemical shifts are a subtle balance of the two contributions. Generally increased 4d, 5s and 5p populations and large HOMO-LUMO gaps have a shielding effect on the 89Y chemical shifts. However, these general trends do not allow a prediction of the chemical shifts without performing the DFT calculations. Thus, we proposed an empirical relationship that provides calculated chemical shifts to an accuracy of about ±6.9 ppm with respect to the experimental ones. This empirical expression will be useful for the rational design of 89Y NMR imaging probes for hyperpolarized imaging. Furthermore, the 89Y NMR chemical shifts are also useful to gain information on the solution structure of the corresponding complexes with Ln3+ ions.

Experimental and Computational Section

General

All chemicals used were of the highest available purity and were not purified further. Solvents used were available commercially and used as received. Hexahydrated yttrium nitrate was obtained from Aldrich. Syntheses performed under microwave radiation were carried out with a Monowave Anton Paar 300 instrument (30 bar/850W). Elemental analyses were performed in a Carlo-Erba EA 1108 microanalyzer. Attenuated total reflection infrared (ATR-FTIR) spectra were recorded on a FP-6100 Jasco spectrometer. ESI-MS experiments of complexes [Y(P2C14Et4)](NO3)3·6H2O and [Y(P2C14Py4)](NO3)3·6H2O were performed on an microTOF(focus) mass spectrometer (Bruker Daltonics, Bremen, Germany). Ions were generated using an ApolloII (ESI) source and ionization was achieved by electrospray. ESI-MS analyses of the remaining complexes were done either at HT Laboratory (San Diego, CA) or at core facilities at UT Southwestern Medical Center at Dallas. The pH values were direct readings as measured on an Orion 720A+ (Thermo Corporation, city, state) or an Ultra Basic 5 pH meter (Denver Instruments, city state).

NMR spectroscopy

One dimensional 1H NMR spectra of compounds [Y(P2C14Et4)](NO3)3·6H2O, [Y(P2C14Py4)](NO3)3·6H2O, [Y(Me2DO2PA)]Cl and [Y(CB-TE2PA)]Cl were recorded at 25°C on a BRUKER Avance III HD 500 spectrometer equipped with an indirect 5mm triple resonance broad band (TBI) 1H/ {BB} /13C probe. Spectra were performed according BRUKER's pulse programs with standard pulse sequences using delay times of 2s, a 30° pulse and 16 scans. The 89Y chemical shifts were obtained using Heteronuclear Multiple Quantum Coherence 1H-89Y (HMQC) experiments with a delay for evolution of long range couplings of 41.66 ms and a 2s delay for relaxation time. For example, in an HMQC experiment the raw data set consisted of 48 fid of 1024 (F2) * 98(F1) complex data points each zero filled to 1k in the F1 dimension prior to Fourier transform with a spectral width of 31846 Hz and 4504 Hz in the F1 and F2 dimensions, respectively. 1H Chemical shifts were expressed in ppm relative to TSP as external reference, in 99.97% D2O, and TMS for others solvents. 89Y chemicals shifts were expressed in ppm relative to Yttrium nitrate, hexahydrate 1M, pH = 4.3 as external reference. The one dimensional 89Y spectrum reference was recorded with a 50 s delay relaxation time, 976 scans, and a spectral window of 5980 Hz. 1H,13C, and thermally polarized 89Y NMR spectra of all other complexes were recorded using a Bruker Avance III 400 MHz spectrometer and a 5 mm sample probe. Hyperpolarized 89Y NMR spectra were collected at a field strength of 9.4 T at 25° on an Oxford unshielded 89 mm magnet with Varian 10 mm low gamma probe. Both hyperpolarized and thermally polarized 89Y NMR spectra were referenced to YCl3 at 0 ppm.

Hyperpolarization Experiments

Hyperpolarization of [Y(EDTA)(H2O)2]- (0.158 M), [Y(EGTA)(H2O)]- (0.200 M), [Y(TTHA)]3- (0.200 M), [Y(NTA)2]3- (0.187 M), [Y(DTPA)(H2O)2]- (0.160 M), [Y(DOTA)(H2O)]- (0.397 M), and [Y(DO3A)(H2O)2] (0.187 M) were generated following a reported method.[7] After 4.16 mL of hyperpolarized [Y(EDTA)(H2O)2]- solution was ejected from the HyperSense, 1 mL was transferred to a 10mm NMR tube to acquire the 89Y NMR data. Hyperpolarized 89Y spectra of [Y(TTHA)]3- were collected at pH 4.8 and 6.9 in buffer solutions.

Crystal Structure Determinations

X-ray diffraction data for compound [Y(P2C14Et4)](NO3)3·6H2O were collected at 173 K on a Bruker Smart-CCD- 6000 using Cu-Kα radiation. All data were corrected by Lorentz and polarization effects. Empirical absorption corrections were also applied.[55] Complex scattering factors were taken from the program package SHELX-97.[56] The structure was solved by direct methods using SIR-92,[57] which revealed the position of all non-hydrogen atoms. The structure was refined on F2 by a full-matrix least-squares procedure using anisotropic displacement parameters for all non hydrogen atoms. The hydrogen atoms bonded to carbon atoms were located in their calculated positions and refined using a riding model. For compounds [Y(Me2DO2PA)]Cl and [Y(CB-TE2PA)]Cl single-crystal X-ray diffraction data were collected with a Xcalibur 2 CCD 4-circle Diffractometer (Oxford Diffraction) fitted with a graphite monochromatedMoKα radiation (λ = 0.71073 Å). Data were collected at 170 K and 298 K. Unit cell determination and data reduction, including interframe scaling, Lorentzian, polarization, empirical absorption, and detector sensitivity corrections, were carried out using attached programs of CrysAlis software (Oxford Diffraction).[58] Structures were solved by direct methods and refined by full matrix least-squares method on F2 with the SHELXL9756 suite of programs. The hydrogen atoms were identified at the last step and refined under geometrical restraints and isotropic U-constraints.

Crystal data and structure refinement details are given in Table S5 (Supporting Information). CCDC 1484094, 1484140 and 1484139 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif. Molecular graphics were generated by using USCF Chimera (version 1.8).[59]

Preparation of the complexes

Yttrium complexes derived from ligands EDTA, EGTA, TTHA, DTPA, DO3A, NTA, DOTA, DOTA-(gly)4, DOTAM-2C, DOTAM-3C, DOTAM-5C, were synthesized following reported procedures[7,60]. MS(ESI-, H2O): m/z 377.00 (100%) [Y(EDTA)]-; 465.07 (100%) [Y(EGTA)-4H]-; 579.20 (30%) [Y(H2TTHA)]-; 601.13 (100%) [Y(HTTHA)+Na]-; 467.00 (52%) [Y(NTA)2-4H]-; 511.07 (88%) [Y(NTA)2+2Na]-; 489.00 (100%) [Y(NTA)(HNTA)+Na]-; 489.5 (100%) [Y(DOTA)]-; 238.4 (100%) [Y(DTPA)]2-/2;773.53 (100%) [Y(DOTAM-2C)]-; 829.60 (100%) [Y(DOTAM-3C)]-; 941.73 (100%) [Y(DOTAM-5C)]-.MS (ESI+): m/z741.87 (100%)[Y(HDOTA-(Gly)4)+Na]+; 455.20 (100%) [Y(DO3A)+Na]+; 571.20 (80%)[Y(DO3A)+3Na+2Cl]+.

General procedure for the syntheses of [Y(P2C14Et4)](NO3)3·6H2Oand [Y(P2C14Py4)](NO3)3·6H2O

A solution of Y(NO3)3·6H2O (0.04 mmol) in methanol (5 mL) was added to a stirred solution of the appropriate ligand (0.04 mmol) in the same solvent (10 mL). The solution was allowed to concentrate and the crude products were recrystallized in water. By slow concentration of the water solutions, crystalline products were obtained, which were filtered off and air-dried.

[Y(P2C14Et4)](NO3)3·6H2O

Yield: 67%;1H NMR (500 MHz, D2O, pD = 7.46, 25 °C): δ=8.00 (t, 3J = 7.8 Hz, 2H, py), 7.51 (d, 3J = 7.8 Hz, 4H, py), 4.68 (d, 2J = 16.3 Hz, 4H, –CH2eq-py), 4.17 (td, 2J = 10.9 Hz, 3J = 6.7 Hz, 4H, –CH2ax-OH), 3.88 (m, 4H, –CH2ax-py), 3.86 (m, 4H, –CH2eq-OH), 3.59 (d, 3J = 9.9 Hz, 4H, –CH2eq-(ethyl)), 3.26-3.19 (m, 4H, –CH2ax-N), 2.68 (td, 2J = 10.9 Hz, 3J = 6.7 Hz, 4H, –CH2eq-N), 2.57 ppm (d, 3J = 9.9 Hz, 4H, –CH2ax-(ethyl));IR (ATR): ν=1607 (s), 1457 (s) (C=C) and (C=N)py, 2868 (m), 2763 (m) (OH), 1329 (s), 827 (m), 736 (m) (NO3-); MS (ESI+): m/z 589.22 (100%) [Y(P2C14Et4)-2H]+, 737.19 (4%) [Y(P2C14Et4)+Na+2(NO3)-H]+; elemental analysis calcd (%) forC26H54N9O19Y: C 35.3, H, 6.2 N; 14.2; found: C 35.1, H 6.4, N 14.3.

[Y(P2C14Py4)](NO3)3·6H2O

Yield: 66%;1H NMR (500 MHz, D2O, pD = 7.46, 25°C): δ=8.46 (t, 3J = 7.9 Hz, 2H, py), 7.97 (d, 3J = 7.9 Hz, 4H, py), 7.56 (t, 3J = 7.7 Hz, 4H, -(py)4), 7.27 (d, 3J = 7.7 Hz, 4H, -(py)4), 7.07 (t, 3J = 6.6 Hz, 4H, -(py)4), 6.87 (m, 4H, -(py)4), 4.74 (d, 3J = 16.6 Hz, 4H, -(CH2eq-py)4), 4.59 (d, 3J = 17.5 Hz, 4H, –CH2eq-py), 4.03 (d, 3J = 16.6 Hz, 4H, -(CH2ax-py)4), 3.78 (d, 3J = 17.5 Hz, 4H, –CH2ax-py), 3.18 (d, 3J = 10.0 Hz, 4H, –CH2eq-(ethyl)), 2.43 ppm (d, 3J = 10.0 Hz, 4H, –CH2ax-(ethyl)); IR (ATR): ν=1602 (s), 1457 (s) (C=C) and (C=N)py, 1324 (s), 817 (m), 828 (m), 760 (m) (NO3-); MS (ESI+): m/z903.3 (70%) [Y(P2C14Py4)(NO3)2]+, 420.6 (100%) [Y(P2C14Py4)(NO3)]2+; elemental analysis calcd (%) forC42H58N13O15Y: C 47.0, H 5.4, N 17.0;found: C 47.0, H 5.7, N 16.8.

[Y(Me2DO2PA)]Cl

This complex was synthesized following a reported procedure.[18] To a solution of Me2DO2PA·4HCl (0.050 mg, 0.081 mmol) and Et3N (66 mg, 0.649 mmol) in MeOH (4 mL) was added a solution of YCl3.6H2O (37 mg, 0.122 mmol) in MeOH (3 mL). The mixture was heated to reflux during 72h and then concentrated to dryness. The residue was stirred in CH3CN (2 mL) at room temperature during 24h. The precipitate was filtered and rinsed with CH3CN (2mL) and Et2O (2mL) to yield 34 mg of a white solid. Addition of an excess of KPF6 to a solution of the complex gave single crystals suitable for X-ray diffraction analysis. Yield 71%; 1H NMR (500 MHz, D2O, pD = 7.46, 25°C): δ=8.31 (m, 2H, py), 8.07 (d, 3J = 7.5 Hz, 2H, py), 7.91 (d, 3J = 9.0 Hz, 2H), 4.86 (d, 2J = 16.5 Hz, 2H, –CH2py), 4.29 (d, 2J = 16.5 Hz, 2H, –CH2py), 3.69 (s br, 4H, –CH2-cyclen), 3.4 – 2.7 (m, 10H, –CH2-cyclen), 2.33 (d, 2J = 14.5 Hz, 2H, –CH2-cyclen), 2.10 ppm (s, 6H, -CH3). 13C NMR (125.77 MHz, D2O, pD = 7.46, 25°C): δ=174.22, 158.99, 152.13 (quaternary C); 145.74, 129.87, 127.06 (tertiary C); 58.47, 56.17 (secondary C); 47.76 ppm (primary C).

[Y(CB-TE2PA)]Cl

This complex was prepared following the previously reported procedure.[21] A mixture containing H2CB-TE2PA·4HCl (50 mg, 0.078 mmol), DIPEA (80 mg, 0.624 mmol) and YCl3·6H2O (47 mg, 0.156 mmol) in n-butanol (10 mL) was heated at 150°C during 4h under microwave radiation. The solvent was evaporated and the solid was washed four times with THF. The complex was obtained as a brown solid (28 mg). Addition of an excess of KPF6 to a solution of the complexgave single crystals suitable for X-ray diffraction analysis. Yield: 58%; 1H NMR (500 MHz, D2O, pD = 5.65, 25 °C): δ=8.21 (t, 3J = 7.9 Hz, 2H, py), 7.96 (d, 3J = 7.9 Hz, 2H, py), 7.82 (d, 3J = 7.9 Hz, 2H, py), 5.11 (d, 2J = 15.3 Hz, 2H, –CH2ax-py), 4.15 (d, 2J = 15.3 Hz, 2H, –CH2eq-py), 3.7-3.6 (m, 4H), 3.10-3.01 (m, 2H), 2.96-2.80 (m, 6H), 2.72-2.69 (m, 2H), 2.50-2.41 (m, 2H), 2.4-2.25 (m, 4H), 2.0- 1.9 (m, 2H), 1.3 ppm (m, 2H). 13C NMR (75.47 MHz, D2O, pD = 5.65, 25 °C): δ=175.80, 159.56, 151.96 (quaternary C); 145.62, 129.08, 126.00 (tertiary C); 71.52, 62.22, 62.10, 60.74, 55.91, 55.62, 26.88 ppm (secondary C).

Computational details

Geometry optimizations were performed in aqueous solution employing DFT within the hybrid meta-GGA approximation with the TPSSh exchange-correlation functional,[61]and the Gaussian 09 package (Revision D.01).[62]In these calculations we used the quasi-relativistic effective core potential ECP28MWB of Preusset al.[63] and the related (8s7p6d2f1g)/[6s5p3d2f1g]-GTO valence basis set for Y,[63,64] in combination with the standard 6-31G(d,p) basis set for the ligand atoms. No symmetry constraints have been imposed during the optimizations. The default values for the integration grid (75 radial shells and 302 angular points) and the SCF energy convergence criteria (10-8) were used in all calculations. The stationary points found on the potential energy surfaces as a result of the geometry optimizations have been tested to represent energy minima rather than saddle points via frequency analysis. Bulk solvent effects (water) were included with the polarizable continuum model using the integral equation formalism (IEFPCM)[65] variant as implemented in Gaussian 09. The universal force field radii (UFF)[66] scaled by a factor of 1.1 were used to define the solute cavities.

The calculations of the 89Y NMR shielding tensors were carried out using the ORCA program package (Version 3.0.1)[67] using the own nucleus as the origin of this gauche origin. Relativistic effects were considered by means of the second order Douglas-Kroll-Hess (DKH2) method,[68] with the all-electron scalar relativistic TZVP-DKH basis set of Pantazis et al for Y.[69]The geometries of the complexes optimized with the Gaussian code as described above were employed. The RIJCOSX approximation[70] was used to speed up the calculations of NMR parameters using the Def2-TZVPP/JK[71] auxiliary basis set as constructed automatically by ORCA. The convergence tolerances and integration accuracies of the calculations were increased from the defaults using the available TightSCF and Grid5 options. Solvent effects (water) were taking into account by using the COSMO solvation model as implemented in ORCA.[72] Chemical shifts were calculated as δ = (σref-σ) using [Y(H2O)8]3+ as a reference. Natural electron configurations were obtained with the NBO 6.0 program developed by Weinhold.[73]

Supplementary Material

Supporting Information

Acknowledgments

The authors are indebted to Centro de Supercomputación de Galicia (CESGA) for providing the computer facilities. M. R.-F., D. E.-G. and C. P.-I. thank Ministerio de Economía y Competitividad (CTQ2013-43243-P and CTQ2015-71211-REDT) for generous financial support. R.T. acknowledges the Ministère de l′Enseignement Supérieuret de la Recherche, the Centre National de la Recherche Scientifique and the “Service Commun” of NMR facilities of the University of Brest. C.P.-I. acknowledges the University of Brest for his invited Professor position. ADS acknowledges financial support from the National Institutes of Health (EB-015908 and CA-115531) and the Robert A. Welch Foundation (AT-584).

Footnotes

Supporting information and ORCID(s) for the author(s) for this article are available on the WWWundergttp://dx.doi.org/.

References

  • 1.a) De Leon-Rodriguez LM, Lubag AJM, Malloy CR, Martinez GV, Gillies RJ, Sherry AD. Acc Chem Res. 2009;42:948–957. doi: 10.1021/ar800237f. [DOI] [PMC free article] [PubMed] [Google Scholar]; b) Terreno E, Castelli DD, Viale A, Aime S. Chem Rev. 2010;110:3019–3042. doi: 10.1021/cr100025t. [DOI] [PubMed] [Google Scholar]
  • 2.a) Aime S, Castelli DD, Crich SG, Gianolio E, Terreno E. Acc Chem Res. 2009;42:822–831. doi: 10.1021/ar800192p. [DOI] [PubMed] [Google Scholar]; b) Caravan P. Chem Soc Rev. 2006;35:512–523. doi: 10.1039/b510982p. [DOI] [PubMed] [Google Scholar]
  • 3.Golman K, Olsson LE, Axelsson O, Mansson M, Karlsson M, Peterson JS. Br J of Radiol. 2003;76:S118. doi: 10.1259/bjr/26631666. [DOI] [PubMed] [Google Scholar]
  • 4.Goodson BM. Annu Rep NMR Spectro. 2005;55:299–323. [Google Scholar]
  • 5.Ardenkjaer-Larsen JH, Macholl S, Johannesson H. Appl Magn Reson. 2008;34:509–522. [Google Scholar]
  • 6.Merritt ME, Harrison C, Kovacs Z, Kshirsagar P, Malloy CR, Sherry AD. J Am Chem Soc. 2007;129:12942–12943. doi: 10.1021/ja075768z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Jindal AK, Merritt ME, Suh EH, Malloy CR, Sherry AD, Kovacs Z. J Am Chem Soc. 2010;132:1784–1785. doi: 10.1021/ja910278e. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Lumata L, Jindal AK, Merritt ME, Malloy CR, Sherry AD, Kovacs Z. J Am Chem Soc. 2011;133:8673–8680. doi: 10.1021/ja201880y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Lumata L, Merritt M, Malloy C, Sherry AD, Kovacs Z. Appl Magn Reson. 2012;43:69–79. [Google Scholar]
  • 10.Holz RC, Horrocks WD. J Magn Reson. 1990;89:627–631. [Google Scholar]
  • 11.Shannon RD. Acta Cryst. 1976;A32:751–767. [Google Scholar]
  • 12.Merbach AE, Helm L, Tóth É, editors. The Chemistry of Contrast Agents in Medical Magnetic Resonance Imaging. 2nd. Wiley; New York: 2013. [Google Scholar]
  • 13.a) Caravan P, Ellison JJ, McMurry J, Lauffer B. Chem Rev. 1999;99:2293–2352. doi: 10.1021/cr980440x. [DOI] [PubMed] [Google Scholar]; b) Verwilst P, Park S, Yoon B, Kim JS. Chem Soc Rev. 2015;44:1791–1806. doi: 10.1039/c4cs00336e. [DOI] [PubMed] [Google Scholar]
  • 14.Popovici C, Fernández I, Oña-Burgos P, Roces L, García-Granda S, Ortiz FL. Dalton Trans. 2011;40:6691–6703. doi: 10.1039/c1dt10194c. [DOI] [PubMed] [Google Scholar]
  • 15.Brun E, Oeser J, Staub HH, Telschow CG. Phys Rev. 1954;93:172–173. [Google Scholar]
  • 16.a) Adam RM, Fazakerley GV, Reid DG. J Magn Reson. 1979;33:655–657. [Google Scholar]; b) Levy GC, Rinaldi PL, Bailey JT. J Magn Reson. 1980;40:167–173. [Google Scholar]
  • 17.del M, Fernández-Fernández C, Bastida R, Macías A, Pérez-Lourido P, Platas-Iglesias C, Valencia L. Inorg Chem. 2006;45:4484–4496. doi: 10.1021/ic0603508. [DOI] [PubMed] [Google Scholar]
  • 18.Rodriguez-Rodriguez A, Esteban-Gómez D, de Blas A, Rodríguez-Blas T, Fekete M, Botta M, Tripier R, Platas-Iglesias C. Inorg Chem. 2012;51:2509–2521. doi: 10.1021/ic202436j. [DOI] [PubMed] [Google Scholar]
  • 19.Rodriguez-Rodriguez A, Esteban-Gómez D, de Blas A, Rodríguez-Blas T, Botta M, Tripier R, Platas-Iglesias C. Inorg Chem. 2012;51:13419–13429. doi: 10.1021/ic302322r. [DOI] [PubMed] [Google Scholar]
  • 20.Castro G, Regueiro-Figueroa M, Esteban-Gómez D, Bastida R, Macías A, Pérez-Lourido P, Platas-Iglesias C, Valencia L. Chem Eur J. 2015;21:18662–18670. doi: 10.1002/chem.201502937. [DOI] [PubMed] [Google Scholar]
  • 21.Rodriguez-Rodriguez A, Esteban-Gómez D, Tripier R, Tircso G, Garda Z, Toth I, de Blas A, Rodriguez-Blas T, Platas-Iglesias C. J Am Chem Soc. 2014;136:17954–17957. doi: 10.1021/ja511331n. [DOI] [PubMed] [Google Scholar]
  • 22.Rodriguez-Rodriguez A, Regueiro-Figueroa M, Esteban-Gómez D, Tripier R, Tircso G, Kálmán FK, Bényei AC, Tóth I, de Blas A, Rodriguez-Blas T, Platas-Iglesias C. Inorg Chem. 2016;55:2227–2239. doi: 10.1021/acs.inorgchem.5b02627. [DOI] [PubMed] [Google Scholar]
  • 23.Roca-Sabio A, Mato-Iglesias M, Esteban-Gómez D, Tóth É, de Blas A, Platas-Iglesias C, Rodríguez-Blas T. J Am Chem Soc. 2009;131:3331–3341. doi: 10.1021/ja808534w. [DOI] [PubMed] [Google Scholar]
  • 24.Regueiro-Figueroa M, Esteban-Gómez D, de Blas A, Rodriguez-Blas T, Platas-Iglesias C. Chem Eur J. 2014;20:3974–3981. doi: 10.1002/chem.201304469. [DOI] [PubMed] [Google Scholar]
  • 25.Castro G, Regueiro-Figueroa M, Esteban-Gómez D, Pérez-Lourido P, Platas-Iglesias C, Valencia L. Inorg Chem. 2016;55:3490–3497. doi: 10.1021/acs.inorgchem.5b02918. [DOI] [PubMed] [Google Scholar]
  • 26.Chang CA, Francesconi LC, Malley MF, Kumar K, Gougoutas JZ, Tweedle MF. Inorg Chem. 1993;32:3501–3508. [Google Scholar]
  • 27.Parker D, Pulukkody K, Smith FC, Batsanov A, Howard AK. J Chem Soc, Dalton Trans. 1994:689–693. [Google Scholar]
  • 28.Benetollo F, Bombieri G, Calabi L, Aime S, Botta M. Inorg Chem. 2003;42:148–157. doi: 10.1021/ic025790n. [DOI] [PubMed] [Google Scholar]
  • 29.Aime S, Botta M, Ermondi G. Inorg Chem. 1992;31:4291–4299. [Google Scholar]
  • 30.Amin S, Voss DA, Jr, Horrocks WD, Jr, Lake CH, Churchill MR, Morrow JR. Inorg Chem. 1995;34:3294–3300. [Google Scholar]
  • 31.Aime S, Barge A, Benetollo F, Bombieri G, Botta M, Uggeri F. Inorg Chem. 1997;36:4287–4289. [Google Scholar]
  • 32.Aime S, Botta M, Fasano M, Marques MPM, Geraldes CFGC, Pubanz D, Merbach AE. Inorg Chem. 1997;36:2059–2068. doi: 10.1021/ic961364o. [DOI] [PubMed] [Google Scholar]
  • 33.Dunand FA, Dichins RS, Parker D, Merbach AE. Chem Eur J. 2001;7:5160–5167. doi: 10.1002/1521-3765(20011203)7:23<5160::aid-chem5160>3.0.co;2-2. [DOI] [PubMed] [Google Scholar]
  • 34.Tóth É, Dhubhghaill OMN, Nesson G, Helm L, Merbach AE. Magn Reson Chem. 1999;37:701–708. [Google Scholar]
  • 35.Zhang X, Chang CA, Brittain HG, Garrison JM, Telser J, Tweedle MF. Inorg Chem. 1992;31:5597–5600. [Google Scholar]
  • 36.Aime S, Barge A, Borel A, Botta M, Chemerisov S, Merbach AE, Muller U, Pubanz D. Inorg Chem. 1997;36:5104–5112. [Google Scholar]
  • 37.Zhang LQ, Fan TT, Wang J, Liu B, Li D, Han GX, Xu R, Zhang XD. Russ J Coord Chem. 2010;36:389–395. [Google Scholar]
  • 38.Peters JA. Inorg Chem. 1988;27:48486–4691. [Google Scholar]
  • 39.Wang J, Fan DM, Zhang XD, Ma R. Acta Chim Sin. 2002;60:536–540. [Google Scholar]
  • 40.Wang J, Wang Y, Zhang ZhH, Zhang XD, Tong J, Liu XZh, Liu XY, Zhang Y, Pan ZhJ. J Struct Chem. 2005;46:895–905. [Google Scholar]
  • 41.Lee SG. Bull Korean Chem Soc. 1996;17:589–591. [Google Scholar]
  • 42.Löble MW, Casimiro M, Thielemann DT, Oña-Burgos P, Fernández I, Roesky PW, Breher F. Chem Eur J. 2012;18:5325–5334. doi: 10.1002/chem.201102636. [DOI] [PubMed] [Google Scholar]
  • 43.White RE, Hanusa TP. Organometallics. 2006;25:5621–5630. [Google Scholar]
  • 44.Geraldes CFGC, Sherry AD. J Magn Reson. 1986;66:274–282. [Google Scholar]
  • 45.Mieville P, Jannin S, Helm L, Bodenhausen G. J Am Chem Soc. 2010;132:5006–5007. doi: 10.1021/ja1013954. [DOI] [PubMed] [Google Scholar]
  • 46.Caravan P, Ellison JJ, Mcmurry TJ, Lauffer RB. Chem Rev. 1999;99:2293–2352. doi: 10.1021/cr980440x. [DOI] [PubMed] [Google Scholar]
  • 47.Regueiro-Figueroa M, Platas-Iglesias C. J Phys Chem A. 2015;119:6436–6445. doi: 10.1021/acs.jpca.5b01728. [DOI] [PubMed] [Google Scholar]
  • 48.Esteban-Gómez D, de Blas A, Rodríguez-Blas T, Helm L, Platas-Iglesias C. ChemPhysChem. 2012;13:3640–3650. doi: 10.1002/cphc.201200417. [DOI] [PubMed] [Google Scholar]
  • 49.Autzchbach J. Phil Trans R Soc A. 2014;372:20120489. doi: 10.1098/rsta.2012.0489. [DOI] [PubMed] [Google Scholar]
  • 50.Graeppi N, Powell DH, Laurenczy G, Zekany L, Merbach AE. Inorg Chim Acta. 1995;235:311–326. [Google Scholar]
  • 51.a) Platas-Iglesias C, Corsi DM, Vander Elst L, Muller RN, Imbert D, Bünzli JCG, Toth E, Maschmeyer T, Peters JA. Dalton Trans. 2003:727–737. [Google Scholar]; b) Balogh E, Mato-Iglesias M, Platas-Iglesias C, Toth E, Djanashvili K, Peters JA, de Blas A, Rodriguez-Blas T. Inorg Chem. 2006;45:8719–8728. doi: 10.1021/ic0604157. [DOI] [PubMed] [Google Scholar]
  • 52.a) Platts JA, Gkionis K. Phys Chem Chem Phys. 2009;11:10331–10339. doi: 10.1039/b822560e. [DOI] [PubMed] [Google Scholar]; b) Greif AH, Hrobarik P, Hrobarikova V, Arbuznikov AV, Autschbach J, Kaupp M. Inorg Chem. 2015;54:7199–7208. doi: 10.1021/acs.inorgchem.5b00446. [DOI] [PubMed] [Google Scholar]
  • 53.Facelli JC. Prog Nucl Magn Reson Spectrosc. 2011;58:176–201. doi: 10.1016/j.pnmrs.2010.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Moncho S, Autschbach J. Magn Reson Chem. 2010;48:S76–S85. doi: 10.1002/mrc.2632. [DOI] [PubMed] [Google Scholar]
  • 55.Sheldrick GM SADABS. Program for Empirical Absoption Correction of Area Detector Data. University of Götingen; Germany: 1996. [Google Scholar]
  • 56.Sheldrick GM. SHELX-97, An integrated system for solving and refining crystal structures from diffraction data. University of Göttingen; Germany: 1997. [Google Scholar]
  • 57.Altornare A, Cascarano G, Giacovazzo C, Gualardi A. J Appl Cryst. 1993;26:343–350. [Google Scholar]
  • 58.CrysAlis software system, version 1.171.28 cycle 4 beta. Oxford Diffraction Ltd; Abingdon, U.K.: 2005. [Google Scholar]
  • 59.Pettersen EF, Goddard TD, Huang CC, Couch GS, Greenblatt DM, Meng EC, Ferrin TE. J Comput Chem. 2004;25:1605–1612. doi: 10.1002/jcc.20084. [DOI] [PubMed] [Google Scholar]
  • 60.Lowe MP, Parker D, Reany O, Aime S, Botta M, Castellano G, Gianolio E, Pagliari R. J Am Chem Soc. 2001;123:7601–7609. doi: 10.1021/ja0103647. [DOI] [PubMed] [Google Scholar]
  • 61.Tao JM, Perdew JP, Staroverov VN, Scuseria GE. Phys Rev Lett. 2003;91:146401. doi: 10.1103/PhysRevLett.91.146401. [DOI] [PubMed] [Google Scholar]
  • 62.Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Scalmani G, Barone V, Mennucci B, Petersson GA, et al. Gaussian 09, Revision D.01. Gaussian, Inc.; Wallingford CT; 2009. [Google Scholar]
  • 63.Andrae D, Haeussermann U, Dolg M, Stoll H, Preuss H. Theor Chim Acta. 1990;77:123–141. [Google Scholar]
  • 64.Martin JML, Sundermann A. J Chem Phys. 2001;114:3408–3420. [Google Scholar]
  • 65.Tomasi J, Mennucci B, Cammi R. Chem Rev (Washington, DC, U S) 2005;105:2999–3093. doi: 10.1021/cr9904009. [DOI] [PubMed] [Google Scholar]
  • 66.Rappe AK, Casewit CJ, Colwell KS, Goddard WA, III, Skiff WM. J Am Chem Soc. 1992;114:10024–10035. [Google Scholar]
  • 67.Neese F. The ORCA program system. Wiley Interdiscip Rev: Comput Mol Sci. 2012;2:73–78. [Google Scholar]
  • 68.a) Barysz M, Sadlej AJ. J Mol Struct (Theochem) 2001;573:181–200. [Google Scholar]; b) Reiher M. Theor Chem Acc. 2006;116:241–252. [Google Scholar]
  • 69.Pantazis DA, Chen XY, Landis CR, Neese F. J Chem Theory Comput. 2008;4:908–919. doi: 10.1021/ct800047t. [DOI] [PubMed] [Google Scholar]
  • 70.a) Neese F, Wennmohs F, Hansen A, Becker U. Chem Phys. 2009;356:98–109. [Google Scholar]; b) Izsak R, Neese F. J Chem Phys. 2011;135:144105. doi: 10.1063/1.3646921. [DOI] [PubMed] [Google Scholar]; c) Petrenko T, Kossmann S, Neese F. J Chem Phys. 2011;134:054116. doi: 10.1063/1.3533441. [DOI] [PubMed] [Google Scholar]; d) Kossmann S, Neese F. Chem Phys Lett. 2009;481:240–243. [Google Scholar]
  • 71.Weigend F, Ahlrichs R. Phys Chem Chem Phys. 2005;7:3297–3305. doi: 10.1039/b508541a. [DOI] [PubMed] [Google Scholar]
  • 72.Sinnecker S, Rajendran A, Klamt A, Diedenhofen M, Neese F. J Phys Chem A. 2006;110:2235–2245. doi: 10.1021/jp056016z. [DOI] [PubMed] [Google Scholar]
  • 73.Glendening ED, Badenhoop JK, Reed AE, Carpenter JE, Bohmann JA, Morales CM, Landis CR, Weinhold F. NBO 6.0. Theoretical Chemistry Institute. University of Wisconsin; Madison, WI: 2013. http://nbo6.chem.wisc.edu/ [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information

RESOURCES