Вы находитесь на странице: 1из 8

In Vivo 19F Magnetic Resonance Spectroscopy

Arend Heerschap
Radboud University Medical Center Nijmegen, Nijmegen, The Netherlands

Fluorine (19 F) is widely present in biomedical compounds. 19 F magnetic resonance (MR) imaging and spectroscopy offer unique possibilities for the
in vivo monitoring of the fate of these compounds and of 19 F-loaded particles and cells after their administration in animal models or humans.
The signal intensity and relaxation times of 19 F spins are applied in MR imaging of tissue perfusion, lung ventilation, tissue oxygenation, and for
cell tracking and targeting. With MR spectroscopy the 19 F chemical shift is employed to examine ion binding, enzyme activity, and metabolism of
fluorinated drugs and metabolites. This article introduces in vivo 19 F MR of animals and humans with a focus on spectroscopy. It describes the specific
properties of the 19 F nucleus that make it suitable for in vivo MR applications. 19 F MR is very sensitive and lacks background signal, which results
in a high contrast-to-noise ratio and specificity. Despite its high intrinsic sensitivity, 19 F MR is challenging, especially in clinical applications, due to
the low number of 19 F nuclei that typically accumulate in tissues. Therefore, whenever possible the number of 19 F nuclei per molecule and/or particle
must be maximized, and the hardware and software for 19 F MR acquisition optimized to obtain the best achievable signal-to-noise ratio.
Keywords: MR spectroscopy, 19 F, 5-fluorouracil, perfluorocarbons, sensitivity

How to cite this article:


eMagRes, 2016, Vol 5: 1283–1290. DOI 10.1002/9780470034590.emrstm1445

Introduction Properties of the 19 F Nucleus for In Vivo


Fluorine (19 F) is a sensitive magnetic resonance (MR) nucleus, Magnetic Resonance
which is an attractive property for biomedical applications. Desirable in vivo MR properties of the 19 F nucleus are that it is a
In animals and humans, however, there are no fluorine atoms spin- 1∕2 isotope, which contributes to narrow resonance lines,
present in soft tissue and 19 F MR can only be performed in and that it occurs at 100% natural abundance. The gyromag-
vivo after administration of compounds or particles that are netic ratio (𝛾) for 19 F nucleus is only slightly less than that of 1 H
labeled with fluorine (Figure 1). Fortunately, fluorine is widely (25.181 vs 26.752 × 107 rad T−1 s), which translates to an intrin-
present in medical compounds so there are many potential sic detection sensitivity of 83% as compared to the proton (1 H)
in vivo biomedical applications for 19 F MR,1 as recognized nucleus and a relative signal-to-noise ratio (SNR) of about 89%,
shortly after the introduction of 1 H MRI for humans at the end assuming that noise increases linearly with frequency.
of the 1970s.2,3 The first in vivo 19 F MR spectroscopy (MRS) In the human body, 19 F is present at significant concen-
study published in 1984 examined the metabolic conversion trations only in bone and teeth, but this is undetectable with
of the fluoropyrimidine 5-fluorouracil (FU), a widely used common MR sequences because of the very short spin–spin
chemotherapeutic drug.4 Since then fluorine-labeled com- relaxation times (T2 ) within these rigid structures.14 Apart
pounds or particles have been applied in a broad range of in from its presence at micromolar levels in blood owing to the
vivo 19 F MR studies. These include the use of 19 F contrast fluorination of tap water and toothpaste, there is no fluorine
agents for tissue perfusion and lung ventilation, for the assess- elsewhere in the body. This absence of a background signal
ment of tissue oxygen tension and hypoxia, for the tracking means that 19 F MR of administered compounds can be per-
of cells loaded with 19 F particles, and in molecular imaging of formed with a very good contrast-to-noise ratio and specificity.
labeled bioactive molecules such as antibodies that can target This differs from 1 H MRI in which high background signals
specific cellular receptors.5–8 may obscure specific image features, and from 1 H MRS in
In nearly all of these studies, information is obtained from which metabolite signals often suffer contamination from
the intensity or relaxation properties of just one 19 F signal at a spurious water and lipid signals.
singlet resonance frequency using MRI methods. 19 F MRS has Owing to its high electronegativity, the chemical shift of 19 F
been used to study the metabolism of fluorine-labeled drugs resonances is very sensitive to the chemical environment of the
and anesthetics, and fluorinated probes whose chemical shift is nucleus. For compounds of biomedical interest it has a range
sensitive to pH, ions, or enzyme activity.9–13 This article intro- of about 150 ppm (Table 1) – about 10-times higher than the
duces in vivo 19 F MR of animals and humans with a focus on ∼15 ppm range spanned by 1 H spectra. Thus 19 F resonances of
spectroscopy. biomedical compounds and their conversion products are often

Volume 5, 2016 © 2016 John Wiley & Sons, Ltd. 1283


A Heerschap

CH3
HN O O
F F
N HN
F O
CH3 N O O N
O O
H3C O H2N OH

HN NH F

OH OH OH OH
O
5-Fluorouracil Capecitabine Deoxy-5-fluorouridine Fluoro-β-alanine

F F
F
F F
F
F O O
F
− F
O O
+ F
N F
O O
N OH F
F O F
F
N O F
F F F
F
F F

Trifluoromisonidazole Perfluoro-15-crown-5-ether

Figure 1. Chemical structure of some typical fluorinated biomedical compounds. Top row: antineoplastic agents 5-fluorouracil and capecitabine, and two of
their metabolic products FBAL (catabolite) and FDUR (anabolite). Bottom row: trifluoromisonidazole, a hypoxia marker and perfluoro-15-crown-5-ether,
a compound used for in vivo pO2 measurements and for cell labeling

well resolved in 19 F MRS, especially at field strengths of 3 T and the same carbon such as in 𝛼-fluoro-𝛽-alanine (FBAL).12,16 For
above. However, in some cases such as for fluoronucleotides, the FU the 1 H–19 F coupling is small compared to the 19 F in vivo
chemical shift differences may be too small to resolve in vivo linewidth and can hence be ignored.
(Figure 2 and Table 1). There is no commonly used chemical
shift reference for 19 F spectra obtained in vivo. The signal posi-
tion of 5-FU is often taken as an internal reference at 0 ppm in
Hardware for In Vivo 19 F MR Spectroscopy
metabolic studies of this compound (Table 1).9 With a 1 H/19 F
dual-tuned probe it is possible to calibrate 19 F frequencies rel- As the 𝛾 for 19 F is close to that of 1 H, the existing instrumen-
ative to that of the water frequency if no suitable endogenous tation of a 1 H MRI machine can often be used for 19 F MR
signal is present.20 Trifluoroacetic acid is often used as an exter- with minimal component adjustments. However, performing
19 F MRS experiments on conventional animal or human MRI
nal reference.
The spin–lattice relaxation times (T1 s) of 19 F spins in dif- systems does require special attention to those hardware and
ferent compounds are quite variable and vary with the main software aspects affecting the SNR. This is because in prac-
magnetic field strength (B0 ) and temperature (Table 1).17 An tice, despite the high relative intrinsic sensitivity of 19 F MR,
interesting property is the sensitivity of the T1 of 19 F spins measurements are often challenged by the limited amount
to the presence of oxygen, especially in perfluorinated com- of the fluorine-labeled compound that can be delivered to a
pounds (PFC; organic compounds in which every 1 H nucleus tissue of interest. This results in a low SNR and sensitivity that
is replaced by a 19 F), which are exploited in noninvasive require large measurement volumes and/or long scan times,
measurements of tissue oxygenation level (pO2 ).12,26 which may be detrimental to clinical applications. In 19 F MRI
In some fluorinated compounds the same or an adjacent car- studies this problem is addressed by maximizing the number
bon atom to which the fluorine is attached has one or more of (chemically equivalent) 19 F atoms per compound or per
protons (Figure 1). The dipolar 1 H–19 F heteronuclear spin–spin particle, for example, by using PFCs such as perfluoro-15-
coupling causes a splitting of the 19 F resonance, which is mostly crown-5-ether (PFCE; Figure 1). However, metabolic studies
of the order of ∼5 Hz if a proton is attached to the adjacent of approved fluorine-bearing pharmaceuticals such as 5-FU
carbon such as in FU and can be 50–100 Hz for protons on and the like often have only one fluorine atom per molecule

1284 © 2016 John Wiley & Sons, Ltd. Volume 5, 2016


In Vivo 19 F Magnetic Resonance Spectroscopy

Table 1. Chemical shifts and reported T 1 relaxation times of some relevant fluorine-containing compounds

Compound Chemical shift (ppm) T1 (ms) solution T1 (ms) rodent T1 (ms) human
Trifluoromisonidazole 97 ∼2500a 370 ± 20b
Trifluoroacetic acid (TFA) 94 2500 ± 60c
Isoflurane CF3 90.5
CHF2 84.2
Perfluoro-15-crown-5-ether (PFCE) 78 ∼2500d 580 to 500e
∼2400f
Capecitabine 6 ∼410 g 250 ± 70h
5-Fluoronucleotides 4–4.5 1100 ± 100i 400 ± 140h
1300 ± 100j
5-Fluoronucleosides 3.5
5-Fluorouracil (FU) 0 3500 ± 760c 2000 ± 200i 380 ± 80k
2940 ± 230l 2300 ± 100j
3400m
𝛼-Fluoro-𝛽-ureidopropionic acid (FUPA) −17 1900 ± 100i
FBAL acid conjugates −17
𝛼-Fluoro-𝛽-alanine (FBAL) −19 2720 ± 260l 2000 ± 400i 1600 ± 200n
1050 ± 210h
2-Fluoro-2-deoxyglucose 𝛼, 𝛽 approx. −30 700–900o
Dihydrofluorouracil (DHFU) −32

Chemical shifts are approximate and may vary with temperature, pH, and salinity. The T1 values in aqueous solution or emulsions are at room temperature.
a Obtained in vitro at 7 T.
b Obtained at 7 T from a subcutaneous prostate tumor model.15
c Obtained at 1.5 T.16
d Obtained at 7 T from emulsion at 37 ∘ C.17
e Obtained at 11.7 T from cell systems.18
f Obtained at 4.3 T from dead mouse tissue. Note that the T from PFCE strongly depends on pO .12,26
1 2
g Obtained at 7 T.19
h Obtained in vivo at 3 T from the human liver.20
i Obtained at 9.4 T from whole body mice.21
j Obtained from subcutaneous tumor at 4.7 T.22
k Obtained in vivo at 1.5 T from the human liver.23
l Obtained at a physiological liver pH 7.4.16
m Obtained at 11.7 T from human plasma.16
n Obtained at 1.5 T from human liver.24
o Obtained from ex vivo mouse tissues after fluorodeoxygluoce (FDG) administration at 6.3 T.25

(Figure 1). Therefore it is extremely important to optimize the The easiest approach for good SNR is a circular 19 F surface
MR hardware for maximum signal detection. coil both for RF transmit and receive, which should be chosen
Because the SNR increases about linearly with B0 (assuming based on the size and location of the tissue of interest (e.g.
optimum detection at each B0 ), choosing the highest field for liver studies, use a receiver coil of diameter approximately
available is advised. In small animal studies field strengths equal to the depth of the signal, positioned adjacent to the
of B0 ≥ 7 T are widely available. For human studies, clinical liver).23 For 1 H shimming and MRI either the volume body
systems operating at 3 T are common and 7 T whole-body coil or a separate surface coil can be used as transmission
systems are available at some institutions.19 An additional and reception. Further improvements have been realized with
benefit of higher fields is an increase in the chemical shift tunable or low-loss dual-frequency 1 H/19 F surface and volume
dispersion in Hertz. The higher SNR and spectral resolution coils that enable easy switching between both frequencies.20,30
available at higher fields may be partly outweighed by increased To benefit from the high sensitivity of local surface coils and
tissue magnetic susceptibility effects, which locally decreases still provide large field-of-view coverage, 1 H/19 F phased array
the B0 homogeneity and shortens the effective transverse coil combinations have also been designed for 19 F MR.19 Note
magnetization decay relaxation times (T2 *).29 that the fluorine present in some plastics such as Teflon used
For 19 F measurements a dedicated RF coil and interface in coil probes can generate broad signal artifacts in 19 F MR
(including preamplifier and transmit/receive switches) is spectra.
required for transmission and detection at the 19 F MR fre- Most clinical MRI/MRS system manufacturers provide spec-
quency. This setup should also have the capability to perform trometers with multinuclear options that include 19 F, but the
1 H MR for shimming the magnetic field for the best possible
standard coil package usually does not include 19 F coils and
homogeneity and 1 H MRI for anatomical localization and one needs to either fabricate or purchase these separately. Until
mapping of the tissue of interest. For optimal SNR the selection recently, 19 F coils for small animal studies were often custom
of an RF coil is at least as important to SNR as the field strength. built, but these are now also commercially available.

Volume 5, 2016 © 2016 John Wiley & Sons, Ltd. 1285


A Heerschap

c d Catabolites 5′DFCR
Capecitabine
1 5 5′DFUR
a b
2
FBAL
FBAL-bile acid conjugate
5 4 3
3
4

1 15 10 5 0 −5
2 6 ppm

116 min
(a) FU

100 min
5′DFCR and 5′DFUR
Anabolites
68 min
Capecitabine

64 min

10 0 −10 −20 −30


(b) ppm (c) 25 0 −25 −50 ppm

Figure 2. Comparison of in vitro and in vivo 19 F MR spectra after application of 5-fluorouracil or its prodrug capecitabine. Each signal in these spectra
represents fluorine in the different chemical environments of each metabolite and the signal integrals are proportional to their tissue levels. The in vivo
spectra exhibit broader resonances and poorer SNR compared to the in vitro spectrum, because of a shorter T 2* due to contributions from tissue and
B0 inhomogeneity and susceptibility effects, along with differences in T 2 processes associated with the tissue molecular-level environment. (a) In vitro
19 F spectrum obtained at 7 T from a perchloric acid extract of C38 murine colon carcinoma tissue, excised from a tumor-bearing mouse 50 min after 5-

fluorouracil (FU) injection. Peak 1, fluoronucleotides (peaks a–d); Peak 2, fluoronucleosides; Peak 3, FU; Peak 4, 𝛼-fluoro-𝛽-ureidopropionic acid (FUPA);
Peak 5, FBAL; Peak 6, 5,6-dihydro-5-fluorouracil (DHFU). The spectrum was recorded from 50 000 acquisitions, with broadband 1 H decoupling and a
pulse repetition time (TR) of 5 s. (Reproduced with permission from Kamm et al.27 ) (b) In vivo 19 F spectrum obtained at 7 T with a 10 mm surface coil
from a C38 murine colon carcinoma implanted subcutaneously in the flank of a female C57BL/6 mouse. The spectrum was acquired in 4 min with a simple
pulse-and-acquire sequence (TR ∼500 ms) about 60 min after intraperitoneal injection of a bolus of FU. (c) In vivo 19 F spectra obtained at 1.5 T with a
pulse-acquire pulse sequence (TR = 470 ms) from the liver of a patient with a colon carcinoma metastasis after oral capecitabine intake. A flexible 19 F coil
with a diameter of 16 cm was used. Sequential spectra were obtained at a time resolution of 4 min, at the indicated times post-capecitabine intake. Inset:
expanded spectral region from −5 to 15 ppm, showing the average of spectra acquired from 60 to 76 min post-capecitabine intake. 5′ DFUR: 5′ -deoxy-
5-fluorouridine. 5′ FDCR: 5′ -deoxy-5-fluorocytidine. The resonance assigned to FBAL-bile acid conjugates may also contain a resonance for FUPA.11,28
(Reproduced from Ref. 29. © AACR, 2003)

19 from different tissues, such as normal liver tissue, liver tumor,


F MR Spectroscopy Acquisition Methods
metastases or the gall bladder.
Once the hardware adaptations are in place, in vivo experiments Given the need for more specific localization to study tumor
can be carried out. After positioning the subject in the scan- metabolism, 19 F MR spectroscopic imaging (MRSI) approaches
ner, the typical 19 F MRS protocol commences by acquiring an have been adopted for small animals and humans.23,28,31–34 The
anatomical 1 H scout image, followed by magnetic field shim- SNR of 19 F images and spectra from the region of interest can be
ming and then acquisition of 19 F data. optimized by using Ernst angle excitation (see The Basics) for
Most of the 19 F MRS studies in animals and humans have simple partial saturation pulse sequences (comprised of pulse-
been performed using simple pulse-and-acquire sequences acquire-TR delay-repeat). This requires knowledge of the T1
wherein spatial localization is limited only by the sensitive values in vivo, some of which are listed in Table 1. An example
volume of a surface detection coil. Adiabatic pulses can be of 19 F MRSI performed at 3 T in the human liver after adminis-
used to obtain a more homogeneous RF field (B1 ) over the tration of the FU prodrug capecitabine is shown in Figure 3. In
tissue of interest when surface coils are used for excitation vivo 19 F MRS signals from fluorinated biocompounds typically
exhibit T2 * values of less than about 15 ms depending on the
(vs volume coil excitation), but these pulses are challenged by
tissue volume and susceptibility effects, but their T2 s are 3–10
bandwidth and power requirements for covering large spectral
times or more longer, potentially enabling fast spin-echo (FSE)
ranges, such as those spanned by signals from both anabolites type acquisitions for MRI, and echo-planar spectroscopic
and catabolites in fluoropyrimidine metabolism (about 25 ppm imaging approaches (see High-Speed Spatial–Spectral Encod-
apart).19,20 Using only the limited localization provided by ing with PEPSI and Spiral MRSI). By interleaving selective
surface coils – which are often much larger than the size of excitation of 19 F spins in different compounds (see also Direct
the targeted region of interest within the tissue or organ – no Water–Fat Imaging Methods: Chemical Shift-Selective and
distinction can be made between metabolite signals originating Chemical Shift-Encoded MRI), this has enabled separate FSE

1286 © 2016 John Wiley & Sons, Ltd. Volume 5, 2016


In Vivo 19 F Magnetic Resonance Spectroscopy

imaging of the fluoronucleotides, FU and FBAL, in an animal phantom, which is then measured under the same acquisition
tumor model after FU administration.21 conditions as the in vivo study.24,35 While the ratio of in vivo
In most 19 F MR studies the spectroscopic dimension is to reference spectra are in direct proportion to the concentra-
discarded and imaging methods are used for signal acquisi- tions, care must be taken to correct for any differences in B1
tion. Even if the spectroscopy information is not required, (or flip-angle) and the coil sensitivity whenever the reference
MR spectroscopic sequences may still be used if large vol- and in vivo signal sources are at different locations. Also any
umes/voxel sizes are needed to maximize the number of 19 F differences in relaxation times between the reference and in
spins contributing to the signal and/or to improve sensitivity situ 19 F compound must be accounted for, unless the signals
by reducing the bandwidth compared to MRI (see The Basics). are acquired under fully relaxed conditions (see Quantifying
An example is the detection of tissue hypoxia by the accu- Metabolite Ratios and Concentrations by Non-1 H MRS).
mulation of nitroimidazoles.15 In applications using PFCs for With a dual tuneable 19 F/1 H coil it is also possible to employ
cellular loading or tissue targeting, standard MRI sequences the water signal intensity as a concentration reference.20 This
with timing parameters optimized for 19 F are employed, using requires a calibration of the ratio of the signal per 19 F to that of
compounds with preferably only a single 19 F resonance, such as 1 H, and knowledge of the tissue water content.
PFCE (Figure 1).18 The chemical shift displacements of signals As 19 F MR has a similar sensitivity as 1 H MR the detection
from compounds with multiresonance spectra due to their limits of the two nuclei are comparable, i.e., in the order of mil-
different frequency offsets are usually considered as artifacts limoles (mM). A lower limit can be assessed from the tissue
of spatial registration that need to be corrected.35,36 However, concentration of 19 F (or 1 H) nuclei in a 1 cm3 voxel that can
an MRI approach making use of the chemical shift difference
be measured in a clinically acceptable time (<20 min, say).35
between two PFCs that were used to label two different cell
Quantitative 19 F MRS studies have shown that a few millimoles
populations was recently demonstrated for cell tracking.37
of 19 F can be detected in 1 cm3 in this time frame at 3 T, as
For 19 F compounds exhibiting heteronuclear 1 H–19 F cou-
would be required for a human study.20,35 However, in studies of
pling, it is possible to collapse the multiplet structures into
fluorinated drugs the tissue concentrations of the compounds
a singlet by the application of broadband RF irradiation
of interest can be substantially lower, requiring larger volumes
applied at the 1 H frequency during data acquisition (see The
to maintain acceptable measurement times (typically 64 cm3 in
Basics).16,34,38 This results in a spectrum with one signal per
chemically distinct fluorine moiety with better SNR, depending the human liver).19,20,34
on the number of peaks in each multiplet that are collapsed. As a rule of the thumb for clinical 3 T studies, in vivo concen-
The in vitro 19 F spectrum shown in Figure 2a was recorded trations of fluorine-containing metabolites must be more than
with 1 H decoupling irradiation. However, applying 1 H decou- about 100 nmol g−1 tissue to be detectable in reasonable mea-
pling for in vivo 19 F MRS is not useful when the heteronuclear surement times (<20 min). For animal studies at 7–9.4 T at least
couplings are smaller than the observed signal linewidths (e.g. 50 nmol g−1 needs to be present to provide a 30 min quantitative
for some 5-FU multiplets). SNR may be further improved measurement in a subcutaneous tumor. In whole mice at 9.4 T,
19 F MRSI studies at resolutions of 5 × 5 mm2 with no spatial
by enhancement using the 1 H–19 F nuclear Overhauser effect
(nOe = 1 + 𝛾 H /2𝛾 F ; see The Basics). This also requires 1 H selection in the third dimension required at least 200 nmol g−1
irradiation, either at the frequency of water protons (involving and 13 min scan-times.21,33 In ex vivo studies 10 nmol can be
intermolecular dipolar interactions) or of protons adjacent detected in 0.5–1 g tissue samples in 1–2 h on a 11.7 T high-
to the fluorine nuclei (involving intramolecular interactions), resolution MRS system.39
and may result in signal enhancements of up to about 50%, Although 19 F MRS is an attractive tool for in vivo phar-
depending on the T1 values and the way the nOe is created.16,34 macokinetic studies, the low SNR per unit time in most
The RF power applied for decoupling and nOe is limited by metabolic studies limits the possibilities for monitoring rapid
safety guidelines for RF power deposition in tissue. kinetics occurring within the time frame of a signal-averaged
acquisition.11 In principle a substantial enhancement of the 19 F
signal in rapid conversions is possible by hyperpolarization,
Quantification and Sensitivity in 19 F MR but this has not yet been performed in vivo.40
Spectroscopy Finally, it should be noted that the uptake and metabolism
As the 19 F signal is proportional to the number of fluorine of fluorinated compounds such as fluoropyrimidine can also be
atoms in the volume that is measured by 19 F MRS, a direct monitored in vivo using positron emission tomography (PET)
quantitative assessment of the amount of labeled chemicals upon administration of 18 F-labeled compounds. The sensitiv-
in the volume is possible. This can be done by comparing the ity of 18 F PET is a factor of 106 –109 higher than the sensitiv-
peak integral of the 19 F signal from the label to that acquired ity of 19 F MRS.41 However, 18 F PET measures a composite sig-
from a fluorine-bearing reference compound. Given the lack of nal from all 18 F metabolites, whereas 19 F MRS resolves signals
suitable endogenous 19 F compounds for this purpose in vivo, a of individual metabolites, enabling separate evaluation. More-
phantom filled with a known concentration of 19 F may be used over, while 18 F-labeled agents require immediate synthesis and
for this purpose. The reference phantom is placed adjacent to administration and decay rapidly, those labeled with 19 F can be
the subject so that its signal is acquired simultaneously with the stored and traced over hours to days post-administration. This
in vivo acquisition. Alternatively, after the in vivo acquisition, property is particularly valuable for the in situ tracking of cells
the subject may be replaced by a similar or larger reference loaded with 19 F particles.5,18,42

Volume 5, 2016 © 2016 John Wiley & Sons, Ltd. 1287


A Heerschap

FBAL

cap
FBAL

Capecitabine

Figure 3. Three-dimensional 19 F MRSI of the liver of a patient after oral intake of capecitabine. At left, the MRSI grid is overlaid on a 1 H image of the
liver. The spectra shown in each voxel are a subset of the spectra from a 10 × 10 × 10 acquisition matrix with a nominal resolution of 2.7 × 2.7 × 2.7 cm3 .
The resonances for capecitabine (cap) and FBAL are indicated. The metabolic maps at right indicate minimal colocalization of the highest intensity signals
from FBAL with those from capecitabine. (Reproduced with permission from Ref. 20. © John Wiley & Sons Ltd., 2007)

Acknowledgments 9. W. Wolf, C. A. Presant, and V. Waluch, Adv. Drug Deliv. Rev., 2000, 41, 55.
10. D. J. McIntyre, F. A. Howe, C. Ladroue, F. Lofts, M. Stubbs, and J. R. Griffiths,
I am most thankful to Yvonne Kamm, Hanneke van Laarhoven,
Cancer Chemother. Pharmacol., 2011, 68, 29.
Dennis Klomp, Mangala Srinivas, Fernando Bonetto, Mark van
Uden, Andor Veltien, Sjaak van Asten, Tom Scheenen, Edwin 11. H. W. van Laarhoven, C. J. Punt, Y. J. Kamm, and A. Heerschap, Crit. Rev. Oncol.
ter Voert and Houshang Amiri for collaborations in performing Hematol., 2005, 56, 321.

and discussing 19 F MR studies of animals and humans. 12. J. X. Yu, V. D. Kodibagkar, W. Cui, and R. P. Mason, Curr. Med. Chem., 2005,
12, 819.
13. H. W. van Laarhoven, D. W. Klomp, M. Rijpkema, Y. L. Kamm, D. J. Wagener, J.
Biographical Sketch O. Barentsz, C. J. Punt, and A. Heerschap, NMR Biomed., 2007, 20, 128.
Arend Heerschap is professor in Biomedical MR at the Radboud Uni- 14. R. F. Code, J. E. Harrison, K. G. McNeill, and M. Szyjkowski, Magn. Reson. Med.,
versity Medical Centre Nijmegen, the Netherlands. His research aims 1990, 13, 358.
to exploit the unique capabilities of MR in ‘molecular’ and ‘biological’
15. D. Procissi, F. Claus, P. Burgman, J. Koziorowski, J. D. Chapman, S. B. Thakur,
imaging and spectroscopy to address biomedical questions and prob-
C. Matei, C. C. Ling, and J. A. Koutcher, Clin. Cancer Res., 2007, 13, 3738.
lems in oncology and intermediate (energy) metabolism. He studies
humans and animals, focusing on tissues such as prostate, brain, liver, 16. P. Bachert, Prog. Nucl. Magn. Reson. Med., 1998, 33, 1.
and skeletal muscle. 17. D. K. Kadayakkara, K. Damodaran, T. K. Hitchens, J. W. Bulte, and E. T. Ahrens,
J. Magn. Reson, 2014, 242, 18.

References 18. M. Srinivas, P. Boehm-Sturm, C. G. Figdor, I. J. de Vries, and M. Hoehn, Bio-


materials, 2012, 33, 8830.
1. D. G. Reid and P. S. Murphy, Drug Discov. Today, 2008, 13, 473.
19. J. S. van Gorp, P. R. Seevinck, A. Andreychenko, A. J. Raaijmakers, P. R. Luijten,
2. G. N. Holland, P. A. Bottomley, and W. S. Hinshaw, J. Magn. Reson., 1977, 28, M. A. Viergever, M. Koopman, V. O. Boer, and D. W. Klomp, NMR Biomed.,
133. 2015, 28, 1433.
3. L. C. Clark Jr, J. L. Ackerman, S. R. Thomas, R. W. Millard, R. E. Hoffman, R. G. 20. D. Klomp, H. van Laarhoven, T. Scheenen, Y. Kamm, and A. Heerschap, NMR
Pratt, H. Ragle-Cole, R. A. Kinsey, and R. Janakiraman, Adv. Exp. Med. Biol.,
Biomed., 2007, 20, 485.
1984, 180, 835.
21. Y. Doi, T. Shimmura, H. Kuribayashi, Y. Tanaka, and Y. Kanazawa, Magn. Reson.
4. A. N. Stevens, P. G. Morris, R. A. Iles, P. W. Sheldon, and J. R. Griffiths, Br. J.
Med., 2009, 62, 1129.
Cancer, 1984, 50, 113.
22. T. P. Gade, W. M. Spees, H. C. Le, K. L. Zakian, V. Ponomarev, M. Doubrovin, J.
5. J. Ruiz-Cabello, B. P. Barnett, P. A. Bottomley, and J. W. Bulte, NMR Biomed.,
2011, 24, 114. G. Gelovani, and J. A. Koutcher, Magn. Reson. Med., 2004, 52, 169.

6. M. Srinivas, E. H. Aarntzen, J. W. Bulte, W. J. Oyen, A. Heerschap, I. J. de Vries, 23. D. W. Klomp, H. W. Van Laarhoven, A. P. Kentgens, and A. Heerschap, Magn.
and C. G. Figdor, Adv. Drug Deliv. Rev., 2010, 62, 1080. Reson. Med., 2003, 50, 303.

7. J. X. Yu, R. R. Hallac, S. Chiguru, and R. P. Mason, Prog. Nucl. Magn. Reson. 24. C. W. Li, W. G. Negendank, K. A. Padavic-Shaller, P. J. O’Dwyer, J. Murphy-
Spectrosc., 2013, 70, 25. Boesch, and T. R. Brown, Clin. Cancer Res., 1996, 2, 339.
8. M. Wolters, S. G. Mohades, T. M. Hackeng, M. J. Post, M. E. Kooi, and W. H. 25. Y. Kanazawa, K. Umayahara, T. Shimmura, and T. Yamashita, NMR Biomed.,
Backes, Invest. Radiol., 2013, 48, 341. 1997, 10, 35.

1288 © 2016 John Wiley & Sons, Ltd. Volume 5, 2016


In Vivo 19 F Magnetic Resonance Spectroscopy

26. B. P. van der Sanden, A. Heerschap, A. W. Simonetti, P. F. Rijken, H. P. Peters, 35. H. Amiri, M. Srinivas, A. Veltien, M. J. van Uden, I. J. de Vries, and A. Heerschap,
G. Stuben, and A. J. van der Kogel, Int. J. Radiat. Oncol. Biol. Phys., 1999, 44, Eur. Radiol., 2015, 25, 726.
649. 36. M. Meissner, M. Reisert, T. Hugger, J. Hennig, D. von Elverfeldt, and J. Leupold,
27. V. J. Kamm, I. M. Rietjens, J. Vervoort, A. Heerschap, G. Rosenbusch, H. P. Hofs, Magn. Reson. Med., 2015, 73, 2225.
and D. J. Wagener, Cancer Res., 1994, 54, 4321. 37. M. Srinivas, J. Tel, G. Schreibelt, F. Bonetto, L. J. Cruz, H. Amiri, A. Heerschap,
28. A. S. Dzik-Jurasz, D. J. Collins, M. O. Leach, and I. J. Rowland, Magn. Reson. C. G. Figdor, and I. J. de Vries, Nanomedicine (Lond.), 2015, 10, 2339.
Med., 2000, 44, 516. 38. B. S. Li, G. S. Payne, D. J. Collins, and M. O. Leach, Magn. Reson. Med., 2000,
29. H. W. van Laarhoven, D. W. Klomp, Y. J. Kamm, C. J. Punt, and A. Heerschap, 44, 5.
Cancer Res., 2003, 63, 7609. 39. N. W. Lutz, B. Naser-Hijazi, S. Koroma, M. R. Berger, and W. E. Hull, NMR
30. L. Hu, F. D. Hockett, J. Chen, L. Zhang, S. D. Caruthers, G. M. Lanza, and S. A. Biomed., 2004, 17, 101.
Wickline, J. Magn. Reson. Imaging, 2011, 34, 245. 40. U. Bommerich, T. Trantzschel, S. Mulla-Osman, G. Buntkowsky, J. Bargon, and
31. G. Brix, M. E. Bellemann, H. J. Zabel, P. Bachert, and W. J. Lorenz, Magn. Reson. J. Bernarding, Phys. Chem. Chem. Phys., 2010, 12, 10309.
Imaging, 1993, 11, 1193. 41. G. Brix, M. E. Bellemann, U. Haberkorn, L. Gerlach, and W. J. Lorenz, Nucl.
32. G. Brix, M. E. Bellemann, L. Gerlach, and U. Haberkorn, Magn. Reson. Imaging, Med. Biol., 1996, 23, 897.
1999, 17, 151. 42. E. T. Ahrens and J. Zhong, NMR Biomed., 2013, 26, 860.
33. H. Kuribayashi, Y. Doi, and Y. Kanazawa, Magn. Reson. Med., 2001, 46, 864.
34. O. Gonen, J. Murphy-Boesch, C. W. Li, K. Padavic-Shaller, W. G. Negendank,
and T. R. Brown, Magn. Reson. Med., 1997, 37, 164.

Volume 5, 2016 © 2016 John Wiley & Sons, Ltd. 1289

Вам также может понравиться