Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

This is the peer-reviewed pre-print version of the following article

“Instrumentation for solid-state dynamic nuclear polarization with magic angle spinning NMR”

which has been published in final form in


Journal of Magnetic Resonance 265, 88-98, 2016, DOI: 10.1016/j.jmr.2015.12.026
http://www.sciencedirect.com/science/article/pii/S1090780716000495
© 2016 Elsevier Inc.
This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0

Instrumentation for solid-state dynamic nuclear polarization


with magic angle spinning NMR
Melanie Rosaya, Monica Blankb, Frank Engelkec*
a
Bruker-Biospin, 15 Fortune Drive, Billerica MA 01730 USA (melanie.rosay@bruker.com)
b
Communications and Power Industries, 811 Hansen Way, Palo Alto, CA 94304 USA (monica.blank@cpii.com)
c
Bruker-Biospin, Silberstreifen 4, 76287 Rheinstetten, Germany (frank.engelke@bruker.com)

Abstract
Advances in dynamic nuclear polarization (DNP) instrumentation and methodology have been key factors in the
recent growth of solid-state DNP NMR applications. We review the current state of the art of solid-state DNP
NMR instrumentation primarily based on available commercial platforms. We start with a general system
overview, including options for microwave sources and DNP NMR probes, and then focus on specific
developments for DNP at 100 K with magic angle spinning (MAS). Gyrotron microwave sources, passive
components to transmit microwaves, the DNP MAS probe, a cooling device for low-temperature MAS, and
sample preparation procedures including radicals for DNP are considered.

1. Introduction
2. Microwave sources for solid-state DNP NMR
3. Solid-state DNP NMR probes and cryogenic MAS
4. DNP radicals and solid-state DNP sample preparation
5. Outlook
6. References

*Corresponding author

1
1. Introduction

The last two decades have witnessed an impressive development and renewed interest in dynamic
nuclear polarization (DNP) in liquids and solids. The present article will focus on developments in the
field of DNP NMR in solids with a particular focus on activities to provide a unique platform for magic
angle spinning (MAS) DNP NMR. The availability of advanced instrumentation for that field,
pioneered by the group of Griffin [1-4] at the Massachusetts Institute of Technology (MIT), has
achieved a sophisticated level of instrumentation and contributed significantly to the propagation
and new applications of solid-state DNP NMR. This progress in instrumentation and methodology is
still continuing such that the present article can only give a summary of the state-of the art with a
brief overview of some achievements and prospective outlook.
In this publication we focus on solid-state NMR enhanced by DNP. The active field of Dissolution DNP
is not included within the present article: although the DNP step in dissolution DNP is also in solid-
state (at lower temperature), the applications are aimed at MRI and liquid-state NMR. This highly
interesting field is covered by several other articles in this special perspectives journal issue.
The discovery of DNP in the 1950’s by Carver and Slichter [5] in solid-state samples (metals) was
followed by many studies of DNP in low temperature physics and liquids and solids NMR [6-8]. The
initial applications of DNP for NMR applications were performed by the groups of Wind, Schaefer
and Yannoni [6,7,58,59] at relatively low NMR and microwave frequency (40 GHz). In the mid-1990’s
Griffin [1-3] started to develop solid-state DNP in solids at low temperatures using gyrotrons as
microwave sources operating beyond 100 GHz. More than one decade later, commercial companies
such as Bruker/CPI and Gycom started to build gyrotrons for DNP, manufactured, for example, by
Bruker with CPI [9], Osaka University [10-12], and Gycom [13-14]. There is a straightforward reason
to use microwave sources which enable relatively high power (on the order of 10 watts) for
frequencies far above 100 GHz, as compared to EPR: NMR DNP probes need to be optimized for
sample size, radio frequency (rf) irradiation, and microwave irradiation. This leads to the
requirement, in particular for MAS probes, that the samples and MAS rotors have geometric
dimensions on the order of or larger than the wavelength (say, about 1 mm at 263 GHz) and
therefore the MAS sample “cavity” is overmoded with a relatively low Q factor for millimeter waves.
From the instrumentation and methodological point of view four fields of relevance should be
considered: (a) DNP microwave sources that fulfill the high demands for spectroscopic techniques,
like gyrotrons, klystrons, or solid-state sources (b) microwave transmission from the gyrotron to the
DNP NMR probe, (c) microwave irradiation of the sample in the DNP NMR probe, (d) DNP sample
preparation including suitable radicals/biradicals particularly adapted to the DNP mechanisms.
2
Figure 1: Schematics of a solid-state DNP NMR system with a gyrotron microwave source (gyrotron
tube in red), microwave transmission line (cyan) and low-temperature NMR probe (green)

A commercial solid DNP platform, as sketched in Fig. 1, contains an NMR spectrometer (including the
NMR magnet), a microwave source for example a gyrotron tube with dedicated magnet plus control
unit with high voltage power supply and gyrotron cooling system, the transmission line for
microwave propagation from the source to the NMR sample, the DNP NMR probe and a cooling unit
providing cryogenic gases for cooling the sample and operating MAS at low temperature. Although
there have been attempts to apply DNP to solid-state NMR experiments at room temperature [6-7],
the majority of solid-state DNP NMR experiments are run at low temperature below 120 K. The main
reason for that is the prolonged electron spin relaxation time at low temperature, a prerequisite to
obtain high DNP enhancement factors.

A variety of components for DNP systems are available from companies such as Revolution NMR with
probes operating according to the cooling principle as introduced by Tycko, et al. utilizing nitrogen
gas for sample rotation and cold helium gas for sample cooling [15-17]. Further Bridge-12 [18] and
Swissto12 [19], offer DNP waveguide components, for example, corrugated waveguides, tapers,
miterbends, and polarization grids. Gyrotrons (with more than 10 W output power) are available
from Bruker/CPI, described in following section, Fukui University [20-22], Gycom [13-14], and within
a research project at EPFL, from Thales [23]. Lower-power millimetre-wave sources, in particular 263
GHz klystron (ca. 1 W) [24,25] and even solid-state sources at 140-263 GHz [4,16,26,27], have been
also demonstrated for applications at helium-cryogenic temperatures. Virginia Diodes, for example,
can provide about 90 mW of cw output power from a solid-state source at 263 GHz [27]. The strong
3
increase of interest in solid-state DNP NMR instrumentation manifests itself also in patent filing
activities. Seeking protection of IP in that field is not restricted to commercial companies such as
Swiss-to-12 [28-31], Bridge-12 [32], Doty [33], or Bruker [34-35], also academic institutions
contribute IP like the patent applications of CNR Roma [36-38], EPFL in Lausanne in collaboration
with Swiss-to-12, and CEA Grenoble [39-41]. In parallel to the publication records of scientific papers
this shows impressively the potential of DNP instrumentation for solid-state NMR today.

In the following sections we focus on a DNP NMR system working with a gyrotron source, operating
at frequencies of either 263 GHz, 395 GHz, or 527 GHz (for 400, 600 and 800 MHz NMR) with the DNP
NMR probe at temperatures around 100 K, including MAS. First, in section 2 an overview of vacuum
electron devices (gyrotron and klystron) is presented followed by a detailed description and
characterization of a gyrotron designed for DNP applications. The transmission of the microwave
beam from its source to the probe and the characteristics of that beam are also described. In section
3 we describe the specifics of solid-state DNP NMR probes, in particular involving MAS and address
briefly cooling techniques and equipment. The success of DNP NMR crucially depends on the DNP
radical used as the source of electron spin polarization - section 4 is dedicated to questions related to
DNP radicals and sample preparation for different application areas. In section 5 we provide an
outlook on near-future prospectives in solid-state DNP NMR.

2. Microwave sources for solid-state DNP NMR


Solid-state and vacuum electronic device sources are available for DNP systems. Vacuum
electronic devices capable of producing the powers, frequencies, and stability suitable for DNP
applications can be broken into two categories, fast-wave and slow-wave devices. For both fast and
slow-wave vacuum electronic sources, energy is transferred from an unbound electron beam to an
electromagnetic wave. In order for this transfer to occur, synchronism between the beam and wave
must be achieved. In fast-wave devices, such as the gyrotron [42-43], the phase velocity of the
electromagnetic wave is equal to or greater than the speed of light and in slow-wave devices, such as
klystrons or helix traveling-wave-tubes, the phase velocity of the wave is less than the speed of light.
Gyrotrons, also known as cyclotron resonance masers, take advantage of the cyclotron resonance
maser instability to transfer energy from an electron beam to an electromagnetic wave. In the
interaction, an annular electron beam, composed of mildly relativistic electrons traveling in helical
paths, interacts with the electromagnetic fields of a circuit or cavity in the presence of an applied
axial magnetic field. In cyclotron resonance masers, the synchronism between the electron beam
electromagnetic fields of the circuit is achieved by choosing the applied magnetic field such that the
4
cyclotron frequency of the electrons is nearly equal to the desired microwave frequency of the
device. The benefits of gyrodevices are derived from the combination of the cyclotron resonance
interaction and the fast-wave interaction circuit. In fast-wave circuits, the electromagnetic field
strength can be quite high, independent of the proximity of the metallic circuit structure. This
enables the electron beam to be situated in regions of high field, for optimum coupling, without
necessarily placing the beam too close to the circuit. The interaction mode is selected by the
magnitude of the applied magnetic field, which determines the electron cyclotron frequency, and the
placement of the electron beam. If a high-order waveguide or cavity mode is selected, the transverse
dimensions of the interaction structure can be several times the free-space wavelength. The larger
interaction structure enhances the ability of the device to generate high output powers and
increasingly greater frequencies. However, the necessity of a superconducting magnet to generate
the magnetic field required for interactions at frequencies above 30 GHz can be construed as a
disadvantage of the device. DNP gyrotrons have been developed from 140-527 GHz with output
power of 10-100 W [1-4, 9, 10-14, 46-48].
In traditional slow-wave devices, such as klystrons [44-45], the interaction circuits are designed
to reduce the phase velocity of the electromagnetic wave to values below the speed of light so that
synchronism between the beam and wave can be maintained. The electric field strength falls off
rapidly with distance from the circuit structure. Typical transverse circuit dimensions required to
effectively slow the phase velocity of the wave are on the order of 10% of the free space wavelength.
These small circuit sizes, necessary to achieve the interaction in a slow-wave device, severely limit
the peak and average power capabilities, particularly at millimeter wave frequencies. Klystrons have
been developed for DNP applications at up to 263 GHz with 1-2 W of cw output power [24-25].
Similarly to gyrotrons, klystrons are driven by high voltage power supplies with substantial beam
energy (kilowatts) and associated cooling requirements.
Solid-state sources are compact, easy to use, and require no special facility requirements. A
low frequency synthesizer (typically 9-20 GHz) is followed by set of multipliers and amplifier to reach
target frequency and power. Several sources can be combined for higher power, but mixers at 263
GHz frequency have high insertion loss and component heating is one of the challenges in extending
to higher power. Solid-state sources have been used for DNP experiments at helium temperature
[4,16,26] and are commonly used in EPR spectrometers, however to date, the output power of solid-
state sources has not been sufficient for MAS DNP experiments above 77K.
Continuous wave (CW) gyrotron oscillators capable of producing 50 W at 263 GHz and 395 GHz,
and 20 W 527 GHz have been developed by Bruker and CPI for 400 MHz, 600 MHz, and 800 MHz
NMR systems, respectively. In Figure 2, a solid model and photograph of a typical DNP gyrotron are
shown. An annular electron beam is produced by a single-anode magnetron injection gun operating

5
at cathode voltages in the 14 – 18 kV range with cathode currents up to 200 mA. The magnetic field
required for the gyrotron interaction at the second harmonic of the electron cyclotron frequency,
approximately 5 T, 7 T, and 10 T for the 263 GHz, 395 GHz, and 527 GHz devices, respectively, is
produced by a cryogen-free superconducting magnet, which is not shown in the figure. The electron
beam is accelerated toward the anode, which is at ground potential, and travels in a tapered beam
tunnel region as its diameter is compressed by the increasing magnetic field.

VAC ION COLLECTOR


PUMP

ADJUSTABLE
MIRROR
WINDOW

STEERING/SHAPING
MIRRORS

LAUNCHER

CAVITY

ANODE-BEAM
TUNNEL

SINGLE-ANODE
MAGNETRON
INJECTION GUN

Figure 2: Solid model and photograph of a typical 395 GHz, 50 W gyrotron.

The interaction cavity is located at the peak of the axial magnetic field where the beam
diameter is at its minimum. The high-order transverse electric (TE) interaction mode is carefully
selected based on several criteria including the electron beam diameter in the cavity, the power
loading on the cavity walls, and competition with other cavity modes. Low-order cylindrical cavity
modes, such as the TE11 or TE01 modes, are not practical because of the small-diameter electron
beam that would be required for the interaction as well as the small diameter cavity that would lead
to large ohmic power densities on the cavity walls and would be difficult to effectively cool. Higher
order modes offer the possibility of larger cavities with lower power densities on the cavity walls

6
excited by larger diameter electron beams. The interaction modes for the 263 GHz, 395 GHz and
527 GHz gyrotrons are the TE11,2, the TE10,3, and TE14,3 modes, respectively.
Because these high-order interaction modes are not suitable for efficient transmission in
smooth or corrugated waveguides, an internal mode converter is used to transform the interaction
mode to a Gaussian beam. The mode converter consists of a helically cut launcher with specially
designed wall-perturbations to increase the Gaussian mode content of the launched beam , as well
as four or five mirrors that serve to both shape the Gaussian beam so that the waist is the suitable
size for transmission through the vacuum window and steer the beam to the window . The final
mirror has the capability of being moved radially and tilted in the horizontal and vertical directions to
ensure that the beam exits through the center of the single-disk alumina vacuum window and is
travelling perpendicular to the window plane. The internal converter also serves to separate the
generated electromagnetic wave from the electron beam, which is guided by the magnetic field and
continues along the gyrotron axis until it is incident on the walls of the collector. The cryogen-free
gyrotron magnet design allows for more compact magnet and shorter path from microwave cavity to
output window than previous generation liquid cooled 263 GHz gyrotron [9]. The vacuum is
maintained by a 2 liter/second vac ion pump that uses the fringing field of the superconducting
magnet rather than the more typically employed permanent magnet.

70 395.18
60 395.17
Microawve Power (W)

50 395.16
Frequency (GHz)

40 395.15
30 395.14
20 395.13
10 395.12
0 395.11
14 14.5 15 15.5 16
Cathode Voltage (kV)

Figure 3: Measured output power (blue diamonds) and frequency (red squares) as a function of
cathode voltage for a 395 GHz gyrotron. All parameters other than cathode voltage are held fixed.

Rigorous tests of each gyrotron are performed to demonstrate the required performance
characteristics including a output power, the ability to tune power smoothly from 50 W to less than 3

7
W, the power and frequency stability, and the Gaussian output beam characteristics and quality.
Typical measured results for a 395 GHz gyrotron are shown in Figure 3, where the output power and
frequency as a function of cathode voltage is plotted. As seen in the figure, the power can be
smoothly varied from 60 W to 3 W by reducing the cathode voltage from 15.7 kV to 14.2 kV and
leaving all other parameters fixed. Output power can also be adjusted by varying the electron beam
current. Figure 4 shows infrared (IR) images of the output beam from a 395 GHz gyrotron. The
output beam is incident on a target that can be moved in a plane perpendicular to the window.
Images at distances 51, 61, 71, and 81 cm from the gyrotron output window are shown in the figure.
The cross hairs represent the position of the window center and the circle shown overlaying each
image is 5 cm in diameter. As seen in the figure, the output beam is a high-quality Gaussian beam
which is centered on the window and travelling perpendicular to the window plane. A modal
decomposition of the measured output beam shows greater than 97% overlap with the beam that is
ideal for injection into the 17 mm diameter corrugated waveguide. The beam waist expansion is
plotted as a function of target distance and shows good match to theoretical expansion of gaussian
beam in free space. As part of the experimental demonstration, the output beam from the window
is injected into the first section of corrugated waveguide and additional IR images of the beam exiting
the guide are made. Images of the beam from the corrugated guide typically show greater than 98%
overlap between the measured beam and the HE11 waveguide mode.

(a) (b)

51 cm 61 cm

71 cm 81 cm

Figure 4. Output beam characterization of the 395 GHz gyrotron output beam. (a) infrared images:
the cross hairs represent the position of the window center and the circle is 5 cm in diameter. (b)
beam expansion: the solid line represents the theoretical expansion of the TEM00 mode from a
17mm diameter corrugated guide and the red diamonds and green squares show the measured
vertical and horizontal beam radii.

8
The gyrotron output beam is transmitted to the NMR probe via corrugated waveguide to retain high
gaussian beam purity and minimize power losses in the transmission line. [49]. The corrugations (1/4
λ depth and width with 1/3 λ period) are cut for 263 GHz or 440 GHz with the later used on the 395
and 527 GHz assembly due to the broadband nature of corrugated waveguide. The waveguide inner
diameter (ID) is 19.3 mm at 263 GHz and 17 mm at 440 GHz. Three 90 degrees miter bends are
included to complete the path from gyrotron window to NMR probe base with a corrugated taper at
the probe base to transition to the probe waveguide (7.6 mm ID for 263 GHz and 7.7 mm ID for 395
and 527 GHz). A directional coupler in the transmission line allows for power and frequency
measurements during DNP experiments. Low-loss beam transmission has been confirmed with
power measurements at the gyrotron output waveguide, NMR probe base and end of probe
waveguide. For example, at 395 GHz just 10% loss was measured over 3030 mm total length of 17
mm ID waveguide and 3 miter bends. An additional 20% power loss is measured across the taper,
700 mm length of probe waveguide and two miter bends by the sample. Similar results are obtained
at 263 and 527 GHz. Figure 5 shows the DNP signal enhancement at 395 GHz on a standard sample
with AMUPOL polarizing agent as a function of microwave power. In this case, 15 W of microwave
power is sufficient to reach saturation point. Samples with different dielectric constants, microwave
penetration depth, polarizing agents, sample temperature, and MAS frequency can vary in their
power requirement to reach the plateau.

150
DNP Signal Enhancement

100

50

0
0 5 10 15 20
Microwave Power (W)

Figure 5. 395 GHz DNP signal enhancement as a function of microwave power at the end of the
probe waveguide. Sample of 1 M 13C Urea in glycerol d8/D2O/H2O (60/30/10) with 10 mM AMUPOL
at 108 K, 12 kHz MAS, 3.2 mm sapphire rotor.

9
3. Solid-state DNP NMR probes and cryogenic MAS
The dominant technique and accompanying instrumentation in solid-state NMR requires fast sample
spinning at the magic angle (MAS). Henceforth in our excursion we will focus on MAS NMR probes
suitable for low-temperature and DNP, in the following abbreviated as LT.MAS DNP probes. There
are applications with static but oriented samples, e.g. oriented biomembranes, which aim at DNP
NMR as well [50-52]. Much of the principal probe design, for example, the corrugated wave guide
inside the probe, the miter bends and the radial irradiation scheme has been derived from earlier
work in the Griffin group [46-48]. The microwave beam arrives at the bottom of the LT.MAS DNP
probe (cf. Fig. 1) is reflected by a 90° miterbend, propagates upwards through a taper and cylindrical
corrugated wave guide, changes direction through one or more miterbends and finally enters the
MAS system with the NMR coil, MAS rotor and sample inside orthogonal to the rotor axis (radial
irradiation scheme). Other schemes, for example an axial irradiation scheme are principally also
possible. Both schemes have their own advantages and disadvantages. The axial scheme, i.e., the
microwave beam is incident along the rotor axis [20,21,34], is very sensitive against coupling
conditions and geometric constraints imposed by the MAS air bearing system. The radial scheme, as
discussed in the following, depends on the geometry of the rf coil, the latter acting as a diffraction
grating, and the geometry and material properties of the MAS rotor. There are quite a few probe
designs and microwave irradiation schemes reported in the literature. Revolution NMR offers the
commercialization of probes [72] according to the Thurber-Tycko design [15-17], where MAS is run
with cryogenic nitrogen and the sample is cooled with cryogenic helium. Doty [22,33] has proposed
probe designs addressing cold NMR coils and cold or warm samples including DNP capabilities in such
a probe for a wide sample temperature range from 30K to 300K and higher. At MIT probes have
been build based on a balanced rf transmission line that allow remote tuning and matching [53]. Last
but not least Maly, et al. at Bridge-12 [54-55] came up with the idea to incorporate NMR DNP probe
and gyrotron within the same cryomagnet in such a way to arrive at a very compact system.
In order to successfully design, construct, and build NMR DNP MAS probes and equipment to
efficiently cool the NMR sample to cryogenic temperatures under MAS, several technological
challenges have to be overcome. The cooling system must produce cold gas flows for MAS with
typical gas flow rates on the order of 30 to 100 standard liters per minute, depending on the MAS
system and the MAS speed. With gaseous nitrogen temperatures down to ca. 90 K can be achieved,
depending on the gas pressures for bearing and drive and the MAS spinning rate applied. The heat
exchanger design needs to address these boundary constraints. An example for such a heat
exchanger suitable for MAS at low temperature has been described in the NMR literature in [56]. The
possibility to lower the temperature below the condensation point of liquid nitrogen by using
cryogenic helium is even more promising for many DNP applications. This requires heat exchangers
10
for helium to operate for MAS conditions, which are different from those for nitrogen. An early
example is Potter’s proposal [57]. MAS with cryogenic helium has been accomplished by Yannoni and
coworkers [58-59]. Samoson reported on MAS with cryogenic helium [60] as well as Levitt, et al. in
[61]. These experiments were conducted with helium heat exchangers that did not form a closed-
loop system and consequently very high helium consumption. Doty [62] described a heat exchanger
system that works over a wide variable temperature range. Thurber, et al. [15-17] linked DNP and
cryogenic helium MAS. Fujiwara , et al. [63-64] went one step further and introduced a closed-loop
system for cryogenic helium MAS. For such a closed-loop helium system much effort has been spent
also at CEA Grenoble in the group of dePaepe, et al. [65] demonstrating MAS DNP at cryogenic
temperatures using a closed-loop helium system allowing much higher MAS speeds than previously
possible at cryogenic nitrogen temperatures or with similar setups with closed-loop helium MAS
described in refs. [63-64] .

From the NMR rf side, MAS DNP NMR probes are very similar to regular solid-state MAS probes. They
are equipped with two or three channels, one tuned to 1H, the other two to heteronuclei like 13C, 15N,
29
Si, 27Al. The particularity of DNP MAS probes consists of the specific design to enable the irradiation
of the sample by millimeter waves (high-frequency microwaves) of a given polarization of the electric
field. The DNP NMR enhancement factors, measured by the ratio of NMR signal amplitudes between
cw microwave irradiation switched on and off, critically depend on the local microwave field
amplitude at the site of unpaired electron spins inside the sample. The MAS system with its NMR rf
coil, the rotor and the sample inside the rotor have a geometry which is not close to a structure
which would be ideal for a single fundamental mode of a resonant cavity. The system is highly
overmoded and for that reason it is still essential to obtain a fairly detailed picture of the microwave
field amplitude distribution. It is difficult to achieve that in situ with the complete MAS system and rf
coil involved, though simulations of the electromagnetic field distribution turn out to be very
beneficial. It is nevertheless required to validate such simulations by experiment, at least to verify
simulation data vs. experimental data for a simplified system. Such a validation is shown in Fig. 6,
which represents the forward scattering of a millimeter wave incident through an aperture on an
3.2mm NMR rf solenoidal coil. The coil acts like a grating and leads to multiple diffraction of the wave
behind the coil. The figure shows on the left the experimentally measured electric field amplitude of
the propagating wave in two orthogonal planes behind the coil, on the right side an electromagnetic
simulation using the simulation software CST Microwave Studio (CST GmbH, Darmstadt). Details on
the measurements of field maps and the simulations of such can be found in [66]. There is earlier
work on electromagnetic simulations of mm wave propagation and reflection of other coil
geometries [67] and millimeter wave propagation through radially irradiated MAS rotors [68], the

11
latter demonstrating the diffractive influence of the MAS rotor material (zirconia ceramics or
sapphire). In a study by Nanni, et al. [69] the millimeter wave field distribution in a complete MAS
system was investigated. In the following some of the subtleties in the millimeter wave field
distribution inside a real MAS system are discussed. The system studied by electromagnetic
simulations using CST Microwave Studio (CST GmbH, Darmstadt) is shown in Fig. 7. On the left side a
photograph of the upper part of a widebore 3.2mm MAS DNP probe is displayed, on the right side
the corresponding axial cross section through the MAS system. As indicated, for the 3.2mm MAS
system, the millimeter wave beam enters through an aperture, propagates radially through the NMR
rf coil and the MAS rotor.

Figure 6: Experimental vs. simulated millimeter wave forward scattering field amplitude distribution
at 263 GHz revealing the field amplitude in two different planes as indicated. The comparison serves
for validating the simulation result (simulation data: Armin Purea, Bruker Biospin 2012)

12
Figure 7: Left – upper part of the interior of an LT.MAS DNP probe with MAS system, miter-bended
waveguide upper end and pneumatic sample insert/eject system. Right – Axial cross section through
a 3.2mm MAS system showing the plane of the incident millimeter-wave front and the interior of an
MAS system constructed for millimeter-wave irradiation in direction along the rotor radius (radial
irradiation)

Fig. 8 demonstrates more examples of field distributions shown as snapshots for three different
frequencies – 263 GHz, 395 GHz, and 527 GHz. It becomes apparent that the mm wave field is far
from being spatially homogeneous, rather it appears periodic, depending on the wavelength, the
dielectric material, the detailed diffraction pattern, and reflection of the beam. Standing waves occur
generated by forward beam propagation and backward propagation caused by reflection. These
standing wave patterns represent one possibility to improve the millimeter wave amplitude inside
the sample. Further, if one keeps in mind that the MAS rotor spins with several kHz it becomes clear
that an unpaired electron spin at a given spatial position inside the rotor experiences a time
dependent electromagnetic field on two time scales – the fast one given by the resonant microwave
frequency and a slow one given by MAS. When one detects the DNP NMR enhancement one
measures an ensemble average of spins and microwave fields over the sample volume and an
additional slow modulation time average by MAS. With this perspective it becomes difficult to
postulate simple straightforward relationships between local features in the sample and the
millimeter wave field. It is rather the millimeter wave field average over the sample that is important.

13
Figure 8: Simulation of the microwave magnetic field magnitude distribution in a radial plane (top)
and an axial plane (bottom) inside a 3.2mm MAS system for three different frequencies of an
incident mm wave beam (incident from the right side) with the mm wave E field polarized parallel to
the rotor axis (simulation data Armin Purea, Bruker Biospin 2014).

In a recent article, Kubicki et al. [70] reported that the DNP enhancement factor of frozen solutions
of glycerol/water or TCE with various radicals (TekPol, bCtbK, bTbK, AMUPOL, TOTAPOL) and a
dispersion of dielectric particles (powdered KBr, Sapphire, NaCl, PTFE, CaF2) embedded in the frozen
solution significantly increases as compared to the bulk solution (without dielectric particles). It could
be excluded that the effect originates from different microwave heating characteristics when
comparing frozen bulk solutions and solutions with particles. Since the increase of DNP enhancement
factors strongly depends on the kind of dielectric particles embedded, it seems plausible that this
dielectric effect comes from the diffractive behavior of the incident millimeter wave beam inside the
frozen solution/particles, such that the diffraction (strongly depending on the spatial distribution of
the dielectric properties) and scattering/reflection should be dependent on the dielectric constant of
the particles. The simulated magnetic field magnitude distribution as shown in Fig. 9a reveals that
this is indeed the case. For a given filling factor of 0.63 it becomes evident, that locally the magnetic
field magnitude may increase, caused by the random diffraction, reflection, and superposition of mm
waves by the dielectric particles. But it is not only the local field distributions that change. As exemp-

14
Figure 9: (a) Magnitude distribution of the millimeter wave magnetic field inside a 3.2mm MAS rotor
filled with frozen solution of H2O/D2O/glycerol (left) and filled with frozen solution and dielectric
particles embedded, filling factor 0.63 (right three panels). The frozen solution has a real part of the
dielectric constant equal to 2.5 (left), the dielectric constant of the particles is varied from 2.5 to 11
(right three panels). (b) Spatially averaged millimeter wave magnetic field magnitude for a frozen
solution with dielectric constant of 2.5 and dielectric particles with dielectric constants varied from
values of 1 to 11 for various filling factors. The solid curves are meant to guide the eye (simulation
data: Armin Purea, Bruker Biospin, 2014).

15
lified in Fig. 9b also the spatially averaged magnetic mm field magnitude becomes dependent on the
dielectric constant of the embedded particles and of the filling factor. For dielectric constants larger
than 2.5 and up to 11 one finds a systematic increase of the millimeter magnetic field by a factor of
ca. 2 in accordance to and as an explanation for the increased DNP enhancement factors observed
experimentally.

As we have discussed now with case studies, some example results shown in Figs. 6-9, there are
several important geometry and material factors that determine the performance of DNP
experiments: (i) coupling geometry of the millimeter wave beam into the MAS system, (ii)
propagation of the millimeter wave beam through NMR rf coil and MAS rotor materials, (iii) the
dielectric properties of the sample itself, and (iv) the radical molecules used as source of electron
spin polarization. These factors should be known since they appear as boundary conditions when one
investigates the detailed microscopic mechanisms of DNP.

4. DNP radicals and DNP sample preparation

Samples for DNP experiments must contain a paramagnetic center, an EPR-active spin often referred
to as the polarizing agent, which is either added to the sample or an endogenous species. The first
demonstration of DNP in frozen aqueous media with a focus toward biological applications was
demonstrated by the Griffin group using 4-amino TEMPO nitroxide radical in a mixture of water and
glycerol [73], a common EPR solvent. The water/glycerol solution forms a cryogenic glass which is
beneficial for microwave transmission and required for keeping the polarizing agent well dispersed
upon freezing. The water/glycerol/nitroxide matrix was demonstrated to be applicable to soluble
proteins [74] and extended to membrane proteins and virus particles [75]. The polarizing agents
themselves have greatly evolved since the original 4-amino TEMPO experiments. Another ground-
breaking step by the Griffin group was the introduction of biradical polarizing agents [76], two
covalently bound nitroxides, allowing for much higher DNP efficiency and at a lower total
paramagnetic concentration. These biradical polarizing agents have been further investigated and
optimized by the groups of Griffin and Swager at MIT [77] and Tordo at the University of Aix-
Marseille [78, 79]. A particular challenging category of biological samples has been lipid bilayers
which often show low DNP efficiency; recent progress with spin-labelled lipids offers a promising
outlook [80,81]. The ultimate goal in DNP experiments for solid state NMR applications is sensitivity-

16
per-unit-time and additional factors such as polarization build-up time, paramagnetic quenching [82],
and depolarization upon MAS [83,84] must also be considered.

Another set of key contributions, this time toward material science applications, were made by the
group of Emsley at ENS Lyon who demonstrated the first applications of DNP on porous materials
[85] introducing incipient wetness impregnation and the use of organic solvents for sample
preparation [86], and optimized biradicals [87]. The fields of biological and materials applications are
covered in details in several other articles in this perspective JMR issue.

We conclude, as we transition to outlook, with a brief mention of recent work on polarizing agents
for applications at high field (800 MHz, 527 GHz), where the DNP efficiency of traditional cross effect
experiments with nitroxide biradicals drops severely [88]. Narrow-line radicals, such as bdpa and
trityl, with the Overhauser mechanism offer opportunities for more favorable scaling with increasing
frequency and reduced microwave power requirements [89]. The introduction of trityl/nitroxide
biradicals has yielded the highest DNP signal enhancements at 527 GHz to date, with DNP signal
enhancements actually increasing with frequency, and provided valuable insight into the factors
contributing to efficient DNP [90].

5. Outlook
We close our survey of instrumentation aspects of solid-state DNP by taking a tentative look to the
next-future developments with some of these conclusions resulting from our own work at present
time. Without the intention of putting any strategic priority to the order of items in the following
list, the present and near-future development of instrumentation for solid-state DNP MAS NMR
might include:

(i) Faster MAS at low temperature by adapting the small diameter MAS systems to the cryogenic
environment necessary for solid-state DNP.
(ii) MAS at cryogenic temperatures below 77K with closed-loop or low helium consumptions systems.
The first steps in that direction have already been taken or are presently in progress by various
groups [15-17, 63-65, 71-72].
(iii) Dedicated probe hardware for DNP on static oriented biomembrane samples [50-52] taking into
account the peculiarity of sample geometry and structure of typical oriented bio-membrane samples
(iv) Solid-state MAS DNP employing microwave sources with powers lower than gyrotron power
levels (10-50W) to provide lower-cost, economic instrumentation solutions also with reduced

17
infrastructure requirements. The first steps have also been taken in this area with klystrons and
solid-state sources as presented in the introduction.
(v) MAS DNP at fields even higher than 800 MHz/527GHz.
(vi) Improvement and optimization of the coupling of the incident microwave field to the MAS NMR
sample by applying more sophisticated resonator designs and principles from quasioptical systems.
(vii) Solid-state DNP with microwave sources capable of short (nanosecond to microsecond
timescale) phase coherent pulses, development of pulsed DNP methods, and/or the availability of
sweepable magnets.
(viii) New developments and optimizations in the field of DNP radicals.

Faster MAS (i) is almost self-explanatory for solid-state NMR spectroscopists – the reasons for it are
the same as for solid-state NMR close to room temperature: higher spectral resolution. These
developments are based on 1.3 and 1.9mm MAS systems and preliminary results already show good
prospect for DNP MAS applications, with high DNP signal enhancement obtained at spinning speeds
at 25 kHz and higher. MAS with cryogenic helium at temperatures below 77 K (ii) has three aspects.
First, experiments at these low temperatures promise even higher DNP enhancement factors.
Second, spinning with helium gas offers the potential for faster spinning than with cold nitrogen gas.
Third, the microwave power requirements at lower temperature become less severe such that lower-
power sources (iv) could be used. Solid-state NMR of oriented biomembranes are very often
characterized by a notoriously low signal-to-noise ratio, thus in this particular subfield (iii) the
benefits of successful DNP experiments can hardly be overestimated. The search and development of
microwave sources different from the gyrotron with smaller power levels (iv) has mainly economic
reasons, nevertheless for wider adoption also depends on the possibilities to go to lower
temperatures (ii), the accomplishments of probe developments (vi), and research in radical chemistry
(viii). The quest for ever higher static NMR fields (v) applies to DNP NMR as it does to NMR without
DNP, very often driven by the need of higher spectral resolution. Probe developments (vi) have been
so far in their “infant stage”. With the current requests for higher MAS speeds, lower temperatures,
and improved microwave coupling there is more to be expected in the next few years. Coherent
pulse capability (vii) in DNP NMR appears to be the ambition of genuine NMR/ESR/DNP
spectroscopists with the potential to truly unify ESR and NMR. This field is covered by another
contribution in this special issue. Last but not least, DNP takes polarization from electron spins and
transfers it to the nuclear spin system. Thus the chemistry of radicals (viii) providing the spin
polarization appears as one of the most important molecular engineering tools for efficacious DNP
NMR experiments.

18
Acknowledgements

The authors are indebted to Dr. Werner Maas, Leo Tometich, Patrick Saul, Bryce Smith, Christian
Reiter, Dr. Armin Purea, Dr. Kevin Felch, and Philipp Borchard.

6. References

[1] L.R. Becerra, G.J. Gerfen, R.J. Temkin, D.J. Singel, R.G. Griffin, Dynamic nuclear polarization with a
cyclotron resonance maser at 5 T, Phys. Rev. Lett. 71 (1993) 3561-3564.

[2] L.R. Becerra, G.J. Gerfen, B.F. Bellew, J.A. Bryant, D.A. Hall, S.J. Inati, R.T. Weber, T.F. Prisner,
A.E. McDermott, K.W. Fishbein, K.E. Kreischer, R.J. Temkin, D.J. Singel, R.G. Griffin,
A spectrometer for dynamic nuclear polarization and electron paramagnetic resonance at high
frequencies, J. Magn. Reson. A 117 (1995) 28-40.

[3] G.J. Gerfen, L.R. Becerra, D.A. Hall, R.G. Griffin, R.J. Temkin, D.J. Singel, High frequency (140GHz)
dynamic nuclear polarization: polarization transfer to a solute in frozen aqueous solution, J. Chem.
Phys. 102 (1995), 9494-9497.

[4] V. Weis, M. Bennati, M. Rosay, J.A. Bryant, R.G. Griffin, High-field DNP and ENDOR with a novel
multiple-frequency resonance structure, J. Magn. Reson. 140 (1999) 293-299.

[5] T.R. Carver, C.P. Slichter, Polarization of nuclear spins in metals, Phys. Rev. 92 (1953), 212-213.

[6] R.A. Wind, M.J. Duijvestijn, C. Van der Lugt, A. Manenschijn, J. Vriend, Applications of
dynamic nuclear polarization in 13C NMR in solids, Prog. NMR Spectrosc. 77 (1985) 33-67.

[7] R.A. Wind, Dynamic nuclear polarization and high-resolution NMR of solids, Encyclopedia
of Magnetic Resonance, eds. R.K. Harris, R.E. Wasylishen, Wiley, Chichester 2007.

[8] Themed issue: Dynamic Nuclear Polarization, Phys.Chem.Chem.Phys. 12 (22) (2010), 5725-5928.

[9] M. Rosay, L. Tometich, S. Pawsey, R. Bader, R. Schauwecker, M. Blank, P.M. Borchard,


S.R. Cauffman, K.L. Felch, R.T. Weber, R.J. Temkin, R.G. Griffin, W.E. Maas, Solid-state
dynamic nuclear polarization at 263 GHz: spectrometer design and experimental results,
Phys.Chem.Chem.Phys. 12 (2010) 5850-5860.

[10] Y. Matsuki, H. Takahashi, K. Ueda, T. Idehara, I. Ogawa, M. Toda, H. Akutsu, T. Fujiwara, Dynamic
nuclear polarization experiments at 14.1T for solid-state NMR, Phys.Chem.Chem.Phys. 12 (2010)
5799-5803.

[11] R. Ikeda, Y. Yamaguchi, Y. Tatematsu, T. Idehara, I. Ogawa, T. Saito, Y. Matsuki, T. Fujiwara,


Broadband continuously frequency tunable gyrotron for 600 MHz DNP-NMR spectroscopy, Plasma
Fusion and Research 9 (2014), 1206068.

[12] T. Idehara, Y. Tatematsu, Y. Yamaguchi, E.M. Kkhutoryan, A.N. Kuleshov, K. Ueda, Y. Matsuki, T.
Fujiwara, The development of 460 GHz gyrotrons for 700 MHz DNP NMR spectroscopy, J. Infrared
Milli Terahz Waves 36 (2015) 613-627.

19
[13] V. Denysenkov, M. J. Prandolini, M. Gafurov, D. Sezer, B. Endeward, T.F. Prisner, Liquid-state
DNP using a 260 GHz high power gyrotron, Phys.Chem.Chem.Phys. 12 (2010) 5786-5790.

[14] M.Yu. Glyavin, A.V. Chirkov, G.G. Denisov, A.P. Fokin, V.V. Kholoptsev, A.N. Kuftin, A.G. Luchinin,
G.Yu. Golubyatnikov, V.I. Malygin, M.V. Morozkin, V.N. Manuilov, M.D. Proyavin, E.V. Sokolov,
A.I. Tsvetkov, V.E. Zapevalov, Experimental test of a 263 GHz gyrotron for spectroscopic
applications and diagnostics of various media, Rev. Sci. Instrum. 86 (2015) 054705.

[15] K.R. Thurber, R. Tycko, Biomolecular solid state NMR with magic-angle spinning at 25K, J. Magn.
Reson. 195 (2008) 179-186.

[16] K.R. Thurber, W.-M. Yau, R. Tycko, Low-temperature dynamic nuclear polarization at 9.4 T with a
30 mW microwave source, J. Magn. Reson. 204 (2010) 303-313.

[17] K.R. Thurber, A. Potapov, W.-M. Yau, R. Tycko, Solid state nuclear magnetic resonance with
magic-angle spinning and dynamic nuclear polarization below 25 K, J. Magn. Reson. 226 (2013) 100-
106.

[18] http://www.bridge12.com

[19] http://www.swissto12.com

[20] K.J. Pike, T.F. Kemp, H. Takahashi, R. Day, A.P. Howes, E.V. Kryukov, J.F. MacDonald, A.E.C. Collis,
D.R. Bolton, R.J. Wylde, M. Orwick, K. Kosuga, A.J. Clark, T. Idehara, A. Watts, G.M. Smith,
M.E. Newton, R. Dupree, M.E. Smith, A spectrometer designed for 6.7 and 14.1 T DNP-enhanced
solid-state MAS NMR using quasi-optical microwave transmission, J. Magn. Reson. 215 (2011) 1-9.

[21] K.J. Pike, Quasi-optical techniques for high-frequency DNP, 4th International DNP Symposium
Elsinore (Denmark), August 2013,
http://www.dnpsymposium.org/public/conferences/1/schedConfs/1/program-en_US.pdf, p. 115.

[22] F. Horii, T. Idehara, Y. Fujii, I. Ogawa, A. Horii, G. Entzminger, F.D. Doty, Development of DNP-
enhanced high-resolution solid-state NMR system for the characterization of the surface structure
of polymer materials, J. Infr. Millimeter. THz Waves 33 (2012) 756-765.

[23] S. Alberti, J.-Ph. Ansermet, K.A. Avramides, F. Braunmueller, P. Cuanillon, J. Dubray, D. Fasel,
J.-Ph. Hogge, A. Macor, E. de Rijk, M. Da Silva, M.Q. Tran, T.M. Tran, Q.Vuillemin, Experimental
study from linear to chaotic regimes on a terahertz-frequency gyrotron oscillator, Phys. Plasmas
19 (2012) 123102.

[24] W-M. Yau, K.R. Thurber, R. Tycko, Synthesis and evaluation of nitroxid-based oligoradicals for
low-temperature dynamic nuclear polarization in solid state NMR, J. Magn. Reson. 244 (2014) 98-106
and A. Potapov, W.-M. Yau, R. Ghirlando, K.R. Thurber, R. Tycko, J. Am. Chem. Soc. 137 (2015),
8294-8307.

[25] D. Berry, H. Deng, R. Dobbs, P. Horoyski, M. Hyttinen, A. Kingsmill, R.MacHattie, A.Roitman,


E. Sokol, B. Steer, Practical Aspects of EIK Technology, IEEE Trans. Electronic Devices 61 (6) (2014)
1830-1835.

[26] B.D. Armstrong, D.T. Edwards, R.J. Wylde, S.A. Walker, S. Han, A 200 GHz dynamic nuclear
polarization spectrometer, Phys.Chem.Chem.Phys. 12 (2010) 5920-5926.

20
[27] http://vadiodes.com

[28] A. Macor, J.-Ph. Ansermet, E. de Rijk, “Passive components for millimeter, submillimeter and
terahertz electromagnetic waves by piling up successive layers of material”, patent
application (PCT) WO 2012/076994 A1, June 2012.

[29] A. Macor, E. de Rijk, D. Fivat, “Flanges for connection between corrugated wave-guiding
modules“,patent application (PCT) WO 2014/174494 A2, October 2014.

[30] A. Macor, E. de Rijk, G. Annino, “Over-moded resonant cavity for magnetic resonance based on a
photonic band gap structure”, patent application (PCT) WO 2013/057688 A1, July 2013.

[31] E. de Rijk, A. Macor, J.-P. Hogge, S. Alberti, J.-P. Ansermet, “Note: Stacked rings for terahertz
waveguiding”, Rev. Sci. Instr. 82, 066102, 2011.

[32] J.R. Sirigiri, T. Maly, “Integrated high-frequency generator system utilizing the magnetic field of
the target application”, US patent 8,786,284 B2, July 2014.

[33] D. Doty, “A tunable microwave resonator for static dynamic nuclear polarization”, patent
application (PCT) WO 2015/107512 A1, July 2015.

[34] A. Purea, F. Engelke, A. Krahn, “Microwave coupler for optimizing an NMR probe head for MAS-
DNP“,patent application (PCT) WO 2015/018640 A1, February 2015.

[35] A. Krahn, F. Engelke, “Microwave resonator with distributed Bragg reflector”, patent
application (PCT) WO 2015/04586 A1, February 2015.

[36] G. Annino, A. Macor, E. De Rijk, S. Alberti, “Magnetic resonance hyperpolarization probe head”,
patent application (PCT) WO2013/000508 A1, January 2013.

[37] G. Annino, A. Macor, E. De Rijk, S. Alberti, “Magnetic resonance hyperpolarization and multiple
irradiation probe head”, patent application (PCT), WO 2013/000964 A1, January 2013, patent
application (PCT) WO2013/000508 A1, January 2013.

[38] G. Annino, A. Macor, E. De Rijk, S. Alberti, “Magnetic resonance hyperpolarization and multiple
irradiation probe head”, US patent application, US 2014/17988 A1, May 2014.

[39] E. Bouleau, G. de Paepe, ”Probe, device, and method for nuclear magnetic resonance analysis
with magic-angle spinning”, patent application (PCT), WO2014/102347 A1, July 2014.

[40] E.Bouleau, G. de Paepe, ”Probe, device, and method for nuclear magnetic resonance analysis”,
patent application (PCT), WO2014/102348 A1, July 2014.

[41] E. Bouleau, G. de Paepe, ”Probe, device, and method for nuclear magnetic resonance analysis
with magic-angle spinning”, patent application (PCT), WO2014/102348 A1, July 2014.

[42] G.S. Nusinovich, Introduction to the Physics of Gyrotrons, Johns Hopkins University Press,
Baltimore, MD, 2004.

21
[43] K. Felch, K., B.G. Danly, H.R. Jory, K.E. Kreischer, W. Lawson, B. Levush, R.J. Temkin,
Characteristics and Applications of Fast-Wave Gyrodevices, Proceedings of the IEEE, Vol. 87 (5)
(1999), 752-781.

[44] A.S. Gilmour, Jr., Klystrons, Traveling Wave Tubes, Magnetrons, Cross-Field Amplifiers, and
Gyrotrons, Artech House, Norwood, MA, 2011.

[45] R.H. Varian, S.F. Varian, A High Frequency Oscillator and Amplifier, J. Appl. Phys.10 (1939) 321.

[46] A.B. Barnes, M.L. Mak-Jurkauskas, Y. Matsuki, V.S. Bajaj, P.C.A. van der Wel, R. DeRocher,
J. Bryant, J.R. Sirigiri, R.J. Temkin, J. Lugtenburg, J. Herzfeld, R.G. Griffin, Cryogenic sample exchange
NMR probe for magic angle spinning dynamic nuclear polarization, J. Magn. Reson. 198 (2009) 261-
270.

[47] A.B. Barnes, E. Markhasin, E. Daviso, E.A. Nanni, S.K. Jawla, E.L. Mena, R. DeRocher, A. Thakkar,
P.P. Woskov, J. Herzfeld, R.J. Temkin, R.G. Griffin, Dynamic nuclear polarization at 700 MHz/460 GHz,
J. Magn. Reson. 224 (2012) 1-7.

[48] P.P. Woskov, V.S. Bajaj, M.K. Hornstein, R.J. Temkin, R.G. Griffin, Corrugated waveguide and
directional coupler for cw 250 GHz gyrotron DNP experiments, IEEE Trans. MTT 53 (2005) 1863-1869.

[49] J. L. Doane, (1985) in: Infrared and Millimeter Waves, ed. K. J. Button, Academic Press, Inc.,
Editon edn., 1985, vol. 13, pp. 123-170.

[50] E.S. Salnikov, O. Ouari, E. Koers, , H. Sarrouj, , T. Franks, M. Rosay, S. Pawsey, C. Reiter,
P. Bandara, H. Oschkinat, P. Tordo, F. Engelke, B. Bechinger, Developing DNP/solid-state NMR
spectroscopy of oriented membranes, Appl. Magn. Reson. 43 (2012) 91-106.

[51] E.S. Salnikov, E. Glattard, H. Sarrouj, C. Aisenbrey, O. Ouari, P. Tordo, F. Engelke, F. Aussenac,
B. Bechinger, Lipid-mediated polypeptide interactions in membranes: case study on the synergism
between linear cationic antimicrobial peptides, Biophysical Journal 106 (2) (2014), Suppl. 1, 441a.

[52] E. S. Salnikov, H. Sarrouj, C. Reiter, C. Aisenbrey, A. Purea, F. Aussenac, O. Ouari, P. Tordo,


I. Fedotenko, F. Engelke, B. Bechinger, Solid-state NMR/Dynamic Nuclear Polarization of
polypeptides in planar supported lipid bilayers, J. Phys. Chem. B 119 (2015) 14574-14583.

[53] E. Markhasin, J. Hu, Y. Su, J. Herzfeld, R.G. Griffin, Efficient, balanced, transmission line rf
circuits by backpropagation of common impedance nodes, J. Magn. Reson. 231 (2013) 32-38.

[54] J.R. Sirigiri, T. Maly, L. Tarricone, A compact 395 GHz gyrotron for dynamic nuclear polarization,
IRMMW-THz-2011 – 36th International Conference on Infrared, Millimeter, and Terahertz Waves,
6104951.

[55] T. Maly, J.R. Sirigiri, Simplified THz instrumentation for high-field DNP-NMR spectroscopy, Appl.
Magn. Reson. 43 (2012) 181-194.

[56] P. Allen, F. Creuzet, H.J.M. de Groot, R.G. Griffin, Apparatus for low-temperature magic-angle
spinning NMR, J. Magn. Reson. 92, 614-617, 1991.

[57] W. H. Potter, Apparatus to rotate samples rapidly at temperatures less than 2 K in high
transverse magnetic fields, Rev. Sci. Instrum. 42 (1971) 618-625.

22
[58] A. Hackman, H. Seidel, R.D. Kendrick, P.C. Myhre, C. Yannoni, Magic-angle spinning NMR at near-
liquid helium temperatures, J. Magn. Reson. 79 (1988) 148.

[59] P.C. Myhre, G.G. Webb, C. Yannoni, Magic angle spinning nuclear magnetic resonance near
liquid-helium temperatures. Variable-temperature CPMAS spectra of the 2-norbornyl cation to 6 K,
J.Am. Chem.Soc. 112 (1990) 8991-8992.

[60] A. Samoson, T. Tuherm, J. Past, A. Reinhold, T. Anupold, I. Heinmaa, New horizons for magic-
angle spinning NMR, Topics Curr. Chem. 246 (2005) 15-31.

[61] M. Concistre, O.G. Johannessen, E. Carignani, M. Geppi, M.H. Levitt, Magic-angle spinning NMR
of cold samples, Acc. Chem. Research 46 (2013) 1914-1922.

[62] F.D. Doty, G.S. Hosford, J.B. Spitzmesser, J.R. Bittner, An ultra-compact laminar-flow cryogenic
heat exchanger, Adv. Cryo. Eng. 37 A (1992) 233-240.

[63] Y. Matsuki, K. Ueda, T. Idehara, R. Ikeda, I. Ogawa, S. Nakamura, M. Toda, T. Anai, T. Fujiwara,
Helium-cooling and spinning dynamic nuclear polarization for sensitivity-enhanced solid-state NMR
at 14T and 30K, J. Magn. Reson. 225 (2012) 1-9.

[64] Y. Matsuki, T. Idehara; Y. Tatematsu, J. Sirigiri, S. Nakamura; T. Fujiwara, Closed-Cycle Helium-


Cooling MAS NMR Probe System for Dynamic Nuclear Polarization at 16.4 T, in: Experimental
Nuclear Magnetic Resonance Conference, Asilomar 2015.

[65] E. Bouleau, P. Saint-Bonnet, F. Mentink-Vigier, H. Takahashi. J.-F. Jacquot, M. Bardet,


F. Aussenac, A. Purea, F. Engelke, S. Hediger, D. Lee, G. de Paepe, Pushing sensitivity limits using
dynamic nuclear polarization with closed-loop cryogenic helium spinning, Chem. Sci. 6 (2015) 6806-
6812.

[66] E. de Rijk, Terahertz passive components for dynamic nuclear polarization – nuclear magnetic
resonance applications, PhD thesis, Ecole Polytechnique Federale de Lausanne 2013.

[67] A. Macor, E. de Rijk, G. Annino, J.-Ph. Ansermet, THz-waves channeling in a monolithic saddle
coil for dynamic nuclear polarization enhanced NMR, J. Magn. Reson. 212 (2011) 440-449.

[68] J-H. Ardenkjaer-Larsen, G.S. Boebinger, A. Comment, S. Duckett, A. S. Edison, F. Engelke,


C. Griesinger, C. Hilty, H. Maeda, G. Parigi, T. Prisner, E. Ravera, J. van Bentum, S. Vega, A. Webb,
C. Luchinat, H. Schwalbe, L. Frydman, Facing and overcoming sensitivity challenges in biomolecular
NMR, Angewandte Chemie Int’l Edition 54 (2015) 2-26.

[69] E.A. Nanni, A.B. Barnes, Y. Matsuki, P.P. Woskov, B. Corzilius, R.G. Griffin, R.J. Temkin,
Microwave field distribution in a magic angle spinning dynamic nuclear polarization NMR probe,
J. Magn. Reson. 210 (2011) 16-23.

[70] D.J. Kubicki, A.J. Rossini, A. Purea, A. Zagdoun, O.Ouari, P. Tordo, F. Engelke, A. Lesage,
L. Emsley, Amplifying dynamic nuclear polarization of frozen solutions by incorporating dielectric
particles, J. Am. Chem. Soc. 136 (2014), 15711-15718.

[71] Songi Han, Surface characterization by solid and solution DNP enhanced NMR, in : 5th
International DNP Symposium Egmomd aan Zee (NL), 2015.

[72] News headline on Revolution NMR, LLC website, https://www.revolutionnmr.com

23
[73] G.J. Gerfen, L.R. Becerra, D.A. Hall, R.G. Griffin, R.J. Temkin D.J. Singel, High Frequency (140 GHz)
dynamic nuclear polarization: Polarization transfer to a solute in frozen aqueous solution, J. Chem.
Phys. 102 (1995) 9494-9497.

[74] D.A. Hall, D.C. Mauss, G.J. Gerfen, S.J. Inati, L.R. Becerra, F.W. Dahlquist, R.G. Griffin,
Polarization-Enhanced NMR Spectroscopy of Biomolecules in Frozen Solution, Science 276 (1997)
930-932.

[75] M. Rosay, A.-C. Zeri, N.S. Astrof, S.J. Opella, J. Herztfeld, R.G. Griffin, Sensivity-Enhanced NMR of
Biological Solids: Dynamic Nuclear Polarization of Y21M fd Bacteriophage and Purple Membrane,
J. Am. Chem. Soc. 123 (2001) 1010-1011.

[76] K. N. Hu, H. H. Yu, T. M. Swager, R. G. Griffin, Dynamic Nuclear Polarization with Biradicals
J. Am. Chem. Soc. 126 (2004) 10844-10845.

[77] K.-N. Hu, C. Song, H.-h. Yu, T. M. Swager and R. G. Griffin, High frequency dynamic nuclear
polarization using biradicals: multi-frequency EPR lineshape analysis, J. Chem. Phys. 128 (2008)
052321.

[78] Y. Matsuki, T. Maly, O. Ouari, H. Karoui, F. Le Moigne, E. Rizzato, S. Lyubenova, J. Herzfeld, T.


Prisner, P. Tordo, R.G. Griffin, Dynamic Nuclear Polarization using a Rigid Biradical, Angew. Chem.
Int. Ed. Engl. 48 (2009) 4996-5000.

[79] C. Sauvée, M. Rosay, G. Casano, F. Aussenac, R.T. Weber, O. Ouari, P. Tordo, Highly efficient,
water-soluble polarizing agents for dynamic nuclear polarization at high frequency, Angew Chem Int
Ed. Engl. 52 (2013) 10858-10861.

[80] A.N. Smith, M.A. Caporini, G.E. Fanucci, J.R. Long, A method for dynamic nuclear polarization
enhancement of membrane proteins, Angew. Chem. Int. Ed. Engl. 24 (2015) 1542-1546.

[81] C. Fernández-de-Alba C, H. Takahashi, A. Richard, Y. Chenavier, L. Dubois, V. Maurel, D. Lee, S.


Hediger, G. De Paëpe, Matrix-free DNP-enhanced NMR spectroscopy of liposomes using a lipid-
anchored biradical. Chemistry 21 (2015) 4512-4517.

[82] B. Corzilius, L.B. Andreas, A.A. Smith, Q.Z. Ni, R.G. Griffin, Paramagnet induced signal quenching
in MAS-DNP experiments in frozen homogeneous solutions, J. Magn Reson. 240 (2014) 113-123.

[83] K.R. Thurber and R. Tycko, Perturbation of nuclear spin polarizations in solid state NMR of
nitroxide-doped samples by magic-angle spinning without microwaves, J. Chem. Phys. 140 (2014)
184201.

[84] F. Mentink-Vigier, S. Paul, D. Lee, A. Feintuch, A. Hediger, S. Vega, G. DePaëpe, Nuclear


depolarization and absolute sensitivity in magic-angle spinning cross effect dynamic nuclear
polarization, Phys. Chem. Chem. Phys. 17 (2015) 1824-1836.

[85] A. Lesage, M. Lelli, D. Gajan, M.A. Caporini, V. Vitzthum, P. Mieville, J. Alauzun, A. Roussey, C.
Thieuleux, A. Mehdi, G. Bodenhausen, C. Coperet, L. Emsley, Surface enhanced NMR spectroscopy by
dynamic nuclear polarization, J Am Chem Soc. 132 (2010) 15459-15461.

[86] A. Zagdoun,A.J. Rossini, D. Gajan, A. Bourdolle, O. Ouari, M. Rosay, W.E. Maas, P. Tordo, M. Lelli,
L. Emsley, A. Lesage, C. Coperet (2012). Non aqueous solvents for DNP surface enhanced NMR
spectroscopy, Chem Commun (Camb) 48 (2012) 654-656.
24
[87] A. Zagdoun, G. Casano, O. Ouari, M. Schwarzwälder, A.J. Rossini, F. Aussenac, M. Yulikov, G.
Jeschke, C. Copéret, A. Lesage, P. Tordo, L. Emsley, Higher MW/longer T1e polarizing agents,
J. Am. Chem. Soc. 135 (2013) 12790-12797.

[88] D. Mance, P. Gast, M. Huber, M. Baldus, K.L. Ivanov, The magnetic field dependence of cross-
effect dynamic nuclear polarization under magnet angle spinning, J. Chem. Phys., 142 (2015) 234201.

[89] T.V. Can, M.A. Caporini, F. Mentink-Vigier, B. Corzilius, J.J. Walish, M. Rosay, W.E. Maas, M.
Baldus, S. Vega, T.M. Swager, R.G. Griffin, Overhauser effects in insulating solids, J. Chem. Phys. 141
(2015) 064202.

[90] G. Mathies, M.A. Caporini, V.K. Michaelis, Y. Liu, K.N. Hu, D. Mance, J.L. Zweier, M. Rosay, M.
Baldus, R.G. Griffin, Efficient dynamic nuclear polarization at 800 MHz/527 GHz with trityl-nitroxide
biradicals, Angew. Chem. Int. Ed. 54 (2015) 11770-11774.

25

You might also like