+ All Categories
Home > Documents > Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1...

Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1...

Date post: 08-Jun-2021
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
21
Hyperpolarized 3 He and perfluorocarbon gas diffusion MRI of lungs Mark S. Conradi a,b, * , Brian T. Saam c , Dmitriy A. Yablonskiy a,b , Jason C. Woods a,b a Department of Physics, Washington University, St Louis, MO 63130, USA b Department of Radiology, Washington University, St Louis, MO 63110, USA c Department of Physics, University of Utah, Salt Lake City, UT 84112, USA Received 7 October 2005 Available online 10 March 2006 Keywords: Diffusion; Imaging; Hyperpolarized; 3 He; Lungs Contents 1. Introduction .................................................................................... 63 1.1. Hyperpolarization of 3 He ..................................................................... 64 1.2. Using HP 3 He in MRI experiments .............................................................. 65 1.3. Rapid imaging of ventilation ................................................................... 65 1.4. Low-field imaging .......................................................................... 66 1.5. Determination of regional oxygen partial pressure (pO 2 ) in the lung ...................................... 66 2. 3 He restricted diffusion measurements ................................................................. 66 2.1. Diffusion ................................................................................. 66 2.2. Method .................................................................................. 67 2.3. Results ................................................................................... 68 3. Fundamental measurement of 3 He gas anisotropic diffusion in human lung and evaluation of lung microstructure .......... 70 3.1. Theory of 3 He gas anisotropic diffusion ........................................................... 70 3.2. Relationship between 3 He gas ADC and airway size ................................................. 71 3.3. Methods .................................................................................. 72 3.4. Preliminary findings ......................................................................... 73 4. Magnetization tagging and long-range diffusivity of 3 He .................................................... 74 4.1. Lung structure ............................................................................. 75 4.2. Diffusion at long distances in lung ............................................................... 75 4.3. Spatial modulation of longitudinal magnetization .................................................... 75 4.4. Long-range diffusion in lungs .................................................................. 76 4.5. In vivo Imaging of D sec in dogs with unilateral emphysema ............................................ 77 4.6. Imaging D sec in explanted human lungs with emphysema .............................................. 77 4.7. In vivo Imaging of D sec in humans .............................................................. 77 5. Fluorine-19 ADC measurements ..................................................................... 78 5.1. Rationale ................................................................................. 78 5.2. Methods .................................................................................. 79 5.3. Results ................................................................................... 79 5.4. Transition to human imaging ................................................................... 80 Acknowledgements ............................................................................... 80 References ..................................................................................... 80 1. Introduction A remarkable development at the interface of physics and biomedical science over the past 10 years has been the use of hyperpolarized (HP) noble gases to perform MRI of the lung Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–83 www.elsevier.com/locate/pnmrs 0079-6565/$ - see front matter q 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.pnmrs.2005.12.001 * Corresponding author. Address: Department of Physics, Washington University, St Louis, MO 63130, USA. Tel.: C1 314 935 6418. E-mail address: [email protected] (M.S. Conradi).
Transcript
Page 1: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

Hyperpolarized 3He and perfluorocarbon gas diffusion MRI of lungs

Mark S. Conradi a,b,*, Brian T. Saam c, Dmitriy A. Yablonskiy a,b, Jason C. Woods a,b

a Department of Physics, Washington University, St Louis, MO 63130, USAb Department of Radiology, Washington University, St Louis, MO 63110, USAc Department of Physics, University of Utah, Salt Lake City, UT 84112, USA

Received 7 October 2005

Available online 10 March 2006

Keywords: Diffusion; Imaging; Hyperpolarized; 3He; Lungs

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

1.1. Hyperpolarization of 3He . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

1.2. Using HP 3He in MRI experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

1.3. Rapid imaging of ventilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

1.4. Low-field imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

1.5. Determination of regional oxygen partial pressure (pO2) in the lung . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

2. 3He restricted diffusion measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

2.1. Diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

2.2. Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

2.3. Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

3. Fundamental measurement of 3He gas anisotropic diffusion in human lung and evaluation of lung microstructure . . . . . . . . . . 70

3.1. Theory of 3He gas anisotropic diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

3.2. Relationship between 3He gas ADC and airway size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

3.3. Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

3.4. Preliminary findings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

4. Magnetization tagging and long-range diffusivity of 3He . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

4.1. Lung structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

4.2. Diffusion at long distances in lung . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

4.3. Spatial modulation of longitudinal magnetization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

4.4. Long-range diffusion in lungs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

4.5. In vivo Imaging of Dsec in dogs with unilateral emphysema . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

4.6. Imaging Dsec in explanted human lungs with emphysema . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

4.7. In vivo Imaging of Dsec in humans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

5. Fluorine-19 ADC measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

5.1. Rationale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

5.2. Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

5.3. Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

5.4. Transition to human imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

0079-6565/$ - see front matter q 2006 Elsevier B.V. All rights reserved.

doi:10.1016/j.pnmrs.2005.12.001

* Corresponding author. Address: Department of Physics, Washington

University, St Louis, MO 63130, USA. Tel.: C1 314 935 6418.

E-mail address: [email protected] (M.S. Conradi).

1. Introduction

A remarkable development at the interface of physics and

biomedical science over the past 10 years has been the use of

hyperpolarized (HP) noble gases to perform MRI of the lung

Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–83

www.elsevier.com/locate/pnmrs

Page 2: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–8364

air space. Such imaging is made possible through laser-optical

pumping, which can improve the magnetic resonance

sensitivity of certain noble-gas isotopes having non-zero

nuclear spin by several orders of magnitude. The two most

important isotopes are the spin one-half nuclei 3He and 129Xe,

because of their long intrinsic T1 relaxation times. We recall

that the conventional thermal polarization Pth in an applied

magnetic field B0 is given by PthZmB0/kBT, where m is the

nuclear magnetic moment, kB is the Boltzmann constant, and T

is the absolute temperature in Kelvins. For fully relaxed

protons in a conventional 1.5 T applied field for whole-body

MRI, we have PthZ6!10K6, which suffices for magnetic

resonance (MR) imaging of protons in water. Gases delivered

at pressures of z1 atm are over 2000 times less dense than

protons in water with a corresponding decrease in MR

sensitivity, but hyperpolarization yields an enhancement of 4

to 5 orders of magnitude (PhypZ10–50%), more than

compensating for the density difference. Indeed, highly

polarized 3He can remain sufficiently sensitive to MRI even

if it constitutes only a fraction of the inhaled gas mixture. Once

introduced in vivo to the lungs, depending on the breathing

maneuver and the inhaled mixture, HP gases relax with a time

T1Z20–30 s, where the relaxation is dominated by interaction

with paramagnetic oxygen. In the glass cell used for

hyperpolarization, T1 values are much longer, ranging from

20 to 40 h.

Because the function of the lung is gas exchange, it is hardly

surprising that regionally specific information about inspired

gas should be at least as relevant to the study of lung

physiology and disease as images of the lung-tissue structure

(e.g. hydrogen MRI or X-ray CT). Results from several

research groups have indeed demonstrated the potential for

HP-gas MRI to enhance our understanding of lung function,

with further potential impact on disease treatment, surgical

planning, and drug development. (Lung diseases such as

chronic obstructive pulmonary disease (COPD) and asthma

affect tens of millions of people in the US alone [1]). Almost

immediately from the time that the first animal- and then

human-lung images were demonstrated [2–5], several groups

have explored pulse-sequence techniques and contrast mech-

anisms that go well beyond static spin-density imaging and that

make use of the unique physical properties of these gases to

address physiologically relevant questions.

This article deals largely with one such contrast mechanism:

diffusion (i.e. Brownian motion) of gas in the lung. Rapid gas

diffusion can cause problems ranging from limited image

resolution to signal attenuation due to the presence of bulk

susceptibility and/or imaging gradients. However, as with

relaxation due to the presence of oxygen (see below), there is a

flip side to the story that enables meaningful physiological

information to be obtained. Indeed, an early conclusion in HP3He MRI, based on the fact that diffusion does not limit signal

intensity or resolution as much as might be expected, was that

airway and alveolar boundaries restrict the diffusive motion of3He in healthy lung [6,7]. The apparent diffusion coefficient

(ADC) of the gas can thus be a powerful indicator of local

airway and alveolar architecture, which is itself often

profoundly affected by lung disorders, such as COPD [8].

Most of the imaging discussed here involves HP 3He, as it

has been most widely used to date in studies of lung ADC.

However, we also consider in this regard other gases that may

be used as MRI signal sources. Other noble-gas isotopes

(principally 129Xe) can be hyperpolarized and have been used

for lung imaging [9]. Although 129Xe has only 1/3 the magnetic

moment of 3He and has generally been more technically

challenging to polarize in large quantities, it provides unique

contrast in that it is taken up by the blood from the lung and

exhibits a wide chemical shift range when dissolved in a

variety of tissues [10]. The excitement generated by HP-gas

MRI has also contributed to a renewed interest in the use of

inert fluorinated gases to image the lungs [11,12]; these gases

have a number of features that somewhat mitigate the

disadvantage of a small (thermally generated) nuclear

polarization, and they are technically easier to handle and

use, particularly for in vivo studies.

Before turning exclusively to diffusion MRI, we will first

briefly discuss how hyperpolarized gas is generated and

imaged. We will also highlight (though by no means

exhaustively), some of the developments in the last 10 years

across the field of lung imaging with HP gases. More complete

reviews of the field may be found in Refs. [13,14]. The

subsequent sections on diffusion MRI are based largely on

research done by the group at Washington University, where it

has been a major effort for the past 8 years. We will start with

how simple measurements of 3He ADC can be used to gauge

the regional severity of tissue destruction due to emphysema.

We then discuss a more detailed model of airway architecture

that can be tested by investigating 3He diffusion anisotropy.

The fact that gases diffuse so rapidly affords the opportunity to

explore lung structure on many different length scales, and the

next section discusses experiments to measure long-range 3He

diffusion in the lung. Finally, we present some recent work

involving 19F diffusion MRI.

1.1. Hyperpolarization of 3He

3He is a stable non-radioactive isotope of helium with

nuclear spin 1/2 and a gyromagnetic ratio about 25% smaller

than the proton; it has z1 ppm natural abundance but is

available in pure form as the result of collection from tritium

decay. The nucleus can be polarized to values approaching

unity by transfer of angular momentum from circularly

polarized laser light. There are two basic schemes for this

transfer: metastability-exchange optical pumping (MEOP)

[15,16] and spin-exchange optical pumping (SEOP) [17]. In

both schemes, the 3He gas typically resides in a glass vessel

(cell) through which the laser light is directed. In MEOP, a

radio-frequency discharge is ignited in the cell to create a

population of 3He atoms in the triplet-2S metastable electron

state. Optical pumping with 1083 nm laser light, corresponding

to transitions from the triplet-2S to certain triplet-2P states,

leads to a large electron polarization in the metastable state,

which is immediately transferred to the nucleus via hyperfine

Page 3: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–83 65

coupling. Rapid electron-exchange collisions then produce

nuclear polarized 3He atoms in the ground state. In SEOP, an

alkali metal (usually rubidium) serves as an intermediary in the

angular momentum-transfer process. Laser light is absorbed by

the Rb vapor at a wavelength of 795 nm, corresponding to the

first principal dipole transition (5S1/2K5P1/2), causing the Rb

valence electron to become highly polarized. A density of Rb

vapor appropriate to the amount of available laser light is

produced in the cell by heating it to 160–200 8C. The angular

momentum is then transferred to the 3He nucleus via a

hyperfine interaction (zero-quantum transition) that takes place

during Rb–3He binary collisions. In both methods, the

mechanism for polarization transfer is easily turned off (in

MEOP by terminating the discharge and in SEOP by cooling

down the cell and condensing the Rb vapor), leaving the

angular momentum stored in the form of polarized 3He nuclei.

The longitudinal relaxation time of the gas (usually dominated

by collisions with the cell walls) can, with some attention given

to how the cell is fabricated, be many 10 s of hours, allowing

sufficient time for storage and transport of the hyperpolarized

gas to an MRI scanner. We note that the SEOP method can also

be used to hyperpolarize the other stable and abundant spin-1/2

noble-gas isotope, 129Xe.

Generally speaking, MEOP has historically had the

advantage of an intrinsically faster production rate and largest

maximum 3He polarization, routinely producing quantitiesw1

STP-liter/hour at O60% polarization [18]. However, the RF

discharge requires that the gas be polarized at pressures

w1 Torr and then compressed to atmospheric pressures without

causing significant polarization loss [19]. In addition, the lasers

used in MEOP have been somewhat more expensive and more

difficult to maintain. Initial large-scale polarization systems

have been large and expensive, but much more portable and

inexpensive MEOP systems have been recently demonstrated

[20]. By contrast, SEOP systems have the advantage of cheap

powerful portable lasers and the ability to operate at helium

pressures between 1 and 10 atmbut a slower intrinsic production

rate; a typical system requiresmany hours to produce 1STP-liter

of HP 3He [21]. SEOP is a photon-limited process, and the

introduction in the 1990s of compact inexpensive diode-laser

arrays, capable of several tens of watts at 795 nm, was a crucial

technological advance that allowed the production of quantities

of HP 3He sufficient for human-lung MRI. Recently, the

simultaneous use of both rubidium and potassium metals in a

SEOP cell has been shown to improve the HP 3He production

rate by an order of magnitude [22]. Combined with ever more

powerful lasers, which can nowbe frequency-narrowed to better

match the alkali-metal absorption profile [23], this so-called

‘hybrid’ SEOP has reduced the gap in intrinsic production

capacity relative to schemes employing MEOP.

1.2. Using HP 3He in MRI experiments

The major technical difference in the use of HP gases in

MRI, as compared to conventional thermal signal sources, is

that the magnetization is independent of the applied magnetic

field and does not recover thermally. Rather, it continually

decays towards a negligibly small thermal-equilibrium value.

Even before being administered to a subject, HP gas must be

handled carefully in order to avoid catastrophic relaxation due

to interactions with foreign surfaces, molecular oxygen [24],

magnetic-field gradients [25], and ambient fluctuating mag-

netic fields. The gas is usually kept in a glass cell that is known

to result in a wall-relaxation time (10 s of hours) that is

sufficiently long. The cell is positioned in a solenoid or similar

homogeneous alignment field, preferably with external

electromagnetic shielding. The gas is typically released from

the cell to a flexible plastic enclosure from which the subject

inhales just prior to imaging. Subjects typically breathe in an

anoxic mixture containing HP 3He, perhaps with a subsequent

inhalation of room air, in order to maximally avoid oxygen-

induced relaxation [24].

An inhaled bolus of HP 3He possesses all of its useable

magnetization from the outset, and this magnetization must be

rationed appropriately over the number of images acquired and

according to the k-space-traversal scheme employed. There is

also a strong desire to use as much of the hard-won

magnetization as efficiently as possible. Since there is no

thermal recovery, imaging speed is limited only by gradient-

switching speed and available receiver bandwidth. Spin echo

sequences are generally avoided, because use of a large number

of inexact 1808 rf pulses would eventually tip most of the non-

renewable magnetization into the transverse plane where it

would dephase rapidly. The use of gradient-recalled echoes

(GRE) combined with low-flip-angle excitation (e.g. the

FLASH sequence [26]) was the canonical choice for early

work in this field. The disadvantage of using GRE sequences is

that the transverse magnetization is limited by T�2 . Although T�

2

is longer for 3He in the gas space than for protons in lung tissue

[27], it is still limited to about 20 ms at 1.5 T by the magnetic

susceptibility effects of the large air–tissue interface in the lung

[28,29]. In addition, rapid diffusion of 3He through the imaging

gradients themselves can also contribute to transverse-signal

attenuation [30].

1.3. Rapid imaging of ventilation

Much of the early progress in lung imaging with HP 3He

came in the form of improved resolution in static images of

ventilation at breath-hold. A two-dimensional FLASH

sequence was typically employed for human imaging, with

in-plane resolution of a few millimeter and total acquisition

times of about 1 s [31]. As spectacular as these early images

were, it could be argued that much of the same information

might be obtained with other, less technically demanding

methods, e.g. the use of inert fluorinated gases or oxygen-

enhanced MRI of lung parenchyma. However, the large single-

shot signal intensity and the absence of any recovery-time

limitation afforded by HP gas make it uniquely suitable for

imaging the dynamics of gas flow. Imaging ventilation in real

time during the breathing cycle requires more efficient use of

the transverse magnetization through short effective echo times

and/or multiple-echo sequences. Johnson and co-workers

extensively developed the use of radial k-space acquisition

Page 4: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–8366

(RA) and projection reconstruction with HP 3He in small

animals [32,33]. Although they consume somewhat larger

amounts of magnetization (not as much of a problem with

small animals as with humans), RA sequences have very short

effective echo-times, which minimize susceptibility and

diffusion losses, and they oversample near kZ0, which tends

to minimize motion artifacts (e.g., from heartbeat). Rapid

repetitive imaging of the human lung [30] was first

demonstrated using echo-planar imaging (EPI) [34], whereby

the transverse magnetization from one excitation is repeatedly

refocused by sinusoidal read-out gradients to acquire succes-

sive lines in k-space. Saam et al. [30] showed that in the low-

flip-angle limit, an increase in linear voxel dimension leads to a

sixth-power increase in the number of images that can be

acquired with one bolus of inhaled HP gas. The larger voxel

size also significantly reduces the signal attenuation due to

diffusion through the correspondingly smaller imaging

gradients. Using a coarse-grid (32!64) two-dimensional EPI

sequence, they were able to acquire an image from a single rf

excitation every 40 ms. These images showed gas filling the

lung, the gravitational dependence of ventilation, and some

washout characteristics. There were some artifacts, mostly due

to susceptibility effects around the pulmonary vessels. The use

of segmented EPI, where only a limited number of echoes are

acquired per excitation, can reduce artifacts at the cost of using

more magnetization. In spiral-acquisition scanning [35,36], as

with RA, the effective echo time is short; each view starts at the

center of k-space but spirals outward to the edge of the plane.

Since a given spiral provides a more uniform sampling of

k-space than a single radial acquisition, the entire image can be

updated after each acquisition by combining it with some

number of previous interleaved acquisitions (k-space window-

ing technique). Salerno et al. [36] generated images of

inspiration with an effective temporal resolution of 15 ms

and produced spectacular cine loops of inflowing inspired gas

in both healthy subjects and patients with cystic fibrosis.

1.4. Low-field imaging

Because the magnetization in HP gases is independent of the

applied field, one expects the signal-to-noise ratio (S/N) to be

approximately independent of applied field in the regime where

noise from the sample dominates coil noise (true for chest MRI

down to fields of 0.1 T and even lower) [37]. For lung imaging,

low fields should be particularly advantageous since they

reduce the magnetic susceptibility gradients in the lung [38].

Durand and co-workers have imaged human lungs at 100 mT

[39], taking advantage of very long transverse coherence times

to use a single-shot multiple-spin-echo pulse sequence (RARE)

[40], for which acquisition of the multiple CPMG echoes is

limited by (the much longer) T2 instead of T�2 . Mair and co-

workers have recently demonstrated human lung images in an

open-access very low-field (!5 mT) system [41]. These

systems are generally not commercially engineered, and

there are some difficulties unique to low-field imaging,

including low-frequency noise and undesirable gradients

concomitant to the imaging gradients that can cause image

artifacts [42,43]. The overall image quality is improving but

has yet to rival what has been achieved at high fields.

1.5. Determination of regional oxygen partial pressure (pO2)

in the lung

The presence of molecular oxygen is generally considered a

limitation for HP-gas MRI, because it causes rapid relaxation

of the gas and is, in fact, the dominant relaxation mechanism in

the lung in vivo. However, Deninger et al. [44] have

demonstrated the use of this relaxation mechanism to quantify

regional oxygen partial pressure (pO2) in the lung, by

measuring regional differences in relaxation rate of inspired3He gas. Later adaptations of this technique [45] have been

used to assess in porcine lung the regional ventilation-

perfusion ratio (VA/Q), an important physiological parameter

for characterizing lung disease [46].

2. 3He restricted diffusion measurements

2.1. Diffusion

Any discussion of 3He MR measurements of restricted

diffusivity in lungs necessarily involves the disease pulmonary

emphysema [8]. In emphysema, the average size of the

compartments which restrict the gas motion is enlarged and

the restricting barriers (alveolar walls) become more porous,

leading to an increase in the apparent diffusion coefficient

(ADC). In the opinion of our group, 3He ADC measurements

for mapping the extent and severity of emphysema in human

lungs (in vivo) is the application of hyperpolarized gas imaging

which is most likely to find widespread clinical application. It

appears that this opinion is shared by a majority of the field

[47].

In healthy human lungs, there are approximately 300

million alveoli, the smallest subdivisions of the air space

[48]. Each alveolus is an open vessel of approximately 300 mmdiameter (0.3 mm). The alveolarized space accounts for about

93% of the total gas volume. In emphysema, the compartments

are enlarged by destruction of elastin, the protein responsible

for the elastic tension (spring return) of the lung. Destruction of

alveolar walls also results in fewer compartments. Fig. 1

presents microscope images of healthy and emphysematous

lung tissue at the same scale (after excision, freezing, sampling,

and slicing of thin sections), demonstrating the changes in lung

microstructure in emphysema. So, at the broadest level,

emphysematous lungs have fewer and larger compartments;

therefore one expects (and finds) the ADC of 3He gas to

increase in emphysema, being less restricted [49]. A more

sophisticated and anatomically correct picture of lung structure

is presented in the next section. It is important to remark that3He ADC is sensitive to and reports upon features of sizes

much smaller than the image resolution. This aspect of

diffusion or q-space imaging has been noted previously [50].

The free diffusivity D0 of3He gas dilute in N2 or air has been

measured to be 0.88 cm2/s [51]. Typically, a bolus of

approximately 0.45 l STP 3He is mixed with sufficient N2

Page 5: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

Fig. 1. Microscope images at 5! of tissues fixed at inflation from healthy (H) and emphysematous (E) lungs. The larger characteristic dimensions in the

emphysematous tissue are evident and result in less restriction to gas diffusion.

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–83 67

(0.55–1.55 l) for a breathhold; thus, in a lung of 6 l total

capacity, the 3He concentration is about 7%, close to the

infinite dilution limit for D0. The gradient pulses employed are

typically tZ2 ms in duration (we note that these are long

gradient pulses, so the time interval across which the

displacement is measured is not well defined) [50]. In this

time interval, the root mean square free displacement in one

direction isffiffiffiffiffiffiffiffiffiffi2D0t

por 600 mm (0.6 mm). This is much larger

than the alveolar size, indicating that the gas atoms will do a

thorough exploration of the airway and nearby alveoli during

the diffusion measurements. Thus 3He diffusive motion in

healthy lungs is expected to be highly restricted: that is, the

ADC will be a small fraction of D0. For a heavier, larger gas

species such as xenon or C2F6 or C3F8 (see Section 5), the

smaller D0 makes the situation somewhat different (less

restricted).

2.2. Method

The MRI measurement of 3He ADC uses a gradient-echo

version of the famous Stejskal-Tanner pulsed field gradient

sequence [52]. Because the hyperpolarized 3He spin magne-

tization is not renewable, rf p-pulses are to be avoided, as

discussed above. The pulse sequence is shown in Fig. 2. The

bipolar diffusion sensitizing gradient (BDSG) is omitted to

make an unweighted image and is present to make the diffusion

weighted image. The timings of the two (with and without

Fig. 2. Pulse sequence for measuring 3He ADC using a gradient echo with two

values of b (one is bZ0). Here, b is the diffusion-sensitizing gradient weight.

The bipolar diffusion sensitizing gradient pulse (BDSG) is detailed at right. The

time durations used here are tZ500 ms, dZ1800 ms, and DZ1800 ms. At left,

only the bs0 acquisition is shown; the bs0 and bZ0 acquisitions (with

BDSG pulse at zero amplitude) are taken alternately.

BDSG) are held the same, so that the only difference between

the two resulting images is diffusion weighting. Typical time

durations of the features of the BDSG are given in Fig. 2.

For a fixed quantity of hyperpolarized gas, the imaging time

and S/N of multi-slice two-dimensional and three-dimensional

acquisitions are predicted to be the same (for n partitions in

three-dimensional, each spin is subjected to n times as many rf

pulses as for n-slice two-dimensional acquisitions, so the

nutation angle must be 1=ffiffiffin

pas large, in the small-angle limit;

the smaller S/N of each echo is exactly compensated by the

signal averaging effect of n times as many echoes). We have

chosen multi-slice two-dimensional acquisitions to minimize

motion effects, as each slice is acquired in about 0.7 s, for 32

phase encodes and two values of the BDSG.

The BDSG occupies a time interval in the sequence during

which the imaging gradients are fully rewound (slice select) or

not yet applied (phase encode and readout). Likewise, the

BDSG is fully rewound to kZ0 or not yet applied when the

imaging gradients are applied. Thus, there are no cross terms

between the BDSG and the imaging gradients and analysis of

the data is simplified.

The two images, with (W) and without (WO) the BDSG, are

generated simultaneously as much as possible, to avoid

artifacts from bulk motion, either patient motion or incomplete

breathhold. This means acquiring the same phase encode in the

same slice, with and without, in immediate succession (about

11 ms apart). So the order of operations from largest to smallest

is slice, phase encode, with and without BDSG.

In our earliest work, we worried about the influence of bulk

motion, which should be approximated well as constant

velocity over the duration (2–4 ms) of the BDSG [49]. To

eliminate attenuation from constant velocity motion of the

spins, a gradient pulse of structure CKKC was used. For

constant velocity motion (but not stochastic or diffusive

motion), the phase shift generated by the first half of the

gradient waveform (CK) is cancelled by the phase shift from

the second half waveform (KC). Compared to a BDSG of the

simplerCK structure with the same duration of each lobe, the

CKKC pulse has twice the overall duration and only twice

the b-value (see below). Compared to a CK pulse of the

doubled duration which would have an 8! larger value of b,

the (CKKC) pulse is inefficient. We abandoned use of the

velocity compensated pulse (CKKC) to allow data

Page 6: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

Fig. 3. Coronal 3He ventilation (gray) and restricted diffusion (color) images of

a 22 kg Yorkshire pig. The slice thickness is 10 mm. The gas diffusion in the

bulk of this healthy lung is much smaller than the unrestricted diffusion evident

in the trachea (largest airway).

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–8368

acquisition with shorter gradient echo times (TE) of 7.2 ms,

because there was no evidence of artifacts with the simpler

CK pulse (see below). We note that gas diffusion is about 104

larger than for water, so b-values for gases need be only 10K4

as large, making measurement of gaseous diffusion compara-

tively easy.

The mean squared displacement hDx2i for motion at constant

velocity V is V2t2 while for purely diffusive motion hDx2iZ2Dt.

Thus, at sufficiently short times, the diffusive displacements are

always larger than those from constant velocity motion. For

example, in a typical time of 2 ms with a restricted diffusivity

of 0.20 cm2/s as measured in healthy human lungs [49], the

root mean square displacement is about 0.3 mm. This

corresponds to a large constant velocity of 150 mm/s. At

breathhold, velocities in the lung and chest cavity are likely to

be a very small fraction of this. The calculation suggests that

compensation for constant velocity motions is not required. We

further note that constant velocity motion of all the spins in a

voxel, as in patient chest motion, would give rise to an overall

phase shift from the BDSG but no attenuation. However,

incomplete breathhold would cause a non-uniform flow

velocity and therefore would result in signal attenuation.

Analysis of the diffusion data is simple. The only difference

between the two images is diffusive attenuation, so for each

voxel

Iw=Iwo Z eKbD (1)

so that

D Z ð1=bÞlnðIwo=IwÞ: (2)

Here the value of b, the weight of the diffusion-sensitizing

gradient, is bZg2G2d2Df, where the times d and D are defined

in Fig. 2, f is a dimensionless number determined by the BDSG

pulse shape, and G is the amplitude of the BDSG. Often there

are regions of lung that ventilate poorly and receive too little3He so that the S/N is inadequate. To eliminate such voxels

from the diffusion map, a lower intensity limit of 2.5 times the

average noise level is used on both raw images, with (W) and

without (WO) the BDSG. Voxels falling below this level are

presented as gray; for all other voxels, diffusion is presented on

a color scale.

Fig. 4. Transverse-slice 3He images of a healthy volunteer (H) and a patient

with severe emphysema (E). The gray-scale images show the distribution of

inspired 3He at breathhold and the color images are maps of restricted

diffusivity. In the emphysematous lungs, the gas distribution is less uniform and

the apparent diffusion coefficient (ADC) is much larger.

2.3. Results

Fig. 3, left, presents a coronal slice ventilation or spin-

density image of a 22 kg Yorkshire pig together with the

apparent diffusivity on the right. The animal was deeply

anesthesized so that an external ventilator was used until the

bolus of 3He and N2 was administered manually; a complete

breathhold was affected by closing the valve in series with the

tracheal insert tube. The diffusion map demonstrates that the

gas diffusivity is highly restricted (D/D0%0.2) everywhere

except in the large airways. In the trachea, D is essentially

equal to D0 (0.88 cm2/s), with some smaller airways evident as

less enhanced D through the partial volume effect (airway is

thinner than image slice). The uniformity of D in the rest of the

lung shows the excellent S/N here.

Ventilation (left) and restricted diffusivity (right) images

from a healthy volunteer (H) and a patient with severe

emphysema (E) are presented in Fig. 4. Like most of our human

images, these are transverse (axial) slices to permit easy

comparison with X-ray CT images, which are presently

considered the gold standard radiological technique for

characterizing emphysema. The ventilation or spin density

image of the emphysema patient is considerably less uniform in

intensity than for the healthy lungs. Some of the intensity

variation in the image of the healthy volunteer is due to a

spatial dependence of the sensitivity of the homebuilt rf coils.

The diffusivity maps do not suffer from the rf inhomogeneity,

because they use a ratiometric method as in Eq. (2). The images

clearly show the restriction of diffusion in the healthy lungs,

with D/D0 of about 0.22. In the emphysematous lungs, the

diffusivity is substantially increased and is spatially dependent.

The spatial variations in D follow the pattern one expects by

comparison to X-ray CT (not shown), with regions of low

Page 7: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

Fig. 6. 3He images of ventilation and ADC in two emphysema patients, top and

bottom. As discussed in the text, poor ventilation does not always accompany

elevated ADC, and well-ventilated regions can display unexpectedly high

ADC. The color scale is the same as in Fig. 4.

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–83 69

X-ray attenuation (and hence low tissue density) having high3He diffusion. Comprehensive quantitative comparison of CT

and 3He diffusion, performed by identifying the same region of

lung in the two image sets, finds a strong correlation between

these quantities [53].

A histogram of diffusivity in 15 healthy volunteers and 33

patients with severe emphysema is presented in Fig. 5. The

peak value of D in healthy lungs is near 0.20 cm2/s and is

surprisingly narrow given the number of subjects. The peak

from emphysematous subjects is greatly elevated and much

broader, presumably reflecting a range of disease severity. The

inescapable conclusion is that 3He diffusivity cleanly separates

healthy and severely emphysematous lungs. An open question

is how early emphysema can be detected and over how short of

a time span its progress can be measured using 3He ADC; this

last question is particularly relevant to testing the efficacy of

new drugs that target emphysema.

Spin-density and diffusion images from two patients with

severe emphysema appear in Fig. 6. The images demonstrate

that spin-density or ventilation images and diffusivity maps

report different aspects of structure and function of the lungs.

That is, they deliver independent information. In general, one

expects that the most diseased lung regions will have the

poorest ventilation (because of air trapping) and the highest3He diffusivity. Indeed, this is true in many cases. In

emphysema patients with a history of cigarette smoking (but

not alpha-one disease [8]), we commonly find the ventilation

decreases and ADC increases from bottom to top (inferior to

superior) of the lung, the usual direction of increasing extent of

disease. But the upper two panels of Fig. 6 show a counter-

example. There in the right lung (at left in the figure), a region

of poor ventilation nevertheless shows only moderately

elevated ADC (yellow). In another counter-example at bottom

of Fig. 6, there is a well-ventilated region of right lung (at lower

left in the figure) that has very high ADC (blue). Thus, the

distribution of gas at breathhold and the ADC report on

Fig. 5. Histogram of restricted diffusivity ADC in healthy (H) and

emphysematous (E) lungs, taken from data on normal volunteers and patients

with severe emphysema. The vertical scale is proportional to the number of

imaging voxels at each ADC. The (H) and (E) distributions are well separated.

different aspects of the lung; the two kinds of images report

distinct information.

Fig. 6 is a good example of non-uniform ventilation,

showing that most severely emphysematous lungs contain

regions that ventilate so poorly that they receive inadequate3He for the diffusion measurement (i.e. inadequate S/N). To

some extent, having the patient rebreathe the 3He–N2 gas

mixture in and out of a flexible vessel (plastic bag) reduces this

effect. But strict limits should be placed on the total time

breathing an anoxic mixture for patients with severe disease.

With some patients, the distribution of inhaled 3He is so non-

uniform that the images ‘do not look like lungs’.

Other groups and our own have examined the dependence of

the measured diffusivity on several factors, including age

within a pool of healthy subjects, the level of lung inflation, the

direction of the BDS gradient, and the time duration of the BDS

gradient pulse [54]. Studies indicate an excellent reproduci-

bility of 3He diffusion results for repeated measurements on a

subject [55,56]. There appears to be little or no dependence on

diffusion direction (global anisotropy, quite distinct from the

microscopic anisotropy discussed in Section 3) [57]. The

diffusivity in healthy subjects decreases slowly with increasing

diffusion time [54], from 0.22 to 0.15 cm2/s for diffusion times

of 1.8 to 5.8 ms (note that much longer diffusion times of order

seconds give much smaller diffusion coefficients, as discussed

in Section 4). While we have not carefully studied the

dependence of D upon the level of inflation, it appears that D

increases with lung volume (total volume of gas in lung), but

more slowly than we expected, with the mean diffusivity

increasing about 25% between a tidal inhalation (functional

residual capacity plus 0.7 l) to a full inhalation (total lung

volume), an approximate doubling of the lung volume.

Page 8: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

Fig. 7. Schematic structure of two levels of respiratory airways. Open spheres

represent alveoli forming an alveolar sleeve around each airway. Each

respiratory airway can be considered geometrically as a cylindrical object

consisting of a tube embedded in the alveolar sleeve. The diagram defines inner

(r) and outer (R) radii (as in Fig. 1 in Ref. [65]). Inset schematically represents

the structure of the same airways in emphysema, scaled down for clarity.

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–8370

It is not yet clear what spatial resolution is required to

obtain a clear picture of the extent and severity of

emphysema in each patient. Our diffusion maps generally

show trends from top to bottom and/or variations between

lobes, suggesting that the fairly large voxels we use in vivo

(7!7!10 mm3) are still much smaller than the relevant

length scales of the variation of disease severity. While one

might think to always obtain images at a better resolution

(smaller voxels) than needed and then simply average

voxels to obtain fewer, larger cells for better S/N, this

procedure (i) results in inferior S/N and (ii) requires extra

imaging and breathhold time, compared to simply imaging

at the desired, coarser resolution in the first place. The

wasting of imaging time is obvious, with extra slices and

extra phase encodes to acquire. The S/N argument can be

derived from general equations [37] of the signal-to-noise in

MR imaging, but we present here a quick version. We

compare the S/N of the entire imaging field obtained by (a)

imaging with M!N!P voxels (either in three-dimensional

or multi-slice two-dimensional, because they give equal S/N

here as remarked earlier) or (b) a non-imaging measurement

(i.e. a single-voxel experiment). We recall that the

integrated intensity of all voxels in an entire image is

given by the amplitude of the kZ0 datum; all other

k-values do not enter into the sum across all voxels. Thus,

the imaging approach (a) must use MNP rf pulses so each

rf pulse is of nutation angle a, of order 1=ffiffiffiffiffiffiffiffiffiffiffiMNP

pradians, to

consume most of the 3He magnetization throughout the

many rf pulses. The non-imaging measurement (b) can use

a single rf pulse of 908, or approximately aZ1 radian.

Thus, the kZ0 datum of the imaging approach has a

smaller S/N by 1=ffiffiffiffiffiffiffiffiffiffiffiMNP

p; hence, for obtaining only the kZ0

point which gives the spatial integral across the entire

sample, the non-imaging approach is superior. The same

argument can be generalized to show that higher S/N is

obtained when the imaging is performed at the desired

resolution, as opposed to imaging at a higher-than-needed

resolution and averaging voxels afterwards. Thus, it is

important to decide on an optimum spatial resolution, based

on the characteristics of the disease.

Besides in vivo measurements on healthy volunteers

and patients with emphysema, we have obtained images

of restricted diffusion in canine lungs [58,59] (using

elastase-treated dogs) and from emphysematous human

lungs excised after lung transplant surgery [60]. The dogs

were imaged after each of three elastase treatments; after

the final treatment and 3He imaging, the dogs were

sacrificed and the lungs were removed, frozen, and sliced

thin for observation under the microscope. The surface

area to volume ratio S/V and 3He diffusivity D showed

the expected correlation, with S/V decreasing and D

increasing as emphysema progressed [58]. Similar

measurements on explanted human emphysematous lungs

have been made but not yet published [60].

Our group has also performed 3He diffusion imaging of

live mice [61]. Techniques for delivering 3He with each

breath are used, to obtain adequate S/N. These methods

were developed [62] by Hedlund and Johnson for rats and

showed enhanced 3He diffusion in elastase-treated rats [63].

Likewise, the results of Dugas et al. [61] on mice show

enhanced diffusivity in elastase-treated mice, but did not

show elevation of D in mice that had been exposed over

months to cigarette smoke. We note however, that no

verification of pulmonary emphysema by other techniques

was obtained for these smoking mice.

3. Fundamental measurement of 3He gas anisotropic

diffusion in human lung and evaluation of lungmicrostructure

Here, we present a mathematical model relating the

measurements of 3He gas diffusivity and lung microstructure

and report in vivo measurements of airway structure at the sub-

acinar level in human lung in healthy subjects and patients with

emphysema.

3.1. Theory of 3He gas anisotropic diffusion

In the branching tree model [64], the hierarchy of the

airway tree begins at the trachea and leads through bronchi

and bronchioles to the terminal bronchiole that feeds each

acinus—the major gas exchange unit in the lung. In humans

there are fourteen generations of airways prior to the

terminal bronchioles and another nine inside the acini [64].

Gas ventilation in the trachea, bronchi, bronchioles and

terminal bronchioles occurs by convection (bulk flow),

while diffusion is the primary ventilation mechanism beyond

the terminal bronchioles—in the acini, where about 93% of

gas resides [48]. According to Haefeli-Bleuer and Weibel

[65] essentially all airways in the acinus are decorated by

alveoli forming an alveolar sleeve. Thus the structures we

focus upon here are cylindrical airways covered by alveolar

sleeves, as schematically represented in Fig. 7. In humans,

intra-acinar airways branch dichotomously over about nine

generations, and the internal airway radius r falls from 250

Page 9: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–83 71

to 135 mm, whereas the outer radius R (including the sleeve

of alveoli) remains constant at 350 mm [65].

The characteristic free diffusion length l0Zffiffiffiffiffiffiffiffi2D0

pt

(0.56 mm for tZ1.8 ms, as applies here) is much larger

than the average alveolar radius of 0.15 mm; 3He atoms can

diffuse out of alveoli and across the airways in the 1.8 ms

time duration of the MR ADC measurement. During the

same time, most 3He atoms will remain inside the same

airway. Thus, in our model of gas diffusion in lung, we

consider airways rather than alveoli as the elementary

geometrical units. We approximate the airways as long

cylinders-either smooth (trachea, bronchi and broncheoli) or

covered with alveolar sleeves (respiratory broncheoli,

alveolar ducts and alveolar sacs). The alveolar walls as

well as the walls of alveolar ducts and other branches of the

airway tree serve as obstacles to the path of diffusing 3He

atoms and reduce the 3He diffusivity. Crucially, these

restrictions are substantially less along the airway axis than

perpendicular to it. Because gas motion along the axis of an

airway is less restricted than perpendicular to the axis,

diffusion in the lung is anisotropic. We show that this

anisotropy manifests itself in the MRI signal even though

each imaging voxel contains a very large number of

differently-oriented airways that cannot be resolved by

direct imaging. In particular, the anisotropy of diffusion

results in non-exponential MR signal decay as a function of

the weight b of the diffusion-sensitizing gradients [51] (see

just below Eq. (2)), allowing the diffusion rates along and

across the airways to be separately determined.

If the diffusion-sensitizing gradient is applied along or

perpendicular to the tube axis, the signal attenuation can be

written in the form of Eq. (1) with DZDL or DT, the

longitudinal or transverse ADC, respectively. For the more

general case of an airway with principal axis tilted from the

field gradient direction by angle q, the ADC can be presented as

ADCðqÞZDL cos2qCDT sin2q: (3)

With the spatial resolution of several millimeters currently

available with 3He MRI, each voxel contains hundreds of

airways with different orientations. For each individual airway

with orientation q, the signal attenuation is exponential with

respect to b, according to Eq. (1). Because of the ADC

dependence on orientation angle q in Eq. (3), after summing the

signals over all airways, the signal decay becomes non-mono-

exponential. This problem is mathematically similar to the

problem of water diffusion in randomly-oriented uniaxial

layers [50]. Because of the large number of acinar airways in

each imaging voxel, their orientation distribution function

g(q)Zsin q/2 can be taken as uniform. Therefore, the signal S

can be written as

S Z S0

ðp0

dqsin q

2expKbðDL cos2qCDT sin2qÞ� �

Z S0 expðKbDTÞp

4bDAN

� �1=2

F½ðbDANÞ1=2

� �(4)

where F(x) is the error function and we have introduced the

anisotropy of ADC

DANZDLKDT: (5)

Eq. (4) assumes that all airways are similar; i.e. all

airways have the same geometrical parameters and,

consequently, the same values of DT and DL. Ideally, the

expression for signal S should be further averaged with

respect to the different geometrical parameters of the

airways. To keep the number of parameters small in the

model, we assume that the diffusivities DT and DL already

represent averaged values.

Eq. (4) describes the non-mono-exponential dependence

of diffusion attenuated signal on the value of b. Hence,

the apparent diffusion coefficient, ADC, defined as ADCZKln(S/S0)/b, is a function of b. It can be easily demonstrated that

ADCZ�D Z

1

3DL C

2

3DT; bDL; bDT/1;

�D ZDT; bDL[1:

8><>: (6)

Apparently, in healthy lungs, DT/DL. The ADC at large b

values decreases to the value of DT because the signal from

airways oriented with a component along the gradient gives a

much smaller contribution as compared to the signal from

airways oriented perpendicular to the direction of the diffusion

sensitizing gradient.

3.2. Relationship between 3He gas adc and airway size

Because alveoli are open polygons with openings

nearly the same size as their diameters (Fig. 7), the gas

diffusion perpendicular to the acinar airway direction can

be approximated by the transverse diffusion in open

smooth tubes. The characteristic radius R here is the

major or larger radius of the airway. For the apparent

diffusion coefficient in the transverse direction with

respect to the tube (airway) axes and for the gradient

pulse waveform and timing of Fig. 2, we [51] found the

following theoretical expression:

DT Z16D0x

4Rh

2

wðh; 3Þ

Xj

bK41j

ðb21jK1ÞQ

b21j

2hx2R;h;3

!: (7)

Here we have introduced dimensionless parameters xR, h,

and 3 for the tube radius and gradient pulse timing (see

Fig. 2)

xR ZR

l0; hZ

D

d; 3Z

t

d; (8)

and the function w is defined as

wðh;3ÞZ hK1

3C3 1K2hCh3K

7

63C

8

1532

� �: (9)

Here, b1j is the jth (non-zero) root of the equation

J 01ðxÞZ0, where J 0

1 is the first derivative of the first order

Page 10: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

Fig. 8. Transverse 3He diffusivity DT as a function of the airway external radius

R. Solid black line represents results obtained according to Eq. (7), and dots

represent computer Monte Carlo simulations. D0 is the 3He free diffusion

coefficient in air, and l0Z(2D0D)1/2 is the characteristic diffusion length. For

the free diffusion coefficient of infinitely diluted 3He in N2 or air, D0Z0.88 cm2/s, and the gradient waveform parameters DZdZ1.8 and tZ0.5 ms

described in Fig. 2, the characteristic free displacement is equal to l0Z0.56 mm.

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–8372

Bessel function and the function Q(a, h, 3) is given by

Qða;h;3ÞZ 1K4

33K

2

a332½expðKa3ÞCa3K1�

C4

a332sin h2 a3

2

� ! expðKað1K3ÞÞK2 sin h2 að1K3Þ

2

� �expðKahÞ

� �;

(10)

where aZD0b21jd=R

2.

The limit of strongly restricted diffusion corresponds to the

case when the free atom root mean square displacement l0during the time D is much larger than R, i.e. xR/1. In this

limit, Q(a,h,3)z(1K43/3), the sums in Eq. (7) can be

calculated exactly, and the transverse ADC turns out to be

inversely proportional to D0 (motional narrowing)

DTyD0

7 1K433

�h2

12wðh;3Þx4Re R4

D0

; xR/1: (11)

In the case xR[1, the free atom displacement during the time

D is much smaller than R, and therefore only the atoms residing

in the cylindrical shell within the distance l0 from the

cylindrical surface ‘sense’ the boundaries. Hence, in this

limit, DT can be approximated as

DTyD0 1Kc1ðh;3Þ

xR

� �; (12)

where c1(h,3) is a numerical coefficient depending on the

waveform parameters. For the waveform parameters given in

Fig. 2, c1z0.65. Naturally, when xR/N, the value of DT

approaches that for the free-diffusion coefficient D0, as expected.

To derive the relationship in Eq. (7) between transverse

diffusivity and tube radius for given parameters of gradient

waveform, we used a theoretical approach that relies on the

assumption of a Gaussian distribution of the phases accumu-

lated by the precessing spins [66,67]. This approach is based on

the assumption of a random character of the phase accumulated

by the nuclei and, in fact, is identical to the random walk

approximation used in the pioneering papers [68,69]. Since,

then it has been applied to describing NMR signal behavior

under the constant or pulsed field gradient in some restricted

geometries (between two parallel barriers, in a cylinder and in a

sphere) in numerous papers (see, for example [50] and

references therein). To test the applicability of the Gaussian

approximation in our case, we have verified our theoretical

result against computer Monte-Carlo random walk simulations.

The ratio DT/D0 versus the reduced radius xR, calculated by

means of Eq. (7), is plotted in Fig. 8 along with the results of

computer simulations for the present gradient waveform

parameters listed in Fig. 2. This figure shows that for xR!0.5

(R!l0/2Z0.28 mm), DT, the transverse ADC, is very small—

less than 5% of the free diffusion coefficient D0. Qualitatively,

the very small value of DT results from a motional averaging

effect during each half (G) of the gradient pulse. In the

physiologically most relevant interval l0/2!R!2l0 (0.28 mm

!R!1.1 mm), the transverse diffusivity DT grows sharply to

about 65% of D0. With further increase in the tube radius R, the

value of DT increases very slowly and approaches the 3He free

diffusion coefficient D0 as 1/R. This means that for all airways

with radius less than 1.1 mm, 3He diffusivity is substantially

restricted (DT!0.65D0).

Themain feature restricting diffusion along the acinar airways

leading to a reduction in the value of DL, the longitudinal ADC,

compared to the free diffusion coefficient D0 is the presence of

alveolar sleeves as in Fig. 7. Longitudinal diffusion of particles

(3He atoms) located in the open area of the tube can be considered

as unrestricted, whereas diffusion of particles within the external

cylinder are significantly restricted due to the alveolar structure.

In fact, from the point of view of longitudinal diffusion, the

alveolar sleeves play the role of ‘traps’, effectively reducing the

longitudinal diffusion coefficient.

If the trapped atoms in the alveolar sleeves cannot exchange

(or can exchange only slowly) with the ‘free’ atoms in the

center of the tube, the average longitudinal diffusion DL

becomes equal to D0 (r/R)2, based on averaging over the

numbers of atoms in each region and their long-time limit

diffusivities (0 and D0). So, with exchange occurring, one has

D0ðr=RÞ2!DL!D0: (13)

Simulations of longitudinal diffusion confirm this expression. For

realistic dimensions, the simulations indicate that DL is about

mid-way between the limits expressed in Eq. (13).

3.3. Methods

Eq. (4) is the basis for separation of longitudinal and

transverse ADC values. Indeed, by collecting a series of MR

images with different b and fitting Eq. (4) to the data on a pixel-

Page 11: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

D DL DT R

N1

P1

P2

Fig. 9. Single-slice maps of diffusivities for a normal subject (N1) and two

patients with severe emphysema (P1 and P2). From left to right the columns

display the orientationally-averaged diffusivity �D, the longitudinal ADC value

DL, the transverse ADC value DT, and the mean airway radius R. The color

scale on the right represents diffusivity coefficients in centimeter2/second and

airway radii in millimeter. Each color corresponds to 0.05 unit. Brown arrows

point to an area of emphysematous lung with minor airway destruction, pink

arrows to an area with moderate airway destruction, and green arrows to a lung

area with severe emphysema. The small high-diffusivity regions in N1 are the

two bronchi below their branching from the trachea.

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–83 73

by-pixel basis, we can create maps of the values of the

transverse and anisotropic ADC, DT and DAN. The mean �D and

longitudinal DL are then obtained from Eqs. (6) and (5), and the

mean airway radius R is obtained from Eq. (7). We use an

imaging technique for 3He diffusion measurement with six

two-dimensional gradient echo sequences as in Fig. 2,

combined together so that each has its own identical small-

angle RF pulse, and its own identical slice selection, phase

encode, and read-out gradients. Data are collected in an

interleaved manner by collecting the same line in k-space for

all six images prior to stepping to the next line, ensuring

reduced sensitivity to motion artifacts. The standard central

reordering of phase encoding is used to reduce the possible

influence of signal decay during acquisition. All six sequences

except the first one include diffusion-sensitizing gradients with

increasing amplitudes, Gm, and consisting of one bipolar pulse

pair as in Fig. 2. The corresponding values of b are 0, 1.5, 3,

4.5, 6 and 7.5 s/cm2. The diffusion gradient is applied

perpendicular to the long axis of the body. For a typical

experiment, we use a slice thickness of 20 mm and an in-plane

resolution of 7!7 mm (225!450 mm field of view with 32!64 k-space samples). Each of the 32 lines in k-space uses an RF

excitation of about 78, allowing for repeated acquisition from

the same hyperpolarized spins. The gradient echo time in all

sequences is TEZ7.2 ms.

All images were acquired with a 1.5 T whole body Siemens

Magnetom Vision scanner. Homemade double-tuned RF

Helmholtz coils were used to transmit and receive the MRI

signal at the 1H and 3He resonance frequencies. Switching of the

operating frequency was performed without moving the subject,

allowing for better registry and comparison between 3He and 1H

scout images. After the RF coils were placed above and below the

chest, the subject was positioned supine in theMRmagnet. First,

scout images were obtained using conventional proton MRI.

These proton images were used to select the slices and

orientations for the 3He images, and for anatomic reference to

the 3He images. Then, the hyperpolarized 3He gas mixed with

nitrogen was delivered to the subject through a plastic ventilator

tubing connected to a mouthpiece. Imaging was performed

during breathhold at the end of full inhalation. Four slices were

obtained from each subject in less then 10 s, except for one

normal volunteer with only three slices because of inadequate

S/N. Diffusivity maps were obtained by fitting Eq. (4) to the

experimental data on a pixel-by-pixel basis using Bayesian

probability theorywith uninformative prior probabilities. Normal

volunteers and patients with severe emphysema and selected for

lung volume reduction surgery were studied.

3.4. Preliminary findings

In all of the subjects, both normal and emphysema patients,

gas diffusivity is anisotropic with the mean longitudinal ADC

being usually two to three times as large as the mean transverse

ADC. Representative maps of ADCs and the mean radii of

acinar airways are shown in Fig. 9 for one normal subject and

two patients with severe emphysema.

The following points are evident from our results. In healthy

subjects, the transverse diffusivity is strongly restricted, with

the mean value of DT almost eight times smaller than the 3He

free diffusion coefficient in air (D0Z0.88 cm2/s). The maps

defining DT and the resulting external radii R of the acinar

airways (including the alveolar sleeves as depicted in Fig. 7)

are highly homogeneous. The mean R is about 0.36–0.37 mm

in normal subjects. Given that our in vivo measurements were

made during full inhalation, this result is in remarkable

agreement with measurements on normal, excised lungs of

mean RZ0.35 mm [65]. In these subjects, the mean values of

DL, the longitudinal ADC, are less than half of the 3He free

diffusion coefficient. This is mainly the result of the restrictions

to diffusion imposed by the walls separating neighboring

alveoli along the same airway, as depicted in Fig. 7. In the

healthy subjects, both DT and DL decrease slightly from apices

to base; similar variation is present in some patients. As our

subjects are supine, gravity effects do not explain this variation.

In patients with severe emphysema, nearly all transverse

ADC maps show increased DT as compared to normal subjects,

but the diffusion is still restricted (DT!D0). The increase in DT

is consistent with an increase in the mean airway radius R. The

value of DL is also substantially elevated, becoming practically

unrestricted in some parts of the lungs (DLyD0). This effect is

consistent with the limits in Eq. (13) discussed above and with

an inflation of the airways which results in r approaching the

value of R (see inset of Fig. 7).

The relationship Eq. (7) depicted in Fig. 8 between

transverse ADC and airway radius R was derived under the

assumption that the diffusing 3He atoms cannot penetrate

through the alveolar walls. However, alveolar walls always

have pores. In normal lung the number of pores (known as

pores of Kohn) is very small and they are generally smaller

than 10 mm in diameter [70]; hence, their effect on DL and DT

and Eq. (7) is negligible. However, in emphysematous lung

many more pores (known as fenestrae) of variable sizes occur

Page 12: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–8374

in alveolar walls [71]. The destruction of alveolar walls in

emphysema will contribute to the above-discussed increases in

the longitudinal and transverse ADC values. Thus, in the light

of tissue destruction in emphysema, Eq. (7) should be

considered as an approximation and the calculated values of

R should be regarded as an apparent radius of the airways.

The orientationally averaged ADC values �D from Eq. (6) are

similar to two-b-value ADC measurements for healthy subjects

and patients with severe emphysema [49,72,73]. However, the

previous work employed only two-b values and effectively

assumed exponential signal decay as a function of b through

the diffusion-sensitizing gradient strength. As is evident from

the present results and analysis, the ADC in lungs determined

from the two-b method depends on b-value and approaches the

true orientation-average value �D from below only in the limit

of vanishing b. That is, the true �D is the slope of the curve (ln of

signal as a function of b) at bZ0, while the two-b-value of

ADC is the slope taken between two points. The �D values

reported herein provide a more objective result because they do

not depend on the diffusion-sensitizing gradient strength.

The above analysis allows us to draw some general

conclusions about gas diffusivity in normal and emphysema-

tous lungs and to develop some notions about emphysema

progression. Our patient pool has been selected for lung

volume reduction surgery and therefore has very hetero-

geneous presentations of the disease. As a result, most of the

DT and DL maps in the patients are highly inhomogeneous,

showing regions of nearly normal as well as very abnormal

lung tissue. This provides an opportunity to follow the

dynamics of lung destruction through the progression of

emphysema by examining the variations within each patient, as

well as between patients. For example, the brown arrows in

Fig. 9 point to a quasi-normal area of lung in Patient 1. Here,

transverse diffusivity is only 30% increased as compared to a

normal lung (see similar area in the image above). This

corresponds to a mean airway external radius of RZ0.42 mm,

according to Eq. (7) and Fig. 8. However, the longitudinal

diffusivity in this area of the lung is increased by about 60%.

Pink arrows point to an area of lung that has an intermediate

level of emphysema (Patient 2)—here DT is increased by

almost 100% (corresponding to RZ0.52 mm) while DL is

elevated by about 80%, nearly equal to the unrestricted value.

Up to this stage we can envision that emphysema progresses by

expansion of airways without substantial destruction of

alveolar walls; this picture is entirely consistent with recent

findings [74]. Green arrows point to a highly emphysematous

lung region in Patient 2 where the transverse ADC value is

0.62 cm2/s—more than four times as large as in normal lung

while the longitudinal ADC is only 90% elevated. This

indicates that according to Eq. (7) and Fig. 8 the mean external

airway radius has become larger than 1 mm. The large increase

in apparent radius R and the concomitant decrease in

anisotropy (DL/DT) are consistent with substantial tissue

destruction in this region. These data indicate that the value

of DL is the more sensitive parameter for identifying early

stages of emphysema, while DT is the more sensitive parameter

for identifying lung tissue destruction as emphysema pro-

gresses (because DL cannot increase beyond D0).

In summary, in this section we have described a further

refinement to the technique of 3He diffusion MRI that

incorporates a model for the lung airway architecture and

allows the measurement of ADC anisotropy. Analysis of the

non-exponential signal decay on a pixel-by-pixel basis yields

separate values for the ADC along and across the acinar

airways, despite the fact that individual airways are too small to

be resolved directly. A mathematical model links the

transverse ADC and the mean airway radius R. Our in vivo

measurements of R in normal lungs are in excellent agreement

with previous ex vivo results. The results demonstrate

substantial differences between healthy and emphysematous

lung at the acinar level and may provide new insights into

emphysema progression.

4. Magnetization tagging and long-range diffusivity of 3He

The methods of Sections 2 and 3 follow the diffusive decay

of transverse spin magnetization over times of a few

milliseconds, corresponding to diffusion distances of a few

tenths of a millimeter. While these methods are quite

successful, the diffusion times are limited by T�2 , making it

impossible to probe the lung structure at longer distances with

transverse magnetization. An important reason to study

diffusion at longer length scales is to understand the

interconnections of the airways in health and disease. In

particular, collateral ventilation pathways (direct connections

between generationally distant parts of the lung tree) can best

be assessed by long-distance diffusion measurements.

There are two strategies for studying diffusion across longer

distances: (1) reduce the external field Bo to a point where T�2 is

much longer, since the decay time of transverse magnetization,

T�2 , is predicted to be inversely proportional to Bo (or even to

B20 in the motional averaging regime), or (2) tag the

longitudinal magnetization, which decays with a relatively

long time constant T1 (approximately 25 s for 3He in healthy

lungs in vivo). The problems associated with low-field imaging

were discussed in Section 1; preliminary images have been of

more modest signal-to-noise than expected [41]. Here, we

focus on tagging the longitudinal magnetization of 3He as a

method for measuring long-range diffusivity. Because of the

bifurcating nature of airways in lung (Fig. 10), the measured

diffusivity is dependent on the length-scale of the measurement

and thus on the time-scale. The long-range diffusivity may be

altered differently with different types of lung-tissue destruc-

tion and is a significantly different measure of lung structure [8]

than the short-range diffusivity [49,75]. The ADC measured

over times of seconds and distances of centimeters is denoted

Dsec to distinguish it from the ADC over milliseconds and

fractions of a millimeter, Dmsec, discussed in the two sections

above. We emphasize that the different time and distance scales

make these ADC values different, not the use of transverse

versus longitudinal spin magnetization. Further, for Dmsec

measurements, the gradient produces modulation of the

magnetization with a (minimum) wavelength smaller than the

Page 13: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

Fig. 10. Pictoral representation of lung, showing the approximate 24 levels of

bifurcating airways and illustrating the conducting and diffusive (respiratory)

zones for gas transport. The diffusive zone contains about 93% of the gas

volume.

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–83 75

linear dimension of a voxel, while for striping measurements of

Dsec the wavelength is 6–12 times the voxel dimension in the

read-out direction.

4.1. Lung structure

Healthy lungs are composed of approximately 24 levels of

bifurcating airways (Fig. 10); these are singly connected,

meaning that only one path along the airways exists between

any two points in the lung [48]. Some collateral, bypass

channels exist in human lungs via the pores of Kohn and

channels of Lambert, but these pathways have high resistance

and are not considered to be important in normal respiration

[76,77]. The conducting airways (approximately the first 15

levels) are essential but only represent about 7% of the total gas

volume. The last 9–10 levels are short in length and comprise a

region of diffusive transport where oxygen and carbon dioxide

exchange with blood through the walls of the alveoli; these

regions are the acini [48]. (We note that the ‘front’ between

convective and diffusive transport can change slightly between

inspiration and expiration and is altered somewhat because of

motion of the adjacent, beating heart [78].) The mean linear

size of an acinar unit is about 7 mm, with individual acinar

airways having lengths of approximately 1 mm [65].

4.2. Diffusion at long distances in lung

The branching nature of airways in lung, in particular that

arbitrary points are singly connected by unique airway paths,

necessitates that atoms must travel from one acinar airway to

the next for diffusion distances greater than 1 mm; over

centimeter distances, the change in direction at each airway

branch results in a tortuous path. Diffusion between arbitrary

points one or more centimeters apart requires that atoms must

travel from one acinus to another, connecting via a common

conducting airway node at level 14 (or lower-numbered) in the

tree. Thus to travel between alveolar sacs at the 24th level in

different acini, atoms must travel up at least 10 levels of

branching airways and then back down [48]. The apparent

diffusion measured over such distances is thus much more

restricted and occurs over times of several seconds. We and

others have determined that this diffusion coefficient Dsec is

about 0.02 cm2/s for explanted healthy human lungs and

in vivo dog lungs [75,79,95].

In emphysematous lungs, tissue destruction is known to

create progressively more alternate (collateral) routes for gas

motion between arbitrary points [80]. Diffusion over centi-

meter distances thus can proceed through the airways and the

collateral paths in parallel, so that Dsec is increased by the

collateral paths. The existence of these collateral pathways is

the basis for a new idea for minimal surgery that relieves

dyspnea by relieving trapped gas in patients with severe

emphysema [81]. Characterization of the collateral pathways

by long-range gas diffusion measurements would potentially be

very useful in planning for such a procedure. Magnetization

tagging with hyperpolarized gas is the only currently available

method capable of imaging diffusive gas motion and collateral

pathways at these distances, and is due to the long relaxation

time constant T1.

4.3. Spatial modulation of longitudinal magnetization

Spatial modulation of the longitudinal spin magnetization

Mz was originally used to track cardiac or thoracic motion by

establishing a coordinate grid locked into the tissue [82,83].

The primary advantage of this type of imaging is that encoding

or labeling of position is in the longitudinal magnetization,

which is characterized by the longest relaxation time constant

in the spin system, T1. Magnetization tagging of gases uses the

same imaging techniques as those used for cardiac tagging, but

here the motion of tagged magnetization in lungs at breathhold

is diffusive. The random (diffusive) motion of the 3He gas

atoms results in attenuation of the spatial modulation, allowing

measurements of diffusion over times limited only by T1,

approximately 25 s in vivo in human lungs and many minutes

ex vivo where O2 is excluded [75,79].

Sinusoidal spatial modulation of the longitudinal magneti-

zation, as opposed to a more nearly square-wave modulation,

offers advantages in that attenuation of sinusoidal contrast can

be easily interpreted in terms of a single diffusion length and

because it can be easily prepared with two rf pulses separated

by a gradient pulse. The result of this preparation is

sinusoidally modulated magnetization with wavelength l

(Figs. 11 and 12). For a gradient pulse of amplitude G and

effective duration t, the value of l is given by lZ2p/gGt, with

g being the spin magnetogyric ratio. If the two rf pulses are

Page 14: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

Fig. 11. Idealized cross-section of a tagged image at increasing waiting times t,

with t0!t1!t2. In (a), diffusion alone is represented, without influence of T1

processes or rf consumption of magnetization by imaging pulses. The spatial-

average magnetization is unchanged, but the extent of modulation is attenuated

by the diffusion. In (b) the influence of T1 and the rf imaging pulses is depicted,

with no diffusion. Here the shape of the cross-section is unchanged, because the

magnetization is everywhere reduced by the same factor.

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–8376

each of flip-angle q with the same rf phase, the fraction of

longitudinal magnetization Mz remaining at position x after the

second rf pulse is obtained from Bloch’s equations:

Mz Z cos2qKsin2q cosðgGxtÞ: (14)

The case with two 45 rf pulses gives

Mz Z ½1KcosðkxÞ�=2 (15)

with khgGtZ2p/l. This delivers 100% modulation of the

magnetization while avoiding the ambiguity of sign reversal

that occurs with gradient echo (magnitude) imaging. After the

magnetization tagging pulses, the spatially modulated magne-

tization is homogenized by diffusion according to the diffusion

equation [84]

vMz=vt ZDsecV2Mz: (16)

Here Dsec represents the effective diffusivity of the gas and T1

relaxation has been ignored for simplicity;P2 is the Laplacian

operator, v2/vx2 Cv2/vy2 Cv2/vz2. The relevant solution to the

above equation is given by

Mzðx;tÞZACBðtÞcosðkxÞZACBð0ÞcosðkxÞeKRt; (17)

Fig. 12. Images of time-evolving, sinusoidally modulated magnetization in canine l

(Dsec); note the different scales for each. Each grayscale image advances 1.36 s. In

lavage to induce emphysema. The healthy lung (at left) has a Dsec of 0.015 cm2/s (12

less restriction, as apparent in both the diffusion map and the rapid disappearance

where the decay rate constant R is determined by

RZDseck2Z4p2Dsec=l

2. Thus, the long-range diffusivity

Dsec can be calculated from measurements of R, since l is

known. Both the decay rate R and resulting Dsec are

independent of the initial modulation depth B(0), so exact

calibration of the rf tagging pulses is not essential. The effect of

both T1 relaxation and the consumption of longitudinal

magnetization Mz by subsequent rf imaging pulses (used to

inspect the decaying modulated magnetization and ignored in

Eqs. (16) and (17)) is to drive Mz uniformly toward zero (recall

the equilibrium Mz is virtually zero compared to hyperpol-

arized values), independent of position x as in Fig. 11. We

define the fractional sinusoidal modulation FM(x), as B(t)/A

from Eq. (17); this ratio is unaffected by T1 and rf pulses,

allowing long-range diffusion to be measured without

corruption from these effects. This is valid provided that T1

and rf pulses are uniform over the length scale l. The rf field

amplitude B1 is a slowly-varying function of position in the rf

coil, so B1 inhomogeneity is not expected to distort the

decaying sinusoidal magnetization over the relevant scale of l

(2–3 cm in our measurements).

The details of the data analysis of the successive images

showing the decay of tagged magnetization as in Figs. 11 and

12 has been published [75]. Essentially, the relative stripe

amplitude FM is determined at each voxel using a Fourier

integral over a single wavelength. The values of FM are fitted

to exp(KRt) separately for each voxel.

4.4. Long-range diffusion in lungs

The first magnetization tagging measurements with hyper-

polarized 3He in lung were reported by Owers-Bradley, et al.

[79] The tagging was also used to image respiratory motion,

and a global, lung-average diffusion was measured from the

decay of tagged magnetization across the entire lungs, without

imaging [83]. Their method calculated Dsec from decay rates at

several different tagging wavelengths. They found RZACDk2

with a constant D (which we call Dsec) for values of l from 1 to

4.5 cm. The constant decay rate A accounts for T1 relaxation.

Their result of DsecZ0.02 cm2/s in a healthy human subject in

this range of wavelengths is one tenth of typical values of

restricted diffusivity measured over milliseconds (denoted

Dmsec here). This dramatic decrease in diffusivity over larger

distances is consistent with the idea that gas atoms must travel

from one acinus to another through the maze of airways via a

common node on the airway tree. We show below in several

ungs, with accompanying color diffusion maps at short- (Dmsec) and long-times

these animals one lung (at right in the images) has been treated with elastase

! smaller than Dmsec in the same lung), and the emphysematous lung has much

of the modulation in the grayscale images.

Page 15: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

Fig. 13. Decay of sinusoidal modulation (grayscale) and resultant long-range

diffusion map (color) in a nominally healthy, excised donor lung. Even after

6 s, there is little attenuation of the modulation; Dsec is measured to be

0.017 cm2/s nearly uniformly.

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–83 77

studies of healthy lung that Dsec is typically about a factor of 10

smaller than Dmsec. We note that an alternative method for

measuring long-range diffusion involves observing the non-

exponential decay of magnetization within a single slice that is

repeatedly imaged; one can extract the rate of diffusion of spin

magnetization from neighboring regions into the imaged slice

[85]. One could also monitor the ‘washout’ of a slice into

adjacent slices with zero magnetization or the modulation of

slice-to-slice contrast. The method described here appears to be

a more direct measurement of Dsec.

4.5. In vivo Imaging of Dsec in dogs with unilateral emphysema

Our group has acquired images of Dsec via magnetization

tagging in live dogs with elastase-induced emphysema in one

lung only [75,86]. This unilateral emphysema model allowed

simultaneous control experiments in the left lungs of each

animal. Restriction to diffusion of 3He gas was significant in

the control lungs, as expected, and the measured long-range

diffusivity was about 12!smaller than short-range diffusivity

in the same lungs (0.015 versus 0.19 cm2/s). Fig. 12

demonstrates an obviously faster decay of magnetization

tagging in the emphysematous lung over the control lung;

there were marked increases (300%) of Dsec in the emphyse-

matous lungs compared to control lungs. More modest

increases in Dmsec (160%) were observed in the same animals

(Fig. 12).

This increase in Dsec for canine, elastase-induced emphy-

sema is strikingly smaller than that observed in several human

patients with severe emphysema (Fig. 14 and Ref. [75]). We

have demonstrated that this canine model reflects diffuse

panacinar emphysema, similar to that often encountered in

patients with aK1 antitrypsin deficiency. Although this is

different than typical smokers’ emphysema, the canine images

show that our technique for measuring Dsec can be utilized

in vivo and reflects a significant increase in collateral pathways

in emphysema via an increase in long-range 3He diffusion over

control lungs.

4.6. Imaging Dsec in explanted human lungs with emphysema

Explanted lungs offer distinct advantages for 3He imaging

and functional lung research: they can be held at inspiration for

several minutes, and the lack of saline (with attendant rf loss)

provides for the use of high-sensitivity coils for multiple

measurements on one bolus of gas. We imaged Dsec in

explanted human lungs from transplant recipients with

advanced COPD and in healthy donor lungs that were rejected

for transplant due to recipient mismatch [95].

The lungs with advanced COPDwere removed at transplant,

fitted with a bronchial connection to laboratory tubing, sealed

of leaks, and purged of O2 (the paramagnetic effects of which

depolarize 3He) with N2. The donor lungs were prepared

similarly. With the lungs at approximately functional residual

capacity, the lungs were repeatedly ventilated with 300-mL

tidal volumes of hyperpolarized 3He to mix the gas for

diffusion imaging. After there was sufficient gas in all areas of

the lung, Dsec was measured by preparing sinusoidally

modulated magnetization and then repeatedly imaging with

FLASH.

Fig. 13 clearly demonstrates the small decay of modulation

in a normal donor (control) lung. A value of 0.017 cm2/s was

measured for Dsec nearly uniformly in the 20-mm, approxi-

mately sagittal slice; this value is a factor of ten smaller than

Dmsec measured in the same lung. In stark contrast, Fig. 14

shows similar images in an emphysematous lung. Decay of the

sinusoidal modulation is apparent already at the second image,

taken 0.33 s after the first. In particular the upper lobe (at left in

each image) shows marked decay, as reflected in the map of

Dsec (0.6 cm2/s in this region, an increase by a factor of 30 over

the control lung).

4.7. In vivo Imaging of Dsec in humans

Measurement of the diffusivity from inspection of magne-

tization-tagged images is significantly affected by bulk flow, so

complete breath hold during the entire experiment time is

essential. We note that proper breath hold for other diffusion

imaging modalities is also important, but small amounts of flow

will affect Dmsec much less than Dsec, because Dsec measures

displacements over such long times (see discussion in

Section 2). Since tagging decay must be measurable after the

approximately 10 s of breath hold, we have not increased

in vivo tagging wavelengths beyond 3 cm. Shown in Fig. 15 are

inspection images of tagged magnetization in a healthy

volunteer. In these images we see slow decay of the tagging

contrast and some heterogeneity—the apex decays more

quickly than the base. We suspect that this may be mixing

driven by cardiac motion.

In summary, measurements of the ADC over centimeter

length-scales (Dsec) is reduced in healthy lungs by a factor of

about 50 from the free diffusivity Do. This large restriction is

attributed to the tortuous network of airways. In some regions

of severely emphysematous lungs, Dsec values approaching Do

have been measured, demonstrating the large range of increase

possible for this measure of lung microstructure. The increase

in Dsec in emphysema is due to increases in collateral

Page 16: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

Fig. 14. Decay of sinusoidal modulation (grayscale) and resultant long-range diffusion map (color) in an emphysematous, explanted human lung. The rate of

diffusion of 3He is very high compared to normal lungs, especially in the upper lobe (at left in the images). Each grayscale image advances by 0.33 s.

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–8378

ventilation pathways, so Dsec may be useful in distinguishing

different phenotypes of emphysema.

5. Fluorine-19 ADC measurements

5.1. Rationale

This review is focused on the diffusion of hyperpolarized3He gas in lungs. Nevertheless, recent results on the diffusion

of perfluorinated gases in lungs are relevant. Imaging of these

inert gases, primarily C2F6, C3F8, and possibly SF6 and CF4,

offers advantages over 3He: laser-driven polarizing apparatus is

not involved, resulting in a simpler and less expensive

technique; such gases gain S/N by having as many as six

equivalent 19F spins per molecule and a short T1 to allow rapid

signal averaging; and the gases may be mixed with oxygen to

allow continuous breathing (as opposed to breath hold of a

single bolus). The major drawback is that the perfluorinated

gases offer much less image S/N in a given acquisition time

compared to hyperpolarized 3He. In addition, long-distance

diffusion measurements as described in Section 4 for 3He, are

impossible for gases with nuclei of short T1. It remains to be

seen whether the perfluorinated gases will enter into wide-

spread use.

Several groups have reported spin-density or ventilation

images from 19F MRI using SF6 [11,12]. This gas has an

exceptionally short T1 and T2 of about 1.8 ms at room

temperature and 1 atm pressure [87]. We note that the first

images of the air-spaces of lungs used CF4 in dogs, by

Lauterbur et al. [88]; CF4 has a similar short relaxation time as

SF6 [87].

However, diffusion measurements to determine the local

severity of emphysema in lungs impose further requirements.

First, the T2 must be sufficiently long to allow the diffusivity D

to be measured with the available gradient strengths. As the

Fig. 15. Tagging decay in a healthy volunteer, in vivo; each grayscale image progres

coefficient is consistent with ex vivo studies by us and in vivo studies in Nottingh

quickly than at the base of the lungs. This may be due to cardiogenic mixing at br

goal of all this work is clinical implementation, this means

gradients of no more than 40 mT/m, for today’s whole body

MRI instrument. Second, the diffusion must be measured using

a sufficiently long diffusion time, t, so that the diffusive

motions of the gas molecules are substantially restricted by the

lung microstructure. Diffusion measurements on gases of very

short T2 necessarily use short diffusion times t and will report a

value very close to the free diffusivity D0 of the gas (see

below). For both reasons, the gases C2F6 and C3F8 (with T1ZT2 of 10 and 20 ms, respectively) are more suitable for

measurements of diffusion in human lungs. We note recent

measurements of D in rat lungs using SF6 gas [89].

The reduced diffusivity, D(t)/D0, is sketched in Fig. 16 as a

function offfiffit

p(after scaling of this variable to dimensionless

units) [90]. The ratio of surface area to gas-phase volume of the

porous structure or lung, S/V, can be used to define the

reciprocal of the characteristic length or feature size. Thus, the

dimensionless variable along the horizontal axis,

xh ðS=VÞffiffiffiffiffiffiffiD0t

p, is approximately the ratio of the free (unrest-

ricted) root mean square displacement in time t to the feature

size, (S/V)K1. In healthy human lungs [74], S/V is approxi-

mately 200 cmK1, the reciprocal of 50 mm. For x/1,

essentially no gas molecules contact a restricting wall in time

t, so DZD0. For x[1 molecules make many collisions with

walls during the measurement duration t and D approaches D0/

a, where a is the tortuosity as defined in the standard theory of

porous structures. (An interesting issue is that lungs with their

bifurcating branches do not truly have a characteristic length,

so they do not strictly possess a unique value of tortuosity a, as

would a sponge for example.) For intermediate values of

diffusion time (that is, x less than or of order unity), a famous

approximation [91] exists between the diffusivity and t,

DðtÞ=D0y1Kð4=9ffiffiffip

pÞðS=VÞ

ffiffiffiffiffiffiffiD0t

p: (18)

ses in time by 2.9 s. The rate of decay is low, and the extracted average diffusion

am [79]. Some heterogeneity is visually apparent; gas in the apex mixes more

eath hold.

Page 17: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

Fig. 16. Universal plot of time-dependent diffusivity D(t) scaled by free

diffusivity D0 as a function offfiffit

p(scaled) in a typical porous structure; S/V is

the ratio of surface area to gas volume. At small times, only a fraction of gas

atoms collide with the walls, resulting in a linear decrease of D withffiffit

p. At

large times, the diffusivity is reduced by the tortuosity factor. For practical

diffusion times t, the gases C2F6 and C3F8 are in the small time region while3He is in the large time limit.

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–83 79

This expression describes the linear portion of the graph in

Fig. 16. Aside from the numerical factors, this relationship has

a simple physical explanation [92]. For 3He, the very large D0

of 0.88 cm2/s (dilute in N2 or air) results in typical x values of

25, for typical t values of 1–5 ms as is common (see above);

this places 3He measurements well into the tortuosity limit and

well outside the realm of applicability of Eq. (18). For C2F6 or

C3F8 (D0Z0.035 or 0.022 cm2/s, respectively [90]), the x

values are approximately 3–4 (see Fig. 16). Thus, imaging

measurements of restricted diffusion using these heavy, large

gas molecules should allow the surface-to-volume ratio to be

determined on a voxel-by-voxel basis, provided D0 is known.

Measurement of the local S/V should be a particularly valuable

characterization of lung microstructure in emphysema; we

know of no competing non-invasive method for measuring S/V

in lungs with spatial resolution. We note that the approximate

relationship of Eq. (18) is model-free; thus the interpretation in

terms of S/V is comparatively unambiguous.

Fig. 17. Images of restricted diffusion of C3F8 in freshly excised healthy (H)

and emphysematous (E) human lungs. The diffusion in (H) is more restricted, as

indicated by the color bar in centimeter2/second at right. The small arc of signal

below lung (E) is from gas in the tubing connecting to the lung. Each image

shown was acquired in 6 min. The partitions shown are 32 mm thick with 5.5!

5.5 mm in-plane resolution.

5.2. Methods

The images and results presented here used 100%

concentration of perfluorinated gas in excised lungs, to avoid

regulatory issues. It is our opinion that explanted emphysema-

tous human lungs are superior models of the disease in live

humans, compared to animal models. A high Q solenoidal rf

coil oriented sideways in the magnet bore (1.5 T) was used.

Three-dimensional data acquisition results in much more

efficient (faster) signal averaging than sequential-slice two-

dimensional acquisition. Typically, voxels were 5.5!5.5!32 mm3. For C3F8, only the 6 equiv. perfluoromethyl spins

were excited. The two center 19F spins are chemically shifted

by 42 ppm, about 2380 Hz at 59.85 MHz, and were placed in a

null of the rf spectrum of the pulses [90].

5.3. Results

The excised lungs were exhausted of air and very uniformly

refilled with 100% C2F6 or C3F8 gas using a bell jar apparatus,

designed to function with the same pressure inside and outside

the lung. Thus, there should be essentially the same amount of

gas in each image voxel. Indeed, spin-echo images (formed

using an rf p-pulse) obtained without diffusion attenuation are

strikingly uniform in signal intensity [90]. However, gradient-

echo images show intensity variations, most notably as regions

of low intensity due to dephasing effects across the large voxels

employed. These artifacts become larger at longer gradient

echo times (typical values are 4–10 ms), as expected for

susceptibility/dephasing effects. However, as remarked below,

the use of spin echoes in human subjects will involve a

substantial increase in deposited rf energy (SAR).

Restricted diffusivity results for C3F8 are presented in

Fig. 17, comparing a normal donor lung to an explanted

severely emphysematous lung. Both panels show individual

partitions (three-dimensional equivalent of slices); outside the

chest cavity, lungs take on quite different shapes. Judging from

the color bar, the diffusion of C3F8 is distinctly smaller (more

restricted) in the healthy lung than in the emphysematous lung.

This result shows the potential of 19F MR to characterize

emphysematous lung changes.

The histogram of Fig. 18 displays the data from C2F6 on two

normal donor lungs (nominally healthy) and eight explanted

lungs with severe emphysema. The data from the two groups

are well resolved with relatively little overlap. The average

ADC is 0.017 cm2/s for the normal lungs and 0.032 cm2/s for

the emphysematous group. Analysis of the raw data indicates

that a significant component of the standard deviation (see

figure caption) is from measurement noise. The factor of nearly

two change in the mean diffusivity between healthy and

emphysematous lungs is promising.

Page 18: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

Fig. 18. Histogram of C2F6 ADC in two healthy (H) and eight emphysematous

(E) excised lungs. The vertical scale is correct for (E) but has been adjusted for

the healthy lungs to make the peak heights comparable. The average ADC and

standard deviation are 0.017 (0.0037) for healthy and 0.032 (0.0094) for

emphysema, all in centimeter2/second.

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–8380

5.4. Transition to human imaging

The primary issues for in vivo human imaging are S/N, rf

energy deposition in the chest (SAR), and safety of the gaseous

agent. This last question is outside the scope of this review, but

the perfluorinated gases are exceptionally inert, appear to have

no toxicity, and have only small anesthetic activity (e.g. C2F6 is

similar to argon and only twice as active as nitrogen [93]). We

note that C3F8 is currently used to inflate microspheres in the

ultrasound contrast agent Optison, which may ease the

approval process for use of C3F8 in lung imaging of human

subjects.

On a per-spin basis, the solenoidal rf coil used with the

excised lungs here is estimated from phantom measurements to

be approximately 13 dB (!20 in power units) more sensitive

than a simple Helmholtz pair in a live subject. Some of the loss

of sensitivity in vivo comes from the smaller filling factor,

some from the long-recognized inferior performance of

transverse rf coils, and a substantial portion from the rf losses

due to saline conductivity in the chest. Thus, the rf power

required for specific nutation angles in a given time is increased

by a factor of 20 and, by reciprocity, 20 times as much signal

averaging time is required to accumulate a specified final S/N.

The need to maintain the deposited rf power within safety

guidelines pushes the technique (i) away from spin echoes

towards gradient echoes and (ii) away from the gases such as

SF6 and CF4 with very short T1 and T2. Consider the spin echo

sequence with p/2 and p rf pulses of equal durations: the ppulse has twice the amplitude and four times the rf energy as

the p/2 pulse. Thus, a spin echo sequence deposits five times

the energy as a gradient echo sequence.

The (S/N)2 accumulated in a given averaging time is

proportional to the fraction of real time that the spin signal is

present. Approximately, each signal persists for time T�2 and

can be repeated every T1; thus (S/N)2 is proportional to T�

2 =T1.

Since each of the gases considered here has nearly equal values

of T2, T�2 , and T1, all should yield the same S/N in a given time,

for equal numbers of equivalent spins per molecule. There is no

advantage to the very fast T1 (and T2) of SF6 and CF4; in fact,

these gases would have to be used with very high pulse

repetition rates to compete with C2F6 and C3F8 because of their

short T2 values, leading to excessive rf power dissipation. As

noted earlier, the short T1 gases are also not suitable for

measuring restricted diffusion. We note that gases with even

longer T1ZT2 than C3F8 would suffer from T�2 becoming

smaller than T2, yielding less efficient accumulation of (S/N)2.

The present results demonstrate that the restricted diffusiv-

ity of C2F6 and C3F8 can distinguish healthy and emphysema-

tous lung tissue and may be able to provide a spatially resolved

measurement of the surface area to volume ratio, S/V [94]. This

method would not require hyperpolarization apparatus and

would be more readily adoptable at medical sites. Based on the

imaging times used here with the high sensitivity rf solenoid

coil and the expected factor of 20 increase in averaging time, it

appears that additional improvements in S/N will be needed for

this technique to be practical. The available avenues for

improvement include use of a higher static field strength than

the present 1.5 T and use of a superior rf coil such as a phased

array.

Acknowledgements

The authors gratefully acknowledge the input and assistance

of their many colleagues and collaborators. In particular, the

medical knowledge of Joel D. Cooper, Steven S. Lefrak, and

David S. Gierada was invaluable. Rick Jacob lead the effort on

the use of C2F6 and C3F8 gases and Yulin V. Chang performed

many of the 3He and 19F measurements. A. L. Sukstanskii is

responsible for the mathematics of the diffusion in randomly-

oriented cylinders. S.S. Gross performed the computer

simulations of restricted gas diffusion. The 3He work was

supported in part by an NIH grant to DAY, R01 HL70037; the19F effort was partly supported by the Gas Enabled Medical

Innovations fund. Finally, we thank the patients, lung

transplant recipients, and the families of lung donors for their

important contributions to the research.

References

[1] T.L. Petty, G.G. Weinmann, Building a national strategy for the

prevention and management of and research in chronic obstructive

pulmonary disease: National Heart, Lung, and Blood Institute workshop

summary, JAMA 277 (1997) 246–253.

[2] M.S. Albert, G.D. Cates, B. Driehuys, W. Happer, B. Saam, C.S. Springer

Jr., A. Wishnia, Biological magnetic resonance imaging using laser-

polarized 129Xe, Nature 370 (1994) 199–201.

[3] R.D. Black, H.D. Middleton, G.D. Cates, G.P. Cofer, B. Driehuys,

W. Happer, L.W. Hedlund, G.A. Johnson, M.D. Shattuck, J.C. Swartz, In

vivo He-3 MR images of guinea pig lungs, Radiology 199 (1996) 867–

870.

[4] J.R. MacFall, H.C. Charles, R.D. Black, H. Middleton, J.C. Swartz,

B. Saam, B. Driehuys, C. Erickson, W. Happer, G.D. Cates,

G.A. Johnson, C.E. Ravin, Human lung air spaces: potential for MR

imaging with hyperpolarized He-3, Radiology 200 (1996) 553–558.

[5] M. Ebert, T. Großmann, W. Heil, W.E. Otten, R. Surkau, M. Leduc,

P. Bachert, M.V. Knopp, L.R. Schad, M. Thelen, Nuclear magnetic

Page 19: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–83 81

resonance imaging with hyperpolarised helium-3, Lancet 347 (1996)

1297–1299.

[6] X.J. Chen, H.E. Moller, M.S. Chawla, G.P. Cofer, B. Driehuys,

L.W. Hedlund, G.A. Johnson, Spatially resolved measurements of

hyperpolarized gas properties in the lung in vivo. Part I: diffusion

coefficient, Magn. Reson. Med. 42 (1999) 721–728.

[7] G.A. Johnson, G.P. Cofer, L.W. Hedlund, R.R. Maronpot, S.A. Suddarth,

Registered 1H and 3He magnetic resonance microscopy of the lung,

Magn. Reson. Med. 45 (2001) 365–370.

[8] J.B. West, Pulmonary Pathophysiology, fourth ed., Williams & Wilkins,

Baltimore, MA, 1992, pp. 1–62.

[9] J.P. Mugler, B. Driehuys, J.R. Brookeman, G.D. Cates, S.S. Berr,

R.G. Bryant, T.M. Daniel, E.E. de Lange, J.H. Downs, C.J. Erickson,

W. Happer, D.P. Hinton, N.F. Kassel, T. Maier, C.D. Phillips, B.T. Saam,

K.L. Sauer, M.E. Wagshul, MR imaging and spectroscopy using

hyperpolarized 129Xe gas: preliminary human results, Magn. Reson.

Med. 37 (1997) 809–815.

[10] K. Ruppert, J.F. Mata, J.R. Brookeman, K.D. Hagspiel, J.P. Mugler III,

Exploring lung function with hyperpolarized 129Xe nuclear magnetic

resonance, Magn. Reson. Med. 51 (2004) 676–687.

[11] D.O. Kuethe, A. Caprihan, E. Fukushima, R.A. Waggoner, Imaging lungs

using inert fluorinated gases, Magn. Reson. Med. 39 (1998) 85–88.

[12] W.G. Schreiber, B. Eberle, S. Laukemper-Ostendorf, K. Markstaller,

N. Weiler, A. Scholz, K. Burger, C.P. Heusel, M. Thelen, H.-U. Kauczor,

Dynamic 19F-MRI of pulmonary ventilation using sulfur hexafluoride

(SF6) gas, Magn. Reson. Med. 45 (2001) 605–613.

[13] H.E. Moller, X.J. Chen, B. Saam, K.D. Hagspiel, G.A. Johnson,

T.A. Altes, E.E. de Lange, H.-U. Kauczor, Magnetic resonance imaging

of the lungs using hyperpolarized noble gases, Magn. Reson. Med. 47

(2002) 1029–1051.

[14] J.C. Leawoods, D.A. Yablonskiy, B. Saam, D.S. Gierada, M.S. Conradi,

Hyperpolarized 3He gas MR imaging, Concepts Magn. Reson. 13 (2001)

277–293.

[15] F.D. Colegrove, L.D. Schearer, G.K. Walters, Polarization of 3He gas by

optical pumping, Phys. Rev. 132 (1963) 2561–2572.

[16] T.R. Gentile, R.D. McKeown, Spin-polarizing 3He nuclei with an arc-

lamp-pumped neodymium-doped lanthanum magnesium hexaluminate

laser, Phys. Rev. A 47 (1993) 456–467.

[17] T.G. Walker, W. Happer, Spin-exchange optical pumping of noble-gas

nuclei, Rev. Mod. Phys. 69 (1997) 629–642.

[18] J. Becker, M. Ebert, T. Grossmann, W. Heil, D. Hofmann, H. Humblot,

M. Leduc, E.W. Otten, D. Rohe, K. Siemensmeyer, R. Surkau, M. Steiner,

F. Tasset, N. Trautmann, Realization of a broad band neutron spin filter

with compressed, polarized 3He gas, Nucl. Instrum. Methods Phys. Res.,

Sect. A 384 (1997) 444–450.

[19] J. Becker, W. Heil, B. Krug, M. Leduc, M. Meyerhoff, P.J. Nacher,

E.W. Otten, T. Prokscha, L.D. Schearer, R. Surkau, Study of mechanical

compression of spin-polarized 3He gas, Nucl. Instrum. Methods Phys.

Res., Sect. A 346 (1994) 45–51.

[20] T.R. Gentile, G.L. Jones, A.K. Thompson, R.R. Rizi, D.A. Roberts,

I.E. Dimitrov, R. Reddy, D.A. Lipson, W. Gefter, M.D. Schnall,

J.S. Leigh, Demonstration of a compact compressor for application of

metastability exchange optical pumping of 3He to human lung imaging,

Magn. Reson. Med. 43 (2000) 290–294.

[21] R.E. Jacob, S.W. Morgan, B. Saam, 3He spin exchange cells for magnetic

resonance imaging, J. Appl. Phys. 92 (2002) 1588–1597.

[22] E. Babcock, I. Nelson, S. Kadlecek, B. Driehuys, L.W. Anderson,

F.W. Hersman, T.G. Walker, Hybrid spin-exchange optical pumping of3He, Phys. Rev. Lett. 91 (2003) 123003.

[23] B. Chann, I. Nelson, T.G. Walker, Frequency-narrowed external-cavity

diode-laser-array bar, Opt. Lett. 25 (2000) 1352–1354.

[24] B. Saam, W. Happer, H. Middleton, Nuclear relaxation of 3He in the

presence of O2, Phys. Rev. A 52 (1995) 862–865.

[25] L.D. Schearer, G.K. Walters, Nuclear spin-lattice relaxation in the

presence of magnetic-field gradients, Phys. Rev. 139 (1965) A1398–

A1402.

[26] A. Haase, J. Frahm, D. Matthaei, W. Hanicke, K.-D. Merboldt, FLASH

Imaging: rapid NMR imaging using low flip-angle pulses, J. Magn.

Reson. 67 (1986) 258–266.

[27] M. Kveder, I. Zapancic, G. Lahajnar, R. Blinc, D. Suput, D.C. Ailion,

K. Ganesan, C. Goodrich, Water proton NMR relaxation mechanisms in

lung tissue, Magn. Reson. Med. 7 (1988) 432–441.

[28] D.A. Yablonskiy, Quantitation of intrinsic magnetic susceptibility-related

effects in a tissue matrix. Phantom study, Magn. Reson. Med. 39 (1998)

417–428.

[29] X.J. Chen, H.E. Moller, M.S. Chawla, G.P. Cofer, B. Driehuys,

L.W. Hedlund, J.R. MacFall, G.A. Johnson, Spatially resolved

measurements of hyperpolarized gas properties in the lung in vivo part

II: T2*, Magn. Reson. Med. 42 (1999) 729–737.

[30] B. Saam, D.A. Yablonskiy, D.S. Gierada, M.S. Conradi, Rapid imaging of

hyperpolarized gas using EPI, Magn. Reson. Med. 42 (1999) 507–514.

[31] E.E. de Lange, J.P. Mugler III, J.R. Brookeman, J. Knight-Scott,

J.D. Truwit, C.D. Teates, T.M. Daniel, P.L. Bogorad, G.D. Cates, Lung

air spaces: MR imaging evaluation with hyperpolarized 3He gas,

Radiology 210 (1999) 851–857.

[32] M. Viallon, G.P. Cofer, S.A. Suddarth, H.E. Moller, X.J. Chen,

M.S. Chawla, L.W. Hedlund, Y. Cremillieux, G.A. Johnson, Functional

MR microscopy of the lung using hyperpolarized 3He, Magn. Reson.

Med. 41 (1999) 787–792.

[33] X.J. Chen, M.S. Chawla, L.W. Hedlund, H.E. Moller, J.R. MacFall,

G.A. Johnson, MR microscopy of lung airways with hyperpolarized 3He,

Magn. Reson. Med. 39 (1998) 79–84.

[34] P. Mansfield, Multi-planar image formation using NMR spin echoes,

J. Phys. C 10 (1997) L55–L58.

[35] M. Viallon, Y. Berthezene, V. Callot, M. Bourgeois, H. Humblot,

A. Briguet, Y. Cremillieux, Dynamic imaging of hyperpolarized 3He

distribution in rat lungs using interleaved-spiral scans, NMR Biomed. 13

(2000) 207–213.

[36] M. Salerno, T.A. Altes, J.R. Brookeman, E.E. de Lange, J.P. Mugler III,

Dynamic spiral MRI of pulmonary gas flow using hyperpolarized 3He:

preliminary studies in healthy and diseased lungs, Magn. Reson. Med. 46

(2001) 667–677.

[37] W.A. Edelstein, G.H. Glover, C.J. Hardy, R.W. Redington, The intrinsic

signal-to-noise ratio in NMR imaging, Magn. Reson. Med. 3 (1986) 604–

618.

[38] M. Salerno, J.R. Brookeman, E.E. de Lange, J.P. Mugler III,

Hyperpolarized 3He lung imaging at 0.5 and 1.5 Tesla: a study of

susceptibility-induced effects, Magn. Reson. Med. 53 (2005) 212–216.

[39] E. Durand, G. Guillot, L. Durasse, G. Tastevin, P.J. Nacher, A. Vignaud,

D. Vattolo, J. Bittoun, CPMG measurements and ultrafast imaging in

human lungs with hyperpolarized helium-3 at low field (0.1 T), Magn.

Reson. Med. 47 (2002) 75–81.

[40] J. Hennig, A. Nauerth, H. Friedburg, RARE imaging: a fast imaging

method for clinical MR, Magn. Reson. Med. 3 (1986) 823–833.

[41] R.W. Mair, M.I. Hrovat, S. Patz, M.S. Rosen, I.C. Ruset, G.P. Topulos,

L.L. Tsai, J.P. Butler, F.W. Hersman, R.L. Walsworth, 3He lung imaging

in an open acces, very-low-field human magnetic resonance imaging

system, Magn. Reson. Med. 53 (2005) 745–749.

[42] D.A. Yablonskiy, A.L. Sukstanskii, J.J.H. Ackerman, Image artifacts in

very low magnetic field MRI: the role of concomitant gradients, J. Magn.

Reson. 174 (2005) 279–286.

[43] D.G. Norris, J.M.S. Hutchison, Concomitant magnetic field gradients and

their effects on imaging at low magnetic field strength, Magn. Reson.

Imaging 8 (1990) 33–34.

[44] A.J. Deninger, B. Eberle, M. Ebert, T. Großmann, W. Heil, H.U. Kauczor,

L. Lauer, K. Markstaller, E. Otten, J. Schmiedeskamp, W. Schreiber,

R. Surkau, M. Thelen, N. Weiler, Quantification of regional intrapul-

monary oxygen partial pressure evolution during apnea by 3He MRI,

J. Magn. Reson. 141 (1999) 207–216.

[45] R.R. Rizi, J.E. Baumgardner, M. Ishii, Z.Z. Spector, J.M. Edvinsson,

A. Jalali, J. Yu, M. Itkin, D.A. Lipson, W. Gefter, Determination of

regional VA/Q by hyperpolarized 3HeMRI, Magn. Reson. Med. 52 (2004)

65–72.

Page 20: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–8382

[46] J.B. West, P.D. Wagner, Ventilation-perfusion relationships in:

R.G. Crystal, J.B. West, E.R. Weibel, P.J. Barnes (Eds.), The Lung:

Scientific Foundations, vol. 2, second ed., Lippincott/Raven, Philadel-

phia, PA, 1997, pp. 1693–1709.

[47] International Workshop on Pulmonary Imaging, University of Pennsyl-

vania, Philadelphia.

[48] J.B. West, Respiratory Physiology—The Essentials, fifth ed., Williams &

Wilkins, Baltimore, MD, 1995.

[49] B. Saam, D.A. Yablonskiy, V.D. Kodibagkar, J.C. Leawoods,

D.S. Gierada, J.D. Cooper, S.L. Lefrak, M.S. Conradi, MR Imaging of

diffusion of He-3 gas in healthy and diseased lungs, Magn. Reson. Med.

44 (2000) 174–179.

[50] P.T. Callaghan, Principles of Nuclear Magnetic Resonance Microscopy,

Clarendon Press, Oxford, 1991.

[51] D.A. Yablonskiy, A.L. Sukstanskii, J.C. Leawoods, D.S. Gierada,

G.L. Bretthorst, S.S. Lefrak, J.D. Cooper, M.S. Conradi, Quantitative

in vivo assessment of lung microstructure at the alveolar level with

hyperpolarized 3He diffusion MRI, Proc. Natl Acad. Sci. USA 99 (5)

(2002) 3111–3116.

[52] E.O. Stejskal, J.E. Tanner, Spin diffusion measurements: spin echoes in

the presence of a time-dependent field gradient, J. Chem. Phys. 42 (1965)

288–292.

[53] T. Tanoli, D.A. Yablonskiy, T. Bae, J.D. Cooper, J.C. Hogg, J.C. Woods,

M.S. Conradi, Correlation of CT and 3He MRI evidence of emphysema in

human lungs, in preparation.

[54] M. Salerno, J.R. Brookeman, J.P. Mugler III, Time-dependent hyperpol-

arized 3He diffusion MR imaging: initial experience in healthy and

emphysematous lungs, Proceedings of the International Society for

Magnetic Resonance in Medicine, Ninth Meeting, 2001, p. 950.

[55] A.E. Morbach, K.K. Gast, J. Schmiedeskamp, A. Dahmen, A. Herweling,

C.P. Heussel, H.-U. Kauczor, W.G. Schreiber, Diffusion-weighted MRI

of the lung with hyperpolarized helium-3: a study of reproducibility,

J. Magn. Reson. Imaging 21 (2005) 765–774.

[56] P. Akeson, I. Casselbrant, As Presented at 2004 International Workshop

on Pulmonary Imaging, University of Pennsylvania, Philadelphia, PA.

[57] J.P. Mugler III, J.R. Brookeman, J. Knight-Scott, T. Maier, E.E. de Lange,

P.L. Bogorad, Regional measurement of the 3He diffusion coefficient in

the human lung, Proceedings of the International Society for Magnetic

Resonance in Medicine, Sixth Meeting, 1998, p. 1906.

[58] J.C. Woods, D.A. Yablonskiy, C.K. Choong, J.C. Hogg, J.D. Cooper,

M.S. Conradi, Measuring the progression of emphysema in a canine

model with 3He diffusion MRI, Proc. Int. Soc. Magn. Reson. Med. 13

(2005) 47.

[59] T.S. Tanoli, J.C. Woods, K.T. Bae, M.S. Conradi, J.C. Hogg, J.D. Cooper,

D.A. Yablonskiy, Hyperpolarized 3He diffusion MRI of acinar airways in

canines with induced emphysema: comparison with computed tomogra-

phy, Proc. Int. Soc. Magn. Reson. Med. 13 (2005) 1804.

[60] J.C. Woods, C.K. Choong, D.A. Yablonskiy, K. Chino, J. Bentley, J.

Wong, J.A. Pierce, J.D. Cooper, P.T. Macklem, M.S. Conradi, J.C. Hogg,

Comparison of morphometric data and 3He diffusion MR in emphyse-

matous human lungs, submitted for publication.

[61] J.P. Dugas, J.R. Garbow, D.K. Kobayashi, M.S. Conradi, Hyperpolarized3He MRI of mouse lung, Magn. Reson. Med. 52 (2004) 1310–1317.

[62] L.W. Hedlund, G. Cofer, S. Owen, G.A. Johnson, MR-compatible

ventilator for small animals: computer controlled ventilation for proton

and noble gas imaging, Magn. Reson. Imaging 18 (2000) 753–759.

[63] X.J. Chen, L.W. Hedlund, H.E. Moeller, M.S. Chawla, R.R. Maronpot,

G.A. Johnson, Detection of emphysema in rat lungs using magnetic

resonance measurements of 3He diffusion, Proc. Natl Acad. Soc. 97

(2000) 11478–11481.

[64] E. Weibel, Design of airways and blood vessels considered as branching

trees, in: N.S. Cherniack (Ed.), The Lung: Scientific Foundations, Raven

Press, New York, 1991, pp. 711–720.

[65] B. Haefeli-Bleuer, E.R. Weibel, Morphometry of the human pulmonary

acinus, Anat. Rec. 220 (1988) 401–414.

[66] D.C. Douglass, D.W. McCall, Diffusion in paraffin hydrocarbons, J. Phys.

Chem. 62 (1958) 1102–1107.

[67] C.H. Neuman, Spin echo of spins diffusing in a bounded medium,

J. Chem. Phys. 60 (1973) 4508–4511.

[68] E.L. Hahn, Spin echoes, Phys. Rev. 80 (1950) 580–594.

[69] H.Y. Carr, E.M. Purcell, Effects of diffusion on free precession in nuclear

magnetic resonance experiments, Phys. Rev. 94 (1954) 630–638.

[70] A. Nagai, W.M. Thurlbeck, K. Konno, Responsiveness and variability of

airflow obstruction in chronic obstructive pulmonary disease. Clinico-

pathologic correlative studies, Am. J. Respir. Crit. Care Med. 151 (1995)

635–639.

[71] A. Nagai, W.M. Thurlbeck, Scanning electron microscopic observations

of emphysema in humans. A descriptive study, Am. Rev. Respir. Dis. 144

(1991) 901–908.

[72] M. Salerno, J.R. Brookeman, E.E. deLange, J. Knight-Scott, J.P.Mugler III,

Detection of regional microstructural changes of the lung in emphysema

using hyperpolarized 3He diffusion MRI, in: Eighth Conference of

International Society for Magnetic Resonance in Medicine, Denver, 2000.

[73] D.A. Yablonskiy, B. Saam, D.S. Gierada, S.S. Lefrak, J.D. Cooper, M.S.

Conradi, Diffusion lung imaging of hyperpolarized 3He in patients with

emphysema: changes in the structure of Alveoli, in: RSNA 84th Scientific

Assembly and Annual Meeting, Chicago, IL, 1999.

[74] H.O. Coxson, R.M. Rogers, K.P. Whittall, Y. D’Yachkova, P.D. Pare,

F.C. Sciurba, J.C. Hogg, A quantification of the lung surface area in

emphysema using computed tomography, Am. J. Respir. Crit. Care Med.

159 (1999) 851–856.

[75] J.C. Woods, D.A. Yablonskiy, K. Chino, T.S.K. Tanoli, J.D. Cooper,

M.S. Conradi, Magnetization tagging decay to measure long-range 3He

diffusion in healthy and emphysematous canine lungs, Magn. Reson.

Med. 51 (2004) 1002–1008.

[76] J.C. Hogg, P.T. Macklem, W.M. Thurlbeck, The resistance of collateral

channels in excised human lungs, J. Clin. Invest. 48 (1969) 421–431.

[77] P.B. Terry, R.J. Traystman, H.H. Newball, G. Batra, H.A. Menkes,

Collateral ventilation in man, N. Engl. J. Med. 298 (1978) 10–15.

[78] L.A. Engel, P.T. Macklem, Gas mixing and distribution in the lung, Int.

Rev. Physiol. 14 (1977) 37–82.

[79] J.R.Owers-Bradley,A.Bennattayalah, S.Fichele, J.C.McGloin,R.Bowtell,

P.S. Morgan, Diffusion and tagging of hyperpolarized 3He in the lungs,

Proceedings of the 10th Annual Meeting ISMRM, Honolulu, 2002, p. 2016.

[80] P.T. Macklem, Airway obstruction and collateral ventilation, Physiol.

Rev. 51 (1971) 368–436.

[81] H.F. Lausberg, K. Chino, A. Patterson, B.F. Meyers, P.D. Toeniskoetter,

J.D. Cooper, Bronchial fenestration improves expiratory flow in

emphysematous lungs, Ann. Thorac. Surg. 75 (2003) 393–398.

[82] L. Alex, L. Dougherty, MR imaging of motion with spatial modulation of

magnetization, Radiology 171 (1989) 841–845.

[83] J.R. Owers-Bradley, S. Fichele, A. Bennattayalah, C.J.S. McGloin,

R.W. Bowtell, P.S. Morgan, A.R. Moody, MR tagging of human lungs

using hyperpolarized 3He gas, J. Magn. Reson. Imaging 17 (2003) 142–

146.

[84] H.C. Torrey, Phys. Rev. 104 (1956) 563.

[85] S. Fichele, M.N. Paley, N. Woodhouse, P.D. Griffiths, E.J. van Beek,

J.M. Wild, J. Magn. Reson. 174 (2005) 28–33.

[86] C.K. Choong, P.D. Toeniskoetter, J.D. Cooper, H.F. Lausberg, K.T. Bae,

J.A. Pierce, J.C. Hogg, Exp. Lung Res. 30 (2004) 319–332.

[87] J.A. Courtney, R.L. Amstrong, A nuclear spin relaxation study of the spin-

rotation interaction in spherical topmolecules,Can. J. Phys. 50 (1972) 1252–

1261.

[88] P.A. Rinck, S.B. Peterson, P.C. Lauterbur, NMR-imaging von fluorhalti-

gen substanzen, Fortschr. Rontgenstr. 140 (1984) 239–243.

[89] J.M. Perez-Sanchez, R.P. de Alejo, I. Rodriguez, M. Cortijo, G. Peces-

Barba, J. Ruiz-Cabello, In vivo diffusion weighted 19F MRI using SF6,

Magn. Reson. Med. 54 (2005) 460–463.

[90] R.E. Jacob, Y.V. Chang, C.K. Choong, A. Bierhals, D.Z. Hu, J. Zheng,

D.A. Yablonskiy, J.C. Woods, D.S. Gierada, M.S. Conradi, 19F MR

imaging of ventilation and diffusion in excised lungs, Magn. Reson. Med.

54 (2005) 577–585.

[91] P.P. Mitra, P.N. Sen, L.M. Schwartz, Short-time behavior of the diffusion

coefficient as a geometrical probe of porous media, Phys. Rev. B 47

(1993) 8565–8574.

Page 21: Hyperpolarized He and perfluorocarbon gas diffusion MRI of ...saam/reprints/Hediffrev.pdf1 relaxation times. We recall that the conventional thermal polarization P th in an applied

M.S. Conradi et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 48 (2006) 63–83 83

[92] T.M. de Swiet, P.N. Sen, Decay of nuclear magnetization by bounded

diffusion in constant field gradient, J. Chem. Phys. 100 (1994) 5597–5604.

[93] K.W. Miller, W.D.M. Paton, E.B. Smith, R.A. Smith, Physiochemical

approaches to the mode of action of general anesthetics, Anesthesiology 36

(1972) 339–351.

[94] M.S. Conradi, M.A. Bruns, D.A. Yablonskiy, A.N. Sukstanskii,

S.S. Gross, J.C. Leawoods, Feasibility of diffusion-NMR surface-to-

volume measurements tested by computer simulations, J. Magn. Reson.

169 (2004) 196–202.

[95] J.C. Woods, D.A. Yablonskiy, C.K. Choong, K. Chino, J.A. Pierce,

J.C. Hogg, J. Bentley, J.D. Cooper, M.S. Conradi, P.T. Macklem,

Long-range diffusion of hyperpolarized 3He in explanted normal and

emphysematous human lungs via magnetization tagging, J. Appl.

Physiol. 99 (2005) 1992–1997.


Recommended