2010 Medphyslib Ray Hashman Hashemi 2
2010 Medphyslib Ray Hashman Hashemi 2
2010 Medphyslib Ray Hashman Hashemi 2
To my MRI fellows over the last quarter century who raised many of
the questions answered in this book. At UCSD I now go through “The
Basics” with the MRI fellows from cover to cover over the course of
the year.
W G B
W G B
Preface
The first two editions of MRI: The Basics were well received, and we
trust enhanced MRI knowledge and practical application. This current
edition of the The Basics spans the gamut from basic physics to
multiuse MR options to specific applications. Although MRI is a maturing
technology, it, nevertheless, continues to challenge radiologists,
residents, and technologists. More improvements and new features are
constantly being developed and applied producing diagnostic images
faster, with higher image quality and reproducibility. This power,
however, invariably gives the MR user more options, and these various
options must be sorted and thought through in order to give optimal
results.
In response to advancements in MRI, we have updated the book and
added completely new chapters addressing parallel imaging, cardiac
MRI, and MR Spectroscopy. These areas are used with increasing
regularity in MR imaging, and we hope that you will glean new insight
into these features and applications in order to enhance your practice.
Again, we are pleased to present to you the third edition of MRI: The
Basics, and hope that it will be a trusty companion during your MRI
journey.
Ray Hashman Hashemi MD, PhD
[+]
Authors
-
Dedication
-
Preface to the First Edition
-
Preface
-
Acknowledgments
↑
TABLE OF CONTENTS
[-]
Part I - Basic Concepts
[+]
1 - Introductory Math
[+]
2 - Basic Principles of MRI
[+]
3 - Radio Frequency Pulse
[+]
4 - T1, T2, and T2*
[+]
5 - TR, TE, and Tissue Contrast
[+]
6 - Tissue Contrast: Some Clinical Applications
[+]
7 - Pulse Sequences: Part I (Saturation, Partial Saturation, Inversion
Recovery)
[+]
8 - Pulse Sequences Part II (Spin Echo)
[+]
9 - Fourier Transform
[+]
10 - Image Construction: Part I (Slice Selection)
[+]
11 - Image Construction: Part II (Spatial Encoding)
[+]
12 - Signal Processing
[+]
13 - Data Space
[+]
14 - Pulse Sequence Diagram
[+]
15 - Field of View
[+]
16 - k-Space: The Final Frontier!
[+]
17 - Scan Parameters and Image Optimization
[+]
18 - Artifacts in MRI
[-]
Part II - Fast Scanning
[+]
19 - Fast Spin Echo
[+]
20 - Gradient Echo: Part I (Basic Principles)
[+]
21 - Gradient Echo: Part II (Fast Scanning Techniques)
[+]
22 - Echo Planar Imaging
[+]
23 - New Scanning Features
[+]
24 - Parallel Imaging
[+]
25 - Tissue Suppression Techniques
[+]
26 - Flow Phenomena
[+]
27 - MR Angiography
[+]
28 - Cardiac MRI
[+]
29 - MR Spectroscopy in the Brain
[+]
30 - High Performance Gradients
[+]
31 - The Many Combinations of MRI
↑
BACK OF BOOK
-
Appendix A: Suggested Readings
[+]
Appendix B - Answers
[+]
Appendix C
[+]
Index
1
Introductory Math
Introduction
In this chapter, we review some of the basic mathematical concepts
that are used in MR imaging. We don't want to scare you away so we
keep things as simple as possible. An understanding of these basic
concepts will help the reader a great deal to comprehend the subtleties
of MR imaging and obtain the necessary tools for manipulating the scan
parameters to improve the quality of the images.
It is not that important to memorize these mathematical formulas;
what's crucial is the understanding of the concepts behind these
formulas. In this chapter, we hope to emphasize the most important
mathematical concepts of MRI physics.
Sinusoidals
Consider a right triangle (having a right angle) with sides a and b and
hypotenuse c and angle x formed by a and c (Fig. 1-1). We can define
sin x (read sine of x), cos x (read cosine of x), tan x (read tangent of x),
cotan x (read co-tangent of x), and arctan x (read arc-tangent of x) in
terms of a, b, and c:
Figure 1-1. A rigt triangle with sides a and b and hypotenuse c
angle x formed by a and c.
Table 1-1
sin x = b/c
cos x = a/c
tan x = sin x/cos x = a/b
arctan b/a = arctan(tan x) = x (Eqn. 1-1)
The variable x can be represented in degrees, that is, 45°, 90°, and
180°, or it can be represented in radians, that is, π/4, π/2, and π,
where π = 180°. Table 1-1 shows x versus. sin x, cos x, and tan x, where
≅ 1.4 so /2 ≅ 0.7 and ≅ 1.7 so /2 ≅ 0.85.
Let's plot x versus sin x (Fig. 1-2). This is called a sinusoidal function.
What about cos x? (Fig. 1-3). Let's now draw cos x and sin x on a single
graph (Fig. 1-4). We can appreciate the symmetry between sin x and
cos x. The difference between the two functions is that sin x is shifted
to the right of cos x by 90°. Later, when we talk about phase and phase
shifts, this mathematical concept will become more important. We can
think of sin x as being cos x with a phase difference of 90°.
Figure 1-4. Sin (x) and cos (x) plotted on the same graph.
So (sin x)2 + (cos x)2 = 1 If we go back to our graph of sin x and cos x
(Fig. 1-4), we can see graphically that because of the phase difference
between the cos x and sin x, the sum of their squares will always equal
1. Another way of looking at sine and cosine is to consider a circle with
a radius of 1 (Fig. 1-5). To understand this concept, it is necessary to
bring up the concepts of vectors, imaginary numbers, and exponentials.
Vector. We'll designate a vector by using a letter such as v with an
arrow above it ( ). This concept will become important later on in the
understanding of resonance of spins and dephasing. A vector is a
mathematical entity that has both a magnitude and a direction. For
example, speed is not a vector—it only has magnitude.
The vector that we've drawn in the circle (Fig. 1-5) has a magnitude of
1. The angle between the vector and the horizontal axis is denoted x. If
we draw perpendiculars from the vector both horizontally and
vertically, we'll get two components of the vector:
a. The horizontal component of the vector would correspond to cos x.
(Remember that the ratio a/c in Figure 1-1 is cos x.)
b. The vertical component of the vector would correspond to sin x.
(Remember that the ratio b/c in Figure 1-1 is sin x.)
Imaginary Numbers. A positive number n2 has two square roots, + n and
− n. For example,
the tangent of the angle between the vector and the x-axis, we get tan
θ = 3/2 = imaginary/real In other words, the ratio of the imaginary part
to the real part gives us the tangent of the angle. The magnitude of
the vector (sometimes called the modulus) is given by the Pythagorean
theorem
When you're dealing with a complex number, if you take the ratio of the
imaginary part to the real part, you get some sort of measure of the
angle of the vector. If you sum the squares of the imaginary and real
part, you get the magnitude of the vector (squared).
Example
If the vector in the circle makes 3 revolutions/sec, then the vector has
a frequency of 3 Hz. Therefore, ω = (2π)(3) = 6π = 18.85 radians/sec.
Sometimes a signal (such as a sine wave) is represented in the following
way: S(t) = sin (ωt) = sin (2πft) Here, we can say that the signal is a
sine wave with a frequency of ω. So, frequency = f = ω/2π.
Example
Draw the signal sin (ωt) versus time t assuming f = 1 Hz (i.e., T = 1 sec
or ω = 2π radians/sec). Thus, sin (ωt) 5 sin (2πt). This equation is
illustrated in Figure 1-8:
when t = 0, then ωt = 0, resulting in
sin ωt = sin 0 = 0;
when t = 1/4, then ωt = (2π) (1/4) = π/2,
resulting in sin ωt = sin π/2 = 1;
when t = 1/2, then ωt= (2π)(1/2) = π,
resulting in sin ωt = sin π = 0; etc.
Phase. Now, let's talk about phase. Consider two sine waves, with one
shifted slightly compared with the other (Fig. 1-9). The two sinusoids
have the same frequency—they oscillate at the same rate—but one of
them is shifted just a little from the other. Suppose they are shifted
apart from each other by a time interval = τ = 1 sec. Suppose further
that the period of 1 cycle = T = 4 sec (Fig. 1-9).
From this, we see that sine and cosine of any given vector have a phase
difference of 90°.
Exponentials
With exponential functions (ex), the letter e is the base of the natural
logarithm with a numerical value of e equals 2.7182818 ≅ 2.72 First,
let's consider the values of ex for various values of x:
for x = 1, e1 = 2.72
for x = 0, e0 = 1
for x = ∞, e−∞ = 0
The nice thing about this exponentially decaying function is that you
can stop anywhere on the curve and the same equation will apply. No
matter how many time constants you extend out along the decay curve,
you still end up with a calculable percentage value of the original
signal.
Table 1-2
t 0 1 2 3 4 5 … ∞
On the recovery curve (1 − e−t/τ), things are just the opposite (Fig. 1-
16). If you draw a tangent at the origin along the exponential recovery
curve (1 − e−t/τ), the point of intersection on the maximum line
corresponding to complete recovery occurs at one time constant (τ). At
this point, we have recovered 63% of the original signal: for t = τ, (1 − e
−t/τ) = 1 − e−1 ≅ 1 − 0.37 = 0.63 At two time constants (2τ), we will have
Sinc Function
There is another function that looks somewhat like Figure 1-17. It is
called a sinc function and is expressed as (sin t)/t:
Question: What is sinc t at time t = 0?
Answer: sinc (0) = sin (0)/0 = 0/0, which is indeterminate. However,
using the principles of differential equations and limits, it can be shown
that sinc (0) = 0/0 = 1 (refer to Question 1-3, at the end). From here
on, sinc (t) is an oscillating wave of t (Fig. 1-18). However, the
envelope of this wave is not an exponential function. This is what a
Radio Frequency (RF) pulse generally looks like. (The frequency or
Fourier transform [FT]
Euler's Equation
This equation describes the vector that we saw in Figure 1-5 with an
angle = θ and a magnitude = 1. The symbol i is the imaginary unit .
Equation 1-5 expresses a complex exponential function of an imaginary
number (iθ) in terms of the sine and cosine functions of the angle θ.
Now consider the Euler's equation for the complex signal eiωt (using ωt
instead of θ):
This formula has a “real” part and an “imaginary” part. The real part
(cos ωt) is what we are interested in because it corresponds to the
measured signal. We use the imaginary part (sin ωt) because it makes
the mathematics simpler (believe it or not!). At the end, we ignore the
imaginary part and keep the real component corresponding to the
actual signal.
Each value of (eiωt) is a vector that spins around at an angular
frequency of ω (Fig. 1-19). It is important to understand the concept of
angular frequency because later, when we talk about proton precession,
we'll see that the principle of precessional frequency (which is an
angular frequency) is used frequently.
Logarithms
The logarithm (log) is sort of the inverse of an exponential, that is, log
(ex) = x
The base of log is usually 10. However, a logarithm can have any base.
Let's say that the log of a number y with base a is equal to x: logay = x
Then, taking the exponential of both sides, we obtain ax = y Thus, logay
= x ⇔ ax = y That is, if the log of y with base a equals x, this implies
that a to the power of x equals y. Example log2 8 = 3 ⇒ 23 = (2)(2)(2)
= 8 The log of a number x to base e is also denoted as the ln of that
number, or “the natural log” of x. logex = lnx (ln) is just the notation
for log to the base e (e is the base of the natural logarithm, e =
2.71…). ln e = loge e = 1 This means log of e to its own base is 1.
Properties of Exponentials
Mathematical operations done with exponentials include (ex is e to the
power of x):
Properties of Logarithms
a. The logarithm of the product of two numbers (A and B) is equal to
the sum of their logarithms. log (A · B) = log A + log B This holds for
any base. So ln (A · B) = ln A + ln B
b. The log of x to a power a is log xa = a log x This also holds for any
base, so logb xa = a logb x ln xa = a ln x
Example
Let's solve the equation: 1/2 = e−t/T2 for t. (This would be the
formula for finding the time t when a function [with a first-order
decay constant T2] decays to half of its initial value.)
1. Take ln of both sides: ln (1/2) = ln e −t/T2
2. Remember that 1/2 = 2−1, so ln (1/2) = ln (2−1) = − ln 2
3. Now from Equation (b) ln (e−t/T2) = −t/T2 ln (e) so −ln 2 = −t/T2 ln (e)
4. Remember that ln (e) = logee e = 1 (log of a number to its own base =
1), so −ln 2 = −t/T2 (1) or −ln 2 = −t/T2
5. Multiply both sides by −1: ln 2 = t/T2 or loge 2 = t/T2 or 0.693 = t/T2
or t = 0.693
This formula also calculates the half-time (or t1/2) in nuclear medicine.
Key Points
Understanding a few mathematical concepts will help
you a great deal in understanding MRI physics.
1 Four sinusoidal functions were discussed: sin x, cos
x, tan x, cotan x tan x = sin x/cos x cotan x = 1/tan
x
2 The angle (or arc) on which the above sinusoidal
wave is applied is defined as arcsin, arccos, arctan,
arccotan For example, arctan(tan x) = x, arcsin(sin
x) = x, etc.
3 A vector possesses magnitude and direction. Force,
for example, has magnitude (weight) and direction.
Another example of a vector is velocity, which has a
speed and a direction.
4 Imaginary numbers are represented by a real and
an imaginary component: c = a + ib where i is the
imaginary unit .
5 A function f of a variable x is designated f(x) and
represents variations in f as x is varied.
6 A signal is a function of time.
7 A periodic signal is a function of time that repeats
itself after a certain period T.
8 The (linear) frequency of a periodic signal is
defined as f = 1/T, where T is the period.
9 The angular frequency ω is defined as ω = 2πf.
10 A periodic signal over one period represents one
cycle.
11 An example of a periodic signal is cos ωt = cos
(2πft).
12 Phase represents the offset between two periodic
signals of the same frequency. For example, cos (ωt)
and cos (ωt + θ) have the same frequency (ω) but
are out of phase by θ.
13 The function et is an exponential function of time.
It is actually an exponentially growing function. It is
1 at t = 0 and ∞ at t = ∞. At t = 1, e1 = e =
2.7182818 ≅ 2.72.
14 The function e−t is also an exponential function of
time. It is an exponentially decaying function. It is 1
at t = 0 and 0 at t = ∞. At t = 1, e−1 = 0.37.
15 The function e−t/τ is an exponentially decaying
function with a time constant or decay constant τ.
The value of the signal at time constant τ is 37% (e−1
= 0.37) of its previous value. The signal is practically
zero after five time constants (e−5 ≅ 0).
16 The function eiωt is given by Euler's equation as
eiωt = cos ωt + i sin ωt which represents a vector of
radius 1 spinning at an angular frequency ω (in
radians/sec).
17 The sinc function is defined as sinc t = sin t/t
which is 1 at t = 0. An ideal RF pulse is a sinc wave
because its Fourier transform (as we will see later)
has a perfect rectangular shape.
18 A logarithm (base 10) of a variable y is
represented as log y and is related to an exponential
as if log y = x then 10x = y
19 The natural logarithm (base e) of y is designated
ln y. Therefore if ln y = x then ex = y Having
understood the above mathematical concepts, the
reader can now read and understand the remainder
of this book with greater ease. As mentioned
previously, it is not that important to memorize the
formulas, but rather to understand the concepts
behind them.
Questions
1-1 Draw the following functions versus x (from − ∞ to + ∞)
(a) e−x cos x
(b) sin x/x ≡ sinc (x)
where f and g are functions of x and f' and g' are their derivatives,
that is, [ f' (x) = d/dx · f (x)]; “lim” denotes the “limit” as x
approaches 0.
Using the above fact, show that
Hint: d/dx (sin x) = cos x and d/dx (xn) = n · xn−1 (n = any integer)
Figure 1-20. A-B
1-4
(a) What is the value of an exponential function e−t/T at one
time constant T?
(b) How about at two time constants 2T?
(c) What is the ratio of (b) to (a)?
1-7 What is (a) sin (0°); (b) sin (30°); (c) sin (90°); (d) sin (180°); (e)
cos (0°); (f) cos (60°); (g) cos (90°); (h) cos (180°)?
2
Basic Principles of MRI
Introduction
In this chapter, we will discuss the basic principles behind the physics of
magnetic resonance imaging (MRI). Some of these principles are
explained using Newtonian physics, and some using quantum
mechanics, whichever can convey the message more clearly. Although
this might be confusing at times, it seems to be unavoidable. In any
case, we'll try to keep it straightforward.
Nuclear magnetic resonance (NMR) is a chemical analytical technique
that has been used for over 50 years. It is the basis for the imaging
technique we now call MRI. (The word nuclear had the false
connotation of the use of nuclear material; thus, it was discarded from
the MR lexicon, and “NMR tomography” was replaced by the phrase
magnetic resonance imaging [MRI].)
106 (1 10-12 (1 pm
1020 MeV) )
103 (1
Ultraviolet 1017 keV) 10-9 (1 nm )
100 (1
Infrared 1014 eV) 10-6 (1 µm)
Electromagnetic Waves
To understand MRI, we first need to understand what an
electromagnetic wave is. Table 2-1 demonstrates the characteristics of
a variety of electromagnetic waves, including X-ray, visible light,
microwaves, and radio waves. All electromagnetic waves have certain
fundamental properties in common:
Table 2-2
Wave
Frequency Energy Length
1.7-3.6 × 1018 30-150
X-ray Hz keV 80-400 pm
Visible light
(violet) 7.5 × 1014 Hz 3.1 eV 400 nm
Visible light
(red) 4.3 × 1014 Hz 1.8 eV 700 nm
20-200
MRI 3-100 MHz MeV 3-100 m
Table 2-3
The other thing that we know from quantum theory is that atomic
nuclei each have specific energy levels related to a property called
spin quantum number S. For example, the hydrogen nucleus (a single
proton) has a spin quantum number S of 1/2:
S (1H) = 1/2
The number of energy states of a nucleus is determined by the formula:
Number of energy states = 2S + 1
For a proton with a spin S = 1/2, we have
Number of energy states = 2 (1/2) + 1 = 1 + 1 = 2 Therefore, a hydrogen
proton has two energy states denoted as − 1/2 and + 1/2. This means
that the hydrogen protons are spinning about their axis and creating a
magnetic field. Some hydrogen protons spin the opposite way and have
a magnetic field in just the opposite direction. The pictorial
representation of the direction of proton spins in Figure 2-6 represents
the two energy states of the hydrogen proton. Each one of these
directions of spin has a different energy state.
Table 2-4
1H 1/2 42.6
Magnetic Susceptibility
All substances get magnetized to a degree when placed in a magnetic
field. However, the degree of magnetization varies. The magnetic
susceptibility of a substance (denoted by the Greek symbol χ) is a
measure of how magnetized they get. In other words, χ is the measure
of magnetizability of a substance.
To develop a mathematical relationship between the applied and
induced magnetic fields, we first need to address the confusing issue
regarding the differences between the two symbols encountered when
dealing with magnetic fields: B and H. We caution the reader that the
following discussion is merely a simplification; an advanced physics
textbook will have details on the theory of electromagnetism. The field
B is referred to as the magnetic induction field or magnetic flux
density, which is the net magnetic field effect caused by an external
magnetic field. The field H is referred to as the magnetic field
intensity. These two magnetic fields are related by the following:
B = µH or µ = B/H
where µ represents the magnetic permeability, which is the ability of a
substance to concentrate magnetic fields. The magnetic susceptibility χ
is defined as the ratio of the induced magnetic field (M) to the applied
magnetic field H:
M = χH or χ = M/H
Furthermore, χ and µ are related by the following:
µ = 1 + χ
making sure that the units used are consistent.
It requires the use of gradient coils, which will be discussed later in this
chapter.
Figure 2-11. In MRI, an RF wave, or an RF pulse, is transmitted
into the patient, and a signal is received from magnetized spins
(protons) in the body.
Types of Magnets
First of all, magnets can be categorized in terms of their field strength;
five types exist:
1. Ultrahigh field (4.0 to 7.0 T); mainly used for research
2. High field (1.5 to 3.0 T)
3. Midfield (0.5 to 1.4 T)
4. Low field (0.2 to 0.4 T)
5. Ultralow field (<0.2 T)
Next, we can categorize magnets in terms of their design; three main
types exist:
1. Permanent magnets
2. Resistive magnets
3. Superconducting magnets
1. Permanent magnets (mainly seen with OPEN MRI scanners such as
Hitachi AIRIS) always stay on and cannot be turned off. They have
the advantage of lower cost and lower maintenance (they require
no cryogens for cooling).
2. Resistive magnets (such as the 0.23 T Philips Panorama and the
Fonar 0.6 T Standup) are based on the electromagnetic principle
that electric current running through a coil produces a magnetic
field. These magnets can be turned off and on.
3. Superconducting magnets are a form of electromagnets. These
magnets operate near absolute zero temperature (e.g., 4.2 K or
−270°C). Consequently, there would be almost no resistance in
their wires. This, in turn, allows us to use very strong electric
currents to generate a high magnetic field without generating
significant heat (hence the name superconducting). To achieve
these ultralow temperatures, cryogens (such as liquid nitrogen
and/or liquid helium) are required (which are very expensive).
Most of the scanners available today are superconducting magnets.
have their own small magnetic fields and they are all spinning about
their own axes. Each one of the magnetic fields is called an MDM and is
denoted by the symbol µ. The axes of the MDMs are arranged in a
random way, and they all cancel each other out. If we add up all the
dipole moments, the net magnetic field will be zero (Fig. 2-13). This
result occurs in the absence of any external magnetic field (B0).
Figure 2-13. In the absence of an external magnetic field B0, net
magnetization is produced.
T1 Relaxation Time
The time constant of the curve in Figure 2-16 is denoted T1. Therefore,
the growth of magnetization M occurs with a time constant T1
Precession
When a proton is placed in a large magnetic field, it begins to “wobble”
or precess. When we take a single proton spinning about its axis, but
Each proton spins much faster about its own axis than it rotates or
precesses around the axis of the external magnetic field.
Figure 2-19. In the presence of an external magnetic field B0, a
proton not only rotates about its own axis but also “wobbles”
about the axis of B0.
Larmor Equation
The rate at which the proton precesses around the external magnetic
field is given by an equation called the Larmor equation:
ω = γB0 (Eqn. 2-1)
where ω is the angular precessional frequency of proton, γ the
gyromagnetic ratio, and B0 the strength of external magnetic field.
The angular frequency ω can be expressed in Hertz (Hz) or radians per
second, depending on the units used for γ. If γ is in terms of MHz/T,
then ω (or, actually, the linear frequency f) is expressed in terms of
MHz. The gyromagnetic ratio γ is a proportionality constant that is fixed
for the nucleus with which we're dealing. For hydrogen protons, γ(H) =
42.6 MHz/T.
Example
If the magnetic field strength is 1 T, the precessional frequency of
hydrogen is
(42.6)(1) = 42.6 MHz
As the external magnetic field strength increases, the precessional
frequency of the hydrogen proton also increases, that is, at 1.5 T
the precessional frequency of hydrogen is
(42.6)(1.5) ≅ 64 MHz
Remember that MRI involves the RF portion of the electromagnetic
spectrum, in the range of 3 to 100 MHz. This range is caused by the
precessional frequency ranges of the hydrogen protons for the
magnetic field strengths we use clinically, that is,
for B0 from 0.2T → 3T, ω = 8.5 MHz → 128 MHz
Coils
A coil is an electrical device generally composed of multiple loops of
wire (Fig. 2-20) that can either generate a magnetic field (gradient
coil) or detect a changing (oscillating) magnetic field as an electric
current induced in the wire (RF coil). Several types of coils are used in
MRI, including the following:
1. Gradient coils
a. Imaging gradient coil
b. Shim coil
Figure 2-20. A: “Ideal” loops of coils positioned perpendicular to
changing transverse magnetization rotating in x-y plane. This
changing magnetization induces a voltage in the coils, which
causes current to flow (arrows). B and C: Loops of coils gradually
molded to fit inside bore of magnet while retaining sensitivity to
changing transverse magnetization.
Figure 2-21. The gradient coils exist in MRI, one along each
direction (x, y, and z).
1. The slice-select gradient
2. The phase-encoding gradient
3. The frequency-encoding or readout gradient
For an axial image, these would correspond to Gz, Gy, and Gx,
respectively. They are discussed at length in chapters to come.
Shim Coils. These coils are used to create a more uniform external
magnetic field B0. Keep in mind that inhomogeneities in the external
field are undesirable and can cause artifacts, especially when a
gradient-echo or a chemical fat suppression technique is used. Shim
coils help to minimize (although not totally eliminate) such variations.
a factor of .
Solenoid Coils. These coils can be wrapped around the patient and
increase SNR. These coils are usually used in lower field magnets (e.g.,
open scanners), which have a vertical magnetic field orientation (rather
than a horizontal orientation in higher field scanners).
Phased-Array Coils. These coils contain multiple small surface coils
that are positioned on either side of the anatomy of interest. These
coils allow faster scanning with finer details. An example is the pelvic
array coil that allows exquisite visualization of pelvic structures.
Plane of Imaging
Selection of the gradient coils along x-, y-, or z-axis is arbitrary.
Imaging in different (axial, sagittal, coronal, or oblique) planes in MRI is
different from CT and is possible simply by appropriate assignment of
the gradient coils while the patient is always positioned in the magnet
with his or her long axis along the long axis of the scanner (Fig. 2-22).
For instance, assuming the z-axis to be along the long axis of the
magnet in the craniocaudal (CC) direction, the y-axis to be in the
posteroanterior (PA) direction, and the x-axis to be from right to left
Table 2-5
Phase- Frequency-
Slice-Select Encoding Encoding
Gradient Gradient Gradient
Axial z y x
Sagittal x y z
Coronal y x z
Key Points
In this chapter, we have discussed the basic
principles behind MR. We have talked about
electromagnetic waves, proton spins, external
magnetic fields, and longitudinal magnetization. We
briefly introduced the parameter T1 that is an
inherent property of a tissue. Let's summarize:
1 Electromagnetic waves (as the name applies) have
two components: an electric component (E) and a
magnetic component (H or B). These two
components are perpendicular to each other and 90°
out of phase.
2 The propagation component C is perpendicular to
both the E and B components, all traveling at the
speed of light (c = 3 × 108 m/sec).
3 In MRI, it is the magnetic component that interests
us. The electric component merely generates heat.
4 Electromagnetic waves are periodic functions of
time, oscillating at a frequency of
ω = 2πf
where ω is the angular frequency (in radians per
second) and f is the linear frequency (in cycles per
second or Hz).
5 Many types of electromagnetic waves exist
throughout the electromagnetic spectrum: X-rays,
visible light, microwaves, radiofrequencies, and so
on.
6 The frequencies used in MRI fall in the RF range (3
to 100 MHz). They are therefore called RF pulses.
7 Spinning charged particles generate an
electromagnetic field.
8 An example of the previous point in the body is the
hydrogen proton (1H).
9 The magnetic components of hydrogen protons
behave like bar magnets. This behavior is referred to
as a magnetic dipole moment (MDM).
10 In general, all particles with an odd number of
electrons in their covalent orbit have this property
(i.e., they can generate a magnetic field).
11 Although many different protons exist in the
body, hydrogen protons are dealt with in MRI
because they are the most prevalent in the body
(particularly in H2O, which comprises 60% of the
body).
12 The main magnetic field in MRI is denoted B0.
13 When a patient is placed in a magnetic field B0,
some of the protons are aligned parallel to B0 and
some antiparallel to it, but more are parallel than
antiparallel, producing a net magnetization
(longitudinal magnetization).
14 These protons also oscillate or precess about the
axis of the external magnetic field.
15 The frequency of precession of protons is
described by the Larmor equation ω0 = γB0 where γ
is the gyromagnetic ratio (in MHz/T). Therefore, the
stronger the magnetic field, the faster the protons
precess about it.
16 Magnetic susceptibility refers to the ability of a
substance to get magnetized when placed in a
magnetic field.
P.30
17 Three types of substances with different magnetic
susceptibility effects were discussed: diamagnetic,
paramagnetic, and ferromagnetic.
18 There are five types of magnets based on field
strength: ultralow field, low field, midfield, high field,
and ultrahigh field.
19 There are three types of magnets based on
design: permanent magnets, resistive magnets, and
superconducting magnets.
20 Most existing scanners are high-field,
superconducting magnets (these require liquid
cryogens like liquid helium and nitrogen for cooling).
21 Most “open”-type MR scanners are permanent or
resistive magnets; they require no cryogens and thus
have low maintenance. However, they usually have
lower field strengths and thus generate less signal.
22 To create an image, RF pulses are transmitted
into the patient. These pulses flip the longitudinal
magnetization and generate a signal from the
patient.
23 RF pulses flip the longitudinal magnetization Mz
away from the z-axis: 90° pulses flip Mz by 90°; 180°
pulses by 180°; and partial RF pulses by α, which is
less than 90°.
24 The received signal has no spatial information.
Three types of gradient coils (sliceselect, readout or
frequency-encoding, and phase-encoding gradients)
are employed for the purpose of spatial
discrimination.
25 Different types of coils are used: body coil, head
coil, and surface coil.
26 Surface coils are used for smaller body parts
(e.g., joints) to increase the signal and reduce the
noise (thus increasing the signal-to-noise ratio).
27 The rate at which the longitudinal magnetization
recovers from the transverse plane (after having
been flipped by a 90° pulse) is given by the time
parameter T1. This parameter also describes the rate
at which protons are magnetized when they are
placed in an external magnetic field.
28 The equation for this recovery at any time t is
given by
1 − e−t/T1
which is an exponential growth curve.
What do we need to create an image from the
information received from the patient? This process is
initiated by the use of an RF pulse and is discussed in
the next chapter.
Questions
2-1 Calculate the Larmor frequency of a proton at the following
magnetic field strengths:
(a) 0.35 T
(b) 0.5 T
(c) 1 T
(d) 1.5 T
(e) 2 T
(f) 3 T
(the gyromagnetic ratio, γ, of a proton ≅ 42.6 MHz/T).
2-2 T/F The rate at which protons are magnetized when placed in a
magnetic environment is the same as the rate of recovery of
longitudinal magnetization.
2-3 T/F Proton density represents the density of all the protons in
the tissue.
2-4 T/F When placed in a magnetic field, protons will line up with
that field immediately.
2-8 T/F The rate at which protons precess about the main magnetic
field is faster than that about their own axes.
2-9 T/F When placed in a magnetic field, all the protons in the body
will line up with the field.
Avogadro's law, there are over 1023 molecules per gram of tissue. Thus,
in each gram of tissue, there will be 1017 (i.e., 1023/106) excess
hydrogen protons pointing north.
2This is in fact the convention used throughout this book for simplicity.
3
Radio Frequency Pulse
Introduction
In the last chapter, we discussed the concept of longitudinal
magnetization. However, we have not yet addressed the issue of
receiving a signal from the patient. We can only transmit and receive
signals that oscillate (like an AC voltage). In addition, we're only
sensitive to oscillations along certain axes. Because the longitudinal
magnetization is not an oscillating function (like a DC voltage), it
cannot be read by a receiver. In addition, we're not sensitive to
oscillations along the z-axis. Consequently, this magnetization needs to
be “flipped” into the transverse x-y plane (where it can oscillate or
“precess” about the z-axis) to generate a readable signal. This is the
purpose of the radio frequency (RF) pulse.
RF Pulse
Suppose that a patient is in the magnet. Then we transmit an RF pulse.
What happens? Remember that an RF pulse is an electromagnetic
wave. Initially, all the spins are lined up along the axis of the external
magnetic field B0 about which they are precessing (Fig. 3-1). Then we
transmit an RF pulse. In a three-dimensional (x, y, z) coordinate
system, the direction of the external magnetic field always points in
the z direction. Thus, the net magnetization vector M0 will also point in
the z direction (Fig. 3-2).
Figure 3-1. An RF pulse is transmitted after the protons have been
exposed to the external magnetic field B0.
At what rate will these protons precess around this new magnetic field?
The new precessional frequency will be
where B1 is the weaker magnetic field associated with the RF pulse.
magnetic field B1 and not with the larger magnetic field B0.
Another point of clarification: remember that before the RF pulse, the
protons precess about the z-axis but they are out of phase and hence
have no net transverse component. After the RF pulse, the protons are
introduced to a new magnetic field B1 (also oscillating at frequency
ω0). Consequently, they will also tend to line up with the new magnetic
field and will then be in phase. This, in effect, creates transverse
magnetization. As more and more protons line up, phase “coherence”
increases, as does transverse magnetization. Simultaneously, as
previously discussed, the B1 field also causes a spiral downward motion
of the protons. These two factors explain the process of flipping.
Going back to the three-dimensional coordinate system (Fig. 3-4), the
vector M0 (the net magnetization in the direction of the protons aligned
along the external magnetic field) begins to precess about the x-axis in
the z-y plane. Depending on the strength of the RF pulse B1, and its
duration τ, we can determine the flip angle (i.e., the fractional angle
of a single precession):
Figure 3-4. A certain amount of time after the application of the
RF pulse, the magnetization vector is partially “flipped” toward
the x-y plane, forming an angle θ with the z-axis.
precession around the z-axis that would slowly spiral down into the x-y
plane (Fig. 3-6). This nutational motion is the result of the two
precessional motions happening simultaneously.
Figure 3-5. An outside observer looking at the coordinate system
sees rapid precession of protons and B1 about the z-axis.
Figure 3-6. The observer outside the coordinate system sees a
spiral motion of the magnetization vector toward the x-y plane.
The protons that are aligned with the external magnetic field are in
two energy states (Fig. 3-9). Those in the lower energy state (E1) are
lined up with (i.e., parallel to) the magnetic field B0, and those in the
higher energy state (E2) are aligned in the opposite direction. After a
90° RF pulse, some of the protons from the lower energy state are
boosted to the higher energy state. This happens only at the Larmor
frequency.
MATH: The Larmor equation can be derived from the following
principle. The energy difference between E1 and E2, denoted δE, is
given by
and thus,
or, alternatively,
f = (2µ/h)B0
With
γ = 2µ/h (in Hz/T)
where
At equilibrium after the protons are placed in the magnetic field, the
number of protons in the low energy state (north-pointing) is greater
than the number in the high energy state (south-pointing), resulting in
the longitudinal magnetization vector M0 (Fig. 3-10A). As energy is
added by the RF pulse to flip the north-pointing protons to the higher
energy state, the number of protons in both states can be equalized.
When this occurs, a measurable longitudinal magnetization vector no
longer exists. In addition, the RF pulse causes the spins to begin
precessing in phase with each other. The vector sum of the in-phase
northand south-pointing precessing protons lies in the transverse plane
(Fig. 3-10B). This transverse magnetization precesses at the Larmor
frequency.
The angular frequency at which the protons rotate 90° about the x-axis
is given by the Larmor equation: ω1 = γB1 where, again, B1 is the
magnetic field associated with the RF pulse. As stated previously, the
phase, that is, the number of degrees of
This equation shows that if we keep the RF pulse on for time τπ/2, the
magnetization vector is flipped 90°.
180° Pulse
A 180° pulse has twice the power (or twice the duration) of a 90° pulse,
as shown in Equation 3-5. After a 180° RF pulse, the longitudinal
magnetization vector is inverted, and the spins begin to recover from
−M0. After a 180° RF pulse, the excess north-pointing spins are boosted
from the low energy state to the high energy state.
Partial Flip
In the case of a partial flip (<90°), the component of magnetization
ending up in the x-y plane (i.e., Mxy) is less than the magnitude of the
original magnetization vector M0 (Fig. 3-11).
In fact, Mxy = M0 sin θ A partial flip is achieved by decreasing either the
strength or the duration of the RF pulse, according to Equation 3-5.
Such flip angles are common in gradient echo (GRE) imaging, as will be
discussed in later chapters.
Key Points
1 An RF (radio frequency) pulse is a brief
electromagnetic burst with frequencies in the radio
frequency spectrum.
2 Like all electromagnetic waves, an RF pulse has
associated magnetic and electric fields. We are
interested in the magnetic component B1. (The
electric component causes tissue heating.)
3 The purpose of the RF pulse in MRI is to flip the
longitudinal magnetization.
4 This flip is done by first causing the protons to
precess in phase about the axis of the external field
(B0), as well causing them to precess about the axis
of the RF field (B1). The result is a spiral motion of
spins toward the x-y plane, called nutation.
5 The flip angle is a function of the RF strength (B1)
and its duration (τ) and could be 180°, 90°, or a
fraction thereof (i.e., a partial flip 90°), depending on
the clinical application.
Questions
3-1 T/F The flip angle is determined by the duration of the RF pulse
and its power.
3-3 T/F The immediate action of the RF pulse is to cause the protons
to precess in phase.
3-5 T/F An RF pulse stands for radio frequency pulse, which is a form
of an electromagnetic wave with frequencies in the radio frequency
spectrum.
3-6 T/F A 180° pulse has 10 times the power of a 90° pulse.
Introduction
We have already introduced relaxation times T1, T2, and T2*. In this
chapter, we discuss the physical properties behind them and see what
conditions cause them to be increased or decreased. As we have
discussed before, T1 and T2 are inherent properties of tissues and are
thus fixed for a specific tissue (at a given magnetic field strength). The
parameter T2*, however, also depends on inhomogeneities in the main
magnetic field, but again is fixed for a specific tissue within a given
external magnetic environment.
T1 Relaxation Time
The term relaxation means that the spins are relaxing back into their
lowest energy state or back to the equilibrium state. (Equilibrium by
definition is the lowest energy state possible.) Once the radio
frequency (RF) pulse is turned off, the protons will have to realign
with the axis of the B0 magnetic field and give up all their excess
energy.
T1 is called the longitudinal relaxation time because it refers to the
time it takes for the spins to realign along the longitudinal (z)-axis. T1
is also called the spin-lattice relaxation time because it refers to the
time it takes for the spins to give the energy they obtained from the RF
pulse back to the surrounding lattice in order to go back to their
equilibrium state.
Realize that the recovery of magnetization along the z-axis and the
decay of magnetization within the x-y plane are two independent
processes occurring at two different rates (Fig. 4-5). Take a simple
exponential process. We would expect the rate at which this process
decays in the x-y plane to be the same as that at which it grows along
the z-axis (Fig. 4-6). This is not the same in the MR system we are
discussing because this system involves a much more complicated
process. T2 decay occurs 5 to 10 times more rapidly than T1 recovery
(Fig. 4-7). To understand this, we need to understand the concept of
dephasing.
Dephasing. After the 90° RF pulse is turned off, all spins are in phase;
they are all lined up in the same direction and spinning at the same
frequency ω0. There are two phenomena that will make the spins get
out of phase: interactions between spins and external field
inhomogeneities.
Figure 4-4. The graph of transverse magnetization with the decay
rate of T2 and RF off.
Figure 4-5. The recovery of longitudinal magnetization and the
decay of transverse magnetization occur at the same time but are
independent of each other with the RF off.
Figure 4-8. Two adjacent protons, one aligned with the field and
the other against it.
If we wait long enough, the three protons in the x-y plane will get
completely out of phase. The net magnetic field within the x-y plane
will then go to 0.
Therefore, at time t = 0, all spins are in phase, and their vector sum
will be at maximum magnitude. As the spins begin getting out of phase
with each other, their summation vector will become smaller and
smaller. When all the spins are completely out of phase with each
other, their vector sum will be zero. The effect of spin-spin interaction
depends to a degree on the proximity of the spins to each other. For
example, in water (H2O), the protons are separated more widely than
they are in a solid tissue. Hence, the dephasing effect of spin-spin
interaction might not be as prominent in H2O as it is in a solid tissue.
The term 1/T is the relaxation rate with units of sec−1 (recall that 1/T
is a frequency).
The relaxation rate (1/T2*) depends on the relaxation rate of the tissue
(1/T2) plus the magnetic field inhomogeneity of the external magnet. If
we have a perfect magnet that does not introduce any inhomogeneity,
then δB = 0 and T2* = T2. The newer systems have less magnetic field
inhomogeneity, thus making the T2* effects less strong; however,
complete homogeneity is not possible. Hence, there will always be
some T2* effect.
Figure 4-13. T2 and T2* decay curves.
Key Points
1 The rate of recovery of the longitudinal
magnetization is given by T1.
2 The rate of decay of the transverse magnetization
is given by T2.
3 The rate of decay of the FID is given by T2*.
4 T1 is 5 to 10 times greater than T2.
5 T2* is always less than T2.
6 T2 is a result of spin-spin interactions (internal
tissue inhomogeneities), whereas T2* is dependent
on both internal and external (main magnetic field)
inhomogeneities.
7 FID is produced by a rotating magnetic field, which
induces an electric current in a stationary coil.
Questions
4-1 True/false questions:
(a) T2* depends on the external magnetic field inhomogeneity.
(b) T2 depends on the external magnetic field inhomogeneity.
(c) T2 depends on T2*.
(d) T2* depends on T2.
1T2 also depends on diffusion (i.e., how rapidly spins spread out and
Introduction
In previous chapters, we discussed the roles of T1 and T2, longitudinal
and transverse magnetization, and the radio frequency (RF) pulse.
Obviously, by doing the procedures described in the previous chapters
only once, we won't be able to create an image. To get any sort of
spatial information, the process must be repeated multiple times, as
we shall see shortly. This is where TR and TE come into play. The
parameters TR and TE are related intimately to the tissue parameters
T1 and T2, respectively. However, unlike T1 and T2, which are inherent
properties of the tissue and therefore fixed, TR and TE can be
controlled and adjusted by the operator. In fact, as we shall see later,
by appropriate setting of TR and TE, we can put more “weight” on T1
or T2, depending on the type of clinical application.
How do we actually measure a signal? With the patient in a large
magnetic field (Fig. 5-1A), we apply a 90° RF pulse, and the
magnetization vector flips into the x-y plane (Fig. 5-1B). Then, we turn
off the 90° RF pulse, and the magnetization vector begins to grow in
the z direction and decay in the x-y plane (Fig. 5-1C). By convention,
we apply the RF pulse in the x direction, and for that reason, in a
rotating frame of reference, the vector ends up along the y-axis (Fig. 5-
1B).
Figure 5-1. A: Longitudinal magnetization before the RF pulse. B:
Immediately after the RF pulse, the magnetization vector is flipped
into the x-y plane. C: After a certain time period, Mz has recovered
by a certain amount while Mxy has decayed by a different amount.
cannot make an image from one signal. (An exception to this statement
is echoplanar imaging [EPI], which is performed after one RF pulse—see
Chapter 22.)
Figure 5-5. The time interval between two successive 90° RF
pulses is denoted TR.
Received Signal
Let's now take a look at the signal we are receiving (S). Because we are
only applying a series of 90° pulses, the signal will be a series of FIDs:
1. At time t = 0, the initial signal will be a strong FID similar to that
shown in Figure 5-7A.
2. At time t = TR, the signal will be slightly less in magnitude but will
also be an FID (Fig. 5-7B).
3. At time t = 2TR, the signal will be equal in magnitude to that in
Figure 5-7B (Fig. 5-7C).
Because the T1 recovery curve is given by the formula 1 − e−t/T1, if we
could measure the signal immediately after the RF pulse is given with
no delay, then each FID signal would be proportional to 1 − e−TR/T1
(This cannot really happen in practice.) Up to now, the signal S is given
by the formula S ∝ 1 − e−TR/T1
Remember that the word “signal” is really a relative term. The signal
that we get is a number without dimension, that is, it has no units. If
we are dealing with a tissue that has many mobile protons, then,
regardless of what the TR and T1 of the tissue are, we'll get more
signals with more mobile protons (see Chapter 2). Thus, when
considering the signal, we must also consider the number of mobile
protons N(H).
For a given tissue, the T1 and the proton density are constant, and the
signal received will be according to the above formula. If we measure
the FID at time TR immediately after the application of the second 90°
RF pulse, it will measure maximal and be equal to N(H) (1 − e−TR/T1).
Therefore, the FIDs that are acquired at TR intervals (i.e., 1TR, 2TR)
are maximal if they can be measured right after the 90° pulse, that is,
right at the beginning of the FID. However, in reality, we have to wait a
certain period until the system electronics allows us to make a
measurement.
Now we have to put the two curves together because both T1 recovery
and T2 decay processes are occurring simultaneously (Fig. 5-9).
Figure 5-9. The recovery and decay curves plotted on the same
graph.
Let's go back to the T1 recovery curve. After the 90° RF pulse, the spins
are flipped into the x-y plane. After a time interval TR, the amount of
received longitudinal magnetization is
Because the T1s of tissue A and tissue B are different, the short TR
brings out the difference in contrast between tissue A and tissue B.
Thus, for short TR, the two tissues can be differentiated on the basis of
different T1s. In other words, we get T1 tissue contrast with short TR.
Let's pick two different TEs (Fig. 5-11B). Here we have two TEs:
1. Short TE = TE1
2. Long TE = TE2
Question: Which one of the TEs in Figure 5-11B results in more tissue
contrast between tissue A and tissue B?
Answer: TE2 gives us more contrast.
Let's again look at the formula for signal intensity (Equation 5-4): SI =
N(H)(e−TE/T2*)(1 − e−TR/T1)
If TE is very short (close to zero), then e −TE/T2* approaches 1. TE →0 →
e−TE/T2*→e0 = 1 Then signal intensity = N(H)(1)(1 − e−TR/T1) = N(H)(1 − e
−TR/T1)
This means that, with a very short TE, we get rid of the T2* effect in
the equation. Therefore, we eliminate (or, again, in reality, reduce) the
T2* effect by having a very short TE.
Key Points
1. Long TR reduces T1 effect.
2. Short TR enhances T1 effect.
3. Short TE reduces T2* (T2) effect.
4. Long TE enhances T2* (T2) effect.
Questions
5-1 In the graph in Figure 5-12, the T1 and T2 curves are plotted
simultaneously for convenience. Assume the following values for T1
and T2: for white matter (WM): T1 = 500, T2 = 100 msec for
cerebrospinal fluid (CSF): T1 = 2000, T2 = 200 msec Also assume a
spin density N = 100 for both WM and CSF.
(a) For a TR = 2000 msec, find the relative signal intensities for
WM and CSF (i.e., points A and B on the graph).
(b) Calculate the crossover TE where WM and CSF have identical
T2 weighting (point C).
(c) Now, calculate the signal intensities of WM and CSF for TE =
25 (first echo) and TE = 100 msec (second echo), and the ratio
CSF/WM.
(d) Repeat (a) to (c) for TR = 3000 msec and observe how one
gets more T2 weighting in the second echo (higher ratio
CSF/WM).
(e) Now, calculate the signal intensities for TR = 3000 and TE =
200 msec.
Notice that despite relative loss of signal for both WM and CSF,
the ratio CSF/WM actually increases, indicating more T2
weighting (i.e., CSF gets brighter on the images).
5-2 Suppose that at 1.0 T, the approximate T1 and T2 values for the
following tissues are as follows.
5-3 A longer TR
(a) increases T1 weighting
(b) reduces T1 weighting
(c) increases T2 weighting
(d) reduces T2 weighting
5-4 A longer TE
(a) increases T1 weighting
(b) reduces T1 weighting
(c) increases T2 weighting
(d) reduces T2 weighting
5-5 Calculate the signal N(H)(1 − e−TR/T1) e−TE/T2 for the following
theoretical situations:
(a) TR = ∞
(b) TE = 0
(c) TR = ∞ and TE = 0
Introduction
In the previous chapter, we talked about T1 and T2 weighting in terms
of the time parameters TR and TE. Now let's discuss the T1 and T2
characteristics of the following tissues and see what physical properties
affect them:
1. H2O
2. Solids
3. Fat
4. Proteinaceous material
T2 Characteristics
The T2 characteristics of a tissue are determined by how fast the
proton spins in that tissue dephase. If they dephase rapidly, we get a
short T2. If they dephase more slowly, we get a longer T2.
H2O. Because of the structure of the water molecule (H-O-H) and
because of the sparsity of these molecules, spin-spin interaction among
the hydrogen protons is minimal. Therefore, dephasing occurs at a
much slower rate in water compared with other tissues. The T2
relaxation time for H2O is, therefore, long. Remember that T2 decay is
caused either by external magnetic field inhomogeneities or by spin-
spin interactions within or between molecules. In H2O, the effect of
one hydrogen proton on another is relatively small. The distance
between hydrogen protons both within each molecule and between
adjacent molecules is relatively large, so there is little spin-spin
interaction and, therefore, less dephasing.
Solids. The molecular structure of solids is opposite to that of pure
water. It is a very compactly structured tissue, with many interactions
between hydrogen protons. This large number of spin-spin interactions
results in more dephasing. Thus, the T2 for solids is short.
Fat and Proteinaceous Material. The structure of these materials is
such that there is less dephasing than in solids but more dephasing than
in water. Therefore, T2 for proteinaceous material or fat is
intermediate.
T1 Characteristics
The T1 of a tissue has to do with the way the protons are able to give
off their energy to the surrounding lattice, or to absorb the energy from
the lattice. It turns out that the most efficient energy transfer occurs
when the natural motional frequencies1 of the protons are at the
Larmor frequency (ω0). Recall that the Larmor frequency is
proportional to the strength of the magnetic field: ω0 = γBo
ω (H2O) ω0
Solids. Hydrogen protons in solids have lower natural motional
frequencies than do water protons. The natural motional frequencies of
hydrogen protons in solids are somewhat slower than the Larmor
frequency for hydrogen.
ω (solids) < ω0
Fat. Hydrogen protons in fat have natural motional frequencies that are
almost equal to the Larmor frequencies used for MRI.
ω (fat) ≈ ω0
This result is caused by the rotational frequency of the carbons around
the terminal C-C bond. Because this frequency is near the Larmor
frequency, the efficiency of energy transfer from the protons to the
lattice or from the lattice to the protons is increased, thus decreasing
T1.
Proteinaceous Solutions. The foregoing discussion on the T1 and T2
characteristics of fluids such as water applies only to pure water (or
bulk phase water). However, most of the water in the body is not in
the pure state but is bound to a hydrophilic macromolecule such as a
protein.
Such water molecules form hydration layers around the macromolecule
and are called hydration layer water (Fig. 6-1). These bound H2O
molecules lose some of the freedom in their
1. White matter is bright. The myelin sheath acts like fat; with more
efficient energy exchange, it has a shorter longitudinal relaxation
than does gray matter.
2. Gray matter is intermediate: without myelin, it acts more like a
typical solid tissue.
3. CSF is dark: like water, it has inefficient energy exchange and thus
the same long longitudinal relaxation, T1.
Let's add the T2 decay curves to the T1 recovery curves (Fig. 6-5):
1. CSF, like H2O, has the least dephasing, and thus the longest T2.
graphs of T1 and T2? Let's talk about two different tissues (Fig. 6-7):
1. CSF
2. White matter
CSF has a higher proton density than white matter, so it has a higher
maximum limit on the T1 recovery curve. White matter has a lower
proton density than CSF, but its T1 is shorter. The two recovery curves
cross at the point where white matter and CSF have the same intensity
(TR ≈ 2500 msec).
For the mathematically interested reader, this TR is
the solution to the following equation: 1.0 (1 − e-
TR/2650
) = 0.61 (1 − e-TR/510) or e-TR/2650 −0.61 e-TR/510
−0.39 = 0 using the T1 and N(H) values for white
matter and CSF from Table 6-1, resulting in a TR of
approximately 2500 msec (2462 msec to be exact!).
Figure 6-8. Recovery and decay curves for WM and CSF for a short
TR.
T1W: Short TR/Short TE
2. Now, draw the T1 and T2 curves again and this time pick a long TR.
Remember that CSF has a greater proton density than white matter, so
it will have a higher plateau value than white matter, which is brought
out by the long TR (Fig. 6-9). Then draw the T2 decay curves, keeping
in mind that CSF has a longer T2 than white matter. If we now pick a
very short TE (TE1), the two signals are driven by their respective
proton densities: CSF will have greater intensity than white matter
(i.e., 39%; Table 6-1). At this point, the difference in intensity reflects
their (true) proton density differences (assuming a very short TE).
Figure 6-9. Recovery and decay curves of WM and CSF for a long
TR.
before the crossover point for CSF and white matter T2 decay), then
edema has the highest signal intensity.
Figure 6-10. Recovery and decay curves of WM, CSF, and edema for
a long TR.
Fat and
Proteinaceous
H2O/Fluids Solids Material
T1 ω ω0 ω < ω0 ω ≈ ω0
Inefficient Inefficient
energy energy Efficient energy
transfer transfer transfer
Intermediate Muscle
GM a
(oxyHgb)
WM
a
a-d represent breakdown products of hemoglobin (a,
oxyhemoglobin; b, deoxyhemoglobin; c, intracellular
methemoglobin; d, extracellular methemoglobin; e,
hemosiderin). Abbreviations: GM , gray matter; WM , white
matter; SI, signal intensity; Hgb, hemoglobin; IC, intracellular;
EC, extracellular.
P.66
Figure 6-12. Axial T1 (A), proton density (B), and fast spin echo
T2 (C) images of the brain. Note that on the T1 image the white
matter is brighter than the CSF due to the shorter T1 of white
matter. However, the CSF is brighter on the proton density image
due to its higher proton density. The white matter becomes even
darker on the T2 compared with the CSF secondary to additional
T2 differences. The arrow points to a subarachnoid hemorrhage,
whereas the arrowhead points to a small subdural hemorrhage.
Note that the hemorrhage is isointense to CSF on the T2 due to a
crossover point being achieved, whereas the hemorrhage is
brighter on the T1 and on the proton density-weighted images
due to the shorter T1 versus CSF.
P.67
Figure 6-13. Axial T1 (A), proton density (B), and T2 (C) images
of the brain again show the relative signal differences in normal
structures between the three different sequences. The arrow
points to a small intraventricular meningioma that has typical
signal close to gray matter on all sequences.
P.68
Figure 6-14. Sagittal T1 (A), proton density (B), and T2 (C)
images of the lumbosacral spine show dark signal in the L5 and
S1 vertebral bodies on the T1 sequence due to the longer T1 of
edema compared with the shorter T1 of the fatty marrow at the
other levels. The CSF and the intervertebral discs are normally
dark on T1 due to fluid and/or dessication. The proton density
and T2 images show the L5/S1 disc and a few others to be
bright, as is the CSF. Note that the CSF is brightest on the T2.
The combination of both abnormal bone marrow signal and the
adjacent bright disc represent osteomyelitis and discitis,
whereas bright discs at other levels with adjacent normal
marrow represent hydrated discs.
P.69
Figure 6-15. Sagittal T1 (A), proton density (B), and T2 (C)
images of the lumbosacral spine show normally bright signal in
all the vertebral bodies through L4; however, the L5 vertebral
body and the sacrum have very bright signal on all sequences.
The proton density and T2 images were acquired with a fast spin
echo technique. This patient had radiation therapy for cervical
cancer with subsequent complete fatty replacement of the
marrow at L5 and the sacrum as opposed to the normal fatty-
containing marrow at other levels.
P.70
P.71
Figure 6-17. (continued)
P.72
Figure 6.19. Axial proton density (A) and T2 (B) images
demonstrate superficial gyriform dark signal from superficial
siderosis (hemosiderin) in a patient with a history of
subarachnoid hemorrhage (arrows).
P.73
Figure 6-20. Axial T1 with fat saturation without gadolinium (A)
and fast spin echo T2 (B) images show bright T1 and relatively
dark T2 signal in the lumen of the small bowel consistent with
proteinaceous solutions (arrows). Additionally, the patient has a
right endometrioma (arrowhead) with predominantly bright T1
and dark T2 signal secondary to recurrent hemorrhage.
Questions
6-1 T/F Hydration layer water has a shorter T1 than bulk water.
6-2 Match (i) short T1 and T2; (ii) short T1, long T2; (iii) long T1,
short T2; and (iv) long T1, long T2 with
(a) air
(b) fat
(c) water
(d) intracellular methemoglobin
(e) extracellular methemoglobin
6-3 T/F The most efficient energy transfer occurs at the Larmor
frequency.
6-4 Match the following: (i) short TR and TE; (ii) long TR and TE; (iii)
short TR, long TE; (iv) long TR, short TE with
(a) T1 weighted
(b) T2 weighted
(c) intermediate weighted
Introduction
A pulse sequence is a sequence of radio frequency (RF) pulses applied
repeatedly during an MR study. Embedded in it are the TR and TE time
parameters. It is related to a timing diagram or a pulse sequence
diagram (PSD), which is discussed in Chapter 14. In this chapter, we
discuss the concepts of saturation and consider pulse sequence partial
saturation, saturation recovery, and inversion recovery (IR). In the next
chapter, we'll talk about the important spin-echo pulse sequence.
Figure 7-1 illustrates the notations used for three types of RF pulses
throughout this book.
Saturation
Immediately after the longitudinal magnetization has been flipped into
the x-y plane by a 90° pulse, the system is said to be saturated.
Application of a second 90° pulse at this moment will elicit no signal
(like beating a dead horse). A few moments later, after some T1
recovery, the system is partially saturated. With complete T1 recovery
to the plateau value, the system is unsaturated or fully magnetized.
Should the longitudinal magnetization only be partially flipped into the
x-y plane (i.e., flip angles less than 90°), then there is still a
component of magnetization along the z-axis. The spins in this state are
also partially saturated.
Figure 7-1. The notation for 90°, 180°, and partial flip pulses.
Partial Saturation Pulse Sequence. Start with a 90° pulse, wait for a
short period TR, and then apply another 90° pulse. Keep repeating this
sequence. The measurements are obtained immediately after the 90°
RF pulse. Therefore, the signal received is a free induction decay (FID).
Let's see how this looks on the T1 recovery curve (Fig. 7-2). At time t =
0, flip the longitudinal magnetization 90° into the x-y plane. Right after
that, the longitudinal magnetization begins to recover. Wait a time t =
TR, and repeat the 90° pulse. Initially, at time t = 0, the longitudinal
magnetization is at a maximum. As soon as we flip it, the longitudinal
magnetization goes to zero and then immediately thereafter begins to
grow. At time t = TR, the longitudinal magnetization has grown but has
not recovered its plateau before it is flipped into the x-y plane again.
(Note that the length of the longitudinal magnetization vector before
the second 90° RF pulse is less than the original longitudinal
magnetization vector.)
Now, with a third 90° RF pulse, we again flip the longitudinal
magnetization into the x-y plane. Again, the longitudinal magnetization
goes to zero and immediately begins to recover.
Again, at time 2TR, it is less than maximum but is equal to the previous
longitudinal magnetization (at time TR). Each subsequent recovery time
TR after each subsequent 90° pulse will also be the same. Thus, the
maximum FID occurs at time t = 0 after the first 90° RF pulse, and all
subsequent FIDs will have less magnitude but will have the same value.
Figure 7-4. In inversion recovery, the time between the 180° pulse
and the 90° pulse is denoted TI.
Before we apply the 180° pulse, the magnetization vector points along
the z-axis. Immediately after we apply the 180° pulse, the
magnetization vector is flipped 180°; it is now pointing south (−z),
which is the opposite direction (Fig. 7-5).
We then allow the magnetization vector to recover along a T1 growth
curve. As it recovers, it gets smaller and smaller in the −z direction
until it goes to zero, and then starts growing in the +z direction,
ultimately recovering to the original longitudinal magnetization.
After a time TI, we apply a 90° pulse. This then flips the longitudinal
magnetization into the x-y plane. The amount of magnetization flipped
into the x-y plane will, of course, depend on the amount of longitudinal
magnetization that has recovered during time TI after the original 180°
RF pulse. We measure this flipped magnetization. Therefore, at this
point we get an FID proportional to the longitudinal magnetization
flipped into the x-y plane. Also, at this point, we begin the regrowth of
the longitudinal magnetization. Recall that for a typical T1 recovery
curve, the formula for the exponential growth of the curve is 1 − e−t/T1
However, when the magnetization starts to recover from −M0 instead of
zero (Fig. 7-6), the formula for recovery is 1 − 2e−t/T1
Exercise
Verify the above formula mathematically.
At time t = 0, Signal intensity (SI) = 1 − 2e−0/T1 = 1 − 2(1) = − 1 So at
time t = 0,
Magnitude of signal intensity = − 1.
At t = ∞ (infinity), Signal intensity = 1 − 2e−∞/T1 = 1 − 2(0) = + 1
So at time t = ∞, the signal is maximal. These values correspond to the
graph in Figure 7-6.
Null Point. The point at which the signal crosses the zero line is called
the null point. At this point, the signal intensity is zero. The time at
this null point is denoted TI(null). We can solve the equation
mathematically for TI(null), at which point the signal intensity is zero:
Signal intensity = 0 = 1 − 2e−T1/T1
The solution to this equation is (see Question 7-1 at the end): TI(null) =
(loge2) T1 = (ln 2) T1 ≈ 0.693 T1
Let's go back and re-examine the recovery curves. Actually, there are
two different exponentially growing curves occurring sequentially (Fig.
7-7):
1. Recovery after the 180° RF pulse
2. Recovery after the 90° RF pulse
1. The T1 recovery curve following the 180° pulse starts at −M0 and
grows exponentially according to the formula: M0 (1 − 2e−T1/T1)
Figure 7-6. The recovery curve in IR is given by the formula 1 −
2e−t/T1.
At this null point for fat, if we draw the T2 decay curves, fat starts at
zero and will stay at zero. There will be no transverse magnetization
from fat in the x-y plane, and water will have its usual T2 decay curve.
In effect, we have suppressed the fat signal. Therefore, after a 180°
inverting pulse, we wait a time TI = 0.693 T1 (fat) and we give the 90°
pulse. All other tissues will have longitudinal magnetization that will
flip into the x-y plane and give off a signal according to their T2 curves.
However, at its null point, fat will not have any longitudinal
magnetization to flip into the x-y plane and thus will not have any
signal.
The term STIR is called short TI inversion recovery because fat has a
very short T1; therefore, a very short TI must be chosen to null it (at
high field [1.5 T] this TI is 140 msec, whereas at midfield [0.5 T] it is
100 msec). Fat will reach its null point before white matter, gray
matter, H2O, or edema (Fig. 7-11).
Key Points
We have discussed three types of pulse sequences:
saturation recovery, partial saturation, and inversion
recovery (IR). The latter is very important because it
allows suppression of any tissue by selecting TI to be
0.693 times the T1 of that tissue: TI(null) = 0.693 ×
T1 This subject is further elaborated in Chapter 25 on
tissue suppression techniques.
A partial saturation sequence results in T1 weighting
(short TR and TE). A saturation recovery, however,
results in PD weighting (long TR, short TE).
Questions
7-1
(a) Given an inversion recovery (IR) pulse sequence (Fig. 7-6),
prove that TI that “nulls” or “suppresses” a certain tissue is
equal to 0.693 × T1 (tissue), that is, TI(null) = 0.693 × T1
Figure 7-12.
8
Pulse Sequences Part II (Spin Echo)
Introduction
This chapter focuses on the most frequently used pulse sequence—the
spin echo (SE) pulse sequence. When the concept of dephasing was
discussed in previous chapters, we brought up two main causes: (i)
external magnetic inhomogeneity, and (ii) inherent spin-spin
interactions. The SE pulse sequence eliminates the former by an
additional refocusing or rephasing 180° radio frequency (RF) pulse. By
using the SE pulse sequence, we can eliminate dephasing caused by
fixed external magnetic field inhomogeneities. (We can't eliminate
spin-spin interactions because they are not fixed, i.e., they fluctuate
randomly.)
Analogy
Let's consider the analogy of three runners running around the track
(Fig. 8-2). Initially, they start out at the same point. After they run for
a time τ, they are no longer together—one is running faster and gets
ahead of the others, and one is running slower, falling behind the
others.
At this time, if we make the runners turn around and run the opposite
way, each one will still be running at the same speed (precessing at the
same frequency in the case of the spins). They have just changed
direction and are running back to where they started. Each one will
then run the same distance if they run the same amount of time τ.
Therefore, at time 2τ, they will all come back at the starting point at
the same time and will be back together in phase. The action of making
the runners change direction is done with the use of a 180° refocusing
pulse in the case of the spins.
Thus, at a certain time τ after the 90° pulse, when the spins have
gotten out of phase, a 180° pulse is applied. Now all the spins flip 180°
in the x-y plane and they continue precessing, but now in the opposite
direction (Fig. 8-3). Let's look at the pulse sequence diagram (Fig. 8-4).
We start off with a 90° RF pulse to flip the spins into the x-y plane. We
wait a time τ and apply a 180° RF pulse. Then we wait a long time, TR,
and repeat the process.
If we draw the free induction decay (FID) after the 90δ pulse, we see
that the FID dephases
very rapidly due to the T2* effect related to external magnetic field
inhomogeneities and spin-spin interactions. The spins get out of phase.
After time τ we apply the 180δ refocusing pulse. After an equal time τ
they will be completely in phase again, and the signal will reach a
maximum.
Figure 8-1. Three magnetization vectors in three slightly magnetic
environments. In (A) they are in phase and their vector sum is
three times each individual vector. In (B) they are slightly out of
phase, yielding a smaller net vector. In (C) vector 1 and 3 cancel
each other out because they are 180δ out of phase, leaving only
vector 2.
1. Time τ is the time from the 90° RF pulse to the 180° RF pulse.
2. Time τ is also the time from the 180° RF pulse to the point of
maximum rephasing, that is, the echo.
3. We call 2τ the echo delay time (time to echo)—TE: the time after
the 90° pulse when we get maximum signal again.
4. The 180° pulse is, therefore, called a refocusing or rephasing pulse.
We can apply a second 180° pulse. Now, instead of one 180° pulse
following the 90° pulse, we have two 180° pulses in sequence after a
90° pulse (Fig. 8-5). After the first echo, the spins will begin to dephase
again. A second 180° pulse applied at time τ2 after the first echo will
allow the spins to rephase again at time 2τ2 after the first echo and a
second echo is obtained. Each echo has its own TE.
1. The time from the 90° pulse to the first echo is TE1.
2. The time from the 90° pulse to the second echo is TE2.
Ideally, we would like to regain all the signal from the original FID. In
practice, it can't happen. We are able to regain the signal lost due to
fixed external magnetic field inhomogeneities by applying a refocusing
180° pulse, but dephasing caused by spin-spin interaction of the tissue
cannot be regained. If we join the points of maximum signal due to
rephasing as a result of the 180° pulses, we will get an exponentially
decaying curve with a time constant given by T2. Therefore, the decay
of the original FID and the decay of each subsequent echo is given by e
−t/T2*, whereas the decay of the
Example
Take TR = 2000, and TE = 40 and 80 msec. Here, τ1 = 20, so that TE1 = 2
τ1 = 40, and TE2 = 80 = TE1 + 2τ2 = 40 + 2τ2; then 2 τ2 = 40 and τ2 = 20.
So τ1 = τ2 in symmetric echoes.
Asymmetric Echoes
If τ1 ≠ τ2, then we get asymmetric echoes.
Example
Take TR = 2000, TE = 30 and 80 msec.
Here TE1 = 2(τ1) = 30 msec, so τ1 = 15 msec.
TE2 = 80 msec = TE1 + 2 (τ2) = 30 msec + 2τ2 = 80 msec.
Then 2τ2 = 50 and τ2 = 25 msec.
So τ1 ≠ τ2 in asymmetric echoes.
Question: What does the 180° pulse do to the longitudinal
magnetization?
Answer: It inverts it. However, at time TE/2 (on the order of 10 msec),
the recovered longitudinal magnetization is negligible and its inversion
does not cause any significant signal loss. In fact, at t = TE/2, we have
Mz = M0 (1 - e-TE/2TR) ≅ 0
because TE/2 TR, so that e-TE/2TR ≅ 1.
Tissue Contrast
As discussed in Chapter 6, tissue contrast in SE depends primarily on TR
and TE. There are three types of tissue contrast:
1. T1 weighted (T1W)
2. T2 weighted (T2W)
3. Proton density weighted (PDW; also called “balanced,”
“intermediate,” and “spin density”)
Let's see what TR and TE must be for these three imaging scenarios
(Table 8-1):
1. For T1 weighting, we want to eliminate the T2 effect and enhance
the T1 effect.
a. To eliminate (reduce) the T2 effect, we want a short TE.
b. To enhance the T1 effect, we want a short TR.
c. The signal is then proportional to N(H) (1 - e-TR/T1).
2. For T2 weighting, we want to eliminate the T1 effect and enhance
the T2 effect.
a. To eliminate (reduce) the T1 effect, we want a long TR.
Table 8-1
TR TE Signal (Theoretical)
Key Points
1 The SE (spin echo) pulse sequence is composed of
a 90° excitation pulse followed by one or more 180°
rephasing pulses.
2 The purpose of the 180° pulse is to eliminate the
dephasing effects caused by external magnetic field
inhomogeneities by rephasing the spins at the time
of echo (TE).
P.89
3 The resultant echo then depends on T2 decay
rather than T2* decay, as seen with the FID (free
induction decay).
4 Table 8-2 summarizes the tissue contrast in SE
with respect to TR and TE.
Table 8-2
Short TE Long TE
8-2 Match
(i) T1W (ii) T2W (iii) PDW with
(a) short TR and short TE
(b) long TR and short TE
(c) long TR and long TE
8-3 T/F
The 180° pulses totally eliminate the dephasing of spins in the
transverse plane.
9
Fourier Transform
Introduction
Fourier was an 18th century French mathematician. His picture, along
with the Fourier transform of his picture, is shown in Figure 9-1. The
Fourier transform (FT) is a mystery to most radiologists. Although the
mathematics of FT is complex, its concept is easy to grasp. Basically,
the FT provides a frequency spectrum of a signal. It is sometimes easier
to work in the frequency domain and later convert back to the time
domain.
Let's start by saying that we have a signal g(t), with a certain waveform
(Fig. 9-2). This signal is basically a time function, that is, a waveform
that varies with time. Now, let's say we have a “black box” that
converts the signal into its frequency components. The conversion that
occurs in the “black box” is the Fourier transform. The FT converts the
signal from the time domain to the frequency domain (Fig. 9-2). The FT
of g(t) is denoted G(ω). (The frequency can be angular [ω] or linear
[f].)
The FT is a mathematical equation (you don't have to memorize it). It is
shown here to demonstrate that a relationship exists between the
signal in the time domain g(t) and its Fourier transform G(ω) in the
frequency domain:
where ω = 2πf.
We are already familiar with the term (e−iωt) from Chapter 1. This is
the term for a vector spinning with angular frequency ω. The formula
integrates the product of this periodic function and g(t) with respect to
time. It also provides another function, G(ω), in the frequency domain
(Fig. 9-2).
One interesting thing about the Fourier transform is that the Fourier
transform of the Fourier transform provides the original signal. If the
Fourier transform of g(t) is G(ω), then the Fourier transform of G(ω) is
g(t):
FT provides the range of frequencies that are in the signal. Here are
some examples of functions and their Fourier transforms:
Example 1
The cosine function: cos (ω0t).
Obviously this signal (Fig. 9-3) has one single frequency. The frequency
could be any number. The Fourier transform is a single spike
representing the single frequency in the frequency domain (because of
its symmetry, we also get a similar spike on the opposite side of zero1).
The Fourier transform in this case tells us that there is only a single
frequency because it shows only one frequency spike at ω0 and is zero
everywhere else on the line. We can, for simplicity, ignore the
symmetric spike
on the negative side of zero and just consider the single spike on the
positive side to tell us that there is a single frequency (ω0). The spike
represents frequency and amplitude of the cosine function.
Figure 9-1. A: Picture of the French mathematician Fourier. B: The
magnitude of Fourier's two-dimensional Fourier transform. C: The
phase of his Fourier transform. (Reprinted with permission from
Oppenheim AV. Signals and systems. Prentice Hall; 1983.)
Figure 9-2. The FT of g(t), designated G(ω).
Example 2
The sinc wave: sinc(ω0t) = sin(ω0t)/(ω0t)
The Fourier transform of this signal (Fig. 9-4) has a rectangular shape
and shows that the signal
Example 3
Let's consider two frequencies:
1. cos ωt and
2. cos 2ωt (which is twice as fast as cos ωt)
The signal cos(2ωt) oscillates twice as fast as cos(ωt) (Fig. 9-5). If we
add them up, we get a complex signal (Fig. 9-6). If we were just given
the signal in Figure 9-6, we would have no idea that it is the sum of two
cosine waves.
Question: What is a good way to figure out the frequencies of which the
signal is composed?
Figure 9-4. The FT of a sinc function (sinc ω0t = sin ω0t/ω0t) has a
rectangular shape. The two ends of this rectangle are at ω0 and −
ω0 (where ω0 is the frequency of the sinc function).
Answer: The Fourier transform of this complex signal (which we know is
the sum of two cosine waves, one twice as fast as the other) contains
two spikes, one twice as far from the origin as the other (Fig. 9-6). This
FT, then, demonstrates the composition of the signal in terms of its
frequencies.
Example 4
Let's now have a complex signal with two cosine waves, with the second
cosine wave not only twice as fast, but with twice the amplitude as
well. Again, by looking at the signal in Figure 9-7, we have no idea what
it is composed of, but by looking at the frequency spectrum (the FT of
the signal), we can tell the composition of the signal: in this case, two
separate cosine waves with differing frequencies and differing
amplitudes. The FT provides the frequency spectrum of a signal with its
amplitudes.
Example 5
Let's consider the FT of a sine wave: sin ωt. The FT of a sine wave is
different from the FT of a cosine wave:
1. FT of a cosine wave is symmetric with two symmetric spikes on either
side of zero (Fig. 9-8a).
Figure 9-5. The FT of cos 2ωt has two spikes: one at 2ω and one
at −2ω. (Here, only the positive frequencies are shown.)
Figure 9-6. The FT of cos ωt + cos 2ωt has two sets of spikes at
±ω and ±2ω. By looking at the FT signal, it is easy to figure out
its composition in the time domain.
Figure 9-7. The signal cos ωt + 2cos 2ωt and its FT. The FT has
again two sets of spikes: one at ±ω and one at ±2ω (but with
twice the magnitude).
Figure 9-8. A: FT of cos ωt. B: FT of sin ωt. Here the spikes have
opposite polarities and are also imaginary [iA in (B) as opposed to
(A)].
Figure 9-9. sin ωt is an example of an odd function where the
value of the signal at −t is the negative of its value at +t.
What does this mean? Let's say that we have the rectangular function
(Fig. 9-11a). We said that this is composed of an infinite number of sine
and cosine waves:
Key Points
The FT, as intimidating as it may appear, represents
a simple concept. Every signal (in the time domain)
is composed of a series of frequencies. The FT is a
way of representing that signal in terms of its
frequencies. The FT also allows mathematical
manipulations performed in the frequency domain,
which are sometimes easier than in the time domain.
The one-to-one relationship between a signal and its
FT allows reconstruction of the original signal from its
FT.
In other words, the FT represents a function in the
frequency domain whose amplitude varies with the
frequencies present in the signal. The bandwidth
(BW) is simply a measure of the range of frequencies
present in the signal (in Hz or in radians/sec).
Questions
9-1 T/F The FT of an FT equals the original signal.
9-2 T/F
(a) It is always easier to perform calculations in the frequency
domain.
(b) It is always easier to perform calculations in the time
domain.
amplitude (in the form of iA, where i is the imaginary unit and A is
the amplitude of the spike). This concept is rather important in
understanding the symmetry that exists in k-space, discussed in
Chapters 13 and 16.
10
Image Construction: Part I (Slice
Selection)
Introduction
The signals received from a patient contain information about the
entire part of the patient being imaged. They do not have any
particular spatial information. That is, we cannot determine the
specific origin point of each component of the signal. This is the
function of the gradients. One gradient is required in each of the x, y,
and z directions to obtain spatial information in that direction.
Depending on their function, these gradients are called
1. The slice-select gradient
2. The readout or frequency-encoding gradient
3. The phase-encoding gradient
Depending on their orientation axis they are called Gx, Gy, and Gz.
Depending on the slice orientation (axial, sagittal, or coronal), Gx, Gy,
and Gz can be used for slice select, readout, or phase encode.
A gradient is simply a magnetic field that changes from point to point—
usually in a linear fashion. We temporarily create magnetic field
nonuniformity in a linear manner along all three axes to obtain
information about position.
Figure 10-1. Selecting a slice of a certain thickness.
First, we'll consider the slice-select gradient, which is the easiest of all
to understand. Once a slice has been selected, we worry about the
problem of in-plane spatial encoding, that is, discriminating position
within the slice. As we'll see shortly, the principles behind slice
selection and spatial encoding in MRI are different from the principles
used in computerized tomography (CT).
signal, and we have no idea yet from where exactly in the body the
signal is coming.
The frequency of the RF pulse is given by the Larmor frequency:
ω0 = γB0
If we transmit an RF pulse that does not match the Larmor frequency
(the frequency of oscillation at magnetic field B0), we won't excite any
of the protons in the patient.
However, if we make the magnetic field vary from point to point, then
each position will have its own resonant frequency. We can make the
magnetic field slightly weaker in strength at the feet and gradually
increase in strength to a maximum at the head (Fig. 10-2). This effect
is achieved by using a gradient coil.
Let's say the magnetic field strength is 1.5 T at the center, 1.4 T at the
feet, and 1.6 T at the head. Then, the foot of the patient will
experience a weaker magnetic field than the head. Therefore, a
gradient in any direction (x, y, or z) is a variation in the field along that
axis in some fashion (the most common form of which is linearly
increasing or decreasing). If we now transmit an RF pulse of a single
frequency into the patient, we will receive signals corresponding to a
line in the patient at the level of the magnetic field corresponding to
that frequency (according to the Larmor frequency), but it will be an
infinitely thin line. What we need to do is transmit an RF pulse with a
range of frequencies—a bandwidth of frequencies.
Figure 10-2. Slice thickness is determined by the slope of the
gradient.
Figure 10-3. The relationship between field strength and Larmor
frequency in determining slice thickness and position.
What will the RF pulse look like in the frequency domain? First, note
that for a 1.5-T magnet, the Larmor frequency is 64 MHz for hydrogen
protons:
Slice Thickness
Let's now talk about how we establish slice thickness (Fig. 10-7). The
same principle would apply if we were to image from the base of the
skull to the vertex rather than from the patient's head to toe. We
establish a magnetic field strength gradient so that at the midpoint of
the field of study (in this instance, the entire body), the field strength
is 1.5 T; at the low end of the gradient (the foot), the field strength
will be 1.4 T; and at the high end of the gradient, the field strength will
be maximum at 1.6 T. These magnetic field strengths also correspond
to different frequencies. Using the Larmor equation, we can calculate
that approximately:
1.6 T ˜ 68 MHz
1.5 T ˜ 64 MHz
1.4 T ˜ 60 MHz
Review
RF Pulses. There are two types of RF pulses:
1. Nonselective
2. Selective
Figure 10-13. The sinc function and its FT. The bandwidth (BW) is
2fmax where fmax is the frequency of the sinc function.
Sinc RF Pulse
Earlier, we talked about one type of RF pulse in which we had a sinc
wave in the time domain. The sinc function is mathematically
expressed as
sinc (t) = sin (t)/t
This simply means that the oscillating function sin t is divided by t.
Therefore, because t goes into an oscillating function (sin t), the result
will be an oscillating function. As t goes from a large value to a small
value, the result of (sin t/t) will get larger and will reach maximum
when t approaches zero. The rectangular transform in the frequency
domain has a positive maximum frequency (fmax) and a negative
maximum frequency (-—fmax), as shown in Figure 10-13. The bandwidth
is thus 2 × (maximum frequency), that is,
BW = 2fmax
This is only one type of selective RF pulse. There are other types of
selective RF pulses.
Gaussian RF Pulse
The first generation of MR machines used an RF pulse that had a
Gaussian shape (Fig. 10-14a).
Going back to the sinc wave, in reality we can't go all the way to
infinity in time when transmitting a signal. We have to truncate the
signal and deal with a certain finite time domain (Fig. 10-16). The
Fourier transform of this truncated signal gives the “ripples” effect on
the square wave as we have seen in Chapter 9. The more we truncate
the signal, the more “ripples” we get. Ripples are also called
“overshoot” and “undershoot” artifacts.
Bandwidth. Bandwidth is a measure of the range of frequencies. We
know that for a 1.5-T magnet, the Larmor frequency is about 64 MHz.
The RF pulse that we generate is in the RF range, but its bandwidth is
in the audible frequency range.
This is analogous to a radio. When we tune into the FM station KSON
97.3 FM in San Diego, are we really getting 97.3 MHz of sound waves? If
the signal you receive on your radio
has this high a frequency, you can't hear it (we can't hear a MHz
frequency signal [maybe dolphins can, but humans can't!]). However,
the frequency that we receive is actually in the audible range. Each
station has a certain frequency range that they deal with, and the
bandwidth that they use is nearly the same for all radio stations in the
audible frequency range; it is about 1 to 2 kHz (Fig. 10-17).
Now, the bandwidth of 1 kHz is 100,000 times smaller than the 97.3
MHz frequency. The audible frequency range gets modulated to the
center frequency (e.g., FM 97.3 MHz). The modulated frequency is
then transmitted. The antenna receives this modulated signal. The
signal then goes into the radio and is demodulated (Fig. 10-18).
Depending on which station we tune into (e.g., KSON FM at 97.3 MHz),
we get that range of frequencies demodulated back to zero frequency.
Thus, the bandwidth of a signal transmitted by a radio station always
stays at about 1 kHz, but it is transmitted as a 1-kHz bandwidth
modulated to, say, 97.3 MHz. Then, in the radio, it is demodulated back
to zero center frequency, still with a bandwidth of about 1 kHz. What
we send and what we receive are in the same frequency range. In
between, the frequency gets modulated from 0 to 97.3 MHz. We are
just changing the center frequency. Everything else stays the same.
The reason radio transmission does this is that each radio station is only
allowed a narrow bandwidth (in the kHz range) for transmission.
However, we cannot transmit a kHz frequency. It just doesn't travel
very far, and radio transmissions have to travel for miles. Besides, the
signals from different stations will get all mixed up if they all work in a
small frequency range in the order of a few kilohertz. Therefore, the
kHz frequency bandwidth is modulated to the MHz range, still
maintaining the same kHz bandwidth. Now with the MHz frequency as a
carrier, the narrow 1-kHz bandwidth can be transmitted over long
distances.
In MRI, the center frequency is the Larmor frequency. Thus, the RF
pulse that we transmit into the patient is centered at the Larmor
frequency of 64 MHz at 1.5 T (Fig. 10-19). However, the bandwidth of
frequencies in the RF pulse is very narrow. Again, for simplicity, we will
assume that the bandwidth of the RF pulse has been demodulated to a
center frequency of zero.
Slice-Select Gradient. Let's go back to the slice-select gradient. We
purposely create a linear magnetic nonuniformity, so the foot will
experience a weaker magnetic field than the head. The slope of
magnetic field versus distance is called the gradient. A gradient is a
measure of change of magnetic field with distance. We can have a
linear gradient or a nonlinear gradient. The gradients that we use in
MRI are usually linear. (In fact, nonlinearities may be present in the
gradients that cause geometric distortion artifacts [see Chapter 18].)
By creating this gradient, different magnetic fields are experienced
from the foot to the head
Key Points
We have seen how one goes about selecting a slice in
the body. This is done via a slice-select gradient. To
vary the slice thickness, we can either vary the
bandwidth of the transmitted RF pulse or the slope of
the gradient. Thus, we can decrease the slice
thickness by
1 decreasing the bandwidth of the RF pulse, or
2 increasing the slice-select gradient.
Because the RF profiles are not ideal and may have
side lobes or tails, if you try to have contiguous
slices, you'll run into a problem called cross-talk.
Basically, the transmitted signals in the frequency
domain (i.e., their Fourier transforms) will overlap
and “cross-talk.” To avoid this, you must introduce
gaps between the slices. This is done by excluding a
certain range of frequencies (i.e., bandwidths) in the
transmitted RF pulses. The larger the gaps, the less
cross-talk you get, but the chance of missing a lesion
within the gaps is greater.
The center frequency for each transmitted bandwidth
is like a carrier frequency around which the desired
bandwidth is centered, much like what goes on in
radio-communication. Once a slice is selected, the
question arises as to how to determine the pixels
within that slice. This is the topic of the next chapter.
Questions
10-1 (a) The range of frequencies included in an RF pulse is referred
to as its bandwidth (BW). Suppose that an RF pulse has frequencies
ranging from −500 to 500 Hz (i.e., BW = 1,000 Hz = 1 kHz). Now, to
achieve a slice thickness of 5 mm, determine the amplitude of the
slice-selection gradient.
(b) What is the minimum achievable slice thickness given a minimum
RF BW = 426 Hz and a maximum gradient Gz = 10 mT/m?
Hint: ω = γB so Δω = BW = γΔB. Now, B = Gzz so that ΔB = GzΔz. Thus
BW = γΔB = γGzΔz or Δz = BW/(γGz), where Δz = slice thickness and γ
= 42.6 MHz/T.
Introduction
In the last chapter, we learned how to select a slice and how to adjust
its thickness. However, we did not address the question of from where
within a particular slice each component of the signal comes. In other
words, we still don't have spatial information regarding each slice. To
create an image of a slice, we need to know how much signal comes
from each pixel (picture element) or, more accurately, each voxel
(volume element). This is the topic of spatial encoding, of which there
are two parts: (i) frequency encoding and (ii) phase encoding.
Frequency Encoding
After selecting a slice, how can we get information about individual
pixels within that slice? As an example, consider a slice with three
columns and three rows, for a total of 9 pixels. This slice is selected
using a selective 90° pulse (Fig. 11-1). We turn the Gz (slice-select
gradient) on during the 90° pulse and turn it off after the 90° pulse.
Figure 11-1. Spin-echo pulse sequence diagram.
The Gx gradient is applied during the time the echo is received, that
is, during readout.
Let's now assign some magnitude numbers to the pixels in the matrix
(Fig. 11-3).
The numbers in each pixel and their specific location is what we
ultimately want to discover because this corresponds to an image. We
want to recreate this image using MRI.
Initially, all the protons in this section experience the same frequency
of precession. Let's call that frequency ω0. Now let's assign each pixel
its frequency at a specific point in time before turning on the Gx
gradient, while they all still have the same frequency, and combine it
with the magnitude we've assigned to each pixel (Fig. 11-4). For
simplicity, we'll use a cosine wave as the received signal. In reality, the
received signal is a more complicated one, such as a sinc wave.
Each pixel has a designated magnitude (amplitude), and they all have
the same precessional frequency ω0 (except those pixels that have zero
signal amplitude). Without any gradient in the x direction, this is the
signal we're going to get. The signal will be the sum of all the signals
from each pixel.
The sum of the amplitudes = (0) + (1) + (-2) + (1) + (2) + (0) + (1) + (0) +
(1) = 4 The frequency is the same for each pixel (i.e., ω0), as is the
shape of the signal (i.e., cos ω0t). So the sum of the pixels = signal
from whole slice = 4 cos ω0t We know that, in reality, the signal is more
complex. For example, the signal is a decaying signal with time such as
a sinc wave, but, for simplicity, let's accept that we are dealing with
The signal that we get now is still the sum of all the individual signals;
however, now each column of pixels has a different frequency, so we
can algebraically only add up the ones that have the same frequency,
as follows: Column #1: 0 + (cos ω0−t) + (−2 cos ω0−t) = −cos ω0−t Column
#2: (cos ω0t) + (2 cos ω0t) + 0 = 3 cos ω0t Column #3: (cos ω0+t) + 0 +
(cos ω0+t) = 2 cos ω0+t
Sum of signals = (−cos ω0-t) + (3 cos ω0t) + (2 cos ω0+t)
Let's look at the Fourier transform of the signal before the Gx gradient
is applied (Fig. 11-6A) and look at it again after the Gx gradient is
applied (Fig. 11-6B). For a cosine wave, the Fourier transform is a
symmetric pair of spikes at the cosine frequency, with an amplitude
equal to the magnitude of the signal. (Remember that this is simplified.
Usually, we deal with a band of frequencies, i.e., the bandwidth, as
opposed to a single frequency. However, right now, for simplicity, we
leave it as a single frequency, with its Fourier transform as a single
spike.)
Now the computer can look at the Fourier transform and see that we
are now dealing with three different frequencies:
1. The center frequency comes from the central column, and the
amplitude of the frequency spike represents the sum of the
amplitudes of the pixels in that column, that is, (3 cos ω0t).
2. The higher frequency comes from the column to the right, and the
amplitude of that frequency spike represents the sum of the
amplitude of the pixels in that column, that is, (2 cos ω0+t).
3. The lower frequency comes from the column to the left, and the
amplitude of that frequency spike represents the sum of the
amplitude of the pixels in that column, that is, (−cos ω0−t).
The way frequency encoding works is that frequency and position have
a one-to-one relationship:
frequency ↔ position
So far we have done some spatial encoding and have extracted some
information from the slice.
We are now able to decompose the slice matrix into three different
columns (Fig. 11-7). That is, we have three different shades of gray
corresponding to the three columns.
Figure 11-5. When the matrix is exposed to the frequency
gradient, different frequencies result in each column: ω0, ω0+, and
ω0−.
So, now we've done our job in the x direction. The next thing we want
to do is to decompose the individual columns into their three individual
pixels (i.e., work in the y direction). The two ways of doing this are as
follows:
1. Back projection
2. Two-dimensional Fourier transform (2DFT)
Back Projection. If we think in terms of CT imaging and apply
gradients, we start out with an area we want to image and apply a
gradient (Fig. 11-8A). Then we can rotate the gradient by an angle θ
and reapply the gradient (Fig. 11-8B). We can continue this to complete
360°, and
Advantage
1. It is possible to pick a small field of view (FOV).
Disadvantages
1. This technique is very dependent on external magnetic field
inhomogeneities (i.e., it is sensitive to ΔB0).
2. This technique is also very sensitive to the magnetic field gradients.
If the gradient is not perfect, you get artifacts.
Because of these disadvantages, this technique was given up.
Advantages
1. Lack of sensitivity to external magnetic field inhomogeneities
2. Lack of sensitivity to gradient field inhomogeneities
Figure 11-9. The phase-encode gradient Gy is applied along the y-
axis. It is usually applied between the 90° and the 180° pulse or
between the 180° pulse and the echo. The first Gx gradient is used
for offsetting any phase shift induced during frequency readout
(further discussion in Chapter 14).
Phase Encoding
In the 2DFT technique, in addition to using the Gz gradient for slice
selection and the Gx gradient for encoding in the x direction, we add
another gradient Gy in the y direction. This is called the phase-
encoding gradient (Fig. 11-9).
We turn on the Gy gradient before we turn on the Gx (readout)
gradient. It is usually applied right after the RF pulse or just before the
Gx gradient or anywhere in between.
This is precisely what the clock diagram shows: Prior to being exposed
to a magnetic field gradient, all the protons in each pixel are in phase
with each other, oscillating at the same frequency. (Note that we have
eliminated magnitude from the clock diagram. All we are concerned
about for
up and down are reflected in that phase value. Hence the term phase
encoding.
Remember that the Gy gradient is turned on before reading out the
signal. So when we read the signal, we turn on the Gx gradient, which,
as we learned from our earlier discussion, allows us to frequency
encode in the Gx direction (Fig. 11-10C). With the Gx gradient on, the
middle column protons won't experience any change in their
precessional frequency, so their frequency is unchanged. However, as
you can see, each pixel in the middle column already has a distinct
phase shift, which had occurred when the Gy gradient was on, and this
phase shift persists.
With the Gx gradient turned on, the protons in the columns to the right
of midline will experience a greater magnetic field so that all the
protons in this column will have a faster precessional frequency.
However, we notice that each pixel in this column was already out of
phase with the other pixels in the column due to the phase shift that
had occurred when the Gy gradient was on. Therefore, protons in each
pixel of the right column shift the same amount (because they all have
the same increased frequency). However, because they shift from a
unique position, they will then each move to a phase shift that is
different for each pixel.
Likewise, the protons in the column to the left of midline will
experience a lower precessional frequency with the Gx gradient on,
but, again, we notice that each pixel in this column is already out of
phase with the other pixels in the column due to the phase shift that
had occurred when the Gy gradient was on. So, again, after the Gx
gradient is turned on, each pixel will move to a specific phase shift
distinct for each pixel. In summary, x position is represented by a
unique frequency and y position by a unique phase.
Figure 11-11. The received signal after application of both Gx and
Gy for the 3 × 3 matrix used in previous examples.
Note, however, that the center of the data space does not represent
the center of your picture. Each signal has information in it about the
entire picture. Remember that each signal that goes into each row of
the data space is the sum of all the signals from individual pixels in the
slice.
The information in the data space is in the time domain (so it is not as
scary as it looks). In fact, it is in the time domain in both directions:
the received signal is displayed over a period of time (t), and signals in
two successive rows are obtained at every TR (Fig. 11-15).
The information in this data space has not yet been digitized. In fact, a
digitized version of this information is the true k-space. We will see in
a later chapter (Chapter 16) that the “digitized” k-space is in a spatial
frequency domain. But let's not get confused now! Let's see how we
can digitize this information in the data. This is accomplished via
sampling.
Sampling
The signal that goes into the data space has been phase encoded and
frequency encoded, but it has not yet been sampled (more on this in
the next chapter).
When we describe a matrix of, say, 256 × 192, what do we mean by it?
If we only have 192 phase-encoding steps, why do we have 256 (instead
of 192) frequency encodes if we only need one frequency encode for
every encoding step?
Actually, the 256 number refers to the different number of frequencies
we have for each phase-encoding step. These two steps are thus
independent from one another. For example, let's look at a 4 × 5
matrix. This means that we do five different frequencies for each
phaseencoding step, with the center column having no change in
frequency and two different frequencies for the columns on either side;
there will be
How do we go from the data space to the desired image? The answer is
via Fourier transformation. The Fourier transform of the signals in
Figure 11-15 is a set of spikes, which in a three-dimensional space looks
like Figure 11-16. More on this in chapters to come.
Key Points
In the previous chapter, we learned how to select a
slice using a slice-select gradient Gz. In this chapter,
we saw how to determine the pixel values in a slice
using two gradients: (i) a frequency-encoding (or
readout) gradient Gx and (ii) a phase-encoding
gradient Gy.
The Gx gradient is the same strength for each echo,
and since it is applied during the readout (i.e., during
reception of the echo), it alters the Larmor frequency
along the x-axis. This provides specific information
along the x-axis. Each TR interval contains one
readout (Gx) per slice.
The Gy gradient, however, is applied in increments
between the 90° RF pulse and each echo. Since this
gradient is applied at a time apart from the echo, it
does not change the echo frequency, but merely
induces a phase shift. Thus, for every slice, each TR
interval contains one phase-encoding step (i.e., one
unique value of Gy). This process completes one line
in k-space corresponding to the selected Gy. This
process is repeated Ny times to fill all of k-space.
We have not yet discussed the mechanism of
performing frequency- and phase-encoding
operations. This is the topic of the next chapter.
Questions
11-1 Match
(i) Gx (ii) Gy (iii) Gz with
(a) applied during the echo
(b) applied during the RF transmission
(c) applied between the RF and the readout
11-3 What is the phase increment for 128 phase-encode steps (i.e.,
Ny = 128)?
11-4 Match
(i) position along x-axis
(ii) position along y-axis with
(a) phase-encode gradient Gy
(b) frequency-encode gradient Gx
(c) absolute phase φy
(d) absolute frequency fx
Introduction
Signal processing refers to analog and/or digital manipulation of a
signal. The signal could be an electric current or voltage, as is the case
in magnetic resonance imaging (MRI). Image processing is a form of
signal processing in which the manipulations are performed on a
digitized image. Analog-to-digital conversion (ADC) is a process by
which a time-varying (analog) signal is converted to a digitized form
(i.e., a series of 0s and 1s) that can be recognized by a computer. An
understanding of signal processing requires a basic understanding of the
concept of frequency domain and Fourier transform (FT) because most
of the “processing” of a signal is accomplished in the frequency domain
and, at the end, the results are converted back into the time domain.
One of the key concepts in signal processing, as we shall see shortly, is
the Nyquist sampling theorem. An understanding of the sampling
procedure allows one to appreciate the relationship between the
samples of a signal (in the time domain) and its bandwidth (in the
frequency domain). Once this concept is grasped, the issue of aliasing
(wraparound) artifact can be explained very easily. A knowledge of
signal processing will also help the reader understand the more
complicated, newer fast scanning pulse sequences presented in later
chapters.
Sequence of Events
First, let's summarize what has been discussed so far. Figure 12-1
illustrates a summary of a spin-echo pulse sequence. The following is a
summary of the sequence of events:
1. We have 90° and 180° pulses separated by a time of TE/2 msec.
2. After a time of TE millisecond after each 90° radio frequency (RF)
pulse, we get an echo.
3. We turn on the slice-selective gradient (Gz) during the two RF
transmissions. This causes a linear gradient of magnetic field along
the z-axis. By choosing an RF pulse with an appropriate frequency
and bandwidth, we can select a slice at a particular position with a
particular thickness.
As an aside, consider Figure 12-2. Plotting field strength versus position
(Fig. 12-2), the gradient is represented as a sloped line. The slope of
the line is a constant that we call G. The value y along this line with
slope G at point x is y = Gx. This is a simple linear equation. Figure 12-
2B plots gradient strength versus time. Figures 12-2A and 12-2B are
used interchangeably to illustrate a linear gradient with strength G.
4. Right before we receive the echo, we apply a phase-encoding
gradient (Gy). The symbol for the phase-encoding gradent is in Figure
12-3. This symbol denotes the multiple phase-encoding steps that are
necessary as we cycle through the acquisition. Remember that one of
the phaseencode steps may be performed without any gradient.
Figure 12-1. Spin-echo pulse sequence diagram.
There may also be time taken up by other events that occur before the
RF pulse (such as presaturation pulses) that we include as overhead
time (To). Thus,
To choose the next slice, we keep the same magnetic gradient Gz, but
we choose a bandwidth at a higher or lower center (or Larmor)
frequency to flip the protons 90° in a different slice. The bandwidth is
the same as the first slice, but the center frequency is different.
We choose the same phase-encoding gradient Gy so we get the same
amount of dephasing in this next slice. We sample the echo with the
same frequency-encoding gradient as we did with the first slice.
Because we still have time during the dead time to acquire a third
slice, we repeat everything and again choose a 90° RF pulse of a
different Larmor frequency so we can flip the protons of a different
slice into the transverse plane. The signal from each slice will be
placed in a different k-space.
Examples
1. TR = 1000, TE = 35 msec, Ts = 10 msec, To = 10 msec (short TE)
Aliasing
We don't want to take too few samples per cycle, that is, we don't want
the sampling interval to be too wide because then the square waves
will overlap and aliasing will result (Fig. 12-17).
Why is the FT of a continuous sinc function a single wave, whereas that
of its sampled version shows repetitive square waves (Fig. 12-18)? The
reason for this is mathematical. We will try to give an explanation that
doesn't involve the complicated mathematics behind it.
Figure 12-17. When the sampling interval is too long (i.e., not
enough samples are taken), the rectangles may overlap (causing
aliasing).
Figure 12-18. The FT of a sinc function is a rectangle, but that of
its sample variant is a series of rectangles.
Figure 12-19. To sample a continuous, analog signal, a series of
spikes (called delta functions) is multiplied by the signal.
Let's now take another example in which we have fewer than four
samples. Let's try two samples per cycle (Fig. 12-28A). In this case,
1. The period is still 4 sec.
2. δTs = sampling interval is 2 sec.
3. 1/δTs = 1/2 = 0.5 Hz.
Let's see the FT of this example (Fig. 12-28B). Frequency is still 1/4.
The sampling interval is
now 2 sec, so 1/δTs = 1/2 Hz. We show two spikes on either side of 1/2
Hz, which is the frequency range around the center frequency of 1/2
Hz.
We see now that the pair of spikes centered at 1/2 Hz come right next
to the original pair of spikes of 1/4 Hz. They practically overlap. If we
pass this FT through an LPF to eliminate the higher frequencies, we still
recover the two frequency spikes, which are the FTs of the original
cosine signal; however, they will be increased in amplitude because of
the overlap (Fig. 12-29).
Now let's see what happens if we take even fewer samples (Fig. 12-
30A). Here,
1. The period is still 4 sec.
2. But δTs = 3 sec.
3. 1/δTs = 1/3 Hz.
4. The frequency is still 1/4 cycles/sec.
So, let's look at the FT (Fig. 12-30B). If we put this FT through an LPF,
the spikes associated with the center frequency (1/δTs) will interfere
with the original signal (Fig. 12-31). We will not only have the original
spikes but also have two extra spikes that are closer to the center. Now
we have the sum of two cosine waves rather than the spikes of the
single original cosine wave. We have the spikes of another cosine wave,
which are closer together, meaning that its transform is a cosine wave
of lower frequency. This is called aliasing.
What we wanted to approximate by sampling was the original signal,
having a frequency of 1/4 Hz. But as a result of undersampling, we also
obtain an undesired signal with a much lower frequency (Fig. 12-32).
Aliasing: An Analogy. As an analogy, let's consider a stage-coach wheel
in a Western movie. Sometimes, it appears as if the wheel is turning in
the reverse direction. Let's see how this happens. When we take a
motion picture, we are actually taking samples in time. Let's take one
point on the wheel and take a sample at time t = 0 (Fig. 12-33). At a
later time (t1), we take another frame. The point rotates a certain
distance forward as the wheel turns. When we watch this in a motion
picture we see this point rotating in a clockwise direction.
Figure 12-29. An LPF applied to the previous FT allows recreation
of the original signal (although the height of the spikes is doubled).
Figure 12-30. When too few samples are taken (A), the spikes will
get mixed up (B).
Figure 12-31. When a (low-pass filter) LPF is passed through the
previous example, two sets, instead of one set, of spikes are
produced, which would yield a different signal than the original
cosine function. This is called aliasing.
Figure 12-32. When too few samples are taken, the perceived
(aliased) frequency is different from the actual (original)
frequency.
Now let's see what happens when we undersample (Fig. 12-34). Let's
say the first frame starts at the same point as before at t = 0. But on
the next frame, we wait longer (t2) to sample and don't sample until
the point almost comes around to its original spot. We wait a similar
time for the next sample and the point again comes around almost to
the spot on the second frame. If we take a motion picture of this, the
wheel will appear as if it's turning counterclockwise. This is because we
are undersampling. This is an example of aliasing. The wheel is actually
turning in a forward direction, but because of undersampling, it
appears to be doing something else.
Aliasing comes from the term alias: a fake name. We have a real
frequency, but because of undersampling, we appear to have a fake
frequency. In the previous example in which the cosine wave was
undersampled, the real signal was a cosine wave with a frequency of
1/4 Hz,
but the “alias” cosine wave had a lower frequency (Fig. 12-31). In the
example of the wagon wheel, the real rotation was in a clockwise
direction, but the alias was in a counterclockwise direction, probably
with a slower rotation.
Figure 12-35. Nyquist law: To avoid aliasing, the maximum
frequency in the signal (fmax) should be less than half the sampling
frequency (1/δTs). In other words, the sampling interval (δTs)
should be at least twice the minimum period (1/fmax). That is, at
least two samples per cycle (corresponding to the highest
frequency in the signal) are required to avoid aliasing.
This is easy to see in the diagram (Fig. 12-35). We want the sampling
rate to be at least equal to the sum of the maximum frequencies of
adjacent boxes (to keep the boxes from overlapping).
Figure 12-36. The original signal can be reconstructed with a
minimum of two samples per cycle.
Digitized k-Space
We've already shown how we fill lines of the data space with our signal,
with each line performed using a different phase-encoding step. What
we actually put into each line of the data space are the sampled data
from each signal. Once we put these samples into each line of data
space, we complete the data space.
All MRI machines operate under the same principle when it comes to
sampling the signal. A minimum of two samples per cycle is taken and
put into the data space (Fig. 12-39A). If more than two samples per
cycle are taken, then the bandwidth is wider, but it won't give us a
better approximation of the signal (Fig. 12-39B). If we take less than
two samples per cycle, the adjacent bandwidths may overlap and cause
aliasing (Fig. 12-39C).
Figure 12-39. A: Operating at Nyquist frequency. B: When more
samples are taken, the BW will be increased. C: When too few
samples are taken, aliasing may occur (Ts3 > Ts1 > Ts2).
Signal-to-Noise Ratio
When the bandwidth is narrowed, the signal-tonoise ratio (SNR or S/N)
is increased. (This topic is discussed at length in a later chapter.) SNR is
inversely proportional to the square root of the bandwidth. SNR is also
proportional to the volume of the pixel and to the square root of the
number of phase-encoding steps (Ny), the number of frequency-
encoding steps (Nx), and the number of excitations (NEX):
Example 1
a. At 1.5 Tesla, it typically takes 8 msec to perform one readout: Ts = 8
msec.
b. We have a matrix of, say, 256 × 256 pixels.
By convention, the first number refers to the number of frequency-
encoding steps. The second number refers to the number of
phaseencoding steps. What is the bandwidth? We know that
BW = 1/δTs
What is δTs?
δTs = sampling time/number of samples we test = 8 msec/256 samples
BW = 1/δTs = 1/(8 msec/256) = 256/8 msec = 256/0.008 sec = 32,000 Hz
= 32 kHz = ± 16 kHz
This is a typical bandwidth for a typical readout with 256 frequency-
encoding steps and a sampling time of 8 msec. Therefore, a frequency
bandwidth of 32 kHz (±16 kHz) is a fairly typical frequency bandwidth
that we deal with in routine imaging. This means that the bandwidth
extends to +16 kHz on the right and -16 kHz on the left side of the
center frequency.
Example 2
What would happen if we go to a 512 × 512 matrix, that is, with 512
frequency-encoding steps?
a. We could either have a larger frequency bandwidth:
b. Or, we could double the sampling time and keep the sampling
interval (and the bandwidth) the same.
(We'll get into the relationship between the field of view and
bandwidth later.)
Key Points
In this chapter, we have presented the basic concepts
of signal processing, of which image processing is a
subset. As mentioned in the introduction, the
understanding of these concepts is crucial in
understanding the intricacies of image optimization,
which is one of the goals of every imager. Again,
memorizing the formulas is not as important as
understanding the concepts behind them.
Let's summarize:
1 ADC (analog-to-digital conversion) is the process in
which an analog (time-varying) signal is encoded to a
digital signal (containing a series of binary numbers
0s and 1s) represented as bytes (8 bits) in the
computer.
2 This is done by sampling the signal.
3 To be able to reconstruct the original signal from its
discrete samples, the Nyquist law must be satisfied.
Otherwise, aliasing will occur.
4 The Nyquist theorem states that the sampling
frequency must be at least twice the highest
frequency present in the signal.
5 Stated differently, if you take the waveform of the
component signal with highest frequency (remember
that each signal is a composite of many different
signals with varying frequencies), then at least two
samples per cycle are required to avoid aliasing.
6 Therefore, aliasing occurs because of
undersampling. To ensure that aliasing will not
happen, MR scanners may automatically perform
oversampling.
7 The bandwidth (BW) is defined as the range of
frequencies in the signal.
8 BW = 1/δTs, where δTs is the sampling interval
(interval between two samples).
9 At Nyquist frequency, BW = 2ωmax, where ωmax is
the highest frequency in the signal, so that δTs =
1/BW = 1/(2ωmax).
10 Sampling time Ts = Nx · δTs, where Nx is the
number of frequency encodes.
11 SNR is given by
12 So, SNR is ↑ if BW ↓
13 A narrower BW is used when a higher SNR is
desired (e.g., on the second echo of a dualecho SE
image).
14 BW is ↓ if δ Ts is ↑(i.e., less samples are taken),
which may cause aliasing!
15 Now, if ↑ Ts is ↑ then Ts = Nx · δTs is ↑, which
causes TE to ↑. Because
Number of slices ≅ TR/TE then a narrower BW will
reduce the coverage.
Questions
12-1 According to the Nyquist theorem, to avoid aliasing:
(a) At most, two samples per cycle corresponding to the highest
frequency are required.
(b) At least two samples per cycle corresponding to the highest
frequency are required.
(c) At most, two samples per cycle corresponding to the lowest
frequency are required.
(d) At least two samples per cycle corresponding to the lowest
frequency are required.
12-3 T/F The bandwidth (BW) is the inverse of the sampling interval
(δTs).
Introduction
Before we can understand k-space, we need to discuss the data space,
which is a matrix of the processed image data. To most radiologists,
kspace is in the twilight zone! We have already seen some of the basic
concepts of the data space and k-space in the previous chapters. In this
chapter, we will learn some of the properties of k-space in greater
detail. The understanding of k-space is crucial to the understanding of
some of the newer MRI fast scanning techniques such as fast spinecho
(FSE) and echo planar imaging (EPI).
Figure 13-1. The data space (with both axes as time variables) is
an “analog” version of k-space.
6. For the second row in the data space, we do the exact same thing,
except in this step the signal is obtained using a largar phaseencoding
gradient in the y-axis.
Remember that the phase-encoding gradient causes dephasing of the
signal. Therefore, the signal for the second line of the data space will
be similar in shape to the first signal (because both are signals from the
same slice of tissue, just obtained at a different time) but smaller in
magnitude than the first signal (because it undergoes additional
dephasing due to the phase-encoding gradient). Thus, when we draw
this signal into the second line in the data space, we see that it is
similar in shape to the first signal, but slightly weaker—because it's
been dephased.
The signal that goes into the last line of the data space (+128) will be
almost flat because it has undergone maximum dephasing; likewise, as
we alternate to the signals placed below the zero line (i.e., −1, −2, …,
−127), a certain symmetry results. For instance, line (−1) is similar in
strength to line (+1) in that, whereas line (+1) experiences mild
dephasing due to a slight increase in magnetic field strength, line (−1)
experiences similar mild dephasing due to a slight decrease in magnetic
field strength. Likewise, the signal that goes into the first line in the
data space (−127) will be almost flat due to maximum dephasing in the
opposite direction of line (+128).
Remember that each line in the data space contains the signal obtained
from the entire image slice during a single TR. Each TR is obtained
using a different phase-encoding step in the y-axis.
Question 1: How long does it take to go from one row in the data space
to another?
Answer: It takes the time of one TR.
Question 2: How long does it take to go from one point (sample) in a
row to the next point (sample) in the same row of the data space?
Answer: It takes the time spent between samples, that is, the sampling
interval (ΔTs).
Question 3: How long does it take to fill one row of the data space?
Answer: Let's say that ΔTs [all equal to] 31 µsec and that there are 256
(N) samples along the readout axis. The sampling time is Ts = (ΔTs)(N) =
(31µsec)(256) = 7.9 msec
Therefore, it takes about 8 msec to fill one line of the data space. In
general, it takes Ts = Nx · ΔTs
to fill one line of the data space.
Question 4: How long does it take to fill one column of the data space?
Answer: It is the acquisition time Np × TR, where Np is the number of
phase-encoding steps.
Let's say TR = 3000 msec and Np = 256.
Acquisition time = (3000 msec) (256)
= 12.8 min
If TR = 500 msec, then
Acquisition time = (500) (256) [all equal to] 2 min
It takes several milliseconds to fill one row of the data space. But it
takes several minutes to fill the columns of the data space.
Motion Artifacts
The preceding concept is one of the reasons why motion artifacts
manifest themselves mainly in the phase-encoding direction. In other
words, it takes much longer to gather the signal in the phase-encoding
direction than in the frequencyencoding direction, leaving more time
for motion to affect the image in the phase direction. Another reason,
as we shall see later, is that motion in any direction results in a phase
change; thus, motion artifact propagates along the phase-encoding
direction.
Properties of k-Space
Center of k-Space. The center of the data space contains maximum
signal. This finding is caused by two factors:
1. Each of the signals has its maximum signal amplitude in the center
column (Fig. 13-3). Recall that when we apply the 180° refocusing
pulse, the dephased signal begins to rephase and reaches maximum
amplitude when the protons are completely rephased. It then
decreases in amplitude as the protons dephase once more.
The middle column in the data space corresponds to the center of
each individual echo, and the more peripheral columns refer to the
more peripheral segments of the echoes: columns to the left of
center depict rephasing of the echoes toward maximal amplitude in
the data space; columns to the right of center depict dephasing of
the echoes away from maximal amplitude in the data space.
Therefore, as we go further out to the more peripheral columns, the
signal weakens. The most peripheral points in the signal to the left
are the weakest point of the signal as the signal just begins to
rephase (Fig. 13-3). Likewise, the most peripheral point in the signal
to the right is the weakest point after the signal has been refocused
and then has regained maximum dephasing.
2. The maximum amplitude occurs in the center row because this line is
obtained without additional dephasing due to phase-encoding
gradients; subsequent rows with progressively larger phase-encoding
gradients have weaker signal amplitude.
Therefore, because the middle row has the strongest of all echoes and
the middle column contains all the peaks of the echoes, the center
point of the data space contains maximum amplitude, that is, maximum
signal-to-noise ratio (SNR) (Fig. 13-4).
Figure 13-3. The peripheral points in the signal have the weakest
amplitude; the center point has the maximal amplitude.
Figure 13-5. A: The original raw data (k-space) of (B) the original
image (midline sagittal T1-weighted image of the brain).
the signal (echo), ring artifacts are introduced in the Fourier transform.
Therefore, by eliminating the samples in the periphery of the data
space, the sharp interfaces in the image are degraded and the image
gets coarser. In other words, the fine detail of the image is
compromised when the edges of k-space are excluded. Figure 13-8B is
the image corresponding to the periphery of k-space (Fig. 13-8A).
Figure 13-7. A: The FT of an ideal sinc function is a rectangle. B:
The FT of a truncated sinc function has ring down effects.
Once we have all the data in k-space, we take the Fourier transform of
k-space to get the image.
Question 1: Why is the Fourier transform of k-space the desired image?
Answer: Because there is a one-to-one relationship between frequency
and position in the x direction and between phase-encoding gradient
strength1 and position in the y direction.
Question 2: Why is there a one-to-one relationship between frequency
and position?
Answer: Because, in our method of spatial encoding, we picked a linear
gradient in the x direction that correlated sequential frequency
increments with position; likewise, we picked a linear gradient in the y
direction that correlated sequential phase gradient increments with
position in the y direction (Fig. 13-9).
Thus, the center of the field of view of the image experiences no
frequency gradient and no phase gradient, and the points in the
periphery of the image experience the highest frequency and phase
gradients. In other words, there is a 1:1 relationship between frequency
and position in the image.
In summary, the frequency- and phaseencoding gradients provide the
position of a signal in space. They tell us which pixels each component
of the signal goes into in the slice under study.
Question: How are the shades of gray determined?
Answer: The shades of gray are determined by the magnitude or
amplitude of the signal (actually its Fourier transform) at each pixel.
Recall that the image of k-space looks like a series of concentric circles
of alternating intensity on a two-dimensional surface. If we now
incorporate amplitude as a third dimension, we would have the areas of
greater amplitude coming off the surface of k-space toward us like a
“warped” image (Fig. 13-10).
k-Space Symmetry. One step needs to be completed after receiving the
signal and before placing it in k-space that we have so far ignored. This
step is called phase-sensitive detection. We want to take the echo
signal, which is on a carrier frequency, shift it to zero frequency, and
divide the signal into its real (cosine) and imaginary (sine) components.
First, we start off with the signal that is being frequency- and phase-
shifted around a carrier frequency of 64 MHz for a 1.5-T magnet.
However, it's hard to tell whether a signal has been frequency- or
phase-shifted unless we “ground” the signal back to “zero.”
2. The data space with “imaginary” (sine) data: the signal that has (sin
ω0t) subtracted from it and brought back to 0 frequency.
We now have two data spaces (Fig. 13-12): one with cosine data (real);
one with sine data (imaginary). Both have data centered at 0
frequency. In the data space with cosine data, we know that a great
deal of symmetry exists. A cosine function is an example of an even
function. If we look at a cosine function, we see that there is symmetry
to the right and left of zero. In addition, there is symmetry above and
below zero. Thus, if we put a pixel in a line of the data space to the
right of the 0 column, and above the 0 line (point a), the symmetry of
the cosine function would make us unable to discriminate between the
other (a) positions. The computer
couldn't tell the difference between any of the four pixel positions. This
is why we use the sine version of the data space.
Figure 13-12. A, B: Spatial direction in k-space. The imaginary k-
space provides a series of left-right or up-down directions.
Now look at the data space with the sine data (Fig. 13-12B). Again, the
pixel is in the same place as the cosine data space; it is in a line of the
data space above the 0 line and to the right of the 0 column. However
(unlike the cosine data space), here we can distinguish it from the pixel
below the 0 line (−b). We can also differentiate it from the pixel to the
left of the 0 column (−b).
Why are these pixels different in the sine data space? Let's review the
sine function (sin ω0t) of the two pixels to the right of the 0 frequency
(Fig. 13-13A). The sine function is an example of an odd function
because of its inherent antisymmetry. The sine function changes
polarity above and below the 0 line. This allows us to differentiate the
two pixels.
Now, let's examine the sine function (sin ω0t) above the zero line of the
data space (Fig. 13-13B). Again, because of inherent antisymmetry of
the sine function on either side of the 0 frequency, the pixels will have
opposite polarity.
Complex Numbers
We said in Chapter 1 that a complex number can be divided into its real
(cosine) and imaginary (sine) parts. If we consider the cosine function
as the real component and the sine function as the imaginary
component, then we can add the sine and cosine together to get the
magnitude of the signal as well as its direction.
So now let's add up the data of the 4 pixels (Fig. 13-14). Because of the
changing polarity of the sine function, when we add the sine function
to the cosine function, we can distinguish the direction of the four
pixels (whereas with cosine function alone, we couldn't tell the
direction).
In the lines above 0 phase encoding:
Pixel a − ib is to the left of the 0 frequency.
Pixel a + ib is to the right of the 0 frequency.
In the lines below 0 phase encoding:
Pixel a + ib is to the left of the 0 frequency.
Pixel a − ib is to the right of the 0 frequency.
Conjugate (Hermitian) Symmetry. The conjugate of a complex number
a + ib is the complex number a − ib (i.e., with the same real component
but a negative imaginary component). From this and from Figure 13-14,
it is clear that k-space possesses conjugate symmetry, also known as
Hermitian symmetry.
Half NEX (½ NEX). In a “½ NEX”2 technique (half-Fourier in phase), we
acquire the data from the upper half of k-space and construct the
lower part mathematically (Fig. 13-15), thus reducing the scan time.
The trade-off is a reduced SNR by a factor of , to be exact (see
Chapter 17). Due to the presence of phase errors in the data, the
symmetry previously discussed may not be perfect. This is why when
employing such techniques, a few extra rows in the center of k-space—
which contains maximum signal—are always added to
allow for such phase corrections. That is, slightly more than 50% of k-
space must be sampled to maintain phase information.
Figure 13-13. A: Because the phase gradients corresponding to the
top and bottom half of k-space generally have opposite polarities,
the values in the corresponding imaginary k-space also will have
opposite polarities. B: Because the sine function is an odd function,
the left half of the signal is the reverse of the right half; thus the
corresponding points in the imaginary k-space will also have
opposite signs.
Fractional Echo. In fractional echo, only the right half of the echo is
sampled, and the left half is constructed based on the right half (Fig.
13-16). (This allows TE to be shorter for fast scanning techniques like
turbo FLASH and Fast SPGR—see Chapter 21.)
¼ NEX. Because of the conjugate symmetry discussed previously,
theoretically you should be able to create an image using only one
quadrant of the combined real and imaginary data spaces (Fig. 13-17).
That is, you should be able to construct the entire k-space data from
only one quadrant. In reality, however, due to the presence of data
acquisition errors, perfect symmetry does not exist and doing so may
lead to phase errors and image distortion, which is probably why this
technique is not being used.
magnitude by .
This concept can be applied to the real and imaginary components of
the image (Fig. 13-18) to generate the magnitude and phase images
(Fig. 13-18).
The magnitude image (modulus) is what we deal with most of the time
in MRI. The phase image is used in cases in which the direction is
important. An example is phase contrast MR angiography, in which the
phase indicates the direction of flow, i.e., up versus down, anterior
versus posterior, or left versus right. In summary, tangent (phase angle)
= (imaginary/real), or phase angle = arctan (imaginary/real), and
Figure 13-18. The FTs of the real and imaginary k-spaces provide
the real and imaginary images, respectively. The real and
imaginary images are used to create magnitude and phase images.
We too can look at the phase image. In flow imaging, the phase image
is a velocity image and indicates magnitude and direction. For
example, in imaging the cerebrospinal fluid (CSF) flow through the
aqueduct, flow in the antegrade direction could be black, and flow in
the retrograde direction could be white on the phase images. Thus,
phase images in phase-contrast studies display the direction of flow
(Fig. 13-19).
Figure 13-20. An example of a 2 × 4 matrix exposed to a frequency
gradient and no phase gradient.
Because the phase is never zero, we can combine the “real” image and
the “imaginary” image to get a composite image, which is the image
that we look at when we read an MR study. The modulus is the image
we look at; it combines the data corresponding to the Fourier transform
of the real and imaginary data spaces:
image = modulus
k-Space: An Example. The following is an example of a 2 × 4 matrix:
the first row. This will result in no change in the numbers of the first
row. But the numbers in the second row will be 180° phase shifted
(i.e., they will be the negative of the original numbers).
Figure 13-22. An example of a 4 × 4 matrix. The phase increments
here are 0°, 90°, 180°, and 270°. (In general, the phase increment
is 360°/Ny, where Ny is the number of phase-encoding steps.)
If we use the clock analogy to evaluate phase shift, the spins in the top
row, experiencing no phase shift, will all point upward. The spins in the
second row (which is experiencing a 180° phase shift) will all be
pointing downward (Fig. 13-21).
Thus, whereas the values in the first row will remain unchanged, the
values in the second row will be 180° reversed from what they were
with no phase shift.
Figure 13-23. The effect of phase shifts on pixel values for two
and four rows.
Row 1
Row 2 before 180° phase shift
Row 2 after 180° phase shift
of 180°, we get the negative value of the original number because cos
180° = −1. For a different phase angle, we would get a fraction of the
original value (from 0 to 1, or from 0 to −1, whatever cos θ is).
Figure 13-24. The previous 2 × 4 example now exposed to a
frequency gradient and a phase gradient (i.e., 180° phase shift).
With this phase-encoding step, let's see what the Fourier transform
would be in the frequency-encoding direction (Fig. 13-24). The
amplitude in columns 1 to 3 remains unchanged from the 0 phase
readings. However, there is a change in the amplitude of column 4:
With zero phase, column 4 adds up to +2 (because 1 + 1 = 2).
With 180° phase, column 4 adds up to 0 (because 1 − 1 = 0).
With the change in phase, the second column is the sum of (+1) and
(−1), which is equal to 0. We still don't know what the original value of
each pixel was, but we do have a different total value in the Fourier
transform of this second line in k-space when we compare it with the
first line.
Now, let's do a little mathematics. The following is like solving two
equations with two unknowns; solve for a, b, c, and d below:
Let's look back at the pixel values in the phase = 0 line and
1. Designate the pixels in the first column (a and b).
2. Designate the pixels in the fourth column (c and d).
Now look at the pixel values in the 180° phase line, and
1. Designate the pixels in the first column (a and − b).
Note that pixel (a) remains the same as 0 phase line because neither
experiences any phase change, but pixel (−b) is negative because it
experiences a 180° phase shift compared with pixel (b) in 0 phase
line.
2. Designate the pixels in column four as (c and −d).
Again note that pixel (c) remains the same as 0 phase line because
neither experiences any phase change. However, pixel (−d) is
negative because it experiences a 180° phase shift compared with
pixel (d) in phase = 0 line.
a + b = 1 c + d = 2
a − b = 1 c − d = 0
Add 2a = 2 Add 2c = 2
a = 1 c = 1
b = 0 d = 1
Key Points
We have introduced the often intimidating topic of k-
space. k-Space can initially be thought of as the
“data space” (which can be thought of as an “analog”
k-space), with each line in it representing a sampled
version of the received signal (the echo). In the data
space, the coordinates are in time. (Horizontal scale
is on the order of the sampling interval and the
vertical scale is on the order of TR.) The Fourier
transform of k-space is the desired image.
There is, however, one more step that comes after
obtaining the data space and before construction of
the true k-space, having to do with the concept of
“spatial frequencies,” which we shall discuss in
Chapter 16.
Questions
13-1 T/F The number of rows in the data space equals the number of
phaseencoding steps.
13-4 T/F Each row of data space contains one of the received signals
(echoes).
13-5 T/F The axes of the data space are in the time domain.
13-7 T/F The right half of the data (or k) space is the mirror image of
the left half.
13-8 T/F The center of the data space is directly related to the
center of the image.
gradient strength—and not just the phase. This is because the rows in
data space are differentiated by different phase-encoding gradient
strengths Gy.
2½ NEX is a misnomer because we are really halving the number of
Introduction
A pulse sequence diagram (PSD) illustrates the sequence of events that
occur during magnetic resonance imaging (MRI). It is a timing diagram
showing the radio frequency (RF) pulses, gradients, and echoes. Having
a good knowledge of the PSD will help the reader follow complicated
pulse sequences with more ease and understand the interplay among
various scan parameters.
PSD of an SE Sequence
Having been exposed to the concept of gradients, we are now able to
illustrate a complete PSD for, say, a spin-echo (SE) sequence (Fig. 14-1).
Everything in the figure looks like what we have discussed before
except for a few adjustments in
Figure 14-1. A spin-echo PSD.
they come back into phase with the second gradient pulse.)
Figure 14-2. The slice-select gradient Gz is followed by a negative
lobe to refocus the spins.
1b. When the 180° pulse is applied, a sliceselect gradient may or may
not be applied. This is optional and depends on whether a single slice
or multiple slices are being acquired. But before and after the 180°
pulse, we apply a socalled crusher gradient so that the 180° pulse has
a tri-lobed shape (Fig. 14-3). This is just a triviality. When the 180°
pulse is applied, it may have elements that are not exactly 180° leading
to unwanted additional transverse magnetization. This may result in
the echo not focusing at time TE, as we would expect. So these
“crusher” gradients are applied to offset that error. (The first lobe is
used to balance the third; the second is slice selective; the third
destroys the free induction decay [FID] from the unwanted transverse
magnetization.)
2. There is an important adjustment in the readout gradient (Gx). If we
just apply a gradient while we're reading out the echo, we end up
dephasing everything (Fig. 14-4). By the time we get to the middle of
the signal, the signal intensity will be decreased because of the
dephasing caused by the gradient, and by the time we get to the end of
the signal, there will be maximum dephasing and so much signal loss
that there may not be any signal to read!
The answer lies in the fact that in Figure 14-7 the positive pre-readout
Gx gradient comes before the 180° refocusing pulse. After the Gx pre-
readout gradient, we'll have a positive phase difference. This phase
difference will stay constant until the 180° pulse is applied (Fig. 14-8).
After the 180° RF pulse is applied, the phase difference will be
reversed. Then it will remain constant until the Gx gradient is applied.
Then, the spins will begin going back in phase, reaching a zero phase
difference at time TE, and going out of phase afterwards. So, we get
the same thing with both the positive and negative pre-readout Gx
gradients, depending on where in the pulse sequence the gradient is
applied.
Figure 14-7. The first lobe of Gx can be applied either after the
180° as a negative lobe (as in previous figures) or as a positive lobe
prior to the 180° pulse.
Acquisition Time
In previous chapters, we talked about multislice imaging and we said
that there was a lot of “dead” time between the end of the echo and
the next 90° RF pulse (Fig. 14-9). We can use this “dead” time to our
advantage to process other slices.
Figure 14-8. This diagram demonstrates how the spins get back in
phase at the center of the echo when an additional refocusing
positive gradient is applied prior to the 180° pulse.
Each cycle is called an excitation. The term NEX stands for the Number
of Excitations (also known as NSA—Number of Signal Averages). So the
acquisition time depends on
1. TR (the time to do one line of the data space)
2. Ny (the number of phase-encoding steps)
3. NEX (the number of times we repeat the whole sequence)
Acquisition time = (TR) (Ny) (NEX)
This formula is for a conventional SE sequence. Notice that the number
of slices doesn't even enter the equation. This is somewhat
counterintuitive if one is used to the principles of scan time in CT
because, in CT, the more slices we obtain, the longer the sequence will
be. This is not necessarily true in MR because of the fact that we can do
multiple slices within the time of
Key Points
We have discussed the topic of pulse sequence
diagram (PSD) and illustrated one example for SE
imaging. In the chapters to come, we will see
examples of more complicated PSDs. Of course, the
PSD does not tell us all the parameters used in MR
imaging, such as the field of view (discussed in the
next chapter), but it offers an algorithm or
prescription for performing the study.
Questions
14-1 The acquisition time depends on which of the following? (one
or more)
(a) TR
(b) TE
(c) Nx
(d) Ny
(e) NEX
14-2
(a) Calculate the acquisition time for a multiacquisition SE
sequence with TR = 2000, NEX = 2, Ny = 128.
(b) Repeat (a) for single slice acquisition of 10 slices. Is this
practical?
Introduction
A pulse sequence diagram (PSD) provides a timing algorithm for the
sequence of events that is carried out during an MR study. However, the
operator must specify the dimensions of the desired part of the body
under study. This is the subject of this chapter, namely, the field of
view (FOV). As we shall see shortly, there is a limitation as to how small
we can make the FOV, depending on the maximum strength of the
gradients and the bandwidth of the received signals.
The magnetic field along the x-axis is Bx. The value of Bx is given by the
linear equation: Bx = (Gx)x This equation shows that the value of the
magnetic field at any point along the gradient (Gx) is the slope of the
gradient (Gx) times the distance x along the x-axis (Fig. 15-2). Let's
multiply both sides of the equation by the gyromagnetic ratio γ: γ · Bx =
γ(Gx)x
Recall that (γBx) is the Larmor equation, which relates the magnetic
field strength to the frequency: frequencyx = γBx This equation states
that the frequency of oscillation at any point along the x-axis is
proportional to the magnetic field strength at that point, that is, fx =
γBx or fx = γ(Gx)x In other words, the frequency at any point along the
x-axis is proportional to the slope of the gradient (Gx) multiplied by the
position along the x-axis.
Figure 15-1. An image with axes x and y. The frequencyencode
gradient Gx causes the center of the field of view (FOV) to have
magnetic field strength B0 and the right and left ends to have
strengths greater and less than B0, respectively.
Figure 15-2. The gradient Gx describes a linear equation Bx = Gx ·
x. Therefore, at x = 0, no net magnetic field is added to the
system, whereas at a positive value of x, a positive value of
magnetic field is added to the main field.
Let's see what happens at each end of the FOV (Fig. 15-3). At the right-
side end of the FOV (i.e., at x = FOV/2), the frequency is maximum
(call it fmax) because the Gx gradient, and therefore the magnetic field
strength, is maximum.
The formula for frequency is fx = γ(Gx)x Now, for fmax, the distance
along the x-axis is ½ FOV. So, fmax = γ(Gx)FOV/2 Remember this is after
we subtract the center frequency. Therefore, these measurements are
centered around the zero frequency. At the opposite end of the
gradient, we have −fmax: −fmax = −γ(Gx)FOV/2 What is the range of
frequencies? The frequency range is from −fmax to +fmax, that is,
Frequency range = −fmax → +fmax = ±fmax = 2fmax Another term for the
range of frequencies is bandwidth (BW). Thus, BW = ±fmax = 2fmax If we
take the frequency at the right-most side of the image and at the left-
most side of the image, we get the range of frequencies, or the
bandwidth. We already know that, for maximum frequency, fmax =
γ(Gx)FOV/2 Because BW = 2fmax we can conclude that BW = γ · Gx · FOV
We, therefore, see a dependent relationship among the field of view,
bandwidth, and gradient strength. Let's now solve the equation for the
FOV in the x direction:
Top ↓ FOV:
1. ↑ Gradient
2. ↓ Bandwidth
There are limits as to how strong you can make the gradient and there
are also limits as to how low you can make the bandwidth.
Question: What is the minimum FOV possible?
Answer: It is the minimum bandwidth divided by the maximum
gradient:
Gmax and BWmin are specific for each machine. For example, for a GE
Echospeed Plus 1.5-T scanner, maximum gradient strength = 23 mT/m
minimum bandwidth = ±4 kHz = 8 kHz Thus, the minimum FOV is
approximately: 8 kHz/(42.6 MHz/T × 23 mT/m) ≅ 0.8 cm Conversely, to
increase the FOV, we can
1. Increase the BW or
2. Decrease the gradient
To ↑ FOV:
1. ↓ Gradient
2. ↑ Bandwidth
(Remember that by increasing bandwidth, we get decreased signal-to-
noise ratio.)
Key Points
We have discussed the interesting relationship
among the FOV, BW, and gradients: FOV = BW/(γ ·
G) As we saw, there is a limit as to how small one
can make the FOV, depending on the minimum
allowable BW and the maximum possible gradient
strength:
Questions
15-1 If the minimum FOV = 30 cm for a frequency-encoding gradient
Gx = 5 mT/m, what would the minimum FOV be for a stronger Gx =
10 mT/m? (i.e., does a stronger Gx reduce or increase the minimum
FOV?)
Introduction
This chapter will summarize some concepts that we've already
discussed and clarify some of the fine points of k-space. Remember that
we have, up to this point, referred to the data space as an “analog” k-
space, and we said that the Fourier transform of the data space is the
image (Fig. 16-1).
This is, in fact, correct, but there is a problem with this concept: the
matrix in the data space is very asymmetric. In the frequency-encoding
direction, the interval between two samples (i.e., the sampling interval
ΔTs) is on the order of microseconds, so that the total time to take all
the samples (i.e., the sampling time Ts) is on the order of milliseconds.
The time intervals in the phaseencoding direction, however, are each
on the order of one TR (i.e., on the order of seconds). The total time to
obtain all the data in the phaseencoding direction is the scan time for
one acquisition (on the order of minutes).
Figure 16-1. The data space (with axes ΔTs and TR) is in the time
domain. k-Space (with axes Δky) is in the spatial frequency domain
and is derived from the data space. The Fourier transform of k-
space is the image (with axes x and y), which is in the frequency
domain.
Thus, in the data space we have a matrix whose x-axis is on the order
of milliseconds and whose y-axis is on the order of minutes. This would
give us a very asymmetric matrix. The true kspace is the same matrix
as the data space, but with a different scale. Recall that
In the last chapter, we talked about the field of view (FOV). We derived
Equation 16-1, showing the relationship among the FOV, the bandwidth
(BW), and the gradient strength. According to this formula, the FOV is
equal to the BW divided by the product of γ and G. We also know from
a
and
we can derive a new formula for the FOV:
Now consider the FOV in the x and y directions. If we consider the FOV
in the x direction, the formula tells us that we need to take the
gradient strength and the sampling interval in the x direction.
Therefore,
Key Points
The true k-space is a mathematically manipulated
variant of the data space, with axes referred to as
“spatial frequencies.” Therefore, k-space is in a
“spatial” frequency domain. Spatial frequencies kX
and kY are inversely proportional to distance (with
units of cycles/cm). The Fourier transform of k-space
is the desired image.
The spatial frequencies kX and kY are expressed as:
kx = γ · Gx · tx ky = γ · Gy · ty with units in cycles/cm.
Questions
16-1 T/F Spatial frequencies have units 1/distance (cycles/cm).
16-2 T/F
(a) The axes in k-space are designated kx and ky.
(b) The axes in k-space are in the frequency domain (with units
1/time or cycle/sec).
16-3 T/F The Fourier transform of the k-space produces the desired
image.
16-5 T/F
(a) The center of k-space contributes to maximum image
contrast.
(b) The periphery of k-space contributes to image details.
Introduction
In this chapter, we will discuss all the important parameters in MR
imaging that the operator can control and adjust. We will then see how
these changes influence the image quality. Every radiologist is
comfortable with a particular set of techniques; therefore, “custom-
made” techniques can be achieved only if the radiologist is aware of
the parameters and trade-offs involved.
which makes sense because Ny × NEX × Ts is the total time the machine
is “listening” to the spin echoes.
Since Ts = Nx · ΔTs and ΔTs = 1/BW, then Ts = Nx/BW.
Voxel Volume
If we increase the voxel size, we increase the number of proton spins in
the voxel and, therefore, increase the signal coming out of the voxel
(Fig. 17-1). The voxel volume is given by Voxel volume = Δx · Δy · Δz
where Δx = pixel size in the x direction, Δy = pixel size in the y
direction, and Δz = slice thickness.
NEX (Number of Excitations or Acquisitions). NEX stands for the
number of times the scan is repeated. Let's say we have two signals (S1
and S2), corresponding to the same slice (with the same Gy). There is
constant noise (N) associated with each signal (N1 = N2 = N). If we add
up the signals (assuming S1 = S2 = S), we get S1 + S2 = 2S However, if we
add up the noise, we get
This formula does not make sense at first glance. Why do we get N
and not 2N? The answer has to do with a somewhat complicated
statistical concept and the so-called random Brownian motion theory,
which deals with the spectral density of the noise.
This is where the factor comes from. However, you don't need to
know the underlying math—you just need to understand the concept. In
summary:
The resulting signal will be twice the original signal. The resulting
noise, however, will be less—it will be the square root of 2 multiplied by
the noise N.
In other words, if we increase the number of acquisitions by a factor of
2, the signal doubles and noise increases by , for a net 2/ = ;
thus, SNR increases by a factor of .
Therefore, ↑ NEX by a factor of 2 → ↑ SNR by a factor of .
Think of the NEX as an averaging operation that causes “smoothing”
and improvement in the image quality by increasing the signal to a
greater degree (e.g., factor 2) relative to the increase in the noise
(e.g., factor ). As another example, increasing NEX by a factor of 4
results in an increase of signal by 4 and an increase of noise by or 2.
Thus, SNR increases by 4/2 or twofold.
Ny (Number of Phase-Encoding Steps). The same concept holds for Ny.
That is, similar to NEX, there is a 41% ( ) increase in SNR when Ny is
doubled. As with NEX, when the number of phase-encode steps
doubles, signal doubles and noise increases (randomly) by (for a net
increase in SNR).
Bandwidth. An inverse relationship exists between BW and SNR. If we
go to a wider bandwidth, we include more noise, and the SNR
decreases. If we decrease the bandwidth, we
SNR in 3D Imaging
In 3D imaging, we have the same factors contributing to SNR, plus an
additional phaseencoding step in the z direction (Nz):
From this equation, you can see why SNR in 3D imaging is higher than
that in 2D imaging. Specifically,
Another way to look at SNR is to say that SNR depends on only two
factors:
1. Voxel size
2. Total sampling time
Sampling time (Ts) is the time that we sample the signal. Therefore, it
makes sense that the more time we spend sampling the signal, the
higher the SNR will be. Let's look again at the formula for SNR (in 2D
imaging):
so
Example
For FOV = 250 mm and a 256 × 256 matrix Nx = Ny = 256 Pixel size
(x) = FOVx/Nx = 250/256 ≅ 1 mm in x direction. Pixel size (y) =
FOVy/Ny = 250/256 ≅ 1 mm in y direction. In the x direction, there
are two ways of increasing resolution (for a given FOV):
1. Increase Nx by reducing the sampling interval ΔTs (i.e., by increasing
the BW) and keeping the sampling time Ts fixed (recall that Ts = Nx ·
ΔTs). The advantage here is no increase in TE; the trade-off is a
reduction in SNR (due to increased BW).
2. Increase Nx by lengthening Ts and keeping ΔTs (and thus BW) fixed.
Here, the SNR does not change, but the trade-off is an increased TE
(due to a longer Ts) and less T1 weighting (this is only a concern in
short echo delay time imaging).
Acquisition Time
The acquisition time or scan time, as we have seen previously, is given
by Scan time = TR · Ny · NEX where Ny is the number of phase-encoding
steps (in the y direction).
For fast spin-echo (FSE) imaging (discussed in detail in Chapter 19), the
above is modified to FSE time = TR · Ny · NEX/ETL where ETL = echo
train length (4, 8, 16, 32).
For 3D imaging, the scan time is given by Time (3D) = TR · Ny · Nz · NEX
where Nz is the number of phase-encoding steps (partitions) in the z
direction. In other words, Time (3D) = Nz · time(2D) Multiplication by
such a large number (e.g., Nz = 32 to 64 or 128) might at first seem to
result in an excessively long scan time for 3D imaging, but the TR used
in 3D gradient-echo imaging is approximately 100 times smaller (order
of 30 msec) compared with the TR used in conventional spin-echo
imaging; we can perform a 3D scan in a reasonable time. Recently, 3D
FSE imaging (discussed in Chapter 19) has also become feasible.
Example
1. Calculate the acquisition time of an SE sequence with TR = 3000
msec, Ny = 256, and NEX = 1.
Solution: Scan time = 3000 × 256 msec = 768 sec = 12.8 min
2. Calculate the acquisition time of an FSE sequence with the above
parameters and an ETL of 8
3.
a. Calculate the acquisition time of a 3D gradient-echo sequence with
TR = 30 msec, Ny = 256, NEX = 1, and Nz = 60.
Solution: Scan time = 30 × 256 × 1 × 60 msec = 460.8 sec = 7.68 min
b. If TR = 300 in the previous example, then the scan time = 76.8 min
= 1 hr and 16.8 min, which is, obviously, impractical. Hence, 3D
techniques use gradient-echo sequences employing a very short TR.
TE (Time to Echo)
Question: What happens if we increase or decrease TE?
Answer:
1. By increasing TE, we:
(a) increase T2 weighting
(b) increase dephasing and thus decrease SNR (according to the T2
decay curve)
(c) decrease number of possible slices (decrease coverage), because
number of slices ≈ TR/TE
(d) no change in scan time (unless, of course, the coverage is not
adequate and either longer TR or extra acquisitions are required)
2. The reverse is true for decreasing TE:
(a) decrease T2 weighting and increase T1 or proton density weighting
(b) increase SNR (less dephasing). However, if TE is reduced by reducing
Ts (i.e., increasing BW), SNR may be reduced!
(c) increase coverage
(d) no change in scan time
Question: What causes lengthening of the minimum TE?
Answer:
1. TE should be long enough so that the side lobes of the 180° pulse do
not interfere with the side lobes of the FID or the echo (Fig. 17-3).
Remember that we need a Fourier transform of the RF pulse with a
square shape to be able to get ideal contiguous slices. To do this, the
RF must be a sinc wave (sinc t = sin t/t) with as many side lobes as
possible. This, in turn, will lengthen the 90° and 180° RF pulse.
2. If TE is so short that it allows interference between the 180° RF
pulse and the FID, an
FID artifact (or zipper artifact) will appear along the zero frequency
line.
Figure 17-3. To avoid overlapping of the FID and the side lobes of
the 180° pulse, you need to increase TE. This increase is one cause
of lengthening the minimum TE.
TI (Inversion Time)
As we saw in Chapter 7, inversion recovery sequences employ an
additional 180° pulse before the 90° pulse.
Table 17-1
TR TE
Advantages
1. Can suppress various tissues by selecting the appropriate TI. More
specifically, as we saw in Chapter 7, if TI = 0.693 × T1 (tissue x) then
tissue x is “nulled” or “suppressed.”
2. STIR (short TI inversion recovery) sequences suppress fat by selecting
TI = 0.693 × T1 (fat) Since at 1.5 Tesla, T1 of fat is approximately 200
msec, then to null fat, we must select TI = 0.693 × 200 ≅ 140 msec
3. FLAIR (fluid-attenuated inversion recovery) sequences suppress fluid
by selecting TI = 0.693 × T1 (fluid)
This sequence is used, for example, in the brain to suppress
cerebrospinal fluid (CSF) to increase the conspicuity of periventricular
hyperintense lesions such as multiple sclerosis plaques. Since at 1.5 T,
T1 of CSF is approximately 3600 msec, then to null CSF, we have to
select TI = 0.693 × 3600 ≅ 2500 msec
Disadvantages
1. Decreased SNR
2. Decreased coverage (by a factor of about 2 due to the presence of
the extra 180° pulse)
Key Points
In this chapter, we discussed the important and
practical factors that influence the quality of MR
imaging. To improve the quality of the images, it is
crucial to have a firm grasp of the parameters that,
directly or indirectly, affect the scan. We introduced
the primary and secondary parameters that are used
to determine MR images (refer to the Introduction in
this chapter). In a nutshell, the name of the game is
“trade-offs.” Often, one cannot gain advantage in one
area without sacrificing another.
Questions
17-1 For a TR = 1500 msec, 2 NEX, and a 128 × 128 matrix, calculate
the scan time for
(a) a single slice
(b) 10 slices (performed one at a time)
(c) 10 slices performed using a multislice (multiplanar)
acquisition
(b)
(c) Ny
(d)
17-9 For a 128 square matrix and an FOV of 25 cm, the pixel size is
about
(a) 0.5 mm
(b) 1 mm
(c) 1.5 mm
(d) 2 mm
(b)
(c) Ny
(d)
Introduction
MRI, as with any other imaging modality, has its share of artifacts.
It is important to recognize these artifacts and to have the tools to
eliminate or at least minimize them. There are many sources of
artifacts in MRI. These are summarized as follows:
1. Image processing artifact
a. Aliasing
b. Chemical shift
c. Truncation
d. Partial volume
2. Patient-related artifacts
a. Motion artifacts
b. Magic angle
3. Radio frequency (RF) related artifacts
a. Cross-talk
b. Zipper artifacts
c. RF feedthrough
d. RF noise
4. External magnetic field artifacts
a. Magnetic inhomogeneity
5. Magnetic susceptibility artifacts
a. Diamagnetic, paramagnetic, and ferromagnetic
b. Metal
6. Gradient-related artifacts:
a. Eddy currents
b. Nonlinearity
c. Geometric distortion
7. Errors in the data
8. Flow-related artifacts
9. Dielectric effects
Let's discuss this list in more detail.
Figure 18-1. For a given FOV and gradient strength, the maximum
frequency fmax corresponds to the edges of the FOV. Any part
outside the FOV will experience a higher frequency. The higher
frequencies outside the FOV may be aliased to a lower frequency
inside the FOV. This will cause a wraparound artifact.
For example, if the higher frequency were 2 kHz higher than (fmax), it
would be recognized as 2 kHz higher than (−fmax), and therefore its
information would be “aliased” to the opposite side of the image—the
side of the FOV that corresponds to the lowest frequencies (Fig. 18-1).
The part of the body and arm on the left side of the patient that is
outside the FOV and is exposed to a higher magnetic field will have
spins oscillating at a frequency higher than (fmax). Thus, it will be
identified as a structure on the right side of the patient—that side of
the image associated with lower frequencies.
Likewise, the arm and body outside the FOV on the right side of the
patient will experience spins oscillating at frequencies lower than
(−fmax) and will also be incorrectly recognized by the computer. For
example, if the lower frequency were 2 kHz lower than (−fmax), it
would be recognized as 2 kHz lower than (fmax), and its information
would be “aliased” to the opposite side of the image—the side of the
FOV that corresponds to the higher frequencies. This process is also
called wraparound—the patient's arm gets “wrapped around” to the
opposite side.
The computer can't recognize frequencies outside the bandwidth (which
determines the FOV). Any frequency outside of this frequency range is
going to get “aliased” to a frequency that exists within the bandwidth.
The “perceived” frequency will be the actual frequency minus twice
the Nyquist frequency.
f (perceived) = f (true) −2f (Nyquist)
Why then do we usually see wraparound in the phase-encoding
direction? Remember that the number of phase-encoding steps is
directly related to the length of the scan time. The phase-encoding
steps can be lowered by shortening the FOV in the phase-encoding
direction versus the frequencyencoding direction also known as
rectangular FOV (see Chapter 23). If the FOV is shortened too much in
this direction versus the actual extent of the body then wraparound will
occur. Figure 18-2 contains an example of wraparound.
3D Imaging. Wraparound artifact can also be seen in 3D imaging in all
three directions.
1. It can be seen along the x and y directions, as with spin-echo
imaging.
2. It can also be seen along the slice-select (phase-encoded) direction
at each end of the slab (e.g., the last slice is overlapped on the first
slice, as in Figs. 18-3 and 18-4).
Example
Suppose the frequency bandwidth is 32 kHz (±16 kHz). This means
that if we're centered at zero frequency, the maximum frequency
fmax = +16 kHz and minimum frequency (−fmax) = −16 kHz (Fig. 18-
1). If we have a frequency in the arm (outside
Figure 18-2. Axial T1 (A) and PD (B) images of the lumbar spine
demonstrate aliasing of the arms in Figure B (arrows). Figure B was
obtained with a smaller FOV resulting in aliasing artifact. Also note
that the patient has a filum terminale lipoma (black arrows in A
and B).
Figure 18-3. 3D gradient-echo T1 image without gadolinium of the
abdomen shows slice direction aliasing with the kidneys appearing
to be in the lungs (arrows). Also note that there is aliasing in the
phase-encoding direction (anteroposterior) from the inferior
abdominal image's anterior subcutaneous tissue “wrapping around”
posteriorly (arrowheads).
if we use a coil that only covers the area within the FOV, we will only
get signal from those body parts within the maximum frequency
range, and no aliasing will result. This type of coil is called a surface
coil. We also use a surface coil to increase the signal-to-noise ratio
(SNR).
Figure 18-4. 3D coronal gradient-echo T1 image of the brain
shows slice direction aliasing of the anterior skull on to the brain.
Figure 18-5. To avoid aliasing, increase the FOV.
H2O. Actually, the protons in H2O precess slightly faster than those in
fat. This difference is only 3.5 ppm. Let's see what this means by an
example.
Figure 18-6. In no phase wrap, aliasing is avoided by doubling the
FOV in the y direction and, at the end, discarding the unwanted
part of the image.
Figure 18-7. Sagittal STIR image (A) of the cervical spine with
craniocaudal phase-encoding direction demonstrates aliasing of the
brain onto the upper thoracic spine. (B) The same image after no
phase wrap was applied. Truncation artifact is also seen (arrows).
Example
Consider a 1.5-T magnet. The precessional frequency is as follows:
Example
We now have a 0.5-T magnet. The precessional frequency of protons
in a 0.5-T magnet is 1/3 of a 1.5-T magnet. The frequency difference
is then
1/3(220 Hz) = 73 Hz
Therefore, at 0.5 T, the difference in precessional frequency of the
hydrogen protons in fat and in H2O is only 73 Hz. In other words, if
we use a weaker magnet, we will get less chemical shift.
How does this affect the image? Chemical shift artifacts are seen in the
orbits, along vertebral endplates, in the abdomen (at organ/fat
interfaces), and anywhere else fatty structures abut watery structures.
In a 1.5-T magnet, the sampling time (Ts) is usually about 8 msec. Let's
take 256 frequency points in the frequencyencoding direction.
BW = N/TS = 256/8 msec BW = 32 kHz
These formulas show that the entire frequency range (i.e., bandwidth)
of 32 kHz covers the whole length of the image in the x direction.
Because we have the FOV of the image in the x direction divided into
256 pixels, each pixel is going to have a frequency range of its own,
that is, each pixel has its own BW:
BW/pixel = 32 kHz/256 BW/pixel = 125 Hz
(This representation of BW on a “per pixel” basis is used by Siemens
and Philips. It has the advantage of less ambiguity than the ±16 kHz
designation, should the “±” be deleted.) Thus, each pixel contains 125
Hz of information (Fig. 18-8). Stated differently, the pixel bin contains
125 Hz of frequencies. Now, because fat and H2O differ in the
precessional frequency of hydrogen by 220 Hz at 1.5 T, how many pixels
does this difference correspond to?
Pixel difference = 220 Hz/125 Hz/pixel ≈ 2 pixels
Figure 18-8. At 1.5 T with a BW of 32 kHz and 256 pixels, there
will be about 125 Hz/pixel (32 kHz/256 = 125 Hz), that is, there is
125 Hz of information in each pixel. This may be a better way of
describing the BW of a scanner because there is no ± confusion.
This means that fat and H2O protons are going to be misregistered from
one another by about 2 pixels (in a 1.5-T magnet using a standard ±16
kHz bandwidth). (Actually, it is fat that is misregistered because
position is determined by assuming the resonance property of water.) If
pixel size δx = 1 mm, this then translates into 2 mm misregistration of
fat.
MATH: For the mathematically interested reader, it can be shown that
Chemical shift
Example—Vertebral Bodies
With frequency-encoding direction—in this case, going up and down
(and “up” having higher frequency) —the fat in the vertebral body
would be misregistered down, making the lower endplate
bright due to overlap of water and fat and the top endplate dark due
to water alone (Fig. 18-10). If we increase the pixel size, the
misregistration artifact will increase.
out of phase every 2.25 msec. This is called a chemical shift effect of
the second kind.
Boundary Effect. If the selected TE is 2.25, 6.75, 11.25, 15.75 msec,
and so on, fat and water protons will be out of phase and a dark
boundary will be seen around organs that are surrounded by fat (such as
the kidneys and muscles). This result is called the boundary effect,
bounce point artifact, or India Ink Etching, which is caused by
chemical shift of the second kind. This type of imaging is referred to as
out-of-phase scanning, referring to the fact that at these TEs, fat and
water spins will be 180° out of phase. This phenomenon does not just
occur along the frequency-encoding axis (as with the chemical shift
artifact of the first kind) because it is a result of fat and water protons
phase cancellation in all directions (Fig. 18-17). (Boundary effect does
not occur in conventional SE techniques because of the presence of the
180° refocusing pulse, which is absent in GRE techniques.)
Figure 18-17. In-phase (A) and out-of-phase (B) spoiled gradient
T1 images show the “boundary effect” on the out-of-phase images
circumferentially at every fat/water interface (arrows in B). Also
note that a left adrenal adenoma loses signal substantially on the
out-of-phase image (arrowhead).
Remedy
1. Make fat and H2O in phase by picking appropriate TE.
2. Increase the BW (trade-off: decreases SNR).
3. Use fat suppression.
Truncation Artifact (Gibbs Phenomenon). This artifact occurs at high
contrast interfaces (e.g., skull/brain, cord/cerebrospinal fluid (CSF),
meniscus/fluid in the knee) and causes alternating bright and dark
bands that may be mistaken for lesions (e.g., pseudo syrinx of the
spinal cord or pseudo tear of the knee meniscus).
The cause is inability to approximate exactly a steplike change in the
signal intensity due to a limited number of samples or sampling time.
The ripples in Figure 18-18 are responsible for
the parallel bands seen at such sharp interfaces. This artifact is seen
mostly in the phase direction (because we typically have few pixels and
lower resolution in phase compared with frequency). Incidentally, the
correct term is “truncation artifact.” “Gibbs phenomenon” refers to
the infinitely thin discontinuity that still persists with an infinite
number of pixel elements. Figures 18-7, 18-19, and 18-20 contain
examples of truncation artifact.
Figure 18-18. Truncation artifact causes a ring-down effect
because the FT of a truncated sinc function has ripples at its
edges.
Remedy
1. Increase sampling time (↓ BW) to reduce the ripples. (Remember, a
wider signal in time domain means a narrower one in frequency
domain.)
Patient-Related Artifact
This artifact is caused by voluntary or involuntary patient motion and
by the patient's anatomy. Pulsating motion in vessels is also an
Example
The aorta pulsates according to the heart rate. If the heart rate is
HR = 60 beats/min = 60 bpm = 1 beat/sec
then the period of motion = T(motion) = 1 sec.
Figure 18-23. Axial STIR image of the neck showed marked
ghosting from both arteries and veins in the phase-encoding
direction (anteroposterior).
Remedy
1. Use spatial presaturation pulses to saturate inflowing protons and
reduce the artifacts.
Figure 18-24. Axial spoiled gradient image of the abdomen shows
both positive (black arrow) and negative (white arrow) “ghost”
artifacts from the aorta. The negative “ghost” mimics a vertebral
body lesion.
Figure 18-25. Axial T2 image of the thoracic spine shows positive
“ghost” artifacts from the CSF that simulate lung nodules.
Remedy
1. Patient instruction: Don't move! (probably the most useful remedy)
2. Respiratory compensation (RC) (uses chest wall motion pattern to
reorder scan and minimize motion)
3. Use of glucagon in the abdomen to reduce artifacts due to bowel
peristalsis
4. Sedation
5. Pain killers
6. Faster scanning (FSE, GRE, EPI, etc.); sequential 2D rather than 3D
scanning (see Fig. 18-27 for an example).
CSF Flow Effects. Dephasing of protons due to CSF motion may
sometimes simulate a lesion. Flow compensation techniques can reduce
this effect. Examples include the following:
1. Pseudo aneurysm of basilar artery due to pulsatile radial motion of
CSF around it (Fig. 18-28).
2. Pseudo MS plaques in the brainstem due to CSF flow in the basal
cisterns.
3. Pseudo disc herniation, again secondary to CSF flow.
Remedy
1. Be certain that “lesions” are seen on all pulse sequences (artifacts
tend to only be seen on one image).
2. Use cardiac gating.
3. Use flow compensation.
Figure 18-28. Axial T2 image shows marked signal void around the
basilar artery mimicking an aneurysm in this 3-year-old patient.
Figure 18-29. Sagittal PD (A) and T2 (B) fat-saturated images of
the knee show magic angle artifact as seen by increased signal on
the short TE PD image (arrow in A), whereas the tendon itself is
not thickened and has dark signal on the T2 image (B). Joint
effusion is also seen. (Courtesy of D. Beall, MD, San Antonio,
Texas.)
RF-Related Artifacts
Cross-talk. We have already discussed this issue in previous chapters.
The problem arises from the fact that the Fourier transform (FT) of the
RF pulse is not a perfect rectangle but rather has side lobes (Fig. 18-
31). We shall use a simpler version of the RF profile, as in Figure 18-32.
If we consider two adjacent slices, there will be an overlap in the FT of
their RF pulses (Fig. 18-32). Cross-talk causes the effective TR per slice
to decrease (due to saturation of protons by the RF signals for adjacent
slices). Thus, more
T1 weighting will result (this is particularly problematic for PD- and T2-
weighted images). Also, due to reduced effective TR, the SNR will
decrease.
Figure 18-30. Angled sagittal PD (A) and T2 (B) fat-saturated
images of the shoulder demonstrate magic angle artifact of the
intra-articular biceps tendon. There is increased signal on the PD
image (A), whereas the tendon has dark signal and overall normal
appearance on the T2 image. Acromioclavicular joint high signal is
from osteoarthritis changes. (Courtesy of D. Beall, MD, San
Antonio, Texas.)
Figure 18-31. The actual RF has a finite time span, yielding side
lobes or rings. A Gaussian RF pulse has a Gaussian FT.
Figure 18-34. The larger the interslice gap is, the less cross-talk is
observed.
Remedy
1. Gaps can be introduced between adjacent slices (Fig. 18-33).
2. Two acquisitions with 100% gaps can be interleaved.
3. The RF pulse can be lengthened to achieve a more rectangular pulse
profile.
Let's discuss these in more detail:
1. If we increase the gap between slices, we reduce cross-talk (Fig.
18-34). The tradeoff is an increase in the unsampled volume and
the increased potential for missing a small lesion located within
the gap.
2. It doesn't matter which way we order the slices (we can do slice 1,
then slice 3, then slice 2, etc.). Adjacent slices still will be sharing
a certain frequency range and cause cross-talk. The only way to
eliminate cross-talk is to do two separate sequences each with a
100% gap, such as:
Figure 18-35. The closer the profile of the RF pulse (actually its
FT) is to a rectangle, the better we can achieve contiguous slices
without encountering cross-talk.
Remedy
1. Increase the TE (increases the separation between the FID and the
180° RF pulse).
2. Increase slice thickness (δz). This in effect results from selecting a
wide RF BW, which narrows the RF signal in the time domain, thus
lowering chances for overlap.
Stimulated Echo. This artifact also appears as a narrow- or wide-band
noise in the center along the frequency-encoding axis. The mechanism
is similar to FID artifacts. In this case, imperfect RF pulses of adjacent
slices or imperfect 90° −180° −180° pulses of a dual-echo sequence
form a stimulated echo that may not be phase-encoded, thus appearing
in the central line along the frequency-encoding axis.
Figure 18-37. FID artifact. The side lobes of the 180° and the FID
may overlap, causing a zipper artifact at zero frequency along the
phase direction.
Figure 18-38. RF feedthrough causes a zipper artifact at zero
frequency along the phase direction.
Remedy
1. Use spoiler gradients.
2. Adjust the transmitter.
3. Call the service engineer.
RF Feedthrough Zipper Artifact. This artifact occurs when the
excitation RF pulse is not completely gated off during data acquisition
and “feeds” through the receiver coil. It appears as a “zipper” stripe
along the phase-encoding axis at zero frequency (Fig. 18-38).
Remedy. Alternate the phase of the excitation RF pulses by 180° on
successive acquisitions; the averaged phase-alternated excitations will
essentially eliminate RF feedthrough.
RF Noise. RF noise is caused by unwanted external RF noise (e.g., TV
channel, a radio station, a flickering fluorescent light, patient
electronic monitoring equipment). It is similar to RF feedthrough
except that it occurs at the specific frequency (or frequencies) of the
unwanted RF pulse(s) rather than at zero frequency (Fig. 18-39).
Remedy
1. Improve RF shielding.
2. Remove monitoring devices if possible.
3. Shut the door of the magnet room!
Gradient-Related Artifacts
Eddy Currents. Eddy currents are small electric currents that are
generated when the gradients are rapidly switched on and off (i.e., the
resulting sudden rises and falls in the magnetic field produce electric
currents). These currents will result in a distortion in the gradient
profile (Fig. 18-48) and in turn cause artifacts in the image.
Nonlinearities. Ideal gradients are linear. However, as in other aspects
of life, there is no such thing as an ideal gradient. These nonlinearities
cause local magnetic distortions and image
Remedy
1. Delete the discrete error and average out the neighboring data.
2. Simply repeating the sequence solves the problem.
Figure 18-51. Axial T2-weighted image of the cervical spine on a
0.23 T magnet shows diagonal lines coursing throughout the image
related to a single data point error in k-space.
Flow-Related Artifacts
Motion artifacts were discussed previously, including periodic flow
artifacts. Other flow-related phenomena are discussed in Chapters 26
and 27.
Dielectric Effects
As the wavelength of the radiowave approaches the dimensions of the
body part being imaged, there can be areas of brightening and
darkening due to standing waves. This is most pronounced at 3 T and
above. Since the body is a conducting medium, the artifact is often
called “dielectric effect” (Fig. 18-52). It seems to be worse in large
body parts, that is, the abdomen, and seems to be quite common when
ascites is present. The solution for dielectric effects is parallel
transmission or “transmit SENSE.”
Questions
18-1 Regarding chemical shift artifact:
(a) protons in fat resonate at 3.5 ppm higher than protons in
water
(b) at 1.5 T it is about 220 Hz
(c) at 1.5 T and for a 32 kHz BW and 256 × 256 matrix, it is about
2 pixels
(d) all of the above
(e) only (b) and (c)
18-2
(a) Determine the chemical shifts (in terms of numbers of pixels)
for the following situations (assume 256 frequency-encoding
steps):
BW
50 kHz
10 kHz
4 kHz
18-8 T/F Fat and water protons get out of phase at TE of odd
multiples of 2.25 msec.
18-10
(a) Calculate the separation (in pixels and mm) between aortic
ghosts for TR 500 msec, NEX 1, Ny 128, HR 80 bpm, and FOV 20
cm.
(b) What is the maximum number of ghosts you could potentially
see within the FOV?
18-13 CSF flow can lead to all of the following artifacts except:
(a) pseudo MS plaques in the brainstem
(b) pseudo disc herniation
(c) pseudo basilar artery aneurysm
(d) pseudo syrinx
Introduction
In this chapter, we will discuss the elegant and cunning technique of
fast spin echo. This technique was first proposed by Hennig et al.1 and
was called RARE (rapid acquisition with relaxation enhancement).
However, it is commonly referred to as fast spin echo (FSE) or turbo
spin echo (TSE). Different manufacturers have different names for it
(Table 19-1).
Consider the pulse sequence diagram in Figure 19-1. This pulse
sequence can be used for either a CSE or an FSE study.
Table 19-1
Manufacturer Name
get to the last echo anyway. In a dual-echo CSE sequence, the first echo
is always “free”—it doesn't cost any time. (However, as we shall see
later, that isn't true with FSE.)
Figure 19-1. A spin-echo PSD with eight echoes. The echo spacing
(ESP) is 17 msec.
Thus, for each k-space in CSE, we repeat the TR 256 times (each at a
different phase-encoding gradient) and fill in the k-space for each echo
with 256 different lines. For an eight-echo train, we get eight different
images.
Figure 19-2. In FSE, k-space is filled eight lines at a time in one
shot (within one TR).
k-spaces, one for each echo, we will have one k-space using the data
from all eight echoes.
Figure 19-3. After two shots, 16 lines in k-space will be filled.
Within the time of one TR (one “shot”), we will get eight lines, one
from each echo, in the single k-space. With the next TR, we'll
accumulate eight more lines, one from each echo, and we'll also put
them into the same k-space (Fig. 19-3).
For each shot/TR, we will fill in another eight lines into the single k-
space. Because we have a total of 256 lines in k-space, and because
during each TR we are filling in eight lines of k-space at a shot, we only
have to repeat the process 32 times (i.e., 256/8 = 32) to fill 256 lines of
k-space.
In CSE, it took one TR for each line of k-space. Therefore, in a CSE
study, we have to repeat the TR 256 times. In this manner, we have cut
the time of the study by a factor of eight.
Example
Acquisition time of CSE sequence with TR = 3000 msec, Ny = 256, NEX =
1 is
SE time = (TR)(number of phase-encoding steps)(NEX)
= (3000)(256)(1)msec
= 12.8 min
This time for the FSE sequence with TR = 3000 msec is
FSE time = (TR)(number of phase-encoding steps/ETL)(NEX)
= (3000)(256/8)(1)msec
= 1.6 min
In this example, we have shortened the time of the study from 12.8 min
to 1.6 min, a factor of eight times faster than the CSE study.
In FSE, before each 180° pulse, we place a different value of the phase-
encoding gradient. For the 180° pulse before the echo we choose as the
TEeff (in this case, 102 msec), we use a phaseencoding gradient with
the lowest strength. Each subsequent phase-encoding step will have a
gradient with more and more amplitude. This increase will result in the
most signal coming from the echo at 102 msec (because this signal is
obtained with a minimum phase gradient) and
decreasing signal from all the other echoes because their signals are
obtained with increasing phase-encoding gradients (Fig. 19-6).
Figure 19-5. The 32 lines in the center of k-space correspond to
the echoes associated with the weakest phase gradients.
Figure 19-6. The echo corresponding to TEeff of 102 msec has the
largest peak.
In the next TR, we again pick a phase-encoding step for the sixth echo
that will be close to zero gradient and all the other phase-encoding
steps will be close to their previous respective gradient values, so that
again the maximum signal during this TR will be obtained from echo 6
(TE = 102 msec) and progressively weaker signals will be obtained from
the other echoes.
The signals from the sixth echo (from the 1st to the 32nd TR) will all be
placed in the center of k-space (Fig. 19-5). The signals from the other
echoes will be placed in the other slabs. The echoes that experience
progressively greater phase-encoding gradients (and therefore less
signal) fall into slabs further away from the center slab, and those
echoes experiencing the weakest phase-encoding gradients (and
therefore having more signal) are placed closer to the center slab. k-
space is organized so that the greatest amount of signal comes from the
center of k-space and the least amount of signal comes from the
periphery of k-space.
Therefore, if we choose an ETL of 8 echoes, we will have 8 slabs in k-
space, with each slab containing 32 lines from 32 shots. Echo slab
corresponds to a different echo. Let's see what the echo looks like (Fig.
19-6).
Because the center slab belongs to the lowest phase-encoding gradient,
it will have the least amount of dephasing. The signal received at TE
=102 msec will have the greatest amplitude. As we go farther away
from 102 msec in either direction, the signal amplitude will get
progressively smaller because the phase-encoding gradients get
progressively larger.
By definition, the maximum signal comes at the TEeff time. But we still
get echoes from the other TEs that do not help our contrast. The
signals from these other echoes are all in the same k-space. Even
though the respective signal amplitudes from the other TEs are
progressively smaller, the further away in time they are from the TEeff;
they still contribute to the contrast from the TEeff. This is why it is
called an effective TE and not a true TE.
In a way, what we are doing is averaging the echoes, although it is a
weighted average. By appropriately picking the slabs, we put most of
the weight on the echo corresponding to 102 msec (the TEeff) and less
weight on the other echoes. As we go away from the center slab of k-
space, we are reducing the weighting, that is, we are reducing the
contribution of the data on that slab to the effective echo.
The previous example gives us a long TR/long TE, which is a T2-
weighted image. Now we want to do a proton density-weighted image
with a long TR/short TE and a TEeff of approximately 30 msec (Fig. 19-
7). In this case, we would assign the center slab of k-space to
correspond to the second echo, that is, TE =
Remember that with the TEeff of 34 msec, we are still getting the
cumulative signal from the entire echo train of eight echoes. So even
information from an echo corresponding to a TE of 136 msec (8 × 17 =
136) is contributing to the signal of the TEeff of 34 msec, which we don't
want. Therefore, for a T1-weighted study, we usually pick a smaller ETL
such as 4. With an ETL of 4, we would only do four phase-encoding
steps. Thus, k-space would only have four slabs, and the longest echo
contributing to the contrast of the shortest TEeff (e.g., 17 msec) would
be the echo at 68 msec (4 × 17 = 68). This will eliminate the T2 effect
on a T1-weighted image caused by the signal contribution of the longer
echo times.
Question: What happens to the time it takes to do an FSE study when
we decrease the ETL from 8 to 4?
Answer: The time of the study will be increased by a factor of 2.
Question: If we have an ETL of 8, how many lines of k-space do we fill
in with each TR?
Answer: We fill in eight lines of k-space with each TR, each line going
to a separate slab within k-space.
Question: How many times do we have to repeat the TR to fill up k-
space with 256 lines?
Answer: In general, we have to repeat it by the following ratio:
Trade-offs
What are the trade-offs of FSE imaging?
Slice Coverage. As we increase the ETL to increase the speed of the
exam, we also decrease the number of slices we can do in a study (Fig.
19-8). In one TR, with eight ETL, we can fit in so many slices. If we use
the same TR, but now use an ETL of 16, it will take twice as long to
receive the echo (i.e., 16 × 17 msec = 272 msec) because now we have
to accumulate data from 16 echoes (each a multiple of 17 msec).
Because the time it takes to accumulate 16 lines of k-space is double
the time it takes to accumulate eight lines, we can fit in only half the
number of slices into one TR.
Therefore, the trade-off is that we decrease the number of slices as we
increase the ETL.
↑ETL→↑speed ↔↓coverage(↓number of slices)
One way to get around this is to increase TR. Although increasing TR
increases scan time, we save so much time by using a long ETL
compared with CSE that we can afford a longer TR. We don't need to be
limited any more to a TR of 3000 msec. We can go up to 4000 to 6000
msec, get more coverage, and still save time.
Let's say we need a coverage of fifteen 5-mm slices with a 2-mm gap to
cover an area of interest (like the brain).
Example 2
Let's choose TR = 3000 msec, ETL = 8, NEX = 1, Ny = 256.
The number of slices we can do in any TR depends on the length of the
longest echo (we'll leave out sampling time for now):
Number of slices ≤ TR/TE
In the case of an ETL of 8, the longest echo is 136 msec (17 × 8 = 136).
So,
Number of slices ≤ TR/TE = 3000 msec/136 msec [all equal to] 22
This formula is regardless of TEeff we pick. The scan time for this
example would be
= (3000 msec)(256)(1)/8
= 1.6 min
Therefore, we can do 22 slices in 1.6 min with an ETL of 8.
Example 3
Let's now choose a different ETL:
ETL = 16, NEX =1, Ny=256, TR = 3000 msec
With an ETL of 16 (since echoes are multiples of 17 msec), the longest
echo would be 272 msec (16 × 17 = 272). So,
Number of slices = TR/TE
= 3000 msec/272 msec [all equal to] 11
With an ETL of 16, we can do only 11 slices, but in 0.8 min. The scan is
faster, but we have limited coverage, and we have not accomplished
the minimum of 15 slices we needed for the exam.
Example 4
To increase the coverage with an ETL of 16, let's now increase the TR:
TR = 4500 msec, ETL = 16, Ny=256, NEX=1 Now,
Number of slices ≤ TR/TE
= 4500/272 msec [all equal to] 16
Multi-echo FSE
Consider the case of eight echoes again. With CSE, we would have eight
k-spaces, and the first seven echoes would come “free.” That's not true
for FSE. Every single one of the echoes is used to fill in a line in a k-
space. If we want to do double-echo imaging, we have twice as many
lines in k-space to fill (in two k-spaces). Thus, we have to either (i) give
up half of the echoes per ETL, in which case the scan time is going to
be doubled or (ii) repeat the scan twice, and again the scan time is
increased. Therefore, regardless of the way we do it, it's going to cost
us time.
In FSE, the first echo is no longer “free!”
There are three ways to get a double-echo image with FSE:
1. Full echo train
2. Split echo train
3. Shared echo
Full Echo Train. In a full echo train, all echoes in the train contribute
to the image. Thus, the full ETL is completed for effective TE1 (TE1eff)
before TE2eff is performed. In other words, two separate concatenated
sequences are acquired. For example, for an
ETL of 8 and a 256 × 256 matrix (Fig. 19-9), 32 echo trains would be
required to fill k-space (8 × 32 = 256).
Figure 19-9. In a full echo train, the entire echo train is completed
for TE1eff and TE2eff.
Split Echo Train. In a split echo train, the first half of the echo train
contributes to the image with TE1eff and the second half to TE2eff
(thus, two k-spaces are created). For example, for an ETL of 8 (Fig. 19-
10), only four echoes would be applied to each effective TE. Therefore,
64 trains would be required to fill k-space (4 × 64 = 256).
Shared Echo. In a shared echo approach, the first and last echoes in
the train are emphasized for TE1 and TE2, respectively, and the echoes
in between are shared for both images. This approach has the
advantage of shorter ETL compared with a full or split echo train
approach, allowing more slices to be acquired for a given TR (since the
number of slices is roughly determined by TR divided by the product of
ESP and ETL, i.e., number of slices [all equal to] TR/[ETL × ESP]).
Figure 19-10. In a split echo train, the first half of the echoes is
used for TE1eff and the second half for TE2eff.
In Figure 19-11, an ETL of 5 is used and four lines of k-space are filled
per TR, three of which are the same for the first and second echo
images. (Thus, there is some overlap of the “information” in the two
images, compared with the four split echo approach.) Filling four
echoes per pass provides the same efficiency as a split echo approach,
that is, 64 echo trains will
be required to fill k-space (4 × 64 = 256). However, the shorter ETL
allows 60% more slices to be acquired in the same time. Let's prove this
mathematically:
Figure 19-11. In a shared echo approach, the first and last echoes
are emphasized for TE1 and TE2, respectively, and the rest are
shared between the two echoes. In this example, the ETL is 5 and
the central three echoes are shared.
Advantages of FSE
1. The scan time is decreased (which allows faster scanning).
2. The signal-to-noise ratio (SNR) is maintained because we still have
256 phaseencoding steps.
3. The increased speed allows for highresolution imaging in a reasonable
amount of time. An example is 512 × 512 imaging through the
internal auditory canals with very long TR.
4. Motion artifacts will be less severe. Because the 180° pulses are
evenly spaced, there is a natural even-echo rephasing effect. For
instance, cerebrospinal fluid (CSF) motion artifacts are much less
severe on FSE than on CSE images.
5. The rephasing from the multiple 180° pulses leads to less distortion
from metallic objects on FSE images (see Chapter 18; also see the
discussion below on magnetic susceptibility).
Figure 19-12. Sagittal proton density (A) and T2 (B), and axial
proton density (C) and T2 (D) images of the lumbosacral spine
were acquired using a shared echo approach. Note that the CSF is
brighter than would be expected in a CSE proton density
acquisition. The fat is bright on the T2 images, which is also
typical of the FSE technique. This patient has minimal L5/S1
anterolisthesis that is secondary to bilateral dysplastic facets
(arrows in C and D).
Disadvantages of FSE
1. Reduced coverage, that is, decreased number of slices.
2. Contrast averaging (k-space averaging) so that
a. CSF is brighter on proton density-weighted FSE images. This is
caused by the effect of averaging all the echoes into a single k-
space. We are still contributing data from very long TEs into
the proton density image, even though the weighting is toward the
lower TE. Therefore, we will have some T2 effect (i.e., bright CSF)
on a short TEeff. To alleviate this problem, either use a shorter ETL
(to exclude longer TEs) or a higher BW (to decrease ESP and the
minimum TEeff).
b. Pathology: MS plaques and other lesions at the brain/CSF
interface may be missed on FSE. CSF appears brighter on FSE
proton density images (as discussed above) and, therefore, the
distinction between CSF and periventricular high-intensity plaques
is more difficult. As in (a), to alleviate this problem, use a shorter
ETL, which helps exclude longer TEs in the echo (fluid-attenuated
inversion recovery [FLAIR] has generally solved this problem).
3. Magnetization transfer (MT or MTC) effect in FSE. MTC is
inadvertently present in FSE. This is caused by the presence of
multiple, rapid 180° pulses containing off-resonant frequencies.
When MT (discussed in more detail in Chapter 25) is intentionally
produced, an RF pulse of 500 to 3000 Hz off the bulk water resonance
saturates protein-bound water in the broad peaks on either side of
the bulk phase water peak. Because the 180° pulses are rapid (in the
time domain), they have a broad BW (in the frequency domain), thus
containing frequencies off the bulk water resonance frequency. These
frequencies suppress proteinbound water like a fat saturation pulse
suppresses fat.
4. Normal intervertebral discs are not as bright on T2-weighted FSE
images compared with CSE. This is caused by the MT effects in FSE
previously discussed. They diminish the contrast between desiccated
(usually dark) and normal (usually bright) discs.
5. Magnetic susceptibility effects will be less than with CSE. This is
caused by decreased dephasing from closely spaced (refocusing) 180°
pulses that leave little time for spins to dephase as they diffuse
through regions of magnetic nonuniformity. In FSE, the signal loss is
minimized because of the rephasing effects of multiple 180° pulses.
Therefore, T2-weighted FSE images are less sensitive to magnetic
susceptibility effects such as metal or hemorrhage (e.g.,
deoxyhemoglobin and hemosiderin) than are T2-weighted CSE images
(Fig. 18-47).
6. Fat is bright on T2-weighted FSE images. This is due to suppression
of diffusionmediated susceptibility dephasing caused by the closely
spaced 180° pulses2 (Fig. 19-13). You could do a fat-saturated FSE to
decrease the intensity of fat.
Example 5
Performing a fat-saturated T2-weighted FSE sequence of the knee to
suppress fat in the marrow will allow bone marrow edema to stand out
against a dark marrow background. This technique increases the
sensitivity of detecting bone bruises (contusions) by about 30%3 (Fig.
19-14).
3D FSE
With the advent of high performance gradients (see Chapter 30), 3D
FSE imaging in reasonable scan times is available. This imaging is
particularly useful for the brain, cervical spine, and lumbar spine where
bright CSF (T2- weighted) images are required in more than one plane.
The basic idea in 3D imaging (3D techniques in connection with
gradient-echo imaging is discussed in Chapter 20) is to have a
phaseencoding gradient not only in the y direction but also in the z
direction (Fig. 19-15). Therefore, the multiple slices in 2D FSE are
replaced by multiple slabs. Crusher gradients (see Chapter 14) are
applied before and after every 180° pulse (which is slice selective).
Each echo is first phase encoded, then sampled, and finally phase
unwound (i.e., a rewinder gradient is applied—see Chapter 21).
Consequently, the scan time will be T(3D FSE) = (TR × NEX × Ny ×
Nz)/ETL where Ny and Nz are the number of phaseencoding steps along
the y and z axes.
Example 6
What is the scan time for a T2-weighted 3D FSE technique with TR =
3000 msec, ETL = 64, NEX = 1, Ny = 256, and Nz = 32?
Answer:
T = 3 sec × 1 × 256 × 32/64 = 384 sec = 6 min, 24 sec
which is very reasonable.
Advantages of 3D FSE
1. Higher SNR (compared with 2D FSE)
2. High (1-mm) isotropic resolution (Fig. 19-16)
Fast IR
By adding a 180° inversion pulse prior to FSE, fast IR can be achieved.
For more details, refer to the discussion on fast FLAIR in Chapter 25.
Fast STIR is similar to fast FLAIR except that the TI is chosen to null fat
instead of fluid.
Key Points
FSE imaging provides almost all the advantages of
conventional spin-echo imaging at a faster speed.
The basic idea behind FSE is utilization of multiple
180° refocusing pulses, which allows filling multiple
lines in k-space for a single TR. The ETL is defined as
the number of 180° pulses and echoes in the train.
As a result, compared with CSE, the acquisition time
T in FSE is reduced by a factor of ETL:
T(FSE) = T(CSE)/ETL
The increased speed also allows for other features
not achievable by CSE techniques, such as high-
resolution imaging of, for example, the internal
auditory canals with very long TRs and, more
recently, 3D imaging using FSE.
Questions
19-1 Advantages of FSE include all of the following except
(a) increased speed
(b) decreased ferromagnetic susceptibility artifacts
(c) decreased motion artifacts
(d) increased number of possible slices
19-7 Calculate the scan time for a 3D FSE technique with TR = 4000
msec, TE = 100 msec, Nx = 128, Ny = 128, Nz = 32, NEX = 1, ETL =
64.
Why fat is bright in RARE and fast spinecho imaging. J Magn Reson
Imaging. 1992;2:533-540.
3For more details, see Kapelov SR, Teresi LM, Bradley WG, et al. Bone
Introduction
In this chapter, we introduce the gradient-echo (GRE) pulse sequence.
It is also called gradientrecalled echo (GRE), the reason for which will
become clear later in the chapter. The major purpose behind the GRE
technique is a significant reduction in the scan time. Toward this end,
small flip angles are employed, which, in turn, allow very short
repetition time (TR) values, thus decreasing the scan time.
Consequently, such techniques are also referred to as partial flip angle
techniques. One of the most important applications of GRE is the ability
to employ three-dimensional (3D) imaging, thanks to the higher speed
of GRE due to very short TRs. The major differences between GRE and
spin-echo (SE) sequences are explained in this chapter.
Gradient-Recalled Echo
As mentioned in the introduction, the purpose of the GRE technique is
to increase the speed of the scan. Recall that the scan time for
“conventional” techniques is given by
Question: In the absence of a 180° pulse, how does one form an echo?
Answer: One way is to measure the free induction decay (FID) instead.
However, it's impractical to do so because the FID comes on too early
and decays more rapidly than we can spatially encode the signal. We
need time to apply a phase-encoding gradient and prepare the signal
for frequency encoding. We also need time to let the RF pulses and
gradients die down before doing anything else. To this end, we will
intentionally dephase the FID and rephase (or recall) it at a more
convenient time, namely, at TE. This is accomplished via a refocusing
gradient in the x direction and is illustrated in Figure 20-5. This
gradient has an initial negative lobe that intentionally dephases the
spins in the transverse plane and thus eliminates the FID. It is then
followed by a positive lobe that rephases the spins, thus restoring the
FID in the form of a readable echo. The area under the negative lobe is
equal
to half the area under the positive lobe. The refocusing occurs at the
midpoint of the positive lobe. Figure 20-6 illustrates how the FID is
refocused at TE. In other words, the FID is first eliminated and then
recalled at time TE; hence the name gradient-recalled echo.
Figure 20-6. The bilobed gradient causes dephasing of the FID and
its recalling at time TE (at the center of the positive lobe).
Because there is no 180° rephasing pulse, the rate of decay is given
by T2* (instead of T2).
Slice Excitation
The other difference between GRE and SE is that in GRE imaging, TR
may be too short to allow for processing of other slices. For this reason,
short TR GRE techniques may acquire only one slice at a time. This is
referred to as sequential scanning. (In the next chapter, we will
introduce variations of GRE that allow multiplanar imaging by
lengthening the TR.) In other words, in a sequential mode (single-slice)
GRE, the total scan time is given by
Magnetic Susceptibility
The lack of a 180° refocusing pulse results in greater dephasing of spins
compared with conventional SE. This in turn results in greater
sensitivity to magnetic susceptibility effects. This increased sensitivity
can be both problematic (e.g., increased artifact at the air/tissue
interfaces) and advantageous (e.g., when searching for subtle
hemorrhage), depending on the application (Figs. 20-7 and 20-8).
Figure 20-7. Axial T1 gradient echo out of phase (TR 110/TE 2)
(A), and T1 gradient echo in phase (TR 110/TE 4) (B) images show
marked “blooming” artifact (arrows) from the patient's known
spinal fixation hardware. The out-of-phase image (A) has less
artifact than the in phase image (B) due to the shorter TE. Axial T2
FSE (C) and ½ NEX SSFSE T2
Figure 20-7. (continued) (D) images have less artifact due to
multiple 180° refocusing RF pulses with the SSFSE having the least
due to the longest echo train length. Finally, an axial T2 FSE image
with fat saturation (E) shows ferromagnetic artifact (arrows) and
magnetic field changes resulting in paradoxical water saturation
(arrowheads) and very poor fat saturation.
Tissue Contrast
Figure 20-10 depicts a generic GRE pulse sequence diagram (PSD).
There are three operatorcontrolled parameters that affect the tissue
contrast: α, TR, and TE. Let's discuss how each of these plays a role in
this matter.
First, consider a small flip angle α (e.g., 5° to 30°). As can be seen in
Figure 20-11, a small flip angle results in a large amount of (persistent)
longitudinal magnetization after application of the RF pulse. This result
implies that complete recovery of the longitudinal magnetization to its
initial value will take much less time than following the 90° RF pulse of
an SE sequence. Consequently, given two tissues with different
Finally, let's discuss the TE factor. This parameter's role in GRE is similar
to the SE technique. That is, a short TE reduces T2* weighting and
enhances T1 or proton density weighting.
Now, with the addition of the flip angle α, this equation is modified as
follows:
Signal-to-Noise Ratio
The SNR in a GRE technique is decreased per echo (compared with SE)
due to shorter TRs in GRE; however, more echoes are obtained in GRE
per unit time, somewhat compensating for the former effect.
Thus, W = (Iip + Iop)/2 F = (Iip − Iop)/2 allowing only water (W) or fat
(F) imaging. This is the method of Dixon, which is particularly useful for
spectroscopic fat saturation at low field strengths.
3D GRE Volume Imaging
3D imaging with contiguous thin slices is feasible by using GRE
techniques. This type of imaging is accomplished by an addition of a
phase-encoding step (Nz) in the slice-select direction (z-axis).
Consequently, the total scan time is now
Example
Calculate the scan time for a 3D GRE technique in the C-spine with
the following parameters: TR 30/TE 13/α5°/NEX 2/256 × 192/Nz 64
Scan time = (TR)(NEX)(Ny)(Nz) = (30)(2)(192)(64) = 737280 msec =
737.28 sec = 12 min, 17 sec A typical PSD is shown in Figure 20-18. The
major difference here is the addition of an RF pulse for selection of a
slab (i.e., a slab-select gradient), as well as application of phase-
encoding steps in the z direction (i.e., slice encoding). This is depicted
in Figure 20-19.
Volume imaging can be performed using isotropic (cubic, with δx = δy =
δz) voxels or anisotropic (noncubic) voxels. The advantage of the
former is the ability to perform high-quality reformation in any plane of
choice. 3D imaging is now also possible using newer FSE techniques that
employ high performance gradients. More on this is in a later chapter.
Advantages of 3D GRE
1. Rapid volume imaging of thin contiguous slices without cross-talk
(Figs. 20-20 and 20-21)
2. Reformation capabilities (especially if isotropic)
Figure 20-19. The three gradients in 3D GRE are illustrated here.
Advantages of GRE
1. Increased speed
2. Increased sensitivity to magnetic susceptibility effects of hemorrhage
(allowing better detection compared with SE)
Disadvantages of GRE
1. Decreased SNR caused by (a) small α, reducing the transverse
magnetization, and (b) very short TR, not allowing sufficient recovery
of the longitudinal magnetization.
2. Increased magnetic susceptibility artifacts (caused by lack of a 180°
refocusing pulse), most noticeable at air-tissue interfaces such as in
the region of the paranasal sinuses or the abdomen (Fig. 18-43).
3. T2* decay because there are no 180° rephasing pulses. This results in
increased sensitivity to magnetic field inhomogeneities, intravoxel
dephasing, and magnetic susceptibility artifacts (Fig. 18-41).
4. Compared with FID imaging, the GRE technique (which is really an
FID-recalled technique) uses a longer TE, thus reducing the SNR
caused by increased T2* decay.
5. Introduction of chemical shift effects of the second kind, resulting in
a dark band around organs with water-fat interfaces, such as the
kidneys, liver, spleen, etc.
Key Points
1 The objective of GRE techniques is to reduce the
scan time.
2 Scan time is proportional to TR; a very small TR is
possible in GRE imaging.
3 Because a small TR does not allow reasonable
recovery of the longitudinal magnetization (thus
significantly diminishing SNR), a partial flip angle α
(<90°) should be used.
4 Because TR may be too short to allow complete
dephasing of the spins in the transverse plane, a
residual transverse magnetization may remain before
the next cycle.
5 A refocusing gradient (readout direction) is
employed that eliminates the original FID and later
recalls it at the echo time TE (hence the name
gradient-recalled echo, or GRE).
6 Because TR may be too short to acquire other
slices within one TR period, GRE techniques often
perform one slice at a time. (In the next chapter, we
will discuss multiplanar GRE techniques.)
7 As a result, the scan time is also proportional to
the number of slices obtained, that is, scan time =
TR × Ny × NEX × (number of slices)
8 The tissue contrast is a function of the flip angle α,
the repetition time TR, and the echo time TE. A
simplified way of presenting the results is
summarized in Table 20-1.
Table 20-1
Small Large
α ↑ PDW ↑ TI W
TR ↑ T2 *W ↑ T1W
TE ↑ PDW ↑ T2 *W
Questions
20-1 T/F Regarding chemical shift of the second kind, at 1.5 T, fat
and water protons get out of phase at TE = 2.2, 6.7 msec, etc.
20-6 T/F The longer the TR, the more T2* weighting will be
achieved.
20-7 Calculate the scan time for a GRE with TR = 30 msec, NEX = 2,
Ny = 256, for (a) a single slice and (b) 15 slices.
20-8 T/F Magnetic susceptibility is less with GRE than with CSE.
20-9 T/F The reason a 180° pulse is not used in GRE is to reduce the
scan time.
20-10 The SNR in 3D GRE is equal to the SNR in 2D GRE times the
factor:
(a) Nz
(b)
(c) Ny
(d)
20-11 T/F GRE sequences use a partial flip angle because the TR is
too small to allow adequate recovery of the longitudinal
magnetization to allow sufficient SNR.
1984;153:189-194.
21
Gradient Echo: Part II (Fast Scanning
Techniques)
Introduction
In the last chapter, the technique of gradientecho imaging was
introduced. In this chapter, several gradient-echo techniques will be
discussed, including GRASS (gradient-recalled acquisition in the steady
state)/FISP (fast imaging with steady-state precession), SPGR (spoiled
GRASS)/FLASH (fast low-angle shot), and SSFP (steady-state free
precession)/PSIF (opposite of FISP). Although every manufacturer uses
a different acronym, the underlying concepts are the same. We will
also discuss a multiplanar (MP) variant of these GRE techniques (e.g.,
MPGR-MP FISP, MPSPGR/MP FLASH). Finally, faster versions of these
techniques are introduced (e.g., Fast GRASS [FGR]/Turbo FISP, Fast
SPGR [FSPGR]/Turbo FLASH) as well as their MP versions (e.g.,
FMPGR/Fast MP FISP and FMPSPGR/Turbo MP FLASH).
Nomenclature
Table 21-1 contains a summary of the important acronyms used by three
major manufacturers: General Electric (GE), Siemens, and Philips. Also
refer to the list of abbreviations in the Glossary. As an example, GE
uses the prefix “Fast” and Siemens uses “Turbo” to denote similar fast
scanning techniques, be it gradient-recalled echo or spin echo. (The
acronyms are spelled out in the Glossary.)
GRASS/FISP
It was mentioned in the last chapter that, in contrast to SE, in GRE
there may be residual transverse magnetization at the end of each
cycle remaining for the next cycle. This residual magnetization reaches
a steady-state value after a few cycles and is denoted Mss.
This residual, steady-state magnetization is added to the transverse
magnetization created by the next α radio frequency (RF) pulse and
thus increases the length of the vector in the x-y plane (Fig. 21-1). This
then yields more T2* weighting. In other words, tissues with a longer T2
have a longer Mss than do tissues with shorter T2.
Actually, to preserve this steady-state component, an additional step
needs to be taken in the pulse sequence. A so-called rewinder gradient
is applied in the phase-encoding direction at the end of the cycle to
reverse the effects of the phase-encoding gradient applied at the
beginning of the cycle (i.e., it “unwinds” the former effect). In other
words, the rewinder gradient is nothing but the opposite of the phase-
encoding gradient (Fig. 21-2). For instance, if gradient +3 is applied for
phase encoding, the rewinding gradient would be −3.
SPGR/FLASH
The word “spoiling” refers to the elimination or “spoiling” of the
steady-state transverse
Table 21-1
GE Siemens Philips
1. By applying RF spoiling
2. By applying variable gradient spoilers
3. By lengthening TR
RF Spoiling. RF spoiling is the method used in SPGR and is illustrated in
Figure 21-3. In this scheme, a phase offset is added to each successive
RF pulse. This causes a corresponding phase shift in successive Mss
vectors. By maintaining a constant phase relationship between the
transmitter and the receiver (achieved via a phase-locked circuit),
successive Mss vectors cancel each other out. A pulse sequence diagram
(PSD) for SPGR is shown in Figure 21-4. In this scheme, rewinder
gradients are naturally not used because their purpose is to preserve
the steady-state magnetization, thus defeating the purpose of spoiling.
An example is seen in Figure 21-5.
Variable Gradient Spoilers. Spoiling can also be achieved by using
gradient spoilers. This is accomplished by introducing an additional
gradient with variable strengths from cycle to cycle (Fig. 21-6).
SSFP/PSIF
This technique is harder to comprehend. It yields heavily T2 (not T2*)-
weighted images. The PSD is shown in Figure 21-7. The idea here is that
each α RF pulse contains some 90° and some 180° pulses embedded in
it. Therefore, in Figure 21-7, α1 acts like a 90° excitation pulse and α2
like a 180° rephasing pulse. This yields a pulse sequence similar to spin
echo (SE) whereby an echo is formed at α3. Because it is difficult to
read a signal and transmit α3 at the same time, the echo is actually
recalled 9 msec prior to α3 by using an appropriate gradient. Note that
the echo corresponding to α1 is formed between α2 and α3.
Interestingly, in this scheme, TE is larger than TR (and smaller than 2TR
by 9 msec), which is somewhat counterintuitive. The rewinder
gradients are also shown in the diagram. The rewinder gradient is one
cycle ahead of the phase-encoding gradient due to the mechanism
described in the foregoing discussion. (In this technique, any two
successive RF pulses can be employed to create an SE.)
Figure 21-7. A PSD for SSFP/PSIF. Each α pulse contains some 180°
pulse embedded in it that acts like a refocusing pulse. This, in
turn, will result in a spin echo (SE) at the time of the next α pulse.
Hence, contrast is determined by T2 (not T2*).
SSFP/PSIF Tissue Contrast. The SSFP sequence provides heavily T2 (not
T2*)-weighted images with increased scan speed without the use of
dedicated excitation and rephasing pulses.
Advantages of SSFP/PSIF
1. Decreased dephasing due to inhomogeneities in B0 compared with
GRASS and SPGR
2. Decreased magnetic susceptibility artifacts compared with GRASS and
SPGR
3. Decreased chemical shift artifacts (dark bands) compared with GRASS
and SPGR
Disadvantages of SSFP
1. Decreased signal-to-noise ratio (SNR) secondary to the use of longer
TEs (TE > TR)
2. Increased sensitivity to nonstationary tissue
Multiplanar Techniques
The GRASS and SPGR sequences can be performed using an MP
technique by selecting a long TR (several hundred milliseconds). These
are termed MPGR (multiplanar GRASS or multiplanar gradient
recalled)/MP FISP and MPSPGR (multiplanar SPGR)/MP FLASH. As
mentioned previously, this long TR causes spoiling of the steady-state
component in the transverse plane, thus making GRASS and SPGR
possess similar features.
We can still achieve both T1 and PD/T2* weighting depending on the
flip angle α, as discussed previously. To reiterate, a small α yields PD
weighting, whereas a large α yields T1 weighting. More specifically, at
small flip angles, MPGR and GRASS behave fairly similarly, whereas at
larger angles, MPGR tends to be more T1 weighted than GRASS because
MPGR uses a long TR.
Advantages of Long TR
1. SNR is increased, because the longitudinal magnetization has more
time to recover completely.
2. MP scanning is feasible because a long TR allows acquisition of other
slices during the dead time within one TR period (similar to SE).
3. It is also possible to perform multiecho imaging (e.g., a short TE and
a long TE) similar to the multiecho, MP technique in SE. However, the
second echo tends to get degraded in GRE secondary to rapid T2*
decay.
4. A long TR, as discussed in Chapter 27, reduces saturation effects
(seen with very short TRs caused by incomplete recovery of the
longitudinal magnetization). Consequently, larger flip angles can be
used. Although larger flip angles cause more saturation effects, the
presence of a longer TR counterbalances it. Larger flip angles
obviously increase the length of the transverse magnetization and
thus the SNR.
Example
1. Determine the scan time for obtaining 15 slices using Fast
SPGR/Turbo FLASH (one slice at a time) when TR = 10 msec, TE =
min, Ny = 256, NEX = 1: Time = (10)(256)(1)(15) = 38,400 msec = 38.4
sec
2. Determine the scan time for obtaining 15 slices using MP Fast
SPGR/Turbo FLASH (multislice) when TR = 100 msec, TE = min, Ny =
256, NEX = 1: Time = (100)(256)(1) = 25,600 msec = 25.6 sec
Disadvantages
1. Decreased SNR and CNR caused by ultrashort TRs (less so in degree
with MP techniques)
2. Increased chemical shift artifacts of the second kind at very low TEs
(namely, at TE = 2.2 msec, 6.6 msec, etc.)
Table 21-2
Spleen 400-500
CSF 700-800
Flow Imaging
GRE scanning is generally performed one slice at a time (except in the
MP situation); therefore, each slice is an entry slice. Consequently,
flowrelated enhancement (FRE) applies to every single slice, and
vessels appear bright on GRE images.
The basic concept is that no saturated flowing protons enter the slice
so that flipping these protons yields maximum signal. This is the basic
concept behind 2D or 3D time of flight (TOF) MR angiography. This
topic is discussed at length in Chapters 26 and 27.
Key Points
1 Several GRE techniques are available:
GRASS/FISP, SPGR/FLASH, and SSFP/PSIF (Table 21-
1).
2 In GRASS, the residual transverse magnetization is
preserved via a rewinder gradient, contributing to
increased T2* weighting.
3 In SPGR/FLASH, the residual transverse
magnetization is “spoiled” by introducing phase shifts
in the successive RF pulses. This causes reduced T2*
and increased T1 weighting.
4 Spoiling can also be accomplished via gradient
spoilers or by lengthening the TR.
5 In SSFP/PSIF, heavily T2-weighted images are
obtained. GRASS/FISP and SPGR/FLASH represent
gradient-recalled FID sequences, whereas SSFP/PSIF
represents a gradient-recalled SE (spin-echo)
sequence. Interestingly, in this technique, TE is
larger than TR and usually is smaller than 2TR by 9
msec.
6 MP variants of the above are also available (e.g.,
MPGR/MP FISP and MPSPGR/MP FLASH) by
employing a longer TR (over 100 msec).
7 Fast GRE techniques (e.g., FGR/Turbo FISP and
FSPGR/Turbo FLASH) provide additional speed. This
is accomplished by using a fractional RF, fractional
echo, and fractional NEX, as well as a wider BW
(shorter sampling time Ts).
8 Combined fast and MP GRE techniques (e.g.,
FMPGR and FMPSPGR/Turbo MP FLASH) render
multiple slices with increased SNR in a fast mode.
9 The characteristics of GRASS/FISP, SPGR/FLASH,
and SSFP/PSIF are summarized in Table 21-3.
Table 21-3
GRE
Technique SNR CNR Comments
Best
GRASS possible Preserves steady-
/FISP Highest T2* state component
Best
SPGR possible Spoils steady-
/FLASH Intermediate TI W state component
Gradient-
Provides recalled SE ; TR
SSPF/PSIF Lower T2W < TE < 2TR
Questions
21-1 Spoiling of the residual transverse magnetization can be
accomplished by
(a) gradient spoilers
(b) RF spoilers
(c) long TR
(d) all of the above
(e) only (a) and (b)
Introduction
In the last three chapters, we discussed some of the fast scanning
techniques, namely, the fast spin echo (FSE) and gradient-recalled echo
(GRE) techniques along with their “fast” variants. In this chapter, we
will discuss the echo planar imaging (EPI) technique, which is the
fastest MRI technique available.
Types of EPI
Two main types exist: single-shot EPI and multishot EPI. Earlier single-
shot EPI techniques used a constant phase-encode gradient. Newer
techniques use a “blipped” phase-encode gradient, referred to as
“blipped EPI.”1
Single-Shot EPI. In single-shot EPI, all the lines in k-space are filled by
multiple gradient reversals, producing multiple gradient echoes in a
single acquisition after a single radio frequency (RF) pulse, that is, in a
single measurement or “shot.”
To achieve this, the readout gradient must be reversed rapidly from
maximum positive to maximum negative Ny/2 times (e.g., 256/2 = 128
times) during a single T2* decay (e.g., 100 msec). Each lobe of the
readout gradient above or below the baseline corresponds to a separate
ky line in k-space. Therefore, the number of phase-encode steps Ny is
equal to the sum of the positive and negative lobes of the readout
gradient. The area under the Gx lobe determines the field of view (FOV)
—the larger the area, the smaller the FOV. Apparently, single-shot EPI
places tremendous demands on the gradients with respect to maximum
strength Gmax and minimum rise time tRmin (i.e., the maximum slew
rate Gmax/tR) as well as on the analogto- digital converter (ADC). In
general, ADCs with maximum bandwidths (BWs) on the order of MHz are
required rather than the kHz maximum BWs used for conventional spin
echo (CSE).
In earlier EPI methods, the phase-encode gradient was kept on
continuously (Fig. 22-1)
Figure 22-3 provides a PSD for blipped EPI in which the phase-encode
gradient is applied briefly Ny number of times when the readout
gradient is zero.
Contrast in EPI
Contrast in EPI depends on the “root” pulsing sequence (which is
similar to preparation prepulses in GRE techniques). For instance, to
achieve SE contrast, a 90°-180° SE root sequence is applied before the
EPI module. Similarly, a partial flip RF pulse (<90°) before the EPI
module provides gradient-echo contrast. A 180°-90°-180° IR root prior
to EPI provides inversion recovery contrast. In addition, diffusion
gradients can be added for EPI-diffusion imaging.
Figure 22-7. An SE-EPI PSD for the original EPI. A 90°-180° root
pulsating sequence is applied before the EPI module.
Artifacts in EPI
N/2 Ghost Artifacts. Even with blipped EPI, phase errors may result
from the multiple positive and negative passes through k-space (i.e.,
alternating polarity of the readout gradient). “Ghost” artifacts of the
main image may appear along the phase axis, not caused by motion as
in CSE, but by eddy currents, imperfect gradients, field
nonuniformities, or a mismatch between the timing of the odd and even
echoes. Since the ghosts are derived from half the data (even or odd),
they are called N/2 ghosts. See Figure 22-8 for examples.
Remedy. Minimize eddy currents; proper tuning of the gradients.
Susceptibility Artifacts in EPI. Magnetic susceptibility effects in EPI
may result in variations in frequencies and phase errors. This effect is
reduced for multishot EPI because phase
errors have less time to build up. Advantages of multishot EPI over FSE
is that it has contrast much closer to CSE and it has greater sensitivity
to magnetic susceptibility effects such as hemorrhage compared with
FSE.
Figure 22-8. Both (A) and (B) are b0 images from an EPI sequence
of the brain in different patients. (A) has a marked amount of “N
over 2” artifact, whereas (B) has more subtle findings (arrows).
Advantages of EPI
1. Scan time is approximately 100 msec or less.
2. Cardiac and respiratory motion won't pose problems any longer.
Figure 22-11. Diffusion tensor showing reorientation of
coordinate system so new z-axis (z) is parallel to the main white
matter tract in the voxel.
Disadvantages of EPI
1. Because fat-water chemical shift artifacts (of the second kind [Dixon
effect]) can be problematic with such short TEs, fat suppression with
presaturation techniques is always required for EPI.
2. Because of rapid on-and-off switching of the gradients, there is a
potential for generating an electric current or voltage in the patient,
thus causing peripheral nerve stimulation where the patient may
experience a crawling sensation in the skin also known as
“formication.” This shock is caused by the well-known fact in
electromagnetic theory that rapid changes in a magnetic field (i.e.,
dB/dt for the mathematically oriented reader) induce an electric
current (E) in a conductor (in this case the patient).
Figure 22-17. Axial EPI DWI image of the liver with a low b value
(b = 50 sec/mm2) demonstrates a bright T2 hemangioma (arrow).
Also note that the low b value provides excellent suppression of
signal in the liver vessels. This is a classic example of “T2 shine
through.”
Figure 22-18. Short-axis segmented GRE-EPI dynamic
postgadolinium adenosine stress study image shows decreased
signal in the inferolateral wall consistent with ischemia (normal
signal was seen on the resting study).
Key Points
Echo planar imaging (EPI) is the fastest MR
technique that is widely available. It has become
particularly important in the imaging of acute CVAs.
It employs a train of oscillating frequency-encoding
gradients, thus rendering a sinusoidal k-space
traversal after a single RF pulse. Consequently, each
slice can be imaged in a matter of milliseconds (free
of any motion artifact), and the entire study can be
accomplished in a matter of seconds.
Questions
22-1 T/F Contrast in EPI depends on the root pulsing sequence.
22-3 Regarding multishot echo, all the following are true except
(a) It places less stress on the gradients compared with single-
shot EPI.
(b) Phase errors have less time to build up compared with single-
shot EPI.
(c) It takes longer to perform than singleshot EPI.
(d) It is less susceptible to motion artifacts.
Radiology. 1994;192:600-612.
23
New Scanning Features
Introduction
In this chapter, we discuss some of the more recent techniques used by
newer MR scanners with more advanced software. The following is a
summary of these features and their functions:
1. To increase speed
a. Fractional number of excitations (NEX)
b. Fast spin echo (FSE)
c. Fast gradient-echo techniques
d. Parallel imaging (discussed in the next chapter)
2. To reduce TE
a. Fractional echo
b. Fractional radio frequency (RF)
3. To increase resolution (without time penalty)
a. Asymmetric field of view (FOV)
4. To reduce aliasing
a. No phase wrap
b. No frequency wrap
5. To increase coverage
a. Phase-offset RF pulses
6. To achieve contiguous slices
a. Contiguous slices
b. 3D acquisition
7. To achieve saturation
a. Spatial saturation
b. Spectral (chemical) saturation
8. To increase signal-to-noise ratio (SNR)
a. Low bandwidth
9. Reduce motion
a. Periodically rotated overlapping parallel lines with enhanced
reconstruction (PROPELLER)
Remember that FSE and fast gradient-echo techniques are separate
pulse sequences, whereas the other features can be added to any pulse
sequence.
Fractional NEX
See Figure 23-1, and also refer to Chapter 13.
Mechanism
1. Only a portion of k-space is used (e.g., ½ NEX, ¾ NEX [actually, the
number of phase encodes Ny is reduced, not NEX]). Reconstruction is
based on inherent symmetry of k-space along the phase axis.
2. Slightly more than half of k-space is used (called overscan) for phase
correction.
3. The center of k-space is usually included because it contains the
strongest signals.
Advantages
1. Increased speed (caused by reduced Ny)
Disadvantages
1. Decreased signal-to-noise ratio (SNR)
2. May increase artifacts
Applications
1. Used for localizer (scout) images
Fractional Echo
See Figure 23-2.
Mechanism
1. Only a fraction of the received echo is sampled (feasible because of
the symmetry of the echo about TE and symmetry of k-space along
frequency axis).
Advantages
1. TE can be reduced
2. SNR is improved in early echoes (less T2 decay)
3. Improves T1 weighting (reduces T2 effect)
4. May decrease flow artifacts and susceptibility effects
Figure 23-2. Fractional echo. Note outer dashed box is the full
echo while the inner dashed box represents the fractional echo.
Applications
1. T1-weighted images
2. To reduce flow artifacts and magnetic susceptibility effects
NOTE: Using a higher bandwidth can also result in a lower minimum TE,
however, at a SNR loss.
Asymmetric FOV
See Figure 23-4.
Mechanism
1. Rectangular FOV is used (FOV is typically reduced in phase direction
because Ny is typically less than Nx).
2. May get square or rectangular pixels.
Advantages
1. With rectangular FOV, we can obtain resolution of, say, 512 × 512
matrix in the time it takes to perform a 512 × 256 acquisition.
2. Increases speed while maintaining resolution.
3. Useful when the anatomy being imaged is asymmetric (smaller) in
phase direction (e.g., the spine).
Disadvantages
1. Reduced SNR compared with full FOV
2. May cause wraparound in phase direction
Figure 23-4. Asymmetric FOV. When a rectangular FOV is used,
FOV typically is reduced in the phase direction. By acquiring more
phase-encoding steps, a higher resolution can be achieved. In this
example, you can acquire 256 phase-encoding steps and still get a
512 resolution.
Applications
1. Spine
2. Extremities
3. Abdomen
2. Various low-pass filters (LPFs) and band-pass filters (BPFs) are also
used to get rid of unwanted high frequencies in the signal.
Advantages
1. Avoids wraparound in the frequencyencoding direction
2. No increase in the scan time
Disadvantages
1. May decrease SNR because, by increasing the number of samples, the
sampling interval is reduced and thus the BW is increased (remember
BW = 1/δTs and SNR ∝ ).
2. This is done internally, that is, the operator has no control over it.
Applications
1. Almost always automatically turned on during routine scanning
Phase-Offset RF Pulses
See Figure 23-6.
Mechanism
1. Simultaneously excites two slices with two RF pulses that have a
phase offset
Advantages
1. Doubles the number of slices per TR (without increasing time)
2. Can achieve more T1 weighting by reducing TR
Disadvantages
1. May increase wraparound artifact
Figure 23-6. Phase-offset RF pulses. Two RF pulses with different
phases excite two slices simultaneously.
Applications
1. When a short TR is desired, and a large area is studied.
2. Gadolinium-enhanced studies in which shorter TR enhances
paramagnetic effects of contrast.
2. Eliminates the need for interleaving (which doubles the scan time
when correctly performed with two acquisitions)
Disadvantages
1. Increases TE (longer RF pulses)
2. Fewer slices per TR
Applications
1. When gaps are not desirable
3D Acquisition
Mechanism
1. Gradient-echo (GRE) technique over a 3D volume.
2. 3D FSE T2 is an option.
3. Requires a phase-encoding gradient along the slice-select z-axis.
4. Contiguous slices (zero gaps).
Advantages
1. Increased SNR (due to acquisition from larger volume)
2. Allows thin, high-resolution contiguous slices and/or overlapping
slices
3. Allows for high-quality reformation in any plane
Disadvantages
1. May introduce wraparound artifacts at each end of the volume (end
of the slab) caused by the presence of phase encoding in the z
direction (Figs. 18-3 and 18-4).
2. Scan time now incorporates the number of phase-encoding steps
along the z-axis (e.g., 28, 60) into the formula as well. That is to say,
Acquisition time(3D GRE) = T(3D GRE) = TR · NEX · Ny · Nz = Nz · T(2D
GRE) but because TR is very short, this is acceptable. For 3D FSE, this
formula becomes Acquisition time(3D FSE) = T(3D FSE) = TR · NEX · Ny
· Nz/ETL = Nz · T(2D FSE) where ETL can be chosen to be very large.
Applications
1. C-spine
2. MR angiography (e.g., circle of Willis)
3. Dynamic gadolinium abdomen
Advantages
1. It resolves tissues with similar T1 values (such as fat and gadolinium-
enhanced tumors).
2. This technique does not have any influence on the signal from tissues
other than the one being suppressed (in contrast, IR affects the
contrast of all tissues).
Disadvantages
1. Since this approach employs a frequencyselective technique, it
suffers from sensitivity to magnetic field inhomogeneities (see Figs.
18-42, 18-43, and 20-7).
2. Requires extra time (thus lengthening TR and increasing the scan
time).
3. Increases the number of RF pulses, causing extra RF heating.
Key Points
Many new MR techniques are reviewed in this
chapter. These, along with their major trade-offs, are
summarized below.
Advantage Feature Disadvantage
Questions
23-1 Speed can be increased by
(a) FSE or GRE techniques
(b) fractional NEX
(c) reducing BW
(d) all of the above
(e) only (a) and (b)
23-4 T/F Spatial saturation pulses are used to minimize phase ghosts
and flow artifacts.
Introduction
Parallel imaging is an approach for reducing scan time and includes
techniques such as sensitivity encoding (SENSE) and generalized
autocalibrating partially parallel acquisition (GRAPPA). Vendor
implementations of these techniques are also known as array spatial
and sensitivity encoding technique (ASSET), autocalibrating
reconstruction for Cartesian sampling (ARC), integrated parallel
acquisition technique (iPAT), and rapid acquisition through parallel
imaging design (RAPID).
Concepts
The “parallel” in parallel imaging refers to the fact that each coil in
the phased array receives data at the same time (i.e., in parallel).
Recall from Chapter 2 that a phased array is made up of many small
coils, with 32 or more available on modern systems.
Parallel imaging works by knowing the local sensitivity of each coil in
the phased array. The field of view (FOV) is made intentionally too
small in the phase-encoding direction, and the resulting aliasing is
unwrapped using this information. The basic concepts were described in
the early 1990s, and the first successful clinical implementation was
simultaneous acquisition of spatial harmonics (SMASH). SMASH has now
been superseded by more sophisticated techniques and the main
approaches in clinical use today are SENSE and GRAPPA.
A parallel imaging acquisition with an acceleration factor of 2 acquires
data only for every other line of k-space. The corresponding images
have a ½ FOV and exhibit aliasing artifact. The SENSE and GRAPPA
approaches work either by unwrapping the aliasing in the images or,
equivalently, by filling in the missing lines of k-space (Fig. 24-1).
Both approaches make use of the different spatial sensitivities of the
coils in the phased array. Due to their distinct spatial locations, each
coil has different view of the object (Fig. 24-2). Knowing their spatial
locations and sensitivity profiles allows the aliasing to be undone by
postprocessing.
Figure 24-1. Image domain (top) and the associated k-space
(bottom), which indicates how aliasing in the image domain is
equivalent to undersampling in k-space. The SENSE method
unwraps the aliasing in the image, but the same result can be
obtained using GRAPPA to fill in lines of k-space.
Figure 24-2. Aliased images obtained from an 8-coil phased array.
Note each image has the characteristic spatial sensitivity of the
coil, such that signals near the coil are strong and those far away
are weak. This basic information allows us to work out which
signals are in the right location and which signals have been aliased
from ½ FOV away.
Prescan or Autocalibration
There are two common ways to determine the spatial sensitivities of
the coils in the phased array. SENSE and GRAPPA differ in the way that
spatial sensitivity maps are determined and how the postprocessing is
done.
Prescan. A large 3D volume acquisition lasting a few seconds run
immediately after the localizers. This gives low-resolution images of
the entire region inside the scanner and provides spatial sensitivity
maps for all subsequent scans.
Autocalibration. A few extra phase-encoding lines are acquired within
each parallel imaging scan. These are used in postprocessing to train
the algorithm how to generate missing k-space lines for that specific
scan only.
Advantages/Disadvantages
1. Prescan only needs to run one time so there is little time spent on
sensitivity measurements.
2. Autocalibration calibrates on every scan so it can tolerate patient
repositioning between scans.
3. Autocalibration costs a few extra phaseencodes so acceleration
factors will be slightly lower (e.g., 1.9 rather than 2). Note the SNR
decrease is proportionate (e.g., decrease rather than ).
SENSE. Sensitivity encoding generally uses a prescan to measure
sensitivity of the coils. The individual images are divided by a body coil
image to remove “contaminating” structure, leaving just the spatial
sensitivity of the coils (or sensitivity maps). These are shown in Figure
24-3; for simplicity, only two coils in the phased array are shown.
Since we used an acceleration factor of 2, we know the aliased pixels
come from two locations at a distance of exactly ½ an FOV apart. We
also know the pixels are weighted by the sensitivity of the coil, for
example, pixels close to the coil will have a higher weighting than
pixels further away. By knowing the aliasing pattern and the spatial
sensitivities, we can write a linear equation and solve it to obtain the
values of the two pixels (Fig. 24-3). The process is repeated for all
pixels.
Because the coil sensitivity maps and aliased images are represented in
the spatial domain, SENSE is often described as an image-domain
parallel imaging technique.
GRAPPA. GRAPPA uses autocalibration to provide information about the
coil sensitivities. For an acceleration factor of 2, alternate lines of k-
space are skipped but a few additional lines are acquired at the center
for calibration. The overall acceleration factor is slightly reduced
although the calibration data contribute to the SNR so the SNR is
proportionate to the overall acceleration factor. A schematic of the k-
space data is shown in Figure 24-4.
In GRAPPA, we seek to recreate the missing k-space lines using a
combination of its neighbors, often just two neighboring lines. Note the
SMASH technique mentioned above uses just one neighboring line.
The key is to work out how to generate missing lines, and this is where
the calibration data are needed. If the calibration data consist of (say)
line 2, then we need to find what linear combination of lines 1 and 3
can give the closest approximation to line 2. We must also specify a
coil, so the full procedure is to determine what
combination of lines 1 and 3 from all coils gives the best approximation
to line 2 from coil 1. Similarly, we must determine what combination
gives the best approximation to line 2 of coil 2.
Key Points
We have looked at the concepts behind parallel
imaging, which uses phased-array coils to accelerate
scans, and taken a close look at two popular
methods, SENSE and GRAPPA. We saw that each coil
in a phased array produces an image with a
characteristic spatial sensitivity and that
postprocessing can undo the aliasing introduced by
reducing the FOV in the phase direction, for a
corresponding saving in scan time. We noted that
acceleration factor of 2 is readily obtainable on
current systems; this is likely to increase as special
purpose coils for high accelerations are developed.
Questions
24-1 What does the word “parallel” refer to in parallel imaging?
(a) The need to use fast computers for postprocessing
(b) The simultaneous data acquisition by the coils
(c) The blend of phase-encoding and phased-array coils
(d) The arrangement of the coils needed to make it work
Introduction
One of the beauties of MR, especially with some of the new features, is
the ability to image a body part while “suppressing” the signal coming
from a certain selected tissue. This suppression allows perturbation of
tissue contrast to enhance the signal coming from tissues of greater
interest (such as a pathology). Two types of tissues are commonly
suppressed in clinical practice: fat and water.
Suppression Techniques
Several suppression techniques are available, some of which are listed
below.
1. Inversion recovery techniques
2. Chemical (or spectral) saturation or frequency-selective
presaturation
3. Spatial presaturation in the field of view (FOV)
Inversion Recovery (IR) Techniques. This technique was discussed at
length in Chapter 7. The IR pulse sequence diagram (PSD) is shown in
Figure 25-1. By appropriate selection of TI (time to inversion), we can
nullify or suppress a certain tissue. In fact, as we saw in Chapter 7, if TI
= (ln 2) [T1(tissue x)] = 0.693 T1(tissue x) then tissue x is nulled. Thus,
TI can be chosen to null fat or water or any other desired tissue
depending on the application (Fig. 25-2).
STIR (Short TI Inversion Recovery). This is the IR technique that is used
to suppress fat.
Example
What is the TI used in STIR? At 1.5 T, T1 of fat is approximately 200
msec. Then TI = 0.693 × 200 ≅ 140 msec FLAIR (Fluid-Attenuated
Inversion Recovery). This is an IR technique that nulls fluid. For
example, this sequence is used in the brain to suppress cerebrospinal
fluid (CSF) to bring out the periventricular hyperintense lesions, such as
multiple sclerosis (MS) plaques (Fig. 25-3).
Example
What is the TI used in FLAIR? At 1.5 T, T1 of CSF is approximately
3600 msec. Then TI = 0.693 × 3600 ≅ 2500 msec Fast FLAIR. STIR
sequences are typically performed with a fast spin-echo (FSE)
technique; however, there are some novel changes that must be
performed to apply FSE technique to the normally slow FLAIR sequence
and achieve CSF suppression in a fast manner.1 Figure 25-4 depicts a
schematic representation for Fast FLAIR. In this scheme, the following
parameters are used:
TR = 10,000 msec
TI = 2500 msec
FSE with ETL = 8
TEeff = 112 msec
Figure 25-1. Inversion recovery (STIR shown). The recovery curves
following the 90° and 180° pulses are illustrated for fat and water.
In Figure 25-4, there are two packets of 15 slices each. During the first
5000 msec, 15 slice-selective 180° inverting pulses are followed 2500
msec (TI) later by 15 slice-selective FSE readout in 2500 msec. During
the second 5000 msec, recovery occurs in the first 15 slices (for a total
TR of 10,000 msec), and the process is repeated on the second set of 15
interleaved slices. In all, 30 slices are acquired.
A TR of 10,000 msec allows almost complete longitudinal recovery of
CSF. This relatively long time period also makes it possible to perform
multislice interleaving during inversion and readout. The inversion
“period” is the time during which 15 slice-selective 180° pulses are
applied. It is the time TI (2500 msec in this case) from the first 180°
inverting pulse to the 90° pulse at the beginning of the readout period.
FSE readout takes 136 msec (8 × 17 msec) for each slice. The readout
“period” is the time (also 2500 msec) during which 15 sliceselective
readouts are performed. The recovery period is the time from the slice-
selective 90° pulse
Disadvantages of IR
1. Tissues with similar T1 values are all suppressed and thus cannot be
differentiated (e.g., fat and blood or gadolinium-enhanced tumors
—Figs. 25-5 and 25-6)
Figure 25-6. Coronal T1 postgadolinium image with fat saturation
(A) shows a large enhancing hemangioma in the left shoulder of a
young patient with Kesselbach-Merritt syndrome. Pregadolinium
coronal FSE T2 image (B) shows areas of bright signal; however, a
postgadolinium STIR image (C) shows signal dropout from the T1
shortening effects of gadolinium that make the lesion's T1
equivalent to that of fat.
Advantages
1. Resolves tissues with similar T1 values (such as fat and gadolinium-
enhanced tumors, or fat and blood products—Fig. 25-5)
2. No influence on the signal from tissues other than the one being
suppressed (in contrast, IR affects the contrast of all tissues)
Figure 25-7. Spectral presaturation. A frequency-selective
presaturation pulse is applied before the excitation pulse to
eliminate longitudinal magnetization for a specific tissue such as
fat or water.
Disadvantages
1. Suffers from sensitivity to magnetic field inhomogeneities because it
employs a frequency-selective technique (Fig. 25-9)
2. Requires extra time (thus lengthening TR and increasing the scan
time)
3. Increases length of RF pulses, causing extra RF heating
Spatial Presaturation. Spatial presaturation is done generally to reduce
motion-and flowrelated artifacts of structures adjacent to—or actually
within the FOV next to—the region of interest. Examples include
Figure 25-8. Axial T2 FSE image with chemical (spectral) fat
saturation shows suppression of fat signal. There is an intermediate
brightness left adrenal ganglioneuroma (arrows), which extends
toward the midline.
Key Points
Tissue suppression is an important feature of MRI—it
allows improved tissue contrast and enhanced lesion
detectability. Generally, two tissues are subject to
suppression: fat and water. The two major
suppression techniques are inversion recovery (IR)
and chemical (spectral) saturation. Each technique
has its own advantages and disadvantages. The
selection of the type of technique depends on the
clinical application. Other tissues may be saturated
(e.g., protein-bound water and flowing blood).
Major fat suppression techniques include
1 STIR
2 Chemical (spectral) fat sat
Major fluid suppression techniques include
1 FLAIR
2 Chemical (spectral) water suppression
Techniques for suppression of protein-bound water
include
1 MT technique for background suppression
2 MT effect in FSE
Effects of presaturation technique in the FOV include
1 Saturation bands reduce flow artifacts (e.g., in
spine imaging)
2 Saturation pulses remove venous or arterial flow
(e.g., in MRA or MRV)
Questions
25-1 In an IR sequence, the TI to null a tissue is given by
(a) 0.693 T2
(b) 0.693 T1
(c) (1/0.693) T1
(d) (1/0.693) T2
Introduction
Unlike computerized tomography (CT) (with or without contrast), in
which the appearance of flowing blood in a vessel is predictable, the
appearance of flowing blood in MRI is much more complicated. Flowing
blood or cerebrospinal fluid (CSF) can appear dark or bright depending
on numerous factors, including, but not limited to, the following:
1. Velocity
2. Pulse sequence (e.g., spin echo [SE] vs. gradient echo [GRE])
3. Position of the slice containing the vessel relative to the rest of the
slices
4. Contrast (TR and TE)
5. Echo number (even or odd)
6. Slice thickness
7. Flip angle
8. Gradient strength and rise time
9. Use of gradient moment nulling techniques
10. Use of spatial presaturation pulse
11. Use of cardiac gating
12. Chance of cardiac gating (pseudogating)
Types of Flow
In Chapter 18, we discussed two types of motion: random and periodic.
Blood and CSF flow have a periodic-type motion. Furthermore, flow of
blood can be divided into the following:
1. Laminar flow
2. Plug flow
3. Turbulent flow
4. Flow (stream) separation/vortex flow
These types of flow are depicted in Figure 26-1. Let's discuss these
separately.
Laminar Flow. This type of flow is seen in most normal vessels and has
a parabolic profile. If the lumen radius is R, then the velocity v at
position r would be
v(r) = Vmax (1 - r2/R2)
where Vmax is the maximum velocity in the center of the lumen.
Therefore, the average velocity in the lumen is given by
Vave = Vmax/2
Plug Flow. This type of flow is usually seen only in large vessels (aorta)
with high velocities. The velocity across the lumen is constant, thus
yielding a flat velocity profile:
v(r) = Vmax = Vave = constant
Turbulence. This phenomenon is observed in abnormal vessels (e.g.,
distal to stenosis) or at bifurcations during which a random motion of
fluid elements is observed. Vortex flow and large-scale recirculation
zones (i.e., eddies) are other terms that refer to such random motions.
Flow (Stream) Separation. This phenomenon is observed near the wall
of a vessel (e.g., in the proximal internal carotid artery) in which some
of the flow is separated from the main streamline.
Question: What determines laminar versus turbulent flow?
Figure 26-1. Types of flow. A: Laminar flow. B: Plug flow. C:
Turbulent flow (distal to stenosis). D:Flow (stream) separation.
Table 26-1
Aorta 140 ± 40
or
v = d/t
The distance each proton in flowing blood has to travel within a slice is
the slice thickness δz. The time between a 90° pulse and a 180° pulse
is one-half TE.
Thus, if the velocity of flowing blood is
v = δz/(1/2 TE)
then protons flowing into the slice would be exposed only to the 90°
pulse, and not to the 180° pulse, thus forming no signal or SE. Let's give
this velocity an arbitrary name vm, that is,
vm = δz/(1/2 TE)
However, if the velocity were 0 (i.e., for stagnant blood), then an SE
would be formed. If the velocity falls in between, only a fraction of the
protons would form an SE. The fraction of protons escaping the 180°
pulse is given by
v/vm = v(1/2 TE)/δz
Thus, the fraction of protons receiving both RF pulses is
1-v/vm = 1-v(1/2 TE)/δz
The signal intensity is, therefore, proportional to
I α 1-v(1/2 TE)/δz
Figure 26-3 illustrates the linear relationship between signal intensity
(I) and velocity v. It is clear from this graph that if the velocity of
flowing blood is at least equal to vm (i.e., v ≥ vm), then a flow void is
observed.
Figure 26-3. A plot of signal intensity versus velocity. The higher
the velocity, the more the signal loss.
Figure 26-4. Axial PD image shows TOF losses in the right internal
carotid and basilar arteries (black arrows); however, no TOF loss or
“flow void” is seen in the left internal carotid artery (white
arrow). Patient had left internal carotid artery occlusion.
Example
Suppose that the slice thickness δz = 1 cm and TE = 50 msec. What is
the velocity at which flow void is observed?
Using the above formula, the velocity should be at least
vm = Δz/(1/2 TE) = 1 cm/(25 msec) = 1000 cm/25 sec = 40 cm/sec
which is seen in an artery.
Therefore, vm is larger for thicker slices and smaller TEs, and vice
versa (see Fig. 26-4 for an example).
N.B.: Remember that TOF losses only apply to SE imaging and not to
GRE imaging
Dephasing
There are many causes of dephasing. One important cause is the so-
called odd-echo dephasing. This phenomenon results in signal void on
the first and other odd echoes. The protons in a voxel do not move at
the same velocity across the lumen in laminar flow; thus, each
precesses at a different frequency and accumulates a different phase.
(Also see the discussion below on evenecho rephasing.)
Another cause of dephasing is intravoxel dephasing. Because of
laminar flow, different velocities may exist within a voxel, thus leading
to phase dispersion (incoherence) and signal loss. How to decrease
intravoxel dephasing and increase SNR:
1. Decrease the voxel size (increase spatial resolution), either by
increasing the matrix (trade-off: will reduce signal-to-noise ratio
[SNR]) or by reducing the FOV (trade-off: may cause wraparound).
2. Reduce the TE (e.g., use a fractional echo).
3. Add flow compensation (FC) techniques (see below).
Even-Echo Rephasing
This phenomenon is somewhat the opposite of odd-echo dephasing. It
only occurs in SE imaging with symmetric echoes (i.e., when the second
echo delay time is twice the first one: when TE2 = 2 TE1, e.g.,
30/60/90/120 and 40/80/120/160). The result is higher signal intensity
for even echoes compared with odd echoes. (We shall see later that in
GRE or SE, FC techniques basically yield “even-echo rephasing” on the
first echo to minimize signal losses.)
To see how this works, we first need to learn about the relationship
between phase and velocity. Recall that along the readout gradient,
ω = γBX = γGX
Now, for a constant velocity v, the position at time t is given simply by
x = vt
Thus, combining the above two, we get
ω = γGvt
Now, phase change Δφ and angular frequency ω are related by
δφ = ωδt
so that
φ = ∫ωdt = ∫(γGvt)dt = γGv∫t dt = γGv(t2/2)
From this we can see the following:
1. Phase φ and velocity v are proportional.
MATH: Let's try to prove the above fact mathematically. Assume that τ
= 1/2 TE. The phase gain at time TE/2 is kτ2, where k is a constant (k =
1/2 γGv). Right after the 180° pulse, the phase is -kτ2. Now, at time TE
(i.e., at 2τ), the phase gain will be k[(2τ)2 - (τ)2] = 3kτ2. (The
mathematically oriented reader will recognize this as the integral of ∫
2kτ from point τ to 2τ.) Thus, the net phase gain will be 3kτ2 + (-kτ2) =
2kτ2. Similarly, at 3/2 TE (or 3τ), that is, at the second 180° pulse, the
phase gain is k[(3τ)2 + (2τ)2] = 5kτ2, with a net gain of 5kτ2 + 2kτ2 =
7kτ2. Immediately after this second 180° pulse, the phase will be -7kτ2.
In a similar fashion, the phase gain at the time of the second echo (i.e.,
4τ) will be k[(4τ)2 - (3τ)2] = 7kτ2. Therefore, the final net phase gain
will be 7kτ2 + (-7kτ2) = 0. That is to say, the net phase shift of flowing
blood on the second echo is zero. This is illustrated in Figure 26-6.
Thus, protons in flowing blood lose their coherence on the first echo
(causing odd-echo dephasing) and subsequently regain coherence on
the second echo (causing even-echo rephasing). This pattern yields a
higher signal of flowing blood on the second echo. It only works for
symmetric echoes, not for asymmetric ones. (Interested readers can go
through the previous math and prove this fact for themselves.)
Diastolic Pseudogating
During a cardiac cycle, blood flow is more rapid during systole and
slower in diastole. Thus, in diastole, a higher intravascular signal is
observed. (Remember that higher flow results in more TOF losses.)
When cardiac gating is used, each slice is (theoretically) acquired at a
fixed point in the cardiac cycle (although different slices in a multislice
acquisition are obtained at different points in the cycle). Consequently,
a vessel traversing several slices will demonstrate varying signal
intensities at different slices. In a cardiac-gated sequence, TR must be
a multiple of the heart rate (HR). For example, if HR = 60 bpm (1 beat
per second = 1 Hz), then TR = 1000, 2000, 3000, etc. In general,
TR = 1/HR
with appropriate units.
protons can penetrate adjacent slices and yield FRE. Obviously, as these
inflowing protons travel through the imaging volume, they are
subjected to more and more RF pulses and become more and more
saturated. Thus, FRE is always maximum at the entry slice and becomes
gradually fainter in deeper slices (a good distinguishing point from an
intraluminal thrombus). Now, how far FRE can penetrate the imaging
volume has to do with the direction of flow and the direction of slice
excitation.
Figure 26-7. Effect of flow-related enhancement (FRE) for slow
flow. Unsaturated protons produce more signals. As the velocity
increases, a larger portion of unsaturated inflowing protons will
replace the previously partially saturated protons, thus increasing
signal intensity.
GE FC Flow compensation
FC can be applied along each and all the three coordinates x, y, and z.
Figure 26-15. Series of T1 images of the femoral vessels acquired
superior to inferior demonstrates the phenomena of countercurrent
flow-related enhancement. Note that on the first few images (the
most inferior slices) there is very bright signal in the
countercurrent venous flow (white arrows), whereas the arteries
remain dark (black arrow in first image). Notice that as the slices
become deeper within the volume, there is less signal until it is
nearly dark on the last image. Note that the periphery is brighter
than the center as opposed to the schematic in Figure 26-13. This
is because higher velocities in the center preferentially result in
more TOF loss in this patient, while if the flow is slower, the center
will have higher signal as seen in Figure 26-13.
Questions
26-1 Normal intravascular flow is usually:
(a) turbulent flow
(b) plug flow
(c) laminar flow
(d) all of the above
(e) none of the above
26-7 T/F The Reynolds number (Re) can predict laminar versus
turbulent flow.
26-8 T/F FRE penetrates deeper slices when flow is cocurrent rather
than countercurrent.
Magnetic Resonance Imaging, 3rd ed, vol 1. St. Louis: Mosby; 1999.
27
MR Angiography
Introduction
In this chapter, we will discuss the topic of MR angiography (MRA). As
in the last chapter, MRA may at first appear very complicated, but we'll
try to present the major concepts in a simplified fashion. There are
three main MRA techniques:
1. TOF (time-of-flight) MRA
2. PC (phase contrast) MRA
3. CE (contrast-enhanced) MRA
TOF and PC techniques can be performed using two-dimensional Fourier
transform (2DFT) or three-dimensional FT (3DFT). CE MRA is performed
with the 3D technique. Thus, there are a total of five different
methods:
1. 2D-TOF MRA
2. 2D-PC MRA
3. 3D-TOF MRA
4. 3D-PC MRA
5. 3D-CE MRA
Each of these techniques lends itself to a different type of clinical
application.
TOF MRA
TOF MRA is based on flow-related enhancement (FRE; discussed in the
previous chapter) in a 2D or 3D gradient-echo (GRE) technique.
(Remember that in GRE imaging, TOF losses do not play an important
role.) Usually, flow compensation is used perpendicular to the vessel
lumen.
2D-TOF MRA. Figure 27-1 depicts a typical pulse sequence for 2D-TOF
MRA. A presat (presaturation) pulse is applied above or below each slice
to eliminate the signal from vessels flowing in the opposite direction.
Usually a short TR (about 50 msec), a moderate flip angle (45° to 60°),
and a short TE (a few millisecond) are used. An example is seen in
Figure 27-2.
3D-TOF MRA. Figure 27-3 depicts a pulse sequence diagram (PSD) for a
3D-TOF MRA. Here, a slab of several centimeters (usually about 5 cm) is
obtained that contains up to 60 slices.
PC MRA
PC MRA is based on the fact that the phase gain of flowing blood
through a gradient is proportional to its velocity (assuming constant
velocity). We saw in the previous chapter that phase (φ) and velocity
(v) are related by φ = ∫ωdt = ∫(γGvt)dt = 1/2γGvt2 Therefore, knowing
the phase at any point in time allows us to calculate the velocity.
The most common method to employ PC MRA is by the use of a bipolar
gradient (Fig. 27-4A). This process is called flow encoding. Because the
two lobes in this bipolar gradient have equal area, no net phase change
is observed by stationary tissues (Fig. 27-4A). However, flowing blood
will experience a net phase shift proportional to its velocity (assuming
a constant flow velocity) (Fig. 27-4B). This is how flow is distinguished
from stationary tissue in PC MRA. Figures 27-5 and 27-6 illustrate the
PSDs for 2D-PC and 3D-PC MRA, respectively.
There are several features unique to PC MRA, as the following
discussions demonstrate.
Question: What are the “magnitude” image and the “phase” image?
Answer: In PC MRA, you not only get an image of the blood vessels
(magnitude image); you also get an image that shows you the direction
of flow (phase map). The phase image would tell you whether the flow
is right-left, superior-inferior, or anterior-posterior. An example is
determination of hepatofugal versus
Advantages of PC MRA
1. The capability to generate magnitude and phase images
2. Superior background suppression
3. Less sensitive to intravoxel dephasing or saturation effects
Figure 27-7. Two phase contrast images (phase maps) through the
root of the aorta in a patient with aortic stenosis. A: VENC = 150
cm/sec; B: VENC = 300 cm/sec. Note in (A) the bright central
signal (arrow) that may indicate regurgitant flow; however, the
signal is surrounded by darker signal with abrupt transition from
black pixel to white pixel diagnostic of aliasing. Increasing the
VENC in (B) eliminated the aliasing. No bright signal indicative of
regurgitation was observed and an accurate maximal velocity
determination was now possible.
Disadvantages of PC MRA
1. Takes longer to do
2. More sensitive to signal losses caused by turbulence and by dephasing
on vessel turns (e.g., carotid siphon)
3. The need to guess the maximum flow velocity in order to select an
optimum VENC
Figure 27-8. Coronal (A) and axial (B) thick slab (5 cm) 2D-PC MRV
(magnetic resonance venogram) of the brain in normal patients.
CE MRA
CE MRA is different than either TOF or PC imaging since CE MRA is
primarily dependent on the T1 properties of gadolinium in the
vasculature rather than the flow properties per se. This technique has
been made possible due to the advent of high performance gradients
(more in Chapter 30) that permit very rapid GRE imaging and the use of
the paramagnetic contrast agent gadolinium. This allows imaging to be
achieved during the transit time and hence peak T1 shortening of the
gadolinium. This technique is therefore very dependent on the precise
timing of the arrival of the bolus of gadolinium in the vessel of interest.
The plane of imaging is usually in the plane of the vessel (usually
coronal) as opposed to 2D-TOF technique in which the imaging plane is
usually orthogonal to the vessel of interest. Imaging this way increases
coverage while maximizing resolution. Since this technique is more
reliant on the T1 properties than any flow properties, it is very
resistant to dephasing artifacts that are seen in some of the other
techniques.
Figure 27-10. Axial 3D phase contrast MIP (maximum intensity
projection) image of the renal vessels shows normal flow without
stenosis. This is a magnitude image reflecting higher velocities in
the renal arteries represented by brighter signal.
Advantages
1. Rapid technique
2. Resistant to dephasing (e.g., from turbulent flow)
3. Large field of view with good resolution
4. Excellent SNR
Disadvantages
1. Dependent on timing (artifacts or venous contamination may occur)
2. Requires intravenous injection for administration of gadolinium
3. No directional information
See Figures 27-11 through 27-14 for examples.
Table 27-1 contains a summary of some of the major clinical
applications of the five methods of MRA discussed above.
Figure 27-12. Coronal source images (A and B) show a focal high-
grade stenosis of the left proximal main renal artery (arrow in A),
whereas the origin of the right main renal artery is normal (arrow
in B). Note that because these are thin slices, the images do not
usually show the entire vessels. The MIP (C) image in this patient
shows the left renal artery stenosis (arrow) and an accessory right
renal artery (arrowhead).
Table 27-1
Figure 27-16. CE MRA MIP image (A) and comparison 2D-TOF MIP
(B) show occlusion of the left, internal carotid artery in the same
patient as Figure 27-2.
Disadvantages of MIP
The major drawback of MIP is that bright structures other than flowing
blood may potentially be included in the mipped image. Examples are
fat, subacute hemorrhage, and the posterior pituitary gland (Fig. 27-
17). This problem is mainly with TOF MRA and not with PC MRA.
Figure 27-17. Sagittal MIP (A) image shows worrisome signal for
aneurysm near the anterior communicating artery and the internal
carotid terminus. Additional sagittal T1 (B) shows the high signal to
be pituitary hemorrhage and not flow-related enhancement.
Saturation Effects
Saturation effects refer to the gradual loss of longitudinal
magnetization caused by repeated excitation radio frequency (RF)
pulses. This, in turn, leads to loss of signal (and thus reduced SNR). This
problem usually arises in a 2D acquisition in which flowing blood has to
travel within (rather than through) a slice or in a 3D acquisition in
which the blood travels through a thick imaging volume (or slab). In
such a situation, saturation effects may cause the distal portion of a
vessel not to be included in the image.
Black-Blood MRA
Black-blood MRA is a subset of TOF techniques, which accentuates the
TOF losses resulting in flowing blood appearing dark rather than bright.
Rapidly flowing blood (arterial flow), as discussed in the previous
chapter, demonstrates TOF signal losses. Slowly flowing blood (venous
flow) has higher intensity. Various flow presat pulses and dephasing
methods via gradients are employed in this technique to render flowing
blood black. Note that the maximum intensity projection is replaced by
a minimum intensity projection algorithm.
Advantages
1. This technique does not overestimate the degree of stenosis as much
as bright-blood TOF MRA.
2. Dephasing in vessel turns that mimic stenosis with bright-blood TOF
MRA is not a problem here.
Disadvantages
1. Calcified plaques may also be dark and therefore invisible. Thus, this
technique may underestimate the degree of stenosis.
2. Other black materials (such as air or cortical bone or calcification)
may mimic flow.
Fresh-Blood Imaging
3D-fresh blood imaging (FBI) MRA is a new technique based on the fact
that vessel signal intensity is dependent on blood flow (or cardiac
phase) in T2-weighted images. The early systolic phases (0 ˜ 200 msec
after the R wave) show low signal intensity for arteries (high velocity
with TOF losses) and high signal intensity for veins (lower velocities
with no significant TOF loss), whereas the diastolic phases (400 ˜ 600
msec after the R wave) show high signal intensity for both arteries and
veins (no significant TOF loss). Bright-blood MRA is achieved by
subtracting systolic images from diastolic images. The 3D-FBI method
employs an electrocardiography (ECG)-gated 3D half-Fourier fast
spinecho sequence triggered for systolic and diastolic acquisitions. The
ECG triggering time is an important factor influencing the blood signal
intensity in the vessel of interest. An “ECG-prep scan” is typically used
to produce 2D half-Fourier FSE single-slice images at various triggering
delay times. An appropriate ECG delay time is determined for the
vessel of interest and later applied in the 3D half-Fourier FSE
acquisition synchronized by ECG gating for every slice encoding. An
example is seen in Figure 27-25.
Advantages
1. Non-contrast technique; gadolinium not required
Figure 27-25. Coronal FBI MIP image of the lower extremity
shows abrupt occlusion of tibioperoneal trunk, peroneal and
tibial arteries, as well as geniculate collaterals.
Disadvantages
1. Sensitive to selection and timing of the triggered acquisitions
2. Motion artifacts and blurring
3. Longer scan time than CE-MRA
Key Points
1 Three main MR angiography (MRA) techniques
exist: (i) time-of-flight (TOF), (ii) phase contrast
(PC), and (iii) contrast enhanced (CE).
2 Both TOF and PC MRA can be performed in 2D or
3D.
3 TOF MRA is based on FRE.
4 PC MRA is based on velocity-induced phase
changes.
5 PC MRA techniques allow acquisition of magnitude
and phase images.
6 Phase images provide information regarding the
direction of flow (not provided by TOF MRA).
7 VENC (velocity encoding) is a parameter that needs
to be input by the operator when doing PC MRA. It
represents the maximum flow velocity before aliasing
occurs.
8 If VENC is too low, aliasing is more likely to occur.
If VENC is too high, slow flow and small vessels are
not well visualized. Therefore, low VENC is good for
visualization of slow (venous) flow and small
branches. High VENC is good for highvelocity
(arterial) flow.
9 CE MRA is based on a rapid 3D GRE sequence that
is timing sensitive and takes advantage of the
powerful T1 shortening of gadolinium.
10 CE MRA is the most resistant to dephasing of the
bright-blood techniques.
11 Time-resolved imaging of contrast kinetics
(TRICKS) provides rapid multiphase imaging while
retaining much of the SNR. TRICKS is useful for high-
flow malformations or differential flow conditions
(e.g., lower extremity run offs).
12 MIP (maximum intensity projection) is an
algorithm used in MRA in which the highest intensity
dots in space are connected to generate a three-
dimensional-looking image of the vessels.
13 3D imaging is subject to saturation effects (as
blood enters deeper slices).
14 Saturation effects can be minimized by several
methods: (i) decreasing the flip angle, (ii) increasing
TR, (iii) using gadolinium, (iv) MOTSA, or (v) TONE.
15 MOTSA (multiple overlapping thin-slab
acquisition) uses several overlapping thin slices to
reduce saturation effects. A potential problem is
Venetian blind artifacts.
16 TONE (tilted optimized nonsaturating excitation)
uses a ramped RF flip angle (small α at the entry
slice and larger α at the exit slice) to reduce
saturation effects.
17 TOF MRA tends to overestimate the degree of
stenosis (due to dephasing effects).
18 Black-blood MRA is based on TOF signal losses.
Instead of maximum intensity projection, a minimum
intensity projection algorithm is used. This technique
overcomes the problem of overestimating the degree
of stenosis.
19 Fresh-blood imaging (FBI) is a new TOF bright
arterial signal MRA technique that takes advantage of
the differing velocities and TOF losses between
systole and diastole.
Questions
27-1 The main MRA techniques include
(a) TOF MRA
(b) PC MRA
(c) CE MRA
(d) all of the above
27-8 Match
i. ramped RF
ii. Venetian blind artifact
iii. deeper FRE
iv. magnitude and phase images
v. low VENC
with
(a) MOTSA
(b) TONE
(c) aliasing
(d) countercurrent flow
(e) PC MRA
27-9 T/F In the MIP algorithm, the pixel with the highest intensity is
selected for each slice.
Introduction
Cardiac MRI is arguably the most difficult MRI examination to perform.
Cardiac MRI must overcome not just respiratory motion, but also
cardiac motion that cannot be suspended for the image. Additionally,
there are a variety of nomenclature approaches to the pulse sequences,
some of which are fairly unique to cardiac imaging while others are
used in other organ systems all contributing to difficulty in
understanding cardiac MRI. These different nomenclature systems are
superimposed on the option of static versus cine imaging, and
functional or physiologic imaging which all combine to challenge not
just the patient and MRI technologist, but the MR physicist and
radiologist.
Static Imaging
All of the previously described pulse sequences in this book have been
or are used to produce static cardiac images. These images are usually
divided into two categories: (i) bright blood and (ii) dark blood.
Unfortunately, the appearance of blood as either bright or dark is a
consequence of the specific pulse sequence and not a particular
sequence one can select on
the scanner. Blood signal has dark T1 and bright T2 signal intrinsically.
Knowledge of the intrinsic signal of blood combined with superimposed
time-of-flight effects (time-of-flight [TOF] loss or flow-related
enhancement [FRE]) in a particular pulse sequence determines whether
blood is bright or dark.
Fast Spin Echo and HASTE. Spin-echo imaging has generally been
abandoned due to excessive scan times, and it is now replaced by
either FSE or HASTE that provides good anatomic detail, intrinsic dark-
blood signal due to TOF loss, and appropriate T1 or T2 weighting based
on TR and TE. FSE with ECG gating and either breath-hold or navigator-
echo gating result in good image quality; however, the drawback is
lengthy scan times. HASTE sequences have shorter scan times and are
usually obtained in a single R-R interval; however, the SNR will be less
due to 1/2NEX signal averages.
Although FSE and HASTE sequences are usually dark blood, some areas
of slow blood flow from physiologic, pathologic, or in plane flow can
result in minimal expected TOF loss and unexpected bright-blood signal
confounding interpretation. This bright signal can be minimized by
using a double-inversion recovery (DIR) sequence. This sequence uses
a nonslice-selective 180° RF pulse followed immediately by a slice
selective 180° RF pulse and an appropriate T1 that is influenced by the
R-R interval (T1 usually 600 msec) that nulls the signal of inflowing
blood, resulting in a more robust black blood effect (Figs. 28-4 and 28-
5). Due to some intrinsic changes in the position of the slice from the
initial 180° RF pulses and the actual filling of k-space about 600 msec
later, the slice-selective 180° RF pulse is usually twice the nominal
thickness of the slice to ensure that all the myocardium is reinverted.
Additionally, to allow full recovery of the longitudinal magnetization, k-
space is usually filled every other R-R interval unless the patient is
bradycardic.
Figure 28-4. Prospective gated single slice/single phase (mid-
diastole) fast spin-echo DIR sequence with an echo train length = 6.
Notice the shorter trigger delay (TD) driven by the requirement to
initiate the pulse sequence with the nonselective 180° inversion
pulse (dashed line in pair) immediately followed by a slice-
selective 180° inversion or “reinversion” pulse (solid line in pair).
This schematic shows acquisition of echoes every R-R interval,
which occurs only in bradycardic patients. The R-R interval is
usually shorter requiring echo acquisition every other R-R interval
or heartbeat.
each image is acquired in the same cardiac phase. Static images are
either acquired as a single slice/single phase (usually mid-diastole) or a
multislice/single phase where multiple slices are acquired in an
interleaved fashion. The single image/single phase technique fills k-
space during mid-diastole and results in less motion since it minimizes
changes of cardiac phase due to beat-tobeat variability; however,
usually only one slice or at most two slices can be acquired during a
single breath-hold (Fig. 28-2). Multislice/single phase is more efficient,
but has more motion effects due to beat-to-beat variability and
misregistration between slices due to the changing position of the heart
at different phases of the cardiac cycle.
Figure 28-5. A: Direction of longitudinal magnetization
immediately after the nonselective and slice-selective 180° RF
pulses. These pulses are usually applied during late diastole
immediately after the R wave is detected. Solid arrows represent
longitudinal magnetization in the myocardium, while dashed lines
represent blood. B: Direction of longitudinal magnetization
immediately before the acquisition of echoes and filling of k-space.
The TI is usually chosen to coincide with acquisition of k-space
during mid-diastole. Solid arrows represent amplitude of
longitudinal magnetization in the myocardium, while dashed lines
represent blood. Notice that the dashed arrows that were in the
imaging slice in (A) have now moved to the left with the direction
of flow and will not give off signal due to their movement outside
the imaging slice. The prior dashed arrows that were inverted have
now become nearly zero in longitudinal magnetization (gray dashed
arrows).
signal; however, the contrast between the blood pool and the
myocardium will be decreased due to the enhancing myocardium.
Cine Imaging
Cine imaging is similar to static imaging; however, instead of getting a
single image for a single slice, you obtain a series of images obtained at
different phases within the cardiac cycle for a single slice resulting in a
single slice/multiphase acquisition. This permits visualization of the
motion of the heart, valves, and any abnormal structures within the
heart. GRE and True FISP sequences
provide this capability since FSE sequences take too long to acquire the
required multiple phases per slice. In cine imaging, the user defines
how many phases within the cardiac cycle per slice (usually 15 to 25)
are acquired. The goal for temporal resolution between different
phases should be around 50 msec. For example, a patient with an R-R
interval of 1000 msec (60 beats/min) with a cine sequence with 20
phases will have a temporal resolution of 1000 msec/20 phases = 50
msec/phase. The benefit of increased phases is better temporal
resolution and more precise physiologic calculations (see later
discussion) with smoother cine loop viewing; however, the drawback is
that it will take longer to scan. The relationship between phases, heart
rate, and views per segment is given below.
Figure 28-9. The multiple TRs in the IR-prepped GRE or True FISP
delayed enhancement sequences are fewer in this example for
clarity, but usually number 20 to 25; in the case of a single-shot
sequence, the number of TRs would equal all the lines of k-space.
The time to inversion is variable or sliding along the train of TRs;
however, the nominal TI is in the middle of the TR train where the
lowest phase-encode step is applied. The solid line in the
longitudinal magnetization (Mz) indicates abnormal enhancing
myocardium, while the dashed line reflects normal myocardium.
Notice that ideally the normal myocardium crosses the null point
exactly in the center of the train of TRs. Finally, there is no
acquisition of k-space in the next R-R interval due to the need to
allow time for longitudinal magnetization recovery.
Example: How long will it take to perform a prospective cine sequence
using a True FISP sequence with 20 phases in the cardiac cycle, R-R
interval, or time of segment acquisition = 1000 msec with a 20%
arrhythmia reject window (800 msec of scanning per R-R), TR = 4 msec,
and 160 phase-encode steps or views, and NEX = 1.
800 msec/20 phases = 40 msec/phase (temporal resolution)
(40 msec/phase)/(4 msec/TR) = 10 TRs or 10 views/segment
160 views/10 views/segment = 16 segments
Total scan time = (16 segments)(1000 msec/segment) = 16,000 msec
or 16 sec
If the TR is longer (GRE for instance) or if more phases are required,
then the total scan time will
Overview
Proton magnetic resonance spectroscopy (MRS) is an MR-based chemical
analytical technique that can be performed on any high field magnet.
Rather than providing images, it usually provides spectra consisting of
individual peaks, the chemical shift of which from a universal standard
helps identify the species comprising the peak. The normal spectrum
from gray matter is different from the normal spectra from white
matter (Fig. 29-1). MRS is the equivalent of the nuclear magnetic
resonance (NMR) used in organic chemistry—only now performed on a
patient in a whole-body MR magnet.
MRS is a functional technique. It can identify abnormalities not seen on
conventional MRI, for example, invasion of a glioblastoma multiforme
(GBM) into adjacent brain where there is no enhancement or T2
abnormality on MRI. It can distinguish between two or more
abnormalities that appear the same on MRI, for example, recurrent
GBM versus radiation necrosis.
As in NMR, the area under a given peak is proportional to the number of
protons contributing to the peak. So, taking ethanol (CH3— CH2—OH) as
an example, the area under the methyl (CH3) peak would be 3 (in
relative units), the area under the methylene (—CH2—) peak would be
2, and the area under the hydroxyl (—OH) peak would be 1. Since the
peaks generally have the same shape, this ratio of peak areas also holds
for peak heights. In general, peak height is proportional to the
concentration of a given species. MRS requires a species to be present
in at least a 1 mM concentration to be seen.
The above discussion is somewhat simplified. In actuality, the protons
on adjacent carbons influence the net magnetic field on each other. So
the magnetic field experienced by the methyl protons in ethanol would
be influenced by the adjacent methylene protons. In quantum
mechanical sense, these two protons can either both point up, both
point down, or one up and one down or one down and one up. When
they both point up the net magnetic field experienced by the adjacent
methyl protons is slightly higher than B0 and when they both point
down, it is slightly lower than B0, which shifts the peaks. Up-down and
down-up cancel each other and results in the same peak position as one
would have without any adjacent protons. This is known as “spin-spin
splitting,” and it results in a big peak with area or peak height of 3 to
be split into three peaks (called a “triplet”) of area 1-2-1 (Fig. 29-2).
Spin-spin splitting or “J-coupling” smears the chemical shift, making
peak identification more difficult. It increases with increasing TE.
MRS in the brain is generally performed in conjunction with MRI which
identifies the abnormality. For single voxel techniques, a volume of 8 cc
is generally recommended at 1.5 T. A smaller voxel can be used but the
signal-to-noise ratio (SNR) will be reduced (Fig. 29-3). Since peak height
is generally proportional to field strength, a smaller voxel can be used
at 3 T, reducing partial volume averaging. Generally, a second spectrum
is acquired in the contralateral normal brain.
Figure 29-1. Normal spectra for white matter (1) and gray matter
(2). Note: the higher Cho level in normal white matter (third
highest peak).
Patterns of Disease
Solid tumors have elevated Cho and decreased NAA. When they become
necrotic, lipid and lactate become elevated as well (Fig. 29-11).
Although elevated Cho/NAA is a sensitive marker for cancer, it is
nonspecific. The same pattern is seen in tumefactive multiple sclerosis
and acute disseminated encephalomyelitis.
MRS can distinguish recurrent GBM (elevated Cho/NAA) from radiation
necrosis (no elevated Cho but elevated lipid) (Fig. 29-9B). It can
distinguish necrotic lymphoma (elevated Cho) from toxoplasmosis (no
elevated Cho).
By placing the voxel outside the area of enhancement or T2
prolongation, MRS can distinguish a metastasis (no elevated Cho) from
an infiltrating glioma (elevated Cho) (Fig. 29-12).
Figure 29-12. A-C: Gastric metastasis. Cho/Cr is not elevated
outside area of enhancement. D-F: GBM. Cho/Cr is elevated both
in (D) and outside (E) enhancement (Cho is the tallest peak in the
spectra; compare with normal Cho/Cr in F). Elevated Cho/Cr in (E)
represents infiltration of tumor outside area of enhancement.
Key Points
1 MRI images any mobile proton-containing species.
MRS shows up to 10 different proton-containing
chemical species, allowing more specific diagnoses.
2 MRS shows abnormalities, for example, infiltration
of a GBM, not seen on MRI.
3 J-coupling decreases peak height linearly with TE,
for example, glutamate and glutamine, which are
seen jointly as Glx on the left-hand shoulder of NAA
(2.0 to 2.5 ppm).
4 Species are identified based on their chemical shift
in parts per million (ppm) relative
P.351
to tetramethylsilane at 0 ppm (by definition).
5 In a normal brain, NAA (2.0 ppm) indicates normal
neurons, Glx (2.0 to 2.5 ppm) represents glutamate
and glutamine, Cr (3.0 ppm) represents energy, Cho
(3.2 ppm) represents membrane turnover, and mI
(myoinosotol at 3.5 ppm) is elevated in glial tumors
and hyperosmolar states. Choline is elevated in all
tumors and a few other states, for example, multiple
sclerosis. Lactate (1.3 ppm) indicates anaerobic
glycolysis seen in stroke and GBMs. Lipid (0.9 to 1.2
ppm) is seen with membrane breakdown seen in
infection, radiation necrosis, and tumors.
Questions
29-1 Which of the following metabolites is normally not seen?
(a) NAA
(b) Lactate
(c) Cho
(d) Cr
(e) mI
29-5 mI is elevated in
(a) Alzheimers
(b) Hyperosmolar states
(c) Glial tumors
(d) All of the above
29-7 T/F Performing MRS after giving gadolinium raises the Cho/NAA
ratio.
Introduction
This chapter briefly discusses the new technology of high performance
gradients. As you know by now, gradients have many purposes,
including slice selection, spatial encoding, flow compensation (FC), and
spoiling, rewinding, and presaturation. It should be fairly obvious that
every time you use a gradient, you are lengthening the pulse cycle
(thus increasing minimum TE).
Consider the two gradients in Figure 30-1A and 30-1B. The gradient in
Figure 30-1A has half the strength of the one in Figure 30-1B but twice
the duration. Thus, both these gradients have the same area (shaded
area). They both achieve exactly the same result (e.g., phase shift) on
stationary spins, but the second one is twice as fast and allows one to
reduce the echo delay time TE. Therefore, the first requirement of a
high performance gradient is a higher maximum strength.
When we discuss high performance gradients, not only do we want to
achieve a stronger magnitude or strength, but we also want the
maximum strength to be achieved in as short a time as possible (i.e., a
short rise time) to minimize the duration of the gradient. Therefore,
the second issue is how fast a gradient can reach its plateau (Fig. 30-2).
The ratio of the maximum gradient (Gmax) to the rise time (tR) is called
the slew rate (SR).
Advantages
1. Shorter cycles are possible. As discussed previously, a stronger
gradient can be applied over a shorter duration. Consider the
example of FC. Figure 30-3 demonstrates two FC gradients that
achieve the same thing, but the second one has higher strength (G′ >
G) and is thus applied over a shorter period (T′ < T). (As an aside,
because of the quadratic nature of phase accumulation for flowing
spins, there is no linear or inverse linear relationship between
gradient strength and application time, i.e., doubling the gradient
strength does not halve the time of gradient application.) Higher
order motion (e.g., acceleration, jerk) requires
the addition of even more gradient lobes. You can see how high
performance gradients can save a lot of time during each cycle,
allowing for shorter TEs (reducing dephasing) and TRs (for fast
scanning).
Figure 30-1. The gradient in (B) has twice the strength of the
one in (A) but half the duration. Consequently, they both have
exactly the same area and thus achieve the same results for
stationary spins. However, the one in (B) has the advantage of
being faster, a condition that is required for fast scanning.
Key Points
High performance gradients have revolutionized MR
imaging. In short, a higher magnetic strength allows
a shorter duration, thus reducing the cycle time. This
in turn reduces the minimum TE and TR, which
shortens the acquisition time. In summary, high
performance gradients allow many new and improved
features, including faster scanning (including EPI,
GRASE, 3D-FSE, CE MRA, etc.), improved spatial and
temporal resolution, smaller FOV, high-resolution
MRA (both TOF and PC) with improved visualization
of small vessels and slow flow, and improved imaging
of CSF flow, just to name a few.
Questions
30-1 High performance gradients require
(a) high gradient strength (Gmax)
(b) short rise time (tR)
(c) both
(d) none
30-4 Two gradients with different profiles but the same area (Fig.
30-1) will have the same effects (i.e., phase changes) on
(a) stationary spins
(b) flowing spins
(c) both
(d) none
31
The Many Combinations of MRI
Introduction
Thus far we have studied a vast array of MRI sequences that seem
somewhat related to one another, but also, quite different. The pulse
sequences that we have looked at are the basic sequences used in MRI;
however, there are many more combinations of sequences that can be
performed and in this chapter we will look at how these interrelate.
time of 200 to 300 msec to null normal myocardium (Table 31-3 and Fig.
31-3).
Figure 31-1. A: An SSFSE FLAIR image of the brain in an
uncooperative patient. There are corresponding SSFSE (B), T2 (C)
with marked motion, and later FLAIR (D) images when the patient
was more cooperative. There are periventricular white matter
ischemic changes and an old right parietal stroke. This example
demonstrates a novel approach for newer fast imaging sequences
that can be used in uncooperative patients.
(180° pulse) SE
GRE
FSE
EPI
Multiple (single
MT shot)
EPI
SSFSE
GRE sequences) and any additional repetition of the readout (with its
accompanying different phase-encode steps) such as additional 180° RF
pulses in FSE or the alternating gradients in both the frequency and
phase directions in EPI. An overarching idea is that the readout
following a single RF pulse can either fill a single line of k-space,
multiple lines (segmented or multishot
Most gradient sequences fill a single line of k-space per RF; however,
newer techniques are being implemented such as an FGRET (fast
gradient-echo train) that is useful in cardiac perfusion imaging (Table
31-7 and Fig. 31-6). This is a GRE sequence with a segmented filling of
k-space (four lines per RF pulse).
Questions
31-1 T/F A prep pulse is required in every MRI pulse sequence.
4. Bushberg JT, Seibert JA, Leidholdt EM, et al. The Essential Physics
of Medical Imaging. 2nd ed. Philadelphia: Lippincott Williams &
Wilkins; 2002.
7. Erickson SJ, Cox IH, Hyde JS, et al. Effect of tendon orientation
on MR imaging signal intensity: a manifestation of the “magic angle”
phenomenon. Radiology. 1991;181:389-392.
9. Henkelman RM, Hardy PA, Bishop JE, et al. Why fat is bright in
RARE and fast spin-echo imaging. J Magn Reson Imaging.
1992;2:533-540.
10. Hennig J, Nauerth A, Friedburg H. RARE imaging: a fast imaging
method for clinical MR. Magn Reson Med. 1986;3:823-833.
12. Kapelov SR, Teresi LM, Bradley WG, et al. Bone contusions of the
knee: increased lesion detection with fast spin-echo MR imaging
with spectroscopic fat saturation. Radiology. 1993;189:901-904.
13. Oppenheim AV, Willsky AS, Nawab S. Signals and Systems. (2nd
Ed.). New Jersey: Prentice-Hall; 1996.
17. Stark DD, Bradley WG, eds. Magnetic Resonance Imaging. Vol. 1-
3. 3rd ed. St Louis: Mosby; 1999.
18. Ulug AM, Moore DF, Bojko AS, et al. Clinical use of diffusion
tensor imaging for diseases causing neuronal and axonal damage.
Am J Neuroradiol. 1999;20:1044-1048.
22. Simonetti OP, Kim RJ, Fieno DS, et al. An improved MR imaging
technique for the visualization of myocardial infarction. Radiology.
2001; 218:215-223.
23. Foo TK, Bernstein MA, Aisen AM, et al. Improved ejection
fraction and flow velocity estimates with use of view sharing and
uniform repetition time excitation with fast cardiac techniques.
Radiology. 1995;195:471.
Chapter 1
1-1 (a) See Figure 1-17
(b) See Figure 1-18
1-2 ei(x+y) = eixeiy. So cos (x + y) + isin (x + y) = (cos x + isin x)(cos y +
isin y) = (cos x · cos y − sin x · sin y) + i(cos x · sin i + sin x · cos y)
1-3 sinc(0) = sin(0)/0 = limd/dx (sin x/x) xrarr;0
= cosx/1 = cos0/1 = 1/1 = 1 (x = 0)
1-4 (a) e−1 = 0.37 (b) e−2 = 0.14
(c) e−1 = 0.37
1-5 d/dt(Ae−t/T) = A (−1/T)e−t/T, which is −A/T at t = 0. This is the
slope of the line tangent to the curve at t = 0 that crosses the t-axis at
t = T.
1-6 ln (ex) = x = ln 8 = ln 23 = 3ln 2 = 3 × 0.693 = 2.079
1-7 (a) 0 (b) 0.5 (c) 1
(d) 0 (e) 1 (f) 0.5
(g) 0 (h) −1
Chapter 2
2-1 (a) 14.9 MHz
(b) 21.3 MHz
(c) 42.6 MHz
(d) 63.9 MHz
(e) 85.2 MHz
(f) 127.8 MHz
2-2 T
2-3 F (only mobile protons)
2-4 F
2-5 T
2-6 F (speed of light)
2-7 T
2-8 F
2-9 F
2-10 T
2-11 T
Chapter 3
3-1 T
3-2 F
3-3 T
3-4 T
3-5 T
3-6 F (twice)
3-7 T
Chapter 4
4-1 (a) T (b) F (c) F (d) T
4-2 b
4-3 d
4-4 F (by T2*)
4-5 c
4-6 (a) i (b) ii
Chapter 5
5-1 (a) 1.56 (b) 90 msec
(c) 0.72 and 1.05
(d) 1.28, 50 msec, 0.88 and 1.28
(e) 2.10
5-2 H2O/fat = 0.25 and 1.63; CSF/GM = 0.41 and 1.40
5-3 b
5-4 c
5-5 (a) N(H)e−TE/T2 (i.e., ideal T2W)
(b) N(H)(1 − e−TR/T1) (i.e., ideal T1W)
(c) N(H)(i.e., ideal PDW).
5-6 (a) ii (b) i (c) iii (d) iv
Chapter 6
6-1 T
6-2 (a) iii (b) i (c) iv
(d) i (e) ii
6-3 T
6-4 (a) i (b) ii (c) iv
Chapter 7
7-1 Setting 1 − 2e−t/T1 = 0, we get e−t/T1 = 1/2 so −t/T1 = ln (1/2) =
−0.693 so t = 0.693 T1
7-2 SαN(H) (1 − 2e−TI/T1) (1 − e−TR/T1) and N(H) (1 − 2e−TI/T1 − e−TR/T1 +
2e−(TR+TI)/T1) ≅ N(H) (1 − 2e−TI/T1 − e−TR/T1 + 2e−TR/T1) = N(H) (1 − 2e−TI/T1
+ e−TR/T1)
7-3 (a) ii (b) i
7-4 T
Chapter 8
8-1 (a) N(1 − e−TR/T1)e−TE1/T2 and N(1 − e−TR/T1)e−TE2/T2
(b) N(1 − e−TR/T1) e−TE1/T2*
(c) 0.61 and 0.37
8-2 (a) i (b) iii (c) ii
8-3 F (not that due to spin-spin interactions)
Chapter 9
9-1 T
9-2 (a) F (b) F
9-3 T
9-4 T
Chapter 10
10-1 (a) 4.7 mT/m (b) 1 mm
10-2 f
10-3 T
10-4 T
Chapter 11
11-1 (a) i (b) iii (c) ii
11-2 T
11-3 360°/128 = 2.8°
11-4 (i) d (ii) a
11-5 T
Chapter 12
12-1 b
12-2 c
12-3 T
12-4 F(1/ )
12-5 d
12-6 F
Chapter 13
13-1 T
13-2 F (phase-encoding gradient)
13-3 T
13-4 T
13-5 T
13-6 F
13-7 F (there is conjugate symmetry)
13-8 F
Chapter 14
14-1 a, d, e
14-2 (a) 512 sec = 8 min, 32 sec
(b) 5120 sec = 85 min, 20 sec = 1 hr, 25 min, 20 sec!
14-3 b
Chapter 15
15-1 (a) 15 cm (reduces the minimum FOV)
15-2 b
15-3 61°
15-4 d
15-5 F (increases)
15-6 b
15-7 a
Chapter 16
16-1 T
16-2 (a) T (b) F (cycles/cm)
16-3 T
16-4 d
16-5 (a) T (b) T
16-6 T
Chapter 17
17-1 (a) 384 sec = 6 min, 24 sec
(b) 3840 sec = 64 min = 1 hr, 4 min
(c) 6 min, 24 sec
17-2 10 slices
17-7 g
17-8 f
17-9 d
17-10 c
17-11 c
17-12 e
17-13 a
17-14 e
17-15 f
17-16 d
17-17 (a) ii (b) i
Chapter 18
18-1 e
18-2 (a) In terms of number of pixels
(b) In terms of mm
(c) The narrower the BW or the stronger the main magnetic field, the
greater the chemical shift.
18-3 (a) 51.2 (b) 256/51.2 = 5
(c) fewer ghosts
18-4 b
18-5 d
18-6 F (the other way around)
18-7 a
18-8 T
18-9 d
18-10 (a) 48 pixels or 7.5 cm
(b) 2 ghosts
18-11 c
18-12 c
18-13 d
18-14 F (55°)
18-15 a
18-16 c
18-17 e
18-18 T
Chapter 19
19-1 d
19-2 (a) 34 min, 8 sec
(b) 4 min, 16 sec
19-3 F (decreased coverage)
19-4 d
19-5 d
19-6 c
19-7 4 min, 16 sec
19-8 T
Chapter 20
20-1 T
20-2 T
20-3 T
20-4 T
20-5 T
20-6 F (more T1W)
20-7 (a) 15.4 sec
(b) 230.4 sec = 3 min, 50 sec
20-8 F (the opposite)
20-9 T
20-10 b
20-11 T
Chapter 21
21-1 d
21-2 f
21-3 T
21-4 T
Chapter 22
22-1 T
22-2 F
22-3 d
22-4 F
22-5 F
22-6 T
22-7 b
Chapter 23
23-1 e
23-2 d
23-3 a
23-4 T
23-5 T
23-6 F
Chapter 24
24-1 b
24-2 b
24-3 c
24-4 F
24-5 T
Chapter 25
25-1 b
25-2 a
25-3 F
25-4 T
25-5 T
25-6 e
25-7 e
25-8 T
Chapter 26
26-1 c
26-2 F (distal to a stenosis)
26-3 T
26-4 F
26-5 d
26-6 d
26-7 T
26-8 F
26-9 F
26-10 T
26-11 a
26-12 c
26-13 c
26-14 F
26-15 T
26-16 c
Chapter 27
27-1 d
27-2 b
27-3 (a) i (b) i (c) ii (d) ii
27-4 F
27-5 e
27-6 T
27-7 T
27-8 (a) ii (b) i
(c) v (d) iii
(e) iv
27-9 T
27-10 c
Chapter 28
28-1 d
28-2 a
28-3 b
28-4 c
28-5 d
28-6 c
28-7 a
Chapter 29
29-1 b
29-2 a
29-3 c
29-4 c
29-5 d
29-6 F: also elevated in MS and ADEM.
29-7 F: since Cho has a shorter T1 than NAA, the Gd shortens the T1
relatively more for NAA, raising its signal.
29-8 e
Chapter 30
30-1 c
30-2 b
30-3 c
30-4 a
Chapter 31
31-1 F
31-2 T
31-3 F
31-4 e
31-5 T
Appendix C
Abbreviations
α, θ, φ
Alpha, theta, phi. Symbols used to designate an angle
(e.g., flip angle)
Γ
γ. Gyromagnetic ratio (in T/MHz)
Μ
µ. Micron (10-6 m)
Σ
σ. Standard deviation
Τ
τ. Symbol used to designate time
ω, ω0
Omega. Angular (Larmor) frequency (in radians/sec)
2D
Two dimensional
2DFT
Two-dimensional Fourier transform
3D
Three dimensional
3DFT
Three-dimensional Fourier transform
A2D, ADC
Analog-to-digital converter or conversion
ASSET
Array spatial and sensitivity encoding technique
B0
Main external magnetic field
B1
Magnetic field associated with the RF pulse
b-FFE
balanced fast field echo
b-SSFP
balanced steady-state free precession
BW
Bandwidth
CNR
Contrast-to-noise ratio
CSE
Conventional spin echo
DE
Driven equilibrium
EPI
Echo planar imaging
ESP
Echo spacing
ET
Echo train
ETL
Echo train length
FC
Flow compensation
FFE
fast field echo
FFT
Fast Fourier transform
FGR
Fast GRASS
FID
Free induction decay
FIESTA
Fast imaging employing steady-state acquisition
FISP
Fast imaging with steady-state precession
FLAIR
Fluid-attenuated inversion recovery
FLASH
Fast low-angle shot
FOV
Field of view
FSE
Fast spin echo
FSPGR
Fast SPGR
FT
Fourier transform
Gx
Frequency-encoding gradient
Gy
Phase-encoding gradient
Gz
Slice-select gradient
GE
Gradient echo
GM
Gray matter
GMN
Gradient moment nulling
GMR
Gradient-motion rephasing
GRASE
Gradient and spin echo
GRASS
Gradient recalled acquisition in the steady state
GRE
Gradient echo or gradient-recalled echo
HASTE
Half-Fourier acquired single-shot turbo spin echo,
Siemens
Hz
Hertz (1 cycle/sec)
IPAT
Integrated parallel acquisition techniques
IR
Inversion recovery
kHz
Kilohertz
M0
Initial longitudinal magnetization
MAST
Motion artifact suppression technique
MEMP
Multi-echo multiplanar
MHz
Megahertz
mm
Millimeter (10-3 m)
MOTSA
Multiple overlapping thin-slab acquisition
MPGR
Multiplanar GRASS
MP-RAGE
Magnetization-prepared, rapid acquisition gradient echo
MR
Magnetic resonance
MRA
Magnetic resonance angiography
MRI
Magnetic resonance imaging
msec
Milliseconds
MT
Magnetization transfer
MTC
Magnetization transfer contrast
Mxy
Transverse magnetization in the x-y plane
Mz
Longitudinal magnetization
NEX
Number of excitations
nm
Nanometer (10-9 m)
NMR
Nuclear magnetic resonance
NPW
No phase wrap
NFW
No frequency wrap
NSA
Number of signal averages
Nx
Number of frequency-encoding steps in the x direction
Ny
Number of phase-encoding steps in the y direction
Nz
Number of phase-encoding steps in slice-select
direction (in 3D imaging)
PC
Phase contrast
PD
Proton density
PDW
Proton density weighting
PDWI
Proton density-weighted image
pixel
Picture element
POMP
Phase offset multiplanar
pm
Picometer (10-12 m)
ppm
Parts per million
PS
Pulse sequence
PSD
Pulse sequence diagram
PSIF
Opposite of FISP
RARE
Rapid acquisition with relaxation enhancement
Re
Reynolds number
RF
Radio frequency
SE
Spin echo
SENSE
Sensitivity encoding
SMASH
Simultaneous acquisition of spatial harmonics
SNR, S/N
Signal-to-noise ratio
SPGR
Spoiled GRASS
SSFP
Steady-state free precession
SSFSE
Single-shot fast spin echo, GE
STIR
Short TI (or tau) inversion recovery
T1
T1 relaxation time, longitudinal relaxation time, and
spin-lattice relaxation time
T1W
T1 weighting, T1 weighted
T1WI
T1-weighted image
T2
T2 relaxation time, transverse relaxation time, and
spin-spin relaxation time
T2*
T2* relaxation time
T2W
T2 weighting, T2 weighted
T2WI
T2-weighted image
T2*W
T2* weighting, T2* weighted
T2*WI
T2*-weighted image
T
Tesla; period of a periodic signal
TE
Echo delay time (time to echo)
TI
Time to inversion (inversion time)
T0
“Overhead” (dead) time in the pulse cycle
TOF
Time of flight
TONE
Tilted optimized nonsaturating excitation
TR
Repetition time (time of repetition)
tR
Rise time
Ts
Sampling time
TSE
Turbo spin echo
Turbo
Siemens' and Philips' prefix to denote a fast scanning
mode
VB
Variable bandwidth
VENC
Velocity encoding
VEMP
Variable echo multiplanar
voxel
Volume element
WM
White matter
A
Acceleration compensation, 308
Acoustic schwannoma, 231
Acquisition time, 164 165 166 180 181 See also Scan time
Active time, 124 125 127
ADC. See Analog-to-digital conversion (ADC); Analog-to-digital
converter (ADC)
Algebraic manipulation, 173
Aliasing, 131 132 133 134 135 136 132 185 186 187 188
as an analogy, 136 137 138 138
defined, 136
prevention of, 187 188 188
undersampling and, 138 139
Amplitude
of signals, 109 110 151
Analog-to-digital conversion (ADC), 130 defined, 123
Analog-to- digital converter (ADC), 257
Angle
magnitude and, 5 6
of vector, 6 6
Angular frequency (ω), 6 17 26 See also Larmor frequency
Anisotropy, 201
Anisotropy index, 263
Apparent diffusion coefficient, 264
Applied magnetic field
induced and, relationship between, 20
Arc, 3
Array spatial and sensitivity encoding technique (ASSET), 279
Arrhythmia reject window, 329
Artifact(s), 185 186 187 188 189 190 191 192 193 194 195 196
197 198 199 200 201 202 203 204 205 206 207 208 209 210 211
212
central, 203 204 204
chemical shift, 188 189 190 191 192 193 194 195 190 191 192
193 262
CSF flow and, 200 200 212 212
data errors and, 212
in echo planar imaging (EPI) technique, 261 262
external magnetic field, 205 206 207
free induction decay (FID), 204 204
geometric distortion, 105
ghost, 198 199 198 199 208
gradient-related, 207 211 212
image processing, 185 186 187 188 189 190 191 192 193 194 195
196
magic angle, 200 201 202
magnetic susceptibility, 206 207 208 209 210
motion, 147 197 198 199 200 201 198 199 200 201
N/2 ghost, 261
overshoot, 104 See also Ripples effect
partial volume, 196 197
patient-related, 196 197 198 199 200 201
RF-related, 201 202 203 204 205
feedthrough zipper, 205 205
zipper, 203 204 205 204
stimulated echo, 204 205
susceptibility, 261 262
truncation, 189 195 196 196 197
undershoot, 104 See also Ripples effect
zipper, 203 204 205 204
feedthrough, 205 205
Asymmetric field of view (FOV), 270 271 271
Audio frequency, 104 105
Autocalibrating reconstruction for Cartesian sampling (ARC), 279
Auto RF (radio frequency) pulse, 38
Auto shimming, 206
B
Back projections, 111 112 113 113
advantage of, 112
disadvantage of, 112
Band-pass filters (BPF), 272
Bandwidth (BW), 92 99 104 105 106 106 129 130 130
center frequency of, 104 105 106 106
decreasing, and chemical shift artifact, 192 193 194
field of view (FOV) and, 167 168 169 171
Gaussian RF pulse, 104 105
measurement of, 99 100 100
narrower frequency, 101
receiver, 140
and sampling interval (ΔTs), 139 140 141 142 143
signal-to-noise ratio (SNR) and, 177 178
transmission, 140
Bar magnet, 18 18
Bellows device, 327 328 329
b-FFE (balanced fast field echo, Philips), 330
Bi-lobed gradients, 235 235 236
Binary number, 130
Bipolar flow-encoding gradient, 312
bit(s), 130
Black-blood MRA, 323 See also Magnetic resonance angiography (MRA)
fat-saturated, 356
Blood flow, 295 295
normal appearance of, 295 295
Blood motion, 331
Body coil(s), 27
Bounce point artifact, 195
Boundary effect, 195 195
BPF. See Band-pass filters (BPF)
Brain. See also White matter
hematoma in, and tissue contrast, 70 71
hemorrhage in, and tissue contrast, 66 70 72 73
lesions of
tissue contrast and, 61 61
suppression of. See Tissue suppression
tissue contrast in, 60 61
TE (echo delay time), 63 64 63 64
TR (repition time) and, 63 64 63 64
Brownian motion, 177
BW. See Bandwidth (BW)
Byte, 130
C
Cardiac gating, 300
Cardiac motion, 329 330 331
Cardiac MRI, 327
cine gating imaging, 337 338 339 340
cine imaging, 335 336 337
clinical applications, 340
compensation, 327 328 329 330 331
delayed enhancement, 335
GRE and true FISP gating, 335
k-space filling strategies, 331
late gadolinium enhancement, 335
motion effects, 327 328 329 330 331
pulse sequences, 340
static imaging, 332 333 334
Carrier frequency, 130 151
Center frequency, 105
of RF pulses, 129 130 130
Central artifacts, 203 204 204
Cerebrospinal fluid (CSF), 60 294
flow of
artifact caused by, 200 200
imaging of, 156 156
proton density factor for, 62 62 62
recovery and decay curves for TR (repition time) and, 63 64 63 64
suppression of. See Tissue suppression
T1 characteristics, 60 60 62
T2 characteristics of, 60 61 62
Charged nuclei, 19
Chemical saturation pulse, 356
Chemical shift artifact(s), 188 189 190 191 192 193 194 195 190
191 192 193
in EPI, 262
Chemical shift effect of the second kind, 194 195 194
Chemical (Spectral) presaturation, 274 276 288 289 289
advantages of, 288
disadvantages of, 289
CNR. See Contrast-to-noise ratio (CNR)
Cobalt (Co), 21
Cocurrent flow, 303 304
Coil(s)
body, 27
gradient, 26 27 28 98
frequency-encoding (readout) (Gx), 28
phase-encoding (Gy), 28
slice-select (Gz), 28
head, 26
phased-array, 28
quadrature, 28
radio frequency (RF), 26
shim, 28
solenoid, 28
surface, 27
transmit/receive, 27
Combined flow phenomena, 303
Complementary SPAtial Modulation Magnetization (CSPAMM), 337
Complex numbers, 5 153 154 155 156 157 158 159 160
conjugate (Hermitian) symmetry, 153 154
fractional echo, 155 155
magnitude (modulus) and phase image, 155 156 157 156
NEX
½, 153 155 155
¼, 155 155
real and imaginary images, 155 156
Composite signals
sampling of, 141 142 143 142
Conjugate (Hermitian) symmetry, 153 154
Contiguous slice option, 273 273
Contiguous slices, 203 203
Contrast averaging, 226 227 228
Contrast enhanced MRA, 315 354
advantages, 316
disadvantages, 316
Contrast-to-noise ratio (CNR), 287
Conventional spin echo (CSE), 217 218 257
scan time for, 217
Convolution, 131 133
Cosine, 3 4 5
cosine data
data space with, 152 153 153
Cosine function, fourier transform (FT) of, 90 91 92 93 99 100 109
110 134 134
Co-tangent, 3
Countercurrent flow, 303 304
Coverage, 128 129
parameters of, 176 181 182 181
Creatine (Cr), 345
Cross-talk, 101 101 201 202 203 202 203
minimize, 101 102
Crusher gradient(s), 163 163 230
CSE. See Conventional spin echo (CSE)
CSF. See Cerebrospinal fluid (CSF)
Curves
Gaussian, 101 101 104
“Custom-made” techniques, 176
Cycle(s), 6 7
D
Data errors, artifact caused by, 212
Data space, 117 118 119 120 119 145 146 145 171 See also K-space
components of, 152 153 153
definition of, 171
Fourier transform of, 121 121
in time domain, 120
trajectories, 120
Dead time, 127 127 164 165
Decay constant, 9 10
Decay/recovery curves, TR (repetition time) and, 63 64 63 64
Delayed enhancement imaging, 255 356
Delta function, 131 133
Demodulation, 105 106
Deoxyhemoglobin
tissue contrast and, 70
Dephasing, 41 42 43 83 125 146 163 163
intravoxel, 298
odd-echo, 298
in phase spins, 41
90° RF pulse turned off, 41
spin-spin interaction(s), 42 43 83
DFT. See Digital Fourier transform (DFT)
2DFT. See Two-dimensional fourier transform (2DFT)
Diamagnetic substances, 21
magnetic susceptibility in, 21
in MRI, 21
opposite direction in, magnetic field in, 21
reduced magnetic field in, 21
Diamagnetism, 21 206 209
Diastolic pseudogating, 300
Dielectric effects
artifacts caused by, 212 212
Diffusion coefficient, apparent, 264
Diffusion imaging
in echo planar imaging (EPI) technique, 262
Diffusion tensor imaging (DTI)
in echo planar imaging (EPI) technique, 262 263 263
diffusion-weighted imaging, 361
Digital Fourier transform (DFT), 160
Dipolar Hamiltonian, 200
Dipole-dipole interactions, 20 21
Direction, of signal, 153
Discitis, tissue contrast in, 68
Dixon effect, 242 243 244 243
Double inversion recovery (DIR), 332 356
DTI. See Diffusion tensor imaging (DTI)
2D-TOF MRA, 310
3D-TOF MRA, 310
Dysprosium (Dy), 21 206
E
Echo(es), 85 86 108 123 124
asymmetric, 87
delay time. See TE (time to echo)
fractional, 155 155
symmetric, 87 87
Echo planar imaging (EPI), 50 145 354
advantages and disadvantages of, 263 264
artifacts in, 261 262
basic idea in, 257
blipped, 258 259
clinical applications of, 265 266
contrast in, 260 261 261
for diffusion imaging, 262
for diffusion tensor imaging (DTI), 262 263 263
functional, 262 263 264 265
gradient echoes in, 257
k-space trajectory in, 260 260
multishot, 257 259 259
for perfusion imaging, 263 264 265
root pulsing sequence, 260 261 261
scan time in, 260
segmental, 259
sequence, 359
single-shot, 257 258 258
types of, 257 258 259
Echo sampling period (ESP), 260
Echo spacing (ESP), 218 218 219
Echo train length (ETL), 180 218 219 220 221 222 223 220 221 222
223
Eddy currents
artifacts caused by, 207 211
Edema
inversion recovery (IR) imaging of, 78 79 79
magnitude reconstruction, 79 80 81
proton density factor for, 62 62
vasogenic, and tissue contrast, 61 61
Effective TE (TEeff), 219 221
Eigenvalues, 263 264
Eigenvectors, 263 264
diffusion, 265
principal, 263
Electric field, of electromagnetic waves, 16 17
Electromagnetic field, 18 19 20
Electromagnetic spectrum
frequency ranges in, 16 18
windows of, 16 17 18 18
Electromagnetic wave(s)
angular frequency of, 17
characteristics of, 16 17
components of, 16 17 17
electric field of, 16 17
electromagnetic windows and, 18
energy of, 18 18
frequency ranges in, examples of, 18
linear frequency of, 17
magnetic field of, 17
Maxwell's wave theory and, 16 17
RF pulse, 31
speed of light of, 16
Electromagnetic windows, 18
Elliptical-centric technique, 354
Endometrioma, tissue contrast and, 73
Energy levels, of atomic nucleus, 19
Entry phenomenon, 300
EPI. See Echoplanar imaging ( EPI)
ESP. See Echo spacing (ESP)
ETL. See Echo train length (ETL)
Euler's equation, 11 12 12
Even-echo rephasing, 298 299 300
effect, 226
Even function, 152
Fourier transform (FT) of, 94 94
Excitation
definition of, 165
Exponential functions (ex), 8 9 8 9 10
properties of, 13
External magnetic field artifact(s), 205 206 207
External magnetic field (B0), 21 23 24
inhomogeneity, and dephasing, 42 43 83
in Larmor equation, 26 26
with RF pulse transmission, 31 31 32 33
F
Faster imaging, 331
Fast FLAIR (fluid-attenuated inversion recovery), 284 285 286 287
288 286
Fast Fourier transform (FFT), 160
Fast gradient echo train. See FGRET (fast gradient echo train)
Fast imaging with steady-state precession. See FISP (fast imaging with
steady-state precession)
Fast IR, 231
Fast low-angle shot (FLASH), 247 248 249 250 251
characteristics of, 256 256
disadvantages of, 251
multiplanar (MPSPGR), 252
tissue contrast, 248
Turbo, 252 253
Fast spin echo (FSE), 145 257 269 354 356
gating, 332 333
and HASTE, 332
Fast spin-echo (FSE) imaging, 180 218 219 220 221 222 223 224 225
226 227 228 229 230 231
advantages of, 226 227
disadvantages of, 227 228
echo train length (ETL), 218 219 220 221 222 223 220 221 222
223
with higher bandwidths (BWs), 228 230
and magnetic susceptibility, 228
magnetization transfer in, 228
multi-echo, 224 225 226 227 228
slice coverage in, 223 224 223
three-dimensional (3D), 230 231 230 231
trade-offs of, 223 224
Fat
suppression, 79 80 81
clinical applications of, 290 291
T1 characteristics of, 58 59 65 68 73
T2 characteristics of, 57 59 60 65 69
Fat-saturated blackblood technique, 356
Fat saturation pulse, 228
Feedthrough zipper artifacts, 205 205
Felix Bloch, 18
Ferromagnetic substances
cobalt (Co), 21
iron (Fe), 21
magnetic susceptibility in, 21
in MRI, 21
nickel (Ni), 21
permanent magnetism in, 21
Ferromagnetism, 21 206 207 210
FFT. See Fast Fourier transform (FFT)
FGRET (fast gradient echo train), 361 362 362
FID. See Free induction decay (FID)
Field of view (FOV), 112 167 168 169 167 171 172 173
bandwidth and, 167 168 169 171
gradient strength and, 167 168 169 171
increasing, and prevention of aliasing, 188 188
FIESTA (fast imaging employing steady-state acquisition), 330
First-pass technique, 354
FISP(fast imaging with steady-state precession), 247
characteristics of, 256 256
multiplanar, 252
Turbo, 252 253
FLAIR (fluid-attenuated inversion recovery), 284 286
fast, 284
FLASH. See Fast low-angle shot (FLASH)
Flip angle, 34
partial, 233
partial techniques, 233
Flip(ping), 34
90°, 35 36 37
180°, 37 38
partial, 38 38
physical properties behind, 34
Flow
blood, 295 295
combined, 303
concurrent, 303 304
countercurrent, 303 304
laminar, 294 295
plug, 294 295
second order, 308
turbulent, 294 295 298
vortex, 294 295
Flow compensation techniques, 200
Flow encoding, 311
Flow phenomena, 294 295 296 297 298 299 300 301 302 303 304
305 306 307 308
combined, 303
Flow-related enhancement (FRE), 300 301 302 303
Flow (stream) separation, 294 295
Fluid-attenuated inversion recovery (FLAIR), 182
Formication, 264
Fourier series, 94 95 96
Fourier transform (FT), 11 90 91 96 99 See also Fast Fourier
transform (FFT)
of complex signal, 92 93
of cosine function, 90 91 92 93
of cosine function, 134 134
for cosine wave, 99 100
of data space, 121 121
digital (DFT), 160
equation for, 90
of even function, 94 94
versus Fourier series, 94 95 96 95
imaginary images, 156
of k-space, 171 173 174
of odd function, 94 94
real images, 156
of RF pulses, 130 130
of sinc function, 91 92 92 130 131 131 150 150
of sine wave, 92 93 94 93 94
of spikes, 131 133
of truncated sinc function, 105
FOV. See Field of view (FOV)
Fractional anisotropy, 263
Fractional echo, 155 155 269 270 270
Fractional NEX(number of excitations), 268 269 269
Fractional RF, 270 270
FRE. See Flow-related enhancement (FRE)
Free induction decay (FID), 44 45 46 46 49 49 50 51 52 51 52 74
75 75 83 84 85 86 97 163 359
artifacts caused by, 204 204
measurement of, 235 236 236
Frequency, 6 See also Angular frequency (ω)
angular (ω), 6
linear (f), 6
maximum, 167 168
phase and, relationship between, 174
Frequency domain, 101
Frequency-domain parallel imaging technique. See GRAPPA
Frequency encoding, 108 109 110 111 112 113
Frequency-encoding (Gx) gradient, 108 109 124 124 162 163 163
164 167 167 168
Frequency-encoding (readout) (Gx) gradient coils, 28
Frequency oversampling, 188
Frequency range, 168
3D-fresh blood imaging (FBI), 323 324
FSE. See Fast spin echo (FSE)
FSE imaging. See Fast spin-echo (FSE) imaging
FT. See Fourier transform (FT)
Full bandwidth, 178
Full echo train, 224 225 225
Function, mathematical, 6
Functional imaging, 339 340
G
Gadolinium (Gd), 21 206
Gauss, 23
Gaussian curve, 101 101 104 See also Gaussian RF pulse
Gaussian distribution, 177
Gaussian RF pulse, 103
bandwidth, 104 105
slice-select gradient, 105 106
Generalized autocalibrating partially parallel acquisition (GRAPPA), 279
Geometric distortion artifacts, 105 211 211 212
g-factor, 282 283
Ghost artifact(s), 198 199 198 199 208
Gibbs phenomenon. See Truncation artifacts
Glioblastoma multiforme (GBM), 342
Glx, 345
GMN. See Gradient moment nulling (GMN)
Gradient(s), 49 97 105
bi-lobed, 235 235 236
frequency-encoding (Gx), 108 109 162 163 164 167 168
high performance, 228 229 230
linear, 105 123 124
nonlinear, 105
nonlinearities in, artifacts caused by, 207 211 211
slice-select (Gz), 162 163 163
slice-selective (Gz), 123
slope of, slice thickness and, 98 98
Gradient and spin echo. See GRASE (gradient and spin echo)
Gradient coil(s), 26 27 28 49 98
frequency-encoding (readout) (Gx), 28
phase-encoding (Gy), 28
slice-select (Gz), 28
Gradient echo (GRE) imaging, 38
advantages and disadvantages of, 245
chemical shift artifact of second kind (Dixon effect), 242 243 244
243
fast scanning techniques, 247
DE-prepared, 255
IR-prepared, 254 255 255 255
multiplanar, 253
nomenclature, 247 248
tissue-prepared, 254 255 256
for flow imaging, 256
and magnetic susceptibility, 237 237 238 239
multiplanar variant of, 247 252
PSD diagram for, 239
scan time for, 236
signal-to-noise ratio in, 242
slice excitation and, 236
three-dimensional, 244 245
advantages of, 244 245 245
PSD diagram for, 244 244
scan time for, 244
tissue contrast in, 238 239 240 241 242 239 240
Gradient-echo (GRE) pulse sequence, basic principles of, 233 234 235
236 237 238 239 240 241 242 243 244 245
gradient-echo (GRE) sequences, 356
Gradient moment nulling (GMN), 304
Gradient-recalled acquisition in the steady state (GRASS), 247 248 249
251 252 254
Gradient-recalled echo (GRE), 233 234 235 236 234 235 236 330
333 334
Gradient refocusing mechanism, 359
Gradient-related artifacts, 207 208 209 210 211
Gradient spoilers, 248 250
GRAPPA, 281 282
GRASE (gradient and spin echo), 354
GRASS. See Gradient-recalled acquisition in the steady state (GRASS)
Gray matter
proton density factor for, 62 62
T1 characteristics of, 60 60 62
T2 characteristics of, 61 61 62
GRE. See Gradient echo (GRE)
GRE imaging. See Gradient echo (GRE) imaging
Gyromagnetic ratio, 20 26
Gz. See Z gradient (Gz)
Gz gradient. See Slice-selective (Gz) gradient
H
Half-Fourier acquired single-shot turbo spin-echo (HASTE) sequences,
331
and fast spin echo, 332
Half NEX (1/2 NEX), 153 155 155
Hamiltonian, 201
dipolar, 201
HASTE. See Half-Fourier acquired single-shot turbo spin-echo (HASTE)
sequences
Head coils, 27
Hematoma, cerebral, and tissue contrast, 70 71
Hemorrhage
in brain, and tissue contrast, 70
in endometrioma, and tissue contrast, 73
Hemosiderin, 21 206
tissue contrast and, 72
Hermitian symmetry, 153 154
Hertz (Hz), 6
High performance gradient(s), 228 229 230 257
advantages of, 352 354
requirements for, 352
Hydration layer water, 58 59 58
Hydrogen proton
electromagnetic properties of, 18 19 19 20
natural motional frequency of, 57 58
I
Image
optimization of, scan parameters and, 176 177 178 179 180 181
182
real and imaginary components of, 155 156
Image construction
slice selection for, 97 98 99 100 101 102 103 104 105 106 107
Image contrast, parameters of, 176
Image processing
artifacts caused by, 185 186 187 188 189 190 191 192 193 194
195 196
defined, 123
Imaginary images
FT of, 156
Imaginary number(s), 5
Imaginary unit (i), 5
India Ink Etching, 195
Induced magnetic field
applied and, relationship between, 20
Integrated parallel acquisition technique (iPAT), 279
Interslice gap, 181 181
Intravoxel dephasing, 298
Inversion recovery (IR). See also STIR (short TI inversion recovery)
clinical applications of, 78 79 79
magnitude reconstruction, 79 80 80 81
null point in, 77 78 77
Inversion recovery (IR) pulse, 180 182
Inversion recovery (IR) techniques, 284 285 286 287 288
advantages of, 287
disadvantages of, 287 288
Inversion recovery pulse sequence, 76 77 78 79 76 77
Inversion time (TI), 78 79 78 79 80 182 See also STIR (short TI
inversion recovery)
IR. See Inversion recovery (IR)
Iron (Fe), ferromagnetism in, 21
J
“J-coupling,” 342
Jerk compensation, 308
K
Keyhole imaging, 226
Key points
of artifacts, 213
of data space, 161
of fast spin-echo, 231
of field of view (FOV), 169
of Fourier transform, 96
of gradient-echo (GRE), 246
of image construction, 106
of k-space, 174
of mathematical concepts, 13 14
of physical principles of magnetic resonance imaging, 29 30
of pulse sequence diagram (PSD), 166
of pulse sequences (saturation, partial saturation and inversion
recovery), 81
of radio frequency (RF) pulse, 38
of scan parameters and image optimization, 183
of signal processing, 143
of spatial encoding, 121
of spin echo (SE) pulse sequence, 88 89
of T1, T2, T2*, 46
of tissue contrast, 65
of TR, TE, and tissue contrast, 55
k-space, 117 118 119 126 127 145 146 145 171 172 173 174
analog version of. See Data space
center of, 147 148 147 148
conjugate (Hermitian) symmetry, 153 154
in conventional spin echo (CSE), 217 218
digitized, 140 141 141
edges of, 149 150 149 150
example of, 157 158 159 160 157 158 159
Fourier transform (FT) of, 171 173 174
and image, 171 172 173 172
image construction, 150 151 151
images of, 148 149 149
properties of, 147 148 149 150 151 152 153
real and imaginary, FT of, 155 156
symmetry, 151 152 153 152 153
three-dimensional line drawing of, 151 152
unit interval in, 171 172 173 174 172
k-space averaging, 227
L
Laminar flow, 294 295
Lanthanide group, 21
Larmor equation, 26 168 174
derivation of, 36
linear frequency f in, 26
Larmor frequency (ω0), 36 49 57 98 290 See also Angular frequency
(ω)
magnetic field strength and, 98 99 98
Linear frequency (f), 6 17
in Larmor equation, 26
Linear gradient(s), 123 124
Lipids
and radiation necrosis, 348
Logarithms, 12
properties of, 13
Longitudinal magnetization (Mz), 36 37 38 233 234 235
after time interval TR, 52 52
recovery of, 41 41
RF pulse and, 40 41 40 41 48 48 52
saturation recovery and, 74
T2 relaxation time and, 41 41
Longitudinal relaxation time, 40 See also T1 relaxation time
Low bandwidth, 276
Lower frequency
cosine wave of, 136
Low-pass filters (LPF), 134 135 136 134 136 137 272
LPF. See Low-pass filters (LPF)
M
Magic angle artifacts, 200 201 202
Magnet(s)
design of, 23
field strength of, 23
permanent, 23
resistive, 23
superconducting, 23
types of, 23
Magnetic dipole moment (MDM), 20 20 23 24 24
in Larmor equation, 36
Magnetic field (B), 25 33
of electromagnetic waves, 17 17 18
external, 21 23 24 31 31 32 33
inhomogeneity, and dephasing, 42 43 83
Magnetic field intensity (H), 20
Magnetic field strength
Larmor frequency and, 98 99 98
Magnetic flux density (B), 20
Magnetic induction field (B), 20
Magnetic permeability (µ), 20
Magnetic resonance angiography (MRA)
black-blood, 323
contrast-enhanced, 315 354
advantages, 316
disadvantages, 316
phase contrast (PC), 311 312 313
advantages of, 313
disadvantages of, 314
2D- versus 3D-, 314
time of flight (TOF), 310
Magnetic resonance imaging (MRI)
definition of, 16
perfomance of, 21 22 23
physical principles of, 16 17 18 19 20 21 22 23 24 25 26 27 28
29
Magnetic susceptibility artifact(s), 206 207 208 209 210
Magnetic susceptibility (χ), 20 206 207 209 228
in diamagnetic substances, 21
in ferromagnetic substances, 21
and gradient-echo (GRE), 237 237 238 239
in paramagnetic substances, 21
Magnetism
diamagnetism, 21
ferromagnetism, 21
paramagnetism, 21
Magnetizability, measure of, 20
Magnetization. See also Net magnetization
longitudinal (Mz), 36 37 38 40 40 41 41 52 52 74 233 234 235
recovery of, 25
transverse (Mxy), 40 41 40 41 41 45 48 48
residual, in gradient echo (GRE) imaging, 238 239
Magnetization (cont.)
steady-state (Mss), in gradient echo (GRE) imaging, 238 239
Magnetization transfer (MT), 290 291 292
effect in FSE, 290
in fast spin-echo (FSE) imaging, 228
saturation, 322
Magnetization vector, 31 32 33 34 35 36 37 38 34 35 41 44 48 50
51 74 76 77 78 83 84 255
Magnitude
angle and, 5 6
of signals, 125 146 151 153
of the sine wave, 11
of vector, 6
Magnitude image (modulus), 6 155 156 157 156
Magnitude reconstruction, inversion recovery, 79 80 80 81
Mathematical concepts, of MRI physics, 3 4 5 6 7 8 9 10 11 12 13
Maximum frequency, 167 168
Maximum intensity projection (MIP), 319
disadvantages of, 319
Maxwell's wave theory, electromagnetic waves and, 16 17
MDM. See Magnetic dipole moment (MDM)
Meningioma
fast spin-echo (FSE) imaging of, 229
tissue contrast in, 67
Methemoglobin
in tissue contrast, 70 71
Microwaves, characteristics of, 16
Minimum intensity projection algorithm, 323
Mipping, 319
Mobile protons, magnetization and, 25
Modulation, of frequency, 105 106
Modulus. See Magnitude image (modulus)
Moiré fringe, 206 208
Motion
artifacts, 147 197 198 199 200 201 198 199 200 201
blood, 331
cardiac, 329 330 331
in cardiac imaging and solutions, 331
periodic, 197 198 199
random, 197 199 200 199 200
respiratory, 327 328 329
Motion-correction, 329
MOTSA, 321
MP-RAGE, 254
MRI. See Magnetic resonance imaging (MRI)
MRI pulse sequence, 356
MR signal
radio frequency (RF) and, 22
MT. See Magnetization transfer (MT)
Mucinous fluids, 59
Multi-echo fast spin-echo (FSE) imaging, 224 225 226 227 228
Multiple overlapping thin-slab acquisition. See MOTSA
Multislice acquisition, 127 127
Multivoxel techniques, 344 345
Myocardial tagging, 337
Myoinosotol, 345
N
Narrower frequency bandwidth
slice thickness and, 101 102
Natural logarithm, 12
Natural motional frequencies, of protons, 57
Navigator-echo gating, 328
Necrotic lymphoma, 349
Net magnetization, 24 24 25 25 31 32 34
Net magnetization vector (M0), 31 32 35
spiral motion of, 33
in x-y plane (Mxy), 35 35 37 38 38
Newtonian physics, 16
NEX. See Number of excitations (NEX)
NFW. See No frequency wrap (NFW)
N/2 ghost artifact(s), 261
Nickel (Ni), 21
NMR. See Nuclear magnetic resonance (NMR)
No frequency wrap (NFW), 188 271 272 272
No phase wrap (NPW), 188 188
NPW. See No phase wrap (NPW)
NSA. See Number of signal averages (NSA)
Nuclear magnetic moment, 21
Nuclear magnetic resonance (NMR), 16
Nuclear paramagnetism, 21
Nucleus
energy states of, 19
even number of protons in, 19
MDM in, 20
odd number of protons in, 19
paired proton in, 19
Null Point, in inversion recovery, 77 78 77
Number of excitations (NEX), 141 165 177
½, 153 155 155
¼, 155 155
Number of phase-encoding steps (Ny ), 177
Number of signal averages (NSA), 165
Nutation (spiral motion), 33 35
Ny (number of phase-encoding steps), 177
Nyquist frequency, 178 271 354
Nyquist law. See Nyquist sampling theorem
Nyquist sampling theorem, 123 139 140 139 140
O
Oblique planes, 29
Odd-echo dephasing, 298
Odd function, 153
Odd function, Fourier transform (FT) of, 94 94
Osteomyelitis, tissue contrast in, 68
Out-of-phase scanning, 195
Overhead time (To), 125
Oversampling, 140 188
Overscanning, 156 188
Oxyhemoglobin, tissue contrast in, 70
P
Paired protons, 19
Parallel imaging
advantages, 281
autocalibration, 280 281
concept, 279
disadvantages, 281
prescan, 280
Paramagnetic substances
deoxyhemoglobin and, 21
dysprosium (Dy), 21
gadolinium (Gd), 21
hemosiderin, 21
increased magnetic field in, 21
magnetic susceptibility in, 21
methemoglobin and, 21
in MRI, 21
same direction, magnetic field in, 21
Paramagnetism, 21 206 209
nuclear, 21
Partial flip, 38 38
Partial flip angle techniques, 233
Partial saturation pulse sequence, 74 75 75
Partial volume artifacts, 196 197
Pelvis
fast spin-echo (FSE) imaging of, 229
Perfusion imaging, 337 339
in echo planar imaging (EPI) technique, 263 264 265
Periodically rotated overlapping parallel lines with enhanced
reconstruction (PROPELLER), 276 277
Periodic motion, 197 198 199
Period (T), 6 6
Permanent magnets, 23
Phase, 7 8 7
frequency and, relationship between, 174
Phase construction, 79
Phase contrast imaging, 330
Phase contrast (PC) MRA, 311 312 313
advantages of, 313
disadvantages of, 314
2D- versus 3D-, 314
Phased-array coils, 28
Phase difference, 4 163 164
Phase encoding, 113 114 115 116 117 118 119 120 114 115 116
117 118 123 124 125 126
defined, 125
motion artifact and, 198 199
symbol for, 123 125
TR (repetition time) and, 117 118 119 118 119
Phase-encoding (Gy) gradient, 113 114 115 116 117 118 119 113
114 115
coils, 28
Phase image, 155 156 157 156
Phase offset, 7 7
Phase-offset RF pulses, 272 273 273
Phase oversampling, 188
Phase-sensitive detection, 151
Phase shift, 7 8 115 116 117 158 158
Photography, 21 22
Pixel, 108
bandwidth of, 189 190
Pixel bin, 190 194
Pixel size
and chemical shift artifact, 190 191 194
Planck's constant (h), 36
Planck's law, 36
Plane of imaging, 28 29 29
oblique, 29
x, y, and z axes for, 28 28 29
Plug flow, 294 295
Postgadolinium imaging, 331 354
Precession, 25 26 26
frequency of, 33
out of phase, 34
in phase, 34 36
RF pulse and, 31 32 33
Prep pulse, 356 359
Prescan, 38
PRESS (point-resolved spectroscopy) technique, 348
Propagation factor (C), of electromagnetic waves, 17 17
Proteinaceous material
T1 characteristics, 58 59 60 61 62 63 64 59 65
T2 characteristics, 57 59 60 65
Proteinaceous solutions, 58 59 60 61 62 63 64
Proton density, magnetization and, 25
Proton density factor: N(H), 61 62 62
Proton density-weighted image, 76 87 88 88
Proton magnetic resonance spectroscopy (MRS), 342
nuclear magnetic resonance (NMR) and, 342
Proton(s)
energy states of, 19 19 20 36
ensemble of, FID and, 49
even number of, 19
mobile, 25 51
natural motional frequencies of, 57
odd number of, 19
imaging in MR, used for, 20
paired, 19 20
precession of, 25 26 26 31 32 33 34
out of phase, 34
in phase, 34 36 37 40 40 41 43
unpaired, 20 20
wobble of, 25 26 26
Proton spins, direction of, 19
PR (projection reconstruction) acquisition, 316
PSD. See Pulse sequence diagram (PSD)
PSIF. See Steady-state free precession (SSFP)
Pulse sequence diagram (PSD), 74 162 163 164 165 166 310
for 3D gradient-echo (GRE) imaging, 244 244
in echo plannar imaging (EPI), 259 260
for SE pulse sequence, 108 124
spin-echo (SE), 162 163 164 162 163 164 165
Pulse sequence(s), 74 75 76 77 78 79 80 81 83 84 85 86 87 88
See also Radio frequency (RF) pulses
definition of, 74
inversion recovery, 76 77 78 79 76 77
partial saturation, 74 75 75
saturation recovery, 75 76 76
spin-echo (SE), 83 84 85 86 87 88
Pythagorean theorem, 6
Q
Qp/Qs ratios (pulmonary to systemic flow ratio), 340
Quadrature coils, 28
Quantum mechanics, 16 19
Questions
on artifacts, 213 214
on data space, 161
on fast spin-echo, 232
on field of view (FOV), 169 170
on Fourier transform, 96
on gradient-echo (GRE), 246
on image construction, 107
on k-space, 175
on mathematical concepts, 14 15
on physical principles of magnetic resonance imaging, 30
on pulse sequence diagram (PSD), 166
on pulse sequences (saturation, partial saturation and inversion
recovery), 81 82
on radio frequency (RF) pulse, 39
on scan parameters and image optimization, 183 184
on signal processing, 144
on spatial encoding, 121 122
on spin echo (SE) pulse sequence, 89
on T1, T2, and T2*, 47
on tissue contrast, 73
on TR, TE, and tissue contrast, 55 56
R
Radiation necrosis
and lipids, 348
Radio frequency (RF), 11 18
Radio frequency (RF) pulses, 11 12 12 18 22 22 23 31 32 33 34 97
257 359 See also Flipping, partial
90°, 35 36 37 37 48 48 49 74 74 75
180°, 356
180°, 37 38 76 83
refocusing (rephasing), 83 85 87
auto, 38
center frequency of, 129 130 130
definition of, 31
frequency of, 33 34
FT of, 130 130
Gaussian. See Gaussian RF pulse
magnetic field generated by (B1), 32 33
magnetic resonance signal, 22 23
nonselective, 103
selective, 103
sinc, 103 103
Radio waves, characteristics of, 16
Random motion, 197 199 200 199 200
Rapid acquisition through parallel imaging design (RAPID), 279
Rapid acquisition with relaxation enhancement (RARE), 217 See also
Fast spin-echo (FSE) imaging
RARE. See Rapid acquisition with relaxation enhancement (RARE)
Rare earth elements, unpaired electrons in, 21
Readout, 359 360 361 362
Readout gradient. See Frequency-encoding (Gx) gradient
Real images, 155
FT of, 156
Received signal (S), 43 44 49 49 51 51
Receiver bandwidth, 140
Recovery curves. See Decay/recovery curves, TR (repetition time) and
Refocusing mechanism, 83 85 87 162 163 165 359
Relative anisotropy, 263
Relaxation, definition of, 40
Relaxation rate, 46
Relaxation time. See also T1 relaxation time; T2 relaxation time; T2*
relaxation time
longitudinal, 40
spin-lattice, 40
spin-spin, 42
transverse, 42
Repeated excitation radio frequency (RF) pulses, 320
Repetition time. See TR (repetition time)
Rephasing, 83 85 87 163
even-echo, 298 299 300
Resistive magnets, 23
Resolution
parameters of, 176
and signal-to-noise ratio (SNR), 179 180
Resonance, 33 34
Respiratory compensation (ROPE and COPE), 328 329
Respiratory motion, 327 328 329
Rewinder gradient, 230
RF. See Radio frequency (RF)
RF coil, 26
RF pulses. See Radio frequency (RF) pulses
RF (radio frequency) noise, 205 206
Right-hand rule, 44 44
Ring-down effect, 94 95
Ripples effect, 104
Root mean square (rms), 79
Root pulsing sequence, 260
Rotating frame of reference, 34 35 36 37
S
Sampling, 120 121 135 136 137 138 139 135 136 137 138 139
of composite signals, 141 142 143 142
Sampling interval (ΔTs), 171
Sampling time (Ts), 124 140 163
versus sampling interval, 139 140 140
Sat Pulses. See Spatial saturation pulses (Sat Pulses)
Saturated, 300
Saturation, 74 75
partial, 74 75
and prevention of aliasing, 188
Saturation effects, 320 321
Saturation recovery pulse sequence, 75 76 76
Scanners, recent advances in, 268 269 270 271 272 273 274 275
276 277
Scanning, recent advances in, 268 269 270 271 272 273 274 275
276 277
Scan parameters
primary, 176
secondary, 176
Scan time. See also Acquisition time
for conventional spin echo (CSE), 217
for gradient-echo (GRE), 236
SE. See Spin-echo (SE)
SE-EPI, 261 262 263
SE imaging. See Spin-echo (SE) imaging
Sensitivity encoding (SENSE), 279 281 283
SE pulse sequence. See Spin-echo (SE) pulse sequence
Sequential scanning, 236
SEW. See Slice-excitation wave (SEW)
Shared echo approach, 225 226 227 228 226 227
Shim coils, 23 28
Shimming coils (auto shimming), 206
Short TI inversion recovery (STIR), 81 81 182
Shot, 257
Signal, 6 7
definition of, 44
measurement of, 49
oscillating, 45
periodic, 6 7
received, 43 44 45 51 51
sinusoidal, 46 46 49
Signal intensity, 53 54
Signal loss, high velocity, 296 297 297
Signal processing, 123 124 125 126 127 128 129 130 131 132 133
134 135 136 137 138 139 140 141 142 143 144
defined, 123
Signals
amplitude of, 109 110 151
composite, sampling of, 141 142 143 142
digitized sample, 146
direction of, 153
magnitude of, 125 146 151 153
Signal-to-noise ratio (SNR), 27 28 141 148 148 176 177 279 282 283
287
in 3D imaging, 178 179 180
parameters of, 176
resolution and, 179 180
Similarity transform, 263 263
Simultaneous acquisition of spatial harmonics (SMASH), 279
sinc function, 11 11 12
Fourier transform (FT) of, 92 93 94 93 94
FT of, 130 131 131 150 150
Sinc RF pulse, 103 103
sine data
data space with, 152 153 153
sine function, 94 153 154 155
Single-shot technique, 331 334
Single voxel techniques, 342 343 344
Sinusoidals, 3 4 4 5
exponentially decaying, 11 11 49 49
Slew rate (SR), 257 352
Slice excitation, gradient-echo (GRE) and, 236
Slice-excitation wave (SEW), 303
Slice(s)
contiguous, 128
gaps between, 128
interleaved, 128
number of (coverage), 128 129 223 224
sequential, 128
Slice-select gradient coils (Gz), 28
Slice-select gradient (Gz), 103 105 106 162 163 163
Slice selection, 128
Slice-selective (Gz) gradient, 123
Slice thickness, 100 100
cross-talk, 101 101
narrower frequency bandwidth and, 101 102
selection of, 97 97
slice-select gradient, 103
slope of the gradient and, 98 98
ways to change, 101 102 103
Slice-tracking, 329
Small bowel, tissue contrast in, 73
SNR. See Signal-to-noise ratio (SNR)
Solenoid coils, 28
Solids
T1 characteristics of, 58 59 65
T2 characteristics of, 57 59 60 65
Solid tumors, 349
SPAMM. See SPAtial Modulation of the Magnetization
Spatial encoding, 22 23 28 49 97
Spatial frequencies, 145
Spatial frequency domain, 120 173
Spatial frequency(ies), 173 174
SPAtial Modulation of the Magnetization, 337
Spatial presaturation, 289 290 290
Spatial saturation pulses (Sat Pulses), 274 275
Spectral density of noise, 177
Speed of light, of electromagnetic waves, 16
Spikes
FT of, 131 133
sequence of, 131 133
Spin(s), 18 19 20
interactions between, 42 43
out of phase, 34 42 43 83 84
out phase, 125 126
in phase, 34 41 43 83 84 125 126
Spin density, magnetization and, 25
Spin dephasing, 43 44
Spine, lumbosacral, tissue contrast in, 68 69
Spin-echo (SE) imaging, 83 See also Conventional spin echo (CSE); Fast
spin-echo (FSE) imaging
aliasing (wraparound) in, 185 186 186
contrast on, 182 182
Spin-echo (SE) pulse diagram, 83 84 85 86 87 84 85 86
Spin-echo (SE) pulse sequence, 83 84 85 86 87 88 146
asymmetric echoes in, 87
diagram of, 146
pulse sequence diagram (PSD) for, 108 124 162 163 164 162 163
164 165
symmetric echoes in, 87 87
Spin-lattice relaxation time. See T1 relaxation time
Spinning charged particle, 18 19 20 18 20 20
Spin quantum number (S), 19 20
Spin-spin interaction(s), 42 43 46 83
“spin-spin splitting,” 342
Spiral motion, of net magnetization vector (M0), 33
Split echo train, 225 225
Spoiled GRASS images (SPGR), 247 248 249 250 251
characteristics of, 256 256
disadvantages of, 251
Fast, 252 253
multiplanar (MPSPGR), 252
tissue contrast, 248
Spoiling
defined, 247 248
RF (radio frequency), 248 249 250
Steady-state free precession (SSFP), 247 251 252
advantages of, 251
characteristics, 256 256
disadvantages of, 252
tissue contrast, 251
STIR. See Short TI inversion recovery (STIR)
STIR (Short TI Inversion Recovery), 284 285
Subtraction technique, 320
Superconducting magnets, 23
Superparamagnetic substances, 21
Superparamagnetism, 206
Suppression. See also Tissue suppression
Surface coil(s), 27 187 188
Susceptibility artifact(s)
in EPI, 261 262
T
Tangent, 3 6 6 9
T1 characteristics
of cerebrospinal fluid (CSF), 60 60 62 66 68
of fat, 58 59 65 68 69
of gray matter, 60 60 62
of proteinaceous material, 65
of proteinaceous solutions, 58 59 60 61 62 63 64 59 73
of solids, 58 59 65
of various tissues, 65
of water, 58 59
of white matter, 60 60 62 66
T1 weighting, 75 87 88 88
T2 characteristics
of cerebrospinal fluid (CSF), 60 61 62 66 68
of fat, 57 59 60 62
of gray matter, 61 61 62
of proteinaceous material, 57 59 60 65
of proteinaceous solution, 65 73
of solids, 57 59 60 65
of various tissues, 65
of water, 57 59 60
of white matter, 61 61 62 66
T2 weighting, 87 88 88
TE (echo delay time or time to echo), 51 52 53 52 85 86
in partial saturation, 74 75 75
saturation recovery, 75 76 76
and T1 effect, 60
and T2 effect, 60
and T2* effect, 54 54
TE (echo delay time or time to echo) (cont.)
and tissue contrast in, 63 64 63 64
lengthening of
cause of, 181 182 182
effects of, 181
reduction of
effects of, 181
method for, 182
TEeff. See Effective TE (TEeff)
Tesla, 23
Three-dimensional (3D) acquisition, 273 274
Three-dimensional (3D) fast spin-echo (FSE) imaging, 230 231 230 231
Three-dimensional (3D) imaging
aliasing (wraparound) in, 186 187 188 187 188
signal-to-noise ratio (SNR) and, 178 179 180
Three-dimensional gradient-echo (GRE) imaging, 244 245
advantages of, 244 245 245
PSD diagram for, 244 244
scan time for, 244
TI. See Inversion time (TI)
Tilted optimized non-saturating excitation (TONE). See TONE (tilted
optimized nonsaturating excitation)
Time constant(s) (τ), 9 10
Time domain, 120
data space in, 120
Time-of-flight loss, 302 332
Time of flight (TOF) MR angiography, 256 290
Time-resolved imaging of contrast kinetics (TRICKS), 316
Time to echo. See TE (echo delay time)
Timing diagram. See Pulse sequence diagram (PSD)
Tissue contrast, 53 54 53
clinical application of, 57 58 59 60 61 62 63 64
in gradient-echo (GRE), 238 239 240 241 242 239 240
in spin-echo (SE) pulse imaging, 87 88 88 88
T1, 53 54 53 57 58 59 60 61 62 63
T2, 53 54 54
T2*, 54 54
Tissue preparation, 254
Tissue suppression. See also Fat, suppression of techniques for, 284
285 286 287 288 289 290 291 292
TI (time fo inversion). See Inversion time (TI)
To. See Overhead time (To)
TOF MRA, 310
TONE (tilted optimized nonsaturating excitation), 321 322 323 323
Toxoplasmosis, 349
Trade-off, 203
Transmission bandwidth, 140
Transmit bandwidth, 129 130
Transmit/receive coil(s), 27 44 45
Transverse magnetization (Mxy), 35 35 37 38 38
Transverse (Mxy) magnetization
with decay rate of T2, 41 41
received signal and, 45
residual, in gradient echo (GRE) imaging, 238 239
RF pulse and, 40 41 41 48 48
spiral-like decay of, 45
steady-state (Mss), in gradient echo (GRE) imaging, 238 239
T1 relaxation time and, 40 41 40 41
T2 relaxation time and, 41 41
T1 relaxation time, 24 25 40 41 41 43
T2 relaxation time, 41 42 43 41 43
T2* relaxation time, 46
Trigger delay, 330
Triple inversion recovery (IR), 356
“triplet,” 342
TR (repetition time), 50 51 50 62 128
decreased
effects of, 180 181
increased
effects of, 180
long, for multiplanar, 252
maximum number of slices, 129
in partial saturation, 74 75 75
and phase encoding, 117 118 119 118 119
saturation recovery, 75 76 76
and T1 effect, 60
and T2 effect, 60
and tissue contrast in brain, 63 64 63 64
True fast imaging with steady-state precession (True FISP), 334
“True” interleaving technique, 203
Truncation artifacts, 189 195 196 196 197
Ts. See Sampling time (Ts)
TSE. See Turbo spin echo (TSE)
Turbo spin echo (TSE), 217 217 See also Fast spin-echo (FSE) imaging
Turbulent flow, 294 295 298
Two-dimensional fourier transform (2DFT), 113
U
Uncooperative patient, 357
Undersampling, 188
and aliasing, 138 139
Unpaired electron
hemosiderin, at the end stage of hemorrhage, 21
rare earth elements and, 21
Unpaired protons, 20
Unsaturated, 300
V
Variance (σ2) of Gaussian distribution, 177
Vector(s), 4 5 5 12 12
Velocity encoding. See VENC
VENC, 312 313 340
Venetian blind, 321
Vertebral bodies
and chemical shift artifact, 191 191 192 193
View sharing, 337
Views per segment (VPS), 331
Visible light, characteristics of, 16
Vortex flow, 294 295
Voxel, 108
dimensions of, 177 177
size of, and signal-to-noise ratio (SNR), 177 178
Voxel volume, and signal-to-noise ratio (SNR), 177 178 178
W
Water
bound to hydrophilic macromolecule, 58
bulk phase, 58
clinical applications of suppression, 291 292
diamagnetism of, 21
hydration layer water, 58 58
hydrogen in, 20
T1 characteristics of, 58 59 65
T2 characteristics of, 58 59 65
Waveform(s), 6
narrow, 99 99
wide, 99 99
White matter
T1 characteristics of, 60 60 62 66
T2 characteristics of, 61 61 62 66
Wobble, 26
Wraparound. See also Aliasing
definition of, 186
X
X-ray, characteristics of, 16
X-ray imaging, 22 22
Z
z gradient (Gz), 103
Zipper artifact(s), 203 204 205 204
feedthrough, 205 205