26.11.2012 Views

Studio delle prestazioni di un sistema a fosfori per mammografia ...

Studio delle prestazioni di un sistema a fosfori per mammografia ...

Studio delle prestazioni di un sistema a fosfori per mammografia ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

UNIVERSITÀ DEGLI STUDI DI FIRENZE<br />

FACOLTA’ DI MEDICINA E CHIRURGIA<br />

SCUOLA DI SPECIALIZZAZIONE IN FISICA SANITARIA<br />

<strong>Stu<strong>di</strong>o</strong> <strong>delle</strong> <strong>prestazioni</strong> <strong>di</strong> <strong>un</strong> <strong>sistema</strong><br />

a <strong>fosfori</strong> <strong>per</strong> <strong>mammografia</strong> <strong>di</strong>gitale<br />

Tesi <strong>di</strong> Specializzazione in Fisica Sanitaria del<br />

Dr. Simone Busoni<br />

Relatore Dott. Giacomo Belli<br />

Direttore della scuola Prof. Salvatore Romano<br />

Firenze, 29 Ottobre 2002 Anno Accademico 2001/2002


Contents<br />

Introduction 1<br />

1 Computed Ra<strong>di</strong>ography Principles 3<br />

1.1 Projection ra<strong>di</strong>ography principles . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

1.2 The imaging plate IP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

1.2.1 The single and dual-side rea<strong>di</strong>ng . . . . . . . . . . . . . . . . . . . . . 11<br />

1.3 The <strong>di</strong>gitization process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

2 Characterization of a <strong>di</strong>gital ra<strong>di</strong>ography system 15<br />

2.1 Linear systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

2.2 The Modulation Transfer F<strong>un</strong>ction MTF . . . . . . . . . . . . . . . . . . . . . 18<br />

2.2.1 Digital MTF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.3 NPS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

2.4 DQE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

2.4.1 DQE o<strong>per</strong>ating definition. . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

2.5 NEQ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

3 Ex<strong>per</strong>imental setup 31<br />

3.1 The mammography <strong>un</strong>it . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

3.1.1 Dose vs. mAs Relation . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

3.1.2 The Half Value Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

3.2 Determination of the X-ray spectrum . . . . . . . . . . . . . . . . . . . . . . . 35<br />

3.3 The FUJI Computed Ra<strong>di</strong>ography workstation FCR 5000 MA . . . . . . . . . 38<br />

3.4 IP cassette and photostimulable storage plate . . . . . . . . . . . . . . . . . . 40<br />

3.5 Image readout process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

i


3.6 Histogram analysis and IP sensitometric curve . . . . . . . . . . . . . . . . . . 43<br />

4 Measurement of the physical quantities contributing to the DQE 46<br />

4.1 Ex<strong>per</strong>imental characteristic curve . . . . . . . . . . . . . . . . . . . . . . . . . 47<br />

4.2 Pre-sampling MTF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

4.3 EMTF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

4.4 NPS measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

4.5 NEQ measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

4.6 DQE measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

4.6.1 Determination of q . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

4.6.2 Measured DQE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

Conclusions 72<br />

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

Ringraziamenti 77<br />

ii


Introduction<br />

Digital X-ray imaging devices are increasingly used in me<strong>di</strong>cal <strong>di</strong>agnosis and will widely re-<br />

place conventional (analogic) imaging systems in the future. The benefits of acquisition of<br />

ra<strong>di</strong>ological images in <strong>di</strong>gital form became quickly obvious following the introduction of com-<br />

puted tomography by Ho<strong>un</strong>sfield in 1973. These advantages include an increased flexibility in<br />

recor<strong>di</strong>ng images and <strong>di</strong>splay characteristics as well as the possibility of retrieving and trans-<br />

mitting data through comm<strong>un</strong>ications networks (PACS). It was only with the development of<br />

improved detector technologies, more powerful computers, high-resolution <strong>di</strong>gital <strong>di</strong>splays and<br />

commercial and affordable laser devices, that <strong>di</strong>gital ra<strong>di</strong>ography obtained a big <strong>di</strong>ffusion in<br />

other fields, like the standard ra<strong>di</strong>ographic projection imaging. Initially, it was thought that<br />

<strong>di</strong>gital ra<strong>di</strong>ography would have to match the very deman<strong>di</strong>ng limiting spatial resolution <strong>per</strong>for-<br />

mance of film-based imaging. However, film imaging is often limited by a lack of exposure<br />

latitude due to the film characteristic curve and by the noise associated with film granularity<br />

and inefficient use of the incident ra<strong>di</strong>ation. Further ex<strong>per</strong>ience has suggested that a high value<br />

of limiting spatial resolution is not as important as the ability to provide excellent image con-<br />

trast over a wide latitude of x-ray exposures for all spatial frequencies up to a more modest<br />

limiting resolution [1]. A <strong>di</strong>gital ra<strong>di</strong>ographic system can provide such <strong>per</strong>formances as well as<br />

allowing the implementation of computer image processing techniques [2] for the extraction of<br />

useful me<strong>di</strong>cal quantitative information.<br />

Two main <strong>di</strong>gital image acquisition systems have been developed for ra<strong>di</strong>ography clinical<br />

applications: the Computed Ra<strong>di</strong>ography (CR), currently more <strong>di</strong>ffused, and the Direct Digital<br />

Ra<strong>di</strong>ography. The first technology employs an image storage system that only in a second step<br />

is read and <strong>di</strong>gitized, the latter uses a solid state detector which converts imme<strong>di</strong>ately the image<br />

information in <strong>di</strong>gital form.<br />

The Computed Ra<strong>di</strong>ography (CR) system, which is an X-ray imaging system employing<br />

photostimulable phosphor, is the first <strong>di</strong>gital ra<strong>di</strong>ography system of practical use. In the mid<br />

1980s, the CR system was released to the market, and since then great improvements have been<br />

1


made, allowing the <strong>di</strong>ffusion of these technologies in highly specialized fields, as in mammo-<br />

graphical screening where particular <strong>per</strong>formances are required in terms of spatial resolution.<br />

One of the challenges with <strong>di</strong>gital projection ra<strong>di</strong>ography, and in mammography field in partic-<br />

ular, is to improve the detection of abnormalities. This last goal can be obtained acting on two<br />

<strong>di</strong>fferent stages: 1) enhancing the image quality of the acquisition, for instance improving the<br />

S/N ratio and dynamic range, 2) optimizing the <strong>di</strong>splay parameters through image processing<br />

so that as much <strong>di</strong>agnostic information as possible can be extracted from an image read by a<br />

ra<strong>di</strong>ologist.<br />

The work and results described in this thesis have the main goal of fully characterize a<br />

mammographical imaging storage system from a physical point of view, leaving to a further<br />

study the optimization of the <strong>di</strong>splay parameters of clinical images from a rea<strong>di</strong>ng ra<strong>di</strong>ologist<br />

point of view. The definition of the parameters that describe the specific imaging pro<strong>per</strong>ties of<br />

these <strong>di</strong>gital X-ray imaging devices is therefore necessary together with a standar<strong>di</strong>zation of the<br />

measurement procedures employed. To this end a set of measurements have been <strong>per</strong>formed on<br />

the X-ray equipment and on the mammographical station (Fuji FCR 5000 MA). A set of soft-<br />

ware routines was written to analyze the data and to calculate all the main physical parameters<br />

involved in imaging <strong>per</strong>formances of the system and to compare them with result previously<br />

obtained and published in literature.<br />

In Chapter 1 the working principles of a CR system are described, together with a small<br />

summary of the theory of ra<strong>di</strong>ographic imaging. In Chapter 2 the main parameters describ-<br />

ing the <strong>per</strong>formances of a mammographic system are defined and the theory of linear imaging<br />

systems is briefly summarized. The hardware and ex<strong>per</strong>imental setup used during this work,<br />

their characterization and the X-ray spectrum determination are described in Chapter 3. Finally,<br />

the measured <strong>per</strong>formances of the CR system in terms of the physical parameters Modulation<br />

Transfer F<strong>un</strong>ction MTF, Noise Equivalent Quanta NEQ, Noise Power Spectrum NPS and De-<br />

tective Quantum Efficiency DQE are reported in Chapter 4, together with the description of the<br />

procedures and software codes used to evaluate them.<br />

2


Chapter 1<br />

Computed Ra<strong>di</strong>ography Principles<br />

A Computed Ra<strong>di</strong>ography (CR) system is based on a detector (imaging plate) which allows a<br />

delayed photostimulable light emission after an X-ray exposure.<br />

The image formation process in a CR system can be thought as the sum of several steps in-<br />

volving both the image information storage in a detecting device and the subsequent acquisition<br />

and storage of the image as <strong>di</strong>gital data. The block <strong>di</strong>agram of the image formation process is<br />

shown in Fig. 1.1 together with the noise sources dominating each step.<br />

The main steps involved in the PSP image acquisition process are summarized in a pictorial<br />

view in Fig. 1.2.<br />

The image information is retrieved in a de<strong>di</strong>cated system by raster scanning the plate with<br />

a laser to stimulate the luminescence. The light emitted, which is proportional to the absorbed<br />

dose, is then collected by a photomultiplier tube (PMT), <strong>di</strong>gitized and finally stored on a work-<br />

station for subsequent analysis. One of the main advantages of this technology, without forget-<br />

ting those common to all <strong>di</strong>gital ra<strong>di</strong>ography systems, is the possibility to expose the imaging<br />

plate (IP) in the same projection-ra<strong>di</strong>ography facility usually used with conventional screen-film<br />

based cassettes, since the detector holder has the same geometrical characteristics.<br />

In this chapter the main physical processes and CR principles are summarized. At first,<br />

projection ra<strong>di</strong>ography principles and the mechanism of ra<strong>di</strong>ation interaction with matters will<br />

be briefly reviewed, together with the process of image storage in a photo-stimulable phosphor<br />

detector. Finally, the rea<strong>di</strong>ng and <strong>di</strong>gitization steps will be described, with particular emphasis<br />

to the elements and characteristics peculiar to the system <strong>un</strong>der investigation. From now on, if<br />

3


Figure 1.1: Block <strong>di</strong>agram of the image formation process and main noise sources<br />

Figure 1.2: PSP image acquisition and processing.<br />

not otherwise stated, a set of procedures and a <strong>di</strong>agnostic mammographical setup is considered.<br />

1.1 Projection ra<strong>di</strong>ography principles<br />

The ra<strong>di</strong>ographic image is formed by the interaction of X-ray photons with a photon detector<br />

and is therefore a <strong>di</strong>stribution of those photons which are transmitted through the patient and<br />

recorded by the detector. These photons can either be primary photons, which have passed<br />

4


the patient without interacting, or secondary photons, which results from an interaction. The<br />

secondary photons will in general be deflected from their original <strong>di</strong>rection and carry little<br />

useful information. The primary photons, on the other side, give a measure of the probability<br />

that a photon will not interact, and this probability will itself depend upon the sum of the X-ray<br />

attenuating pro<strong>per</strong>ties of all the tissues the photon traverses. The image is therefore a projection<br />

of the tissue attenuating pro<strong>per</strong>ties. The scattered photons create a backgro<strong>un</strong>d signal which<br />

degrades contrast. In most cases the majority of the scattered photons can be removed by<br />

placing an anti-scatter device between the patient and the image receptor. This device can be a<br />

grid formed of a series of lead strips that are parallel to the X-ray <strong>di</strong>rection.<br />

A simple mathematical model of the formation of the ra<strong>di</strong>ographic image can be derived in<br />

the hypothesis of monochromatic X-ray source. Let us consider all the incident photons to have<br />

energy E and to be parallel to the Þ <strong>di</strong>rection. The detector lies in the ÜÝ plane. We assume that<br />

each photon interacting in the receptor is locally absorbed and that the response of the detector<br />

is linear, so that the image can be considered as a <strong>di</strong>stribution of absorbed energy. If there are<br />

Æ photons <strong>per</strong> <strong>un</strong>it area incident on the patient and Á Ü� Ý �Ü�Ý is the energy absorbed in the<br />

area �Ü�Ý of the detector, then:<br />

Á Ü� Ý �Ư �� �� Ê � Ü�Ý�Þ �Þ<br />

� ¯ �×�� �×Ë Ü� Ý� �×� ª �ª��×<br />

(1.1)<br />

where the first term is the primary photons contribution and the second term depends from<br />

the secondary photons. The line integral is over all tissues along the path of the primary pho-<br />

tons reaching the point Ü� Ý and � Ü� Ý� Þ is the linear attenuation coefficient. The scatter<br />

<strong>di</strong>stribution f<strong>un</strong>ction Ë is defined so that Ë Ü� Ý� �� ª �ª���Ü�Ý gives the number of scattered<br />

photons in the energy range � to � �� and the solid angle ª to ª �ª which pass through<br />

the area �Ü�Ý of the detector. The energy absorption efficiency ¯ of the detector is a f<strong>un</strong>ction<br />

of both the photon energy and the angle � between the photon <strong>di</strong>rection and the Þ axis. �× is<br />

the energy of the scattered photons. It is worth noting that if the detector has an efficiency not<br />

equal to one, the path length of the photon inside the detecting material will have an important<br />

effect on the global efficiency. In fact scattered photons will be absorbed more efficiently than<br />

primary photons, so that inefficient receptors will enhance the effects of scatter on the image.<br />

The scatter f<strong>un</strong>ction Ë has a complicated dependence on position and on the <strong>di</strong>stribution of<br />

tissues within the patient. For many applications, it is sufficient to treat it a slowly varying<br />

5


f<strong>un</strong>ction and to replace the very general integral in Eq. 1.1 with the value at the center of the<br />

image. As the scatter will decrease away from the center, this will give a maximal estimate of<br />

the contrast-degra<strong>di</strong>ng effects of the scatter. Eq. 1.1 then simplifies to:<br />

Á Ü� Ý �Ư �� �� Ê � Ü�Ý�Þ �Þ<br />

˯ � � (1.2)<br />

where<br />

�<br />

Ë � Ë � � ª��× �ª��× (1.3)<br />

and<br />

�<br />

¯ � � � ¯ �×�� �×Ë � ��×� ª �ª��×�Ë (1.4)<br />

In practice, it is the ratio of the scattered to primary ra<strong>di</strong>ation that is either measured or<br />

calculated, and an appropriate form of Eq. 1.2 is then:<br />

Á Ü� Ý �Ư �� �� Ê � Ü�Ý�Þ �Þ<br />

Ê (1.5)<br />

The parameters used by the ra<strong>di</strong>ologist to evaluate the image quality are the ra<strong>di</strong>ographic<br />

contrast, the spatial resolution and the noise. The last two terms will be examined more in<br />

detail in the following chapters. The contrast is defined with respect to the target tissue to be<br />

examined. Given the energy ÁÓ�� absorbed by the detector in correspondence of the target and<br />

the energy Á�� � of the surro<strong>un</strong><strong>di</strong>ng area, the contrast C is defined as:<br />

� � Á�� �<br />

Á�� �<br />

ÁÓ��<br />

(1.6)<br />

The term R is one of the responsible factors of the degradation in contrast in the final image.<br />

Other factors that affects the degradation in contrast are the target thickness and the <strong>di</strong>fference in<br />

linear attenuation coefficients. While the former factor is not <strong>un</strong>der the ra<strong>di</strong>ologist control, the<br />

former can be optimized by an appropriate beam quality choice. The contrast decreases rapidly<br />

with increasing photon energy, so that for the best contrast a low photon energy is better. This is<br />

particular important in mammography, where small targets, for example calcifications, must be<br />

detected. A problem arises if we consider the mechanism of interaction of photons with matter.<br />

If the transmission of photons is very low, as is the case with low energy beams, then very few<br />

6


photons will reach the detector and the ra<strong>di</strong>ation dose to the tissue will be very high in order<br />

to obtain a signal to noise ratio similar to the one obtained with a more energetic beam. The<br />

choice of energy will therefore be a compromise between the requirements of low dose and high<br />

contrast [3].<br />

At the energy values typical of ra<strong>di</strong>ography, the most important photons interactions are the<br />

photoelectric effect and scattering. In soft tissue the photoelectric cross section is larger than<br />

the scatter cross section for energies up to about 25 keV. The photoelectric cross section varies<br />

approximately as the fourth power of the atomic number and inversely as the third power of<br />

the photon energy and it shows <strong>di</strong>scontinuities at the absorption edges. The photon scattering<br />

cross section varies more slowly with energy than does the photoelectric cross section and is<br />

approximately proportional to atomic number [4].<br />

The voltage applied to a X-ray tube de<strong>di</strong>cated to mammography is usually between 20 and<br />

30 kVp (Vmax), depen<strong>di</strong>ng on the clinical <strong>di</strong>agnosis and patient morphology. The X-ray ra-<br />

<strong>di</strong>ation is generated by Bremmstrahl<strong>un</strong>g and by an X-ray emission characteristic of the anode<br />

material. The anode material used in mammography X-ray <strong>un</strong>its is molybdenum, since, with<br />

respect to t<strong>un</strong>gsten targets, it produces lower energy X-rays, which are more appropriate for<br />

imaging thinner body sections at high contrast. Usually the spectrum is attenuated by an added<br />

molybdenum filter (see Chapter 3.1 for the system <strong>un</strong>der investigation) which reduces both the<br />

X-ray component above its K-edge of about 20 keV and the lower energy components. This<br />

is important since it reduces the soft component of the spectrum, which contributes only to the<br />

patient dose and not to image contrast.<br />

1.2 The imaging plate IP<br />

The Imaging Plate is the detecting element of a CR system. It is a multilayer film composed<br />

of a support layer and a sensible layer made of photo-stimulable phosphors. Photo-stimulable<br />

phosphors, also known as storage phosphor, is one of the most successful detectors for <strong>di</strong>gital<br />

ra<strong>di</strong>ography to date.<br />

The photo-stimulable phosphor (PSP) stores the absorbed X-ray energy in crystal structure<br />

traps. The trapped energy can be released in a later moment if stimulated by ad<strong>di</strong>tional light<br />

7


energy of the pro<strong>per</strong> wavelength by the process of photo-stimulated luminescence (PSL here-<br />

after). The output signal shows a linear dependence to the absorbed dose. Furthermore the IP<br />

maintains this linearity over a wide range of exposures (four orders of exposure magnitude).<br />

The PSP is placed in a protective, light-tight cassette with the same geometrical characteris-<br />

tics of a film-based ra<strong>di</strong>ographic cassette, so that it can be exposed in a ra<strong>di</strong>ographic equipment.<br />

Using X-ray imaging techniques similar to screen film imaging, a latent image, in the form<br />

of trapped electrons is imprinted on the PSP receptor by absorption of the photons transmit-<br />

ted through the object. At this point, the <strong>un</strong>observable latent image is processed by inserting<br />

the IP cassette into an image reader, where the PSP plate is extracted from the cassette and<br />

raster-scanned with a highly focused red laser light. A higher energy, low intensity blue photo-<br />

stimulated luminescence signal is emitted, the intensity of which is proportional to the num-<br />

ber of x-ray photons that were absorbed in the local area of the receptor. The PSL signal is<br />

channeled to a photomultiplier tube, converted to a voltage, <strong>di</strong>gitized with an analog to <strong>di</strong>gital<br />

converter, and stored in a <strong>di</strong>gital image matrix. After PSP detector is totally scanned, analysis<br />

of the raw <strong>di</strong>gital data locates the <strong>per</strong>tinent areas of the useful image. Scaling of the data with<br />

well-defined computer algorithms creates a greyscale image that mimics the analog film image.<br />

Finally, the image is recorded on film, or viewed on a <strong>di</strong>gital image monitor. The described<br />

steps of the PSP rea<strong>di</strong>ng process are summarized in Fig. 1.3.<br />

PSP devices are based on the principles of photo-stimulated luminescence. When an X-<br />

ray photon deposits energy in the PSP material, the energy can be released by three <strong>di</strong>fferent<br />

physical processes. Fluorescence is the prompt release of energy in the form of light. This<br />

process is the basis of conventional ra<strong>di</strong>ographic intensification screens. PSP imaging plates<br />

also emit fluorescence in sufficient quantity to expose conventional ra<strong>di</strong>ographic film [5] [6],<br />

however this is not the intended method of imaging. PSP materials store some of the deposited<br />

energy in defects in their crystal structure, thus they are sometimes called storage phosphors.<br />

This stored energy constitutes the latent image. Over time, the latent image fades spontaneously<br />

by the process of phosphorescence. If stimulated to light of the pro<strong>per</strong> wavelength, the process<br />

of stimulated luminescence can release the trapped energy. The emitted light constitutes the<br />

signal for creating the <strong>di</strong>gital image [7].<br />

Many compo<strong>un</strong>ds possess the pro<strong>per</strong>ty of PSL. Few of these materials have characteristics<br />

8


Figure 1.3: Main components of a PSP reader.<br />

desirable for ra<strong>di</strong>ography, i.e. a stimulation-absorption peak at a wavelength produced by com-<br />

mon lasers, a stimulated emission peak rea<strong>di</strong>ly absorbed by common photomultiplier tube input<br />

phosphors, and retention of the latent image without significant signal loss due to phosphores-<br />

cence. The compo<strong>un</strong>ds that most closely meet these requirements are alkali-earth halides. Com-<br />

mercial products have been introduced based on Ê��Ð, ����Ö � �Ù , ��� �ÖÁ � �Ù ,<br />

��ËÖ��Ö � �Ù [27]. Trace amo<strong>un</strong>ts of impurities, such as �Ù , are added the PSP to<br />

alter its structure and physical pro<strong>per</strong>ties. The trace impurity is also called an activator. �Ù<br />

replaces the alkali earth in the crystal, forming a luminescence center. Ionization by absorption<br />

of X-rays forms electron/hole pairs in the PSP crystal. An electron/hole pair raises �Ù to<br />

an excited state, �Ù . �Ù produces visible light when it returns to the gro<strong>un</strong>d state, �Ù .<br />

Stored energy (in the form of trapped electrons) forms the latent image.<br />

There are currently two major theories for the PSP energy absorption process and subse-<br />

quent formation of luminescence centers: a bimolecular recombination model [8], and a pho-<br />

tostimulable luminescence complex (PSLC) model [9]. The energy levels and the interaction<br />

mechanisms proposed by the two theories are shown in Fig. 1.4.<br />

Physical processes occurring in ����Ö � �Ù using the latter theory appears to closely<br />

9


Figure 1.4: Energy levels of the PSP.<br />

approximate the ex<strong>per</strong>imental fin<strong>di</strong>ngs. In this model, the PSLC is a metastable complex at<br />

higher energy (F-center) in close proximity to an �Ù �Ù recombination center. X-rays<br />

absorbed in the PSP induce the formation of ”holes” and ”electrons”, which either activate an<br />

”inactive PSLC” by being captured by an F-center, or form an active PSLC via formation and<br />

recombination of ”excitons” explained by ”F-center physics”.<br />

In either situation, the number of active PSLC’s created (number of electrons trapped in the<br />

metastable site) are proportional to the x-ray dose to the phosphor, critical to the success of the<br />

phosphor as an image receptor.<br />

Fa<strong>di</strong>ng of the trapped signal will occur exponentially over time, through spontaneous phos-<br />

phorescence. A typical imaging plate will lose about 25% of the stored signal between 10<br />

minutes to 8 hours after an exposure, and more slowly afterwards. Fa<strong>di</strong>ng introduces <strong>un</strong>certain-<br />

ties in output signal that can be controlled by introducing a fixed delay between exposure and<br />

readout to allow decay of the fast component phosphorescence of the stored signal.<br />

The latent image imprinted on the exposed BaFBr:Eu phosphor corresponds to the activated<br />

PLSC’s (F-centers), whose local population of electrons is <strong>di</strong>rectly proportional to the incident<br />

x-ray flux for a wide range of exposures, typically excee<strong>di</strong>ng 10,000 to 1 (four orders of expo-<br />

sure magnitude). Stimulation of the �Ù F-center complex and release of the stored electrons<br />

requires a minimum energy of � eV, most easily deposited by a highly focused laser light<br />

10


source of a given wavelength. Laser light produced by He-Ne (� � � nm) and solid state<br />

(� � �� nm) sources are most often used. The incident laser energy excites electrons in the<br />

local F-center sites of the phosphor. Accor<strong>di</strong>ng to von Seggern [9], two subsequent energy<br />

pathways within the phosphor matrix are possible-to return to the F-center site without escape,<br />

or to t<strong>un</strong>nel to an adjacent Eu3+ complex. The latter event is more probable, where the electron<br />

cascades to an interme<strong>di</strong>ate energy state with the release of a non-light emitting ”phonon”. A<br />

light photon of 3 eV energy imme<strong>di</strong>ately follows as the electron continues to drop through the<br />

electron orbitals of the �Ù complex to the more stable �Ù energy level.<br />

The readout hardware will be described in Chapter 3.<br />

1.2.1 The single and dual-side rea<strong>di</strong>ng<br />

One of the most important factors in determining the image quality in an X-ray imaging system<br />

is the X-ray utilization efficiency. In fact the noise components can be roughly classified into<br />

two categories: quantum noise, dependent upon X-ray exposure, and fixed noise, independent<br />

from exposure. Quantum noise includes X-ray photon noise which is caused by X-ray spatial<br />

fluctuations, and “light photon noise”, caused by temporal fluctuations of the photo-electrons in<br />

the photomultiplier tube (PMT). Fixed noise includes, for example, IP structural noise, electri-<br />

cal system noise and laser noise. Especially at low spatial frequencies the X-ray photon noise<br />

is dominant, therefore, in order to achieve an improvement in image quality of the CR sys-<br />

tem, reduction of this noise source would be most effective [11]. In order to optimize X-ray<br />

utilization efficiency, X-ray absorption should be increased, together with the photostimulated<br />

luminescence detected by the photomultiplier tube. With the same image rea<strong>di</strong>ng system, in-<br />

creasing the thickness of the IP system would be effective, with the drawback of an acceptable<br />

worsening of the spatial resolution <strong>per</strong>formances due to multiple scattering.<br />

In the past the IP scanning system was based on a single-side rea<strong>di</strong>ng system. However,<br />

in the single-side rea<strong>di</strong>ng method (see Fig. 1.5), as the thickness of the IP is increased, the<br />

light emission in the deep phosphor layers is less likely to be detected by the PMT, because the<br />

photons must pass through thicker layers of material before they can reach the photodetector<br />

located on the front side. Thus the amo<strong>un</strong>t of photostimulated luminescence that can be detected<br />

by the PMT increases but tends to saturate as the phosphor layer is any thicker.<br />

11


Figure 1.5: Single-side rea<strong>di</strong>ng method and cross section of a conventional IP<br />

To overcome this problem the FUJI mammographic system <strong>un</strong>der investigation employs a<br />

particular IP that consists of a thicker photostimulable phosphor layer (320 �m instead of the<br />

usual 230 �m thickness) combined with a transparent support and a dual-side rea<strong>di</strong>ng [10]. The<br />

rea<strong>di</strong>ng system, described in Fig. 1.6, is equipped with a photodetector on the support side as<br />

well, in order to detect light that emission from both sides of the IP.<br />

Figure 1.6: Dual-side rea<strong>di</strong>ng method and cross section of a transparent support IP<br />

The overall result is that the emitted photons correspon<strong>di</strong>ng to the X-ray absorbed in the deep<br />

inside of the phosphor layer can be detected more efficiently by the PMT on the back side. Thus<br />

the X-ray absorption can be increased substantially by adopting a thicker phosphor layer.<br />

The image data detected by the respective PMT are added together with an appropriate<br />

frequency-dependent ad<strong>di</strong>tion ratio and then used as the final image data. The improvement in<br />

terms of NEQ or DQE described in literature is about 30-40 % [11]. The characteristics of the<br />

12


ea<strong>di</strong>ng system and of the IP <strong>un</strong>der investigation are describe more in detail in Chapter 3.<br />

1.3 The <strong>di</strong>gitization process<br />

Digitization is a two step process of converting an analog signal into a <strong>di</strong>screte <strong>di</strong>gital value.<br />

The signal must be sampled and quantized. Sampling determines the location and size of the<br />

PSL signal from a specific area of the PSP receptor, and quantizing determines the average<br />

value of the signal amplitude within the sample area. The output of the PMT is measured at<br />

a specific temporal frequency coor<strong>di</strong>nated with the laser scan rate, and quantized to a <strong>di</strong>screte<br />

integer value dependent on the amplitude of the signal and the total number of possible <strong>di</strong>gital<br />

values by the Analog to Digital Converter (ADC). The ADC converts the PMT signals at a rate<br />

much faster than the fast scan rate of the laser (on the order of 2000 times faster, correspon<strong>di</strong>ng<br />

to the number of pixels in the scan <strong>di</strong>rection). A pixel clock coor<strong>di</strong>nates the time at which a<br />

particular signal is encoded to a physical position on the scan line. Therefore, the ratio between<br />

the ADC sampling rate and the fast scan (line) rate determines the pixel <strong>di</strong>mension in the scan<br />

<strong>di</strong>rection. The translation speed of the phosphor plate in the sub scan <strong>di</strong>rection coor<strong>di</strong>nates with<br />

the fast scan pixel <strong>di</strong>mension so that the width of the line is equal to the length of the pixel (i.e.<br />

the pixels are ”square”).<br />

ADC<br />

sample<br />

Figure 1.7: Schematic <strong>di</strong>agram of the <strong>di</strong>gitization process.<br />

The pixel size is typically between 100 and 200 �m but in mammographical applications a<br />

pixel size of 50 �m is preferred. Typically the ADC converts the signal to a 10,12 or 16 bits<br />

13


<strong>di</strong>gital levels. The system <strong>un</strong>der investigation has a 12 bit <strong>di</strong>gitization ADC prior to implement a<br />

software transformation to 10 bit/pixel image (i.e. 1024 grey levels). Furthermore the FUJI FCR<br />

5000 MA reader uses a logarithmic amplifier on the pre-<strong>di</strong>gitized data. Analog amplification<br />

prior to the final <strong>di</strong>gital conversion reduces quantization errors in the signal estimate. The spatial<br />

sampling process is described in Fig. 1.7.<br />

The <strong>di</strong>stance between two consequent laser beam positions gives the sampling <strong>di</strong>stance,<br />

while the pixel size gives the sampling a<strong>per</strong>ture. In the FCR5000 MA sampling <strong>di</strong>stance and<br />

sampling a<strong>per</strong>ture are equal and correspond to the pixel size. The inverse of twice the sampling<br />

<strong>di</strong>stance corresponds to the Nyquist frequency.<br />

14


Chapter 2<br />

Characterization of a <strong>di</strong>gital ra<strong>di</strong>ography<br />

system<br />

In this chapter the main physical quantities involved in the CR characterization will be defined<br />

and the mathematical formalism required will be developed in the Fourier framework. The first<br />

section is devoted to the definition of linear systems and to the Fourier methods for the image<br />

transfer f<strong>un</strong>ctions. In the following sections the system response f<strong>un</strong>ction will be described as<br />

well as the concepts of Modulation Transfer F<strong>un</strong>ction (MTF), Pre-sampled MTF, Noise Equiv-<br />

alent Quanta and Detective Quantum Efficiency. This last quantity is the one that will be used<br />

as final descriptor of the efficiency and of the global <strong>per</strong>formances of the system <strong>un</strong>der investi-<br />

gation.<br />

2.1 Linear systems<br />

The CR system <strong>per</strong>formances will be analyzed in the spatial frequency domain in order to evalu-<br />

ate the capability to detect structures in clinical images. To this end, Fourier techniques provide<br />

a particularly elegant framework from which we can evolve a description of the formation of<br />

images. A key point in this analysis is the possibility to treat our system as linear. Before<br />

developing the required formalism, the definition of linear system will be recalled. Let suppose<br />

that a bi-<strong>di</strong>mensional input signal � Ü� Ý passing through some optical and electrical system<br />

results in an output � �� � . The system is linear if:<br />

15


¯ multiplying � Ü� Ý by a constant � produces an output �� �� � .<br />

¯ when the input is a weighted sum of two (or more) f<strong>un</strong>ctions, �� Ü� Ý �� Ü� Ý , the<br />

output will similarly have the form �� �� � �� �� � , where �� �� � is the image<br />

of �� Ü� Ý .<br />

Furthermore, a linear system will be space invariant if it possesses the pro<strong>per</strong>ty of stationarity.<br />

A system is stationary if changing the position of the input merely changes the location of the<br />

output without altering its f<strong>un</strong>ctional form. The idea behind much of this is that the output<br />

produced by an optical system can be treated as a linear su<strong>per</strong>position of the outputs arising<br />

from each of the in<strong>di</strong>vidual points of the object. Indeed, if we symbolically represent the linear<br />

transformation of the system as ��, the input and output can be written as<br />

� �� � �Ä�� Ü� Ý � (2.1)<br />

Using the shifting pro<strong>per</strong>ty of the Æ-f<strong>un</strong>ction this becomes<br />

� �� � �Ä<br />

� ��<br />

� Ü �Ý Æ Ü Ü Æ Ý Ý �Ü �Ý<br />

The integral expresses � Ü� Ý as a linear combination of elementary delta f<strong>un</strong>ctions, each<br />

weighted by a quantity � Ü �Ý . It follows from the second linearity con<strong>di</strong>tion that the sys-<br />

�<br />

(2.2)<br />

tem o<strong>per</strong>ator can equivalently act on each of the elementary f<strong>un</strong>ctions, thus obtaining:<br />

� �� � �<br />

�� �<br />

� Ü �Ý Ä Æ Ü Ü Æ Ý<br />

�<br />

Ý �Ü �Ý (2.3)<br />

�<br />

��<br />

�<br />

� Ü Ü �Ý Ý Ä Æ Ü Æ Ý<br />

�<br />

The quantity � Ü �Ý���� � Ä Æ Ü Æ Ý<br />

� �Ü �Ý<br />

� is the response of the linear system to a delta<br />

f<strong>un</strong>ction located at the point Ü �Ý in the input space, the so-called impulse response. In<br />

general it will depend both on the input and output space coor<strong>di</strong>nates. We can define the Point<br />

Spread F<strong>un</strong>ction (PSF) as the normalized impulse response:<br />

ÈË� Ü� Ý� �� � �<br />

ÊÊ<br />

� � � �� �<br />

� Ü� Ý� �� � �Ü�Ý<br />

16<br />

�<br />

� � � �� �<br />

£<br />

(2.4)


where £ is the system gain.<br />

The PSF has a f<strong>un</strong>ctional form identical to that of the image generated by a Æ-pulse input. If<br />

the system is space invariant, a point-source input can be moved along the object plane without<br />

any effect other than changing the location of its image, the f<strong>un</strong>ction being the same for any<br />

point Ü� Ý . So the dependence of the PSF on space variables can only be related to Ü� Ý as<br />

far as the location of its center is concerned. The value of the PSF on the image plane merely<br />

depends on the <strong>di</strong>splacement at that location from the particular image point at which the PSF<br />

is centered (see Fig. 2.1).<br />

Figure 2.1: Convolution of a source composed of Æ f<strong>un</strong>ctions with the PSF. The resulting pattern<br />

is the sum of all the spread out contributions.<br />

In other words it can be stated that:<br />

ÈË� Ü� Ý� �� � �ÈË� � Ü� � Ý (2.5)<br />

then, <strong>un</strong>der the hypotheses of space invariance and linearity,<br />

� �� � �£<br />

��<br />

� Ü� Ý ÈË� � Ü� � Ý �Ü�Ý (2.6)<br />

It is also worth defining the Line Spread F<strong>un</strong>ction (LSF) as the integral of the PSF along<br />

one <strong>di</strong>rection, since this is the quantity that will be used to ex<strong>per</strong>imentally evaluate the spatial<br />

17


frequency pro<strong>per</strong>ties of our system.<br />

ÄË� Ü� �� � �<br />

ÊÊ<br />

Ê<br />

� Ü� Ý� �� � �Ý<br />

� Ü� Ý� �� � �Ü�Ý<br />

The LSF carries information about the system response to a linear input.<br />

as:<br />

(2.7)<br />

The expression in Eq. 2.6 is a bi-<strong>di</strong>mensional convolution integral and is usually expressed<br />

� �� � �£� Ü �Ý ª ÈË� � Ü� � Ý (2.8)<br />

At this point, Fourier theory provides an elegant and powerful framework to deal with the<br />

system analysis, starting from the convolution integral. Let us define the Fourier transforms<br />

� of the f<strong>un</strong>ctions � Ü� Ý ,� �� � and � � Ü� � Ý (� is the non-normalized PSF):<br />

��� Ü� Ý � � � Ù� Ú , ��� �� � � � � Í� Î and ��� Ü� Ý � � À Ù� Ú . The convolu-<br />

tion theorem states that:<br />

or<br />

���� � ��� ª �� � ��� �¡���� (2.9)<br />

� Í� Î �� Ù� Ú ¡ À Ù� Ú (2.10)<br />

In our subsequent analysis a linear and space invariant system will be considered, <strong>un</strong>less<br />

otherwise stated. In case the relation between the input mean value of the quantities <strong>un</strong>der<br />

investigation is not linearly related to the output mean value, the linearity con<strong>di</strong>tion can be<br />

recovered by using the sensitometric curve.<br />

2.2 The Modulation Transfer F<strong>un</strong>ction MTF<br />

In order to estimate the capability of an imaging system to map the amplitudes of input fre-<br />

quency components to the amplitudes of output frequency components, the concept of Modula-<br />

tion Transfer F<strong>un</strong>ction is usually introduced for analog systems. Let us consider a bi<strong>di</strong>mensional<br />

18


wave object � �� � , oscillating with spatial frequencies �� and �� and amplitude � aro<strong>un</strong>d a<br />

constant value �:<br />

� �� � �� ¡ � � � � �� ���<br />

� (2.11)<br />

A highly useful parameter in evaluating the <strong>per</strong>formances of a system is the contrast or modu-<br />

lation Å, defined by:<br />

Å � �Ñ�Ü �Ñ�Ò<br />

�Ñ�Ü �Ñ�Ò<br />

� ��� (2.12)<br />

The image of the object trough the system can be calculated, in the linear space invariant system<br />

approximation, as:<br />

� Ü� Ý �<br />

� �<br />

� �£<br />

� �£<br />

�£<br />

� �<br />

� Ü �� Ý � ¡ � �� � ���� (2.13)<br />

� �<br />

� �<br />

ÈË� Ü �� Ý � � � � � �� ��� ����<br />

ÈË� Ü �� Ý � ����<br />

ÈË� Ü �� Ý � � � � � � � Ü �� � Ý � � � � �Ü ��Ý ���� �£<br />

where £ is the gain of the system. After a change in the integration variables � � Ü � and<br />

� � Ý �, the same expression is:<br />

� Ü� Ý ��£� � � � �Ü ��Ü<br />

� �<br />

ÈË� �� � � � � � �� ��� ���� �£ (2.14)<br />

The integral is the Fourier transform of the PSF and is usually referred to as OTF (Optical<br />

Transfer F<strong>un</strong>ction). The OTF is thus defined as the Fourier transform of the output of a system<br />

which has a delta f<strong>un</strong>ction as its input. The OTF is two-<strong>di</strong>mensional for a two-<strong>di</strong>mensional<br />

image, but for simplicity of nomenclature only the one-<strong>di</strong>mensional case will be considered<br />

here. With a last change of variables �Ü � �� and �Ý � ��, the image � Ü� Ý can be written as:<br />

� � �ÜÜ �ÝÝ<br />

� Ü� Ý ��£ÇÌ� �Ü��Ý �<br />

�£ (2.15)<br />

It appears that the image of a sinusoid is still a sinusoid and, as it could be expected from the<br />

system linearity, with the same spatial frequency of the object; but the amplitude depends both<br />

19


on the gain £ and on the f<strong>un</strong>ction OTF. The OTF is a complex, frequency dependent f<strong>un</strong>ction<br />

described by a module and a phase �:<br />

ÇÌ� � �ÇÌ� �� �� � ÅÌ�� ��<br />

(2.16)<br />

where the MTF (Modulation Transfer F<strong>un</strong>ction) is the OTF module, and � is also known as<br />

Phase Transfer F<strong>un</strong>ction. If the input image is real (as will be always assumed to be the case),<br />

then the real component of the OTF is symmetric about the zero frequency and the imaginary<br />

component is antisymmetric, yiel<strong>di</strong>ng a symmetric MTF. Therefore only one-half of the MTF<br />

is typically reported (the positive frequency). We stress the fact that the MTF acco<strong>un</strong>ts for the<br />

transfer of the modulation between the object and the image, at each spatial frequency. In fact,<br />

the ratio of the object modulation ÅÓ�� and the image modulation Å�Ñ is the MTF:<br />

ÅÓÙØ<br />

Å�Ò<br />

�<br />

�¡£¡ÅÌ�<br />

�£<br />

�<br />

�<br />

� ÅÌ� (2.17)<br />

Furthermore, one of the advantages of using MTF as a <strong>per</strong>formance index of a system is that if<br />

the MTFs for the in<strong>di</strong>vidual independent components in a system are known, the total MTF is<br />

often simply their product.<br />

The more <strong>di</strong>rect method for the MTF measurements is to expose the detector to in<strong>di</strong>vidual<br />

monochromatic signals of known amplitude, thus evaluating the MTF for each point in the<br />

spatial frequency domain. The MTF can be measured using filters with a sinusoidal attenuation<br />

profile. This poses several technological constraints on the realization of the filters and of the<br />

measurement, even if a square wave profile (easier to realize), with the pro<strong>per</strong> correction factor,<br />

is used instead of the sinusoidal profile. A second method relies on the pro<strong>per</strong>ty of signals<br />

having a very steep gra<strong>di</strong>ent in their structure. A f<strong>un</strong>ction with an infinite gra<strong>di</strong>ent, such as a<br />

Dirac delta or a step f<strong>un</strong>ction or a narrow slit, has a Fourier spectrum covering all the frequency<br />

domain, with zero values at most on a <strong>di</strong>screte sample of points. It is <strong>di</strong>fficult to realize a<br />

delta-like input but it is easier to produce a sharp edge or a narrow slit simulating a line, thus<br />

modeling the input with a f<strong>un</strong>ction whose Fourier transform is known. From the output image,<br />

which contains all the spatial frequency information, the OTF can be obtained. In this work a<br />

slit has been used as mono-<strong>di</strong>mensional input filter, to evaluate the frequency response of the<br />

system. In fact the Fourier transform of an ideal, zero width, slit in a <strong>di</strong>mension orthogonal to<br />

20


its <strong>di</strong>rection is a constant f<strong>un</strong>ction covering all the spatial frequency range. The image of a slit<br />

is the so-called Line Spread F<strong>un</strong>ction.<br />

2.2.1 Digital MTF<br />

In a <strong>di</strong>gital system, the sampling procedure usually leads to complication due to <strong>un</strong>dersampling<br />

in the MTF analysis since, as will be explained later, the system response depends on the spatial<br />

frequency content of the images being evaluated. Undersampling in <strong>di</strong>gital systems occurs<br />

when the image is not sampled finely enough to record all spatial frequencies without aliasing.<br />

Undersampling is almost always present to some degree in any real <strong>di</strong>gital imaging device,<br />

and not only makes the physical measurement of MTF more <strong>di</strong>fficult but it also complicates the<br />

pro<strong>per</strong> <strong>un</strong>derstan<strong>di</strong>ng and interpretation of these quantitative measures. The main complications<br />

are related to the fact that, when applying classical analysis to <strong>un</strong>dersampled <strong>di</strong>gital systems,<br />

the MTF do not behave as transfer amplitude of a single sinusoid and the response of a <strong>di</strong>gital<br />

system to a delta f<strong>un</strong>ction is not spatially invariant. There are two main elements <strong>di</strong>stinguishing<br />

the MTF of a <strong>di</strong>gital system from that of an analog system: replication of FTs in frequency<br />

space and the overlapping of FT segments from aliasing 1 when the system is <strong>un</strong>dersampled.<br />

While replication and aliasing overlap are certainly related, they are <strong>di</strong>fferent in the effects they<br />

have on the interpretation of MTF in <strong>un</strong>dersampled <strong>di</strong>gital systems. The replication of FTs<br />

is a result of the infinite sum of sinusoids required to produce a signal comprised of a string<br />

of infinitely sharp delta f<strong>un</strong>ction, i.e. from multiplying the original f<strong>un</strong>ction by the sampling<br />

f<strong>un</strong>ction � ¡ ÁÁÁ Ü� � . The overlap of these FT replications, on the other hand, is the result of<br />

<strong>un</strong>dersampling, by which aliasing causes sampled frequencies above the Nyquist frequency to<br />

contaminate their co<strong>un</strong>terpart frequencies below the Nyquist frequency (see Fig. 2.2).<br />

In order to <strong>un</strong>derstand how these effects reflect quantitatively on the MTF, the terms con-<br />

tributing to the system OTF must be considered. The OTF of a <strong>di</strong>gital system is comprised<br />

of a presampling component and a sampled component. The presampled OTF is the result of<br />

image blurring from geometric considerations (for instance focal spot blurring), analog input<br />

1aliasing is so named because a sampled sinusoid of frequency¡Ù above the Nyquist frequency Ù Æ is identical<br />

in every regard to a sampled sinusoid of frequency ¡Ù below Ù Æ (if at the same phase); thus the higher-frequency<br />

sinusoid takes the alias of the lower frequency.<br />

21


(a)<br />

(b)<br />

(c)<br />

-u N<br />

MTF pre<br />

MTF pre<br />

-u N<br />

MTF<br />

Figure 2.2: System response to a single sinusoids. (a) MTF of an analog system. (b) MTF<br />

of a <strong>di</strong>gital system without <strong>un</strong>dersampling. (c) MTF of a <strong>di</strong>gital system with <strong>un</strong>dersampling.<br />

The solid line in<strong>di</strong>cates the total <strong>di</strong>gital MTF inclu<strong>di</strong>ng the aliasing overlap from adjacent FT<br />

replications.<br />

subsystems (e.g. the IP response) and the a<strong>per</strong>ture f<strong>un</strong>ction of the acquisition device (e.g. the<br />

effective laser beam shape):<br />

ÇÌ�ÔÖ� � ÇÌ���ÓÑ ¡ ÇÌ��Ò�ÐÓ�<br />

u 1<br />

MTF<br />

u 1<br />

MTF<br />

u 1<br />

u 2<br />

u 2<br />

u 2<br />

u N<br />

u N<br />

u<br />

u<br />

u<br />

(2.18)<br />

The image is <strong>di</strong>gitally sampled using the sampling theorem, giving rise to the <strong>di</strong>gital OTF<br />

(ÇÌ����):<br />

ÇÌ���� Ù ��ÇÌ�ÔÖ� Ù £ Û ¡ ×�Ò ÙÛ ℄ £ ÁÁÁ Ù� �<br />

(2.19)<br />

where Û is the width of the image, � is the sampling interval, and £ denotes convolution. The<br />

sampling f<strong>un</strong>ction �¡ÁÁÁ Ü� � is a string of delta f<strong>un</strong>ctions separated by the pixel spacing � in the<br />

Cartesian space; its Fourier transform ÁÁÁ Ù� � gives rise to replication in frequency space.<br />

The factor Û ¡ ×�Ò ÙÛ in frequency space corresponds to the final image width in Cartesian<br />

22


space. Since Û is almost always much greater than the width of the presampling LSF, one can<br />

usually approximate Û×�Ò ÙÛ as Æ Ù . The ÇÌ���� then simplifies to:<br />

ÇÌ���� Ù � � ÇÌ�ÔÖ� Ù £ ÁÁÁ Ù� �<br />

(2.20)<br />

The MTF is defined for analog systems as �ÇÌ� �. As long as there is no aliasing from<br />

<strong>un</strong>dersampling, the same definition can also be applied to <strong>di</strong>gital systems without alteration.<br />

However, when a <strong>di</strong>gital system is <strong>un</strong>dersampled, two conceptual <strong>di</strong>fficulties arise: the <strong>di</strong>gital<br />

MTF no longer describes the amplitude of a single frequency passed by the system (see Fig. 2.2)<br />

and it is phase dependent, and therefore not spatially invariant as required for the stationarity<br />

pro<strong>per</strong>ty of the linear system definition of MTF.<br />

The <strong>di</strong>fficulties arising from the first item can be explained considering that two working<br />

definition of MTF exist. The MTF can be measured as the response of a system to a delta<br />

f<strong>un</strong>ction:<br />

ÅÌ���� Ù � �ÇÌ���� Ù �<br />

�ÇÌ����<br />

�<br />

(2.21)<br />

i.e. the MTF is defined as the frequency output of a system when an input consisting of <strong>un</strong>iform<br />

frequency content is present. On the other side the MTF can be measured as the amplitude<br />

mo<strong>di</strong>fication of a single sinusoid passed by a system:<br />

ÅÌ� ��� Ù � ��Ì��� Ù �<br />

��Ì�Ò Ù �<br />

(2.22)<br />

where �Ì�Ò and �Ì��� are the amplitudes of the sinusoid before and after sampling. These<br />

two working definitions of MTF are equivalent for analog systems and for <strong>di</strong>gital systems at<br />

frequencies <strong>un</strong>affected by aliasing (see for instance Fig. 2.2 (a) and (b)). However the two def-<br />

initions do not agree in <strong>un</strong>dersampled <strong>di</strong>gital systems for frequencies where aliasing causes an<br />

overlap of adjacent MTF replications. In summary, the amplitude response of an <strong>un</strong>dersampled<br />

<strong>di</strong>gital system to a single sinusoid contains replicated but not overlapped values, and is equal to<br />

ÅÌ�ÔÖ�. The standard definition of ÅÌ����, on the other hand, is the response of a system to a<br />

delta f<strong>un</strong>ction input, and contains overlapped values if <strong>un</strong>dersampled. The second item relates to<br />

the fact that ÅÌ���� depends on the phase relation of the sampling comb f<strong>un</strong>ction with respect<br />

to the f<strong>un</strong>ction being <strong>di</strong>gitized, when aliasing occurs [12]. This effect is described in Fig. 2.3,<br />

where the ÅÌ�ÔÖ� response of the system to a Æ f<strong>un</strong>ction is shown. If the sampling comb is<br />

23


(a)<br />

(b)<br />

(c)<br />

-u N<br />

-u N<br />

-u N<br />

MTF p re<br />

|OTF |<br />

d ig<br />

|OTF |<br />

d ig<br />

Figure 2.3: The effects of phase dependence on <strong>di</strong>gital MTF.<br />

aligned exactly with the delta f<strong>un</strong>ction, the resulting sampled �ÇÌ����� is given in Fig. 2.3(b).<br />

However, if the phase of the input delta f<strong>un</strong>ction is shifted by �� with respect to the sampling<br />

comb, the sampled �ÇÌ����� takes the form in Fig. 2.3(c). Values of the shift between and ��<br />

gives interme<strong>di</strong>ate curves. The phase dependence of �ÇÌ����� can be derived mathematically.<br />

If the delta input f<strong>un</strong>ction is at a <strong>di</strong>stance � relative to the origin and the sampling comb f<strong>un</strong>ction<br />

is centered on the origin, the ÇÌ���� is given by:<br />

ÇÌ���� Ù� � ��� � ��Ù ÇÌ�ÔÖ� Ù ℄ £ ÁÁÁ Ù� �<br />

u N<br />

u N<br />

u N<br />

u<br />

u<br />

u<br />

(2.23)<br />

where the factor � � ��Ù represents the phase shift � relative to the origin. The value of ÇÌ����<br />

at any frequency is just the sum of all frequency components that overlap the frequency of<br />

interest due to aliasing:<br />

24


ÇÌ���� Ù� � �<br />

�<br />

�<br />

�<br />

� Ó× �Ù�� Ê Ù� ×�Ò �Ù�� Å Ù� ℄ (2.24)<br />

�<br />

�<br />

�×�Ò �Ù�� Ê Ù� Ó× �Ù�� Å Ù� ℄ ¢<br />

where Ê � Ê��ÇÌ�ÔÖ��,Å � ÁÑ�ÇÌ�ÔÖ��, and the term<br />

�<br />

� acco<strong>un</strong>ts for the anti-<br />

symmetry of the replications of the imaginary part of ÇÌ����. After computing the amo<strong>un</strong>t of<br />

overlap, ÇÌ���� will be the same at each replication; therefore one needs only to <strong>per</strong>form the<br />

calculations in the frequency range ÙÆ to ÙÆ. The amplitude of ÇÌ����, that is ÅÌ����,<br />

is symmetric about zero frequency, so its value only needs to be calculated in the frequency<br />

range 0 to ÙÆ. For each frequency Ù in this range, one can evaluate the sum of all overlapping<br />

frequencies components in Eq. 2.25 and take the amplitude [12]:<br />

�ÇÌ���� Ù� � � � �<br />

�<br />

��<br />

ÅÌ� ÔÖ� Ù�<br />

�<br />

��<br />

�<br />

�� �����<br />

�Ê Ù� Ê Ù�<br />

È��Å Ù� Å Ù� ℄ Ó×� �� Ù� È��Ù� ℄<br />

�<br />

��<br />

�<br />

�� �����<br />

�Å Ù� Ê Ù�<br />

È��Ê Ù� Ê Ù� ℄<br />

¢ ×�Ò� �� Ù� È��Ù� ℄� � Ù � ÙÆ<br />

(2.25)<br />

where �� � if � and � are both odd or both even, and �� � if � and � have opposite<br />

parities. The frequencies in the sum are denoted Ù �Ù �Ù etc., all positive, and are in the ranges<br />

�Ù �ÙÆ, ÙÆ �Ù � ÙÆ etc. These frequencies are defined as:<br />

Ù� � Ù � ÙÆ for i odd (2.26)<br />

Ù� � ÙÆ Ù � ÙÆ for i even (2.27)<br />

Typically only a few terms of each sum will be needed due to the limited bandwidth of<br />

ÇÌ�ÔÖ�. The <strong>di</strong>agram of the terms contributing to the real part of ÇÌ���� is shown in Fig. 2.4.<br />

Finally, the value of the ÅÌ���� is given by:<br />

ÅÌ���� Ù � � � �ÇÌ�ÔÖ� Ù � � �<br />

�ÇÌ�ÔÖ� � � �<br />

25<br />

� Ù � ÙÆ<br />

(2.28)


-2u N<br />

Re { OTF }<br />

<strong>di</strong>g<br />

u1 u2 u3 uN 2uN Figure 2.4: Definition of overlapping frequencies in an <strong>un</strong>dersampled system.<br />

The value of ÅÌ���� outside the range � Ù � ÙÆ can be fo<strong>un</strong>d by reflecting ÅÌ���� Ù � �<br />

about zero frequency and replicating it every ÙÆ in frequency space. The phase dependence<br />

of ÅÌ���� is demonstrated since ÅÌ���� is explicitly a f<strong>un</strong>ction of the phase �. If the <strong>di</strong>gital<br />

system were not <strong>un</strong>dersampled, then there would be no overlap of aliased frequency components<br />

and the sum in Eq. 2.26 would have only one term, leaving �ÇÌ�ÔÖ� Ù � � � � ÅÌ�ÔÖ� Ù in<br />

the range � Ù � ÙÆ, which is independent from phase �. In section 4.3 a method will be<br />

described to overcome this problem by defining a new f<strong>un</strong>ction that takes into acco<strong>un</strong>t the MTF<br />

average over all possible phase values.<br />

2.3 NPS<br />

The noise contribution can be introduced in this analysis by ad<strong>di</strong>ng a statistical term Ò �� �<br />

to the expression for the image � Ü� Ý in Eq. 2.6. We obtain:<br />

� �� � �£<br />

��<br />

� Ü� Ý ÈË� � Ü� � Ý �Ü�Ý Ò �� � (2.29)<br />

A first quantitative analysis can be made by considering Ò �� � as a statistical fluctuation<br />

aro<strong>un</strong>d the average value g(X,Y). The variance � for the statistical process on the image plane<br />

is given by:<br />

� � Ð�Ñ<br />

����� � ��<br />

� �� ��<br />

��<br />

� ��<br />

��<br />

Ò �� � ���� (2.30)<br />

26<br />

u


where � � and � � are the image <strong>di</strong>mensions. The integration can be extended to the entire<br />

plane if we consider Ò �� � � outside the image. Due to the Parseval theorem [13], the<br />

same relation is valid in the spatial frequency domain for the Fourier components �Ò ����� of<br />

Ò �� � :<br />

� � Ð�Ñ<br />

��� �� � � � � �<br />

�<br />

�<br />

��Ò ����� � ������ (2.31)<br />

Starting from Eq. 2.31, the noise spectral density Ï ����� is defined as:<br />

Ï ����� � Ð�Ñ<br />

��� � � � � � � � ��Ò ����� � (2.32)<br />

Finally, in order to take into acco<strong>un</strong>t the fact that the process statistics can depend on multiple<br />

realizations, the average of the noise spectral density is computed over a large ensemble of<br />

realizations, thus obtaining the Noise Power Spectrum (NPS):<br />

ÆÈË ����� � Ð�Ñ<br />

��� � � �<br />

� �� � � ��Ò ����� � � (2.33)<br />

The NPS (or Wiener spectrum) describes the variance in amplitude of each frequency com-<br />

ponent of a system.<br />

2.4 DQE<br />

The Detective Quantum Efficiency DQE is a physical quantity that takes into acco<strong>un</strong>t the imag-<br />

ing system <strong>per</strong>formances both from the noise and from the spatial resolution point of view. In<br />

order to <strong>un</strong>derstand the <strong>un</strong>derlying reasons that lead to the definition of the o<strong>per</strong>ative DQE def-<br />

inition, let us consider the expected value �� ℄and the variance Î �℄of an image signal. Given<br />

the relation � between input and output we obtain the system characteristic curve in absence<br />

of noise:<br />

��ÇÍÌ ℄�� ��ÁÆ℄ (2.34)<br />

The noise can be added as a term Æ�, with a zero expectation value ��Æ�℄ � , such that the<br />

output is:<br />

ÇÍÌ � � ÁÆ Æ� (2.35)<br />

27


for every exposition. The output variance reflects the dependence from the input signal variance<br />

and from the system intrinsic, not eliminable, variance. In order to quantify the worsening of the<br />

image information due to the system, the simple ratio between the output and output variance<br />

is not significant, since it usually compares <strong>di</strong>fferent quantities and does not take into acco<strong>un</strong>t<br />

the f<strong>un</strong>ctional relation between input and output.<br />

Instead a more efficient quantity is usually used, the DQE, defined as [14]:<br />

�É� � ËÆÊ ÓÙØ<br />

ËÆÊ �Ò<br />

where the signal to noise ratio SNR, for the input and output components respectively, is:<br />

ËÆÊ�Ò � �ÁÆ℄<br />

�ÁÆ<br />

�ÁÆ℄<br />

ËÆÊÓÙØ � ¬ ��ÁÆ℄<br />

¬<br />

¬ � ¬<br />

��ÇÍÌ ¬ �ÓÙØ ℄<br />

��ÇÍÌ ℄¬<br />

��ÁÆ℄<br />

�ÓÙØ<br />

¬<br />

¬<br />

¬ �ÁÆ℄<br />

In the ËÆÊÓÙØ definition, the output signal is projected back to the input.<br />

(2.36)<br />

(2.37)<br />

(2.38)<br />

A further interpretation of the DQE concept can be given if we consider a detector work-<br />

ing as co<strong>un</strong>ter, i.e. when input and output are expressed as events number. In the Poisson<br />

<strong>di</strong>stribution vali<strong>di</strong>ty con<strong>di</strong>tions, it is valid:<br />

and<br />

ËÆÊ ÓÙØ �<br />

¬<br />

�ÁÆ℄<br />

¬ ��ÁÆ℄<br />

¬ �ÓÙØ ¬ ��ÁÆ℄<br />

��ÇÍÌ ℄<br />

� �Ò ��ÁÆ℄ (2.39)<br />

ËÆÊ �Ò � �ÁÆ℄<br />

�ÁÆ℄<br />

¬<br />

¬ ��ÁÆ℄<br />

��ÇÍÌ ℄<br />

��ÁÆ℄ (2.40)<br />

��Ò¬ ��ÁÆ ℄ (2.41)<br />

¬ �ÓÙØ where ËÆÊ ÓÙØ has the <strong>di</strong>mensions of a quanta flux. In this case the DQE can be written as:<br />

�É� � ËÆÊ ÓÙØ<br />

ËÆÊ �Ò<br />

��Ò � ¬ ��ÁÆ℄<br />

¬<br />

��ÇÍÌ ℄¬<br />

�ÓÙØ 28<br />

(2.42)


2.4.1 DQE o<strong>per</strong>ating definition.<br />

This definition can be extended to take into acco<strong>un</strong>t the spatial modulation of the input signals if<br />

we consider the spatial frequency dependence. From this point of view, the DQE is the transfer<br />

coefficient, for each spatial frequency, of the signal to noise ratio through the system. It can be<br />

compared to the MTF, which describes the modulation transfer. In order to obtain an efficient<br />

o<strong>per</strong>ating definition of the DQE, it is necessary to rely on the linearity pro<strong>per</strong>ty of the system<br />

<strong>un</strong>der investigation (see Section 2.1). Let us define the input two <strong>di</strong>mensional signal � Ü� Ý (in<br />

this case a dose <strong>di</strong>stribution) and the output of the system � Ü� Ý (which is the dose <strong>di</strong>stribution<br />

map obtained applying the inverse sensitometric curve to the image data). The correspon<strong>di</strong>ng<br />

input and output fluctuations are ¡� Ü� Ý and ¡� Ü� Ý and their respective power spectra<br />

Ï ¡� Ù� Ú and Ï ¡� Ù� Ú . The power spectrum of an ideal system, which does not generate an<br />

intrinsic noise, is related to the input power spectrum through the MTF:<br />

Ï ¡� Ù� Ú �ÅÌ� Ù� Ú ¡ Ï ¡� Ù� Ú (2.43)<br />

The DQE definition based on the average signals:<br />

¬<br />

¬<br />

¬<br />

�� �<br />

��ÇÍÌ ℄<br />

��ÁÆ℄ ¬ ��Ò �ÓÙØ can be extended so to obtain:<br />

�É� Ù� Ú � ÅÌ� Ù� Ú Ï ¡� Ù� Ú<br />

Ï ¡� Ù� Ú<br />

(2.44)<br />

(2.45)<br />

where the generic average signal has been replaced by a monochromatic wave, with spatial<br />

frequencies Ù� Ú . The input output relation is ruled by the ÅÌ� Ù� Ú coefficient, while the<br />

variance is given by the Ù� Ú component of the correspon<strong>di</strong>ng power spectra.<br />

The whole analysis is consistent since, if we introduce a signal of amplitude � Ù� Ú ,we<br />

obtain once more that the DQE is the transfer f<strong>un</strong>ction of the signal to noise ratio:<br />

ËÆÊ ÓÙØ �<br />

ËÆÊ �Ò �<br />

ÅÌ� Ù� Ú � Ù� Ú<br />

Ï ¡� Ù� Ú<br />

� Ù� Ú<br />

Ï ¡� Ù� Ú<br />

�É� Ù� Ú � ËÆÊ ÓÙØ<br />

ËÆÊ �Ò<br />

� ÅÌ� Ù� Ú Ï ¡� Ù� Ú<br />

Ï ¡� Ù� Ú<br />

29<br />

(2.46)<br />

(2.47)<br />

(2.48)


The ex<strong>per</strong>imental data will be analyzed following Eq. 2.45 , as described in Chapter 4.<br />

2.5 NEQ<br />

The quantities described in the previous section lead to the definition of the Noise Equivalent<br />

Quanta, frequently used to evaluate the <strong>per</strong>formances of imaging systems.<br />

Since an ideal detector has the same SNR in the output as in the input, the quantity �ÁÆ ℄<br />

in Eq. 2.41 represents the flux that makes the variance of the output equal to the one of the<br />

real system in an ideal detector. The ideal system in fact works with a bigger flux in input, and<br />

consequently with a better input signal. It can be also written:<br />

�ÁÆ ℄�� �Ò<br />

(2.49)<br />

with �ÁÆ ℄ � �ÁÆ℄ due to the second, a<strong>di</strong>mensional factor smaller than one appearing in<br />

Eq. 2.41. The quantity �ÁÆ ℄ is thus the signal that makes the ideal system equivalent to the<br />

real system from a noise point of view, and it is the so-called Noise Equivalent Quanta NEQ.<br />

From Eq. 2.42 it is straightforward to derive the relation between NEQ and DQE:<br />

�É� � Æ�É<br />

�ÁÆ℄<br />

(2.50)<br />

The physical meaning of the NEQ can be also described from a slightly <strong>di</strong>fferent point of<br />

view. The detector is exposed to a flux �ÁÆ℄ and a ËÆÊÓÙØ is obtained. If the detector were<br />

ideal, the same ËÆÊÓÙØ would be obtained with a flux given by NEQ (� �ÁÆ℄); the ratio of the<br />

two flux gives an o<strong>per</strong>ating definition of the detector efficiency.<br />

30


Chapter 3<br />

Ex<strong>per</strong>imental setup<br />

The entire set of measurements <strong>per</strong>formed in order to evaluate the physical <strong>per</strong>formances of a<br />

complete system of CR mammography, based on a FUJI computed ra<strong>di</strong>ography system, have<br />

been taken at the CSPO (“Centro Stu<strong>di</strong> Prevenzione Oncologica”) in Florence (Italy) with the<br />

su<strong>per</strong>vision of the Physical Health Staff of the “Azienda Ospedaliera Careggi”.<br />

In this chapter the characteristics of each hardware element will be described, with particular<br />

emphasis devoted to the CR <strong>di</strong>gital acquisition system FUJI FCR 5000 MA and to the X-ray<br />

source, based on a mammography facility manufactured by Instrumentarium, in daily clinical<br />

use at CSPO.<br />

3.1 The mammography <strong>un</strong>it<br />

The mammographic <strong>un</strong>it used as X-ray equipment in the present work is an Instrumentarium<br />

model Diamond, with a molybdenum anode and an internal molybdenum filter 25 �m thick (a<br />

25 �m rho<strong>di</strong>um filter and a 500 �m aluminum filter are also available but have not been used in<br />

this investigation). The exit window isa1mmthick beryllium one. Great care has been taken to<br />

<strong>per</strong>form all the measurements in a clinical <strong>di</strong>agnostic-like environment. The X-ray tube voltage<br />

was set to the nominal 28.0 kVp, but a set of measurements have been <strong>per</strong>formed to check the<br />

precision of the control <strong>un</strong>it. A comparison of the value set on the <strong>un</strong>it control and of the same<br />

as measured by a High Voltage Voltmeter (Victory Model Nero) is shown in Tab. 3.1.<br />

The other parameter which can be adjusted by the o<strong>per</strong>ator is the mAs value, that is strictly<br />

31


Console Value Measured Value<br />

(kVp) (kVp).<br />

24.0 ¦ 0.5 24.0 ¦ 0.1<br />

25.0 ¦ 0.5 25.3 ¦ 0.1<br />

26.0 ¦ 0.5 26.3 ¦ 0.1<br />

27.0 ¦ 0.5 27.2 ¦ 0.1<br />

28.0 ¦ 0.5 28.0 ¦ 0.1<br />

29.0 ¦ 0.5 29.1 ¦ 0.1<br />

30.0 ¦ 0.5 30.2 ¦ 0.1<br />

Table 3.1: Comparison between the mammographic <strong>un</strong>it tube voltage as shown by the console<br />

and the correspon<strong>di</strong>ng values measured by an external Voltmeter.<br />

correlated to the dose impinging on the detector.<br />

3.1.1 Dose vs. mAs Relation<br />

The relation between the mAs value set on the Diamond console and the dose on the plate<br />

has been measured with a calibrated ionization chamber. The ionization chamber Model 10x5-<br />

6M has been manufactured by Radcal Corporation and has a sensible volume of 6 cm . It is<br />

connected to a Radcal Converter Mod.9060 and the dose measured values are read by a Radcal<br />

Monitor Controller Mod.9015. The ex<strong>per</strong>imental setup used to obtain the curve dose vs. mAs<br />

is shown in Fig. 3.1.<br />

The X-ray beam has been attenuated by an externally added 30 mm thick PMMA filter<br />

positioned next to the beam beryllium exit window and without the breast compressor. The<br />

source detector <strong>di</strong>stance SDD <strong>di</strong>stance was 62.5 cm and the ionization chamber was positioned<br />

just above the cassette holder. The measured tube voltage was 27.9 kVp with an <strong>un</strong>certainty<br />

of 0.1 kVp. The whole setup is as close as possible to the clinical con<strong>di</strong>tions and the data<br />

we obtained allow a subsequent calibration of Dose vs. Pixel values of the acquired <strong>di</strong>gital<br />

images. It should be stressed that the characterization of the spatial resolution and the <strong>un</strong>iform<br />

dose exposition have been <strong>per</strong>formed with the cassette above the antiscatter grid. Dose values<br />

32


62.8 cm<br />

source-camera<br />

<strong>di</strong>stance<br />

Ionization<br />

chamber<br />

30 mm PMMA<br />

filter<br />

Figure 3.1: Ex<strong>per</strong>imental setup used to measure the relation between Dose and mAs.<br />

correspon<strong>di</strong>ng to a set of mAs values are shown in Tab. 3.2, where the dose values are the<br />

arithmetic mean of at least three independent measurements. The <strong>un</strong>certainties are calculated<br />

as the maximum deviation from the mean.<br />

Just to cross-check, the exposition values for a restricted set of mAs value have been mea-<br />

sured with the Radcal monitor set on the Roengten scale. The exposition and mAs values are<br />

reported in Tab. 3.3. The same values are plotted su<strong>per</strong>imposed to the dose values (in Gray)<br />

in Fig. 3.2, after being rescaled by a factor equal to 8.73 (for photons in Bragg-Gray cavity<br />

approximation 1mR=8.73 �Gy).<br />

3.1.2 The Half Value Layer<br />

The half value layer has been measured positioning aluminum filters close to the X-ray tube<br />

exit. So the X-ray beam is attenuated by an internal 25 �m molybdenum filter, by the 1 mm<br />

beryllium window and by the external aluminum filters. The tube voltage is 27.9 kVp and the<br />

mAS value is set to 12. A second HVL has been measured with a 40 mm PMMA absorber<br />

positioned before the Al filters. In this case the mAs value is 100 mAs while the tube voltage<br />

in <strong>un</strong>changed. The ex<strong>per</strong>imental setup is the same as the one described in Fig. 3.1. The dose<br />

33


mAs Dose (�Gy) mAs Dose (�Gy)<br />

2.0 11.1 ¦ 0.1 80.0 540.7<br />

4.0 25.4 ¦ 0.1 100.0 675.7<br />

6.0 37.0 ¦ 0.2 125.0 843.3<br />

12.0 80.0 ¦ 0.4 150.0 1013<br />

16.0 106.0 ¦ 0.2 175.0 1183<br />

20.0 132.0 ¦ 0.4 200.0 1354<br />

25.0 166.1 ¦ 0.2 250.0 1696<br />

32.0 215.9 300.0 2032<br />

40.0 267.6 350.0 2375<br />

50.0 335.3 400.0 2713<br />

63.0 425.8<br />

Table 3.2: Measured dose values correspon<strong>di</strong>ng to a set of mAs values set on the Diamond<br />

console.<br />

mAs Exposure (mR)<br />

40.0 30.54 ¦ 0.1<br />

80.0 62.05 ¦ 0.2<br />

100.0 77.53 ¦ 0.2<br />

150.0 116.2 ¦ 0.2<br />

Table 3.3: Measured exposure values correspon<strong>di</strong>ng to a set of mAs values set on the Diamond<br />

console.<br />

values measured in both cases are shown in Tab. 3.4 and Tab. 3.5 respectively, and plotted in<br />

Fig. 3.3.<br />

The HVL is 0.276 mm of Aluminum without PMMA, and 0.55 mm Al with the 40 mm<br />

PMMA filter.<br />

34


Figure 3.2: Dose vs. X-ray tube mAs values. Black triangles are ex<strong>per</strong>imental data. The<br />

exposition values measured as cross-check of the f<strong>un</strong>ctionality of the ionization chamber are<br />

su<strong>per</strong>imposed (blue squares).<br />

3.2 Determination of the X-ray spectrum<br />

The X-ray spectrum is obtained using a computer model completely based on measured mam-<br />

mography x-ray spectra. The technique does not require any physical assumption concerning<br />

x-ray, simplifying the ex<strong>per</strong>imental determination of this quantity. The software routine, de-<br />

veloped in IDL framework, allows the user to calculate the realistic polyenergetic spectra at<br />

any voltage between 18 and 40 kV for a molybdenum anode. Ad<strong>di</strong>tional spectral shaping by<br />

elemental filters such as molybdenum, which is routine in mammography, as well ad<strong>di</strong>tional ex-<br />

ternal filters such as PMMA phantoms, can be applied to the raw spectra produced by the model<br />

using the energy dependent Lambert-Beers law with appropriate attenuation coefficients.<br />

Let � ��Πrepresent the photon fluence (photons/mm ) at energy � when a voltage Πis ap-<br />

plied to the x-ray tube. At each energy “bin” (0.5 keV intervals are used), a polynomial f<strong>un</strong>ction<br />

35


Al thickness Dose<br />

(mm) (�Gy)<br />

0.0 1760<br />

0.1 1314<br />

0.2 1031<br />

0.3 834.7<br />

0.4 680.8<br />

0.5 551.0<br />

1.0 241.3<br />

Table 3.4: Dose attenuation values of the beam after the internal Mo and Be filter. Aluminum<br />

foils are externally added.<br />

Al thickness Dose<br />

(mm) (�Gy)<br />

0.0 319.3<br />

0.1 281.3<br />

0.2 248.6<br />

0.3 219.5<br />

0.4 194.6<br />

0.5 169.2<br />

0.6 150.5<br />

0.7 133.3<br />

1.0 93.6<br />

Table 3.5: Dose attenuation values of the beam after the internal Mo and Be filter and 40 mm<br />

PMMA external filter. Aluminum foils are externally added.<br />

was defined as in Eq. 3.1.<br />

¨ ��Î �� ��℄ � ��℄Î � ��℄Î � ��℄Î (3.1)<br />

36


(a) (b)<br />

Figure 3.3: (a) Attenuation curve with Aluminum filters. (b) Attenuation curve with an added<br />

40 mm PMMA filter.<br />

where ����℄ define the polynomial coefficients and are tabulated in literature [15].<br />

The mammographical system <strong>un</strong>der investigation had an ad<strong>di</strong>tional 25 �Ñ molybdenum<br />

internal filter and a 1mm thick beryllium window. Finally, the clinical setup was simulated<br />

ad<strong>di</strong>ng an ad<strong>di</strong>tional 40 mm PMMA (poly-methil-metacrylate) filter near the focal spot of the x-<br />

ray tube. So the spectrum obtained for the molybdenum target without filter has been corrected<br />

following the attenuation law in Eq. 3.2:<br />

¨��ÐØ ��Î �� �� � ¡�¡��ÐØ ¨ ��Î (3.2)<br />

where �� � is the energy dependent mass attenuation coefficient, � is the filter material<br />

density, and ��ÐØ is the thickness of the filter interposed. ¨��ÐØ ��Î is the photon fluence of<br />

the filtered spectrum.<br />

All spectra have been calculated considering V=27.9 kVp. The computed spectrum after<br />

the Mo anode is shown in Fig. 3.4 , while the spectra filtered with the internal 25 �m Mo filter<br />

and Be 1 mm window, and with an externally added 40 mm PMMA, are shown in Fig. 3.4<br />

37


(a) (b)<br />

Figure 3.4: X-ray spectrum computed from ex<strong>per</strong>imental data as described in [15]. (a) No<br />

filtration, 25 �m Mo filter and the final 40 mm PMMA filtered spectrum are considered. (b)<br />

The internal 25 �m Mo filter, a 1 mm Be window and an externally added 40 mm PMMA are<br />

considered.<br />

The spectra obtained agree with the curves fo<strong>un</strong>d in literature. An independent way to test the<br />

vali<strong>di</strong>ty of the model used to calculate the spectra is based on the measurement of the half-value<br />

layers (HVLs) for aluminum (see also [18] for another numerical approach).<br />

3.3 The FUJI Computed Ra<strong>di</strong>ography workstation FCR 5000<br />

MA<br />

The FCR 5000 MA system is expressly de<strong>di</strong>cated to mammography applications. It is capable<br />

of rea<strong>di</strong>ng 50 �m pixel size imaging plates because it employs a dual light collection rea<strong>di</strong>ng<br />

technology to improve the S/N ratio. Furthermore it incorporates high-density rea<strong>di</strong>ng and high<br />

speed data processing. The dual light collection IP rea<strong>di</strong>ng is expected to give a deep contribu-<br />

tion to the improvement of <strong>di</strong>agnostic <strong>per</strong>formance of mammography, in particular with regard<br />

to the <strong>di</strong>agnosis of the shapes of micro-calcifications [10]. The FCR 5000 MA is composed of<br />

38


a cassette reader and a touch screen <strong>di</strong>splay for easy adjust of the readout process parameters<br />

and o<strong>per</strong>ation selection (see Fig. 3.5).<br />

Figure 3.5: The FUJI FCR5000.<br />

The IP handling <strong>un</strong>it furnishes the cassette with shock absorbing clothes to prevent scratch-<br />

ing of the IP. In ad<strong>di</strong>tion to IP cleaning brush rollers positioned on both sides of the IP, it has<br />

also cleaning guides to prevent <strong>di</strong>rtying of the plate and minimize the possibility of image im-<br />

pairment by dust. In the configuration used during this work, the image reader was connected<br />

to an external controller (based on a <strong>per</strong>sonal computer) equipped with an image processor.<br />

The controller was connected also to a film printer and to a graphical workstation by means of<br />

a network system. The graphical workstation was based on a further PC, de<strong>di</strong>cated to image<br />

storage, and a pair of high definition clinical <strong>di</strong>agnosis monitors manufactured by Barco. The<br />

controller has further features, regar<strong>di</strong>ng the image ID registrations and the image acquisition<br />

and processing setup.<br />

The hardware <strong>di</strong>mensions of the FCR5000 MA are 730 mm width, 700 mm depth and 1565<br />

mm height, for a total weight of 280 Kg. The feed/load time specified for the high resolution<br />

39


Figure 3.6: The graphical workstation equipped with a pair of Barco high resolution monitors<br />

and, on the right, the controller.<br />

18cm¢24cm cassette is about 72 sec, correspon<strong>di</strong>ng to a processing capability of about 48<br />

IPs/hour. This rea<strong>di</strong>ng rate is one of the items that need to be improved in order to provide good<br />

clinical use <strong>per</strong>formances, where higher readout rates are desired.<br />

3.4 IP cassette and photostimulable storage plate<br />

The cassette hol<strong>di</strong>ng the imaging plate has been designed to cope with the specifications of a<br />

dual light collection mammographic system. The IP is not easily accessible from the exterior<br />

during normal o<strong>per</strong>ation and the body is a tight light enclosure. Anyway a fully-openable<br />

design is employed to facilitate interior cleaning. For the protection of both IP surfaces, shock<br />

absorbing clothes are affixed on the back surface as well as the front surface. The cassette is<br />

available in two sizes ( � ¢ �cm and � ¢ cm ), with the same <strong>di</strong>mensional standards of<br />

the conventional film plate cassettes, so to allow their use in pre-existing X-ray <strong>un</strong>its. A picture<br />

of the IP cassette used in this work is shown in Fig. 3.7.<br />

A bar-code label is attached to the cassette body for IP identification purposes. When used<br />

with an external controller, the whole system allows a bar code identification procedures that<br />

prevents the user from possible mistakes and makes easier the input of the patient parameters<br />

and the association of the image information with the readout data.<br />

The photostimulable storage plate is comprised of a barium fluorobromide/io<strong>di</strong>de (BaFBr ���I � �)<br />

compo<strong>un</strong>d with an activator (Europium) [16]. The high resolution dual side (HR-BD) rea<strong>di</strong>ng<br />

plate thickness is about 30 % greater with respect to the single side rea<strong>di</strong>ng plate (� �m),<br />

40


Figure 3.7: The IP cassette.<br />

both manufactured by FUJI for computed ra<strong>di</strong>ography. This in order to increase the amo<strong>un</strong>t<br />

of X-ray absorption and to reduce the photon noise. In general, increasing the thickness of the<br />

phosphor layer reduces sharpness. However the FUJI IP compensates this effect by minimizing<br />

the phosphor grain size, which enhances sharpness [10]. A comparison between the single side<br />

rea<strong>di</strong>ng and dual side rea<strong>di</strong>ng IP layer structure is shown in Fig. 3.8.<br />

Figure 3.8: The IP layer structure for the FUJI High Resolution single and dual side rea<strong>di</strong>ng<br />

plates.<br />

For all the measurements done, described in Chapter 4, the same IP has been used. Never-<br />

41


theless a set of cassettes has been exposed to the same dose and the results have been compared<br />

in order to be sure of the <strong>un</strong>iformity of the response over a larger set of IPs. Prior of every ex-<br />

position, the IP is inserted in the cassette reader and submitted to a complete cancellation cycle,<br />

so to minimize possible ghosts effects. Great care has been put in rea<strong>di</strong>ng the IP with the same<br />

delay (approximately 30 seconds) after each exposure session so to reduce the decay effect of<br />

the stored information.<br />

3.5 Image readout process<br />

The IP readout process starts when the cassette is inserted in the FCR5000 MA reader. The IP<br />

is internally extracted from the cassette box. A laser beam is focused on the sensitive part of the<br />

IP and scan the detector surface in a <strong>di</strong>rection parallel to the longer side (scan <strong>di</strong>rection). When<br />

the first row has been read, the plate moves a step correspon<strong>di</strong>ng to a pixel size in a <strong>di</strong>rection<br />

parallel to the shorter side (sub-scan <strong>di</strong>rection) so to start a new row acquisition. The laser<br />

source is a semiconductor device with a spectral emission centered on 660 nm and a power of<br />

60 mW. The light beam is separated in two components by a beam splitter just outside the laser<br />

output window. The less intense components is <strong>di</strong>rected onto a photo-detector that provides<br />

the reference signal necessary to deal with possible intensity fluctuations during the scanning<br />

o<strong>per</strong>ation. This is a key feature since the IP response is a f<strong>un</strong>ction of the laser intensity arriving<br />

on the photostimulable phosphor layer. The main beam is <strong>di</strong>rected towards an optical system<br />

(polygonal mirror, collimating lens, deflection mirror) that has the main goal to steer the bean<br />

on the desired portion of the IP and to maintain an <strong>un</strong>iform speed and focus on the plate. The<br />

laser spot size is about 50 �m, correspon<strong>di</strong>ng to the pixel size <strong>di</strong>mensions. The scan speed has<br />

an up<strong>per</strong> limit set by the phosphor decay time constant, equal to ���s. This poses a minimum<br />

limit to the time required for rea<strong>di</strong>ng an IP with 50 �m pixel size and a 18cm¢24cm size<br />

(3540 ¢4740 pixels), that is about 12 sec. The total read-out time is in any case dominated<br />

by the feed-load-eject mechanical procedure. When the laser reaches the end of the scan line,<br />

it is repositioned to the beginning of a new line by the polygonal mirror. In the meantime the<br />

plate moves in an orthogonal <strong>di</strong>rection with a 50 �m step. This process ends when the entire<br />

detector surface has been exposed to the laser light. The light emitted is proportional to the PSL<br />

42


centers activated in the illuminated area and thus to the dose exposure on the plate, with a linear<br />

dependence. The laser power determines the ratio of the information released by the F-centers<br />

and consequently, the rea<strong>di</strong>ng time and the remaining <strong>un</strong>used information.<br />

The ra<strong>di</strong>ation emitted by the F-center from both sides of the IP is isotropic. In order to<br />

collect most of the emitted ra<strong>di</strong>ation from the up<strong>per</strong> part of the plate, a light guide receives the<br />

<strong>di</strong>rect light and the light deflected by a mirror that has the goal of cover most of the solid angle<br />

as possible. The lower side of the IP is coupled to a second light guide and a second PMT. The<br />

signal from the two sides, pro<strong>per</strong>ly weighted, are added together. The photostimulated emis-<br />

sion is centered at about 400 nm, a wavelength which <strong>di</strong>ffers deeply from the laser stimulating<br />

ra<strong>di</strong>ation, so to avoid interference between the stimulating and detection process. In fact the<br />

photomultiplier tubes have their sensitivity peak at 400 nm.<br />

An analog amplifier stage that applies a logarithmic conversion follows the PMT. This is<br />

necessary to recover the linearity between the photostimulated light and the electrical signal<br />

which had been lost in the PMT photo-electron amplification process [17]. Then the analog to<br />

<strong>di</strong>gital conversion takes place.<br />

The system <strong>un</strong>der investigation <strong>di</strong>gitize the total X-ray induced signal over a range of in-<br />

cident exposures four orders of magnitude wide (from 0.01 mR to 100 mR) keeping a linear<br />

relation between exposure dose and ADC co<strong>un</strong>ts. The 12 bit ADC produces 11 bits of effec-<br />

tive quantization levels prior to normalization of the image 10 bit depth. The output data for a<br />

typical high resolution image used during this analysis are stored in 32 MBytes files.<br />

In order to <strong>per</strong>form a complete cancellation of the information from the IP after the rea<strong>di</strong>ng<br />

process, the plate is exposed to an intense light that removes the remaining image causing the<br />

decay of the metastable sites still present. Due to the dual side rea<strong>di</strong>ng technology, the erasure<br />

lamps act on both the surfaces of the imaging plate.<br />

3.6 Histogram analysis and IP sensitometric curve<br />

The <strong>di</strong>gital output represents the pre-processed data. Histogram analysis is applied to the pre-<br />

processed data to define the wanted versus <strong>un</strong>wanted signals in a scanned image plate for a<br />

particular incident exposure and examination type. As the linear exposure latitude for the imag-<br />

43


ing plate is very wide, a variable rea<strong>di</strong>ng sensitivity (sensitivity number S) is necessary to map<br />

the stimulated luminescence of the imaging plate to a range of output <strong>di</strong>gital numbers within a<br />

10 bit range. A second parameter, the latitude L, in<strong>di</strong>cates the range of stimulated luminescence<br />

(minimum to maximum) that will be included in the output <strong>di</strong>gital number range.<br />

The histogram of the <strong>di</strong>gital values for each pixel, correspon<strong>di</strong>ng to the stimulated lumines-<br />

cence, is created in order to <strong>per</strong>form an optimization of the data. The method used to assign<br />

one of the 1024 grey scale level to the <strong>di</strong>gital value of each pixel is based on the analysis of the<br />

histogram having on the x axis the dose value and on the y axis the final pixel value number.<br />

The elements used by the FUJI histogram analysis are shown in Fig. 3.9.<br />

PV<br />

Q<br />

mRem<br />

[X-ray dose]<br />

Figure 3.9: Histogram analysis and main parameters involved.<br />

The FUJY FCR5000MA system introduces a quantity × that is declared to be related to the<br />

44


dose by the relation [17]:<br />

× � ÐÓ� ¡ �Ó×� ÑÊ�Ñ (3.3)<br />

× is the parameter linearly connected to the PV. The × values <strong>di</strong>stribution histogram is shown<br />

in Fig. 3.9, where the number of image pixels having the same × value is reported on the left y<br />

axis.<br />

Two parameters are used in order to determine linear relation between × and the grey scale<br />

values. The sensitivity index S is defined as<br />

Ë � ¡<br />

� ×<br />

(3.4)<br />

where × is the × value correspon<strong>di</strong>ng to the central point of the grey scale (PV=511 in our<br />

case). The latitude index Ä is defined as:<br />

Ä � ×� ×� (3.5)<br />

where ×� is the minimum × value that corresponds to PV=1023 (response saturation at high<br />

dose) and ×� is the maximum × value that corresponds to PV=0 (response saturation at low dose).<br />

The sensitometric curve is completely determined by the parameters S and L, both shifting the<br />

useful range of the curve along the dose values and the latter alone changing the slope of the<br />

straight line. In fact it can be written for the PV - dose relation:<br />

ÈÎ �<br />

Ä<br />

¡ × �<br />

Ä<br />

� ÐÓ�<br />

Ë<br />

(3.6)<br />

The system <strong>un</strong>der investigation provides three user selectable modes that can be applied to<br />

the histogram: automatic, semiautomatic and fixed mode. In the automatic mode the system<br />

adjust, by means of de<strong>di</strong>cated threshold and <strong>di</strong>fferential algorithms, the L and S values so to<br />

select the useful region of the histogram. In fixed mode the user can select both the L and S<br />

parameter. All the image used for the measurement described in this thesis have been acquired<br />

using the Fixed sensitivity mode, so to keep <strong>un</strong>der control all the parameters necessary to the<br />

off-line analysis. Furthermore a linear transformation curve was used between the input <strong>di</strong>gital<br />

number and the output final pixel value. In other words, in fixed sensitivity mode the user<br />

defines the system speed and the latitude to be used for processing the exposed imaging plate.<br />

Thus the response in this mode is similar to a screen-film cassette. The user can change the<br />

image quality by acting n the mAs and kVp values.<br />

45


Chapter 4<br />

Measurement of the physical quantities<br />

contributing to the DQE<br />

In this chapter the image quality characteristics of the CR system <strong>un</strong>der investigation are re-<br />

ported. Physical criteria, such as modulation transfer f<strong>un</strong>ction (MTF), noise power spectrum<br />

(NPS), noise equivalent quanta (NEQ) and detective quantum efficiency (DQE), were employed<br />

for this evaluation. In section 4.1 the relation between pixel values and dose is described. In<br />

the following sections all the quantities needed for the evaluation of the DQE are calculated,<br />

in order to determine the DQE and to compare it to other systems in use at present. The dose<br />

effects on these quantities are also shown, together with their spatial frequency dependence.<br />

Unless otherwise stated, a 50 �m pixel resolution (High Resolution HR) and a fixed linearity<br />

acquisition mode have been employed for all the data acquired for the analysis and measure-<br />

ments described in this work. Each image was stored in a DICOM format on the PC controlling<br />

the acquisition procedure, and later transferred to a more powerful calculator for subsequent<br />

raw data analysis.<br />

A reference system is chosen where the X and Y axes are coincident with the IP sides, with<br />

the former <strong>di</strong>rected in a <strong>di</strong>rection orthogonal to the thorax side of the breast (shorter side, called<br />

also sub-scan <strong>di</strong>rection) , and the latter parallel to the thorax side of the breast (longer side - scan<br />

<strong>di</strong>rection). The origin is in the right corner close to the breast, with increasing values moving<br />

far away from the patient.<br />

46


4.1 Ex<strong>per</strong>imental characteristic curve<br />

The response curve of the imaging system, in terms of PV as a f<strong>un</strong>ction of exposure dose, has<br />

been evaluated by exposing the IP at several dose levels with a fixed tube voltage of 28 kVp.<br />

The ex<strong>per</strong>imental setup is the same as the one described in Fig. 3.1 since we want to use the<br />

mAs vs. the air kerma calibration curve previously measured. The only <strong>di</strong>fference is that the<br />

ionization chamber was replaced by the IP cassette, while the 30 mm thick PMMA external<br />

filter, close to the beam exit window, remained the same. The cassette was placed just above<br />

the cassette holder, and therefore it was exposed without the anti-scatter grid in between.<br />

The IP has been exposed to the <strong>un</strong>iform X-ray beam for a set of dose values, the o<strong>per</strong>ator<br />

being careful to maintain the reproducibility of the process. After the exposure, the IP was<br />

imme<strong>di</strong>ately read, in order to minimize and to keep the temporal decay effect constant for every<br />

measure, and then cancelled. The acquired data were processed in the linear fix mode, allowed<br />

by the FUJI system software, in order to fully control every step of the image processing mode.<br />

The system was o<strong>per</strong>ating in High Resolution mode, with a pixel size of 50 �m. The sensitivity<br />

S was set to 200 and the latitude L to 4.0, correspon<strong>di</strong>ng to a linear processing curve over the<br />

entire dynamic range of the system. Each DICOM image was read by an IDL software code in<br />

order to access the raw data. The IP is exposed so that its shorter side is parallel to the <strong>di</strong>rection<br />

where the heel effect is bigger. In Fig. 4.1 a row parallel to the <strong>di</strong>rection where the heel effect is<br />

present is shown. In this case the IP was exposed to a <strong>un</strong>iform beam with a dose of about 676<br />

�Gy.<br />

In order to analyze a <strong>un</strong>iformly exposed region (ROI) of the image, and to be allowed<br />

to forget the heel effect, a strip 400 pixels wide (2 cm) and 2000 pixels height (10 cm) was<br />

selected. Furthermore the ROI was centered aro<strong>un</strong>d a point <strong>di</strong>stant 4 cm from the thorax side<br />

laterally centered. The ionization chamber was positioned in the same point during dose-mAs<br />

calibration measurements. The selected area is shown in black in the typical <strong>un</strong>iformly exposed<br />

image reported in Fig. 4.2.<br />

For each dose, the mean PV has been computed as the average value of the PVs of the ROI<br />

previously described. The correspon<strong>di</strong>ng standard deviation SD has been calculated on the ROI<br />

after a plane subtraction, in order to eliminate the first order contribution. If we consider the<br />

plane P that best fits the ROI <strong>un</strong>der investigation and we address each ROI pixel by the indexes<br />

47


Figure 4.1: Section along the shorter side of a <strong>un</strong>iformly exposed IP. The heel effect in this<br />

<strong>di</strong>rection is clear.<br />

Thorax side<br />

Lateral side<br />

Figure 4.2: Uniformly exposed IP. The ROI used for PV-dose calibration is shown in black. The<br />

left longer side is the one closer to the breast.<br />

�� � , the following equations hold:<br />

ÈÎ � Æ ¡ Å<br />

�<br />

��<br />

�<br />

��<br />

ÈÎ���<br />

(4.1)<br />

ÊÇÁÐ�Ò �� � �ÊÇÁ �� � È �� � (4.2)<br />

where N is the number of pixels (400 in this case) along the X <strong>di</strong>rection and M (2000 in this<br />

case) the number of pixels along the Y <strong>di</strong>rection . SD is the standard deviation of ROIÐ�Ò, with<br />

a value decreasing from 5 PV to 1 PV with increasing dose.<br />

48


The PV curve as a f<strong>un</strong>ction of dose together with the least square fit <strong>per</strong>formed with a<br />

f<strong>un</strong>ction f(x) of the type:<br />

� Ü �� ¡ ÐÓ� � ¡ Ü � (4.3)<br />

is shown in Fig. 4.3 (see also 3.6).<br />

Figure 4.3: Pixel Value vs. Dose (�Gy).<br />

The A parameter is connected to the latitude L, since Ä � �� holds. From the fit it<br />

appears that L= ��� ¦ � , in <strong>per</strong>fect agreement with the pre<strong>di</strong>ctions. In Fig. 4.4 the same data<br />

set is shown, with the PV as a f<strong>un</strong>ction of log (Dose), with the goal to stress the linearity of<br />

the system over three orders of magnitude.<br />

From the characteristic curve, it is possible to linearize the PV vs. dose relation and con-<br />

sequently to take benefit from the pro<strong>per</strong>ties of linear systems in the analysis as described in<br />

chapter 2. In this case the linearized PV (hereafter PVÐ�Ò) are calculated as:<br />

ÈÎÐ�Ò � à ¡<br />

ÈÎ<br />

� (4.4)<br />

where K is a normalization parameter non f<strong>un</strong>damental in the subsequent analysis.<br />

49


Figure 4.4: Pixel Value plotted as a f<strong>un</strong>ction of the Log Dose (�Gy).<br />

4.2 Pre-sampling MTF<br />

The overall two <strong>di</strong>mensional MTF of a <strong>di</strong>gital system as described in Chapter 1 can be expressed<br />

by [19]:<br />

ÅÌ� Ù� Ú � ��ÅÌ�� Ù� Ú ¡ ÅÌ�Ë Ù� Ú ℄ £<br />

�<br />

�<br />

�<br />

Ò�<br />

Æ Ù Ñ�¡Ü� Ú Ò¡Ý �<br />

¡ÅÌ�� Ù� Ú ¡ ÅÌ�� Ù� Ú (4.5)<br />

where Ù� Ú are the spatial frequencies and £ denotes the convolution o<strong>per</strong>ator. MTF�(u,v)<br />

,MTFË(u,v), MTF� (u,v) and MTF�(u,v) are the MTFs of analog input, sampling a<strong>per</strong>ture, filter<br />

and <strong>di</strong>splay a<strong>per</strong>ture respectively. The Ñ and Ò are integers, and ¡Ü and ¡Ý are the sampling<br />

<strong>di</strong>stances in the Ü and Ý <strong>di</strong>rections.<br />

The product of MTF�(u,v) and MTFË(u,v) is referred to as pre-sampling modulation trans-<br />

fer f<strong>un</strong>ction MTFÈÊË(u,v) of a <strong>di</strong>gital imaging system [20], which includes the geometric <strong>un</strong>-<br />

sharpness, if not negligible, the detector <strong>un</strong>sharpness, and the sampling a<strong>per</strong>ture <strong>un</strong>sharpness.<br />

The <strong>di</strong>gital MTF (MTF�Á�) of the system can be obtained by the convolution of MTFÈÊË(u,v)<br />

with the comb f<strong>un</strong>ction in the frequency domain. In this work the MTF due to the filter and the<br />

50


<strong>di</strong>splay will be neglected since they are dependent on the o<strong>per</strong>ator choice and hardware setup,<br />

and are in any case easily controlled. It should be noted that the MTF�Á� and the overall MTF<br />

may incorrectly in<strong>di</strong>cate the resolution capability of a <strong>di</strong>gital system because both MTFs might<br />

include a false response due to aliasing, while MTFÈÊ� can characterize inherent resolution<br />

pro<strong>per</strong>ties of a <strong>di</strong>gital imaging system. The effects of aliasing will be described in the next<br />

section, together with the problems deriving from this lack of spatial invariance.<br />

In this work MTFÈÊË is measured by the Fourier transform of a “finely sampled” Line<br />

Spread F<strong>un</strong>ction (LSF), which is obtained with a slit slightly tilted with respect to the IP pixel<br />

grid. The sampling <strong>di</strong>stances used for data acquisition were 0.05 mm (20 pixels/mm), with<br />

a standard HR IP having the size of 18x24 cm (3540x4740 pixels). The signal from each<br />

pixel was <strong>di</strong>gitized with a final 10 bit grey scale. The exposure level and the image acquisition<br />

parameters were chosen in order to provide a good quality slit profile, spanning as much as<br />

possible the ADC range. In the examined cases, the LSF images were acquired with a ¦ �m<br />

wide, �� ¦ � mm long slit camera manufactured by Nuclear Associates, model 07-624 (see<br />

Fig 4.5).<br />

Figure 4.5: The slit used for the LSF acquisition.<br />

The camera was positioned by means of a custom, angle graduated, lead holder , which has<br />

the double f<strong>un</strong>ction of shiel<strong>di</strong>ng the IP from scattered ra<strong>di</strong>ation and of allowing the repeatability<br />

of the measurement and of the slit orientation (see Fig 4.6). The slit was positioned on the<br />

cassette, orthogonally to the central axis the X-ray beam; the ex<strong>per</strong>imental setup is analogue to<br />

the one used for the <strong>un</strong>iform dose expositions; in particular the SDD was 62.5cm, the dose was<br />

about 680 �Gy, the latitude parameter L was 2.5 and the sensitivity S was set to 63. The beam<br />

51


was filtered by a 30 mm block of PMMA, positioned close to the target, in order to approximate<br />

the energy dependent attenuation of breast in a scatter reduced geometry. The <strong>di</strong>stance of the<br />

slit from the IP has been taken into acco<strong>un</strong>t in all the subsequent analysis.<br />

Figure 4.6: The setup used for the LSF acquisition.<br />

The LSF was measured both in the scan and in the subscan <strong>di</strong>rection. In Fig. 4.7 the two<br />

<strong>di</strong>mensional image of the slit is shown. The three <strong>di</strong>mensional profile of the same image is<br />

shown in Fig. 4.8, with the PV on the z axes. The off line linearization process has not been<br />

<strong>per</strong>formed yet.<br />

The slight angle between the slit and the scan or subscan <strong>di</strong>rection allows us to obtain a LSF<br />

more finely sampled than the one that would be reconstructed from a slit parallel to the x and<br />

y axes; in fact, in this last case the sampling <strong>di</strong>stance would be given by the pixel size. The<br />

method to obtain a smaller effective sampling <strong>di</strong>stance is described below. Assume that a slit is<br />

positioned at a slight angle � (usually � Æ ) in the <strong>di</strong>rection <strong>per</strong>pen<strong>di</strong>cular to the scanning (sub-<br />

scanning) <strong>di</strong>rection in deriving the MTF× �Ò (MTF×Ù�× �Ò). In Fig. 4.9 the schematic <strong>di</strong>agram<br />

showing this situation is reported, with the IP pixel grid and the slit projection su<strong>per</strong>imposed.<br />

Because of the slight angulation of the slit, a series of <strong>di</strong>fferent LSFs can be extracted with<br />

respect to the alignment of the slit relative to the sampling coor<strong>di</strong>nate (in the <strong>di</strong>agram, four<br />

<strong>di</strong>fferent LSFs su<strong>per</strong>position of the slit on the pixel grid are shown, since line A is equal to<br />

line E). The LSF of each row is sampled at a <strong>di</strong>stance ¡Ü. From the combined data of each<br />

52


Figure 4.7: Image of the slit used for the LSF and MTF calculation.<br />

(a) (b)<br />

Figure 4.8: Typical slit profile. The PV (z axes) are represented as a f<strong>un</strong>ction of the x-y IP<br />

coor<strong>di</strong>nate.<br />

row, each representing a LSF sampled at a slightly <strong>di</strong>fferent point (<strong>di</strong>fferent phase), a composite<br />

finely sampled LSF can be generated with a smaller effective sampling <strong>di</strong>stance. The number of<br />

points in the LSF is chosen large enough to obtain a sufficient number of values, in the frequency<br />

53


17 13 9 5 1<br />

18 14 10 6 2<br />

19 15 11 7 3<br />

Pixel Grid<br />

Slit Image<br />

20 16 12 8 4<br />

13 9<br />

17 14<br />

10<br />

5 1<br />

18<br />

15<br />

19<br />

16<br />

11<br />

12<br />

6<br />

7<br />

8<br />

3<br />

2<br />

20<br />

4<br />

Figure 4.9: Schematic <strong>di</strong>agram showing the generation of a composite finely sampled LSF. The<br />

LSFs correspon<strong>di</strong>ng to the various alignments of the slit relative to the sampling coor<strong>di</strong>nate are<br />

composed, after a spatial shift, to give the final LSF.<br />

domain, to reconstruct the MTF finely along all the interesting range of spatial frequencies.<br />

So for every slit image it is necessary to evaluate when the LSFs on each rows reproduce<br />

the same pattern. In this work a numerical approach has been chosen but similar results can<br />

be obtained by graphical methods. As it can be inferred from Fig. 4.9, in the hypothesis of<br />

constant slit width and of <strong>un</strong>iform exposition, the maximum <strong>di</strong>gital value M1 of all the rows is<br />

obtained when the slit is just above a pixel center since the light is not <strong>di</strong>stributed between two<br />

adjacent pixels. We select this row as reference line for the angle calculation. Moving from<br />

this point along a <strong>di</strong>rection orthogonal to the slit, the maximum value of each row is lower than<br />

M, reaching a minimum when the slit center is above the line separating two pixels, and then<br />

increases <strong>un</strong>til a new maximum M2 is reached. The <strong>di</strong>stribution of maxima as a f<strong>un</strong>ction of the<br />

column number allows to retrieve the slit angle value. A more efficient method has emerged to<br />

be the search, in each row, of the column number correspon<strong>di</strong>ng to the maximum. The number<br />

Ò of rows having the maximum in the same column is the number of lines that contribute to the<br />

finely sampled LSF without reproducing the same pattern. Furthermore, more than one LSF<br />

can be obtained from a slit image since this method allows an easy reconstruction of the start<br />

and end point of each composite LSF. The slit angle �, i.e. the angle between the slit and the<br />

54


vertical <strong>di</strong>rection in Fig. 4.9, is given by:<br />

� � �Ö Ø�Ò �Ò (4.6)<br />

while the effective sampling <strong>di</strong>stance ¡Ü is:<br />

¡Ü �¡Ü�Ò �¡Ü ¡ Ø�Ò � (4.7)<br />

In the slit image used for the LSF reconstruction the angle was � ¦ � Æ for the scan<br />

<strong>di</strong>rection and � ¦ � � Æ for the sub-scan <strong>di</strong>rection. To independently check the vali<strong>di</strong>ty of<br />

the analysis, the maximum value of each row has been plotted as a f<strong>un</strong>ction of the row number.<br />

Each PV has been normalized to the area <strong>un</strong>der the slit profile at each row to take into acco<strong>un</strong>t<br />

possible non-<strong>un</strong>iformity effects of the IP and im<strong>per</strong>fections of the slit. The <strong>di</strong>stance from two<br />

local mimina is the number Ò (see Fig. 4.10).<br />

Figure 4.10: Maxima of the PV slit values.<br />

We can observe that the graphical method agrees with the numerical one, even if it is less<br />

efficient in the classification of the critical points.<br />

The angle amplitude is a tradeoff between a small angle, which allows to obtain a very<br />

finely sampled LSF, and a big angle that allows to reconstruct more LSFs from the same image<br />

and to average them reducing possible local effect. In this work the minimum angle is given<br />

55


y the pixel size (� �m) and the slit length useful to the analysis (quantified in 6 mm). In<br />

order to fully reconstruct a finely sampled LSF from a single acquisition �Ñ�Ò is about �� Æ ,<br />

but in the analysis an angle between Æ and Æ has always been used. The useful (to the LSF<br />

reconstruction) column width is chosen so to obtain long enough tails without tr<strong>un</strong>cating before<br />

reaching the plateau value.<br />

At this point the IP characteristic curves, relating the PV to the relative exposure, were used<br />

for the linearization. The signal can now be expressed in terms of exposure levels instead of<br />

grey levels. Each finely sampled LSF, one for each Ò rows, is then reconstructed by aligning<br />

each row value with the appropriate shift (a ¡Ü integer multiple). The final LSF is the average<br />

of three LSF. The effect of the finite slit width will be taken into acco<strong>un</strong>t in the MTF calculation.<br />

A typical example of composite LSF is shown in Fig. 4.11; it is clear the effect of fine sampling.<br />

(a) (b)<br />

Figure 4.11: Typical LSF plot. Points deriving from several LSF, extracted from a single im-<br />

age, are su<strong>per</strong>imposed to show the good agreement of the ex<strong>per</strong>imental data. (a) LSF in scan<br />

<strong>di</strong>rection. (b) LSF in subscan <strong>di</strong>rection.<br />

In case of failures in the LSF reconstruction, it has emerged that each error in the choice of<br />

the parameters needed by the reconstruction algorithm has macroscopic consequences on the<br />

LSF quality, so to allow an easy rejection of the case.<br />

56


The reconstructed LSF may include in the tails errors due to the glare and quantization ef-<br />

fects. So an exponential extrapolation was employed for the tails, with a tr<strong>un</strong>cation level of<br />

approximately 0.01 with respect to the maximum of the LSF. In literature ex<strong>per</strong>imental evi-<br />

dences can be fo<strong>un</strong>d that the exponential extrapolation gives better results.<br />

We will see that the tail corrections affect only the low spatial frequencies and that the<br />

overall effect is not significant because the CR is a low noise system. In Fig. 4.12 the LSF with<br />

tail extrapolation is shown.<br />

Figure 4.12: LSF�ÜØÖ.<br />

The mean composite LSF�ÜØÖ is then Fourier transformed to obtain the presampling MTF.<br />

The smaller effective sampling <strong>di</strong>stance allows to have more reconstructed points for the MTFÔÖ�.<br />

The Optical Transfer F<strong>un</strong>ction OTF is calculated as FFT of the LSF. The OTF is then <strong>di</strong>vided<br />

by the factor ×�Ò ��Ï (� is the frequency) to take into acco<strong>un</strong>t the finite slit witdh W. The<br />

MTF is the modulus of the complex f<strong>un</strong>ction OTF.<br />

A comparison between the MTFs obtained from the same slit image, before and after the<br />

tales exponential extrapolation, shows that the two curves <strong>per</strong>fectly agree at the higher frequen-<br />

cies, while the MTF�ÜØÖ has a smoother behavior, as expected, at the lower frequencies (see<br />

Fig. 4.13)<br />

The final MTF× �Ò and MTF×Ù�× �Ò are shown in Fig. 4.14 and Fig. 4.15 .<br />

57


Figure 4.13: Comparison between MTF and MTF�ÜØÖ.<br />

Figure 4.14: MTF× �Ò.<br />

From a comparison between the two curves it emerges that MTF behaves essentially in the<br />

same way for the scan and sub-scan <strong>di</strong>rection, even if they are slightly higher for the sub-scan<br />

slit image. (see Fig. 4.16).<br />

58


Figure 4.15: MTF×Ù�× �Ò.<br />

Figure 4.16: MTF× �Ò and MTF×Ù�× �Ò.<br />

These are the curves that will be used for the subsequent analysis.<br />

59


4.3 EMTF<br />

The phase dependence of ÅÌ���� resulting from <strong>un</strong>dersampling, described in Section 2.2.1,<br />

poses a conceptual problem in that it violates the desired stationarity pro<strong>per</strong>ty of MTF. The<br />

solution adopted in this work is to consider the expectation value of ÅÌ���� averaged over<br />

all phases, which shall be called the expectation MTF (EMTF). The EMTF seems a suitable<br />

descriptor of <strong>di</strong>gital MTF since it satisfies the requirements of being defined approximately as<br />

�ÇÌ����� and being spatially invariant. EMTF is calculated by averaging MTF��� Ù � � over<br />

all possible phase values. For the image of a slit, the phase factor is just the <strong>di</strong>splacement of the<br />

slit center from the origin of the image since all cosinusoids in a slit image have their origins<br />

at the slit center. Furthermore it can be shown that MTF��� ٠� � is the same if � is shifted<br />

by multiples of one pixel [12]. Therefore, the average over all possible phases needs only to<br />

include the range 0 to �. The relation used to evaluate EMTF is then:<br />

�ÅÌ� Ù�<br />

� �<br />

�ÇÌ���� Ù�� � �<br />

��ÅÌ�Ù�� � �� ��<br />

� �ÇÌ���� � � �<br />

(4.8)<br />

where Ù� is the generic spatial frequency, � the phase and �ÇÌ���� Ù�� � � is calculated as<br />

described in Eq. 2.26.<br />

A software code developed in IDL language <strong>per</strong>forms the described calculations with the<br />

sampling interval <strong>di</strong>vided in 10000 steps (i.e. �� � � pixel size). The EMTF curves are<br />

shown in Fig. 4.17 for the scan (a) and sub-scan <strong>di</strong>rection (b). Su<strong>per</strong>imposed are the MTFÔÖ�<br />

curves.<br />

The EMTF was fo<strong>un</strong>d to be very close to the value of ÅÌ�ÔÖ�, with a <strong>di</strong>fference increasing<br />

with spatial frequencies, as shown in Fig. 4.18, but less than 0.5% at the Nyquist frequency.<br />

Conceptually, EMTF is a reasonable compromise to the problem of phase dependance of<br />

<strong>di</strong>gital MTF. Since MTFÔÖ� is the response of a <strong>di</strong>gital system to a delta f<strong>un</strong>ction (slit in this<br />

case), EMTF is the response of the system to a slit averaged over all locations in an image.<br />

EMTF is therefore a more accurate reproduction of the response of the total system to a slit<br />

than is MTFÔÖ�, although EMTF will not necessarily represent the delta-f<strong>un</strong>ction response at<br />

any given location. In the subsequent evaluations of the system NEQ and DQE, the EMTF will<br />

be used. In Fig 4.19 the two EMTF, in scan and subscan <strong>di</strong>rections, are compared.<br />

60


(a) (b)<br />

Figure 4.17: EMTF and MTFÔÖ� (a) scan <strong>di</strong>rection. (b) subscan <strong>di</strong>rection.<br />

4.4 NPS measurement<br />

The Noise Power Spectrum has been measured starting from <strong>un</strong>iformly exposed images. A large<br />

set of exposure levels has been used. Initially the two-<strong>di</strong>mensional NPS has been evaluated in<br />

order to study the presence or absence of any off-axis noise structure.<br />

Initially the image data was linearized to the exposure using the characteristic transforma-<br />

tion curve. A big, square ROI, centered on the point of clinical interest (4 cm away from the<br />

chest), was selected for the subsequent analysis for each image correspon<strong>di</strong>ng to a <strong>di</strong>fferent<br />

exposure level. The <strong>di</strong>mensions of this ROI are such to contain £ smaller ROIs, with 128<br />

pixel side, each shifted from its neighbor by 64 pixels step [21] . A planar trend is subtracted<br />

from each ROI and the Fast Fourier Transform is <strong>per</strong>formed on the flat field data. The NPS is<br />

then calculated as:<br />

ÆÈËÖ�Û Ù� Ú �<br />

� ��Ì��Ð�Ø���Ð� Ü� Ý ℄� �<br />

ÆÜÆÝ<br />

¡Ü¡Ý<br />

(4.9)<br />

where � ��Ì��Ð�Ø���Ð� Ü� Ý ℄� � represents the ensemble average of the Fast Fourier<br />

transformed 128*128 ROIs. ÆÜ and ÆÝ are the number of elements in the Ü and Ý <strong>di</strong>rections<br />

(which are equal to 128 in this case) and ¡Ü and ¡Ý are the pixel pitch (equal to 50 �m).<br />

61


Figure 4.18: Differences between EMTF and MTF.<br />

The NPS used in subsequent calculations is the normalized NPS (ÆÈËÒÓÖÑ�Ð�Þ��), obtained by<br />

scaling the ÆÈËÖ�Û for the mean signal of the £ ROIs (large area signal).<br />

ÆÈËÒÓÖÑ�Ð�Þ�� Ù� Ú �<br />

ÆÈËÖ�Û<br />

large area signal<br />

(4.10)<br />

The two <strong>di</strong>mensional normalized NPS <strong>di</strong>stribution for an image exposed to ���Gy is<br />

shown in Fig. 4.20.<br />

An evaluation of the two-<strong>di</strong>mensional NPS image reveals the presence of an apo<strong>di</strong>zation<br />

filter that significantly reduces the noise power spectrum above 8 lp/mm in the scan <strong>di</strong>rection.<br />

This filter is likely to be necessary to reduce the high frequency components of the decay lag<br />

characteristics of the photostimulated luminescence that occurs during the readout in the scan<br />

<strong>di</strong>rection. This apo<strong>di</strong>zation filter lowers the NPS response, but has little impact on the NEQ and<br />

DQE because the dominant factor above this cut-off frequency is the MTF response, which is<br />

close to zero.<br />

From the 2D NPS, the NPS along the scan and sub-scan <strong>di</strong>rection has been calculated. The<br />

1D NPS along the scan and sub-scan <strong>di</strong>rection is obtained considering a thick slice of four lines<br />

62


Figure 4.19: EMTF scan Vs. EMTF sub-scan.<br />

of the 2D NPS on either side of both scan and sub-scan axis (exclu<strong>di</strong>ng the axis). For each data<br />

value of coor<strong>di</strong>nates (u,v) in the 2D NPS space, the new 1D frequency value was computed as<br />

Ô Ù Ú . After a rebinning of the frequency axis values, the 1D NPS is obtained [22]. The<br />

assumptions for using this technique for estimating the 1D NPS are that the 2D NPS exhibits<br />

moderate ra<strong>di</strong>al symmetry and that the noise data are basically <strong>un</strong>iform within the small annuli<br />

of spatial frequencies used for regrouping the noise data. The measured NPS in the scan and<br />

sub-scan <strong>di</strong>rection are shown in Fig. 4.21 and Fig. 4.22 respectively for the entire set of exposure<br />

levels used in this analysis.<br />

A comparison between the NPS in the scan and sub-scan <strong>di</strong>rection is <strong>per</strong>formed in Fig. 4.23<br />

4.5 NEQ measurement<br />

If the MTF describes the signal response of a system and the NPS describes the amplitude<br />

variance, both at a given frequency Ù, then the ratio of these two quantities, pro<strong>per</strong>ly normalized,<br />

63


Figure 4.20: 2 <strong>di</strong>mensional NPS The dose values is 37 �Gy.<br />

Figure 4.21: NPS as a f<strong>un</strong>ction of exposure dose values in the scan <strong>di</strong>rection.<br />

64


Figure 4.22: NPS as a f<strong>un</strong>ction of exposure dose values in the sub-scan <strong>di</strong>rection.<br />

Figure 4.23: Comparison of the NPSs in the scan and -subscan <strong>di</strong>rections for a set of exposure<br />

values.<br />

gives information about the maximum available signal-to-noise ratio as a f<strong>un</strong>ction of frequency.<br />

65


The Noise Equivalent Quanta was computed, for each exposure level, as [23] [22]:<br />

Æ�É � �<br />

large area signal �ÅÌ� �<br />

ÆÈË �<br />

�ÅÌ� �<br />

�<br />

ÆÈËÒÓÖÑ�Ð�Þ�� �<br />

(4.11)<br />

where � is the spatial frequency and assuming that the measured large area signal is lin-<br />

early related to the input signal. EMTF is the <strong>di</strong>gital expectation MTF, computed numerically<br />

from the pre-sampling MTFs [12] as described in section 4.3. The large area signal is the av-<br />

erage <strong>di</strong>gital value in <strong>un</strong>its of exposure of each ROI after backgro<strong>un</strong>d trend correction. NPS is<br />

the thick-slice one-<strong>di</strong>mensional NPS near the Ù or Ú axis (see section 4.4) and NPSnormalized is<br />

given by NPS <strong>di</strong>vided the large area signal. The curves describing the NEQ as a f<strong>un</strong>ction of<br />

spatial frequency are shown in Fig. 4.24 and Fig. 4.25 for each exposure level. The interested<br />

frequencies are in the range between 0 mm and the Nyquist frequency 10 mm .<br />

Figure 4.24: NEQ scan.<br />

In Fig. 4.26 the NEQ for the scan and sub-scan <strong>di</strong>rections are compared. The scan NEQ<br />

is always lower than the correspon<strong>di</strong>ng sub-scan curves <strong>un</strong>til the frequency is about 8 mm ,<br />

where it inverts its slope, becoming greater than the sub-scan NEQ after the 9 mm frequency.<br />

This behavior reflects a global characteristic of the system <strong>un</strong>der investigation which has an<br />

evident reduction of the NPS along the scan <strong>di</strong>rection, with respect to the sub-scan <strong>di</strong>rection,<br />

66


Figure 4.25: NEQsubscan.<br />

at higher frequencies. The effect is due to an apo<strong>di</strong>zation filter along the scan <strong>di</strong>rection and is<br />

certified in recent works [24] <strong>per</strong>formed on similar systems.<br />

4.6 DQE measurement<br />

The Detective quantum efficiency has been calculated for each exposure level by <strong>di</strong>vi<strong>di</strong>ng the<br />

NEQ by the SNR of the incident x-ray beam, as [22]:<br />

�� � �<br />

�ÅÌ� �<br />

ÆÈËÒÓÖÑ�Ð�Þ�� � Õ<br />

� Æ�É �<br />

Õ<br />

(4.12)<br />

where � is the spatial frequency and Õ is the number of x-ray photons incident on the detector<br />

<strong>per</strong> <strong>un</strong>it area. The last factor to be determined is Õ.<br />

4.6.1 Determination of q<br />

The total number of incident photons <strong>per</strong> <strong>un</strong>it area of the detector at each exposure level, Õ, has<br />

been calculated starting from the X-ray photon fluence <strong>per</strong> exposure <strong>un</strong>it at each energy, from<br />

the X-ray spectral <strong>di</strong>stribution and the exposure level (X). The photon fluence <strong>per</strong> mR � � ,at<br />

67


Figure 4.26: NEQ scan vs NEQ subscan.<br />

energy (�), is best described by the polynomial [23] [25]:<br />

� � � � � �� �� ��� � � ���� � �� � � �� � ¡<br />

� � � (4.13)<br />

In Fig. 4.27 we show the curve obtained fitting X-ray photon fluence <strong>per</strong> mR between the<br />

range of 5 and 35 keV, as taken from published [23] [25] data.<br />

The X-ray spectrum has been evaluated as described in section 3.2. The exposure value on<br />

the IP surface is evaluated from the calibration curve measurements, described in section 3.1.1<br />

, between mAs and dose on the plate.<br />

The total number of incident photons <strong>per</strong> <strong>un</strong>it area of the detector <strong>per</strong> exposure level X is<br />

calculated as <strong>per</strong> Eq. 4.14<br />

�� �<br />

Ê Õ � � � ��<br />

Ê Õ � ��<br />

(4.14)<br />

The Õ value for the 40 mm PMMA filtered beam is 52800 photons/mm /mR, while for<br />

the <strong>un</strong>filtered beam we obtain 37800. For the 30 mm PMMA case the Õ value is 49700<br />

photons/mm /mR.<br />

68


Figure 4.27: The curve is obtained fitting the X-ray photon fluence <strong>per</strong> mR (taken from pub-<br />

lished [23] [25] data) between the range of 5 and 35 keV.<br />

4.6.2 Measured DQE<br />

DQE estimates versus spatial frequency as a f<strong>un</strong>ction of exposure are shown in Fig. 4.28 for the<br />

scan <strong>di</strong>rection and in Fig. 4.29 for the sub-scan <strong>di</strong>rection.<br />

Figure 4.28: DQE scan.<br />

69


Figure 4.29: DQEsubscan.<br />

In Fig. 4.30, a reduced set of the DQE measurements, as a f<strong>un</strong>ction of exposure level, is<br />

shown in order to allow a comparison between the scan and sub-scan <strong>di</strong>rections.<br />

Figure 4.30: DQE scan vs DQE subscan.<br />

The DQE of this de<strong>di</strong>cated mammography CR system is significantly better than the one<br />

70


of earlier conventional CR system with a 100 �m sampling pitch when acquired <strong>un</strong>der similar<br />

con<strong>di</strong>tions [24] [26].<br />

71


Conclusions<br />

The work <strong>per</strong>formed in the context of this thesis has concerned the study of the <strong>per</strong>formances<br />

of the Fuji FCR5000 MA CR system, with particular emphasis devoted to the evaluation of the<br />

physical quantities involved in the image acquisition process. The fulfilled measurements are<br />

particular interesting in the context of the expansion of <strong>di</strong>gital mammography to clinical use.<br />

Several technologies are currently <strong>un</strong>der test with the goal to provide the benefits of <strong>di</strong>gital ac-<br />

quisition, image processing, electronic <strong>di</strong>splay and storage while, at the same time, producing<br />

an image with as good as or better image quality than the state of the art screen-film detectors.<br />

The results obtained here show that a de<strong>di</strong>cated <strong>di</strong>gital mammography system based on pho-<br />

tostimulable storage phosphor technology (CR) has good quantitative capabilities in terms of<br />

MTF, NPS, NEQ and DQE, particularly when compared to previous CR mammographic im-<br />

plementations. The system included new features with respect to previous CR equipments, as a<br />

dual-side imaging plate detector, which allowed photostimulated luminescence to be extracted<br />

from both sides of the imaging plate, a thicker phosphor layer and 50 �m sampling pitch.<br />

The X-ray <strong>un</strong>it has been fully characterized in order to measure the parameters required for<br />

the subsequent analysis (Dose-mAs curve, energy spectrum, voltage calibration). The IP reader<br />

<strong>un</strong>it has been used in fixed sensitivity mode, with a set of latitude and speed parameters decided<br />

by the user and in linearity mode. Uniform dose expositions at <strong>di</strong>fferent dose levels have been<br />

used to measure the linearity response of the system, for the pixel value-dose calibration and for<br />

the NPS evaluation. The raw data have been analyzed off line with a set of software routines,<br />

base on IDL platform, completely developed in the framework of this thesis. The software<br />

also allows the user to calculate the X-ray spectrum in f<strong>un</strong>ction of the attenuation filter and the<br />

voltage applied to the tube.<br />

The MTF has been measured with the slit method. The MTF curve reaches the 5% value of<br />

its maximum at 8 lp/mm, with an improvement with respect to earlier systems [24]. However<br />

this resolution is 30 % less than would be expected from an ideal detector with 10 mm Nyquist<br />

frequency, basically due to the thicker IP layer. Furthermore, in the spatial frequency <strong>di</strong>agnostic<br />

72


ange (i.e. up to 2.5 lp/mm), the MTF of this system is comparable to the MTF of screen-film<br />

based mammography systems [28].<br />

The evaluation of the two-<strong>di</strong>mensional NPS image has revealed, as expected from literature,<br />

the presence of an apo<strong>di</strong>zation filter that reduces the noise power spectrum significantly above 8<br />

lp/mm in the scan <strong>di</strong>rection. This filter is necessary to reduce the high frequency components of<br />

the decay lag characteristics of the photostimulated luminescence that occurs during the readout<br />

in scan <strong>di</strong>rection.<br />

Finally the NEQ and DQE have been measured since, with widespread consensus in the<br />

international comm<strong>un</strong>ity, they are considered to be the main parameters in order to evaluate<br />

the global <strong>per</strong>formances of an imaging system. The NEQ depends both on the noise pro<strong>per</strong>ties<br />

and on the spatial resolution <strong>per</strong>formances of the detector. It is therefore a good can<strong>di</strong>date to<br />

compare <strong>di</strong>fferent systems, in particular in mammography. In fact in this field the main goals<br />

are often to detect relatively large objects (with <strong>di</strong>mension about 1 mm) with low contrast,<br />

(low noise required), or small objects (with <strong>di</strong>mensions of the order of 200 �m) with high<br />

contrast. An accurate evaluation of the NEQ and its dependence on the exposure has been<br />

<strong>per</strong>formed. The DQE of this de<strong>di</strong>cated mammography CR system is significantly better than<br />

the one of earlier conventional CR system with a 100 �m sampling pitch when acquired <strong>un</strong>der<br />

similar con<strong>di</strong>tions [24] [26]. The measured quantities have revealed to be reproducible and<br />

are in <strong>per</strong>fect agreement with the ones fo<strong>un</strong>d in recent literature works, <strong>per</strong>formed on similar<br />

systems [24].<br />

The low throughput of the system, limited by the scan time of about 90 sec for each IP, is still<br />

a limiting factor in the use of this CR equipment in the service of two or more mammography<br />

rooms during clinical use.<br />

The analysis <strong>per</strong>formed here allows to start a new series of stu<strong>di</strong>es, involving the clinical<br />

<strong>per</strong>formances of the system with the aid of de<strong>di</strong>cated phantoms (such as the CD-MAM or sim-<br />

ilar) and of the first clinical cases <strong>un</strong>der the su<strong>per</strong>vision of specialized me<strong>di</strong>cal staff.<br />

73


Bibliography<br />

[1] M.J. Yaffe and J.A. Rowlands, “X-ray detectors for <strong>di</strong>gital ra<strong>di</strong>ography”, Phys. Med. Biol.<br />

42, 1997, 1-39.<br />

[2] M. Yamada, K. Shimura, “Imaging <strong>di</strong>agnosis quality improved by pattern enhancement<br />

processing for mammography (PEM)”, Fuji Computed Ra<strong>di</strong>ography Technical Review<br />

N.11, FUJI Photo Film C., LTD.<br />

[3] S. Webb, “The physics of me<strong>di</strong>cal imaging”, Me<strong>di</strong>cal Science Series, Institute of Physics<br />

Publishing, Bristol and Philadelphia.<br />

[4] G.F. Knoll, “Ra<strong>di</strong>ation Detection and Measurements ”, 1989, Wiley.<br />

[5] H.G. Chotas, J.T. Dobbins III, C.E. Floyd, C.E. Ravin, “Single-exposure conventional and<br />

computed ra<strong>di</strong>ography image acquisition”, Invest. Ra<strong>di</strong>ol. (26) 1991, 438-445.<br />

[6] H. MacMahon, S. Sanada, K. Doi, M. Giger, X.W. Xu, F.F. Yin, S.M. Montner and M.<br />

Carlin, “Direct comparison of conventional and computed ra<strong>di</strong>ography with a dual-image<br />

recor<strong>di</strong>ng technique”, Ra<strong>di</strong>ographics (11) 1991, 259-268.<br />

[7] M. Sonoda, M. Takano, J. Miyahara and H. Kato, “Computed ra<strong>di</strong>ography utilizing scan-<br />

ning laser stimulated luminescence”, Ra<strong>di</strong>ology (148), 1983, 833-838.<br />

[8] K. Takahashi et al., “Mechanism of Photostimulated Luminescence in ���� � �Ù<br />

(X=Cl, Br) Phosphors”, Journal of Luminescence (31/32), 1984, 266-268.<br />

[9] H. von Seggern, T. Voigt, W. Knupfer, G. Lange, “Physical Model of Photostimulated<br />

Luminescence of x-ray Irra<strong>di</strong>ated ����Ö � �Ù ” J Appl. Phys. 64(3), 1988, 1405-<br />

1412.<br />

74


[10] H. Yasuda et al., “Development of high-quality mammographical FCR system based on<br />

dual light collection technology”, Fuji Computed Ra<strong>di</strong>ography Technical Review N.13,<br />

FUJI Photo Film C., LTD.<br />

[11] S. Arakawa et al.,“Novel computed ra<strong>di</strong>ography system with improved image quality by<br />

<strong>di</strong>rect detection of emission from both sides of imaging plate”, SPIE Me<strong>di</strong>cal Imaging<br />

1999 (Physics of Me<strong>di</strong>cal Imaging), Vol. 3659, San Diego, February 1999.<br />

[12] J.T.Dobbins III , “Effects of <strong>un</strong>dersampling on the pro<strong>per</strong> interpretation of modulation<br />

transfer f<strong>un</strong>ction, noise power spectra, and noise equivalent quanta of <strong>di</strong>gital imaging<br />

systems”, Med.Phys.22(2), October 1995, 171-181.<br />

[13] E. Hecht, “Hecht”, Second E<strong>di</strong>tion 1987, Ad<strong>di</strong>son-Wesley Publishing Company.<br />

[14] J.C.Dainty and R.Shaw, “Image science”, London, Academic Press, 1974.<br />

[15] J.M.Boone,T.R.Fewell,R.J.Jennings, “Molybdenum,rho<strong>di</strong>um and t<strong>un</strong>gsten anode spectral<br />

models using interpolating polynomials with application to mammography”, Med. Phys.<br />

24(12), December 1997, 1863-1874.<br />

[16] H. Kato, “Photostimulable phosphor ra<strong>di</strong>ography:Design considerations”, in Specifica-<br />

tion, Acceptance Testing and Quality Control of Diagnostic X-ray Imaging Equipment,<br />

J.A. Seibert, G.T.Barnes, R.G. Gould, Eds. AIP Press, 1994, 731-770.<br />

[17] Fuji Computed Ra<strong>di</strong>ography FCR5000 MA, Rev.020402, Fuji Me<strong>di</strong>cal Systems Italia<br />

s.p.a.<br />

[18] R.G.Waggener et al., “X-ray spectra estimation using attenuation measurements from 25<br />

kVp to 18 MV”, Med. Phys.26(7), July 1999.<br />

[19] H.Fujita et al., “A simple method for determining the modulation transfer f<strong>un</strong>ction in<br />

<strong>di</strong>gital ra<strong>di</strong>ography”, IEEE Transaction on Me<strong>di</strong>cal Imaging, Vol. 11,No.1, March 1992.<br />

[20] M.L.Giger and K.Doi, “Investigation of basic imaging pro<strong>per</strong>ties in <strong>di</strong>gital ra<strong>di</strong>ography.<br />

1.Modulation transfer f<strong>un</strong>ction.”, Med. Phys.11, 1984, 287-295.<br />

75


[21] “Characteristics of <strong>di</strong>gital X-ray imaging devices - Determination of the detective quantum<br />

efficiency”, International Electrotechnical Commission, Document IEC 62220-1CDV, IEC<br />

2002-3-22.<br />

[22] J.T.Dobbins III et al., “DQE(f) of four generation of computed ra<strong>di</strong>ography acquisition<br />

devices”, Med.Phys.22(10), October 1995, 1581-1593.<br />

[23] S.Vedantham et al., “Full breast <strong>di</strong>gital mammography with an amorphus silicon-based flat<br />

panel detector: Physical characteristics of a clinical prototype”, Med. Phys. 27 (3),March<br />

2000, 558-567.<br />

[24] J.A.Seibert,J.M.Boone,V.N.Coo<strong>per</strong>, “Determination of imaging <strong>per</strong>formance of a photo-<br />

stimulable phospor system for <strong>di</strong>gital mammography”, Procee<strong>di</strong>ngs of SPIE VOl.4682<br />

(2002), 447-456.<br />

[25] H.E.Johns and J.R.C<strong>un</strong>ningham, “The Physics of Ra<strong>di</strong>ology”, 4th ed., (Thomas, Spring-<br />

field, IL,1983)<br />

[26] J.A.Seibert,J.M.Boone,V.N.Coo<strong>per</strong>, “Evaluation of a Computed Ra<strong>di</strong>ography Detector for<br />

Digital Mammography”, Med. Phys. 28 (2001) , 1246.<br />

[27] “Acceptance Testing and Quality Control of Photostimulable Storage Phosphor Imaging<br />

Systems”, Report of Task Group N.10, American Association of Physicists in Me<strong>di</strong>cine,<br />

DRAFT Document version 4.1, August 1998.<br />

[28] P.C. B<strong>un</strong>ch, “Objective imaging characteristics of mammography screen-film systems.”,<br />

SPIE 1999, Vol.2708/241-271.<br />

76


Ringraziamenti<br />

Ipiù sinceri ringraziamenti vanno al Dr. Giacomo Belli e alla Dott.ssa Barbara Lazzari <strong>per</strong><br />

la <strong>di</strong>sponibilità, la professionalità e la simpatia con cui hanno reso meno pesanti le corse su e<br />

giù <strong>per</strong> Firenze. Tanta gratitu<strong>di</strong>ne alla Prof.ssa Marta Bucciolini e al Dr. Cesare Gori <strong>per</strong> la<br />

pazienza e la gentilezza nel darmi consigli. Un grazie al Prof. Salvatore Romano, anche <strong>per</strong><br />

aver semplificato quando possibile la burocrazia e i formalismi. Infine <strong>un</strong> sorriso a Francesco<br />

R., del quale è il merito se sono arrivato a scrivere questo lavoro.<br />

77

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!