Molecular Imaging of Inflammation Using Echocardiography. Advances with the Use of Microbubbles


Adhesion molecule

Alternative designation

Localisation

Ligand

Function

Basal expression

Selectin family

P-selectin

CD62P

Endothelial cell

L-selectin

Rolling

Yes

PADGEM

Platelet

PSGL-1
  
GMP-140
 
120 kD PSL
  
E-selectin

CD62E

Endothelial cell

L-selectin

Rolling

Noa

ELAM-1
 
CLA, SSEA-1
    
250 kD ESL-1
  
Ig-supergene family

ICAM-1

CD54a

Endothelial cell

CD11a/CD18

Adherence/transmigration

Yes
 
Leukocyte (myocyte, haematopoietic cell, fibroblast) [13]

CD11b/CD18
 
VCAM-1

CD106

Endothelial cell (myocyte, fibroblast, follicular dendritic cell) [14]

CD49d/CD29

Adherence

No/very low

PECAM-1

CD31

Endothelial cell

PECAM-1

Adherence/transmigration

Yes
 
Leukocytes
   
Platelet
  
MAdCAM-1
 
Endothelial cell

L-selectin

Adherence/transmigration

Yes
 
(Intestine)

CD49d/β7
 

Adapted from Krieglstein et al. [7]

CLA cutaneous lymphocyte antigen, ELAM-1 endothelial leukocyte adhesion molecule-1, ESL-1 E-selectin Ligand-1, GMP-140 granule membrane protein-140, ICAM-1 inter-cellular adhesion molecule-1, MAdCAM-1 mucosal address in cell adhesion molecule-1, PADGEM platelet activation-dependent granule external membrane protein, PECAM-1 platelet endothelial cell adhesion molecule-1, PSGL-1 P-selectin glycoprotein ligand-1, SSEA-1 sialyl stage-specific embryonic antigen-1, VCAM-1 vascular endothelial cell adhesion molecule-1

aBasal expression of E-selectin in resting endothelial cells, where present, is very low or undetectable. However, detectable low levels of constitutive E-selectin expression has been reported in some vascular beds of certain animals in the absence of overt inflammation [15, 16]



Stimuli for the up-regulated expression of endothelial adhesion molecules include tumour necrosis factor, interleukin-1β, γ-interferon, substance P and bacterial endotoxin lipopolysaccharide [18]. Maximal endothelial adhesion molecule expression is generally noted between 3 and 6 h following exposure to the stimulus [7].

Transcription factors such as NF-κB and AP-1 have been shown to regulate the expression of endothelial adhesion molecules. NF-κB also leads to further production of inflammatory mediators, such as tumour necrosis factor, by the recruited leukocytes. Thus a vicious cycle occurs at the site of inflammation with inflammatory mediator induced endothelial adhesion molecule expression, leading to leukocyte recruitment, leading to further production of inflammatory mediators from the recruited leukocytes, leading to further endothelial adhesion molecule expression and leukocyte recruitment [7, 19].

There are species [20] and tissue [16] differences in endothelial adhesion molecule expression. More recently, genetic polymorphisms of endothelial adhesion molecules have been associated with their altered expressions in inflammatory conditions, such as atherosclerosis, rheumatoid arthritis, other connective tissues diseases, vasculitis and diabetes mellitus [20].

The up-regulated expression of endothelial adhesion molecules may be used as molecular markers of endothelial activation or inflammatory activity in a number of important cardiovascular and non-cardiovascular diseases [1, 3133]. Furthermore, imaging of a series of endothelial adhesion molecules may be relevant because the expression profile (combination and level) of individual adhesion molecules may differ in different types and/or stages of diseases e.g., acute vs chronic vs resolving inflammation. For example, there is evidence that E- and P-selectin are involved in the selective recruitment of T-helper 1 cells [22, 23]. In myocarditis, T-helper 1-dominated responses are associated with the induction of myocarditis, whilst T-helper 2-dominated responses are associated with enhanced recovery [21]. Thus it may be possible to determine the status of myocarditis without tissue biopsy, by non-invasively imaging the endothelial adhesion molecule expression profile using quantitative ultrasound molecular imaging.


24.1.1.1 E-Selectin


Classically, E-selectin is expressed only on activated endothelial cells in the post-capillary venules. Its basal expression in resting endothelial cells, where present, is very low or undetectable. However, detectable low levels of constitutive E-selectin expression has been reported in some vascular beds of certain animals in the absence of overt inflammation [15, 16]. In addition to the post-capillary venules (the classical site of E-selectin expression), E-selectin expression has also been found in inflamed myocardial microvasculature [24], renal glomeruli [10] and the aorta [25, 26], in mouse models of Coxsackievirus B3 myocarditis [24], glomerulonephritis [10], and atherosclerosis [25], respectively.

E-selectin interacts with N- and O-glycans that contain the sialyl Lewis x tetrasaccharide or its isomers, such as the glycoprotein L-selectin (in man, not mice), E-selectin ligand-1 (ESL-1) and P-selectin glycoprotein ligand-1 (PSGL-1). These E-selectin ligands are present on neutrophils, monocytes, lymphocytes, myeloid cells, dendritic cells, and eosinophils [15].

E-selectin on the endothelial cell surface may be cleaved, releasing soluble E-selectin that lacks the intact cytoplasmic domain. The normal blood level of soluble E-selectin has been reported to be in the range of 0.1–3 ng/ml [12]. Cell-surface E-selectin may also be internalised into the endothelial cell by endocytosis [27].

E-selectin has been detected in human tissues of cardiac transplant rejection [28], myocarditis [29], dilated cardiomyopathy [29], coronary atheromatous heart disease [30], rheumatoid arthritis [31], inflammatory bowel disease [8], chronic inflammatory disease of the skin [9], cytomegalovirus infection [11] and sepsis [12].

E-selectin’s mono cell type expression and specificity for endothelial activation/inflammation made it an attractive prototype molecule for developmental work in quantitative real-time ultrasound molecular imaging Yeh et al. 2008 [1, 34].



24.1.2 Microbubbles as Ultrasound Contrast Enhancement Agents



24.1.2.1 Ultrasound Imaging for Microbubble Detection


Microbubbles are strong scatterers of ultrasound, readily detectable at typical clinical ultrasound frequencies of 2–13 MHz, despite being <6–10 um in diameter which is significantly smaller than the resolution of ultrasound. The echogenicity of the bubbles is dependent on the bubble size (size distribution), compressibility of the bubbles or gas, shell thickness/elasticity/viscosity, bubble scattering: attenuation ratio, insonating ultrasound frequency and power.

Microbubbles are compressible and may undergo non-linear radial oscillation in an acoustic field generating strong backscattered signals that can be differentiated from tissue signals using special non-linear ultrasound imaging modes (e.g., pulse inversion, Contrast Pulse Sequencing, Power Modulation).

In the acoustic field, microbubbles have different responses depending on the insonating acoustic power (peak negative pressure or mechanical index (MI)). At very low acoustic powers (e.g., <0.1 MPa or MI <0.1), bubbles oscillate linearly (contraction and expansion are similar in amplitude). At low-moderate powers (e.g., 0.1–1 MPa or MI 0.1–1), bubble expansion is larger than contraction (non-linear oscillation), and produces signals whose frequencies are multiples (harmonics) of the incident ultrasound wave. At high powers (e.g., >1 MPa or MI >1), the bubbles are destroyed rapidly, producing a brief non-linear response and harmonic signal of high amplitudes.

Non-linear imaging methods allow sensitive and specific detection of bubbles by evaluating the frequency components due to nonlinear response of the bubbles and suppressing those due to linear response of the tissues. Methods include high or low acoustic power (MI) settings (see below). High acoustic power settings are usually >1 MPa (MI >1), while the low power settings are usually 0.1–0.25 MPa (MI 0.1–0.25). The MI used for contrast ultrasound studies in clinical practice ranges 0.05–1.9.

The high acoustic power imaging modes (0.8–2.5 MPa or MI >0.8) are based on harmonic B-mode, harmonic Doppler or Power Doppler techniques. They cause rapid bubble destruction, associated with a brief emission of intense non-linear bubble signals, giving high bubble sensitivity. These bubble destructive imaging modes exhibit a higher signal-to-noise ratio, better bubble-to-tissue contrast and lesser nonlinear signal propagation in tissues compared with fundamental B-mode or Doppler imaging. However only intermittent single-frame contrast ultrasound images can be acquired, due to the rapid bubble destruction. This precludes continuous real-time imaging/assessment.

The low acoustic power imaging modes (0.05–0.5 MPa or MI 0.05–0.5) are based on pulse inversion (harmonic or Doppler), multi-pulse or coded excitation techniques. They cause minimal bubble destruction, allowing continuous real-time imaging. However, the lower acoustic power used has higher attenuation and lower bubble backscatter signal intensity, which reduces bubble detection sensitivity compared with the high acoustic power techniques. In harmonic mode, the attenuation is higher for returning harmonics than for the incident pulse. This effect cannot be compensated for by a higher amplitude of the incident pulse because it would give rise to: (a) nonlinear signal propagation in tissues, reducing signal-to-noise ratio and bubble-to-tissue contrast; (b) increased bubble destruction. Newer imaging modes based on multi-pulse sequence or coded excitation have increased signal-to-noise ratio and bubble-to-tissue contrast, allowing highly sensitive and specific bubble imaging at low acoustic powers. E.g., Contrast Pulse Sequencing [74] and Power Modulation imaging [75].

In practice, the relationship between the backscatter signal intensity and bubble concentration is exponential, with an approximately linear-range at low bubble concentrations [1, 76, 77]. At higher bubble concentrations, the increase in signal intensity plateaus; and at even higher concentrations, the signal intensity decreases because of: (a) signal saturation due to the limited dynamic range of the ultrasound system; and (b) signal attenuation due to multiple scattering effect [78], absorption, other ultrasound energy dissipating mechanisms such as viscous loss and thermal damping [79].


24.1.2.2 Microbubble Stability


Bubbles decay by dissolution (gas deflation), resulting in a decrease in size and number, and contrast ultrasound signal intensity. Factors influencing the stability of bubbles include the bubble size, characteristics of the bubble gas (low solubility, high molecular weight, low Ostwald coefficient, high saturated vapour pressure or low lipophilicity gives more stable bubbles), diffusivity of the gas across the shell, the shell thickness, surface tension, and solubility (bubbles with water soluble galactose shell last for shorter periods than those with relatively insoluble albumin- or phospholipid-shell), shell stability against mechanical perturbation, and the presence/absence of mechanical perturbation such as ultrasound insonation, shear force and systemic pressures in vivo [1, 3639].


24.1.2.3 Biodistribution and Elimination of Microbubbles In Vivo


Microbubbles are considered as flow tracers which remain within the intravascular space and have rheology similar to that of the red blood cells [35]. The lungs and reticuloendothelial system (RES) are their major routes of elimination. They differ from iodinated or other contrast agents commonly used in scintigraphy, CT, MRI or PET imaging, which diffuse into the interstitium or cells, with often the kidneys as one of the major routes of elimination.

Following iv administration, the bubbles first reach the right heart, then the pulmonary circulation in the lungs, then return to the left heart and onwards to the systemic circulation. They continuously decay (decreasing in size and number) in the blood stream by gaseous deflation, the rate of which is influenced by the surrounding milieu for gaseous diffusion, such as the diffusion gradient and changes in the ambient pressure (such as in the ventricle, lungs or acoustic field) [3639]. Iv administered bubbles undergo first-pass elimination in the lungs by size filtration, whereby those larger than the capillaries (e.g., >6–10 μm diameter) are mechanically trapped in the pulmonary microcirculation. The gas is exhaled out of the lungs unmetabolised. There is no significant sequestration of the bubbles in the lungs [40], presumably because the large trapped bubbles deflate rapidly. This first pass lung extraction has been shown to occur within 20–60 s [41, 42] in man following iv injection, and may decrease the total bubble gas volume by 30–50 %, depending on the size distribution of the bubbles administered. The bubbles are continuously removed from the circulation by RES uptake – primarily in the liver and spleen – involving phagocytosis by Kupffer cells or splenic macrophages, respectively [40, 4348]. This is similar to other macromolecular contrast agents such as liposomes [49], supramagnetic iron oxide particles [50] and quantum dots [51]. The bubble shells are metabolised through pathways dictated by its component composition. Bubble elimination approximated to first-order kinetics (reviewed in Yeh et al. 2010 [1]).

In the pig, a significant proportion of microbubbles are removed by the pulmonary intravascular macrophages lining the lung capillaries [46, 52]. Pulmonary intravascular macrophages are only found in certain species such as cats and some mammals of the order Artiodactyla, which includes pigs, sheep, goats, and cattle. They are absent or negligible in rodents, rabbits, chickens, monkeys or man [53, 54]. The uptake of bubbles by the pulmonary macrophages causes powerful pulmonary vasoconstriction – apparently mediated by thromboxane [55]. An important implication of this is that experiments using certain animals (e.g., pigs) would require pre-treatment with steroids or non steroidal anti-inflammatory drugs, in order to prevent adverse effects from pulmonary vasoconstriction induced by the bubbles. The experiments would need to be designed such that the critical region of the hypothesis being tested will not be affected by these variables.

The biodistribution or elimination of targeting microbubbles are seldom described, however they are expected to be similar to that of the non-targeting bubbles, with the exception that targeting bubbles accumulate additionally at site(s) of their targeted molecule expression [44, 45]. Unlike circulating bubbles, retained bubbles at target sites are eliminated primarily by passive and/or ultrasound mediated gaseous deflation [39]; RES-uptake or rapid gaseous deflation in the lungs do not occur unless the bubbles targeted molecules expressed in the RES or lungs, respectively. Retained bubbles in non-RES tissues may be phagocytosed and degraded by leukocytes [56], but this is not universal [1] and likely dependent on the bubble property (e.g., chemical composition, physico-chemical and biological properties), the biology of the targeted molecule, the targeted cell, tissue or disease process. The elimination of targeting bubbles approximates to first-order kinetics [1]. The half-life of retained bubbles is longer the greater its initial (maximum) concentration [1]. This may be due to a bubble-to-bubble protection phenomenon recently described in vitro [57] and in vivo [1]–as the distance between adjacent bubbles shortens at higher concentrations, there is: (i) increased acoustical interactions between adjacent bubbles causing mechanical responses such that the net diffusion of gas out of the bubble population is reduced; and/or (ii) reduced concentration gradient for gas diffusion out of the bubbles, due to increased gas saturation in the micro-environment surrounding them. Such phenomenon highlights an important pitfall of conventional acoustic quantification methods (see Sect. 24.2.1).


24.1.2.4 Composition and Preparation of Microbubbles


The chemical composition and processing condition of the microbubbles influence their physical chemistry and mechanical properties, thus affecting their stability, yield, charge, size distribution, acoustic response, degree of non-specific adherence/retention in tissues, elimination and toxicity in vivo. Some of these variables are inter-related [1, 5863].

Non-targeting and targeting microbubbles, produced by the pharmaceutical industry and others, which demonstrated ultrasound imaging in vivo are shown in Table 24.2. Detailed information on the composition and processing conditions required for making them is generally lacking in the public domain. For phospholipid-based microbubbles, the phospholipids can be dispersed along with surfactant polyethylene glycol (PEG) stearate in an organic solvent. The solvent is evaporated and replaced with an aqueous buffer. Bubbles are then formed using shear mixing approaches [58], where the mixture is sonicated or agitated vigorously in the presence of the filling-gas. For making targeting bubbles, targeting ligands can be conjugated to bubbles after the bubble formation (Fig. 24.1); or they can be conjugated to one of the shell components (e.g., a phospholipid species) prior to bubble formation [60].


Table 24.2
Examples of microbubbles used in humans (albumin-, phospholipid- or other-shelled bubbles)








































































































































































































































































Microbubble

Composition

Gas

Net charge/zeta potential

Mean size or cumulative size distribution (μm)

Conc (x109/ml)

Application

Clinical trial status

Approval status

Developer

BY963

Phospholipid shell: contains DSPG

Air
     
LVO Doppler (cerebral, central and peripheral vessels)

Phase III (Europe)
 
Byk Gulden

Phase III

Definity (Perflutren, Aerosomes, MRX 115, DMP 115)

Phospholipid shell (8 nm thick): DPPA: DPPC: MPEG5000DPPE in 10:82:8 molar ratio

C3F8

Neg

1.1–3.3

98 % <10 μm

1–1.5

LVO

FDA-filed (USA and Europe), Phase III (Japan)

Licensed

Bristol-Myers Squibb (ImaRx Pharmaceuticals?)

MP

Phase II/III

Hepatic and renal masses

FDA-filed (USA and Europe), Phase III (liver, Japan)

Breast masses

Phase II

Imagent (Imavist, AFO-150)

Phospholipid shell: contains DMPC

C6F14

Neut

5

0.5

LVO

FDA-filed (USA)

Licensed

Alliance Pharmaceutical Corporation + Schering AG

99.8 % <10 μm

MP

Phase II completed

Stress echo (pharmacological stress)

Phase II

Lesions and blood flow abnormalities

Phase II

Prostate and breast

Phase II

Sonazoid (NC100100)

Phospholipid shell

C4F10

Neg

3

1

LVO

FDA-submitted (USA)

Not approved

Nycomed Imaging

MP

Phase II

Visualisation of liver lesions

Phase II completed

SonoVue (BR1)

Phospholipid shell: contains DSPC, DPPG, palmitic acid, PEG4000

SF6

Neg

2.5

>90 % <8 μm

0.1–0.5

LVO

FDA-filed (USA and Europe)

Licensed

Bracco Diagnostics

Macrovasculature

FDA-filed (USA and Europe)

Licensed

MP

Phase II
 

BR55 (VEGFR2 targeting bubble)

Phospholipid shell: contains DSPE-PEG(2000), DSPE-PEG(2000)-peptide

C4F10 + N2
 
1.5 ± 0.1(SD), range 1–3 μm

2

Prostate cancer imaging

Phase 0 (results awaited)
 
Bracco Diagnostics

Echovist (SH U454)

Galactose-based shell

Air
 
99 % < 12 μm, 95 % < 8 μm
     
Licensed

Schering AG

Levovist (SHU 508A)

Galactose-based shell: 99.9 % galactose, 0.1 % palmitic acid

Air

Neg

95 % < 10 μm
 
LVO

Marketed in Europe, etc.

Licensed (EU, Japan, Canada)

Schering AG

Doppler and CFI; vascular

Phase III completed in USA and Japan
 

Albunex (ABX)

Albumin shell (15 nm thick):

Air

−3 mV

4 ± 2 (SD)

0.5–0.8
   
Withdrawn

Mallinckrodt

5 % human serum albumin

Optison (FS069)

Albumin shell:

albumin N-acetyltryptophan, caprylic acid

C3F8

Slight neg

2–4.5

0.5–0.8

LVO

Marketed in USA and Europe

Withdrawn

Molecular Biosystems

93 % <10 μm

MP

Phase II/III

Stress echo

Phase II

Liver imaging

Phase II/III

TCD and kidney

Quantison

Albumin:

Air
 
3.2

1.5

MP

Phase II (suspended)

Not approved/withdrawn

Andaris Ltd

Human serum albumin (recombinant)

Cardiosphere (PB 127)

Bilayer polylactide/albumin shell

N2

Slight neg
   
MP

Phase III completed
 
POINT Biomedical

Sonovist (Sonavist, SHU 563A)

Polycyanoacrylate shell

Air
     
Liver, kidney, MP

Phase II
 
Schering AG

Acusphere (Imagify, AI 700)

Poly(lactic-co-glycolic acid) based shell

C4F10

Neg
   
MP

Phase III
 
Acusphere Inc

EchoGena (QW3600)

Charged surfactant (liquid droplet without shell):

DDFP

Neg

3–10

3–5

LVO

Approved for marketing in Europe

Withdrawn

Sonus Pharmaceuticals

2 % DDFP emulsion containing surfactant (PEG Telomer B) + sucrose

MP

Phase II

Stress echo (exercise and pharmacological stress)

Phase III completed

Chest pain in ER

Phase II

3D echocardiography

Phase III completed

Prostate


Adapted and modified from Yeh et al. [1]

Microbubbles produced in the industry not reaching beyond Phase I clinical trial are omitted (e.g., BR14, Myomap (AIP 201), SonoGen (QW7437), SHU 616A), except BR55 (the only targeting microbubble tested in humans, Phase 0)

Abbreviations: CFI colour flow imaging, Conc concentration, DDFP dodecafluoropentane, DPPA dipalmitoylphosphatidic acid, DPPC dipalmitoylphosphatidylcholine, DPPE dipalmitoylphosphatidalethanolamine, DPPE-MPEG5000 N-(methoxypolyethylene glycol 5000 carbamoyl)-DPPE, DPPG dipalmitoylphosphatidylglycerol, DSPC distearoylphosphatidylcholine, DSPG-Na 3-sn-phosphatidyl-1-DL-glycero-distearyl sodium derivative of phospholipid isolated from soybeans, LVO left ventricular opacification, also known as endocardial border delineation (EBD)), MP myocardial perfusion, PEG4000 macrogol 4000, TCD transcranial Doppler

aEchogen is a liquid-in-liquid fluorocarbon emulsion containing surfactant (PEG Telomer B), where DDFP microdroplets phase shift to bubbles after physical activation in vivo (hypobaric activation by body temperature, NB. boiling point = 29.3 °C) [73]


A319559_1_En_24_Fig1_HTML.jpg


Fig. 24.1
Strategy for converting native (non-targeting) microbubbles to targeting microbubbles for ultrasound molecular imaging. Native bubbles containing biotin or maleimide on its shell surface are conjugated with targeting ligands (e.g., mAb or its F(ab′)2 fragment). Top: traditional biotin-streptavidin conjugation where biotinylated mAb is linked to bubbles containing biotin on its surface, via streptavidin. Bottom: maleimide-thiol conjugation where reduced F(ab′)2 containing thiol (thiol produced from the reduction of interchain disulfide bond) is linked to bubbles containing maleimide on its surface. Abbreviations: mAb monoclonal antibody, NEM N-ethylmaleimide

Conjugation chemistries used in making targeting bubbles include non-covalent biotin-(strept)avidin and covalent maleimide-thiol [1, 64], pyridyldithio-thiol [102], carbodiimide [65], benzotriazole-amine [66], succinate-amine [103] or amine-succinimidyl [67] conjugations. To date, the biotin-(strept)avidin conjugation chemistry is the most commonly used, it being simple, robust, efficient, and forms very strong stable bonds. Although attractive for early developmental work, it is not suitable for human use because streptavidin or avidin is a bacterium Streptomyces avidinii protein or egg white protein, respectively, both known to produce unwanted immunogenic side effects [68]. An alternative maleimide-thiol conjugation chemistry has the advantage of low potential immunogenicity. Conjugate products containing maleimide have been tested in humans without significant immunogenic or toxic side-effects attributable to maleimide per se [69, 70]. This includes the maleimide-PEG-hemoglobin conjugate, Hemospan®, which has completed Phase III clinical trial [70]. The thioether bond formation between the maleimide on the bubble-shell and thiol on the targeting-ligand is strong and rapid at near neutral pH (bond strength in the order of nano newtons; second-order rate constant 0.8–1.2 × 104 M−1 s−1) [71]. The near neutral pH is advantageous in avoiding negative impact on the ligands and bubbles during the conjugation reaction, and in preventing dissociation of the ligands from the bubble-shell in vivo. In 2008, Yeh et al. reported E-selectin targeting bubbles, based on maleimide-thiol conjugation chemistry, that were effective for ultrasound molecular imaging in vivo (Figs. 24.2 and 24.3). These bubbles were stable, echogenic, specific to the targeted molecule and effective for real-time and quantitative ultrasound molecular imaging in vivo. Non-specific retention in non-RES tissues and adverse effects were minimal following iv administration [1, 34, 64]. Elsewhere, an amine-succinimidyl conjugation chemistry has been used in BR55 (a VEGFR2 targeting microbubble) [67], the only targeting bubble tested in humans so far – the results of its Phase 0 clinical trial are still awaited [72].

A319559_1_En_24_Fig2_HTML.jpg


Fig. 24.2
Ultrasound molecular imaging of E-selectin expression in the mouse heart. Sequential Images (a, b). E-selectin targeting microbubbles were administered as a rapid iv bolus via the tail vein at 10s (0:10) to anaesthetised WT (a) and E-selectin KO (b) mice, both pre-treated with lipopolysaccharide to induce systemic inflammation involving the heart. Shown are sequential 14 MHz low MI Contrast Pulse Sequencing (a bubble-specific imaging mode) images of the heart in end-diastole, in parasternal short axis view of the LV. Bubble signal intensities are shown in a heated object scale (orange colour), tissue signals are not displayed. Results: Signals in the LV cavity and myocardium peaked within a few heart beats (<1 s). Signal attenuation in the LV cavity and myocardium decreased with time (e.g., frame 0:30 vs 6:20) as bubble concentrations decreased. Signal in the LV cavity (representing the concentration of free flowing bubbles in the blood pool) decreased more rapidly in the WT than the E-selectin KO mouse. Signal in the myocardium persisted after clearance of bubbles from the blood pool in the WT but not the KO mouse, indicating retention of the E-selectin targeting bubbles by E-selectin expressed in the WT myocardium. Time intensity curves of the LV cavity and myocardium in the WT (c) and E-selectin KO (d) mice. Plotted are the background subtracted mean acoustic units (AU) ± SD for the LV cavity (region of interest (ROI) C) and myocardium (ROI M). Suggested bolus (B), distribution (D) and elimination phases (E) of the bubbles are shown. A second lower peak signal intensity ≈ 40–50s post bubble administration could be observed in most animals (arrow, inset) as ultrasound signal attenuation reduced with decreased bubble concentration in the distribution phase


A319559_1_En_24_Fig3_HTML.jpg


Fig. 24.3
Ultrasound molecular imaging of E-selectin expression in the mouse kidneys. WT and E-selectin KO mice were both pre-treated with lipopolysaccharide to induce systemic inflammation involving the kidneys. Baseline transverse images of the abdomen were obtained by scanning in a cranial-caudal direction, before iv bolus administration of E-selectin targeting microbubbles (i–iv). Imaging of the abdomen was resumed at ~30min post bubble administration (v–viii). Shown are 14 MHz low MI Contrast Pulse Sequence images of the upper abdomen encompassing the liver (i, iii, v, vii) and lower abdomen encompassing the kidneys (K) and spleen (S) (ii, iv, vi, viii). Bubble signal intensities are displayed in a heated object scale (orange colour) superimposed with tissue signals in grey scale. Results: The elimination of the circulating bubbles by RES-uptake could be inferred from the persistence of bubble signals in the liver and spleen, in both the WT and E-selectin KO mice, following the clearance of circulating bubbles from the blood pool. Specific targeting of the bubbles to E-selectin in the kidneys was observed in the WT, shown as the persistence of bubble signals in the kidneys after clearance of the circulating bubbles from the blood pool, which occurred in the WT but not E-selectin KO mice. The targeted bubble signal for E-selectin was located predominantly in the renal cortex, as expected (E-selectin was expressed predominantly on the endothelial cells of the renal glomeruli, which are concentrated in the renal cortex)

The number of targeting ligands per bubble ranged 0.3–80 × 104 (0.6–300 × 103 ligands/μm2), most in the order of 1–10 × 104/bubble (1–10 x 103 ligands/μm2) Table 24.3.



Table 24.3
Phospholipid-shelled targeting microbubbles tested for ultrasound molecular imaging in vivo


























































































Origin (bubble ID, where available)

Conjugation chemistry

Bubble shell composition

Bubble gas

Targeting ligand

Targeting ligand density on bubble (conjugation reaction ratio)

Bubble diameter (μm)

Targeted molecules

Disease model

Tissues imaged

Host & bubble dose (number of bubbles)

Applicationa

References

UVA

Biotin- streptavidin

DSPC (83 mol %)

PEG40-stearate (16 mol %)

DSPE-PEG2000- biotin (1 mol %)

±DiI, DiO or rhodamine DHPE (small unspecified amount)

C4F10

mAb

Peptide

Protein

2–165 k/μm2

60–300 k/bubble

Mean: 1.9–3.4

Range: 94–97 % are <5

α5-integrins

αv-integrins

β1-and β3-integrins

CX3CR-1

H-2Kk (mouse MHC class I H-2Kk protein) on transfected bone-marrow derived endothelial progenitor cell

ICAM-1

VCAM-1

MAdCAM-1

P-selectin

Platelet glycoprotein-1bα

Selectins (not discriminating amongst E-, P-, and L-selectin)

vWF (A1-domain)

Angiogenesis (chronic ischemia, stem cell therapy, tumor)

Coronary thrombosis & angioplasty

Inflammation (atherosclerosis, cardiac transplant rejection, Crohn’s disease (small bowel), ischemia- reperfusion injury)

Aorta

Bowel

Heart

Kidney

Skeletal muscle

Tumor (implanted in brain or flanks)

Matrigel plug (implanted in abdomen)

Mouse: 1–10 × 106

Rat: 2.5–100 × 106

Dog: 108

RT

SQ

Lindner JR et al., Circulation 2001;104:2107

Ellegala DB et al., Circulation 2003;108:336

Leong-Poi H et al., Circulation 2003;107:455; Circulation 2005;111:3248

Weller GE et al., Circulation 2003;108:218; Cancer Res 2005;65:533

Sakuma T et al., Cardiovasc Res 2005;66:552

Bachmann C et al., Gastroenterology 2006;130:8

Villanueva FS et al., Circulation 2007;115:345

Kaufmann BA et al., Circulation 2007;116:276; Eur Heart J 2007;28:2011; J Am Soc Echocardiogr 2010;23:79; Arterioscler Thromb Vasc Biol 2010;30:54

Behm CZ et al., Circulation 2008;117:2902

Kuliszewski MA et al., Cardiovasc Res 2009;83:653

McCarty OJ et al., JACC Cardiovasc Imaging 2010;3:947

Carr CL et al., Arterioscler Thromb Vasc Biol 2011;31:2526

Davidson BP et al., J Am Coll Cardiol 2012;60:1690

Khanicheh E et al., Arterioscler Thromb Vasc Biol 2013;33:2187

Liu Y et al., Circ Cardiovasc Imaging 2013;6:74

Ryu JC et al., Circulation 2013;127:710

UVA

Biotin- streptavidin

DSPC (83 mol %)

PEG40-stearate (16 mol %)

DSPE-PEG3400- biotin (1 mol %)

C4F10

mAb

Protein

Peptide

NA

Mean: 3.2

Selectins (not discriminating between E- and P-selectin)

VCAM-1

VEGFR2

Angiogenesis (inflammation, tumor)

Atherosclerosis

Aorta

Tumor (implanted in hind limb)

Mouse: 1–10 × 106

Not RT

van Wamel A et al., Proc IEEE Ultrason Symp 2007;961 (abstract)

Khanicheh E et al., PLos One 2013;8:e58761

UVA (bubbles loaded with plasmid DNA)

Biotin- streptavidin

DSPC (70 mol %)

PEG40-stearate (13 mol %)

DSPE-PEG2000- biotin (1 mol %)

DSTAP (16 mol %)

cDNA (plasmid containing firefly luciferase gene): 0.04 pg loaded per bubble

±DiI or DiO

C4F10

mAb

NA

Mean: 2.5–3.1

P-selectin

Microbubble targeted gene delivery in ischemia (ischemia- reperfusion)

Skeletal muscle

Mouse: 2 × 108

Not RT

Xie A et al., J Am Coll Cardiol Img 2012;5:1253

Targeson (TargeStar-series)

Biotin- streptavidin

TargeStar/TargeStar-B

 Biotin coated bubbles

 Component details not disclosed

TargeStar-SA

 Streptavidin coated bubbles, containing 2.7 k streptavidin/μm2 (streptavidin on distal tip of PEG)

 Component details not disclosed

Perfluorocarbon (not specified)

mAb

Peptide

scFvs

(Reaction ratio: ligands at >20x excess necessary to saturate streptavidin binding sites)

Mean: 1.9–2.5

Range: 1–8; >98 % are <8

αvβ3-integrin

αIIbβ3 (GPIIb/IIIa)

CD147b

EGFRb

P-selectin

VEGFR2

Angiogenesis (tumor)

Thrombosis

Tumorigenesis

Artery (carotid)

Tumor (prostate)

Tumor (implanted in flank, hind limb or mammary fat pad)

Mouse: 0.84–100 × 106

RT

SQ

Rychak JJ et al., Mol Imaging 2007;6:289

Xuan JW et al., Mol Imaging 2009;8:209

Hu X et al., Invest Radiol 2012;47:398

Sorace AG et al., J Ultrasound Med 2012;31:1543

Knowles JA et al., Arch Otolaryngol Head Neck Surg 2012;138:662

Wang X et al., Circulation 2012;125:3117

Targeson (TS-02-008, bubbles loaded with plasmid DNA)
< div class='tao-gold-member'>

Only gold members can continue reading. Log In or Register to continue

Stay updated, free articles. Join our Telegram channel

Jul 10, 2016 | Posted by in CARDIOLOGY | Comments Off on Molecular Imaging of Inflammation Using Echocardiography. Advances with the Use of Microbubbles

Full access? Get Clinical Tree

Get Clinical Tree app for offline access