Next Article in Journal
Understanding Fenofibrate Release from Bare and Modified Mesoporous Silica Nanoparticles
Next Article in Special Issue
The Evolution and Recent Trends in Acoustic Targeting of Encapsulated Drugs to Solid Tumors: Strategies beyond Sonoporation
Previous Article in Journal
RETRACTED: Zhou et al. Pyrvinium Treatment Confers Hepatic Metabolic Benefits via β-Catenin Downregulation and AMPK Activation. Pharmaceutics 2021, 13, 330
Previous Article in Special Issue
Improved Therapeutic Delivery Targeting Clinically Relevant Orthotopic Human Pancreatic Tumors Engrafted in Immunocompromised Pigs Using Ultrasound-Induced Cavitation: A Pilot Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeted Microbubbles for Drug, Gene, and Cell Delivery in Therapy and Immunotherapy

by
J. Angel Navarro-Becerra
1 and
Mark A. Borden
1,2,*
1
Mechanical Engineering Department, University of Colorado Boulder, Boulder, CO 80309, USA
2
Biomedical Engineering Program, University of Colorado Boulder, Boulder, CO 80309, USA
*
Author to whom correspondence should be addressed.
Pharmaceutics 2023, 15(6), 1625; https://doi.org/10.3390/pharmaceutics15061625
Submission received: 31 March 2023 / Revised: 18 May 2023 / Accepted: 26 May 2023 / Published: 30 May 2023
(This article belongs to the Special Issue Cavitation-Enhanced Drug Delivery and Immunotherapy)

Abstract

:
Microbubbles are 1–10 μm diameter gas-filled acoustically-active particles, typically stabilized by a phospholipid monolayer shell. Microbubbles can be engineered through bioconjugation of a ligand, drug and/or cell. Since their inception a few decades ago, several targeted microbubble (tMB) formulations have been developed as ultrasound imaging probes and ultrasound-responsive carriers to promote the local delivery and uptake of a wide variety of drugs, genes, and cells in different therapeutic applications. The aim of this review is to summarize the state-of-the-art of current tMB formulations and their ultrasound-targeted delivery applications. We provide an overview of different carriers used to increase drug loading capacity and different targeting strategies that can be used to enhance local delivery, potentiate therapeutic efficacy, and minimize side effects. Additionally, future directions are proposed to improve the tMB performance in diagnostic and therapeutic applications.

1. Introduction

Microbubbles (MBs) are gas-filled particles of 1–10 μm in diameter suspended in an aqueous medium. The gas core is highly compressible, making the MB an ideal acoustically responsive agent for ultrasound. However, uncoated MBs are highly unstable owing to their Laplace pressure [1] and tendency to coalesce [2]. Therefore, shells comprising lipid, surfactant, protein, polymer or other materials have been developed to stabilize the MBs for in vitro and in vivo use (Figure 1) [3,4]. Current clinically approved MBs comprise sulfur hexafluoride, perfluorobutane or perfluoropropane gas, and a phospholipid or protein shell (Table 1) [4,5,6]. These commercial MBs have been approved as ultrasound contrast agents for different diagnostic purposes including echocardiography, radiology, and other diagnostic imaging purposes [5,6,7].
Following their clinical translation as ultrasound imaging contrast agents, MB formulations have been engineered as ultrasound-responsive carriers to promote and enhance the local delivery and uptake of a wide variety of drugs [8,9,10,11], genes [12,13,14], and cells [15,16,17,18,19,20] for various therapeutic applications. Many of these applications include the delivery of therapeutic agents to treat the brain [19,20,21,22,23], heart [15,24,25,26], and cancer [27,28,29]. For immunotherapy, MBs can facilitate the delivery of immune cells, cytokines, antigens, and antibodies to promote the activation and infiltration of immune cells from the level of single cells to tissues, organs, and even physiological systems [28,30,31,32]. In cancer, for example, the concept is to induce the modulation and modification of the tumor microenvironment with a subsequent enhanced adaptive immune-cell activation to destroy the primary tumor and its metastases.
At the cellular level, increased uptake of therapeutic molecules into the cytosol has been attributed to sonoporation, i.e., the creation of a transient pore in the cell plasma membrane that increases permeability to molecules into the cytosol [33,34]. This effect is elicited by MB cavitation at the cell surface. MB cavitation refers to volumetric oscillations of the gas core and depends on the frequency, pressure, and waveform of the acoustic pulse. Fortuitously, MBs resonate at clinical ultrasound scanner frequencies, in the 1–10 MHz range. At low acoustic pressures, MBs exhibit stable cavitation, whereby repeated harmonic oscillations induce pushing-and-pulling effects and microstreaming that impart contact and shear forces against the cell membrane (Figure 2A) [35,36]. This mode of operation can be detected by passive acoustic cavitation and cavitation mapping as harmonics of the fundamental diving frequency [37]. At high acoustic pressures, MBs oscillate more violently, leading to rapid collapse and other inertial effects, such as microjets, shockwaves, and fragmentation (Figure 2B) [35,36]. This mode of operation can be detected by passive acoustic cavitation and cavitation mapping as a broadband response, although inertial cavitation can also enhance harmonics [37]. Stable and inertial cavitation can trigger various cellular effects, such as plasma membrane pore formation, cytoskeleton reorganization, transmembrane calcium (Ca2+) influx, potassium (K+) efflux (hyperpolarization), and reactive oxygen species production (Figure 2C). These biological effects lead to enhanced cellular delivery through mechanisms such as drug convection and diffusion through the pores (4–70 kDa) and endocytosis (70–500 kDa) [33,34,35,38].
Table 1. Current clinically approved MB ultrasound contrast agents.
Table 1. Current clinically approved MB ultrasound contrast agents.
Contrast AgentManufacturerIndicationsShellGasConcentration
(MBs/mL)
Size
(Diameter)
Half-Life (min)Volume Dose
(μL/kg) a
Mechanical PropertiesReference
OptisonGE HealthcareLVO/EBDProtein:
HSA
ψ0 = −9.5 to −25.3 mV
C3F85–8 × 1083.0–4.5 μm
(max. 32 μm)
95% < 10 μm
0.5 ± 0.36f = 2–4 MHz
χ = 0.9 N/m
[39,40,41,42,43]
DefinityLantheusLVO/EBD, breast, liver, vascular.Phospholipid:
DPPC, DPPA, DPPE-mPEG5000
ψ0 = −1.1 to −4.2 mV
C3F81.2 × 10101.1–3.3
(max. 20 μm)
98% < 10 μm
2.0 ± 0.310f = 2–6 MHz
χ = 0.5–2.5 N/m
[40,42,43,44,45,46]
SonoVueBraccoLVO/EBD, breast, liver, vascular, urinary tract.Phospholipid:
DSPC, DPPG, PA
ψ0 = −28.3 mV
SF61.5–2.5 × 1081.5–2.5
(max. 20 μm)
99% < 10 μm
1.04 ± 0.1525f = 1.5–2 MHz
χ = 0.2–0.3 N/m
[40,44,47,48,49]
SonazoidGE HealthcareMyocardial perfusion, liver, breast.Phospholipid:
H-EPS
ψ0 = −76 to −82 mV
C4F101.2 × 1091.0–5.0
(max. 10 μm)
99.9% < 7 μm
2.6 ± 0.215f = 4–6 MHz
χ = 0.6 N/m
[50,51,52]
EBD = endocardial border definition; LVO = left ventricular opacification; HSA = human serum albumin; DPPC = 1,2-dipalmitoyl-sn-glycero-3-phosphocholine; DPPA = 1,2-dipalmitoyl-sn-glycero-3-phosphate; DPPE-mPEG5000 = 1,2- dipalmitoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol)-5000]; DSPC = 1,2-distearoyl-sn-glycero-3-phosphocholine; DPPG = 1,2-dipalmitoyl-sn-glycero-3-phospho-(1′-rac-glycerol); PA = palmitic acid; H-EPS = hydrogenated egg phosphatidylserine sodium; ψ0 = zeta potential measured; C3F8 = perfluoropropane; SF6 = sulfur hexafluoride, C4F10 = perfluorobutane; MBs = microbubbles; a Bolus intravenous injection; f = resonance frequency; χ = shell elasticity. Adapted with permission from references [4,5,21].
At the vascular level, increased transport of therapeutic molecules from the vessel lumen into the surrounding tissue parenchyma also occurs by sonoporation, including the mechanisms described above for uptake into the endothelium and then transcytosis or disruption of the endothelium (e.g., rupture of tight junctions to create transport paths between endothelial cells). This effect can be captured by an increase in the Biot number (Bi), which is defined as
B i = K m L t i s s D t i s s
where K m is the mass transfer coefficient, equivalent to the endothelial cell permeability (m/s), and is proportional to the pharmacokinetic first-order rate constant ( K = K m A , where A is the endothelial surface area) between the central and tissue compartments; D t i s s is the drug diffusivity in the tissue; and L t i s s is the diffusion length in tissue, typically the half-length between capillaries (~100 µm). Microbubble sonoporation increases K m in a dose-dependent manner, making diffusion into tissue the limiting transport barrier [53]. Here, the dose includes not only the drug dose, but also the microbubble dose (microbubble volume dose) and ultrasound dose (mechanical index), which can be estimated by the received echo intensity by passive cavitation detection.
Therapeutic agents can be co-administered with MBs or loaded onto the MB shell. This review focuses on shell modification to bioconjugate molecular ligands, drugs (including nucleic acids), nanoparticles and/or cells for therapeutic applications. Articles utilizing so-called “nanobubble” formulations (<1 μm diameter) were excluded from this review. By incorporating specific targeting moieties into the shell, MBs can be engineered to create targeted formulations (tMBs) that exhibit improved pharmacokinetic and acoustic performance, ligand-receptor targeting, biological activation, immunomodulation, and many other functions. With so many options for MB composition, processing, size, microstructure and properties, there is a wealth of opportunity to synthesize innovative targeted MB designs for enhanced performance. Thus, the aim of this review is to summarize the state-of-the-art of current engineered MB formulations and their ultrasound-targeted delivery in different biomedical applications.

2. Microbubble Formulations

2.1. Shell Composition

2.1.1. Phospholipid-Coated Microbubbles

Phospholipid-coated MBs have a thin and soft shell (3–5 nm) (Figure 1) that provides excellent acoustic response for US-contrast imaging. However, for prolonged imaging and drug delivery, the MBs have limited circulation persistence and drug loading capacity [3,54,55]. To optimize MB performance by controlling their structure, properties, targeting, drug loading, and acoustic response, the shell is made up of saturated diacyl phosphatidylcholine lipids (PC) (80–90 mol%) and PEG-lipid emulsifiers (10–20 mol%) [56]. Longer lipid chains increase intermolecular forces between the phospholipids [57], increase shell rigidity [58,59], decrease shell permeability [60,61], and improve MB stability [62,63,64]. This results in superior acoustic stability, stiffer shell elasticity, and reduced fragmentation propensity [62,65]. They also prolong the in vivo half-life [63] and increase the response at the second harmonic frequency [62,65]. Moreover, MBs with long lipid chains can achieve high delivery efficiency for larger molecules [66].
MBs can be quickly eliminated by opsonization, leading to decreased circulation time. To address this issue, buried-ligand architecture has been developed to extend lipid-coated MB circulation time [67,68,69], preserve targeted specificity, and reduce the immunogenic response by blocking complement protein C3b fixation [67,70,71].

2.1.2. Protein-Coated Microbubbles

Protein-coated MBs have a thicker and stiffer layer (15–150 nm) composed of natural proteins (Figure 1). These MBs exhibit moderate rigidity and drug-loading capabilities, as well as good stability and moderate acoustic response [3,72,73]. At present, only Optison is clinically approved and commercially available as a protein-coated MB for UCAs (Table 1). However, the development of novel albumin-tMBs is gaining interest due to their potential for rapid approval and clinical translation as a drug delivery system.
To maximize albumin-MBs’ loading efficiency, the physicochemical properties of drugs play a critical role. Drug loading can be increased by pre-loading the drugs prior to MB synthesis or using drugs with high albumin binding percentages at higher feed ratios [74]. Additionally, chemical modifications such as disulfide bonds and glutaraldehyde cross-linking [75] can link therapeutic molecules or other loaded carriers to albumin-MBs. Physical properties such as size, shape, storage stability, and acoustic response can be tuned by incorporating dextrose [74,76,77], glycerol, and propylene glycol [77], amphipathic molecules [78], polymers [79], PEGylation [80] or increasing the number of thiols groups [81] into the albumin shell. Several in vitro and in vivo studies have employed these methods to propose using of albumin-MBs as targeted carriers [82,83,84,85,86,87] and multimodal contrast agents [84,86,88,89,90,91]. These strategies offer promising options to develop targeted albumin-MB formulations that can enhance their performance in US-image-guided drug delivery applications. Additionally, proteins such as lysozyme and oleosin [92] are alternative candidates that could be used to create new protein-tMB constructs.

2.1.3. Polymer-Coated Microbubbles

Polymer-coated MBs are composed of a thick and stiff layer (50–500 nm) of biodegradable, natural or synthetic polymers (Figure 1). These MBs can encapsulate either hydrophilic and hydrophobic molecules, carry higher drug doses and enhance MB stability during circulation, more so than typical protein and phospholipid MBs. Nevertheless, their incompressibility and rigidity lead to suboptimal US imaging capability and limited acoustic response [3,72,93]. To enhance polymer-MB drug loading capacity and acoustic properties, their chemical composition or polymer molecular weight (MW) is often modified to adjust the shell’s elasticity and thickness. Increasing the drug’s hydrophobicity and molecular weight also boosts loading capacity, stability, and drug release [94]. Proteins [95] or high MW surfactants [96] can also be used to fine-tune MB loading capacity and acoustic response by regulating their stability, size, and shell thickness.

2.2. Gas Composition

The composition of the gas core plays a significant role in the stability of MBs. MBs composed of PFC and SF6 gases, which have lower solubility in blood and higher molecular weight, diffuse more slowly across the MB shell, leading to longer circulation persistence. A recent study has shown that MBs filled with C3F8 or C4F10 were highly stable, exhibited sustained in vivo circulation, and demonstrated more efficient delivery of Evans blue into the brain without causing side effects as compared to SF6-MBs [97].
MBs are now being used as carriers for therapeutic gases, such as NO and O2, to potentially regulate the tumor microenvironment or to enhance immune responses. O2-MBs, for example, have been used to locally release O2 into tumors, thereby improving therapeutic efficacy by reducing hypoxic treatment resistance [98,99]. Additionally, O2-MBs can normalize dysfunctional vessels, enhancing vascular maturity, blood perfusion, and drug penetration [100]. Meanwhile, NO-MBs can control the release of NO to mitigate oxidative stress and apoptosis during ischemia–reperfusion injury [101], enhance the targeted delivery of mesenchymal cells into the infarcted myocardium, and induce regional angiogenic response [15]. In addition, NO-MBs accelerate deep vein thrombosis resolution by reducing platelets and inflammatory cells aggregation, enhancing collagen turnover and stimulating an anticoagulant condition of endothelium [102].
The use of O2- and NO-MBs is still limited due to the properties of the diffusible gases, resulting in lower storage stability and significantly shorter circulation persistence than PFC-MBs. To solve this problem, our group has demonstrated that the lipid shell can be engineered with longer acyl chains (C22:0) to increase O2 payload and enhance local delivery [103]. Additionally, O2-MBs with larger diameters (2–10 µm) have demonstrated an 8-fold increase in half-life compared to smaller diameters (0.5–2 µm) [104]. Recent approaches propose incorporating folate ligands or SPIONs into drug-loaded-O2MBs to accumulate them in a target region by molecular targeting or using an external magnetic field. This incorporation enhances stability, improves drug delivery, and suppresses tumor growth [105,106,107]. Additionally, C4-d-tMBs composed with a gas core combination of PFC/NO have been successfully used to alleviate cardiac allograft rejection by suppressing thrombosis and inflammatory cell infiltrates, while prolonging the survival time two-fold [108].

2.3. Size and Microbubble Volume Dose

The varying components of MBs, including the shell, gas core, size, concentration, and recommended dosage, lead to distinct acoustic responses [109,110], circulation half-life [110,111], and biological effects [111,112,113,114] for both clinically approved MBs (Table 1) and experimental tMB formulations (Table 2, Table 3, Table 4, Table 5 and Table 6). The biological impact of MB acoustic behavior is dose-dependent [115], so selecting dosing strategies for each MB formulation that produce comparable bioeffects is crucial. The MB gas volume dose (MVD) is a unified dose metric that integrates MB size distribution and concentration into a single parameter [22,116]. Our group has demonstrated that MVD can be used to: (1) maximize US imaging contrast and circulation persistence (Figure 3A) [112]; (2) achieve similar, consistent, and comparable molecular delivery across the blood–brain barrier (Figure 3B) [116]; (3) predict, compare, and characterize the MB pharmacokinetics behavior (Figure 3C) [117]; and (4) match the harmonic and broadband cavitation doses regardless of MB size (Figure 3D) [37]. Moreover, matching the gas volume fractions of different commercial MBs with varying sizes and distributions results in equivalent permeabilization effects [118].
The acoustic response of MBs is significantly affected by their size. Changes in size can lead to variations in important parameters such as cavitation threshold [153], radiation force experienced [154,155], and resonance frequency [156,157]. Monodisperse tMB size distributions, combined with acoustic radiation force, have been suggested as a new approach to improve MB targeting, promote ligand–receptor interactions and enhance the sensitivity of US detection in various disease models, including inflammation [158,159,160,161], prostate cancer [162], and fibrosarcoma tumors [67,163]. Our recent results have shown that monodisperse tMBs, driven near their resonance frequency and at low concentrations, can maximize the adhesion efficiency and specificity of tMB [164]. Thus, the use of monodisperse tMB formulations and MVD as a dose metric could be advantageous in future biomedical applications to enhance US-image contrast sensitivity, increase MB adhesion efficiency, maximize MB interaction with endothelial cells, control cavitation behavior, and increase drug delivery efficacy at the target site.

2.4. Methods to Control Microbubble Size Distributions

Regulating the size of MBs is critical to control their acoustic response, generate high contrast and maximize therapeutic efficiency. Therefore, ensuring a narrow size distribution is desirable for future formulations. Although mechanical agitation is currently employed to produce clinically approved MB suspensions, the resulting size distributions are highly polydisperse, which is unsuitable for therapeutic applications. Hence, to engineer monodisperse MB suspensions, research has mainly focused on employing differential centrifugation and microfluidic technologies.

2.4.1. Differential Centrifugation

To obtain stable size-selected MBs with narrow size distributions, our group developed the differential centrifugation method [165]. This method involves two steps: First, highly concentrated, polydisperse MB suspensions (109 to 1010 MBs/mL) are rapidly produced by the sonication method. Second, the MBs are isolated based on their migration in a centrifugal field. The sorting process relies on the application of Stokes’ equation (Equation (2)) which describes the velocity at which a buoyant particle raises relative to the surrounding fluid under conditions of creeping flow:
u i = 2 ( ρ 2 ρ 1 i ) 9 η 2 r i 2 g
where i refers to the MB size index desired for isolation, ri is the MB radius, g is the relative centrifugal force (RCF) required for the MB rise through a column of length L, and η 2 is the effective viscosity.
Size-isolated MB populations obtained by this method can be biochemically targeted with specific ligands for ultrasound molecular purposes [67,68,69,70,164]. Thus, this method represents a simple way to generate size-isolated tMBs suspensions in high yield that can be adopted for therapeutic applications.

2.4.2. Microfluidic Devices

Microfluidic devices, such as flow focusing [2,166,167,168] and T-junctions [169], have emerged as promising one-step methods to generate stable monodisperse MB suspensions in the clinical application range. In these geometries, a gas phase is concentrated between two liquid flows through an orifice. Due to unstable capillary instability, the gas becomes unstable and pinches off, resulting in the release monodisperse MBs. Moreover, these devices, offer precise control over the flow rate, viscosity, and interfacial tension, allowing for the production of MBs with uniform and narrow size distributions at the desired size [170,171,172,173]. However, one must account for a transient Ostwald ripening process when using microfluidic microbubbles [2,167].
The use of microfluidic devices has been proposed for the in situ production of monodisperse drug-loaded MBs for delivery applications. The use of US and MB enhances the drug delivery to ex vivo rat aortic smooth muscle cells, with over 70% of cells internalization under physiological flow and shear stress conditions [174]. Recently, the feasibility of coating monodisperse MBs with liposomes containing quantum dots and a drug model via a biotin-streptavidin linkage using microfluidic devices has been demonstrated [175]. This technology allows the production of the liposome-MB complex at clinically relevant concentrations with high reproducibility. These results suggest the possibility of adapting this method to simultaneously produce future tMB formulations with homogenous ligand and therapeutic molecule distribution, high payloads, and deliver therapeutic agents in real time.

3. Targeted Microbubbles for Therapeutic Applications

MBs can be biochemically targeted by loading their shell with ligands that avidly bind to specific cell receptors, e.g., those overexpressed during disease states. The ligand can be linked to the MB shell by covalent or noncovalent coupling [6,176,177]. The covalent coupling uses PEG-functionalized lipids with thiol-maleimide, disulfide, DBCO-azide, and folate bonds, whereas noncovalent coupling uses electrostatic interactions and biotin-avidin bridge interactions (Figure 4A).
To improve the delivery of drugs, genes, and cells in different biomedical applications, MB targeting can be divided into four strategies: (1) MBs targeted with a therapeutic ligand; (2) drug or gene loaded-MBs targeted with a ligand; (3) drug, gene, or cell loaded-MBs targeted with two different ligands; and (4) MBs conjugated with magnetically or molecularly targeted loaded carriers.

3.1. Microbubbles Targeted with a Therapeutic Ligand

MBs can be loaded with a therapeutic ligand (i.e., where the ligand aids microbubble adhesion and has a therapeutic effect) to improve targeted binding efficiency, spatial delivery, and facilitate real-time imaging of tissue structure, function, and molecular characteristics in neurodegenerative and cardiovascular diseases (Table 2). Zhu et al. [20] demonstrated that administering anti-Aβ-tMBs with neural stem cells is a safe and effective dual delivery approach to accelerate Aβ plaque clearance, increase BDNF expression, and improve learning and spatial memory function impairment in Alzheimer’s disease. Similarly, Gao et al. [119] showed that injecting anti-SDF-1-tMBs with mesenchymal stem cells can simultaneously monitor real-time renal perfusion, promote cell homing, and repair early diabetic neuropathy kidneys. Using GPVI-tMBs [120] and anti-IL8-tMBs [121] can aid in diagnosing atherosclerotic lesions, slowing down atheroprogression, reducing inflammation, and enhancing plaque stability. Additionally, administering nitric oxide-filled MBs that are targeted with anti-Cd4 can help alleviate cardiac allograft rejection by suppressing thrombosis and inflammatory cell infiltrates, while prolonging the survival time two-fold [108].

3.2. Drug or Gene Loaded-MBs Targeted with a Ligand

Drug- or gene-loaded MB can be targeted with a ligand (here, just to aid microbubble adhesion) to enable site-specific delivery, enhance therapeutic efficacy, and enable image-guided treatment (Figure 4B). Antibody/peptide-tMBs loaded with various genes and chemotherapeutic drugs (Table 3) have been proposed in several studies as successful treatments for various tumors, including glioma [122,123], hind limb [124], and breast tumor [127]. Targeted delivery in these therapies acts as both a molecular contrast agent and a delivery system, improving tumor-specific targeting without harming normal tissues, and increasing local agent delivery, circulation times, and median survival times in mouse and rat models.
A recent study has suggested using folate-tMBs loaded with oxygen and paclitaxel as a dual-targeted system for ovarian tumor cells and tumor-associated macrophages, resulting in two-fold increase in drug concentration, apoptosis index, and median survival [105]. Similarly, folate-tMB loaded with pFLuc enhanced gene transfection efficiency in the brain by 1.5-fold higher than MBs without folate conjugation [129].
On the other hand, different studies have demonstrated that RGD-tMBs loaded with fibrinolytic agents, such as urokinase [125] and tPA [128], can improve arterial blood flow and reduce thrombus size locally without causing adverse side effects, making it a promising and safe option for thrombolysis therapy. Furthermore, VCAM-1-tMBs loaded with mir-126 can serve as a non-invasive and risk-free anti-inflammatory therapy to prevent the growth of abdominal aortic aneurysms and assess the inflammatory state of the endothelium [126].

3.3. Drug-, Gene- or Cell-Loaded MBs Targeted with Two Different Ligands

Drug-, gene- or cell-loaded MB can be targeted with two different ligands (dual-tMBs) to achieve higher adhesion efficiency, enhance ultrasound contrast signal, and have more control over drug delivery at the target site (Figure 4C, Table 4). Adipose-derived stem cells conjugated with CD90/ICAM-1-tMBs, known as StemBells, have been proposed as an image-guided treatment for myocardial infarction (MI) [130] and to prevent atherosclerosis acceleration post-MI [131]. Studies have shown that StemBells can improve cardiac function, reach the infarct area without complications, increase cap thickness, decrease intra-plaque macrophage density, and promote the presence of anti-inflammatory macrophages and chemokines in both the plaque and infarcted myocardium, as well as in circulating monocytes.
Additionally, dual-tMBs loaded with genes or chemotherapeutic agents have shown superior performance in treating breast and pancreatic cancer compared to those conjugated with only one ligand [132,133,134]. These studies have shown that dual-tMBs had 2–3 times higher adhesion efficiency, resulting in a 2–4 times greater ultrasound contrast signal for tumor visualization. Additionally, localized release has improved the accumulation of molecules in the tumor region, resulting in tumor suppression 1.5–4 times greater. This has led to reduced side-effects and better drug tolerability.

3.4. MBs Conjugated with Magnetically or Molecularly Targeted Loaded Carriers

MBs can be conjugated with other magnetically [106,107,135,136,137,138,139] or molecularly [140,141,142,143] targeted drug/gene-loaded carriers such as nanoparticles, liposomes, and nanodroplets to increase drug loading capacity, stability in circulation, and delivery efficiency (Figure 4D, Table 5). Drug-loaded superparamagnetic iron oxide nanoparticles (SPIONs) conjugated with MBs have been developed to control drug delivery in brain tumors through magnetic targeting and enabling direct delivery visualization using magnetic resonance imaging (MRI) [107,136]. This conjugation provided significant superparamagnetic/acoustic properties for imaging, resulting in a two-fold increase in MB half-life and US contrast signal, as well as 2–4 times more drug and SPION deposition in the tumor region, leading to enhanced MRI signals.
Recently, a magnetically responsive and ultrasound-sensitive delivery system, called doxorubicin-loaded magneto-liposome MBs, has been proposed for enhanced therapy of pancreatic tumors [139]. This combination resulted in a 1.5-fold improvement in drug penetration at the tumor site and a 2-fold increase in tumor suppression compared to the control group without magnetic targeting. Additionally, formulations of drug- or gene- molecularly targeted liposomes conjugated to MBs have shown the potential to enhance the cytotoxic effects in breast, glioma, and colorectal cancer treatment [140,141,142], by increasing the concentration of encapsulated drug in circulation, promoting drug accumulation in tumors, and reducing toxicity in normal tissues. Moreover, liposomes conjugated to MBs, which are molecularly and magnetically responsive, can increase the targeting efficiency to tumor neovasculature, enhance MRI, enhance US tumor imaging by 2.5-fold, prolong the MB half-life by 5-times, and decrease the tumor growth by 2-fold [138].

4. Targeted Microbubbles for Immunotherapy Applications

The combination of MBs with US has demonstrated promising potential in modulating and modifying the tumor microenvironment. This is achieved by promoting the penetration of immunotherapeutic agents, enhancing blood perfusion, increasing therapeutic delivery, and inducing tumor cell death [28,30,31,32]. To further enhance the performance of MBs in cancer immunotherapy, the use of loaded-tMBs with drugs, genes or cells has emerged as an attractive strategy. This approach enables precise control of immune stimulation, and it enhances the delivery and pharmacokinetics of immunomodulatory agents at the target site. This strategy has been shown for various cancer immunotherapy modalities, including monoclonal antibodies, immune checkpoint inhibitors, adoptive cell transfer, cytokine therapy, and vaccines (Table 6).

4.1. Monoclonal Antibody Immunotherapy

Monoclonal antibody immunotherapy aims to induce cell death by targeting specific antigens, sequences, or epitopes expressed at the disease target site [178]. Therapeutic monoclonal antibodies (mAb) can be administered unconjugated or conjugated with chemotherapeutic drugs and radioisotopes to target tumors and minimize the toxicity effects of conventional chemotherapy [179]. Although mAb immunotherapy has potential therapeutic benefits, poor penetration and heterogenous distribution can impact the therapeutic effectiveness [178,179]. High concentrations are often required, which can result in adverse side effects due to the rapid metabolism and clearance rate through the kidneys [180]. mAb-tMBs + US have been proposed as an image-guided delivery method to increase targeting, enhance local penetration, and potentiate the therapeutic effect of mAbs in different cancer therapies (Table 6). Liao et al. [144] improved glioma treatment by administrating EGFR-tMB + US, resulting in increased tumor vessel permeability and enhanced tumor-suppressing effect 7 days after treatment, with no tumor regrowth in the following 10 days. Kang et al. [145] found that combining anti-DLL4-tMBs + US with DAPT for gastric tumor therapy was more effective than DAPT alone, showing synergistic antitumor proapoptotic effects. These effects were attributed to the regulation of apoptosis-related proteins Bcl-2 and BAX, as well as the tumor suppressor protein P53. Recently, Sun et al. [146] developed a tMB construct that delivers pyropheophorbide sensitizer and therapeutic trastuzumab mAbs for targeted combination of sonodynamic and antibody therapies in gastric cancer. The therapy resulted in enhanced antibody accumulation at the tumor site, increased tumor cell apoptosis and tumor growth inhibition by suppressing AKT phosphorylation.

4.2. Immune Checkpoint Inhibitor Therapy

Immune checkpoint inhibitors (ICI) are mAbs variants that activate T cells by blocking immune checkpoint receptors [178]. Approved ICIs for therapy include programmed cell death protein-1 (PD-1), programmed death ligand-1 (PD-L1), and cytotoxic T-lymphocyte-associated antigen-4 (CTLA-4) [178,181]. However, systemic ICI therapy is limited by severe side effects associated with dosage, low treatment response, and overactivation of the immune response [182]. To address these limitations, a controlled delivery strategy called PD-L1 mAb–tMBs + US has been proposed (Table 6). Kim et al. [147] reported that PD-L1 mAb–tMBs can improve therapeutic efficacy, increase the therapeutic index, reduce toxicity, and avoid immune responses and fatalities associated with PD-L1 mAb systemic administration in the treatment of colon cancer. Ma et al. [148] and Liu et al. [149] both demonstrated that combining PD-L1 mAb–tMBs with chemotherapeutic drugs or loading PD-L1 mAb–tMBs with genes exhibited strong synergistic effects in inhibiting cervical tumor growth, improving survival rates, and reducing side effects compared to using either drug/gene or PD-L1 mAb–tMBs alone. The combination treatment showed better immunological activity, indicated by increased CD8+ T cell infiltration and cytokine expression, ultimately resulting in an effective antitumor immune killing effect.

4.3. Adoptive Cell Immunotherapy

Adoptive cell-mediated immunotherapy (ACT) involves the intravenous transfer in vitro of resident T cells or genetically modified T cells to target tumor antigens and mediate anti-tumor function. The three types of ACT are tumor-infiltrating lymphocytes (TIL), T cell receptor (TCR) gene therapy, and chimeric antigen receptor-modified T cells (CAR-T) [178,183]. However, ACT is limited by the lack of in vivo persistence of transferred cells, toxicities related to lymphodepletion, immune response, and cytokine release [183]. Preliminary studies using MBs targeted with specific antibodies and retroviruses (CD3, CD8, CD45RA, CD62L, CD3/CD28, and CD-19CAR) have shown that tMBs represent a potential method to stimulate [184,185], activate, transduce, and precisely sort specific phenotypes of CAR-T cells [184]. These stimulations lead to greater in vivo persistence, decreased toxicity, and improved antitumor response of adoptively transferred CAR-T cells when compared to CAR-T cells obtained through conventional methods [185].

4.4. Cytokine Immunotherapy

Cytokine-mediated immunotherapy involves the systemic administration of cytokines to enhance the immune response [178]. Commonly used cytokines for immunotherapy in clinics and research include Interferon-alpha (IFN-α), Interleukins (ILs), and granulocyte-macrophage colony-stimulating factor (GM-CSF) [178,186,187]. However, this therapy has several limitations, including low efficacy, high levels of toxicity, and immune response activation [188]. IL-27-tMBs have shown promising results by enhancing cytokine bioactivity, inhibiting prostate tumor growth, and efficiently improving the recruitment of natural killer cells (NKT) and CD8+ cells to the tumor compared to untargeted delivery [189]. Additionally, IL-16–tMBs [190] have been introduced to evaluate myocardial ischemia–reperfusion, detect atherosclerosis, and detect ovarian tumors, respectively.

4.5. Vaccine Immunotherapy

Vaccine-mediated immunotherapy involves administering specific antigens or protein fragments to stimulate an immune response [178]. Different types of vaccines are used, including peptide-based vaccines [191], DNA-based vaccines [192], and cell-based vaccines such as NK cells, dendritic cells (DC), and CAR-T cells [193]. However, the effectiveness of vaccine therapy is limited by tissue-specific antigens, low humoral responses, and heterogenous immune responses [178,194]. Gene- and protein-loaded MBs targeted with tumor-specific antigens offer promising image-guided vaccine immunotherapy for breast and melanoma tumors, as demonstrated by Li et al. [150] and Gao et al. [151]. This approach can increase delivery efficiency, prolong survival rates, activate systemic antitumor immunity, inhibit and delay tumor growth [150,151], reduce systemic toxicity, and inhibit cancer metastasis by bridging the innate and adaptive immune responses [150]. Recently, Jungio et al. [152] introduced the first cell-free vaccine in the form of tMBs with activated DC plasma membranes to enhance breast cancer tumor targeting, reduce tumor growth, and increase survival rates. Studies also showed tumor growth inhibition and/or antigen-specific protection through DCs activated by mRNA-loaded MBs [195] or tMBs conjugated with NK cells to promote the controlled delivery at the target site [196].

5. Challenges of tMBs in Therapy and Immunotherapy

Table 2, Table 3, Table 4, Table 5 and Table 6 highlight the potential of combining drug-, gene- or cell-loaded tMBs with US for treating various cancers and cardiovascular diseases. However, before exploring future clinical applications, it is crucial to address the in vivo safety of tMBs. The diversity of MB compositions, ligands, therapeutic molecules, and US settings applied pose a significant challenge for clinical translation. While the studies report enhanced therapeutic uptake associated with MB cavitation without side effects, evidence of the latter is lacking. Depending on the US and MB parameters, cavitation can induce adverse bioeffects, such as endothelial cell damage, endothelial dysfunction [197,198], vascular rupture [199,200], and petechial hemorrhages [66,201,202,203]. Moreover, acoustic cavitation can cause long-term side effects on the target tissues [35,197]. Therefore, a deeper understanding of how cavitation events interact with cells and tissues, along with the resulting cellular and molecular responses, can lead to the development of novel design strategies that improve treatment efficacy while minimizing potential safety concerns.
The MB doses and US parameters used in current studies are based on the reference dose of clinically approved MBs, but comparing bioeffects among different MB formulations and animal models is challenging due to the differences in size distribution, dose, shell composition, gas core, and half-life [66,109,111,113,114,204]. Strategies for approximating equivalent MB or cavitation doses are necessary to improve understanding of induced bioeffects. While US and MB doses are critical factors in comparing bioeffects, physiological variations also significantly impact biological outcomes [22]. Therefore, understanding MB pharmacokinetics and accounting for physiological variability is essential for developing safe and consistent treatment protocols.
On the other hand, the reported experimental therapy durations are typically shorter than those in clinic applications and success is limited to reporting tumor growth suppression, tumor cell apoptosis, and survival time. However, MB cavitation can trigger the activation of immune cells, cytokine secretion or protein production, which can modulate and modify the tumor microenvironment to fight against cancer [28,30,31,32]. These immune cells, such as T cells, macrophages, neutrophils, fibroblasts, and B cells, participate in the tumor suppression [205]. Therefore, to explain the molecular mechanisms associated with tumor suppression in detail, it is desirable to extend the analysis to characterize the activation and production of these immune components during therapy and immunotherapy protocols. Moreover, the evaluation of only one or four doses of tMB formulations is insufficient, and further pharmacological and toxicological experiments are necessary to identify the optimal dose, administration frequency and protocols, as well as the US parameters, to increase the delivery and maximize therapeutic effects in future studies.
The majority of experimental studies have used murine-origin monoclonal antibodies linked to the MB surface via avidin and biotin coupling, which could potentially cause severe immune responses in human patients [206,207]. To eliminate immunogenic effects, future studies should consider using humanized antibodies, antibody fragments or peptides linked by covalent binding.
Finally, the issue of MB dose is oftentimes ambiguous. We have proposed the use of microbubble volume dose (MVD) as a unifying dose metric that combines MB size distribution and concentration dose to provide a useful metric for correlating pharmacokinetic parameters (e.g., half-life), acoustic response, and the amount of drug delivered to target tissue (Figure 3) [116]. For approved, commercially available MB ultrasound contrast agents, the MVD ranges from 6 to 25 µL/kg (Table 1). However, for targeted MBs, the MVD typically ranges from 1 to 50 µL/kg, although some reports have very low MVD (<0.1 µL/kg), while others have very high MVD (>100 µL/kg). Note that these MVDs were calculated by multiplying the dose given to the animal, usually in number of MBs per animal weight (although sometimes animal weights were not reported, so we had to use average for the species and age). Importantly, there is often a significant difference between the number-weighted distribution (Figure 5D) and volume-weighted distribution (Figure 5E). Note the difference in mean diameters, for example (Figure 5G), even for monodisperse size-isolated microbubbles. For example, as a basis of 1010 MB/mL, the volume fractions estimated by the number and volume distributions are quite different (208, 350, and 698 µL/kg vs. 406, 699, and 1170 µL/kg for 2, 3, and 5 µm diameter MBs, respectively). Additionally, the size distribution is often skewed from a normal distribution, even for monodisperse MBs, owing to Ostwald ripening and other colloidal mechanisms. Therefore, one should take the area-under-the-curve of the MB volume vs. concentration plot (Figure 5F) to determine the volume fraction ( M B ), which can then be used to determine an accurate MVD [117]. Therefore, a major challenge in the field is for researchers to report an accurate MVD for their tMBs, or at least the combination of dose (number of MBs/kg body weight) and volume-weighted mean diameter.

6. Future Directions

This review is focused on tMB formulations, which include direct bioconjugation of a ligand, drug, nanoparticle and/or cell onto the MB shell. One may argue that loading onto the MB is more efficient because the payload is administered at a lower dose and is co-localized with the cavitation event. However, it is currently thought that the regulatory burden of approving a specific MB/drug combination for clinical use makes this approach less financially viable. Additionally, MBs are cleared quickly from circulation (typically 5–10 min) with accumulation primarily in the lung, liver, kidney, and spleen, leading to loss and rapid elimination of the therapeutic cargo before it can become available to the target cells [9,34]. Therefore, current clinical trials favor co-administration of commercial MBs (Table 1) and free drug, either as a single cocktail bolus or a sequence of injections. This co-administration without bioconjugation approach is more amenable as a platform, where a single MB formulation could be used for multiple drug/disease indications. However, this simplistic approach also limits the creativity and potential benefits of MB engineering, several examples of which are reviewed here (Table 2, Table 3, Table 4, Table 5 and Table 6). Therefore, it is advantageous to continue to develop novel targeted microbubble formulations and applications.
As discussed in this review, targeted MBs offer many benefits as innovative agents for ultrasound image-guided drug, gene, and cell delivery in therapy and immunotherapy. While research continues on targeted MBs for research purposes in animal models, more work is necessary to improve their safety, efficacy, and financial viability for clinical translation. Further research and advancements should be made in the following areas:
  • The design and characterization of tMB formulations to achieve high payload capacity, stable drug loading, homogenous and reproducible size distributions and colloidal stability, and strong echogenicity and ultrasound responsiveness.
  • The understanding of tMB pharmacokinetics with the goal of extending circulation time and improving biodistribution, and determination of minimum effective dose and maximum tolerated dose for a given drug and US scheme.
  • The evaluation of tMB interactions with cells and tissues, their correlation to short- and long-term bioeffects in vivo, and molecular description of biological mechanisms induced by tMB cavitation.
  • The optimization of US protocols with consideration of tMB pharmacokinetics and bioeffects to ensure treatment safety and efficacy.
  • The establishment of drug, microbubble, and ultrasound dose metrics to compare therapeutic index between tMB formulations.
Innovations in these areas may help reduce over-reliance on commercially available ultrasound contrast agents and promote clinical translation of engineered tMBs.
However, to justify investment into the clinical translation of tMBs, researchers must show a significant improvement in safety and/or efficacy in comparison to currently approved and commercially available MB formulations (see Table 1). For example, it would be helpful to show an increase in therapeutic index for delivery of the target payload using tMBs vs. commercial MBs (vs. systemic administration without MBs). Additionally, it would be helpful for comparison to report dose optimization in terms of microbubble MVD and pharmacokinetics, along with ultrasound MI, duty cycle, and sonication time. Such characterization and reporting would help the community standardize treatments and optimize passive cavitation monitoring methods.

7. Conclusions

This review highlights recent developments in targeted MB formulations engineered for drug and cell delivery for therapy and immunotherapy. Basic MB formulations are described prior to citing some recent innovative designs for targeted MBs and their applications. The bioconjugation approach has demonstrated promise in enhancing the delivery of various drugs, genes, and cells to target tissues. Issues for future research are discussed, including the need to better define tMB pharmacokinetics and bioeffects, as well as the standardization of MB and US parameters to compare therapeutic index between different MB formulations.

Author Contributions

Conceptualization J.A.N.-B. and M.A.B.; investigation J.A.N.-B.; writing—original draft preparation, J.A.N.-B.; writing—review and editing, M.A.B.; visualization J.A.N.-B.; supervision, M.A.B.; project administration, M.A.B.; funding acquisition, M.A.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the US National Institutes of Health award R01CA239465.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Epstein, P.S.; Plesset, M.S. On the Stability of Gas Bubbles in Liquid-Gas Solutions. J. Chem. Phys. 1950, 18, 1505–1509. [Google Scholar] [CrossRef]
  2. Segers, T.; Lohse, D.; Versluis, M.; Frinking, P. Universal Equations for the Coalescence Probability and Long-Term Size Stability of Phospholipid-Coated Monodisperse Microbubbles Formed by Flow Focusing. Langmuir 2017, 33, 10329–10339. [Google Scholar] [CrossRef]
  3. Sirsi, S.R.; Borden, M.A. Microbubble Compositions, Properties and Biomedical Applications. Bubble Sci. Eng. Technol. 2009, 1, 3–17. [Google Scholar] [CrossRef]
  4. Borden, M.A.; Song, K.-H. Reverse Engineering the Ultrasound Contrast Agent. Adv. Colloid Interface Sci. 2018, 262, 39–49. [Google Scholar] [CrossRef]
  5. Frinking, P.; Segers, T.; Luan, Y.; Tranquart, F. Three Decades of Ultrasound Contrast Agents: A Review of the Past, Present and Future Improvements. Ultrasound Med. Biol. 2020, 46, 892–908. [Google Scholar] [CrossRef]
  6. Borden, M.A.; Dayton, P.A.; Slagle, C.; Walmer, R.W. Ultrasound Contrast Agents. In Molecular Imaging; Elsevier: Amsterdam, The Netherlands, 2021; pp. 639–653. ISBN 978-0-12-816386-3. [Google Scholar]
  7. Stride, E.; Segers, T.; Lajoinie, G.; Cherkaoui, S.; Bettinger, T.; Versluis, M.; Borden, M. Microbubble Agents: New Directions. Ultrasound Med. Biol. 2020, 46, 1326–1343. [Google Scholar] [CrossRef]
  8. Sennoga, C.A.; Kanbar, E.; Auboire, L.; Dujardin, P.-A.; Fouan, D.; Escoffre, J.-M.; Bouakaz, A. Microbubble-Mediated Ultrasound Drug-Delivery and Therapeutic Monitoring. Expert Opin. Drug Deliv. 2017, 14, 1031–1043. [Google Scholar] [CrossRef]
  9. Lammertink, B.H.A.; Bos, C.; Deckers, R.; Storm, G.; Moonen, C.T.W.; Escoffre, J.-M. Sonochemotherapy: From Bench to Bedside. Front. Pharmacol. 2015, 6, 138. [Google Scholar] [CrossRef]
  10. He, J.; Liu, Z.; Zhu, X.; Xia, H.; Gao, H.; Lu, J. Ultrasonic Microbubble Cavitation Enhanced Tissue Permeability and Drug Diffusion in Solid Tumor Therapy. Pharmaceutics 2022, 14, 1642. [Google Scholar] [CrossRef]
  11. Escoffre, J.-M.; Sekkat, N.; Oujagir, E.; Bodard, S.; Mousset, C.; Presset, A.; Chautard, R.; Ayoub, J.; Lecomte, T.; Bouakaz, A. Delivery of Anti-Cancer Drugs Using Microbubble-Assisted Ultrasound in Digestive Oncology: From Preclinical to Clinical Studies. Expert Opin. Drug Deliv. 2022, 19, 421–433. [Google Scholar] [CrossRef]
  12. Negishi, Y.; Endo-Takahashi, Y.; Maruyama, K. Gene Delivery Systems by the Combination of Lipid Bubbles and Ultrasound. DDT 2016, 10, 248–255. [Google Scholar] [CrossRef]
  13. Sirsi, S.R.; Borden, M.A. Advances in Ultrasound Mediated Gene Therapy Using Microbubble Contrast Agents. Theranostics 2012, 2, 1208–1222. [Google Scholar] [CrossRef]
  14. Rychak, J.J.; Klibanov, A.L. Nucleic Acid Delivery with Microbubbles and Ultrasound. Adv. Drug Deliv. Rev. 2014, 72, 82–93. [Google Scholar] [CrossRef]
  15. Tong, J.; Ding, J.; Shen, X.; Chen, L.; Bian, Y.; Ma, G.; Yao, Y.; Yang, F. Mesenchymal Stem Cell Transplantation Enhancement in Myocardial Infarction Rat Model under Ultrasound Combined with Nitric Oxide Microbubbles. PLoS ONE 2013, 8, e80186. [Google Scholar] [CrossRef]
  16. Toma, C.; Fisher, A.; Wang, J.; Chen, X.; Grata, M.; Leeman, J.; Winston, B.; Kaya, M.; Fu, H.; Lavery, L.; et al. Vascular Endoluminal Delivery of Mesenchymal Stem Cells Using Acoustic Radiation Force. Tissue Eng. Part A 2011, 17, 1457–1464. [Google Scholar] [CrossRef]
  17. Sun, T.; Gao, F.; Li, X.; Cai, Y.; Bai, M.; Li, F.; Du, L. A Combination of Ultrasound-Targeted Microbubble Destruction with Transplantation of Bone Marrow Mesenchymal Stem Cells Promotes Recovery of Acute Liver Injury. Stem Cell Res 2018, 9, 356. [Google Scholar] [CrossRef]
  18. Zhu, T.; Huang, X.; Peng, S.; Ye, Y.; Zhao, J. Ultrasound Targeted Microbubble Destruction Promotes the Therapeutic Effect of HUMSC Transplantation on Glaucoma-Caused Optic Nerve Injury in Rabbits. Trans. Vis. Sci. Tech. 2022, 11, 12. [Google Scholar] [CrossRef]
  19. Cui, H.; Zhu, Q.; Xie, Q.; Liu, Z.; Gao, Y.; He, Y.; Tan, X.; Xu, Y. Low Intensity Ultrasound Targeted Microbubble Destruction Assists MSCs Delivery and Improves Neural Function in Brain Ischaemic Rats. J. Drug Target. 2020, 28, 320–329. [Google Scholar] [CrossRef]
  20. Zhu, Q.; Xu, X.; Chen, B.; Liao, Y.; Guan, X.; He, Y.; Cui, H.; Rong, Y.; Liu, Z.; Xu, Y. Ultrasound-targeted Microbubbles Destruction Assists Dual Delivery of Beta-amyloid Antibody and Neural Stem Cells to Restore Neural Function in Transgenic Mice of Alzheimer’s Disease. Med. Phys. 2022, 49, 1357–1367. [Google Scholar] [CrossRef]
  21. Schoen, S.; Kilinc, M.S.; Lee, H.; Guo, Y.; Degertekin, F.L.; Woodworth, G.F.; Arvanitis, C. Towards Controlled Drug Delivery in Brain Tumors with Microbubble-Enhanced Focused Ultrasound. Adv. Drug Deliv. Rev. 2022, 180, 114043. [Google Scholar] [CrossRef]
  22. Song, K.-H.; Harvey, B.K.; Borden, M.A. State-of-the-Art of Microbubble-Assisted Blood-Brain Barrier Disruption. Theranostics 2018, 8, 4393–4408. [Google Scholar] [CrossRef]
  23. Wang, J.; Li, Z.; Pan, M.; Fiaz, M.; Hao, Y.; Yan, Y.; Sun, L.; Yan, F. Ultrasound-Mediated Blood–Brain Barrier Opening: An Effective Drug Delivery System for Theranostics of Brain Diseases. Adv. Drug Deliv. Rev. 2022, 190, 114539. [Google Scholar] [CrossRef]
  24. Unger, E.; Porter, T.; Lindner, J.; Grayburn, P. Cardiovascular Drug Delivery with Ultrasound and Microbubbles. Adv. Drug Deliv. Rev. 2014, 72, 110–126. [Google Scholar] [CrossRef]
  25. Qian, L.; Thapa, B.; Hong, J.; Zhang, Y.; Zhu, M.; Chu, M.; Yao, J.; Xu, D. The Present and Future Role of Ultrasound Targeted Microbubble Destruction in Preclinical Studies of Cardiac Gene Therapy. J. Thorac. Dis. 2018, 10, 1099–1111. [Google Scholar] [CrossRef]
  26. Chen, H.H.; Matkar, P.N.; Afrasiabi, K.; Kuliszewski, M.A.; Leong-Poi, H. Prospect of Ultrasound-Mediated Gene Delivery in Cardiovascular Applications. Expert Opin. Biol. Ther. 2016, 16, 815–826. [Google Scholar] [CrossRef]
  27. Wanigasekara, J.; de Carvalho, A.M.A.; Cullen, P.J.; Tiwari, B.; Curtin, J.F. Converging Technologies: Targeting the Hallmarks of Cancer Using Ultrasound and Microbubbles. Trends Cancer 2021, 7, 886–890. [Google Scholar] [CrossRef]
  28. Tu, J.; Zhang, H.; Yu, J.; Liufu, C.; Chen, Z. Ultrasound-Mediated Microbubble Destruction: A New Method in Cancer Immunotherapy. OTT 2018, 11, 5763–5775. [Google Scholar] [CrossRef]
  29. Li, H.; Zhang, Y.; Shu, H.; Lv, W.; Su, C.; Nie, F. Highlights in Ultrasound-Targeted Microbubble Destruction-Mediated Gene/Drug Delivery Strategy for Treatment of Malignancies. Int. J. Pharm. 2022, 613, 121412. [Google Scholar] [CrossRef]
  30. Han, Y.; Sun, J.; Wei, H.; Hao, J.; Liu, W.; Wang, X. Ultrasound-Targeted Microbubble Destruction: Modulation in the Tumor Microenvironment and Application in Tumor Immunotherapy. Front. Immunol. 2022, 13, 937344. [Google Scholar] [CrossRef]
  31. Ho, Y.-J.; Li, J.-P.; Fan, C.-H.; Liu, H.-L.; Yeh, C.-K. Ultrasound in Tumor Immunotherapy: Current Status and Future Developments. J. Control. Release 2020, 323, 12–23. [Google Scholar] [CrossRef]
  32. Omata, D.; Munakata, L.; Maruyama, K.; Suzuki, R. Ultrasound and Microbubble-Mediated Drug Delivery and Immunotherapy. J. Med. Ultrason. 2022. Online ahead of print. [Google Scholar] [CrossRef]
  33. Yang, Y.; Li, Q.; Guo, X.; Tu, J.; Zhang, D. Mechanisms Underlying Sonoporation: Interaction between Microbubbles and Cells. Ultrason. Sonochem. 2020, 67, 105096. [Google Scholar] [CrossRef]
  34. Deprez, J.; Lajoinie, G.; Engelen, Y.; De Smedt, S.C.; Lentacker, I. Opening Doors with Ultrasound and Microbubbles: Beating Biological Barriers to Promote Drug Delivery. Adv. Drug Deliv. Rev. 2021, 172, 9–36. [Google Scholar] [CrossRef]
  35. Qin, P.; Han, T.; Yu, A.C.H.; Xu, L. Mechanistic Understanding the Bioeffects of Ultrasound-Driven Microbubbles to Enhance Macromolecule Delivery. J. Control. Release 2018, 272, 169–181. [Google Scholar] [CrossRef]
  36. Kooiman, K.; Vos, H.J.; Versluis, M.; de Jong, N. Acoustic Behavior of Microbubbles and Implications for Drug Delivery. Adv. Drug Deliv. Rev. 2014, 72, 28–48. [Google Scholar] [CrossRef]
  37. Martinez, P.; Bottenus, N.; Borden, M. Cavitation Characterization of Size-Isolated Microbubbles in a Vessel Phantom Using Focused Ultrasound. Pharmaceutics 2022, 14, 1925. [Google Scholar] [CrossRef]
  38. Wischhusen, J.; Padilla, F. Ultrasound-Targeted Microbubble Destruction (UTMD) for Localized Drug Delivery into Tumor Tissue. IRBM 2019, 40, 10–15. [Google Scholar] [CrossRef]
  39. Chatterjee, D.; Sarkar, K. A Newtonian Rheological Model for the Interface of Microbubble Contrast Agents. Ultrasound Med. Biol. 2003, 29, 1749–1757. [Google Scholar] [CrossRef]
  40. Ja’afar, F.; Leow, C.H.; Garbin, V.; Sennoga, C.A.; Tang, M.-X.; Seddon, J.M. Surface Charge Measurement of SonoVue, Definity and Optison: A Comparison of Laser Doppler Electrophoresis and Micro-Electrophoresis. Ultrasound Med. Biol. 2015, 41, 2990–3000. [Google Scholar] [CrossRef]
  41. Shi, W.T.; Forsberg, F. Ultrasonic Characterization of the Nonlinear Properties of Contrast Microbubbles. Ultrasound Med. Biol. 2000, 26, 93–104. [Google Scholar] [CrossRef]
  42. Tan, J.-K.Y.; Pham, B.; Zong, Y.; Perez, C.; Maris, D.O.; Hemphill, A.; Miao, C.H.; Matula, T.J.; Mourad, P.D.; Wei, H.; et al. Microbubbles and Ultrasound Increase Intraventricular Polyplex Gene Transfer to the Brain. J. Control. Release 2016, 231, 86–93. [Google Scholar] [CrossRef]
  43. Fix, S.M.; Nyankima, A.G.; McSweeney, M.D.; Tsuruta, J.K.; Lai, S.K.; Dayton, P.A. Accelerated Clearance of Ultrasound Contrast Agents Containing Polyethylene Glycol Is Associated with the Generation of Anti-Polyethylene Glycol Antibodies. Ultrasound Med. Biol. 2018, 44, 1266–1280. [Google Scholar] [CrossRef]
  44. Tu, J.; Swalwell, J.E.; Giraud, D.; Cui, W.; Chen, W.; Matula, T.J. Microbubble Sizing and Shell Characterization Using Flow Cytometry. IEEE Trans. Ultrason. Ferroelect. Freq. Contr. 2011, 58, 955–963. [Google Scholar] [CrossRef] [PubMed]
  45. Helfield, B.L.; Goertz, D.E. Nonlinear Resonance Behavior and Linear Shell Estimates for DefinityTM and MicroMarkerTM Assessed with Acoustic Microbubble Spectroscopy. J. Acoust. Soc. Am. 2013, 133, 1158–1168. [Google Scholar] [CrossRef]
  46. Chatterjee, D.; Sarkar, K.; Jain, P.; Schreppler, N.E. On the Suitability of Broadband Attenuation Measurement for Characterizing Contrast Microbubbles. Ultrasound Med. Biol. 2005, 31, 781–786. [Google Scholar] [CrossRef]
  47. Tu, J.; Guan, J.; Qiu, Y.; Matula, T.J. Estimating the Shell Parameters of SonoVue ® Microbubbles Using Light Scattering. J. Acoust. Soc. Am. 2009, 126, 2954–2962. [Google Scholar] [CrossRef]
  48. Gorce, J.-M.; Arditi, M.; Schneider, M. Influence of Bubble Size Distribution on the Echogenicity of Ultrasound Contrast Agents: A Study of SonoVue? Investig. Radiol. 2000, 35, 661–671. [Google Scholar] [CrossRef]
  49. Wu, S.-K.; Chu, P.-C.; Chai, W.-Y.; Kang, S.-T.; Tsai, C.-H.; Fan, C.-H.; Yeh, C.-K.; Liu, H.-L. Characterization of Different Microbubbles in Assisting Focused Ultrasound-Induced Blood-Brain Barrier Opening. Sci. Rep. 2017, 7, 46689. [Google Scholar] [CrossRef]
  50. Sarkar, K.; Shi, W.T.; Chatterjee, D.; Forsberg, F. Characterization of Ultrasound Contrast Microbubbles Using in Vitro Experiments and Viscous and Viscoelastic Interface Models for Encapsulation. J. Acoust. Soc. Am. 2005, 118, 539–550. [Google Scholar] [CrossRef]
  51. Sontum, P.C. Physicochemical Characteristics of SonazoidTM, A New Contrast Agent for Ultrasound Imaging. Ultrasound Med. Biol. 2008, 34, 824–833. [Google Scholar] [CrossRef]
  52. Landmark, K.E.; Johansen, P.W.; Johnson, J.A.; Johansen, B.; Uran, S.; Skotland, T. Pharmacokinetics of Perfluorobutane Following Intravenous Bolus Injection and Continuous Infusion of SonazoidTM in Healthy Volunteers and in Patients with Reduced Pulmonary Diffusing Capacity. Ultrasound Med. Biol. 2008, 34, 494–501. [Google Scholar] [CrossRef]
  53. Valdez, M.A.; Fernandez, E.; Matsunaga, T.; Erickson, R.P.; Trouard, T.P. Distribution and Diffusion of Macromolecule Delivery to the Brain via Focused Ultrasound Using Magnetic Resonance and Multispectral Fluorescence Imaging. Ultrasound Med. Biol. 2020, 46, 122–136. [Google Scholar] [CrossRef] [PubMed]
  54. Upadhyay, A.; Dalvi, S.V. Microbubble Formulations: Synthesis, Stability, Modeling and Biomedical Applications. Ultrasound Med. Biol. 2019, 45, 301–343. [Google Scholar] [CrossRef] [PubMed]
  55. Al-Jawadi, S.; Thakur, S.S. Ultrasound-Responsive Lipid Microbubbles for Drug Delivery: A Review of Preparation Techniques to Optimise Formulation Size, Stability and Drug Loading. Int. J. Pharm. 2020, 585, 119559. [Google Scholar] [CrossRef]
  56. Borden, M.A. Lipid-Coated Nanodrops and Microbubbles. In Handbook of Ultrasonics and Sonochemistry; Springer: Singapore, 2016; pp. 1075–1100. ISBN 978-981-287-277-7. [Google Scholar]
  57. Borden, M.A. Intermolecular Forces Model for Lipid Microbubble Shells. Langmuir 2019, 35, 10042–10051. [Google Scholar] [CrossRef]
  58. Lum, J.S.; Dove, J.D.; Murray, T.W.; Borden, M.A. Single Microbubble Measurements of Lipid Monolayer Viscoelastic Properties for Small-Amplitude Oscillations. Langmuir 2016, 32, 9410–9417. [Google Scholar] [CrossRef] [PubMed]
  59. Kim, D.H.; Costello, M.J.; Duncan, P.B.; Needham, D. Mechanical Properties and Microstructure of Polycrystalline Phospholipid Monolayer Shells: Novel Solid Microparticles. Langmuir 2003, 19, 8455–8466. [Google Scholar] [CrossRef]
  60. Borden, M.A.; Longo, M.L. Oxygen Permeability of Fully Condensed Lipid Monolayers. J. Phys. Chem. B 2004, 108, 6009–6016. [Google Scholar] [CrossRef]
  61. Pu, G.; Longo, M.L.; Borden, M.A. Effect of Microstructure on Molecular Oxygen Permeation through Condensed Phospholipid Monolayers. J. Am. Chem. Soc. 2005, 127, 6524–6525. [Google Scholar] [CrossRef]
  62. Borden, M.A.; Kruse, D.E.; Caskey, C.F.; Shukui, Z.; Dayton, P.A.; Ferrara, K.W. Influence of Lipid Shell Physicochemical Properties on Ultrasound-Induced Microbubble Destruction. IEEE Trans. Ultrason. Ferroelect. Freq. Contr. 2005, 52, 1992–2002. [Google Scholar] [CrossRef]
  63. Garg, S.; Thomas, A.A.; Borden, M.A. The Effect of Lipid Monolayer In-Plane Rigidity on in Vivo Microbubble Circulation Persistence. Biomaterials 2013, 34, 6862–6870. [Google Scholar] [CrossRef] [PubMed]
  64. Borden, M.A.; Longo, M.L. Dissolution Behavior of Lipid Monolayer-Coated, Air-Filled Microbubbles: Effect of Lipid Hydrophobic Chain Length. Langmuir 2002, 18, 9225–9233. [Google Scholar] [CrossRef]
  65. van Rooij, T.; Luan, Y.; Renaud, G.; van der Steen, A.F.W.; Versluis, M.; de Jong, N.; Kooiman, K. Non-Linear Response and Viscoelastic Properties of Lipid-Coated Microbubbles: DSPC versus DPPC. Ultrasound Med. Biol. 2015, 41, 1432–1445. [Google Scholar] [CrossRef] [PubMed]
  66. Wu, S.-Y.; Chen, C.C.; Tung, Y.-S.; Olumolade, O.O.; Konofagou, E.E. Effects of the Microbubble Shell Physicochemical Properties on Ultrasound-Mediated Drug Delivery to the Brain. J. Control. Release 2015, 212, 30–40. [Google Scholar] [CrossRef]
  67. Borden, M.A.; Streeter, J.E.; Sirsi, S.R.; Dayton, P.A. In Vivo Demonstration of Cancer Molecular Imaging with Ultrasound Radiation Force and Buried-Ligand Microbubbles. Mol. Imaging 2013, 12, 357–363. [Google Scholar] [CrossRef]
  68. Chen, C.C.; Sirsi, S.R.; Homma, S.; Borden, M.A. Effect of Surface Architecture on In Vivo Ultrasound Contrast Persistence of Targeted Size-Selected Microbubbles. Ultrasound Med. Biol. 2012, 38, 492–503. [Google Scholar] [CrossRef]
  69. Borden, M.A.; Sarantos, M.R.; Stieger, S.M.; Simon, S.I.; Ferrara, K.W.; Dayton, P.A. Ultrasound Radiation Force Modulates Ligand Availability on Targeted Contrast Agents. Mol. Imaging 2006, 5, 139–147. [Google Scholar] [CrossRef]
  70. Borden, M.A.; Zhang, H.; Gillies, R.J.; Dayton, P.A.; Ferrara, K.W. A Stimulus-Responsive Contrast Agent for Ultrasound Molecular Imaging. Biomaterials 2008, 29, 597–606. [Google Scholar] [CrossRef]
  71. Chen, C.C.; Borden, M.A. The Role of Poly(Ethylene Glycol) Brush Architecture in Complement Activation on Targeted Microbubble Surfaces. Biomaterials 2011, 32, 6579–6587. [Google Scholar] [CrossRef]
  72. Kiessling, F.; Fokong, S.; Bzyl, J.; Lederle, W.; Palmowski, M.; Lammers, T. Recent Advances in Molecular, Multimodal and Theranostic Ultrasound Imaging. Adv. Drug Deliv. Rev. 2014, 72, 15–27. [Google Scholar] [CrossRef]
  73. Rudakovskaya, P.G.; Barmin, R.A.; Kuzmin, P.S.; Fedotkina, E.P.; Sencha, A.N.; Gorin, D.A. Microbubbles Stabilized by Protein Shell: From Pioneering Ultrasound Contrast Agents to Advanced Theranostic Systems. Pharmaceutics 2022, 14, 1236. [Google Scholar] [CrossRef] [PubMed]
  74. Liu, M.; Dasgupta, A.; Qu, N.; Rama, E.; Kiessling, F.; Lammers, T. Strategies to Maximize Anthracycline Drug Loading in Albumin Microbubbles. ACS Biomater. Sci. Eng. 2021. Online ahead of print. [Google Scholar] [CrossRef] [PubMed]
  75. Ji, J.; Ji, J.; He, X. Preparation of Ultrasound Microbubbles Crosslinked to Albumin Nanoparticles Packaged with Tissue-Type Plasminogen Activator Gene Plasmid and Method of in Vivo Transfection. JEP 2011, 24, 35–41. [Google Scholar] [CrossRef] [PubMed]
  76. Borrelli, M.J.; O’Brien, W.D.; Bernock, L.J.; Williams, H.R.; Hamilton, E.; Wu, J.; Oelze, M.L.; Culp, W.C. Production of Uniformly Sized Serum Albumin and Dextrose Microbubbles. Ultrason. Sonochem. 2012, 19, 198–208. [Google Scholar] [CrossRef] [PubMed]
  77. Chen, J.L.; Dhanaliwala, A.H.; Dixon, A.J.; Klibanov, A.L.; Hossack, J.A. Synthesis and Characterization of Transiently Stable Albumin-Coated Microbubbles via a Flow-Focusing Microfluidic Device. Ultrasound Med. Biol. 2014, 40, 400–409. [Google Scholar] [CrossRef]
  78. Upadhyay, A.; Dalvi, S.V. Synthesis, Characterization and Stability of BSA-Encapsulated Microbubbles. RSC Adv. 2016, 6, 15016–15026. [Google Scholar] [CrossRef]
  79. Lentacker, I.; De Geest, B.G.; Vandenbroucke, R.E.; Peeters, L.; Demeester, J.; De Smedt, S.C.; Sanders, N.N. Ultrasound-Responsive Polymer-Coated Microbubbles That Bind and Protect DNA. Langmuir 2006, 22, 7273–7278. [Google Scholar] [CrossRef]
  80. Upadhyay, A.; Dalvi, S.V.; Gupta, G.; Khanna, N. Effect of PEGylation on Performance of Protein Microbubbles and Its Comparison with Lipid Microbubbles. Mater. Sci. Eng. C 2017, 71, 425–430. [Google Scholar] [CrossRef]
  81. Ma, X.; Bussonniere, A.; Liu, Q. A Facile Sonochemical Synthesis of Shell-Stabilized Reactive Microbubbles Using Surface-Thiolated Bovine Serum Albumin with the Traut’s Reagent. Ultrason. Sonochem. 2017, 36, 454–465. [Google Scholar] [CrossRef]
  82. Liu, X.; Gong, P.; Song, P.; Xie, F.; Miller II, A.L.; Chen, S.; Lu, L. Fast Functionalization of Ultrasound Microbubbles Using Strain Promoted Click Chemistry. Biomater. Sci. 2018, 6, 623–632. [Google Scholar] [CrossRef]
  83. Narihira, K.; Watanabe, A.; Sheng, H.; Endo, H.; Feril, L.B.; Irie, Y.; Ogawa, K.; Moosavi-Nejad, S.; Kondo, S.; Kikuta, T.; et al. Enhanced Cell Killing and Apoptosis of Oral Squamous Cell Carcinoma Cells with Ultrasound in Combination with Cetuximab Coated Albumin Microbubbles. J. Drug Target. 2018, 26, 278–288. [Google Scholar] [CrossRef] [PubMed]
  84. Wang, Y.-H.; Liao, A.-H.; Chen, J.-H.; Chris Wang, C.-R.; Li, P.-C. Photoacoustic/Ultrasound Dual-Modality Contrast Agent and Its Application to Thermotherapy. J. Biomed. Opt. 2012, 17, 045001. [Google Scholar] [CrossRef] [PubMed]
  85. Liou, Y.-R.; Wang, Y.-H.; Lee, C.-Y.; Li, P.-C. Buoyancy-Activated Cell Sorting Using Targeted Biotinylated Albumin Microbubbles. PLoS ONE 2015, 10, e0125036. [Google Scholar] [CrossRef]
  86. Wang, Y.-H.; Chen, S.-P.; Liao, A.-H.; Yang, Y.-C.; Lee, C.-R.; Wu, C.-H.; Wu, P.-C.; Liu, T.-M.; Wang, C.-R.C.; Li, P.-C. Synergistic Delivery of Gold Nanorods Using Multifunctional Microbubbles for Enhanced Plasmonic Photothermal Therapy. Sci. Rep. 2014, 4, 5685. [Google Scholar] [CrossRef]
  87. Porter, T.R.; Xie, F.; Knapp, D.; Iversen, P.; Marky, L.A.; Tsutsui, J.M.; Maiti, S.; Lof, J.; Radio, S.J.; Kipshidze, N. Targeted Vascular Delivery of Antisense Molecules Using Intravenous Microbubbles. Cardiovasc. Revascularization Med. 2006, 7, 25–33. [Google Scholar] [CrossRef]
  88. Liao, A.-H.; Wu, S.-Y.; Wang, H.-E.; Weng, C.-H.; Wu, M.-F.; Li, P.-C. Evaluation of 18F-Labeled Targeted Perfluorocarbon-Filled Albumin Microbubbles as a Probe for MicroUS and MicroPET in Tumor-Bearing Mice. Ultrasonics 2013, 53, 320–327. [Google Scholar] [CrossRef]
  89. Barmin, R.; Rudakovskaya, P.; Gusliakova, O.; Sindeeva, O.; Prikhozhdenko, E.; Maksimova, E.; Obukhova, E.; Chernyshev, V.; Khlebtsov, B.; Solovev, A.; et al. Air-Filled Bubbles Stabilized by Gold Nanoparticle/Photodynamic Dye Hybrid Structures for Theranostics. Nanomaterials 2021, 11, 415. [Google Scholar] [CrossRef] [PubMed]
  90. Maksimova, E.A.; Barmin, R.A.; Rudakovskaya, P.G.; Sindeeva, O.A.; Prikhozhdenko, E.S.; Yashchenok, A.M.; Khlebtsov, B.N.; Solovev, A.A.; Huang, G.; Mei, Y.; et al. Air-Filled Microbubbles Based on Albumin Functionalized with Gold Nanocages and Zinc Phthalocyanine for Multimodal Imaging. Micromachines 2021, 12, 1161. [Google Scholar] [CrossRef]
  91. Yoon, Y.I.; Pang, X.; Jung, S.; Zhang, G.; Kong, M.; Liu, G.; Chen, X. Smart Gold Nanoparticle-Stabilized Ultrasound Microbubbles as Cancer Theranostics. J. Mater. Chem. B 2018, 6, 3235–3239. [Google Scholar] [CrossRef]
  92. Chen, Z.; Chattaraj, R.; Pulsipher, K.W.; Karmacharya, M.B.; Hammer, D.A.; Lee, D.; Sehgal, C.M. Photoacoustic and Ultrasound Dual-Mode Imaging via Functionalization of Recombinant Protein-Stabilized Microbubbles with Methylene Blue. ACS Appl. Bio Mater. 2019, 2, 4020–4026. [Google Scholar] [CrossRef]
  93. Xiong, X.; Zhao, F.; Shi, M.; Yang, H.; Liu, Y. Polymeric Microbubbles for Ultrasonic Molecular Imaging and Targeted Therapeutics. J. Biomater. Sci. Polym. Ed. 2011, 22, 417–428. [Google Scholar] [CrossRef] [PubMed]
  94. Liu, M.; Dasgupta, A.; Koczera, P.; Schipper, S.; Rommel, D.; Shi, Y.; Kiessling, F.; Lammers, T. Drug Loading in Poly(Butyl Cyanoacrylate)-Based Polymeric Microbubbles. Mol. Pharm. 2020, 17, 2840–2848. [Google Scholar] [CrossRef]
  95. Estifeeva, T.M.; Barmin, R.A.; Rudakovskaya, P.G.; Nechaeva, A.M.; Luss, A.L.; Mezhuev, Y.O.; Chernyshev, V.S.; Krivoborodov, E.G.; Klimenko, O.A.; Sindeeva, O.A.; et al. Hybrid (Bovine Serum Albumin)/Poly (N-Vinyl-2-Pyrrolidone-Co-Acrylic Acid)-Shelled Microbubbles as Advanced Ultrasound Contrast Agents. ACS Appl. Bio. Mater. 2022, 5, 3338–3348. [Google Scholar] [CrossRef] [PubMed]
  96. Barmin, R.A.; Dasgupta, A.; Bastard, C.; De Laporte, L.; Rütten, S.; Weiler, M.; Kiessling, F.; Lammers, T.; Pallares, R.M. Engineering the Acoustic Response and Drug Loading Capacity of PBCA-Based Polymeric Microbubbles with Surfactants. Mol. Pharm. 2022, 19, 3256–3266. [Google Scholar] [CrossRef] [PubMed]
  97. Omata, D.; Maruyama, T.; Unga, J.; Hagiwara, F.; Munakata, L.; Kageyama, S.; Shima, T.; Suzuki, Y.; Maruyama, K.; Suzuki, R. Effects of Encapsulated Gas on Stability of Lipid-Based Microbubbles and Ultrasound-Triggered Drug Delivery. J. Control. Release 2019, 311–312, 65–73. [Google Scholar] [CrossRef] [PubMed]
  98. Sun, J.; Yin, M.; Zhu, S.; Liu, L.; Zhu, Y.; Wang, Z.; Xu, R.X.; Chang, S. Ultrasound-Mediated Destruction of Oxygen and Paclitaxel Loaded Lipid Microbubbles for Combination Therapy in Hypoxic Ovarian Cancer Cells. Ultrason. Sonochem. 2016, 28, 319–326. [Google Scholar] [CrossRef] [PubMed]
  99. Eisenbrey, J.R.; Shraim, R.; Liu, J.-B.; Li, J.; Stanczak, M.; Oeffinger, B.; Leeper, D.B.; Keith, S.W.; Jablonowski, L.J.; Forsberg, F.; et al. Sensitization of Hypoxic Tumors to Radiation Therapy Using Ultrasound-Sensitive Oxygen Microbubbles. Int. J. Radiat. Oncol. Biol. Phys. 2018, 101, 88–96. [Google Scholar] [CrossRef]
  100. Ho, Y.-J.; Chu, S.-W.; Liao, E.-C.; Fan, C.-H.; Chan, H.-L.; Wei, K.-C.; Yeh, C.-K. Normalization of Tumor Vasculature by Oxygen Microbubbles with Ultrasound. Theranostics 2019, 9, 7370–7383. [Google Scholar] [CrossRef]
  101. Liang, Z.; Chen, H.; Gong, X.; Shi, B.; Lin, L.; Tao, F.; Wu, Q.; Fang, M.; Li, H.; Lu, C.; et al. Ultrasound-Induced Destruction of Nitric Oxide–Loaded Microbubbles in the Treatment of Thrombus and Ischemia–Reperfusion Injury. Front. Pharmacol. 2022, 12, 745693. [Google Scholar] [CrossRef]
  102. Wang, C.; Yang, F.; Xu, Z.; Shi, D.; Chen, D.; Dai, J.; Gu, N.; Jiang, Q. Intravenous Release of NO from Lipidic Microbubbles Accelerates Deep Vein Thrombosis Resolution in a Rat Model. Thromb. Res. 2013, 131, e31–e38. [Google Scholar] [CrossRef]
  103. Kwan, J.J.; Kaya, M.; Borden, M.A.; Dayton, P.A. Theranostic Oxygen Delivery Using Ultrasound and Microbubbles. Theranostics 2012, 2, 1174–1184. [Google Scholar] [CrossRef] [PubMed]
  104. Reusser, T.D.; Song, K.-H.; Ramirez, D.; Benninger, R.K.; Papadopoulou, V.; Borden, M.A. Phospholipid Oxygen Microbubbles for Image-Guided Therapy. Nanotheranostics 2020, 4, 83–90. [Google Scholar] [CrossRef] [PubMed]
  105. Luo, T.; Sun, J.; Zhu, S.; He, J.; Hao, L.; Xiao, L.; Zhu, Y.; Wang, Q.; Pan, X.; Wang, Z.; et al. Ultrasound-Mediated Destruction of Oxygen and Paclitaxel Loaded Dual-Targeting Microbubbles for Intraperitoneal Treatment of Ovarian Cancer Xenografts. Cancer Lett. 2017, 391, 1–11. [Google Scholar] [CrossRef] [PubMed]
  106. Sheng, Y.; Beguin, E.; Nesbitt, H.; Kamila, S.; Owen, J.; Barnsley, L.C.; Callan, B.; O’Kane, C.; Nomikou, N.; Hamoudi, R.; et al. Magnetically Responsive Microbubbles as Delivery Vehicles for Targeted Sonodynamic and Antimetabolite Therapy of Pancreatic Cancer. J. Control. Release 2017, 262, 192–200. [Google Scholar] [CrossRef] [PubMed]
  107. Fan, C.-H.; Cheng, Y.-H.; Ting, C.-Y.; Ho, Y.-J.; Hsu, P.-H.; Liu, H.-L.; Yeh, C.-K. Ultrasound/Magnetic Targeting with SPIO-DOX-Microbubble Complex for Image-Guided Drug Delivery in Brain Tumors. Theranostics 2016, 6, 1542–1556. [Google Scholar] [CrossRef]
  108. Liao, T.; Li, Q.; Zhang, Y.; Yang, Z.; Huang, Z.; Han, F.; Chen, X.; Yin, T.; Ren, J.; Sun, Q. Precise Treatment of Acute Antibody-Mediated Cardiac Allograft Rejection in Rats Using C4d-Targeted Microbubbles Loaded with Nitric Oxide. J. Heart Lung Transplant. 2020, 39, 481–490. [Google Scholar] [CrossRef] [PubMed]
  109. Sonne, C. Differences in Definity and Optison Microbubble Destruction Rates at a Similar Mechanical Index with Different Real-Time Perfusion Systems. J. Am. Soc. Echocardiogr. 2003, 16, 1178–1185. [Google Scholar] [CrossRef]
  110. Hyvelin, J.-M.; Gaud, E.; Costa, M.; Helbert, A.; Bussat, P.; Bettinger, T.; Frinking, P. Characteristics and Echogenicity of Clinical Ultrasound Contrast Agents: An In Vitro and In Vivo Comparison Study: Comparison of Clinical Ultrasound Contrast Agents. J. Ultrasound Med. 2017, 36, 941–953. [Google Scholar] [CrossRef]
  111. McMahon, D.; Lassus, A.; Gaud, E.; Jeannot, V.; Hynynen, K. Microbubble Formulation Influences Inflammatory Response to Focused Ultrasound Exposure in the Brain. Sci. Rep. 2020, 10, 21534. [Google Scholar] [CrossRef] [PubMed]
  112. Sirsi, S.; Feshitan, J.; Kwan, J.; Homma, S.; Borden, M. Effect of Microbubble Size on Fundamental Mode High Frequency Ultrasound Imaging in Mice. Ultrasound Med. Biol. 2010, 36, 935–948. [Google Scholar] [CrossRef]
  113. Choi, J.J.; Feshitan, J.A.; Baseri, B.; Shougang, W.; Yao-Sheng, T.; Borden, M.A.; Konofagou, E.E. Microbubble-Size Dependence of Focused Ultrasound-Induced Blood–Brain Barrier Opening in Mice In Vivo. IEEE Trans. Biomed. Eng. 2010, 57, 145–154. [Google Scholar] [CrossRef]
  114. Samiotaki, G.; Vlachos, F.; Tung, Y.-S.; Konofagou, E.E. A Quantitative Pressure and Microbubble-Size Dependence Study of Focused Ultrasound-Induced Blood-Brain Barrier Opening Reversibility in Vivo Using MRI: FUS-Induced BBB Opening Reversibility. Magn. Reson. Med. 2012, 67, 769–777. [Google Scholar] [CrossRef] [PubMed]
  115. McMahon, D.; Hynynen, K. Acute Inflammatory Response Following Increased Blood-Brain Barrier Permeability Induced by Focused Ultrasound Is Dependent on Microbubble Dose. Theranostics 2017, 7, 3989–4000. [Google Scholar] [CrossRef]
  116. Song, K.-H.; Fan, A.C.; Hinkle, J.J.; Newman, J.; Borden, M.A.; Harvey, B.K. Microbubble Gas Volume: A Unifying Dose Parameter in Blood-Brain Barrier Opening by Focused Ultrasound. Theranostics 2017, 7, 144–152. [Google Scholar] [CrossRef] [PubMed]
  117. Navarro-Becerra, J.A.; Song, K.-H.; Martinez, P.; Borden, M.A. Microbubble Size and Dose Effects on Pharmacokinetics. ACS Biomater. Sci. Eng. 2022, 8, 1686–1695. [Google Scholar] [CrossRef] [PubMed]
  118. Bing, C.; Hong, Y.; Hernandez, C.; Rich, M.; Cheng, B.; Munaweera, I.; Szczepanski, D.; Xi, Y.; Bolding, M.; Exner, A.; et al. Characterization of Different Bubble Formulations for Blood-Brain Barrier Opening Using a Focused Ultrasound System with Acoustic Feedback Control. Sci. Rep. 2018, 8, 7986. [Google Scholar] [CrossRef] [PubMed]
  119. Gao, Y.; Wu, S.; Li, L.; Wang, G.; Shen, W.; Xu, Y.; Liu, Z.; Zhuo, Z.; Xia, H.; Tan, K. Ultrasound-Targeted Stromal Cell-Derived Factor-1-Loaded Microbubble Destruction Promotes Mesenchymal Stem Cell Homing to Kidneys in Diabetic Nephropathy Rats. IJN 2014, 9, 5639–5651. [Google Scholar] [CrossRef]
  120. Metzger, K.; Vogel, S.; Chatterjee, M.; Borst, O.; Seizer, P.; Schönberger, T.; Geisler, T.; Lang, F.; Langer, H.; Rheinlaender, J.; et al. High-Frequency Ultrasound-Guided Disruption of Glycoprotein VI-Targeted Microbubbles Targets Atheroprogressison in Mice. Biomaterials 2015, 36, 80–89. [Google Scholar] [CrossRef]
  121. Yang, H.; Sun, Y.; Wei, J.; Xu, L.; Tang, Y.; Yang, L.; Zhang, X.; Lu, Y. The Effects of Ultrasound-Targeted Microbubble Destruction (UTMD) Carrying IL-8 Monoclonal Antibody on the Inflammatory Responses and Stability of Atherosclerotic Plaques. Biomed. Pharmacother. 2019, 118, 109161. [Google Scholar] [CrossRef]
  122. Chang, E.-L.; Ting, C.-Y.; Hsu, P.-H.; Lin, Y.-C.; Liao, E.-C.; Huang, C.-Y.; Chang, Y.-C.; Chan, H.-L.; Chiang, C.-S.; Liu, H.-L.; et al. Angiogenesis-Targeting Microbubbles Combined with Ultrasound-Mediated Gene Therapy in Brain Tumors. J. Control. Release 2017, 255, 164–175. [Google Scholar] [CrossRef]
  123. Fan, C.-H.; Ting, C.-Y.; Liu, H.-L.; Huang, C.-Y.; Hsieh, H.-Y.; Yen, T.-C.; Wei, K.-C.; Yeh, C.-K. Antiangiogenic-Targeting Drug-Loaded Microbubbles Combined with Focused Ultrasound for Glioma Treatment. Biomaterials 2013, 34, 2142–2155. [Google Scholar] [CrossRef] [PubMed]
  124. Zhou, Y.; Gu, H.; Xu, Y.; Li, F.; Kuang, S.; Wang, Z.; Zhou, X.; Ma, H.; Li, P.; Zheng, Y.; et al. Targeted Antiangiogenesis Gene Therapy Using Targeted Cationic Microbubbles Conjugated with CD105 Antibody Compared with Untargeted Cationic and Neutral Microbubbles. Theranostics 2015, 5, 399–417. [Google Scholar] [CrossRef] [PubMed]
  125. Nederhoed, J.H.; Ebben, H.P.; Slikkerveer, J.; Hoksbergen, A.W.J.; Kamp, O.; Tangelder, G.-J.; Wisselink, W.; Musters, R.J.P.; Yeung, K.K. Intravenous Targeted Microbubbles Carrying Urokinase versus Urokinase Alone in Acute Peripheral Arterial Thrombosis in a Porcine Model. Ann. Vasc. Surg. 2017, 44, 400–407. [Google Scholar] [CrossRef]
  126. Wang, X.; Searle, A.K.; Hohmann, J.D.; Liu, A.L.; Abraham, M.-K.; Palasubramaniam, J.; Lim, B.; Yao, Y.; Wallert, M.; Yu, E.; et al. Dual-Targeted Theranostic Delivery of MiRs Arrests Abdominal Aortic Aneurysm Development. Mol. Ther. 2018, 26, 1056–1065. [Google Scholar] [CrossRef] [PubMed]
  127. Pu, C.; Chang, S.; Sun, J.; Zhu, S.; Liu, H.; Zhu, Y.; Wang, Z.; Xu, R.X. Ultrasound-Mediated Destruction of LHRHa-Targeted and Paclitaxel-Loaded Lipid Microbubbles for the Treatment of Intraperitoneal Ovarian Cancer Xenografts. Mol. Pharm. 2014, 11, 49–58. [Google Scholar] [CrossRef] [PubMed]
  128. Hua, X.; Zhou, L.; Liu, P.; He, Y.; Tan, K.; Chen, Q.; Gao, Y.; Gao, Y. In Vivo Thrombolysis with Targeted Microbubbles Loading Tissue Plasminogen Activator in a Rabbit Femoral Artery Thrombus Model. J. Thromb. Thrombolysis. 2014, 38, 57–64. [Google Scholar] [CrossRef]
  129. Fan, C.-H.; Chang, E.-L.; Ting, C.-Y.; Lin, Y.-C.; Liao, E.-C.; Huang, C.-Y.; Chang, Y.-C.; Chan, H.-L.; Wei, K.-C.; Yeh, C.-K. Folate-Conjugated Gene-Carrying Microbubbles with Focused Ultrasound for Concurrent Blood-Brain Barrier Opening and Local Gene Delivery. Biomaterials 2016, 106, 46–57. [Google Scholar] [CrossRef]
  130. Woudstra, L.; Krijnen, P.A.J.; Bogaards, S.J.P.; Meinster, E.; Emmens, R.W.; Kokhuis, T.J.A.; Bollen, I.A.E.; Baltzer, H.; Baart, S.M.T.; Parbhudayal, R.; et al. Development of a New Therapeutic Technique to Direct Stem Cells to the Infarcted Heart Using Targeted Microbubbles: StemBells. Stem Cell Res. 2016, 17, 6–15. [Google Scholar] [CrossRef]
  131. Woudstra, L.; Meinster, E.; Van Haren, L.; Kay, A.M.; Koopman, M.; Belien, J.A.M.; Morrison, M.C.; Van Rossum, A.C.; Helder, M.N.; Juffermans, L.J.M.; et al. StemBell Therapy Stabilizes Atherosclerotic Plaques after Myocardial Infarction. Cytotherapy 2018, 20, 1143–1154. [Google Scholar] [CrossRef]
  132. Liu, Y.; Zhou, Y.; Xu, J.; Luo, H.; Zhu, Y.; Zeng, X.; Dong, F.; Wei, Z.; Yan, F.; Zheng, H. Ultrasound Molecular Imaging-Guided Tumor Gene Therapy through Dual-Targeted Cationic Microbubbles. Biomater. Sci. 2021, 9, 2454–2466. [Google Scholar] [CrossRef]
  133. Luo, W.; Wen, G.; Yang, L.; Tang, J.; Wang, J.; Wang, J.; Zhang, S.; Zhang, L.; Ma, F.; Xiao, L.; et al. Dual-Targeted and PH-Sensitive Doxorubicin Prodrug-Microbubble Complex with Ultrasound for Tumor Treatment. Theranostics 2017, 7, 452–465. [Google Scholar] [CrossRef]
  134. Zhang, L.; Sun, L.; Tang, Q.; Sun, S.; Zeng, L.; Ma, J.; Li, X.; Ge, H.; Liang, X. Cascade Drug Delivery through Tumor Barriers of Pancreatic Cancer via Ultrasound in Combination with Functional Microbubbles. ACS Biomater. Sci. Eng. 2022, 8, 1583–1595. [Google Scholar] [CrossRef]
  135. Beguin, E.; Gray, M.D.; Logan, K.A.; Nesbitt, H.; Sheng, Y.; Kamila, S.; Barnsley, L.C.; Bau, L.; McHale, A.P.; Callan, J.F.; et al. Magnetic Microbubble Mediated Chemo-Sonodynamic Therapy Using a Combined Magnetic-Acoustic Device. J. Control. Release 2020, 317, 23–33. [Google Scholar] [CrossRef] [PubMed]
  136. Fan, C.-H.; Ting, C.-Y.; Lin, H.-J.; Wang, C.-H.; Liu, H.-L.; Yen, T.-C.; Yeh, C.-K. SPIO-Conjugated, Doxorubicin-Loaded Microbubbles for Concurrent MRI and Focused-Ultrasound Enhanced Brain-Tumor Drug Delivery. Biomaterials 2013, 34, 3706–3715. [Google Scholar] [CrossRef]
  137. Wang, S.; Guo, X.; Xiu, W.; Liu, Y.; Ren, L.; Xiao, H.; Yang, F.; Gao, Y.; Xu, C.; Wang, L. Accelerating Thrombolysis Using a Precision and Clot-Penetrating Drug Delivery Strategy by Nanoparticle-Shelled Microbubbles. Sci. Adv. 2020, 6, eaaz8204. [Google Scholar] [CrossRef]
  138. Duan, L.; Yang, F.; He, W.; Song, L.; Qiu, F.; Xu, N.; Xu, L.; Zhang, Y.; Hua, Z.; Gu, N. A Multi-Gradient Targeting Drug Delivery System Based on RGD-l-TRAIL-Labeled Magnetic Microbubbles for Cancer Theranostics. Adv. Funct. Mater. 2016, 26, 8313–8324. [Google Scholar] [CrossRef]
  139. Dwivedi, P.; Kiran, S.; Han, S.; Dwivedi, M.; Khatik, R.; Fan, R.; Mangrio, F.A.; Du, K.; Zhu, Z.; Yang, C.; et al. Magnetic Targeting and Ultrasound Activation of Liposome–Microbubble Conjugate for Enhanced Delivery of Anticancer Therapies. ACS Appl. Mater. Interfaces 2020, 12, 23737–23751. [Google Scholar] [CrossRef] [PubMed]
  140. Zhang, J.; Wang, S.; Deng, Z.; Li, L.; Tan, G.; Liu, X.; Zheng, H.; Yan, F. Ultrasound-Triggered Drug Delivery for Breast Tumor Therapy Through IRGD-Targeted Paclitaxel-Loaded Liposome-Microbubble Complexes. J. Biomed. Nanotechnol. 2018, 14, 1384–1395. [Google Scholar] [CrossRef] [PubMed]
  141. Zhao, G.; Huang, Q.; Wang, F.; Zhang, X.; Hu, J.; Tan, Y.; Huang, N.; Wang, Z.; Wang, Z.; Cheng, Y. Targeted ShRNA-Loaded Liposome Complex Combined with Focused Ultrasound for Blood Brain Barrier Disruption and Suppressing Glioma Growth. Cancer Lett. 2018, 418, 147–158. [Google Scholar] [CrossRef]
  142. Ingram, N.; McVeigh, L.E.; Abou-Saleh, R.H.; Maynard, J.; Peyman, S.A.; McLaughlan, J.R.; Fairclough, M.; Marston, G.; Valleley, E.M.A.; Jimenez-Macias, J.L.; et al. Ultrasound-Triggered Therapeutic Microbubbles Enhance the Efficacy of Cytotoxic Drugs by Increasing Circulation and Tumor Drug Accumulation and Limiting Bioavailability and Toxicity in Normal Tissues. Theranostics 2020, 10, 10973–10992. [Google Scholar] [CrossRef]
  143. Charalambous, A.; Mico, V.; McVeigh, L.E.; Marston, G.; Ingram, N.; Volpato, M.; Peyman, S.A.; McLaughlan, J.R.; Wierzbicki, A.; Loadman, P.M.; et al. Targeted Microbubbles Carrying Lipid-Oil-Nanodroplets for Ultrasound-Triggered Delivery of the Hydrophobic Drug, Combretastatin A4. Nanomed. Nanotechnol. Biol. Med. 2021, 36, 102401. [Google Scholar] [CrossRef]
  144. Liao, A.-H.; Chou, H.-Y.; Hsieh, Y.-L.; Hsu, S.-C.; Wei, K.-C.; Liu, H.-L. Enhanced Therapeutic Epidermal Growth Factor Receptor (EGFR) Antibody Delivery via Pulsed Ultrasound with Targeting Microbubbles for Glioma Treatment. J. Med. Biol. Eng. 2015, 35, 156–164. [Google Scholar] [CrossRef] [PubMed]
  145. Kang, M.; Zhang, Y.; Jin, X.; Chen, G.; Huang, Y.; Wu, D.; Li, G.; Shan, J.; Huang, P.; Chen, J. Concurrent Treatment with Anti-DLL4 Enhances Antitumor and Proapoptotic Efficacy of a γ-Secretase Inhibitor in Gastric Cancer. Transl. Oncol. 2018, 11, 599–608. [Google Scholar] [CrossRef] [PubMed]
  146. Sun, L.; Zhang, J.; Xu, M.; Zhang, L.; Tang, Q.; Chen, J.; Gong, M.; Sun, S.; Ge, H.; Wang, S.; et al. Ultrasound Microbubbles Mediated Sonosensitizer and Antibody Co-Delivery for Highly Efficient Synergistic Therapy on HER2-Positive Gastric Cancer. ACS Appl. Mater. Interfaces 2022, 14, 452–463. [Google Scholar] [CrossRef]
  147. Kim, D.; Lee, S.S.; Moon, H.; Park, S.Y.; Lee, H.J. PD-L1 Targeting Immune-Microbubble Complex Enhances Therapeutic Index in Murine Colon Cancer Models. Pharmaceuticals 2020, 14, 6. [Google Scholar] [CrossRef]
  148. Ma, Y.; Han, J.; Jiang, J.; Zheng, Z.; Tan, Y.; Liu, C.; Zhao, Y. Ultrasound Targeting of Microbubble-Bound Anti PD-L1 MAb to Enhance Anti-Tumor Effect of Cisplatin in Cervical Cancer Xenografts Treatment. Life Sci. 2020, 262, 118565. [Google Scholar] [CrossRef]
  149. Liu, Y.; Jiang, J.; Liu, C.; Zhao, W.; Ma, Y.; Zheng, Z.; Zhou, Q.; Zhao, Y. Synergistic Anti-Tumor Effect of Anti-PD-L1 Antibody Cationic Microbubbles for Delivery of the MiR-34a Gene Combined with Ultrasound on Cervical Carcinoma. Am. J. Transl. Res. 2021, 13, 988–1005. [Google Scholar] [PubMed]
  150. Li, X.; Khorsandi, S.; Wang, Y.; Santelli, J.; Huntoon, K.; Nguyen, N.; Yang, M.; Lee, D.; Lu, Y.; Gao, R.; et al. Cancer Immunotherapy Based on Image-Guided STING Activation by Nucleotide Nanocomplex-Decorated Ultrasound Microbubbles. Nat. Nanotechnol. 2022, 17, 891–899. [Google Scholar] [CrossRef] [PubMed]
  151. Gao, X.; Nan, Y.; Yuan, Y.; Gong, X.; Sun, Y.; Zhou, H.; Zong, Y.; Zhang, L.; Yu, M. Gas-filled Ultrasound Microbubbles Enhance the Immunoactivity of the HSP70-MAGEA1 Fusion Protein against MAGEA1-expressing Tumours. Mol. Med. Rep. 2018, 18, 315–321. [Google Scholar] [CrossRef]
  152. Jugniot, N.; Dahl, J.J.; Paulmurugan, R. Immunotheranostic Microbubbles (IMBs)—A Modular Platform for Dendritic Cell Vaccine Delivery Applied to Breast Cancer Immunotherapy. J. Exp. Clin. Cancer Res. 2022, 41, 299. [Google Scholar] [CrossRef]
  153. Chomas, J.E.; Dayton, P.; May, D.; Ferrara, K. Threshold of Fragmentation for Ultrasonic Contrast Agents. J. Biomed. Opt. 2001, 6, 141. [Google Scholar] [CrossRef] [PubMed]
  154. Morgan, K.E.; Allen, J.S.; Dayton, P.A.; Chomas, J.E.; Klibaov, A.L.; Ferrara, K.W. Experimental and Theoretical Evaluation of Microbubble Behavior: Effect of Transmitted Phase and Bubble Size. IEEE Trans. Ultrason. Ferroelect. Freq. Contr. 2000, 47, 1494–1509. [Google Scholar] [CrossRef] [PubMed]
  155. Dayton, P.A.; Allen, J.S.; Ferrara, K.W. The Magnitude of Radiation Force on Ultrasound Contrast Agents. J. Acoust. Soc. Am. 2002, 112, 2183–2192. [Google Scholar] [CrossRef]
  156. Supponen, O.; Upadhyay, A.; Lum, J.; Guidi, F.; Murray, T.; Vos, H.J.; Tortoli, P.; Borden, M. The Effect of Size Range on Ultrasound-Induced Translations in Microbubble Populations. J. Acoust. Soc. Am. 2020, 147, 3236–3247. [Google Scholar] [CrossRef] [PubMed]
  157. Segers, T.; Kruizinga, P.; Kok, M.P.; Lajoinie, G.; de Jong, N.; Versluis, M. Monodisperse Versus Polydisperse Ultrasound Contrast Agents: Non-Linear Response, Sensitivity, and Deep Tissue Imaging Potential. Ultrasound Med. Biol. 2018, 44, 1482–1492. [Google Scholar] [CrossRef]
  158. Wang, S.; Unnikrishnan, S.; Herbst, E.B.; Klibanov, A.L.; Mauldin, F.W.; Hossack, J.A. Ultrasound Molecular Imaging of Inflammation in Mouse Abdominal Aorta. Invest. Radiol. 2017, 52, 499–506. [Google Scholar] [CrossRef]
  159. Rychak, J.J.; Klibanov, A.L.; Ley, K.F.; Hossack, J.A. Enhanced Targeting of Ultrasound Contrast Agents Using Acoustic Radiation Force. Ultrasound Med. Biol. 2007, 33, 1132–1139. [Google Scholar] [CrossRef]
  160. Liu, J.; Zhang, P.; Liu, P.; Zhao, Y.; Gao, S.; Tan, K.; Liu, Z. Endothelial Adhesion of Targeted Microbubbles in Both Small and Great Vessels Using Ultrasound Radiation Force. Mol. Imaging 2012, 11, 58–66. [Google Scholar] [CrossRef]
  161. Wang, S.; Wang, C.Y.; Unnikrishnan, S.; Klibanov, A.L.; Hossack, J.A.; Mauldin, F.W. Optical Verification of Microbubble Response to Acoustic Radiation Force in Large Vessels with In Vivo Results. Investig. Radiol. 2015, 50, 772–784. [Google Scholar] [CrossRef]
  162. Frinking, P.J.A.; Tardy, I.; Théraulaz, M.; Arditi, M.; Powers, J.; Pochon, S.; Tranquart, F. Effects of Acoustic Radiation Force on the Binding Efficiency of BR55, a VEGFR2-Specific Ultrasound Contrast Agent. Ultrasound Med. Biol. 2012, 38, 1460–1469. [Google Scholar] [CrossRef]
  163. Gessner, R.C.; Streeter, J.E.; Kothadia, R.; Feingold, S.; Dayton, P.A. An In Vivo Validation of the Application of Acoustic Radiation Force to Enhance the Diagnostic Utility of Molecular Imaging Using 3-D Ultrasound. Ultrasound Med. Biol. 2012, 38, 651–660. [Google Scholar] [CrossRef] [PubMed]
  164. Navarro-Becerra, J.A.; Castillo, J.I.; Di Ruzza, F.; Borden, M.A. Monodispersity Increases Adhesion Efficiency and Specificity for Ultrasound-Targeted Microbubbles. ACS Biomater. Sci. Eng. 2023, 9, 991–1001. [Google Scholar] [CrossRef] [PubMed]
  165. Feshitan, J.A.; Chen, C.C.; Kwan, J.J.; Borden, M.A. Microbubble Size Isolation by Differential Centrifugation. J. Colloid Interface Sci. 2009, 329, 316–324. [Google Scholar] [CrossRef] [PubMed]
  166. Segers, T.; Lassus, A.; Bussat, P.; Gaud, E.; Frinking, P. Improved Coalescence Stability of Monodisperse Phospholipid-Coated Microbubbles Formed by Flow-Focusing at Elevated Temperatures. Lab Chip 2019, 19, 158–167. [Google Scholar] [CrossRef]
  167. Segers, T.; De Rond, L.; De Jong, N.; Borden, M.; Versluis, M. Stability of Monodisperse Phospholipid-Coated Microbubbles Formed by Flow-Focusing at High Production Rates. Langmuir 2016, 32, 3937–3944. [Google Scholar] [CrossRef]
  168. Dhanaliwala, A.H.; Chen, J.L.; Wang, S.; Hossack, J.A. Liquid Flooded Flow-Focusing Microfluidic Device for in Situ Generation of Monodisperse Microbubbles. Microfluid. Nanofluid. 2013, 14, 457–467. [Google Scholar] [CrossRef]
  169. Khan, A.H.; Jiang, X.; Kaushik, A.; Nair, H.S.; Edirisinghe, M.; Mercado-Shekhar, K.P.; Shekhar, H.; Dalvi, S.V. Combining Ultrasound and Capillary-Embedded T-Junction Microfluidic Devices to Scale Up the Production of Narrow-Sized Microbubbles through Acoustic Fragmentation. Langmuir 2022, 38, 10288–10304. [Google Scholar] [CrossRef]
  170. Gañán-Calvo, A.M.; Gordillo, J.M. Perfectly Monodisperse Microbubbling by Capillary Flow Focusing. Phys. Rev. Lett. 2001, 87, 274501. [Google Scholar] [CrossRef] [PubMed]
  171. Garstecki, P.; Stone, H.A.; Whitesides, G.M. Mechanism for Flow-Rate Controlled Breakup in Confined Geometries: A Route to Monodisperse Emulsions. Phys. Rev. Lett. 2005, 94, 164501. [Google Scholar] [CrossRef]
  172. Anna, S.L.; Bontoux, N.; Stone, H.A. Formation of Dispersions Using “Flow Focusing” in Microchannels. Appl. Phys. Lett. 2003, 82, 364–366. [Google Scholar] [CrossRef]
  173. Wang, H.; Jiang, S.; Zhu, C.; Ma, Y.; Fu, T. Bubble Formation in T-Junctions within Parallelized Microchannels: Effect of Viscoelasticity. Chem. Eng. J. 2021, 426, 131783. [Google Scholar] [CrossRef]
  174. Dixon, A.J.; Dhanaliwala, A.H.; Chen, J.L.; Hossack, J.A. Enhanced Intracellular Delivery of a Model Drug Using Microbubbles Produced by a Microfluidic Device. Ultrasound Med. Biol. 2013, 39, 1267–1276. [Google Scholar] [CrossRef] [PubMed]
  175. Peyman, S.A.; Abou-Saleh, R.H.; McLaughlan, J.R.; Ingram, N.; Johnson, B.R.G.; Critchley, K.; Freear, S.; Evans, J.A.; Markham, A.F.; Coletta, P.L.; et al. Expanding 3D Geometry for Enhanced On-Chip Microbubble Production and Single Step Formation of Liposome Modified Microbubbles. Lab Chip 2012, 12, 4544. [Google Scholar] [CrossRef] [PubMed]
  176. Klibanov, A.L. Ultrasound Molecular Imaging of Cancer: Design and Formulation Strategies of Targeted Contrast Agents. In Molecular Imaging in Oncology; Schober, O., Kiessling, F., Debus, J., Eds.; Recent Results in Cancer Research; Springer International Publishing: Cham, Switzerland, 2020; Volume 216, pp. 319–336. ISBN 978-3-030-42617-0. [Google Scholar]
  177. Klibanov, A.L. Ligand-Carrying Gas-Filled Microbubbles: Ultrasound Contrast Agents for Targeted Molecular Imaging. Bioconjugate Chem. 2005, 16, 9–17. [Google Scholar] [CrossRef]
  178. Sambi, M.; Bagheri, L.; Szewczuk, M.R. Current Challenges in Cancer Immunotherapy: Multimodal Approaches to Improve Efficacy and Patient Response Rates. J. Oncol. 2019, 2019, 1–12. [Google Scholar] [CrossRef] [PubMed]
  179. Cruz, E.; Kayser, V. Monoclonal Antibody Therapy of Solid Tumors: Clinical Limitations and Novel Strategies to Enhance Treatment Efficacy. BTT 2019, 13, 33–51. [Google Scholar] [CrossRef]
  180. Hansel, T.T.; Kropshofer, H.; Singer, T.; Mitchell, J.A.; George, A.J.T. The Safety and Side Effects of Monoclonal Antibodies. Nat. Rev. Drug Discov. 2010, 9, 325–338. [Google Scholar] [CrossRef]
  181. Fares, C.M.; Van Allen, E.M.; Drake, C.G.; Allison, J.P.; Hu-Lieskovan, S. Mechanisms of Resistance to Immune Checkpoint Blockade: Why Does Checkpoint Inhibitor Immunotherapy Not Work for All Patients? Am. Soc. Clin. Oncol. Educ. Book 2019, 39, 147–164. [Google Scholar] [CrossRef]
  182. Singh, S.; Hassan, D.; Aldawsari, H.M.; Molugulu, N.; Shukla, R.; Kesharwani, P. Immune Checkpoint Inhibitors: A Promising Anticancer Therapy. Drug Discov. Today 2020, 25, 223–229. [Google Scholar] [CrossRef]
  183. Rohaan, M.W.; Wilgenhof, S.; Haanen, J.B.A.G. Adoptive Cellular Therapies: The Current Landscape. Virchows Arch. 2019, 474, 449–461. [Google Scholar] [CrossRef]
  184. Fu, C.; Shi, G.; Liu, Y.-T. Manufacturing Anti-CD19 CAR-Tscm Cells for Immunotherapy Using Innovative Microbubble-Based Technologies for Precision Cell Processing. Blood 2021, 138, 3889. [Google Scholar] [CrossRef]
  185. Lustig, A.; Manor, T.; Shi, G.; Li, J.; Wang, Y.-T.; An, Y.; Liu, Y.-T.; Weng, N. Lipid Microbubble–Conjugated Anti-CD3 and Anti-CD28 Antibodies (Microbubble-Based Human T Cell Activator) Offer Superior Long-Term Expansion of Human Naive T Cells In Vitro. ImmunoHorizons 2020, 4, 475–484. [Google Scholar] [CrossRef] [PubMed]
  186. Waldmann, T.A. Cytokines in Cancer Immunotherapy. Cold Spring Harb. Perspect. Biol. 2018, 10, a028472. [Google Scholar] [CrossRef]
  187. Berraondo, P.; Sanmamed, M.F.; Ochoa, M.C.; Etxeberria, I.; Aznar, M.A.; Pérez-Gracia, J.L.; Rodríguez-Ruiz, M.E.; Ponz-Sarvise, M.; Castañón, E.; Melero, I. Cytokines in Clinical Cancer Immunotherapy. Br. J. Cancer 2019, 120, 6–15. [Google Scholar] [CrossRef]
  188. Qiu, Y.; Su, M.; Liu, L.; Tang, Y.; Pan, Y.; Sun, J. Clinical Application of Cytokines in Cancer Immunotherapy. DDDT 2021, 15, 2269–2287. [Google Scholar] [CrossRef]
  189. Figueiredo, M.L.; Figueiredo Neto, M.; Salameh, J.W.; Decker, R.E.; Letteri, R.; Chan-Seng, D.; Emrick, T. Ligand-Mediated Targeting of Cytokine Interleukin-27 Enhances Its Bioactivity In Vivo. Mol. Ther.-Methods Clin. Dev. 2020, 17, 739–751. [Google Scholar] [CrossRef]
  190. Barua, A.; Yellapa, A.; Bahr, J.M.; Adur, M.K.; Utterback, C.W.; Bitterman, P.; Basu, S.; Sharma, S.; Abramowicz, J.S. Interleukin 16- (IL-16-) Targeted Ultrasound Imaging Agent Improves Detection of Ovarian Tumors in Laying Hens, a Preclinical Model of Spontaneous Ovarian Cancer. BioMed Res. Int. 2015, 2015, 1–10. [Google Scholar] [CrossRef] [PubMed]
  191. Malonis, R.J.; Lai, J.R.; Vergnolle, O. Peptide-Based Vaccines: Current Progress and Future Challenges. Chem. Rev. 2020, 120, 3210–3229. [Google Scholar] [CrossRef]
  192. Amara, S.; Tiriveedhi, V. The Five Immune Forces Impacting DNA-Based Cancer Immunotherapeutic Strategy. Int. J. Mol. Sci. 2017, 18, 650. [Google Scholar] [CrossRef]
  193. Saxena, M.; Balan, S.; Roudko, V.; Bhardwaj, N. Towards Superior Dendritic-Cell Vaccines for Cancer Therapy. Nat. Biomed. Eng. 2018, 2, 341–346. [Google Scholar] [CrossRef]
  194. Lin, M.J.; Svensson-Arvelund, J.; Lubitz, G.S.; Marabelle, A.; Melero, I.; Brown, B.D.; Brody, J.D. Cancer Vaccines: The next Immunotherapy Frontier. Nat. Cancer 2022, 3, 911–926. [Google Scholar] [CrossRef]
  195. Dewitte, H.; Van Lint, S.; Heirman, C.; Thielemans, K.; De Smedt, S.C.; Breckpot, K.; Lentacker, I. The Potential of Antigen and TriMix Sonoporation Using MRNA-Loaded Microbubbles for Ultrasound-Triggered Cancer Immunotherapy. J. Control. Release 2014, 194, 28–36. [Google Scholar] [CrossRef]
  196. Song, H.-W.; Lee, H.-S.; Kim, S.-J.; Kim, H.Y.; Choi, Y.H.; Kang, B.; Kim, C.-S.; Park, J.-O.; Choi, E. Sonazoid-Conjugated Natural Killer Cells for Tumor Therapy and Real-Time Visualization by Ultrasound Imaging. Pharmaceutics 2021, 13, 1689. [Google Scholar] [CrossRef]
  197. Navarro-Becerra, J.A.; Franco-Urquijo, C.A.; Ríos, A.; Escalante, B. Localized Delivery of Caveolin-1 Peptide Assisted by Ultrasound-Mediated Microbubble Destruction Potentiates the Inhibition of Nitric Oxide-Dependent Vasodilation Response. Ultrasound Med. Biol. 2021, 47, 1559–1572. [Google Scholar] [CrossRef] [PubMed]
  198. Navarro-Becerra, J.A.; Caballero-Robledo, G.A.; Franco-Urquijo, C.A.; Ríos, A.; Escalante, B. Functional Activity and Endothelial-Lining Integrity of Ex Vivo Arteries Exposed to Ultrasound-Mediated Microbubble Destruction. Ultrasound Med. Biol. 2020, 46, 2335–2348. [Google Scholar] [CrossRef]
  199. Chen, H.; Brayman, A.A.; Evan, A.P.; Matula, T.J. Preliminary Observations on the Spatial Correlation Between Short-Burst Microbubble Oscillations and Vascular Bioeffects. Ultrasound Med. Biol. 2012, 38, 2151–2162. [Google Scholar] [CrossRef] [PubMed]
  200. Chen, H.; Brayman, A.A.; Bailey, M.R.; Matula, T.J. Blood Vessel Rupture by Cavitation. Urol. Res. 2010, 38, 321–326. [Google Scholar] [CrossRef] [PubMed]
  201. Li, P.; Armstrong, W.F.; Miller, D.L. Impact of Myocardial Contrast Echocardiography on Vascular Permeability: Comparison of Three Different Contrast Agents. Ultrasound Med. Biol. 2004, 30, 83–91. [Google Scholar] [CrossRef]
  202. Miller, D.L.; Lu, X.; Fabiilli, M.; Fields, K.; Dou, C. Frequency Dependence of Petechial Hemorrhage and Cardiomyocyte Injury Induced during Myocardial Contrast Echocardiography. Ultrasound Med. Biol. 2016, 42, 1929–1941. [Google Scholar] [CrossRef]
  203. White, P.J.; Zhang, Y.-Z.; Power, C.; Vykhodtseva, N.; McDannold, N. Observed Effects of Whole-Brain Radiation Therapy on Focused Ultrasound Blood–Brain Barrier Disruption. Ultrasound Med. Biol. 2020, 46, 1998–2006. [Google Scholar] [CrossRef]
  204. Sun, T.; Samiotaki, G.; Wang, S.; Acosta, C.; Chen, C.C.; Konofagou, E.E. Acoustic Cavitation-Based Monitoring of the Reversibility and Permeability of Ultrasound-Induced Blood-Brain Barrier Opening. Phys. Med. Biol. 2015, 60, 9079–9094. [Google Scholar] [CrossRef]
  205. Anderson, N.M.; Simon, M.C. The Tumor Microenvironment. Curr. Biol. 2020, 30, R921–R925. [Google Scholar] [CrossRef] [PubMed]
  206. Hernot, S.; Klibanov, A.L. Microbubbles in Ultrasound-Triggered Drug and Gene Delivery. Adv. Drug Deliv. Rev. 2008, 60, 1153–1166. [Google Scholar] [CrossRef] [PubMed]
  207. Jallinoja, V.I.J.; Houghton, J.L. Current Landscape in Clinical Pretargeted Radioimmunoimaging and Therapy. J. Nucl. Med. 2021, 62, 1200–1206. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Illustration of a typical engineered microbubble used in biomedical applications. Microbubbles are particles filled with gas, such as air, oxygen (O2), nitric oxide (NO) or perfluorocarbon (PFC), with a diameter of 1–10 μm. They are stabilized by shells made of phospholipid (3–5 nm thick), protein (15–150 nm thick) or polymer (50–500 nm thick).
Figure 1. Illustration of a typical engineered microbubble used in biomedical applications. Microbubbles are particles filled with gas, such as air, oxygen (O2), nitric oxide (NO) or perfluorocarbon (PFC), with a diameter of 1–10 μm. They are stabilized by shells made of phospholipid (3–5 nm thick), protein (15–150 nm thick) or polymer (50–500 nm thick).
Pharmaceutics 15 01625 g001
Figure 2. Schematic overview of the underlying physical and biological effects produced during microbubble cavitation on the cell membrane. (A,B) Mechanical effects associated with sonoporation, including stable and inertial cavitation of microbubbles. (C) Biological effects induced by microbubble cavitation, which enhance the internalization of drugs.
Figure 2. Schematic overview of the underlying physical and biological effects produced during microbubble cavitation on the cell membrane. (A,B) Mechanical effects associated with sonoporation, including stable and inertial cavitation of microbubbles. (C) Biological effects induced by microbubble cavitation, which enhance the internalization of drugs.
Pharmaceutics 15 01625 g002
Figure 3. Effect of microbubble volume dose (MVD) on microbubble performance. (A) The use of MVD as a dose metric can maximize US imaging contrast (a) and circulation persistence (b) for isolated microbubbles (MBs) of different diameters. Adapted with permission from reference [112]. Copyright © 2023 World Federation for Ultrasound in Medicine & Biology. Published by Elsevier Inc. All rights reserved. (B) The use of MVD results in similar and consistent delivery of Evans blue across the blood–brain barrier (a) compared to using MB concentration as dose metric (b). The delivery effect of size distribution and concentration of MBs can be collapsed by using MVD. Best fit trend lines, determined by linear regression analysis, are represented by black lines for the 2-µm and 6-µm data sets during the initial 5-minute period of BBB disruption (solid circles; right striatum) and the subsequent 5-minute period (empty circles; left striatum). * p < 0.0001. Adapted with permission from reference [116]. (C) In vivo pharmacokinetics of size-isolated MBs demonstrated that MVD has greater impact on half-life (a) and area-under-the-curve (b) than size. * p < 0.05 and ** p ≤ 0.01. Adapted with permission from [117]. Copyright © 2023, American Chemical Society. (D) The in vitro harmonic (a) and broadband (b) cavitation dose (HCD and BCD) versus mechanical index versus gas volume fraction can be aligned by matching the MVD for 2, 3, and 5 μm diameter and polydisperse MBs. Adapted with permission from [37]. © 2023 by the authors.
Figure 3. Effect of microbubble volume dose (MVD) on microbubble performance. (A) The use of MVD as a dose metric can maximize US imaging contrast (a) and circulation persistence (b) for isolated microbubbles (MBs) of different diameters. Adapted with permission from reference [112]. Copyright © 2023 World Federation for Ultrasound in Medicine & Biology. Published by Elsevier Inc. All rights reserved. (B) The use of MVD results in similar and consistent delivery of Evans blue across the blood–brain barrier (a) compared to using MB concentration as dose metric (b). The delivery effect of size distribution and concentration of MBs can be collapsed by using MVD. Best fit trend lines, determined by linear regression analysis, are represented by black lines for the 2-µm and 6-µm data sets during the initial 5-minute period of BBB disruption (solid circles; right striatum) and the subsequent 5-minute period (empty circles; left striatum). * p < 0.0001. Adapted with permission from reference [116]. (C) In vivo pharmacokinetics of size-isolated MBs demonstrated that MVD has greater impact on half-life (a) and area-under-the-curve (b) than size. * p < 0.05 and ** p ≤ 0.01. Adapted with permission from [117]. Copyright © 2023, American Chemical Society. (D) The in vitro harmonic (a) and broadband (b) cavitation dose (HCD and BCD) versus mechanical index versus gas volume fraction can be aligned by matching the MVD for 2, 3, and 5 μm diameter and polydisperse MBs. Adapted with permission from [37]. © 2023 by the authors.
Pharmaceutics 15 01625 g003
Figure 4. Schematic illustration of current strategies to improve the performance of targeted microbubbles (tMBs) in therapeutic applications. (A) Microbubble targeting can be achieved through electrostatic interactions or through the use of PEGylated lipids linked with conjugation groups such as thiol-maleimide, disulfide, DBCO-azide, folate or biotin. (B) Ligand–receptor targeting of drug, gene or cell-loaded MBs with antibodies, peptides or vitamins can improve delivery at the target site. (C) Dual targeting of loaded MBs with antibody-antibody, antibody-peptide, peptide-peptide, or peptide-vitamin conjugation can improve delivery efficiency and spatial targeting. (D) MBs conjugated with molecular or magnetic targeted carriers, such as nanoparticles, liposomes, and nanodroplets, can increase drug loading capacity and delivery efficiency at the target site.
Figure 4. Schematic illustration of current strategies to improve the performance of targeted microbubbles (tMBs) in therapeutic applications. (A) Microbubble targeting can be achieved through electrostatic interactions or through the use of PEGylated lipids linked with conjugation groups such as thiol-maleimide, disulfide, DBCO-azide, folate or biotin. (B) Ligand–receptor targeting of drug, gene or cell-loaded MBs with antibodies, peptides or vitamins can improve delivery at the target site. (C) Dual targeting of loaded MBs with antibody-antibody, antibody-peptide, peptide-peptide, or peptide-vitamin conjugation can improve delivery efficiency and spatial targeting. (D) MBs conjugated with molecular or magnetic targeted carriers, such as nanoparticles, liposomes, and nanodroplets, can increase drug loading capacity and delivery efficiency at the target site.
Pharmaceutics 15 01625 g004
Figure 5. Importance of microbubble volume dose (MVD) as a unifying dose metric. Example of size characteristics for three size-isolated lipid-coated microbubbles: (AC) microscope images of MBs from the three different sizes; (D) number-weighted size distributions; (E) volume-weighted size distributions; (F) MB volume vs. concentration curve used to determine the gas volume fraction by taking the area-under-the-curve; (G) corresponding size metrics. Adapted with permission from [117]. Copyright © 2023, American Chemical Society.
Figure 5. Importance of microbubble volume dose (MVD) as a unifying dose metric. Example of size characteristics for three size-isolated lipid-coated microbubbles: (AC) microscope images of MBs from the three different sizes; (D) number-weighted size distributions; (E) volume-weighted size distributions; (F) MB volume vs. concentration curve used to determine the gas volume fraction by taking the area-under-the-curve; (G) corresponding size metrics. Adapted with permission from [117]. Copyright © 2023, American Chemical Society.
Pharmaceutics 15 01625 g005
Table 2. Microbubbles targeted with a therapeutic ligand used for therapy.
Table 2. Microbubbles targeted with a therapeutic ligand used for therapy.
Ligand TypeLigand
Species
Conjugation ChemistryMicrobubble CompositionUltrasound ParametersAnimal/Disease ModelReference
ProteinSDF-1αEDC/sulfo-NHSShell: DSPC, DPPG,
PEG4000-COOH
Gas: C3F8
Diameter: 1–5 μm
Dose: unknown
f = 4 MHz
MI = 1.5
Time = 1.15 min
Rat/Diabetic nephropathy[119]
GPVI-FcAvidin-biotinMicroMarker™
Shell: Phospholipid
Gas: C4F10/N2
Diameter: 2.3–2.9 μm
Dose: 2.8 × 108 MBs/kg or 5.3 µL/kg MVD
f = 24 MHz
Time = 2 min
Mouse/
Atherosclerosis
[120]
AntibodyAβ 1-42Avidin-biotinShell: DSPC, DPPA, PEG4000
Gas: C3F8
Diameter: 2.13 μm
Dose: 5.8 × 108 MBs/kg or 2.9 µL/kg MVD
f = 3 MHz
MI = 0.8
PRF = 50 Hz
Time = 5 min
Mouse/Alzheimer’s disease[20]
C4d mAbAvidin-biotinShell: DPPC, DSPE-PEG-2000, -Biotin
Gas: C3F8/NO
Diameter: 0.96 ± 0.07 μm
Dose: 5.8 × 106 MBs/kg or 0.003 µL/kg MVD
f = 13 MHz
MI = 0.33
Time = 5 min
Rat/Heterotopic Heart transplant[108]
IL-8 mAbAvidin-biotinUSphere™ Labeler
Shell: Phospholipid
Diameter: 1.8 μm
Dose: 1.2 × 109 MBs/kg or 3.6 µL/kg MVD
f = 1 MHz
MI = 1.5
Power = 5 W
Time = 1 min
Rabbit/
Atherosclerosis
[121]
DSPC = 1,2-distearoyl-sn-glycero-3-phosphocholine; DPPG = 1,2-dipalmitoyl-sn-glycero-3-phospho-(1′-rac-glycerol); DPPA = 1,2-dipalmitoyl-sn-glycero-3-phosphate; PEG4000-COOH = polyethylene glycol-4000-carboxy; DSPE-PEG2000 = 1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol)-2000]; C3F8 = perfluoropropane; C4F10 = perfluorobutane; NO = nitric oxide; MVD = Microbubble Volume Dose; f = frequency; MI = mechanical index; PRF = pulse repetition frequency.
Table 3. Drug or gene loaded-MBs targeted with a ligand used for therapy.
Table 3. Drug or gene loaded-MBs targeted with a ligand used for therapy.
Ligand TypeTargeting
Ligand
Therapeutic MoleculeConjugation
Chemistry
Microbubble CompositionUltrasound ParametersAnimal/Disease ModelReference
AntibodyVEGFR2 mAbpHSV-TKAvidin-biotin
Electrostatic
Shell: DPPC, DPTAP,
DSPE-PEG2000, -Biotin
Gas: C3F8
Diameter: 1.1 ± 0.1 μm
Dose: 1.8 × 1010 MBs/kg or 12.4 µL/kg MVD
f = 1 MHz
PNP = 0.7 MPa
PRF = 5 Hz
Time = 2 min
Mouse/Glioma
tumor
[122]
VEGFR2 mAbBCNUAvidin-biotin
Hydrophobic
Shell: DPPC,
DSPE-PEG2000, -Biotin
Gas: C3F8
Diameter: 1.79 ± 0.13 μm
Dose: 1.4 × 1010 MBs/kg or 41 µL/kg MVD
f = 1 MHz
PNP = 0.7 MPa
DC = 5%
PRF = 5 Hz
Time = 1 min
Rat/Glioma
tumor
[123]
CD105pEZ-M46-ESAvidin-biotin
Electrostatic
Shell: DPPC, Cholesterol,
DSPE-PEG2000, -Biotin
Gas: C3F8
Diameter: 1.44 ± 0.21 μm
Dose: 1.3 × 1010 MBs/kg or 20.6 µL/kg MVD
f = 1 MHz
PNP = 0.7 MPa
DC = 50%
I = 2 W/cm2
PRF = 5 Hz
Time = 30 s
Mouse/Hind limb tumor[124]
RGDUrokinaseElectrostaticSonoVue
Diameter: 1.5–2.5 μm
Dose: 4.4 × 107 MBs/kg or 0.18 µL/kg MVD
f = 1.6 MHz
MI = 1.1
PRF = 24 kHz
Time = 10 min
Pig/Thrombosis[125]
VCAM-1miR-126Avidin-biotinVisualSonics Target-Ready MBs
Shell: Phospholipid
Gas: C4F10/N2
Diameter: 1.5 μm
Dose: 3.7 × 109 MBs/kg or 6.6 µL/kg MVD
f = 10 MHz
MI = 0.66
I = 0.076 W/cm2
Time = 2.5 min
Mouse/Aortic
Aneurysm
[126]
PeptidesLHRaPaclitaxelAvidin-biotin
Hydrophobic
Shell: DPPC, DSPE-PEG2000-Biotin
Gas: C3F8
Diameter: 1.8 ± 0.2 μm
Dose: 1.5 × 109 MBs/kg or 4.5 µL/kg MVD
f = 0.3 MHz
I = 1 W/cm2
DC = 50%
Time = 3 min
Mouse/Breast
cancer
[127]
RGDtPAAmine
Hydrophobic
Shell: DPPC, DSPC
PEG-Amine
Gas: C3F8
Diameter: 2.08 ± 0.93 μm
Dose: 8.0 × 108 MBs/kg or 3.8 µL/kg MVD
f = 2 MHz
MI = 1.4
I = 1.8 W/cm2
DC = 95%
PRF = 15 Hz
Time = 30 min
Rabbit/
Thrombolysis
[128]
VitaminsFolatePaclitaxelHydrophobicShell: DPPC,
DSPE-PEG2000-Folate
Gas: O2
Diameter: 1.81 ± 0.04 μm
Dose: unknown
f = 0.3 MHz
I = 1 W/cm2
Time = 3 min
Mouse/Ovarian cancer
[105]
Folatep-FLucAmide
Electrostatic
Shell: DPPC, DPTAP,
DSPE-PEG2000
Gas: C3F8
Diameter: 3.2 ± 0.1 μm
Dose: 1.8 × 108 MBs/kg or 3.1 µL/kg MVD
f = 1 MHz
PRF = 5 Hz
PNP = 0.7 MPa
Time = 1 min
Rat/
Glioma tumor
[129]
DPPC = 1,2-dipalmitoyl-sn-glycero-3-phosphocholine; DSPE-PEG2000 = 1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol) -2000]; DPPG = 1,2-dipalmitoyl-sn-glycero-3-phospho-(1′-rac-glycerol); PA = palmitic acid; DSPC = 1,2-distearoyl-sn-glycero-3-phosphocholine; DPTAP = 1,2-dipalmitoyl-3-trimethylammonium-propane; C3F8 = perfluoropropane; SF6 = sulfur hexafluoride, C4F10 = perfluorobutane; N2 = nitrogen; O2 = oxygen; MVD = Microbubble Volume Dose; f = frequency; MI = mechanical index; PRF = pulse repetition frequency; I = intensity; DC = duty cycle; PNP = peak negative pressure.
Table 4. Drug, gene, or cell loaded-MBs targeted with two different ligands.
Table 4. Drug, gene, or cell loaded-MBs targeted with two different ligands.
Targeting
Ligands
Therapeutic
Molecule
Conjugation ChemistryMicrobubble CompositionUltrasound ParametersAnimal/Disease ModelReference
CD90 Ab
ICAM-1 Ab
Adipose-derived stem cellsAvidin-biotin
Shell: DSPC, PEG40S,
DSPE-PEG2000, -Biotin
Gas: C4F10
Diameter: 3.5 μm
Dose: 9.3 × 107 MBs/kg or 2.1 µL/kg MVD
f = 1 MHz
PNP: 0.1 MPa
PRF = 1 kHz
DC = 50%
Time = 10 min
Rat/Myocardial
Infarction
[130]
CD90 Ab
ICAM-1 Ab
Adipose-derived stem cellsAvidin-biotinShell: DSPC, PEG40S,
DSPE-PEG2000, -Biotin
Gas: C4F10
Diameter: 3.5 μm
Dose: 4.6 × 107 MBs/kg or 1 µL/kg MVD
f = 1 MHz
PNP: 0.1 MPa
PRF = 1 kHz
DC = 50%
Time = 1 min
Mouse/
Atherosclerosis
[131]
CCR2 Ab
iRGD peptide
shAKT2Avidin-biotin,
Electrostatic
Shell: DSPC, Stearic-PEI600,
DSPE-PEG2000-iRGD-Biotin
Gas: C4F10
Diameter: 1.32 ± 0.22 μm
Dose: 5.3 × 109 MBs/kg or 6.3 µL/kg MVD
f = 1 MHz
PNP: 1.2 MPa
PRF = 1 kHz
DC = 50%
Time = 10 min
Mouse/Breast
cancer
[132]
cRGD peptide
Folate
DoxorubicinAvidin-biotinShell: DSPC,
DSPE-PEG2000, -Biotin
Gas: C3F8
Diameter: 5.8 ± 2.1 μm
Dose: 1.1 × 1010 MBs/kg or 1.1 mL/kg MVD
f = 10 MHz
MI: 0.64
Time = 0.5 min
Mouse/Breast
cancer
[133]
cRGD and cCLT1 peptidePaclitaxelAmide,
Hydrophobic
Shell: DSPC,
DSPE-PEG2000, -COOH
Gas: SF6
Diameter: 1.59 ± 0.54 μm
Dose: 1.0×1010 MBs/kg or 21 µL/kg MVD
f = 1 MHz
MI = 1.17
DC = 10%
I = 2.5 W/cm2
Time = 3 min
Mouse/Pancreatic cancer[134]
DSPC = 1,2-distearoyl-sn-glycero-3-phosphocholine; PEG-40S = Polyoxyethylene (40) stearate; DSPE-PEG2000 = 1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol)-2000]; PEI = Branched Polyethylenimine; C3F8 = perfluoropropane; SF6 = sulfur hexafluoride, C4F10 = perfluorobutane; MVD= Microbubble Volume Dose; f = frequency; MI = mechanical index; PRF = pulse repetition frequency; I = intensity; DC = duty cycle; PNP = peak negative pressure.
Table 5. Microbubbles conjugated with targeted drug-loaded carriers used for therapy.
Table 5. Microbubbles conjugated with targeted drug-loaded carriers used for therapy.
CarrierType of
Targeting
Therapeutic
Molecule
Conjugation ChemistryMicrobubble CompositionUltrasound
Parameters
Animal/Disease ModelReference
NanoparticlesMagnetic5-fluorouracil
Rose Bengal
Avidin-biotinShell: DBPC,
DSPE-PEG2000, -Biotin
Gas: O2
Diameter: 1–2 μm
Dose: 7.0 × 109 MBs/kg or 12.3 µL/kg MVD
f = 1 MHz
PNP = 0.85 MPa
DC = 30%
I = 3.5 W/cm2
PRF = 100 Hz
Time = 3.5 min
Mouse/Pancreatic tumor[106]
MagneticDoxorubicinAmideShell: DSPC, DSPG
DSPE-PEG2000, -Biotin
Gas: C3F8
Diameter: 5.4 ± 1.1 μm
Dose: 5.7 × 108 MBs/kg or 46.7 µL/kg MVD
f = 1 MHz
PNP = 0.3 MPa
DC = 30%
I = 3.5 W/cm2
PRF = 1 Hz
Time = 4 min
Rat/Glioma tumor[107]
MagneticGemcitabine
Rose Bengal
Avidin-biotinShell: DBPC,
DSPE-PEG2000, -Biotin
Gas: O2
Diameter: 1.9 ± 0.4 μm
Dose: 5.4 × 109 MBs/kg or 19.5 µL/kg MVD
f = 1.17 MHz
PNP = 0.7 MPa
DC = 30%
Time = 3.5 min
Mouse/Pancreatic tumor[135]
MagneticDoxorubicinElectrostatic and
hydrophobic
Shell: DSPC, DSPG
DSPE-PEG2000, -Biotin
Gas: C3F8
Diameter: 1.04 ± 0.01 μm
Dose: 9.7 × 1010 MBs/kg or 56.8 µL/kg MVD
f = 0.4 MHz
PNP = 0.325 MPa
Power = 4 W
PRF = 1 Hz
Time = 1.5 min
Rat/Glioma tumor[136]
MagnetictPAElectrostaticShell: SDS
Gas: Air
Diameter: 5.36 ± 1.44 μm
Dose: 6.4 × 1010 MBs/kg or 0.50 µL/kg MVD
f = 18 MHz
DC = 10%
Time = 5 min
Mouse/
Thrombolysis
[137]
Magnetic/
molecular
RGDAmideShell: PVA
Gas: Air
Diameter: 1.37 μm
Dose: 5.4 × 108 MBs/kg or 0.70 µL/kg MVD
f = 30 MHz
Time = 10 min
Mouse/Colon cancer[138]
LiposomesMagneticDoxorubicinMaleimideShell: DPPC,
DSPE-PEG2000-SPDP
Gas: C3F8
Diameter: 4 μm
Dose: 7.0 × 109 MBs/kg or 233 µL/kg MVD
f = 1 MHz
DC = 30%
I = 2 W/cm2
Time = 2 min
Mouse/Pancreatic tumor[139]
Molecular: RGD peptidePaclitaxelAvidin-biotinShell: DSPC,
DSPE-PEG2000, -Biotin
Gas: C3F8
Diameter: 1.5 μm
Dose: 1.1 × 1010 MBs/kg or 17.8 µL/kg MVD
f = 1 MHz
DC = 1%
PRF = 1 Hz
Time = 2 min
Mouse/Breast
cancer
[140]
Molecular:
NGR peptide
shBirc5
Avidin-biotinShell: DPPC, Cholesterol
DSPE-PEG2000-Biotin
Gas: C3F8
Diameter: 2.90 ± 0.38 μm
Dose: 2.1 × 1010 MBs/kg or 265.6 µL/kg MVD
f = 1 MHz
DC = 50%
Power = 1.84 W
Time = 1–5 min
Rat/Glioma tumor[141]
Molecular:
VEGFR2 Ab
IrinotecanAvidin-biotinMicroMarker™
Shell: Phospholipid
Gas: C4F10/N2
Diameter: 1.5 μm
Dose: 4.7 × 109 MBs/kg or 8.2 µL/kg MVD
f = 2.2 MHz
PNP = 0.26 MPa
PRF = 1 Hz
Time = 4 min
Mouse/Colorectal cancer[142]
NanodropletsMolecular:
VEGFR2 Ab
Combretastatin A4Avidin-biotinShell: DPPC, DSPE-PEG2000-Biotin
Gas: C4F10
Diameter: 2.6 ± 1.5 μm
Dose: unknown
f = 2.2 MHz
PNP = 0.26 MPa
PRF = 1 Hz
Time = 5 s
Mouse/Colorectal cancer[143]
DBPC:1,2-dibehenoyl-sn-glycero-3-phosphocholine; DSPG: 1,2-Distearoyl-sn-glycero-3-phospho-rac-glycerol; DSPE-PEG2000: 1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol) -2000]; DSPC: 1,2-distearoyl-sn-glycero-3-phosphocholine; SDS: Sodium Dodecyl Sulfate; PVA = Poly(vinyl alcohol); DPPC = 1,2-dipalmitoyl-sn-glycero-3-phosphocholine; C3F8 = perfluoropropane; C4F10 = perfluorobutane; O2 = oxygen; MVD = Microbubble Volume Dose; f = frequency; PRF = pulse repetition frequency; I = intensity; DC = duty cycle; PNP = peak negative pressure.
Table 6. Targeted microbubbles used for cancer immunotherapy.
Table 6. Targeted microbubbles used for cancer immunotherapy.
Immunotherapy
Modality
Therapeutic
Molecule
Conjugation ChemistryMicrobubbleCompositionUltrasound
Parameters
Animal/Disease ModelReference
Monoclonal
antibody
EGFR mAbAvidin-biotinTargestar™-SA
Shell: Phospholipid
Gas: C4F10
Diameter: 2.5 μm
Dose: 3.6 × 109 MBs/kg or 29.7 µL/kg MVD
f = 400 kHz
PRF = 1 Hz
Power = 5 W
Time = 3–4 min
Mouse/Glioma
tumor
[144]
DLL4 mAbAvidin-biotinTargestar™-SA
Shell: Phospholipids
Gas: C4F10
Diameter: 2 μm
Dose: 5.3 × 108 MBs/kg or 2.2 µL/kg MVD
f = 1 MHz
DC= 50%
I = 2 W/cm2
Time = 1.5 min
Mouse/Gastric
cancer
[145]
Trastuzumab mAbNHSShell: DPSC, DSPE-PEG-2000-NHS, Cholesterol and pyropheophorbide
Gas: SF6
Diameter: 1.654 ± 1.07 μm
Dose: unknown
f = 1 MHz
DC= 50%
I = 2 W/cm2
Time = 5 min
Mouse/Gastric
cancer
[146]
Immune checkpoint inhibitorsPDL-1 mAbNHSShell: DSPC,
DSPE-PEG-2000-NHS.
Gas: SF6
Diameter: 1.06 ± 0.31 μm
Dose: 6.9 × 1010 MBs/kg or 43.3 µL/kg MVD
f = 1.1 MHz
DC = 5%
PRF = 100 Hz
Time = 0.5 min
Mouse/Colon cancer[147]
PDL-1 mAb and CisplatinAvidin-biotin
Unbounded
Shell: DPSC, DSPE-PEG-2000, -Biotin
Gas: C3F8
Diameter: 1.01 ± 0.14 μm
Dose: 1 × 108 MBs/kg or 0.06 µL/kg MVD
f = 1 MHz
DC = 50%
I = 1W/cm2
PRF = 1 kHz
Time = 1.5 min
Mouse/Cervical
cancer
[148]
PDL-1 mAb and miR-34aAvidin-biotin
Electrostatic
Shell: DSPC, DSPE-PEG-2000, DSPE-PEG-2000-Biotin, PEI-600
Gas: C3F8
Diameter: 0.940 ± 0.080 μm
Dose: 4 × 109 MBs/kg or 1.7 µL/kg MVD
f = 18 MHz
DC = 50%
I = 1 W/cm2
Time = 1.5 min
Mouse/Cervical
cancer
[149]
VaccineCD11b mAb and
CGAMP
MaleimideShell: DSPC, DSPE-PEG-2000,
DSPE-PEG-5000-Maleimide
Gas: C4F10
Diameter: 2.6 μm
Dose: 1.4 × 109 MBs/kg or 13.1 µL/kg MVD
f = 1 MHz
DC = 50%
I = 4 W/cm2
Time = 2 min
Mouse/Breast cancer[150]
HSP70-MAGEA1ElectrostaticShell: Span 60 and Tween 80
Gas: SF6
Diameter: 6 μm
Dose: 1.3 × 109 MBs/kg or 144.7 µL/kg MVD
MI = 0.75Mouse/Melanoma tumor[151]
Dendritic cell plasma membrane fragmentsHydrophobicShell: DPPC, DPPA, DSPE-PEG5000
Gas: C3F8
Diameter: 1.21  ±  1.0  μm
Dose: 5 × 108 MBs/kg or 0.5 µL/kg MVD
f = 18 MHz
MI = 0.75
Time = 5 min
Mouse/Breast cancer[152]
DSPC = 1,2-distearoyl-sn-glycero-3-phosphocholine; DPPA: 1,2-dipalmitoyl-sn-glycero-3-phosphate, DPPC: (1,2-dipalmitoyl-sn-glycero-3-phosphocholine); DSPE: 1,2-Distearoyl-sn-glycero-3-phosphorylethanolamine; PEI: Branched Polyethylenimine; DSPE-PEG-2000: (1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-methoxypolyethylene glycol)-2000]; DSPE-PEG-5000: (1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-methoxypolyethylene glycol)-5000]; C3F8 = perfluoropropane; SF6 = sulfur hexafluoride, C4F10 = perfluorobutane; MVD = Microbubble Volume Dose; f = frequency; DC = duty cycle; I = intensity; PRF = pulse repetition frequency; MI = mechanical index.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Navarro-Becerra, J.A.; Borden, M.A. Targeted Microbubbles for Drug, Gene, and Cell Delivery in Therapy and Immunotherapy. Pharmaceutics 2023, 15, 1625. https://doi.org/10.3390/pharmaceutics15061625

AMA Style

Navarro-Becerra JA, Borden MA. Targeted Microbubbles for Drug, Gene, and Cell Delivery in Therapy and Immunotherapy. Pharmaceutics. 2023; 15(6):1625. https://doi.org/10.3390/pharmaceutics15061625

Chicago/Turabian Style

Navarro-Becerra, J. Angel, and Mark A. Borden. 2023. "Targeted Microbubbles for Drug, Gene, and Cell Delivery in Therapy and Immunotherapy" Pharmaceutics 15, no. 6: 1625. https://doi.org/10.3390/pharmaceutics15061625

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop