Next Article in Journal
Longitudinal Serum Protein Analysis of Women with a High Risk of Developing Breast Cancer Reveals Large Interpatient Versus Small Intrapatient Variations: First Results from the TESTBREAST Study
Previous Article in Journal
Tropaeolum majus R2R3 MYB Transcription Factor TmPAP2 Functions as a Positive Regulator of Anthocyanin Biosynthesis
Previous Article in Special Issue
Magnetic Micellar Nanovehicles: Prospects of Multifunctional Hybrid Systems for Precision Theranostics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Current Nanomedicine for Targeted Vascular Disease Treatment: Trends and Perspectives

1
National Institute of Medical Welfare, Kangnam University, Yongin 16979, Korea
2
Department of Biotechnology, The University of Suwon, Suwon 18323, Korea
*
Authors to whom correspondence should be addressed.
These authors equally contributed to this work.
Int. J. Mol. Sci. 2022, 23(20), 12397; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232012397
Submission received: 4 September 2022 / Revised: 4 October 2022 / Accepted: 14 October 2022 / Published: 17 October 2022
(This article belongs to the Special Issue Nanotechnology in Cancer Treatment 3.0)

Abstract

:
Nanotechnology has been developed to deliver cargos effectively to the vascular system. Nanomedicine is a novel and effective approach for targeted vascular disease treatment including atherosclerosis, coronary artery disease, strokes, peripheral arterial disease, and cancer. It has been well known for some time that vascular disease patients have a higher cancer risk than the general population. During atherogenesis, the endothelial cells are activated to increase the expression of adhesion molecules such as Intercellular Adhesion Molecule 1 (ICAM-1), Vascular cell adhesion protein 1 (VCAM-1), E-selectin, and P-selectin. This biological activation of endothelial cells gives a targetability clue for nanoparticle strategies. Nanoparticle formation has a passive targeting pathway due to the increased adhesion molecule expression on the cell surface as well as increased cell activation. In addition, the VCAM-1-targeting peptide has been widely used to target the inflamed endothelial cells. Biomimetic nanoparticles using platelet and leukocyte membrane fragment strategies have been promising techniques for targeted vascular disease treatment. Cyclodextrin, a natural oligosaccharide with a hydrophobic cavity, increase the solubility of cholesterol crystals at the atherosclerotic plaque site and has been used to deliver the hydrophobic drug statin as a therapeutic in a targeted manner. In summary, nanoparticles decorated with various targeting molecules will be an effective and promising strategy for targeted vascular disease treatment.

1. Introduction

Cardiovascular disease (CVD) refers to any disorder that affects the heart or blood vessels. CVD is frequently linked to fatty deposits in the arteries and an elevated risk of blood clots. Atherosclerosis, coronary artery disease (CADs), strokes, peripheral arterial disease, rheumatic heart disease, and other heart and blood vessel disorders are all classified as CVDs [1,2,3,4,5]. They have also been linked to arterial damage in organs such as the brain, heart, kidneys, and eyes [6,7,8,9]. According to the World Health Organization (WHO), CVDs are the leading cause of death worldwide, claiming the lives of 17.9 million people each year. Heart attacks and strokes account for more than four out of every five CVD deaths, with one-third of these deaths occurring before the age of 70 [10]. Current CVD treatment methods include management approaches such as exercise or dietary patterns, traditional medication treatment at the preventive level, and surgery, which places a heavy burden on the patient [11,12,13]. Even though the aforementioned treatments lessen symptoms and decrease mortality rates, both therapeutics have some limitations. Most CVDs are chronic due to tissue inflammation damage accumulated over a long period of time, and therefore, long-term drug administration for therapeutic purposes raises concerns about side effects such as liver failure, the weakening of renal function, and the malfunction of other organs [14,15,16]. In addition, these conventional drug formulations had inherent limitations including low water solubility, potential drug resistance, and poor biological efficiency [17,18,19]. Some patients who suffer from severe progressive vascular diseases need to consider surgical operations. Although significant contributions to reduce the symptoms and mortality of many vascular diseases have been made due to the excellent development of surgical techniques, surgery-related risks including inflammation, restenosis, and long recovery time are still a huge burden on patients. Therefore, establishing novel CVD therapeutics beyond conventional medication methods and surgical interventions are needed to increase the therapeutic effects and reduce the surgery-related risk burden on patients (Figure 1).
In the past decades, nanomedicine has been massively investigated for various disease models including cancer, CVDs, and diabetes [21,22,23,24,25]. Nanomedicine is the use of nanotechnology to accomplish disease treatment at the nanometer (~10–9 m)-size scale. Unlike conventional medicine, it makes the properties of materials suitable for use at the nanometric scale, which often differ from the same substance at a larger size in terms of physics, chemistry, and biology. Furthermore, the nanometric scale is also the size of many biological mechanisms in the human body, allowing nanoparticles (NPs) to potentially cross natural barriers to reach the new delivery sites and interact with DNA or small proteins at various levels, whether in blood or within organs, tissues, or cells. Nanomedicine currently covers a number of healthcare fields including targeted nanotherapeutics [26,27,28], medical imaging [29,30,31], diagnostics [32,33,34], vaccines [35,36], and regenerative medicines [37,38]. In particular, targeted nanotherapeutics can increase the therapeutic effects while reducing off-target side effects. Despite the development of various research using nanomedicine, nanomedicine for CVD purposes is still challenging. Because of the wide range of lesion areas of common vascular diseases, it is complicated to accurately deliver the therapeutic cargos to specific blood vessel lesions. Therefore, a targeted nanomedicine system that can accurately deliver a therapeutic drug to a desired location in a desired amount is expected as a promising treatment method for vascular disease. These approaches will be able to provide a basis for a platform to develop CVD therapeutics (Table 1).
In this review, we will discuss the recent progress of nanomedicine for targeted CVD treatment. Furthermore, current studies, trends, and future perspectives in targeted CVD treatment are also discussed.

2. Properties of NPs for Targeted Drug Delivery and Types of NPs Used for CVD

A number of studies suggested that the physiochemical properties of NPs, including shape, size, composition, charge, and surface chemistry, would typically influence their interaction with biological systems, thus altering their in vivo dissemination [39]. Since NPs are composed of materials designed at the molecular level, they are generally small-scale nanospheres [40]. In this way, they are able to move more spontaneously in the human systems than to big-sized materials. Small-sized nanoscale particles demonstrate exclusive chemical, structural, magnetic, mechanical, and biological characteristics. Particle size and distribution are the main critical features of NPs. They regulate the in vivo distribution, biological consequence, toxicity, and targeting capability of these delivery systems. Moreover, they can affect drug loading and release, and the stability of synthesized NPs. Small-sized particles have a bigger surface area-to-volume ratio; thus, a high number of the drugs made up of smaller particles could lead to prompt drug release, unlike large-sized particles with big cores that permit more of the drug to be encapsulated in each particle, providing a gradual release [41]. Therefore, the regulation of particle size delivers a method of adjusting the drug-release frequency. Along with the particle size, the shape of NPs also affects the intracellular release of therapeutics [42]. Apart from this, the zeta potential of an NP is generally used to distinguish its surface charge property [43]. It reveals the electrical potential of NPs and is affected by the alignment of the NPs and the medium in which they are dispersed. It has been known that NPs with a zeta potential more than ± 30 mV are quite stable in suspension, because the surface charge counteracts their aggregation. Numerous main mechanical characteristics that define the material performance while loading include rigidity, elasticity, hardness strength, and stiffness [44]. Based on the earlier literature, NPs can be categorized as inorganic, organic, or hybrid (Figure 2). While inorganic NPs are favored in shaping diverse functions and characteristics, organic NPs often show better biocompatibility. In the case of organic–inorganic hybrid NPs, the organic functional groups merge the particular features of the inorganic counterparts to develop effective utility for different biological applications [45]. Currently, the application of hybrid NPs for gradual drug release has been attracting increasing interest, especially to enhance the selectivity and efficiency of the drugs through merging the features of inorganic and organic constituents into one single NP system. A nanocarrier is an NP used for the delivery of a cargo, such as a therapeutic particle. The variety of existing NPs for drug delivery is extensive and involves dendrimers, polymeric NPs, carbon nanotubes (CNTs), metallic NPs, or lipid-based structures, for instance liposomes or micelles.

2.1. Organic NPs

2.1.1. Lipid-Based NPs

Generally, lipid-based NP preparations involve solid–lipid NPs, nano-emulsions, and lipid–drug conjugates; each largely consist of physiological lipid analogs with stabilizers such as surfactants [46]. Based on the size of the lipid-based NPs, they are designated as micelles (~10–15 nm), liposomes (~90–100 nm), or polymeric NPs [47,48]. Among these, polymeric NPs have received much attention in CVD nanomedicine owing to their changing nature and feasible reabsorption in the biological system for high efficiency. It is worth mentioning here that polymeric NPs, such as polyglycolic acid, polylactic acid, and polylactic-co-glycolic acid (PLGA), are FDA-permitted polymers. Between these polymers, PLGA has been broadly tested as a drug carrier for CVD treatment [49,50,51]. To date, there are no liposomal drug platforms accessible for the management of CVDs in human systems. However, liposomal alendronate platforms were found to be secure in early-phase clinical trials for infusion during the time of percutaneous coronary interference [52].

2.1.2. Dendrimers

Dendrimers own the benefit of improving the binding space upon adjustment of their outer surface with a few antibodies or ligands for active targeting [53]; moreover, they are able to hold drugs with low solubility [54]. Xue et al. [55] disclosed that the apoptosis induction and infraction size in cardiomyocytes were remarkably suppressed after being given a single intravenous dose of dendrimer, conjugated with a microRNA-1 inhibitor as seen in the myocardial infarcted (MI) mouse system. Currently, NP distribution via intravenous injection with targeting peptides is emerging as a promising therapeutic approach. In agreement, Xue et al. [56] described an initial targeting treatment for MI mouse using the tail vein injection with anti-miR-1 antisense oligonucleotide-loaded myocardium-targeting dendrimer.

2.2. Inorganic NPs

2.2.1. Carbon-Based NPs

CNTs is a well-suited drug carrier for improved diffusion in the cells and also for contributing privileged drug functions [57]. Their electrical, optical, and mechanical characteristics make them a favorable candidate for possible therapeutic applications [58]. Furthermore, some studies have verified the promising possibilities of CNTs in cardiac tissue engineering, for instance, to promote the growth and function of cardiomyocytes [59,60] and stimulate the formation of gap junctions [61,62]. Unlike these works, other examinations have proposed that scaffolds comprising CNT and col-hydrogel might be encouraging injectable biomaterial to distribute cells and drugs for tissue reconstruction in infarcted myocardial tissues [58,63].

2.2.2. Metal NPs

Gold NPs are one of the broadly utilized nanocarriers for the distribution of cardioprotective drugs. The different physiochemical characteristics, for instance, the exceptional stability, greater safety, and robust biocompatibility of several gold NPs, makes them prospective candidates in nano biomedicine [64]. It has been generally acceptable and claimed that conjugated drugs are clinically highly efficient and accurate. Even Simdax has been declared a clinically safe and sanctioned drug for the therapeutics of heart disease. Additionally, Simdax conjugated on gold NPs displays cardioprotective outcomes in vivo rat models with heart failure [65]. In another report, gold NPs given via intravenous injection can also recover myocardial injury in rat models [66]. Comparable to the organic NPs, the intramyocardial injection of a nano-complex (graphene oxide conjugated with the VEGF-165 gene) in the acute MI rat model demonstrates remarkable decrease in infarct size and capillary density improvement [67]. Cardiac tissue regeneration is an approaching therapeutic technique for the handling of CVDs. In addition, laponite NPs loaded with extracellular matrix hydrogel improved the phenotypic expansion of cardiac proteins along with cardiac cell compatibility [68].

2.3. Organic–Inorganic NPs

As of now, significance in the applications of numerous organic–inorganic hybrid NPs has risen excessively. Hybrid NPs combine the characteristics of inorganic and organic building units and form NPs with enhanced physical and chemical attributes, for example, good particle diameter and surface charge [45]. Moreover, Zhu et al. [69] fabricated inorganic–organic hybrid silica NPs for gene transfection delivery in bone marrow-derived mesenchymal stem cells and consequent in vivo cardiac reconstruction.
Figure 2. (A) Outline of the forms of drug delivery systems to treat heart disorders. Figure adapted with permission from Skourtis et al, 2020 [70]. Different NPs generally utilized in CVD therapeutic studies such as organic NPs (B,C), inorganic (DH), and hybrid NPs. Panel reproduced with permission from Fan er al., 2020 [71].
Figure 2. (A) Outline of the forms of drug delivery systems to treat heart disorders. Figure adapted with permission from Skourtis et al, 2020 [70]. Different NPs generally utilized in CVD therapeutic studies such as organic NPs (B,C), inorganic (DH), and hybrid NPs. Panel reproduced with permission from Fan er al., 2020 [71].
Ijms 23 12397 g002

3. Biological Markers in CVDs for Targeted Nanomedicine

The biological marker, also known as a biomarker, represents measurable and quantifiable biological parameters such as a specific enzyme concentration, phenotype distribution, and the expression of certain biological substances [72]. In particular, diagnosis through the discovery of the biomarkers of various CVDs contributes to early detections, risk predictions, novel clinical approaches, and finally decreases the mortality of CVD patients. The expression of these biomarkers is caused by pathophysiological changes that occur as the cardiovascular disease progresses, and these markers are involved in various processes such as inflammation, platelet activation, plaque instability, and systemic stress [73]. There are several biomarkers for the targeted nanomedicine treatment of vascular diseases.
Cholesterol is transported throughout the body by lipoproteins. Lipoproteins can be distinguished from one another most simply by their density, which is directly correlated with the amount of either apolipoprotein A-I (ApoAI) or apolipoprotein B (ApoB). A higher risk of coronary heart disease is linked to high cholesterol levels in lipoproteins containing ApoB, specifically in VLDL (very low-density lipoprotein) and even more so in LDL (low-density lipoprotein) [74]. In numerous types of research assessing the concentration of ApoA-I, it has been demonstrated that ApoA-I-containing lipoproteins (usually HDL, high-density lipoprotein) are negatively linked with CHD [75]. Due to HDL cholesterol’s role in reverse cholesterol transport, which is the process by which peripherally deposited cholesterol is transported back to the liver for elimination, there is evidence that higher levels of HDL cholesterol may lessen the risk of cardiovascular disease [76,77]. In particular, HDLs that are natural targeted NPs of vascular diseases (Figure 3) and transport cholesterol in systemic circulation have been used as targeted nanomedicine and imaging agents [78,79,80,81]. These HDL-based NPs can be used as not only targeted delivery platforms to deliver therapeutic cargos and imaging agents of biomimetic nature, but also anti-inflammatory molecules given their characteristics for CVDs.
Endothelial cell (EC) dysfunction and activation after the penetration of LDL cholesterol induce the overexpression of adhesion molecules including intracellular cell adhesion molecule-1 (ICAM-1), vascular adhesion molecule-1 (VCAM-1), and endothelial leukocyte adhesion molecule-1 (ELAM-1, or selectin). The elevated level of circulating adhesion molecules predicts atherosclerosis and cardiovascular events and it is linked to cardiovascular risk factors [82,83,84]. These inflammation markers are another key biomarker for vascular disease for targeted nanomedicine. VCAM-1, an immunoglobulin superfamily glycoprotein, is expressed on the surface of the activated endothelium in the early stage of atherosclerosis [85,86]. In addition, atherosclerotic lesion development decreased in genetically modified mice with diminished VCAM-1 function [87,88]. Interestingly, MRI and optical imaging identified a linear peptide affinity ligand, VHPKQHR, a known ligand for VCAM-1-decorated NPs colocalized to VCAM-1-expressing ECs compared to non-targeting NPs [89]. This result denotes that VHPKQHR peptide can be a targeting molecule to VCAM-1-overexpressing ECs, which is the early biomarker for atherosclerosis.

4. Strategies for Targeted Nanomedicine for CVDs

A profound understanding is required of the mechanisms associated with disease progression and therapeutic targets and the drugs required to conquer physiological barriers, especially circulation to different organs, and then from tissue to cells, to reach therapeutic targets. A nanocarrier must overcome these physiological barriers for successive targeting in nanomedicine. Commonly, biological areas, for instance, the vascular endothelium or the passageway of NPs via cellular tissues, serve as barriers for the nano-sized materials. The smaller size of NPs permits them to go through the cellular membrane or other barriers, e.g., the blood–brain barrier, for the targeted drug delivery. Due to the steep rise in nanotechnology usage, the use of targeted delivery strategies has dramatically increased for treatment options against various diseases over the past few decades. There are two different targeting strategies that use NPs: passive targeting and active targeting. Passive targeting can be achieved by the enhanced permeability and retention (EPR) effect in cancer treatment [90,91]. The EPR effect represents a general pathophysiological mechanism, whereby non-targeted biological molecules or nano-sized particles can steadily accumulate in the tumor vascularized area due to increased vascular permeability, and thus achieve the passively targeted accumulation of anticancer compounds into inflamed regions and tumor tissues. Passive targeting allows for the efficient localization of NPs within lesion areas, whereas active targeting permits the active uptake of NPs by conjugating targeting molecules to nanoparticles. Either passive or active targeting methods are the best potent solution to increase NP accumulation in the desired area of the lesion in CVDs (Figure 4).

4.1. Passive Targeting

Atherosclerosis is a chronic inflammatory disease that starts with EC activation [92,93,94,95,96]. ECs in the arterial wall activated by chemical, mechanical, and immunological interactions increase the expression of adhesion molecules such as ICAM-1, VCAM-1, E-selectin, and P-selectin (Figure 4A) [97,98,99]. These molecules lead to monocyte adhesion and the differentiation of intimal macrophages. Activated ECs and increased adhesion molecules on ECs allow the preferential penetration of NPs into ECs via the passive targeting method. The critical physiological change of the inflamed region induced by activated ECs for passively targeted NP accumulation is increased vascular permeability [100]. Normal endothelial cells act as endothelial barriers, which can regulate endothelial junctions in the control of homeostasis. In the chronic inflammation of ECs, the normally tight junctions undergo a phenotype change into a leakier status. This phenotype change increases the passive accumulation of the macromolecules with a large molecular weight as well as the NPs through the EPR effect. Although this effect has not received much attention for diverse vascular diseases, a recent study revealed that cyclodextrin NPs have an increased accumulation in the atherosclerotic plaque lesions compared to non-targeting cyclodextrin molecules [101]. The NPs of cyclodextrin polymers of ~10 nm in diameter accumulated at the atherosclerotic plaque site due to their long circulation time and by evading the renal clearance system, whereas cyclodextrin solution was rapidly eliminated from the bloodstream. These findings support the conclusion that NP formations can passively target the lesion area due to the phenotype change of leaky junctions due to activated ECs and increased circulation time.

4.2. Active Targeting

As shown in Figure 4B, the active targeting strategy includes NP uptake by lesion area cells via NP receptor interactions. For the active targeting method, NPs need to be decorated with targeting molecules such as VCAM-1-targeting peptide or biomimetic surface proteins. In addition, recent research revealed that cyclodextrin can help to regress the atherosclerosis via macrophage reprogramming [102] and it can dissolve the cholesterol crystals in the plaque region [103]. These results indicate that cyclodextrin can be a promising targeted therapeutic approach against the development of atherosclerosis.
VCAM-1 is an immunoglobulin superfamily glycoprotein that is expressed on activated endothelial cells, macrophages, and smooth muscle cells (SMCs) and plays a role in the inflammation and development of atherosclerotic plaques [88,97,104]. Since the VHPK peptide (VHPKQHR) has high specificity and affinity to VCAM-1 on the endothelium in atherosclerotic plaque, the VCAM-1-targeting peptide, VHPK peptide, has been widely used for imaging and targeting the atherosclerosis lesion [89,105]. Kheirolomoom and his colleagues developed a VCAM-1-targeting peptide containing lipid NPs to target the atherosclerotic plaque on endothelial cells [106]. The VHPK peptide-bearing lipid NPs containing miR-712 as a therapeutic cargo had a particle size of 167 ± 40 nm and a round shape of single or bilayer membrane. TNFα (3 ng/mL)-pretreated mouse endothelial cells induced an increased VCAM-1 expression on the endothelial cell surface and the VHPK NPs was abundantly observed in the cytosol of the activated cells. These VHPK peptide-conjugated lipid NPs delivered anti-miR-712, which silences miR-712 levels known as pro-atherogenic mechanosensitive miRNAs to targeted inflamed endothelium in a mouse model, resulting the downregulation of plaque development. In addition, VHPKQHR peptide-incorporated NPs have been used for the non-invasive diagnosis of atherosclerosis via magnetic resonance imaging detection [107,108]. FITC (fluorescein isothiocyanate)-labeled VHPKQHR peptide-loaded Fe3O4@SiO2 NPs (FITC-VHP-Fe3O4@SiO2) were constructed for the MRI detection of atherosclerotic regions as an efficient targeted contrast agent. The NPs had good dispersion properties and a positively charged particle surface. The FITC-VHP-Fe3O4@SiO2 has low cytotoxicity against Raw264.7 (immune cells) and MAECs (mouse aortic endothelial cells, non-immune cells) as well as good blood compatibility. The NPs showed a clear increase of T2 image signal intensity on the atherosclerotic plaque region and the significant accumulation of the FITC-VHP-Fe3O4@SiO2 NPs in the dissected aortic vessel samples. These results suggested that VHPKQHR peptide conjugation can be an ideal option to target atherosclerotic plaque. VHPK-conjugated poly(β-amino ester) (pBAE) NPs were used to deliver anti-miR-712 to inflamed ECs. These NPs showed an increased therapeutic effect in an atherosclerotic plaque-developed mouse model, which was designed to induce atherosclerosis in the carotid via partial carotid ligation [109,110]. Overall, the conjugating of VCAM-1-targeting peptide (VHPK peptide) has been investigated as an efficient and promising targeting strategy against inflamed ECs, which is the region from which CVD originates.
Another active targeting method towards atherosclerosis plaque is using biomimetic particles. Rapamycin-loaded NPs, which have the leukocyte surface membrane decoration, leukosome, were developed and evaluated to reduce vascular inflammation [111]. Rapamycin is a promising anti-atherosclerotic agent and is an inhibitor of the mammalian target of the rapamycin (mTOR) pathway [112,113]. After the injection of rapamycin-loaded leukosome NPs, decreased proliferating macrophages in the aorta were reported as well as decreased MCP (monocyte chemoattractant protein)-1, IL (interleukin)-1b, and MMP (matrix metalloproteinases) activity in the aorta. Another study using leukocyte membrane-decorated NPs revealed that biomimetic NPs can target the atherosclerosis [114]. Leukocyte membrane fragments were used to prepare the magnetic nanoclusters and the anti-inflammatory drug simvastatin was embedded to the nanoparticle. These NPs showed excellent anti-atherosclerotic effects by decreasing the inflammation level and oxidative stress, and promoting cholesterol efflux.
Platelet membrane is also used to prepare biomimetic liposomes (platelet-mimetic hybrid liposomes) to make a direct interaction between the NPs and activated ECs [115]. During the progression of atherosclerosis, platelets play a critical role in atherosclerosis initiation to adhere to activated ECs. Platelet-mimetic hybrid liposomes showed a significant increase of plaque accumulation and plaque penetration. These results demonstrated that leukocyte and platelet membrane fragments can be a promising targeting strategy against activated ECs, which is the starting phenotype of various chronic CVDs.
As mentioned earlier, cyclodextrin-based NPs can be a promising targeting molecule for targeted CVD treatment. Cyclic oligosaccharide 2-hydroxypropyl-β-cyclodextrin [116] induces increased cholesterol crystal solubility, preventing atherosclerosis [102]. The cyclodextrin treatment against a murine atherosclerosis model reduced the plaque size of atherosclerosis and promoted plaque regression with a continuous high-cholesterol diet. In addition, a cyclodextrin polymer represented NP formation, which could accumulate at the plaque site [101], and due to the hydrophobic cavity of cyclodextrin, it was used as a hydrophobic drug carrier [103]. Simvastatin, which is widely used as a cholesterol-lowering drug to reduce the risk of CVDs, was loaded into the hydrophobic cavity of cyclodextrin. After lipid coating and homogenization, cargo-switching NPs were prepared, and these NPs showed a cargo-switching ability from loaded statins to cholesterol crystals at the atherosclerotic plaque site. This platform increased the targeted therapeutic effect at the plaque site due to the selective statin release mechanism, and these findings demonstrated that cyclodextrin can be a promising atherosclerotic plaque-targeting material.

5. Challenges and Future Perspectives

Although there are a number of promising results from nanotechnology-based CVD targeted therapy, critical roadblocks still remain in the effective treatment of CVDs, including low efficacy, decreased circulation time, and off-target effects from the delivered NPs. In this review, we have described the current research of targeted nanomedicine for CVD treatment. In the passive targeting method, the nano-sized particles demonstrate strong advantages compared to water-soluble molecules. In addition, more options were discussed for the active targeting method, including VCAM-1-targeting peptides, biomimetic membrane fragment hybridized NPs and liposomes, and cyclodextrin NPs. Despite the development of these targeting strategies against vascular disease, there are still challenging steps for future directions. First, a novel biomarker for vascular disease should be investigated. During the chronic development of vascular disease, extensive genotype and phenotype changes were observed over a long period of time. Therefore, the discovery of new biomarkers may provide an opportunity to find various target substances in targeted nanomedicine-based vascular disease therapy. Second, it is necessary to develop a range of peptide sequences that can target various biomarkers. Currently, the VCAM-1-targeting peptide, VHPKQHR, is the most successful targeting peptide against the early stages of atherosclerosis plaque development. Numerous studies using VCAM-1-targeting peptides showed excellent results in targeting the plaque regions, from non-invasive diagnosis imaging platforms to targeted therapeutic cargo delivery. Likewise, novel peptide sequences to target various biomarkers from the early stage to the severe status of pathology have to be investigated. Third, a new targeting strategy using various types of NPs should be established. As mentioned, current conventional medication methods have their own limitations such as low water solubility, potential drug resistance, and poor biological efficiency; therefore, the aggressive usage of nanotechnology is needed for vascular disease. Moreover, when combining NP strategies, various administrative routes should be devised. The majority of NP treatments were administered via intravenous injection. Nobody wants to carry out intravenous injections for lowering the risk of CVD. In particular, when it comes to oral delivery, NP oral delivery must be improved in terms of their bioavailability during the digestive process. Lastly, personalized medicine techniques must be improved. Due to the chronic inflammation characteristics of various CVDs, the vascular disease treatment usually takes a long time, and it must be safe and stable enough to be used repeatedly. In the near future, improvements in nanomedicine, personalized therapy, and material science are projected to deliver significant functional improvements and EC restoration in patients with vascular diseases. Exploring the relationship among patient systems, disease heterogeneity, and nanomedicine could help conquer the biological obstacles of nanotheranostics. Prior to conducting clinical phase trials, nanoformulations should be intensively certified in preclinical in vivo model systems. Various natural products derived from active drugs and nanodrugs with cardioprotective actions are awaiting future investigation and clinical translations.

Author Contributions

Conceptualization, K.-A.C. and K.R.; writing—original draft preparation, K.-A.C., J.H.K., N.K. and K.R.; writing—review and editing, N.K., K.R., K.-A.C. and J.H.K.; funding acquisition, K.-A.C., N.K. and J.H.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Research Foundation of Korea (NRF-2021R1C1C1013875), the University of Suwon, Research Fund 001072 (JHK), and Kangnam University Research Fund (KAC). Also, this research was supported by Small and Medium Business Technology Development Fund 001329 (JHK).

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. Bonetti, P.O.; Lerman, L.O.; Lerman, A. Endothelial dysfunction: A marker of atherosclerotic risk. Arterioscler. Thromb. Vasc. Biol. 2003, 23, 168–175. [Google Scholar] [CrossRef] [PubMed]
  2. Malakar, A.K.; Choudhury, D.; Halder, B.; Paul, P.; Uddin, A.; Chakraborty, S. A review on coronary artery disease, its risk factors, and therapeutics. J. Cell. Physiol. 2019, 234, 16812–16823. [Google Scholar] [CrossRef] [PubMed]
  3. Geovanini, G.R.; Libby, P. Atherosclerosis and inflammation: Overview and updates. Clin. Sci. 2018, 132, 1243–1252. [Google Scholar] [CrossRef] [PubMed]
  4. Cho, J.G.; Lee, A.; Chang, W.; Lee, M.-S.; Kim, J. Endothelial to Mesenchymal Transition Represents a Key Link in the Interaction between Inflammation and Endothelial Dysfunction. Front. Immunol. 2018, 9, 294. [Google Scholar] [CrossRef] [Green Version]
  5. Libby, P. Inflammation in atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2012, 32, 2045–2051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Libby, P. Vascular biology of atherosclerosis: Overview and state of the art. Am. J. Cardiol. 2003, 91, 3–6. [Google Scholar] [CrossRef]
  7. Lusis, A.J. Atherosclerosis. Nature 2000, 407, 233–241. [Google Scholar] [CrossRef] [PubMed]
  8. Villa-Roel, N.; Ryu, K.; Jo, H. Role of Biomechanical Stress and Mechanosensitive miRNAs in Calcific Aortic Valve Disease. In Cardiovascular Calcification and Bone Mineralization; Aikawa, E., Hutcheson, J.D., Eds.; Springer International Publishing: Cham, Swizerland, 2020; pp. 117–135. [Google Scholar] [CrossRef]
  9. Sontheimer, D.L. Peripheral vascular disease: Diagnosis and treatment. Am. Fam. Physician 2006, 73, 1971–1976. [Google Scholar] [PubMed]
  10. Ghantous, C.M.; Kamareddine, L.; Farhat, R.; Zouein, F.A.; Mondello, S.; Kobeissy, F.; Zeidan, A. Advances in Cardiovascular Biomarker Discovery. Biomedicines 2020, 8, 552. [Google Scholar] [CrossRef]
  11. Stehlik, J.; Kobashigawa, J.; Hunt, S.A.; Reichenspurner, H.; Kirklin, J.K. Honoring 50 Years of Clinical Heart Transplantation in Circulation. Circulation 2018, 137, 71–87. [Google Scholar] [CrossRef]
  12. Zelniker, T.A.; Braunwald, E. Clinical Benefit of Cardiorenal Effects of Sodium-Glucose Cotransporter 2 Inhibitors. J. Am. Coll. Cardiol. 2020, 75, 435–447. [Google Scholar] [CrossRef] [PubMed]
  13. Leopold, J.A.; Loscalzo, J. Emerging Role of Precision Medicine in Cardiovascular Disease. Circ. Res. 2018, 122, 1302–1315. [Google Scholar] [CrossRef]
  14. Ariel, H.; Cooke, J.P. Cardiovascular Risk of Proton Pump Inhibitors. Methodist DeBakey Cardiovasc. J. 2019, 15, 214–219. [Google Scholar] [CrossRef]
  15. Rossello, X.; Pocock, S.J.; Julian, D.G. Long-Term Use of Cardiovascular Drugs: Challenges for Research and for Patient Care. J. Am. Coll. Cardiol. 2015, 66, 1273–1285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Pala, R.; Anju, V.T.; Dyavaiah, M.; Busi, S.; Nauli, S.M. Nanoparticle-Mediated Drug Delivery for the Treatment of Cardiovascular Diseases. Int. J. Nanomed. 2020, 15, 3741–3769. [Google Scholar] [CrossRef]
  17. Chen, J.; Zhang, X.; Millican, R.; Sherwood, J.; Martin, S.; Jo, H.; Yoon, Y.-s.; Brott, B.C.; Jun, H.-W. Recent advances in nanomaterials for therapy and diagnosis for atherosclerosis. Adv. Drug Deliv. Rev. 2021, 170, 142–199. [Google Scholar] [CrossRef] [PubMed]
  18. Germain, M.; Caputo, F.; Metcalfe, S.; Tosi, G.; Spring, K.; Åslund, A.K.O.; Pottier, A.; Schiffelers, R.; Ceccaldi, A.; Schmid, R. Delivering the power of nanomedicine to patients today. J. Control. Release 2020, 326, 164–171. [Google Scholar] [CrossRef]
  19. Ryu, K.; Park, J.; Kim, T.-i. Effect of pH-Responsive Charge-Conversional Polymer Coating to Cationic Reduced Graphene Oxide Nanostructures for Tumor Microenvironment-Targeted Drug Delivery Systems. Nanomaterials 2019, 9, 1289. [Google Scholar] [CrossRef] [Green Version]
  20. Hu, Q.; Fang, Z.; Ge, J.; Li, H. Nanotechnology for cardiovascular diseases. Innovation 2022, 3, 100214. [Google Scholar] [CrossRef]
  21. Kim, B.Y.S.; Rutka, J.T.; Chan, W.C.W. Nanomedicine. N. Engl. J. Med. 2010, 363, 2434–2443. [Google Scholar] [CrossRef] [Green Version]
  22. Irvine, D.J.; Dane, E.L. Enhancing cancer immunotherapy with nanomedicine. Nat. Rev. Immunol. 2020, 20, 321–334. [Google Scholar] [CrossRef] [PubMed]
  23. Ryu, K.; Lee, M.K.; Park, J.; Kim, T.-i. pH-Responsive Charge-Conversional Poly (ethylene imine)–Poly (l-lysine)–Poly (l-glutamic acid) with Self-Assembly and Endosome Buffering Ability for Gene Delivery Systems. ACS Appl. Bio Mater. 2018, 1, 1496–1504. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, J.; Sharma, R.; Ryu, K.; Shen, P.; Salaita, K.; Jo, H. Conditional Antisense Oligonucleotides Triggered by miRNA. ACS Chem. Biol. 2021, 16, 2255–2267. [Google Scholar] [CrossRef] [PubMed]
  25. Matoba, T.; Egashira, K. Nanoparticle-Mediated Drug Delivery System for Cardiovascular Disease. Int. Heart J. 2014, 55, 281–286. [Google Scholar] [CrossRef] [Green Version]
  26. Sun, D.; Chen, J.; Wang, Y.; Ji, H.; Peng, R.; Jin, L.; Wu, W. Advances in refunctionalization of erythrocyte-based nanomedicine for enhancing cancer-targeted drug delivery. Theranostics 2019, 9, 6885–6900. [Google Scholar] [CrossRef] [PubMed]
  27. Pearce, A.K.; O’Reilly, R.K. Insights into Active Targeting of Nanoparticles in Drug Delivery: Advances in Clinical Studies and Design Considerations for Cancer Nanomedicine. Bioconjugate Chem. 2019, 30, 2300–2311. [Google Scholar] [CrossRef]
  28. Lombardo, D.; Kiselev, M.A.; Caccamo, M.T. Smart Nanoparticles for Drug Delivery Application: Development of Versatile Nanocarrier Platforms in Biotechnology and Nanomedicine. J. Nanomater. 2019, 2019, 3702518. [Google Scholar] [CrossRef]
  29. Gao, D.; Guo, X.; Zhang, X.; Chen, S.; Wang, Y.; Chen, T.; Huang, G.; Gao, Y.; Tian, Z.; Yang, Z. Multifunctional phototheranostic nanomedicine for cancer imaging and treatment. Mater. Today Bio 2020, 5, 100035. [Google Scholar] [CrossRef]
  30. Man, F.; Lammers, T.; de Rosales, R.T.M. Imaging Nanomedicine-Based Drug Delivery: A Review of Clinical Studies. Mol. Imaging Biol. 2018, 20, 683–695. [Google Scholar] [CrossRef]
  31. Li, P.; Wang, D.; Hu, J.; Yang, X. The role of imaging in targeted delivery of nanomedicine for cancer therapy. Adv. Drug Deliv. Rev. 2022, 189, 114447. [Google Scholar] [CrossRef]
  32. Chen, W.; Schilperoort, M.; Cao, Y.; Shi, J.; Tabas, I.; Tao, W. Macrophage-targeted nanomedicine for the diagnosis and treatment of atherosclerosis. Nat. Rev. Cardiol. 2022, 19, 228–249. [Google Scholar] [CrossRef]
  33. Wu, P.-H.; Opadele, A.E.; Onodera, Y.; Nam, J.-M. Targeting Integrins in Cancer Nanomedicine: Applications in Cancer Diagnosis and Therapy. Cancers 2019, 11, 1783. [Google Scholar] [CrossRef] [Green Version]
  34. Cano, A.; Turowski, P.; Ettcheto, M.; Duskey, J.T.; Tosi, G.; Sánchez-López, E.; García, M.L.; Camins, A.; Souto, E.B.; Ruiz, A.; et al. Nanomedicine-based technologies and novel biomarkers for the diagnosis and treatment of Alzheimer’s disease: From current to future challenges. J. Nanobiotechnol. 2021, 19, 122. [Google Scholar] [CrossRef] [PubMed]
  35. Thi, T.T.H.; Suys, E.J.A.; Lee, J.S.; Nguyen, D.H.; Park, K.D.; Truong, N.P. Lipid-Based Nanoparticles in the Clinic and Clinical Trials: From Cancer Nanomedicine to COVID-19 Vaccines. Vaccines 2021, 9, 359. [Google Scholar] [CrossRef]
  36. Patra, S.; Singh, M.; Wasnik, K.; Pareek, D.; Gupta, P.S.; Mukherjee, S.; Paik, P. Polymeric Nanoparticle Based Diagnosis and Nanomedicine for Treatment and Development of Vaccines for Cerebral Malaria: A Review on Recent Advancement. ACS Appl. Bio Mater. 2021, 4, 7342–7365. [Google Scholar] [CrossRef] [PubMed]
  37. Zhou, J.; Zhang, Z.; Joseph, J.; Zhang, X.; Ferdows, B.E.; Patel, D.N.; Chen, W.; Banfi, G.; Molinaro, R.; Cosco, D.; et al. Biomaterials and nanomedicine for bone regeneration: Progress and future prospects. Exploration 2021, 1, 20210011. [Google Scholar] [CrossRef]
  38. Garbayo, E.; Pascual-Gil, S.; Rodríguez-Nogales, C.; Saludas, L.; Estella-Hermoso de Mendoza, A.; Blanco-Prieto, M.J. Nanomedicine and drug delivery systems in cancer and regenerative medicine. WIREs Nanomed. Nanobiotechnol. 2020, 12, e1637. [Google Scholar] [CrossRef]
  39. Donahue, N.D.; Acar, H.; Wilhelm, S. Concepts of nanoparticle cellular uptake, intracellular trafficking, and kinetics in nanomedicine. Adv. Drug Deliv. Rev. 2019, 143, 68–96. [Google Scholar] [CrossRef]
  40. Rudramurthy, G.; Swamy, M.; Sinniah, U.; Ghasemzadeh, A. Nanoparticles: Alternatives Against Drug-Resistant Pathogenic Microbes. Molecules 2016, 21, 836. [Google Scholar] [CrossRef] [PubMed]
  41. Redhead, H.M.; Davis, S.S.; Illum, L. Drug delivery in poly(lactide-co-glycolide) nanoparticles surface modified with poloxamer 407 and poloxamine 908: In vitro characterisation and in vivo evaluation. J. Control. Release 2001, 70, 353–363. [Google Scholar] [CrossRef]
  42. Truong, N.P.; Whittaker, M.R.; Mak, C.W.; Davis, T.P. The importance of nanoparticle shape in cancer drug delivery. Expert Opin. Drug Deliv. 2014, 12, 129–142. [Google Scholar] [CrossRef]
  43. Couvreur, P.; Barratt, G.; Fattal, E.; Vauthier, C. Nanocapsule Technology: A Review. Crit. Rev. Ther. Drug Carr. Syst. 2002, 19, 99–134. [Google Scholar] [CrossRef]
  44. Hui, Y.; Yi, X.; Hou, F.; Wibowo, D.; Zhang, F.; Zhao, D.; Gao, H.; Zhao, C.-X. Role of Nanoparticle Mechanical Properties in Cancer Drug Delivery. ACS Nano 2019, 13, 7410–7424. [Google Scholar] [CrossRef]
  45. Haque, S.T.; Chowdhury, E.H. Recent Progress in Delivery of Therapeutic and Imaging Agents Utilizing Organic-Inorganic Hybrid Nanoparticles. Curr. Drug Deliv. 2018, 15, 485–496. [Google Scholar] [CrossRef]
  46. Qi, J.; Zhuang, J.; Lu, Y.; Dong, X.; Zhao, W.; Wu, W. In vivo fate of lipid-based nanoparticles. Drug Discov. Today 2017, 22, 166–172. [Google Scholar] [CrossRef]
  47. Paulis, L.E.; Geelen, T.; Kuhlmann, M.T.; Coolen, B.F.; Schäfers, M.; Nicolay, K.; Strijkers, G.J. Distribution of lipid-based nanoparticles to infarcted myocardium with potential application for MRI-monitored drug delivery. J. Control. Release 2012, 162, 276–285. [Google Scholar] [CrossRef]
  48. Vinhas, R.; Mendes, R.; Fernandes, A.R.; Baptista, P.V. Nanoparticles—Emerging Potential for Managing Leukemia and Lymphoma. Front. Bioeng. Biotechnol. 2017, 5, 79. [Google Scholar] [CrossRef] [Green Version]
  49. Katsuki, S.; Matoba, T.; Koga, J.-i.; Nakano, K.; Egashira, K. Anti-inflammatory Nanomedicine for Cardiovascular Disease. Front. Cardiovasc. Med. 2017, 4, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Pascual-Gil, S.; Simón-Yarza, T.; Garbayo, E.; Prósper, F.; Blanco-Prieto, M.J. Cytokine-loaded PLGA and PEG-PLGA microparticles showed similar heart regeneration in a rat myocardial infarction model. Int. J. Pharm. 2017, 523, 531–533. [Google Scholar] [CrossRef]
  51. Oduk, Y.; Zhu, W.; Kannappan, R.; Zhao, M.; Borovjagin, A.V.; Oparil, S.; Zhang, J. VEGF nanoparticles repair the heart after myocardial infarction. Am. J. Physiol.-Heart Circ. Physiol. 2018, 314, H278–H284. [Google Scholar] [CrossRef] [PubMed]
  52. Danenberg, H.D.; Golomb, G.; Groothuis, A.; Gao, J.; Epstein, H.; Swaminathan, R.V.; Seifert, P.; Edelman, E.R. Liposomal Alendronate Inhibits Systemic Innate Immunity and Reduces In-Stent Neointimal Hyperplasia in Rabbits. Circulation 2003, 108, 2798–2804. [Google Scholar] [CrossRef] [Green Version]
  53. Thomas, T.; Joice, M.; Sumit, M.; Silpe, J.; Kotlyar, A.; Bharathi, S.; Kukowska-Latallo, J.; Baker, J.; Choi, S. Design and In vitro Validation of Multivalent Dendrimer Methotrexates as a Folate-targeting Anticancer Therapeutic. Curr. Pharm. Des. 2013, 19, 6594–6605. [Google Scholar] [CrossRef] [Green Version]
  54. Singh, B.; Garg, T.; Goyal, A.K.; Rath, G. Recent advancements in the cardiovascular drug carriers. Artif. Cells Nanomed. Biotechnol. 2014, 44, 216–225. [Google Scholar] [CrossRef]
  55. Xue, X.; Shi, X.; Dong, H.; You, S.; Cao, H.; Wang, K.; Wen, Y.; Shi, D.; He, B.; Li, Y. Delivery of microRNA-1 inhibitor by dendrimer-based nanovector: An early targeting therapy for myocardial infarction in mice. Nanomed. Nanotechnol. Biol. Med. 2018, 14, 619–631. [Google Scholar] [CrossRef]
  56. Xu, Q.-H.; Guan, P.; Zhang, T.; Lu, C.; Li, G.; Liu, J.-X. Silver nanoparticles impair zebrafish skeletal and cardiac myofibrillogenesis and sarcomere formation. Aquat. Toxicol. 2018, 200, 102–113. [Google Scholar] [CrossRef]
  57. Zhang, W.; Zhang, Z.; Zhang, Y. The application of carbon nanotubes in target drug delivery systems for cancer therapies. Nanoscale Res. Lett. 2011, 6, 555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Gorain, B.; Choudhury, H.; Pandey, M.; Kesharwani, P.; Abeer, M.M.; Tekade, R.K.; Hussain, Z. Carbon nanotube scaffolds as emerging nanoplatform for myocardial tissue regeneration: A review of recent developments and therapeutic implications. Biomed. Pharmacother. 2018, 104, 496–508. [Google Scholar] [CrossRef]
  59. Ahadian, S.; Davenport Huyer, L.; Estili, M.; Yee, B.; Smith, N.; Xu, Z.; Sun, Y.; Radisic, M. Moldable elastomeric polyester-carbon nanotube scaffolds for cardiac tissue engineering. Acta Biomater. 2017, 52, 81–91. [Google Scholar] [CrossRef]
  60. Sun, H.; Tang, J.; Mou, Y.; Zhou, J.; Qu, L.; Duval, K.; Huang, Z.; Lin, N.; Dai, R.; Liang, C.; et al. Carbon nanotube-composite hydrogels promote intercalated disc assembly in engineered cardiac tissues through β1-integrin mediated FAK and RhoA pathway. Acta Biomater. 2017, 48, 88–99. [Google Scholar] [CrossRef]
  61. Martinelli, V.; Cellot, G.; Toma, F.M.; Long, C.S.; Caldwell, J.H.; Zentilin, L.; Giacca, M.; Turco, A.; Prato, M.; Ballerini, L.; et al. Carbon Nanotubes Instruct Physiological Growth and Functionally Mature Syncytia: Nongenetic Engineering of Cardiac Myocytes. ACS Nano 2013, 7, 5746–5756. [Google Scholar] [CrossRef] [PubMed]
  62. Shin, S.R.; Jung, S.M.; Zalabany, M.; Kim, K.; Zorlutuna, P.; Kim, S.b.; Nikkhah, M.; Khabiry, M.; Azize, M.; Kong, J.; et al. Carbon-Nanotube-Embedded Hydrogel Sheets for Engineering Cardiac Constructs and Bioactuators. ACS Nano 2013, 7, 2369–2380. [Google Scholar] [CrossRef] [Green Version]
  63. Sun, H.; Zhou, J.; Huang, Z.; Qu, L.; Lin, N.; Liang, C.; Dai, R.; Tang, L.; Tian, F. Carbon nanotube-incorporated collagen hydrogels improve cell alignment and the performance of cardiac constructs. Int. J. Nanomed. 2017, 12, 3109–3120. [Google Scholar] [CrossRef] [Green Version]
  64. Sperling, R.A.; Rivera Gil, P.; Zhang, F.; Zanella, M.; Parak, W.J. Biological applications of gold nanoparticles. Chem. Soc. Rev. 2008, 37, 1896–1908. [Google Scholar] [CrossRef]
  65. Spivak, M.Y.; Bubnov, R.V.; Yemets, I.M.; Lazarenko, L.M.; Tymoshok, N.O.; Ulberg, Z.R. Development and testing of gold nanoparticles for drug delivery and treatment of heart failure: A theranostic potential for PPP cardiology. EPMA J. 2013, 4, 20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Ahmed, S.M.; Abdelrahman, S.A.; Salama, A.E. Efficacy of gold nanoparticles against isoproterenol induced acute myocardial infarction in adult male albino rats. Ultrastruct. Pathol. 2017, 41, 168–185. [Google Scholar] [CrossRef]
  67. Paul, A.; Hasan, A.; Kindi, H.A.; Gaharwar, A.K.; Rao, V.T.S.; Nikkhah, M.; Shin, S.R.; Krafft, D.; Dokmeci, M.R.; Shum-Tim, D.; et al. Injectable Graphene Oxide/Hydrogel-Based Angiogenic Gene Delivery System for Vasculogenesis and Cardiac Repair. ACS Nano 2014, 8, 8050–8062. [Google Scholar] [CrossRef] [Green Version]
  68. Zhang, Y.; Fan, W.; Wang, K.; Wei, H.; Zhang, R.; Wu, Y. Novel preparation of Au nanoparticles loaded Laponite nanoparticles/ECM injectable hydrogel on cardiac differentiation of resident cardiac stem cells to cardiomyocytes. J. Photochem. Photobiol. B Biol. 2019, 192, 49–54. [Google Scholar] [CrossRef]
  69. Zhu, K.; Wu, M.; Lai, H.; Guo, C.; Li, J.; Wang, Y.; Chen, Y.; Wang, C.; Shi, J. Nanoparticle-enhanced generation of gene-transfected mesenchymal stem cells for in vivo cardiac repair. Biomaterials 2016, 74, 188–199. [Google Scholar] [CrossRef]
  70. Skourtis, D.; Stavroulaki, D.; Athanasiou, V.; Fragouli, P.G.; Iatrou, H. Nanostructured Polymeric, Liposomal and Other Materials to Control the Drug Delivery for Cardiovascular Diseases. Pharmaceutics 2020, 12, 1160. [Google Scholar] [CrossRef]
  71. Fan, C.; Joshi, J.; Li, F.; Xu, B.; Khan, M.; Yang, J.; Zhu, W. Nanoparticle-Mediated Drug Delivery for Treatment of Ischemic Heart Disease. Front. Bioeng. Biotechnol. 2020, 8, 687. [Google Scholar] [CrossRef]
  72. Vasan, R.S. Biomarkers of Cardiovascular Disease. Circulation 2006, 113, 2335–2362. [Google Scholar] [CrossRef]
  73. Lyngbakken, M.N.; Myhre, P.L.; Røsjø, H.; Omland, T. Novel biomarkers of cardiovascular disease: Applications in clinical practice. Crit. Rev. Clin. Lab. Sci. 2019, 56, 33–60. [Google Scholar] [CrossRef]
  74. McQueen, M.J.; Hawken, S.; Wang, X.; Ounpuu, S.; Sniderman, A.; Probstfield, J.; Steyn, K.; Sanderson, J.E.; Hasani, M.; Volkova, E.; et al. Lipids, lipoproteins, and apolipoproteins as risk markers of myocardial infarction in 52 countries (the INTERHEART study): A case-control study. Lancet 2008, 372, 224–233. [Google Scholar] [CrossRef]
  75. The Emerging Risk Factors Collaboration. Major Lipids, Apolipoproteins, and Risk of Vascular Disease. JAMA 2009, 302, 1993–2000. [Google Scholar] [CrossRef] [Green Version]
  76. Movva, R.; Rader, D.J. Laboratory Assessment of HDL Heterogeneity and Function. Clin Chem. 2008, 54, 788–800. [Google Scholar] [CrossRef] [Green Version]
  77. Rosenson, R.S.; Brewer, H.B.; Ansell, B.; Barter, P.; Chapman, M.J.; Heinecke, J.W.; Kontush, A.; Tall, A.R.; Webb, N.R. Translation of High-Density Lipoprotein Function Into Clinical Practice. Circulation 2013, 128, 1256–1267. [Google Scholar] [CrossRef] [Green Version]
  78. Rink, J.S.; Sun, W.; Misener, S.; Wang, J.-J.; Zhang, Z.J.; Kibbe, M.R.; Dravid, V.P.; Venkatraman, S.; Thaxton, C.S. Nitric Oxide-Delivering High-Density Lipoprotein-like Nanoparticles as a Biomimetic Nanotherapy for Vascular Diseases. ACS Appl. Mater. Interfaces 2018, 10, 6904–6916. [Google Scholar] [CrossRef]
  79. Foit, L.; Thaxton, C.S. Synthetic high-density lipoprotein-like nanoparticles potently inhibit cell signaling and production of inflammatory mediators induced by lipopolysaccharide binding Toll-like receptor 4. Biomaterials 2016, 100, 67–75. [Google Scholar] [CrossRef]
  80. Chen, W.; Cormode, D.P.; Fayad, Z.A.; Mulder, W.J.M. Nanoparticles as magnetic resonance imaging contrast agents for vascular and cardiac diseases. WIREs Nanomed. Nanobiotechnol. 2011, 3, 146–161. [Google Scholar] [CrossRef]
  81. Leeper, N.J.; Park, S.-m.; Smith, B.R. High-Density Lipoprotein Nanoparticle Imaging in Atherosclerotic Vascular Disease. J. Am. Coll. Cardiol. Basic Trans Sci. 2017, 2, 98–100. [Google Scholar] [CrossRef]
  82. Ponthieux, A.; Herbeth, B.; Droesch, S.; Haddy, N.; Lambert, D.; Visvikis, S. Biological determinants of serum ICAM-1, E-selectin, P-selectin and L-selectin levels in healthy subjects: The Stanislas study. Atherosclerosis 2004, 172, 299–308. [Google Scholar] [CrossRef]
  83. Hwang, S.-J.; Ballantyne, C.M.; Sharrett, A.R.; Smith, L.C.; Davis, C.E.; Gotto, A.M.; Boerwinkle, E. Circulating Adhesion Molecules VCAM-1, ICAM-1, and E-selectin in Carotid Atherosclerosis and Incident Coronary Heart Disease Cases. Circulation 1997, 96, 4219–4225. [Google Scholar] [CrossRef]
  84. Caterina, R.D.; Basta, G.; Lazzerini, G.; Dell’Omo, G.; Petrucci, R.; Morale, M.; Carmassi, F.; Pedrinelli, R. Soluble Vascular Cell Adhesion Molecule-1 as a Biohumoral Correlate of Atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 1997, 17, 2646–2654. [Google Scholar] [CrossRef]
  85. Iiyama, K.; Hajra, L.; Iiyama, M.; Li, H.; DiChiara, M.; Medoff, B.D.; Cybulsky, M.I. Patterns of Vascular Cell Adhesion Molecule-1 and Intercellular Adhesion Molecule-1 Expression in Rabbit and Mouse Atherosclerotic Lesions and at Sites Predisposed to Lesion Formation. Circ. Res. 1999, 85, 199–207. [Google Scholar] [CrossRef]
  86. Li, H.; Cybulsky, M.I.; Gimbrone Jr, M.A.; Libby, P. Inducible expression of vascular cell adhesion molecule-1 by vascular smooth muscle cells in vitro and within rabbit atheroma. Am. J. Pathol. 1993, 143, 1551. [Google Scholar]
  87. Cybulsky, M.I.; Iiyama, K.; Li, H.; Zhu, S.; Chen, M.; Iiyama, M.; Davis, V.; Gutierrez-Ramos, J.-C.; Connelly, P.W.; Milstone, D.S. A major role for VCAM-1, but not ICAM-1, in early atherosclerosis. J. Clin. Investig. 2001, 107, 1255–1262. [Google Scholar] [CrossRef] [Green Version]
  88. Nakashima, Y.; Raines, E.W.; Plump, A.S.; Breslow, J.L.; Ross, R. Upregulation of VCAM-1 and ICAM-1 at atherosclerosis-prone sites on the endothelium in the ApoE-deficient mouse. Arterioscler. Thromb. Vasc. Biol. 1998, 18, 842–851. [Google Scholar] [CrossRef] [Green Version]
  89. Nahrendorf, M.; Jaffer, F.A.; Kelly, K.A.; Sosnovik, D.E.; Aikawa, E.; Libby, P.; Weissleder, R. Noninvasive vascular cell adhesion molecule-1 imaging identifies inflammatory activation of cells in atherosclerosis. Circulation 2006, 114, 1504–1511. [Google Scholar] [CrossRef] [Green Version]
  90. Attia, M.F.; Anton, N.; Wallyn, J.; Omran, Z.; Vandamme, T.F. An overview of active and passive targeting strategies to improve the nanocarriers efficiency to tumour sites. J. Pharm. Pharmacol. 2019, 71, 1185–1198. [Google Scholar] [CrossRef] [Green Version]
  91. Alavi, M.; Hamidi, M. Passive and active targeting in cancer therapy by liposomes and lipid nanoparticles. Drug Metab. Pers. Ther. 2019, 34, 20180032. [Google Scholar] [CrossRef]
  92. Villa-Roel, N.; Gu, L.; Fernandez Esmerats, J.; Kang, D.-W.; Kumar, S.; Jo, H. Hypoxia Inducible Factor-1α Inhibitor, PX-478, Reduces Atherosclerosis in vivo. Arterioscler. Thromb. Vasc. Biol. 2020, 40, A366. [Google Scholar]
  93. Kumar, S.; Kim, C.W.; Simmons, R.D.; Jo, H. Role of flow-sensitive microRNAs in endothelial dysfunction and atherosclerosis: Mechanosensitive athero-miRs. Arterioscler. Thromb. Vasc. Biol. 2014, 34, 2206–2216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Kumar, S.; Kim, C.W.; Son, D.J.; Ni, C.W.; Jo, H. Flow-dependent regulation of genome-wide mRNA and microRNA expression in endothelial cells in vivo. Sci. Data 2014, 1, 140039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Mudau, M.; Genis, A.; Lochner, A.; Strijdom, H. Endothelial dysfunction: The early predictor of atherosclerosis. Cardiovasc. J. Afr. 2012, 23, 222–231. [Google Scholar] [CrossRef]
  96. Chiu, J.-J.; Chien, S. Effects of disturbed flow on vascular endothelium: Pathophysiological basis and clinical perspectives. Physiol. Rev. 2011, 91, 327–387. [Google Scholar] [CrossRef] [Green Version]
  97. Cybulsky, M.I.; Gimbrone, M.A. Endothelial Expression of a Mononuclear Leukocyte Adhesion Molecule During Atherogenesis. Science 1991, 251, 788–791. [Google Scholar] [CrossRef] [Green Version]
  98. Dong, Z.M.; Chapman, S.M.; Brown, A.A.; Frenette, P.S.; Hynes, R.O.; Wagner, D.D. The combined role of P- and E-selectins in atherosclerosis. J. Clin. Investig. 1998, 102, 145–152. [Google Scholar] [CrossRef] [Green Version]
  99. Collins, R.G.; Velji, R.; Guevara, N.V.; Hicks, M.J.; Chan, L.; Beaudet, A.L. P-Selectin or Intercellular Adhesion Molecule (Icam)-1 Deficiency Substantially Protects against Atherosclerosis in Apolipoprotein E–Deficient Mice. J. Exp. Med. 2000, 191, 189–194. [Google Scholar] [CrossRef]
  100. Claesson-Welsh, L.; Dejana, E.; McDonald, D.M. Permeability of the Endothelial Barrier: Identifying and Reconciling Controversies. Trends Mol. Med. 2021, 27, 314–331. [Google Scholar] [CrossRef]
  101. Kim, H.; Han, J.; Park, J.-H. Cyclodextrin polymer improves atherosclerosis therapy and reduces ototoxicity. J. Control. Release 2020, 319, 77–86. [Google Scholar] [CrossRef]
  102. Zimmer, S.; Grebe, A.; Bakke, S.S.; Bode, N.; Halvorsen, B.; Ulas, T.; Skjelland, M.; De Nardo, D.; Labzin, L.I.; Kerksiek, A. Cyclodextrin promotes atherosclerosis regression via macrophage reprogramming. Sci. Transl. Med. 2016, 8, 333ra50. [Google Scholar] [CrossRef] [PubMed]
  103. Kim, H.; Kumar, S.; Kang, D.-W.; Jo, H.; Park, J.-H. Affinity-Driven Design of Cargo-Switching Nanoparticles to Leverage a Cholesterol-Rich Microenvironment for Atherosclerosis Therapy. ACS Nano 2020, 14, 6519–6531. [Google Scholar] [CrossRef] [PubMed]
  104. O’Brien, K.D.; Allen, M.D.; McDonald, T.O.; Chait, A.; Harlan, J.M.; Fishbein, D.; McCarty, J.; Ferguson, M.; Hudkins, K.; Benjamin, C.D.; et al. Vascular cell adhesion molecule-1 is expressed in human coronary atherosclerotic plaques. Implications for the mode of progression of advanced coronary atherosclerosis. J. Clin. Investig. 1993, 92, 945–951. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Distasio, N.; Salmon, H.; Dierick, F.; Ebrahimian, T.; Tabrizian, M.; Lehoux, S. VCAM-1-Targeted Gene Delivery Nanoparticles Localize to Inflamed Endothelial Cells and Atherosclerotic Plaques. Adv. Ther. 2021, 4, 2000196. [Google Scholar] [CrossRef]
  106. Kheirolomoom, A.; Kim, C.W.; Seo, J.W.; Kumar, S.; Son, D.J.; Gagnon, M.K.J.; Ingham, E.S.; Ferrara, K.W.; Jo, H. Multifunctional nanoparticles facilitate molecular targeting and miRNA delivery to inhibit atherosclerosis in ApoE−/− mice. ACS Nano 2015, 9, 8885–8897. [Google Scholar] [CrossRef]
  107. Xu, W.; Zhang, S.; Zhou, Q.; Chen, W. VHPKQHR peptide modified magnetic mesoporous nanoparticles for MRI detection of atherosclerosis lesions. Artif. Cells Nanomed. Biotechnol. 2019, 47, 2440–2448. [Google Scholar] [CrossRef] [Green Version]
  108. Yadav, M.; Bhayana, S.; Liu, J.; Lu, L.; Huang, J.; Ma, Y.; Qamri, Z.; Mo, X.; Jacob, D.S.; Parasa, S.T.; et al. Two-miRNA-based finger-stick assay for estimation of absorbed ionizing radiation dose. Sci. Transl. Med. 2020, 12, eaaw5831. [Google Scholar] [CrossRef]
  109. Nam, D.; Ni, C.-W.; Rezvan, A.; Suo, J.; Budzyn, K.; Llanos, A.; Harrison, D.; Giddens, D.; Jo, H. Partial carotid ligation is a model of acutely induced disturbed flow, leading to rapid endothelial dysfunction and atherosclerosis. Am. J. Physiol. Heart Circ. Physiol. 2009, 297, H1535–H1543. [Google Scholar] [CrossRef]
  110. Kumar, S.; Kang, D.W.; Rezvan, A.; Jo, H. Accelerated atherosclerosis development in C57Bl6 mice by overexpressing AAV-mediated PCSK9 and partial carotid ligation. Lab. Investig. 2017, 97, 935–945. [Google Scholar] [CrossRef] [Green Version]
  111. Boada, C.; Zinger, A.; Tsao, C.; Zhao, P.; Martinez, J.O.; Hartman, K.; Naoi, T.; Sukhoveshin, R.; Sushnitha, M.; Molinaro, R.; et al. Rapamycin-Loaded Biomimetic Nanoparticles Reverse Vascular Inflammation. Circ. Res. 2020, 126, 25–37. [Google Scholar] [CrossRef]
  112. Mita, M.M.; Mita, A.; Rowinsky, E.K. The molecular target of rapamycin (mTOR) as a therapeutic target against cancer. Cancer Biol. Ther. 2003, 2, 168–176. [Google Scholar] [CrossRef]
  113. Tsang, C.K.; Qi, H.; Liu, L.F.; Zheng, X.S. Targeting mammalian target of rapamycin (mTOR) for health and diseases. Drug Discov. Today 2007, 12, 112–124. [Google Scholar] [CrossRef]
  114. Wu, G.; Wei, W.; Zhang, J.; Nie, W.; Yuan, L.; Huang, Y.; Zuo, L.; Huang, L.; Xi, X.; Xie, H.Y. A self-driven bioinspired nanovehicle by leukocyte membrane-hitchhiking for early detection and treatment of atherosclerosis. Biomaterials 2020, 250, 119963. [Google Scholar] [CrossRef]
  115. Song, Y.; Zhang, N.; Li, Q.; Chen, J.; Wang, Q.; Yang, H.; Tan, H.; Gao, J.; Dong, Z.; Pang, Z.; et al. Biomimetic liposomes hybrid with platelet membranes for targeted therapy of atherosclerosis. Chem. Eng. J. 2021, 408, 127296. [Google Scholar] [CrossRef]
  116. Kaplan, R.N.; Riba, R.D.; Zacharoulis, S.; Bramley, A.H.; Vincent, L.; Costa, C.; MacDonald, D.D.; Jin, D.K.; Shido, K.; Kerns, S.A.; et al. VEGFR1-positive haematopoietic bone marrow progenitors initiate the pre-metastatic niche. Nature 2005, 438, 820–827. [Google Scholar] [CrossRef]
Figure 1. Feasible advances for the diagnosis and treatment of cardiovascular diseases (CVDs). Figure adapted with permission from Hu et al, 2022 [20].
Figure 1. Feasible advances for the diagnosis and treatment of cardiovascular diseases (CVDs). Figure adapted with permission from Hu et al, 2022 [20].
Ijms 23 12397 g001
Figure 3. HDL-like NPs as a biomimetic nanotherapy for CVDs. Reprinted with permission from [78] Copyright 2018, American Chemical Society.
Figure 3. HDL-like NPs as a biomimetic nanotherapy for CVDs. Reprinted with permission from [78] Copyright 2018, American Chemical Society.
Ijms 23 12397 g003
Figure 4. Targeted delivery of NPs. (A) Passive targeting is accomplished by increased nanoparticle adhesion to the adhesion molecules of activated endothelial cells, and (B) the active targeting method needs targeting molecule-decorated nanoparticle and biomimetic NPs.
Figure 4. Targeted delivery of NPs. (A) Passive targeting is accomplished by increased nanoparticle adhesion to the adhesion molecules of activated endothelial cells, and (B) the active targeting method needs targeting molecule-decorated nanoparticle and biomimetic NPs.
Ijms 23 12397 g004
Table 1. List of nanoplatforms for target delivery of various drugs to the sites of cardiovascular diseases (CVDs). Table reproduced from Hu et al., 2022 [20].
Table 1. List of nanoplatforms for target delivery of various drugs to the sites of cardiovascular diseases (CVDs). Table reproduced from Hu et al., 2022 [20].
Drug TypeLoaded DrugNanoplatformsDisordersMechanism of ActionSurface ModificationsModel of Use/AnimalAdministration Route
Statinsatorvastatin
atorvastatin
HA-ATV-NP
Oxi-COS/MM-AT-nps
atherosclerosis
atherosclerosis
suppression of inflammation
suppression of inflammation
hyaluronan
proteins derived from macrophages membrane
in vitro; in vivo, ApoE-/- mice
in vitro; in vivo, ApoE-/- mice
intravenous injection
intravenous injection
Rapamycinrapamycin
rapamycin
PFN1-CD-mnps
liposome
atherosclerosis
atherosclerosis
suppression of inflammation suppression of inflammationprofilin-1 antibody
membrane protein from leukocytes
in vitro; in vivo, ApoE-/- mice
in vitro; in vivo, ApoE-/- mice
intravenous injection
retro-orbital injection
Traditional Chinese medicineSal B, PNSRGD-S/P-lpnsAMI RGD peptide ligandin vivo, SD rats receiving experimental MIintravenous injection
Small molecule agonists/inhibitorsSMI 6877002rHDL NPsatherosclerosisinhibition of monocyte recruitment; suppression of plaque inflammationApoA-Iin vitro; in vivo, ApoE-/- mice, cynomolgus monkeysintravenous injection
Small molecule agonists/inhibitorsSNOSNO-HDL NPsatherosclerosis ApoA-Iin vitro; in vivo, ApoE-/- miceintravenous injection
siRNAsiCamk2gG0-C14 PLGA NPsatherosclerosispromotion of efferocytosisS2P peptide (CRTLTVRKC)in vitro; in vivo, Ldlr-/- miceintravenous injection
miRNAmiR-145PAMatherosclerosispromotion of the contractile VSMC phenotypeMCP1/CCL2in vitro; in vivo, ApoE-/- miceintravenous injection
miRNA switchesmiRNA switchesmRNA-p5RHH nanoparticlerestenosisspecific inhibition of the VSMCs and inflammatory cells in vitro; in vivo, C57BL6/J mice undergoing femoral artery wire injuryintravenous injection
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Choi, K.-A.; Kim, J.H.; Ryu, K.; Kaushik, N. Current Nanomedicine for Targeted Vascular Disease Treatment: Trends and Perspectives. Int. J. Mol. Sci. 2022, 23, 12397. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232012397

AMA Style

Choi K-A, Kim JH, Ryu K, Kaushik N. Current Nanomedicine for Targeted Vascular Disease Treatment: Trends and Perspectives. International Journal of Molecular Sciences. 2022; 23(20):12397. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232012397

Chicago/Turabian Style

Choi, Kyung-A, June Hyun Kim, Kitae Ryu, and Neha Kaushik. 2022. "Current Nanomedicine for Targeted Vascular Disease Treatment: Trends and Perspectives" International Journal of Molecular Sciences 23, no. 20: 12397. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232012397

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