Next Article in Journal
AS1411 Aptamer Linked to DNA Nanostructures Diverts Its Traffic Inside Cancer Cells and Improves Its Therapeutic Efficacy
Previous Article in Journal
A Strategy for Selective Deletion of Autoimmunity-Related T Cells by pMHC-Targeted Delivery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Functionalized Nanoparticles Targeting Tumor-Associated Macrophages as Cancer Therapy

by
Yuanyuan He
1,
Raimundo Fernandes de Araújo Júnior
1,2,3,4,
Luis J. Cruz
1,* and
Christina Eich
1,*
1
Translational Nanobiomaterials and Imaging (TNI) Group, Department of Radiology, Leiden University Medical Center, 2333 ZA Leiden, The Netherlands
2
Postgraduate Program in Health Science, Federal University of Rio Grande do Norte (UFRN), Natal 59064-720, Brazil
3
Cancer and Inflammation Research Laboratory (LAICI), Postgraduate Program in Functional and Structural Biology, Department of Morphology, Federal University of Rio Grande do Norte (UFRN), Natal 59064-720, Brazil
4
Percuros B.V., 2333 CL Leiden, The Netherlands
*
Authors to whom correspondence should be addressed.
Pharmaceutics 2021, 13(10), 1670; https://doi.org/10.3390/pharmaceutics13101670
Submission received: 1 September 2021 / Revised: 2 October 2021 / Accepted: 5 October 2021 / Published: 13 October 2021

Abstract

:
The tumor microenvironment (TME) plays a central role in regulating antitumor immune responses. As an important part of the TME, alternatively activated type 2 (M2) macrophages drive the development of primary and secondary tumors by promoting tumor cell proliferation, tumor angiogenesis, extracellular matrix remodeling and overall immunosuppression. Immunotherapy approaches targeting tumor-associated macrophages (TAMs) in order to reduce the immunosuppressive state in the TME have received great attention. Although these methods hold great potential for the treatment of several cancers, they also face some limitations, such as the fast degradation rate of drugs and drug-induced cytotoxicity of organs and tissues. Nanomedicine formulations that prevent TAM signaling and recruitment to the TME or deplete M2 TAMs to reduce tumor growth and metastasis represent encouraging novel strategies in cancer therapy. They allow the specific delivery of antitumor drugs to the tumor area, thereby reducing side effects associated with systemic application. In this review, we give an overview of TAM biology and the current state of nanomedicines that target M2 macrophages in the course of cancer immunotherapy, with a specific focus on nanoparticles (NPs). We summarize how different types of NPs target M2 TAMs, and how the physicochemical properties of NPs (size, shape, charge and targeting ligands) influence NP uptake by TAMs in vitro and in vivo in the TME. Furthermore, we provide a comparative analysis of passive and active NP-based TAM-targeting strategies and discuss their therapeutic potential.

1. Introduction

In the past decades, cancer has been the leading cause of death worldwide. Although traditional therapies including surgery, radiotherapy and chemotherapy have reduced the incidence of cancer-related deaths, they lack selectivity towards tumor cells [1]. As a result, high doses of drugs are often needed for cancer treatment, leading to the occurrence of cytotoxicity and multi-drug resistance, while the effect on tumor metastasis and advanced tumors is poor. The immune system of treated cancer patients is often weakened, which increases the risk of recurrence and metastasis [2,3]. Thus, there is urgent need for more effective treatment strategies to fight cancer. More recently, exploring the relationship between the tumor and host cells in the TME has become a focus of research. As discussed in several recent reviews, unlimited proliferation, metastasis, immune and apoptosis escape mechanisms of cancer cells are closely related to the TME [4,5,6].
In the past ten years, NP-based drug delivery systems with application in cancer therapy have aroused much interest [7,8,9]. According to the composition of NPs, NPs can be divided into metallic, organic, inorganic and polymeric nanostructures [10]. NPs can be utilized as delivery systems for different types of biomolecules, such as fluorophores, metals, peptides, proteins, oligonucleotides or biomimetic drugs, which gives NPs a multifunctional role in diagnosis and treatment. Controlled drug delivery by NPs can prevent premature drug degradation, improve drug absorption and reduce side effects. Depending on the NPs’ physicochemical properties, such as size, shape, charge and surface chemistry, as well as their optical properties, NPs have been widely used in tumor diagnostics and monitoring [11], drug delivery [12], optical imaging [13], development of synthetic vaccines [14] and the manufacturing of miniature medical devices [15] (Figure 1). In comparison to the bulk material, NPs possess new chemical and physical properties, can increase the biocompatibility and stability of conventional drugs and overcome problems of solubility. In addition, compared with traditional single-modality drugs, NPs can be designed to harbor multiple functions.
Most NPs are biodegradable and possess low cytotoxicity and low immunogenicity in vitro and in vivo [16]. In addition, the possibility to link targeting moieties can be utilized for site-specific delivery in vivo [17,18]. For example, the specific targeting of the BLP25 liposome vaccine to tumor cells mainly depends on the presence of an antibody on the liposome surface [19]. Biodegradable polymer materials with low cytotoxicity and high biocompatibility, such as polylactic acid (PLA), polyglycolic acid (PGA) and poly (lactic-co-glycolic acid) (PLGA), have been approved by the Food and Drug Administration (FDA) as drug delivery systems for vaccine antigens made of proteins, peptides and DNA [20,21]. PLGA is one of the most successfully developed polymers, and has been widely used in the encapsulation and targeted delivery of therapeutic agents to cancer cells to enhance the effect of anticancer therapy [22,23]. Compared to traditional therapies, NP-based therapeutics show longer circulation times in the bloodstream and encapsulated biological agents have longer half-lives than free drugs [24]. Polymeric NPs, such as those made of PLGA, typically show a sustained-release kinetics profile, which leads to uniform and prolonged drug levels in the body and reduced drug toxicity, compared to conventionally administered drugs [23,25]. In addition, PLGA NPs are often functionalized with polyethylene glycol (PEG), which improves the circulation time and half-life in the blood stream, and presents a stable linker to couple-targeting motifs for tumor-specific delivery [14,26,27].
To date, the FDA has approved more than 30 NP formulations for therapy or diagnostics, including metallic NPs, albumin NPs and liposomes, many of them with application in cancer therapy [20]. For example, Abraxane, an albumin-paclitaxel NP formulation, can passively target tumors through the enhanced permeability and retention (EPR) effect [28]. In addition, more than 50 NP formulations are currently in clinical trials and many more are in pre-clinical research [29].
In recent years, the TME has become a target of NP-based delivery systems aiming to prevent tumor progression and survival, to “re-educate” immunosuppressive M2 TAMs towards the M1 type with antitumor properties and to raise an antitumor immune response [7,8,9,30,31]. However, due to individual differences, tumor heterogeneity and complexity of the TME, different results have been reported [32]. Yet, targeting of TAMs in the TME in vivo could be key to the development of effective NP-based cancer treatments. In this review, we give an overview of TAM biology and refer to the latest research in the field of TAM-targeted immunotherapy. We will focus on strategies that utilize NPs to target and shape TAMs during tumor development and progression. In particular, we will pay special attention to how different types of TAMs interact with NPs, and how the physicochemical properties of NPs influence NP uptake.

2. Macrophages in Tumorigenesis

2.1. Tumor-Associated Macrophages (TAMs)

It is well known that the TME promotes tumorigenesis and tumor metastasis in addition to causing resistance to cancer therapy. Tumors are complex biological systems that primarily consist of cancer cells, vascular cells, mesenchymal lineage cells, immune cells of both lymphoid and myeloid origin and fibroblasts [33,34]. The interaction of these cell populations in tumors promotes the formation of the TME. TAMs represent one of the main tumor-infiltrating innate immune cells in the TME, playing a critical role in inflammation, which is well established to promote tumor growth and tumor progression.
TAMs differ both from tissue-resident and bone-marrow-derived macrophages [34,35] (Figure 2). Although macrophages were originally described as arising exclusively from circulating monocyte precursors [36], more recently it was shown that several organs harbor embryonic-derived populations of resident macrophages that maintain and self-renew throughout adulthood [33,37,38]. These tissue-resident macrophages include brain microglia, alveolar, pancreatic, peritoneal, splenic, kidney and gut macrophages, as well as Langerhans and Kupffer cells [33,37] (Figure 2A). There are contradicting reports about the cellular origin of TAMs. For example, it has been reported that TAMs were recruited from CCR2+ inflammatory monocytes in a mouse model of mammary cancer [34]. However, in recent years, with the new concept of the embryonic origin of tissue-resident macrophages, it has been confirmed that both embryonic- and monocyte-derived macrophages constitute the pool of TAMs in the TME [39]. In models of murine pancreatic ductal adenocarcinoma, a large portion of tissue-resident macrophages was shown to originate from embryonic development, and to expand during tumor development [40]. The authors further reported functional heterogeneity among TAM subsets in the TME, which was correlated with their cellular origin. Monocyte-derived TAMs played more potent roles in antigen presentation, while embryonically derived TAMs exhibited a profibrotic transcriptional profile, indicative of their role in producing and remodeling molecules in the extracellular matrix during tumor progression. At the onset of tumor development, TAMs are mainly proinflammatory macrophages which promote tumor progression by inducing the expression of inflammatory cytokines and proangiogenic factors [39,40]. TAMs further support tumor malignancy and metastasis by enhancing the invasion and extravasation ability of tumor cells, inhibiting antigen presentation, as well as stimulating tumor relapse after chemotherapy [41,42,43].

2.2. Macrophage Polarization and Tumor Development

Macrophages show remarkable functional plasticity and undergo specific differentiation in different tissue environments. Tissue-resident macrophages kill foreign pathogens and coordinate leukocyte infiltration and antitumor adaptive immunity through phagocytosis and the degradation of apoptotic cells, microorganisms and neoplastic cells [44]. However, cancer cells that have escaped the immune system slowly form an immunosuppressive TME, where the release of tumor-associated antigens as a consequence of tumor cell death is decreased and cancer cells successfully escape sanctions of the immune system [5]. TAMs are widely present in various tumors, participate in the formation of the TME and often exhibit a variety of polarization states under the influence of hormones, cytokines and regulatory molecules (Figure 3), which affect their role in innate and acquired immunity. In general, TLR and cytokine receptor signaling skews TAMs away from the “classically” activated tumoricidal phenotype (referred to as M1) towards an “alternatively” activated tumor-promoting one (M2) (Figure 3). M1 macrophage activation stimuli are interferon gamma (IFN-γ) and lipopolysaccharide (LPS), which induce strong microbicidal properties, production of the proinflammatory cytokines TNFα, type I interferons (IFNs), IL-1, CXCL1-3, IL-6 and IL-12 [45], and promote an IL-12-mediated Th1 response, which generates responses against intracellular parasites and neoplastic cells [46] (Figure 3A). In contrast, M2 macrophages are divided into type M2a, M2b and M2c, and the activation stimuli are IL-4, IL-10 or IL-13, immune complexes in combination with IL-1β or LPS, TGF-β or glucocorticoid stimulation [47]. A large number of M2-like TAMs reside in the TME and produce anti-inflammatory cytokines, pro-angiogenic and trophic factors, such as nitrate oxide and ornithine, and reduce the secretion of proinflammatory cytokines [48], which altogether have the ability to suppress antitumor immune responses [49], and promote tumor neovascularization as well as tumor growth [50] (Figure 3B).
The differential expression of various TLRs on macrophages is an important factor underlying macrophage polarization [51]. Mechanistically, macrophage polarization is regulated by interferon regulatory factors (IRFs) and the signal transducer and activator of transcription proteins (STAT) signaling pathway [52]. For example, IFN and TLR signals promote the polarization of macrophages towards the M1 type by activating the IRF/STAT1 signaling pathway [53]. Macrophage proliferation is a significant feature of advanced stages of cancer. Giurisato et al. demonstrated that a high proportion of proliferating macrophages in human tumors expresses extracellular-regulated protein kinase 5 (ERK5) [54,55]. The authors showed that ERK5 supported the proliferation of macrophages in melanoma tumor grafts in mice by suppressing p21 expression [56], which led to the halt of the macrophage differentiation program. Inhibition of ERK5 blocked macrophage polarization and reduced metastasis in vivo [57]. In another study, the authors demonstrated that ERK5-deficient TAMs displayed a proinflammatory phenotype. Mechanistically, ERK5 inactivation suppressed the phosphorylation of STAT3, thereby inhibiting the proliferation of melanoma and carcinoma grafts [57]. In contrast, IL-4- and IL-13-induced signaling pathways promote macrophages to polarize towards the M2 phenotype via STAT6 [58]. Apoptotic tumor cells promote macrophage polarization towards the M2 type via the phosphatidylinositol 3-kinase (PI3K)/protein kinase B (Akt) signaling pathway [59]. Activation of the Akt signaling pathway has also been shown to reduce tumorigenesis. Given that Akt subtypes (Akt1 and Akt2) play opposite roles in macrophage polarization, antitumor responses can be adjusted by the activation of different Akt subtypes [60] (Figure 3B). Importantly, M1 and M2 polarization are dynamic processes and the phenotypes of polarized macrophages can be reversed under different physiological and pathological conditions [61]. However, the molecular mechanisms that control the phenotype of macrophages are not fully understood yet.
TAMs resemble the M2 phenotype and play a harmful tumor-promoting role. When cancer occurs, IL-10 and TGF-β secreted by TAMs can prevent an antitumor immune response by inhibiting the function of antigen-presenting cells and effector T cells, leading to the escape of cancer cells and the formation of an immunosuppressive TME [62]. The multifunctional cytokine IL-10 plays a vital role in supporting tumorigenesis and inhibiting the activation of monocytes, T cells and macrophages required for antitumor immunity [63]. Studies in mouse tumor models showed that IL-10 not only inhibits the production of IL-12, but also inhibits the maturation of dendritic cells and limits the cytotoxic T cell response during chemotherapy [64]. Regulatory molecules released by Treg cells, such as IL-35, IL-10, adenosine and TGF-β, inhibit Th1, Th12 and Th17 immune responses, which promote antitumor functions [33,62,65,66]. IFN-γ produced by Th1 cells can up-regulate histocompatibility complex (MHC) class I molecules. The loss or down-regulation of MHC class I molecules on the surface of tumor cells is a tumor immune escape mechanism. Yoo et al. showed that defects in MHC class I molecules cause a poor prognosis in PD-L1-positive head and neck squamous cell carcinoma [67]. In addition, the interaction of PD-1/PD-L1 has been shown to suppress antitumor immunity. For example, Wang et al. showed that the intrinsic PD-1/PD-L1 axis of tumor cells not only inhibits the Akt and extracellular-signal-regulated kinase (ERK)1/2 signaling pathways and tumor growth but also prevents interaction with T cells [68]. In summary, although immune cells, such as T cells, can inhibit tumor growth, their function is inhibited in many cancers. Therefore, how to rationally and effectively reactivate the immune system to kill tumors remains a major challenge.

3. Macrophages in Tumor Therapy

T. Honjo and JP Allison won the Nobel prize in physiology and medicine in 2018 for the discovery of inhibition of negative immune regulation as a form of cancer therapy [69]. Compared to traditional therapies, immunomodulation has great potential in cancer treatment.

3.1. Tumor Therapy Targeting Tumor-Associated Macrophages (TAMs)

In the TME, due to the complex relationship between cancer cells and surrounding immune cells, single-modality cancer treatments are often ineffective. Recruitment of TAMs to the TME has been shown to be associated with resistance to immune-mediated tumor cell killing, and the inhibition of their recruitment or differentiation prevented tumor growth [70] (Figure 4). At present, targeting macrophages has become a focus in the development of current cancer treatments and much clinical and preclinical research is under way. TAM-targeting strategies for cancer therapy can be divided into two main approaches: depletion of TAMs (i.e., by preventing their recruitment) and reprogramming of TAMs.
Cancer cells secrete a variety of cytokines and chemokines to recruit monocytes to the tumor side, including IL-6, IL-8, IL-34, CCL2, CCL3, CCL5, CCL20, CSF-1, CSF-2 and C-X-C motif chemokine 12 (CXCL12). In addition, other cells of the TME can also secrete cytokines/chemokines to recruit monocytes. For example, IL-8 secreted by cancer-associated fibroblasts (CAFs), IL-17 secreted by Th17 cells and CCL2 secreted by mesenchymal stem cells (MSCs). Major targets of current strategies to prevent monocyte recruitment and polarization towards the immunosuppressive M2 state are the CSF-1-CSF-1R, CCL2-CCR2, CXCL12-CXCR4 and CCL5-CCR5 signaling axis [71].

3.1.1. CSF-1/IL-34-CSF-1R

CSF-1R on the surface of TAMs can combine with the ligands CSF-1 and IL-34 [72]. CSF-1 participates in the recruitment of TAMs to tumor tissues and the M2 polarization process of TAMs. The survival, proliferation and function of TAMs depend to a large extent on CSF-1R signaling (Figure 4). Blocking the CSF-1-CSF-1R axis inhibits the polarization of macrophages, thereby reducing tumor cell proliferation and inducing TAM apoptosis [73] (Figure 4). Many preclinical studies demonstrated high efficacies of CSF-1R inhibitors. For example, using mouse glioblastoma as an experimental object, researchers used CSF-1R inhibitors to reduce the expression of Arg1 (arginase 1) and Mrc1 (CD206, mannose receptor) genes in the TME, thereby inhibiting tumor growth and metastasis [74]. CSF-1R inhibitors, such as PLX6134 and GW2580, increased the CD8+ T cell response by inhibiting the CSF-1-CSF-1R signaling axis, eventually leading to TAM apoptosis [75]. Many current clinical treatment strategies targeting the CSF-1/CSF-1R pathway have been extensively developed [76], as summarized in Table 1. For example, in ongoing clinical trials targeting CSF-1, anti-CSF-1R monoclonal antibodies (such as IMC-CS4 in NCT01346358) block the binding of CSF-1 and IL-34 to CSF-1R and abrogate the recruitment and survival of TAMs, leading to TAM apoptosis and inhibition of tumor growth [77]. FPA008, an anti-CSF-1R antibody, is used in several clinical trials to treat advanced cancers, including lung, head, neck, pancreatic and ovarian cancer, as well as renal cell carcinoma and malignant glial tumor (i.e., NCT02526017 and NCT02471716) (Table 1). The first phase of the study, the dose-ranging phase, was well-tolerated by patients. However, current clinical studies have so far not shown that CSF-1R inhibitors, as a monotherapy, delay tumor growth. Combinational therapies of CSF-1R and CXCR2 antagonist demonstrated stronger effects on tumor treatment [78].

3.1.2. CCL2-CCR2

CCL2 secreted by tumor cells can attract Th1 cells, Th17 cells and macrophages, which express high levels of CCR2, to the tumor site [79]. The CCL2-CCR2 axis has been shown to recruit TAMs for the construction and maintenance of an immunosuppressive TME [80,81]. Inhibition of the CCL2–CCR2 interaction can significantly reduce the number of TAMs in the tumor, thereby inhibiting tumor growth and spread [82]. In addition, the CCL2-CCR2 signaling pathway regulates macrophage polarization during cancer treatment. In a mouse model of non-small-cell lung cancer (NSCLC), it has been shown that estrogen receptor alpha (Erα) can activate the CCL2-CCR2 axis to promote macrophage infiltration, M2 polarization and MMP9 production, which increased NSCLC cell invasion [83]. Mechanistically, ERα was shown to bind to estrogen response elements on the CCL2 promoter, thereby increasing CCL2 expression. Using anti-estrogens or CCR2 antagonists to target the CCL2-CCR2 axis may improve outcomes in NSCLC [83]. In a murine breast tumor model, targeting the CCL2-CCR2 axis by complexes of CCR2 siRNAs and TAT cell penetrating peptides enhanced the efficacy of immunotherapy and promoted the reprogramming of TAMs [84]. Preclinical studies of hepatocellular carcinoma have shown that activated hepatic stellate cells recruit macrophages through the CCL2-CCR2 pathway and have the ability to induce M2 phenotypic transformation [85]. In clinical studies, Carlumab (anti-CCL2 monoclonal antibody) was proven to effectively resist the metastasis of prostate cancer by inhibiting the CCL2-CCR2 signaling pathway (NCT00992186) (Table 1). As a CCR2 antagonist, PF-04136309 has been well-tolerated in clinical studies of metastatic pancreatic ductal adenocarcinoma (NCT02732938) (Table 1). The combination of PF-04136309 and the chemotherapeutic drugs nab-paclitaxel/gemcitabine promoted the reduction in TAMs in tumors [86]. Therefore, the CCL2-CCR2 signaling pathway represents a promising target in the process of cancer treatment.

3.1.3. CCL5-CCR5

In the TME, high concentrations of CCL5 can promote the accumulation of immunosuppressive cells and promote immune tolerance. Conversely, CCL5 can also recruit immune effector cells, such as lymphocytes, with high CCR5 expression (Figure 4). Phyllodes breast tumors promoted malignant progression by secreting CCL5 to recruit macrophages, which subsequently turned into TAMs [87]. In preclinical studies using mouse models of ovarian cancer, osteosarcoma, prostate and breast cancer, CCL5 antagonists inhibited cancer cell migration and invasion [88,89,90,91]. Mechanistically, crosstalk between CCR5 signaling and MMP-9 secretion has been reported. Anti-CCL5 blocking antibody or CCL5-shRNA knockdown in ovarian cancer stem-like cells decreased MMP-9 mRNA expression, and consequently reduced invasiveness [91]. Recently, the results of phase 2 clinical trials of CCR5 antagonist maraviroc in hematological malignancies and Hodgkin’s lymphoma showed that not only was the accumulation of monocytes was inhibited, but the migration of cancer cells was also effectively reduced (NCT01785810) (Table 1) [92]. Additionally, expression of the Raf kinase inhibitor protein (RKIP) can regulate TAM recruitment by blocking high-mobility group AT-hook 2 (HMGA2), leading to decreased expression of CCL5 [93]. This shows that targeted recovery of RKIP expression to reduce the secretion of CCL5 in tumor cells might be an important strategy to reduce macrophage infiltration [93].

3.1.4. CXCL12-CXCR4

The CXCR4-CXCL12 axis recruits and keeps TAMs in the tumor’s hypoxic area and contributes to neo-angiogenesis and oxygen delivery [94,95] (Figure 4). Activation of chemokine receptor CXCR4/CXCL12 signaling has been shown to be implicated in TAM recruitment in melanoma [96]. Moreover, high expression of CXCL4 was intimately related to the invasion potential of TAMs in tumor tissues and the development of metastasis in NSCLC [97]. Studies have found that blocking CXCL12 can activate the MAPK/PI3K/AP-1 signaling pathway in colon cancer cells, thereby blocking the growth of tumor cells [98]. Blocking the interaction of CXCL12 with its receptor, CXCR4, is also thought to regulate macrophage infiltration and prevent metastasis. For example, the combination of anti-CXCR4 and anti-PD-1 immunotherapy to regulate TAMs in the microenvironment of glioma cells not only promoted an antitumor immune response, but also increased the level of inflammatory cytokines [99] (Figure 4). This strategy is currently employed in preclinical models, including metastatic pancreatic ductal adenocarcinoma, pancreatic neoplasms and prostate cancer [96,100]. CXCR4 antagonist BL-8040 combined with PD-1 blockers in a phase 2 trial of pancreatic ductal adenocarcinoma enhanced the efficacy of chemotherapy (NCT02826486) (Table 1) [101].

3.1.5. CD47-SIRPα

The “do not eat me” signal CD47, overexpressed on the surface of cancer cells, forms a signaling complex with signal-regulatory protein α (SIRPα), a myeloid-specific immune checkpoint, enabling the escape from macrophage-mediated phagocytosis [102]. Therefore, inhibiting CD47-SIRPα signal transduction promotes the phagocytosis of tumor cells by macrophages and inhibits tumor growth [103]. Some preclinical studies have shown that in tumor models, anti-CD47 antibody treatment not only promoted the phagocytosis of cancer cells by macrophages, but also induced an antitumor cytotoxic T cell immune response [104]. In a mouse model of lymphoma, the bispecific anti-CD47/SIRPα antibody inhibited tumor growth and spread by targeting the multiple myeloid cell types in the tumor [105]. Anti-MUC1 and anti-EGFR (cetuximab) antibodies have achieved effective antitumor effects in carcinoma A549 cancer cells, causing tumor regression by inhibiting the binding of SIRPα and CD47 [106] (Figure 4). Currently, several clinical antitumor therapies targeting the CD47-SIRPα axis have been developed, such as two anti-CD47 monoclonal antibodies (Hu5F9-G4 and CC-90002), a soluble recombinant SIRPα-crystallizable fragment (Fc) fusion protein (TTI-621) [107,108] (i.e., NCT02216409 and NCT02678338) (Table 1). For example, in phase I clinical trials of advanced solid tumors and blood cancers, CC-90002, an anti-CD47 antibody, is used to block over-expressed CD47 on cancer cells, thereby inhibiting the CD47-SIRPα signaling pathway and enhancing the phagocytosis of tumor cells by macrophages [102].
In the TME, re-education of TAMs into M1-like macrophages has gradually become a more effective cancer treatment strategy. CD40 is expressed on dendritic cells, B cells, macrophages and even tumor cells (Figure 4). Under normal conditions, a CD40 ligand (CD40L and CD154) on activated T helper cells interacts with antigen-presenting cells via CD40, resulting in CD8+ T cell activation [109]. Tumor cells can undergo apoptosis or growth inhibition after CD40 ligation [27]. Multiple agonistic CD40 antibodies have been developed for clinical use. Currently, a variety of CD40-targeting agonist monoclonal antibodies are in phase 1 in clinical trials, including CDX-1140 and CDX-301 [110] for the treatment of advanced malignant tumors and solid tumors (Table 1). Responses in melanoma and pancreatic cancer have been reported [111]. Interestingly, blocking CD40 also promotes the polarization of TAMs to M1-like macrophages via the production of IL-12 [109].

3.1.6. PI-3 Kinase γ (PI3Kγ)

PI-3 kinase γ (PI3Kγ), which is highly expressed in immune cells, such as mast cells, neutrophils and eosinophils, is a promising target for stimulating TAM re-polarization (Figure 4). The PI3Kγ signaling pathway activated by Akt and mTOR has the ability to stimulate C/EBPβ activation and inhibit the activation of NF-κB, which not only promotes TAM repolarization, but also inhibits tumor growth [112]. For example, in an animal model of pancreatic ductal adenocarcinoma, inhibiting the activity of PI3Kγ can reprogram TAMs and stimulate CD8 + T cells, thereby inhibiting tumor cell migration [110]. Signal pathway inhibitors for PI3Kγ have also demonstrated anti-metastatic effects in clinical tumor research. For example, PI3K/mTOR inhibitor PF-05212384 (Table 1) enhanced the antitumor effect and reduced tumor metastasis in head and neck squamous cell carcinoma in vivo [113].
Small molecule drugs, such as TLR agonists and cytokines, have attracted much attention in cancer therapy. TLR-mediated signaling pathways can repolarize M2 macrophages into the M1 type during tumor treatment to enhance antitumor immunity. For example, Resimod (TLR agonist), which targets TLR7/8, re-educates TAMs towards the M1 type in human colon carcinoma, while inhibiting tumor growth [114]. Cytokines, such as IL-6, IL-17 and IL-23 produced by the NF-κB or STAT3 signaling pathway, induce cancer proliferation and metastasis by promoting TAMs to inhibit a cytotoxic T cell response [115] (Figure 4). Therefore, NF-κB and STAT3 are also targets for drug therapy. NF-κB/STAT inhibitors can regulate the polarization of macrophages and inhibit tumor growth and metastasis [116,117]. Interestingly, the natural lipids ceramide and palmitic acid were recently shown to be effective inhibitors of the NF-κB or STAT3 signaling pathway, which inhibited the migration capacity of colorectal cancer cells, and at the same time promoted the repolarization of M2 TAMs towards the M1 type [118]. An ongoing clinical trial explores the STAT3/NF-κB/polycytosine kinase inhibitor IMX-110 in combination with low-dose doxorubicin (NCT03382340) to kill cancer cells (Table 1). In this context, NPs are used as drug carriers to deliver IMX-110 to the TME. Such nano-drug delivery systems proved to be advantageous in delivering drugs to the TME and TAMs, compared to systemically administered naked drugs [119].
Class IIa histone deacetylase (HDAC) has the property of regulating the antitumor immunity of macrophages in tumors (Figure 4A). Studies by Lobera et al. have demonstrated that class IIa HDAC inhibitors can promote macrophages to favor the proinflammatory M1 phenotype in the TME [120]. Similarly, research by Li et al. in mouse tumor models showed that HDAC inhibitors can not only reduce the migration of MDSCs to tumor sites, but also enhance the expression of antitumor phenotypes and the activation of T cells [49]. Furthermore, in the breast cancer model of MMTV-PyMT transgenic mice, Guerriero et al. found that class IIa HDAC inhibitors (TMP195) recruit and differentiate macrophages as well as promote their conversion into hyperphagocytic macrophages, thereby reducing tumor burden and tumor lung metastasis [121]. Therefore, the use of class IIa HDAC inhibitors to extract the potential of antitumor function in TAMs is a potential means of cancer treatment.
The crosstalk between tumor cells and macrophages is affected by microRNAs (miRNAs). Tumor-derived factor miRNAs, a kind of small non-coding RNA, can regulate target gene expression after transcription [122] (Figure 4A). According to reports, miRNA dysregulation can induce the occurrence of a variety of malignant tumors, and it has a significant impact on tumor immune escape, invasion, angiogenesis, migration and drug resistance [123,124]. For example, Zhao et al. found through studies on hepatocellular carcinoma specimens that miRNA-144/miRNA-451a promotes the polarization of TAMs to M1 macrophages by targeting hepatocyte growth factor and macrophage migration inhibitory factor [125]. In addition, small interfering RNAs (siRNA) can also target TAMs and modify the polarization and function of M2 TAMs by silencing the corresponding target gene [126] (Figure 4A). For example, Song et al. silenced endothelial growth factor and placental growth factor by delivering VEGF siRNA/PIGF siRNA to M2-type macrophages and breast tumor cells in vivo. The results showed that M2-type macrophages were re-educated towards M1-type macrophages, and the proliferation and migration of breast cancer cells were inhibited [127].
Table 1. Clinical trials of agents that target TAMs for cancer treatment.
Table 1. Clinical trials of agents that target TAMs for cancer treatment.
ActionNCT NumberTumor TypeStatusTargetAgent NamePhaseEffect
Targeting signal transduction pathwaysNCT02404844Breast cancerCompletedPI3K pathwayBKM120
Tamoxifen
Phase 2Well-tolerated, preliminary activity in advanced cancers [128,129]
NCT02384239Breast cancerActive, not recruitingPI3K pathwayPalbociclibPhase 2NA
NCT02058381Breast cancerCompletedPI3K pathwayAlpelisib (BYL719)
Buparlisib (BKM120)
Phase 1bWell-tolerated, preliminary antitumor activity [130]
NCT01298713Breast neoplasmsCompletedPI3K-Akt- mTOR pathwayTamoxifen
Everolimus
Phase 2Well-tolerated, positive correlation [131]
NCT01971515Solid tumorCompletedp70S6K/AktMSC2363318APhase 1NA
NCT02322853Metastatic breast cancerTerminatedMAK pathwayTamoxifen, ralimetinib (LY2228820)Phase 2NA
NCT03458221Ovarian carcinomaNot yet recruitingST pathwayItraconazolePhase 3NA
NCT00189358Ovarian cancer, cancer of the fallopian tube and peritoneal cancerCompletedthe EGFR pathwayZD1839
Tamoxifen
Phase 2No side effects
Ineffective for refractory/drug resistant cancers [132]
NCT02040857Breast cancerActive, not recruitingCDK 4/6PalbociclibPhase 2Well-tolerated [133]
NCT01920061NeoplasmCompletedPI3K/mTOR pathwayPF-05212384 (gedatolisib)Phase 1NA
NCT02228681Endometrial cancerActive, not recruitingmTOR pathwayTamoxifen, everolimus, letrozole and medroxyprogesterone acetatePhase 2Enhanced anticancer efficacy
NCT04318223Metastatic breast cancer and locally advanced breast cancerRecruitingCDK4/6PalbociclibPhase 2NA
NCT02868268NeuroblastomaRecruitingPD-1/PD-L1 pathwayGene panel sequencing of tumor specimensPhase NANA
NCT03382340Pancreatic cancer, breast cancer and ovarian cancerRecruitingSTAT3/NF-kBImx-110Phase 1, 2NA
NCT04504552Oral premalignant lesionsNot yet recruitingPD-1/PD-L1 pathwayAvelumabPhase 2NA
TAM recruitmentNCT02471716Pigmented villonodular synovitis and tenosynovial giant-cell tumorCompletedCSF-1RFPA008Phase 1, 2NA
NCT02526017Advanced solid tumors, not limited to lung cancer, head and neck cancer, pancreatic cancer and ovarian cancerCompletedCSF-1RFPA008Phase 1, 2NA
NCT02323191Solid tumorCompletedCSF-1RO5509554Phase 1NA
NCT02760797NeoplasmsCompletedCSF-1RO5509554Phase 1NA
NCT03153410Pancreatic cancerRecruitingCSF-1IMC-CS4Early phase 1NA
NCT04066244Amyotrophic lateral sclerosisRecruitingCSF-1BLZ945Phase 2NA
NCT02829723Advanced solid tumorsRecruitingCSF-1BLZ945Phase 1, 2NA
NCT02880371Advanced solid tumorsCompletedCSF-1ARRY-382Phase 1, 2NA
NCT01316822Metastatic cancerCompletedCSF-1ARRY-382Phase 1NA
NCT02777710Colorectal cancer, pancreatic cancer, metastatic cancer and advanced cancerCompletedCSF-1PLX3397Phase 1NA
NCT02584647Sarcoma and malignant peripheral nerve sheath tumorsRecruitingCSF-1PLX3397Phase 1NA
NCT02371369Pigmented villonodular synovitis, giant-cell tumors of the tendon sheath and tenosynovial giant-cell tumorActive, not recruitingCSF-1PLX3397Phase 3Mild side effects, improved patient symptoms and functional outcome [134,135,136,137]
NCT01804530Solid tumorTerminatedCSF-1PLX7486 TsOHPhase 1NA
NCT03320330Recurrent malignant solid neoplasm and osteosarcoma
Refractory malignant solid neoplasm and osteosarcoma
Active, not recruitingSEMA4D or CD100Pepinemab (VX15/2503)Phase 1
Phase 2
NA
NCT03557970Recurrent acute myeloid leukemia and refractory acute myeloid leukemiaActive, not recruitingCSF-1JNJ-40346527Phase 2NA
NCT03690986Squamous cell carcinoma of the head and neckRecruitingRecruitingVX15/2503 (ipilimumab) NivolumabPhase 1NA
NCT03769155Pathologic stage IIIB/C/D cutaneous melanoma AJCC v8RecruitingRecruitingVX15/2503 (pepinemab)Phase 1NA
NCT03177460Prostate adenocarcinomaActive, not recruitingCSF-1JNJ-40346527Phase1NA
NCT01572519Relapsed or refractory Hodgkin lymphomaCompletedCSF-1JNJ-40346527Phase 1NA
NCT01597739Arthritis, rheumatoidCompletedCSF-1JNJ-40346527Phase 2NA
NCT02732938Metastatic pancreatic ductal adenocarcinomaTerminatedCCR-2PF-04136309Phase 2Safe and tolerable [86]
NCT01413022Pancreatic neoplasmsCompletedCCR-2PF-04136309Phase 1NA
NCT01785810Hematologic malignancyCompletedCCR5MaravirocPhase 2NA
NCT02826486Metastatic pancreatic adenocarcinomaActive, not recruitingCXCR4BL-8040Phase 2Safety, efficacy and immunobiological effects [101]
NCT00992186Prostate cancerCompletedCCL2CarlumabPhase 2NA
NCT01494688Solid tumorCompletedCSF1-RRG7155Phase 1NA
Targeting TAM activationNCT04331067Triple-negative breast cancerNot yet recruitingTIL and TAMPaclitaxel, carboplatin
Nivolumab, cabiralizumab
Phase 1, 2NA
NCT03285607Breast cancerWithdrawnCSF-1MCS110, cyclophosphamide and paclitaxelPhase 1NA
NCT02435680Triple-negative breast cancerCompletedCSF-1MCS110, carboplatin and gemcitabinePhase 2NA
NCT01643850Giant-cell tumor of the tendon sheathCompletedCSF-1MCS110
Placebo
Phase 2NA
NCT00757757Prostate cancer and bone metastasesTerminatedCSF-1MCS110Phase 1, 2NA
NCT01346358NeoplasmsCompletedCSF-1RIMC-CS4Phase 1NA
NCT03153410Pancreatic cancerRecruitingCSF1-RIMC-CS4Phase 1NA
NCT01309230Ovarian cancerActive, not recruitingBi-shRNA furin and GMCSFVigil™Phase 2NA
NCT02390752Neurofibroma, plexiform, precursor cell lymphoblastic, leukemia–lymphoma, leukemia and acute sarcomaRecruiting PLX3397Phase 1, 2NA
NCT04079712Metastatic large-cell neuroendocrine carcinoma, metastatic neuroendocrine carcinoma, metastatic neuroendocrine neoplasm and metastatic small-cell carcinomaRecruitingPD-L1Cabozantinib
Cabozantinib S-malate
Ipilimumab
Nivolumab
Phase 2NA
NCT02265536Neoplasms and neoplasm metastasisCompleted LY3022855Phase 1NA
NCT04550624Advanced cholangiocarcinomaNot yet recruiting Pembrolizumab injection
Lenvatinib mesylate
Phase 2NA
NCT01444404Advanced malignancy, advanced solid tumors, cancer, oncology and tumorsCompletedCSF-1RAMG 820Phase 1NA
NCT02223312Cancer
Hematologic malignancies
WithdrawnDCTAPA-pulsed DC vaccinePhase 1, 2NA
NCT01217229Hodgkin lymphomaCompleted PLX3397Phase 2NA
NCT01804530Solid tumor, tenosynovial giant-cell tumorTerminatedCSF-1RPLX7486 TsOHPhase 1NA
NCT01596751Metastatic breast cancerCompletedDCPLX3397
Eribulin
Phase 1, 2NA
NCT00637390Ovarian cancer, fallopian tube cancer and peritoneal cancerTerminatedCD52AlemtuzumabPhase 1NA
NCT01525602Solid tumorsCompleted PLX3397
Paclitaxel
Phase 1bNA
Reprogramming TAMsNCT03285607Breast cancerWithdrawnCSF-1MCS110, doxorubicin, cyclophosphamide and paclitaxelPhase 1NA
NCT00492167NeuroblastomaActive, not recruiting Beta-glucan
monoclonal antibody 3F8
Phase 1NA
NCT03954691Cancer of head and neckNot yet recruitingMicroglia, macrophages and NK cells.Brain-tumor-immune cell communicationNANA
NCT02953782Colorectal neoplasms
Solid tumors
CompletedCD47Hu5F9-G4Phase 1, 2NA
NCT03558139Solid tumor and ovarian cancerActive, not recruitingCD47Hu5F9-G4Phase 1NA
NCT02216409Solid tumorCompletedCD47Hu5F9-G4 Safe and tolerable [108,138]
NCT01561911Cancer, neoplasms and lymphomaCompletedCD40Chi Lob 7/4 (a chimeric monoclonal antibody)Phase 1Activated B and NK cells [139]
NCT03922477Acute myeloid leukemiaRecruitingCD47Hu5F9-G4Phase 1NA
NCT00912327Colorectal cancerCompleted Imprime PGGNANA
NCT03248479Acute myeloid leukemia and myelodysplastic syndromesRecruitingCD47Magrolimab
Azacitidine
Phase 1NA
NCT01433172Lung cancer and adenocarcinomaCompletedCD40 and CCL21GM.CD40LCCL21 VaccinationsPhase 1, 2Well-tolerated
NCT01904123Metastatic malignant neoplasm in the brain, metastatic melanoma, recurrent brain neoplasm, recurrent glioblastoma and recurrent malignant gliomaRecruitingSTAT3STAT3 inhibitor WP1066Phase 1NA
NCT02953509Lymphoma, non-Hodgkin’s; lymphoma, large B cell, diffuse; and indolent lymphomaRecruitingCD47Hu5F9-G4Phase 1, 2NA
NCT00911560NeuroblastomaRecruiting Adjuvant OPT-821 in a vaccine containing two antigens (GD2L and GD3L) covalently linked to KLHPhase 1, 2NA
NCT01103635Recurrent melanoma and stage IV melanomaCompletedCD40CD40 agonist monoclonal antibody CP-870893, tremelimumabPhase 1NA
NCT03527147NHL, DLBCL, non-Hodgkin’s lymphoma and diffuse large B cell lymphomaRecruitingCD47Hu5F9-G4Phase 1NA
NCT02157831Solid tumorsCompletedCD40CP-870893Phase 1NA
NCT02225002Advanced solid tumorsCompletedCD40CP-870893Phase 1NA
NCT00711191Pancreatic neoplasmCompletedCD40CP-870893Phase 1Tumoricidal, the depletion of tumor stroma [140]
NCT00607048NeoplasmsCompletedCD40CP-870893Phase 1Safe and no long-term [141]
NCT01456585Adenocarcinoma pancreasCompletedCD40CP-870893Phase 1NA
NCT01103635Recurrent melanoma and stage IV melanomaCompletedCD40CP-870893Phase 1NA
NCT01008527MelanomaCompletedCD40CP-870893Phase 1NA
NCT02482168Solid tumorsCompletedCD40APX005MPhase 1NA
NCT01839604Advanced adult hepatocellular carcinoma and hepatocellular carcinoma metastaticCompleted AZD9150Phase 1Safe and tolerable
NCT02367196Hematologic neoplasmsActive, not recruitingCD47CC-90002Phase 1NA
NCT02641002Leukemia, myeloid, acute and myelodysplastic syndromesTerminatedCD47CC-90002Phase 1NA
NCT02663518Hematologic malignancies and solid tumorRecruitingSIRPαTTI-621Phase 1NA
NCT02890368Solid tumors, mycosis fungoides, melanoma, merkel cell carcinoma, squamous cell carcinoma, breast carcinoma, and papillomavirus-related malignant neoplasmTerminatedSIRPαTTI-621Phase 1NA
NCT02665416Advanced/metastatic solid tumorsCompletedCD40RO7009789Phase 1NA
NCT02760797NeoplasmsCompletedCD40RO7009789Phase 1NA
NCT02304393Solid tumorsCompletedCD40RO7009789Phase 1NA
NCT02588443Pancreatic cancerCompletedCD40RO7009789Phase 1NA
NCT03329950Advanced malignanciesRecruitingCD40CDX-301/CDX-1140Phase 1NA
NCT00899574Breast cancer
Breast neoplasms
CompletedTLR7ImiquimodPhase 2NA
NCT01421017Breast cancer, metastatic breast cancer and recurrent breast cancerCompletedTLR7ImiquimodPhase 1, 2Supplemented the response evaluation criteria in solid tumors [142,143]
NCT04116320Melanoma, breast cancer, squamous cell cancer, non-small-cell lung cancer, cervical cancer, urothelial carcinoma, ovarian cancer, small-cell lung cancer and esophageal cancerRecruitingTLR7ImiquimodPhase 1NA
NCT03196180Cervical intraepithelial neoplasiaActive, not recruitingTLR7ImiquimodPhase 1NA
NCT00319748Breast cancer, ovarian cancer, endometrial cancer and cervical cancerCompletedTLR7852APhase 2NA
NCT00719199Colorectal cancerTerminatedTLR9IMO-2055Phase 1NA
NCT01040832Squamous cell carcinoma of the head and neckCompletedTLR9IMO-2055Phase 2NA
NCT01360827Squamous cell carcinoma of the head and neckTerminatedTLR9IMO-2055Phase 1NA
NCT02829723Advanced solid tumorsRecruitingCSF1RBLZ945Phase 1, 2NA

4. Interaction of Macrophages with NPs

As outlined in Section 3.1, various drugs that target TAMs are currently in preclinical and clinical development (Table 1). However, their application in the clinic has been limited by side effects and the lack of specific targeting properties [144]. To overcome these obstacles, various strategies utilizing NPs loaded with therapeutics and partially functionalized with TAM-targeting motifs have been developed over the last years (Table 2). In this section, we first discuss how macrophages interact with NPs in vitro and in vivo and how their polarization status affects the efficacy of NP uptake. Finally, we discuss the current progress in the development of NPs to deplete or re-educate TAMs for therapeutic purposes.

4.1. NP-Based Cancer Therapies Targeting TAMs

Compared to normal tissues, tumor tissues are rich in blood vessels. The rapid growth of tumors and associated blood vessels leads to a decrease in blood vessel wall density and vascular endothelial space and an increase in capillary permeability, resulting in high permeability for macromolecules and NPs, namely by the enhanced permeability and retention (EPR) effect [145]. With the gradual disclosure of the role of TAMs in tumor tissues, it has been found that NPs are significantly enriched in TAM-rich tumor tissues compared to TAM-deficient tissues, and the amount of NP uptake by a single TAM is much larger than that of a tumor cell [146]. Furthermore, TAMs can engulf NPs in large quantities and carry NPs to the anaerobic area of the tumor, thus making a great contribution to the enrichment of drugs at the tumor site [147]. Considering the distribution of TAMs in tumor tissues and their ability to engulf NPs, NP-based strategies targeting TAMs mainly include: (1) the utilization of TAMs as a drug carrier “reservoir” to carry nanomaterials into the central area of solid tumors in order to infiltrate and enrich drugs at the tumor site, (2) activation of TAM autophagy and (3) targeting TAMs with drugs that lead to the transformation of inhibitory (M2 type) macrophages into protumorigenic macrophages (M1 type), thereby inhibiting tumor growth and metastasis [148] (Table 2).

4.2. NP Uptake by Macrophages

Next to neutrophils and dendritic cells, macrophages belong to a group of specialized phagocytes that continuously sample their environment in the process of host defense. The fact that macrophages have multiple subtypes and take on various phenotypes depending on the microenvironment they reside in, can be exploited by macrophage-targeting NPs [149,150]. The surface of NPs can be modified with targeting motifs for specific receptors and/or cell types. Examples of targeting motifs include opsonins derived from serum, such as complement factors or immunoglobins that induce phagocytosis by Fc receptors and complement receptors, respectively, or antibodies, polypeptides and small molecules (i.e., folic acid, hyaluronic acid) [151,152]. Macrophages internalize larger substances (>500 nm) through phagocytosis (Figure 5). In the TME, internalization of particles coated with ligands or motifs (small, repeated, and conserved biological units), such as on the surface of pathogens, occurs via the process of receptor-mediated endocytosis [153]. Receptor-mediated endocytosis is one of the most important processes by which NPs enter cells (Figure 5). This process is mainly mediated by scavenger receptors, integrin receptors and thrombospondin receptors [154,155]. Specific macrophage receptors recognize unique motifs presented on the surface of pathogens, drugs, DNA, or enzymes, and induce a corresponding signaling cascade. In the TME, the surface of M1 macrophages is rich in TLR2/TLR4, CD80, CD86, CD40, CSF1, IL-1R1, INF-γ-R and MHC-II receptors (Figure 3), which are often used as targets for cancer therapy [156,157,158]. For example, the CD40 agonist antibody CP-870893 with tumor targeting function can activate M1 receptor (CD40) during cancer treatment, thereby causing macrophages to rapidly infiltrate tumors and promote tumor matrix depletion in the immunosuppressive tumor microenvironment [140]. M2-type macrophages usually express high levels of Dectin-1, DC-SIGN, mannose receptor (CD206), scavenger receptor A (CD204), IL receptor, MHC-II, CSF-1R, CD163, CCR2, TLR1/8, Arginase 1 (ARG1), galactose receptor and CXCR2 [159].

4.3. Effect of NP Physical and Chemical Properties on Cellular Uptake

The extent of cellular uptake of NPs is influenced by several properties, such as NP size, surface charge, hydrophobicity, hydrophilicity, incubation time and NP concentration [160,161], as well as interaction with cell surface receptors and serum proteins [162]. Thus, the desired macrophage-targeting NPs for tumor therapy should have a size that is efficiently taken up by macrophages, be biocompatible, possess a large drug loading capacity and a slow release rate in addition to selectively accumulating at the tumor site.

4.3.1. Size

The size of the NPs directly affects the extent of cellular uptake and intracellular routing (Figure 5). For example, particles with a diameter smaller than 500 nm enter the cells by pinocytosis and particles greater than 500 nm, such as bacteria, dead cells and larger sized NPs, are taken up through phagocytosis [163] (Figure 5). For polymeric NPs (PLA/PLGA), it has been shown that changes from a nanometer to a micrometer range (>1 μm) drastically reduce the cellular uptake rate of NPs [13,164,165]. NPs smaller than 150 nm enter the cell via clathrin-mediated endocytosis [166,167]. In addition, NPs with a diameter less than or equal to the size of cell membrane pores often directly penetrate the cell membrane, such as small metal clusters and dendrimers [168,169], or their uptake is facilitated by cell-penetrating peptides that induce the formation of membrane pores and membrane destabilization [170].
Interestingly, NP size has been shown to be an important determinant in inducing either a Th1 or Th2 response. PLGA NPs of 200–600 nm in size were readily taken up by macrophages and induced a Th1 response (characterized by IFN-γ production and up-regulation of MHC class I molecules), albeit at low antibody titers. In comparison, two–eight-micrometer-sized NPs were tethered to the membrane (were not taken up), but elicited high and long-lasting antibody titers from single-point immunization and showed a Th2 response (characterized by IL-4 secretion, up-regulated MHC class II molecules) [165]. This offers an exciting possibility to modulate an immune response. Depending on the co-administration of adjuvants, PLGA NPs could either induce a M1-type or M2-type response in vivo [171].
Once NPs come into contact with plasma or other protein-containing biological fluids, their surface immediately forms a protein corona with biological activity [172], which affects the interaction of NPs with macrophages [173]. Importantly, the protein corona covering the NP’s surface may hinder the specific interaction between targeting ligands and their receptors. For example, Salvati et al. found that a protein corona shielded a transferrin-functionalized NP’s surface from binding to the transferrin receptor [174]. Contrary, the study by Hoppstädter et al. showed that a protein corona adsorbed on the NP’s surface could be exploited to target M2-macrophage-specific surface receptors, which promoted the uptake of NPs [175]. In line with this finding, SiO2 NPs coated with a hard corona consisting of fibrinogen, immunoglobin G, low-density lipoproteins, high-density lipoproteins or the hormone Kallikrein promoted the binding of NPs to M2 macrophages [176]. As the size and surface area of NPs increases, the unspecific adsorption capacity of proteins from body fluids by the NP surface increases and plays a more important role in NP uptake and interaction with cell membrane receptors [177].
In vivo, the size of the NPs determines whether the NPs enter the lymphatic capillaries or remain at the injection site. For example, preclinical studies conducted in mice found that NPs with diameters between 10 and 100 nm quickly reached the kidneys through capillaries and were easily secreted, while NPs with diameters greater than 130 nm could not pass through the glomerular basement membrane [178,179]. Therefore, to avoid clearance by the mononuclear phagocytic system before reaching the TME, and to meet the needs of drug loading, NPs with a diameter in the size range of 100–200 nm are preferred [30]. In addition, NPs with a diameter in the range of 100–200 nm can leak through the tumor’s blood vessels (EPR effect) and are quickly adsorbed by dendritic cells in the lymph nodes [180,181]. Similar to the EPR effect in solid tumors, the vascular permeability of inflamed joints in rheumatoid arthritis allows for passive targeting of NPs to the inflammation side. The targeting efficacy of PEGylated liposomes of different sizes (70, 100, 200 and 350 nm) and surface charges were evaluated in vivo in a mouse model by near-infrared fluorescence imaging [182]. The results showed that liposomes with 100 nm diameter and a slight negative charge had better in vivo circulation times and inflamed joint targeting than other liposomes did [182].

4.3.2. Shape

It has been demonstrated that the shape (spherical, oblate, cubic, worm-shaped, rod-shaped) of the NPs plays a vital role in their interaction with the cell membrane [183,184,185]. Li et al. showed that the shape of NPs affects the immune response of macrophages and the efficacy of passive and targeted NP uptake by macrophages [186]. Spherical glyco-NPs were internalized by clathrin- and caveolin-mediated endocytosis, while cylindrical glyco-NPs were mainly dependent on clathrin-mediated endocytosis. Interestingly, the authors found that longer and cylindrical NPs preferentially induced an IL-6-dependent inflammatory response, compared to shorter cylindrical NPs. Similar, Niikura et al. reported that the size and shape of AuNPs coated with West Nile virus membrane protein for vaccine purposes determined the type of inflammatory response [187]. Rod-shaped AuNPs induced significant levels of inflammasome-dependent cytokine secretion (IL-1β, IL-18), while spherical AuNPs significantly induced the inflammatory cytokines TNF-α, IL-6, IL-12 and GM-CSF.
A comparative uptake study in human leukocyte populations found that macrophages take up small rod-shaped AuNPs (15 and 50 nm) with greater ease compared to spherical AuNPs [188]. A possible explanation for why differently shaped NPs induced different cellular responses was provided by Hinde et al. [189]. The authors found that polymeric NPs with different shapes (but identical surface charges) moved across various cellular barriers (plasma membrane, endosomal and lysosomal membranes as well as nuclear envelopes) [189]. However, rods and worms, but not micelles and vesicles, could diffuse to the nucleus via passive diffusion, thus improving their nuclear access rate [189]. In summary, NP shape is an important point to consider when designing NP systems to deliver drugs across cell membranes to macrophages.

4.3.3. Surface Charge

The surface charge of NPs plays a key role in determining the interaction with the cell membrane, the mechanism of NP uptake, intracellular localization, immune response and cytotoxicity. For example, the study by Du et al. showed that a positive surface charge promoted NP uptake by Caco-2 cells (human colon adenocarcinoma cell line) and small intestinal epithelial cells in vitro and in vivo, respectively [190]. There are many molecules with different charges in biological systems. In vitro studies have found that phosphatidylserine phospholipids, the major components of the plasma membrane, carry anions that cause the cell membrane to be negatively charged [191]. Therefore, cationic NPs interact with the cell membrane through electrostatic interaction. In vitro studies using RAW 264.7 murine macrophages found that both positively and negatively charged NPs showed higher uptake rates after opsonization in fresh mouse serum, but positively charged NPs induced higher cytotoxicity rates [192]. In vivo biodistribution studies demonstrated that undesirable liver uptake was increased for highly positively or negatively charged NPs, likely due to active phagocytosis by macrophages (Kupffer cells) in the liver. However, liver uptake was low and tumor uptake was high when the surface charge of NPs was slightly negative. This study suggests that a slight negative charge may be introduced to the NP surface to reduce the undesirable clearance by the reticuloendothelial system (RES) and to increase tumor localization [192]. Similar results have been obtained for chitosan NPs, demonstrating that NPs with slight negative charges and a size of 150 nm accumulated in tumors more efficiently [193].
Getts et al. showed that cationic NPs can induce TAM proinflammatory responses; anionic NPs are significantly less immunogenic than cationic NPs [194]. NPs are usually coated with polymers (such as PEG) or zwitterionic structures to improve their biocompatibility, increase their stability and extend their circulation time in the body [24,195,196].

4.3.4. Surface Hydrophobicity

NPs hydrophobicity directly affects the interaction with cells in vivo. Hydrophobic NPs are easily covered by protein complement forming a corona, which further enhances the surface hydrophobicity of NPs [197]. For example, Qianhui et al. found that hydrophobic NPs easier adsorb proteins, while hydrophilic NPs have a higher protein exchange rate [197]. Coating NPs with PEG and its analogs can not only prevent the formation of a protein corona [197], but also enhances the NPs hydrophilicity and increases the possibility of reaching the tumor site.

4.4. Effect of Macrophage Polarization Status on NP Uptake

The literature suggests that the polarization status of macrophages severely affects NP uptake. Jones et al. showed that mouse strains that are prone to Th1 immune responses clear 300 nm cylindrical PEG hydrogel NPs at a slower rate than Th2-prone mice. The NP uptake capacity was increased when macrophages from Th1 strains were differentiated towards M2 macrophages, while M1 macrophages were more likely to take up NPs larger than 500 nm in diameter [198]. Thus, the polarization status of macrophages influences NP clearance capacity. Hoppstadter et al. investigated the uptake of fluorescent silica NPs in different types of macrophages. The results showed that M2 macrophages obtained from lung tumors absorbed more NPs than M1 macrophages isolated from the surrounding lung tissue [175]. MacParland et al. measured the uptake of 100-micrometer-sized AuNPs by different macrophage subtypes, and the results showed that M2c macrophages had the strongest affinity for ingesting AuNPs, followed by M2 macrophages and M1 macrophages, while monocytes had the weakest affinity for NPs [199]. The uptake pattern of silica NPs by M2c macrophages and M1 macrophages in vitro was similar to that of AuNPs [200,201]. Thus, these studies suggest that M2 macrophages promote the internalization of NPs [200].
In contrast to these findings, Herd et al. found a higher uptake of Stöber silica NPs of 130 nm in size in M1 macrophages compared to M2 macrophages [200]. Importantly, they also found NPs to preferentially accumulate in liver and spleen tissues in vivo in macrophages with a Th1/M1 phenotype response. An explanation for the preferential uptake of silica NPs by M1 macrophages is that some M1 surface receptors have been implicated in promiscuous poly anionic and silanol binding in vivo [200]. In addition, a study by Pajarinen et al. found that the uptake of titanium NPs by M1 macrophages in vitro was stronger than that of M2 macrophages [202].
In summary, while the polarization status of macrophages clearly determines the efficacy of NP uptake, controversy remains on which macrophage subset takes up more NPs in vitro and in vivo. This controversy might be explained by differences in physicochemical properties of the NPs (size, shape and material) tested in the studies, which can lead to preferential uptake behavior in M1 or M2 subtypes.

4.5. Modulation of TAMs by NPs via Passive Targeting In Vivo

Table 2 summarizes various strategies for targeting TAMs in the course of cancer immunotherapy by different types of NPs. In vivo, NPs need to overcome many physiological barriers, including clearance by immune cells in the liver and spleen, various endothelial membranes, interstitial spaces, the plasma membrane, endosomal/lysosomal membranes and finally the nuclear membrane and pore complex [203]. Upon intravenous injection, a fraction of NPs reaches the TME primarily via the EPR effect, while another part is taken up by the macrophages of the liver and spleen, which are in direct contact with the bloodstream [204,205]. In addition, the abnormal structure of blood vessels in tumors allows the penetration of NPs, while defects in the lymphatic system facilitate the retention of NPs in the TME, where they slowly release their payload, such as antitumor drugs [206,207]. Several TAM-targeting strategies exploit the EPR effect and/or the intrinsic phagocytotic capacity of TAMs, thus circumventing the need for specific targeting moieties attached to the surface of NPs. In this section, we discuss NP-based passive targeting strategies to TAMs that do not utilize specific targeting motifs.
Liposomes have been exploited in many studies as delivery system for drugs that deplete or modulate TAMs or inhibit their recruitment. For example, Li et al. demonstrated that liposome loaded with C6-ceramide can mitigate the immunotolerant TME of liver cancer by decreasing the frequency of TAMs and the ability of TAMs to suppress the antitumor immune response by modulating ROS signaling and increasing the M1 phenotype [31]. NPs can protect siRNA molecules from rapid degradation in the blood stream. Leuschner et al. used CCR2-siRNA-loaded liposomes of 70–80 nm in size to silence CCR2 mRNA expression, which is required for monocyte recruitment to tumors [208]. After intravenous injection, the liposomes showed a half-life of 8.1 min and accumulated in the spleen and bone marrow, where they delivered siRNA into monocytes expressing high levels of Ly6C, which are precursors of TAMs. This liposome formulation potently decreased the number of TAMs and inhibited tumor growth [208]. Similar, Shen et al. developed monocyte-targeted, siCCR2-loaded, cationic lipid-assisted NPs and reduced the recruitment of monocytes to tumor tissues [209]. As discussed in Section 3.1, STAT3 inhibitors have been shown to promote the M1 phenotype of TAMs. Jose et al. manipulated liposomes to co-deliver curcumin and STAT3 siRNA to treat melanoma. The curcumin/siRNA-loaded liposomes inhibited tumor growth compared to either liposome-curcumin or STAT3 siRNA alone. The antitumor activity of curcumin/siRNA-loaded liposomes was associated with suppressed STAT3 activity and the repolarization of M1 macrophages towards the M1 type [210]. Other drugs, whose therapeutic effect against TAMs has benefited from NPs, are clodronate and bisphosphonates [211]. TAMs play a pivotal role in tumor growth by promoting angiogenesis, and clodronate encapsulated in liposomes (clodrolip) efficiently depleted these phagocytic cells in the murine F9 teratocarcinoma and human A673 rhabdomyosarcoma mouse tumor models, resulting in the significant inhibition of tumor growth [212]. Similar, clodronate encapsulated in liposomes administered into a mouse model of melanoma depleted TAMs and inhibited the monocyte chemoattractant protein 1 (MCP-1) pharmacologically, thereby reducing macrophage recruitment [213]. The bisphosphonate zoledronate has previously been shown to promote the M1 type in macrophages cultured in vitro in a medium conditioned with soluble factors of breast cancer cells [214]. When encapsulated into liposomes, it enhanced levels of the proinflammatory signals iNOS and TNF-α. Rajan et al. showed that PEGylated liposomes encapsulating alendronate efficiently targeted macrophages in the immunocompetent TC-1 murine tumor model, reduced tumor growth and increased disease-free survival [215]. VEGF is a major inducer of tumor angiogenesis and is also required for macrophage recruitment (Table 2). Clodronate encapsulated in liposomes for cancer therapy, in combination with angiogenesis inhibitors, such as anti-VEGF antibodies, has been shown to be more efficient than single-modality treatments, targeting the recruitment hematopoietic precursor cells that stimulate tumor growth [212].
In addition, VEGF and placental growth factor (PIGF), which are overexpressed in M2 TAMs and breast cancer cells, have been shown to work synergistically in mediating tumor progression and immunosuppression. Song et al. developed combinational antitumor immunotherapy using PEG and mannose doubly modified trimethyl chitosan NPs to deliver VEGF and PIGF siRNA to both M2 TAMs and breast cancer cells for antitumor immunotherapy [127]. These pH-sensitive NPs showed prolonged blood circulation times and efficiently accumulated in M2 TAMs and breast cancer cells via mannose-mediated active targeting and passive targeting, respectively. The co-delivery of siVEGF and siPIGF reduced the proliferation of tumor cells and reversed the TME from pro-oncogenic to antitumoral and suppressed breast tumor growth and lung metastasis.
In another approach, PEGylated liposomes functionalized with anti-CD40 and CpG oligonucleotides were developed. CpG oligonucleotides are ligands for Toll-like receptors (TLR) and potent immunostimulatory agents, whereas CD40 ligation triggers a signaling mechanism to promote an antitumor T cell response [216]. Anti-CD40/CpG-liposomes successfully sequestered anti-CD40 and CpG in vivo, significantly inhibited tumor growth and induced a survival benefit in a B16F10 murine model of melanoma (Table 2).
Polymeric NPs have been extensively used to target TAMs. Jung et al. formulated epoxide-modified lipid–polymer hybrid NPs termed 7C1 containing CX3CL1 siRNA to block monocyte recruitment [217]. Their results showed that 7C1-CX3CL1 siRNA significantly reduced the CX3CL1 level in vivo and consequently also reduced Ly6Clo monocyte recruitment. This study revealed an immunosuppressive function of Ly6Clo monocytes (Table 2). Recently, Cavalcante et al. demonstrated that naked or hyaluronic (HA)-coated PeiPLGA NPs encapsulating methotrexate injected intratumorally modified the tumorigenic course of 4T1 breast cancer by modulating the TME, leading to a reduction in primary tumor size and metastases. Interestingly, the application of HA-PeiPLGA NPs promoted the repolarization of M2 macrophages towards the M1 type, reduced the levels of IL-10 and down-regulated the levels of STAT3 and NF-κB [117]. In addition, HA-PeiPLGA NPs reduced the levels of CCL22, IL-10 and TGF-β expression, as well as the recruitment of Tregs and enhanced the cytotoxicity of CD8 T cells in the TME. Another strategy aimed at modifying TAMs in the TME into tumor-suppressive macrophages by local delivery of IL-12. To this end, Wang et al. synthesized microenvironment-responsive poly β-amino ester-based NPs with IL-12 as a payload [218]. The NPs exhibited enhanced tumor accumulation, and significantly extended the circulation half-life and therapeutic efficacy of encapsulated IL-12 compared to free IL-12.
Polymeric chitosan NPs have also been used as an intranasal delivery system for siRNAs targeting galectin-1 expression for the treatment of glioblastoma multiforme (GBM). Galectin-1 is overexpressed in GBM and drives chemo- and immunotherapy resistance. Interestingly, when delivered intranasally and after reaching the TME via nose-to-brain transport, galectin-1 siRNA chitosan NPs silenced galectin-1 in the TME and induced a remarkable switch in the TME composition, including a reduction in macrophage polarization from M1 towards the M2 phenotype during GBM progression [219].
As discussed in Section 3.1, TLR signaling can modulate macrophage polarization. Rodell et al. identified that R848, a TLR7/TLR8 agonist, efficiently induces the M1 phenotype in vitro. Based on this finding, they developed R848-loaded cyclodextrin NPs [220]. After in vivo administration, the NPs accumulated in macrophage-rich tissues over time and demonstrated high polarization efficiency in multiple tumor models. Moreover, R848-loaded cyclodextrin NPs protected mice in tumor rechallenge experiments. Furthermore, the combination of R848-loaded cyclodextrin with an anti-PD-1 checkpoint blockade greatly improved immunotherapy response rates.
Next to liposomes and polymeric NPs, inorganic NPs, which are characterized by their small sizes, large surface areas and various surface modifications, have also attracted attention for TAM-based cancer immunotherapy. For example, Zhao et al. showed that poly(I:C)-loaded amino-modified ferumoxytol NPs could effectively enhance the antitumor immune response, thereby improving cancer immunotherapy and promoting the regression of primary and metastatic melanoma [221]. In vitro studies further showed that ferumoxytol NPs activated the NF-κB signal transduction pathway in macrophages and promoted the M1 phenotype, which prevented tumor cell growth [221].
Multi-walled carbon nanotubes (MWCNTs) belong to the family of nanovectors and have been evaluated as a potential delivery system for brain tumor therapy in the GL261 murine intracranial glioma model. After intratumoral injection, nearly 10–20% of total cells demonstrated MWCNT internalization. Importantly, most MWCNT uptake occurred by TAMs, suggesting that MWCNTs could potentially be used as a non-toxic vehicle for targeting TAMs in brain tumors [222]. Mesoporous silica nanoparticles (MSNs) have been extensively studied for systemic drug delivery. Kwon et al. developed MSNs of 200 nm in size with a larger pore sizer than conventional MSNs, allowing the encapsulation of larger therapeutic biomolecules [223]. XL-MSNs were used to deliver IL-4 in vivo to macrophages and to polarize them towards the anti-inflammatory M2 type (Table 2). Upon intravenous injection, XL-MSNs targeted phagocytic myeloid cells, including macrophages, and successfully triggered M2 macrophage polarization in vivo, demonstrating the clinical potential of XL-MSNs for the modulation of the immune system via targeted delivery of cytokines.
Inhibition of CSF1R signaling has been shown to suppress macrophage infiltration and to reduce tumor growth [224]. Ramesh et al. developed dual-kinase inhibitor-loaded supramolecular nanoparticles (DSNs) for the concurrent and sustained inhibition of the CSF1R and MAPK signaling pathways [225]. When tested in the aggressive 4T1 breast cancer model, the DSNs accumulated in TAMs at a significantly higher concentration, increased M1-like phenotype at a significantly higher proportion and improved antitumor efficacy as compared to a combination of single-inhibitor NPs or small-molecule inhibitors.
Several NPs harbor intrinsic properties that regulate the balance between M1 and M2 macrophages in vitro and in vivo by modulating TAM-related signaling pathways. For example, surface functionalization (-COOH and -NH2) of polystyrene NPs has been shown to impair the expression of the scavenger receptor CD163 and the inhibitor cell surface transmembrane glycoprotein CD200 receptor (CD200R) on M2 macrophages, as well as the release of IL-10, without affecting the expression of M1 markers [226]. Feito et al. used graphene oxide nanosheets (FITC-PEG-graphene oxide) modified with PEG amine functional groups and a fluorescein isothiocyanate label to explore their effect on the polarization of murine peritoneal macrophages [201]. They found an increase in M2 macrophage phenotype after nanosheet treatment, likely promoting the percentage of M2 macrophages for tissue repair [201]. Interestingly, while Au NPs were reported to have intrinsic properties that promote the M2-like phenotype in RAW 264.7 cells in vitro, CaCO3-encapsulated Au NPs induced the expression of the M1 biomarker and inflammatory cytokines [227]. In another study, hydroxyl dendrimers were combined with the CSF-1R inhibitor BLZ945 (D-BLZ) to target TAMs in brain glioblastoma. The authors showed that the conjugation of BLZ945 to dendrimers enabled sustained release in intracellular and intratumor conditions. A single systemic dose of D-BLZ targeted to TAMs decreased the protumorigenic phenotype of TAMs and promoted the infiltration of cytotoxic T cells, resulting in prolonged survival and an ameliorated disease burden compared to free BLZ945 [224]. In summary, targeting NPs to TAMs does not necessarily require the presence of targeting motifs on the surface of NPs. Due to several factors, such as the EPR effect, the positioning of TAMs in the TME, as well as their phagocytotic nature, passive targeting strategies have been shown to lead to the accumulation of NPs in TAMs in vivo. However, these strategies are not TAM-specific and lead to the global uptake of NPs in TME cellular subsets, including immune cells. Thus, whether passive targeting is suitable for the delivery of therapeutics to TAMs must be carefully evaluated.

4.6. Targeting NPs to TAMs via Specific Surface Receptors

The elimination or modification of TAMs by targeting NPs to TAM surface receptors is regarded as a promising method in cancer immunotherapy. To date, NPs encapsulating DNA, vaccines, oligonucleotides and imaging agents to TAMs have been targeted to macrophages via membrane receptors (Table 2). Mannose receptors, scavenger receptors (CD163) and CD200R are highly expressed on the surface of M2 macrophages, and thus represent potent targets for NP delivery to TAMs [47].

4.6.1. CD163

For example, antibody-decorated liposomes loaded with doxorubicin could be specifically targeted to TAMs in vivo after injection into mice via the hemoglobin scavenger receptor CD163 [228]. Specific depletion of CD163+ TAMs resulted in a massive infiltration of activated T cells, the mobilization of inflammatory monocytes and tumor regression.

4.6.2. Mannose Receptor

In addition, carbohydrates, such as mannose, galactose and glucomannan, can also be employed as targeting motifs. Mannose has been shown to continuously promote the internalization of mannosylated molecules [100]. Therefore, decorating NPs with mannose to target the mannose receptor for the selective delivery of therapeutics is a common method in the field of drug delivery. Experiments have shown that primary macrophages took up 4-fold more siRNA when pH-responsive polymeric micelles encapsulating siRNA were mannosylated compared to naked NPs [229]. Several polymeric NPs, including PLGA, have been decorated with mannose to target the mannose receptor on TAMs. Shi et al. formulated mannose-decorated PLGA NPs to co-deliver ICG, ammoniumhydrogencarbonate (NH4HCO3) and titanium dioxide (TiO2) to tumors [230]. Upon an external laser trigger, the NPs promoted the formation of reactive oxygen species, which led to the reprogramming of TAMs towards the antitumor M1 phenotype in vitro and in vivo. Zhu et al. decorated acid-sensitive PEG-modified PLGA NPs with mannose and examined their uptake by TAMs in vivo [231]. The authors demonstrated that the NPs efficiently targeted TAMs via mannose–mannose receptor interactions after acid-sensitive PEG shedding in the acidic TME, while their uptake by normal macrophages in the spleen and liver was significantly reduced due to effective PEG shielding at neutral pH. In another example, researchers designed cationic nanohydrogel NPs containing α-mannosyl functional groups and CSF-1R siRNA to specifically target TAMs through the mannose receptor [232]. SiRNA-loaded mannosylated NPs preferentially targeted mannose receptors on M2 macrophages and effectively reduced expression of CSF-1R. The targeted immunomodulatory response only occurred in M2 macrophages and did not affect the expression state in M1 macrophages [232]. Research by Niu et al. showed that when sufficient TAMs are present in the tumor, PEGylated, mannose-modified PLGA-DOX NPs could effectively deliver DOX to the tumor site, thereby inhibiting tumor growth and reducing the number of TAMs in TME [233]. In another example, hybrid NPs made of Pluronic® -F127 polymer and tannic acid with varying mannose densities showed superior uptake behavior in M2-polarized U937 macrophages in vitro [234]. Metformin (Met), a popular drug used to treat diabetes, also exhibits anticancer activity. Recently, Met has been found to efficiently repolarize M2-like TAMs towards the M1 phenotype to inhibit tumor growth and metastasis through the AMPK-NF-κB signaling pathway [235]. Wei et al. developed mannose-modified macrophage-derived microparticles (Man MPs) encapsulating metformin [236]. Met@Man-MPs efficiently targeted M2-like TAMs and repolarized them into the M1-like phenotype. In addition, Met@Man-MPs increased the recruitment of CD8+ T cells to the TME and decreased the immunosuppressive infiltration of myeloid-derived suppressor cells and Tregs.
Wang et al. developed bioresponsive polymeric NPs (P3AB) that could selectively eliminate TAMs. The NPs consisted of PEG, PLGA and a MMP-cleavable peptide that formed the shell, while a bisphosphonate–glucomannan conjugate with affinity for macrophage mannose receptors formed the core. While in this strategy the NPs were not targeted specifically to TAMs, this strategy utilized MMPs overexpressed in TAMs to cleave the NP shell to expose the bisphosphonate–glucomannan conjugates in order to eliminate TAMs. Systemically administered P3AB polymeric complexes were taken up by TAMs and led to a potent reduction in TAM viability [237], effectively inhibiting tumor growth and restoring local immunosurveillance in vivo.

4.6.3. C-Type Lectin

TAMs express high levels of macrophage galactose-specific C-type lectin (MGL) [238]. Decorating NPs with galactose moieties can facilitate the recognition and uptake of NPs by TAMs. Han et al. developed baicalin (a flavone glycoside)-loaded PLGA NPs containing an antigenic peptide (Hgp 10025–33, Hgp) and a Toll-like receptor 9 agonist (CpG), coated with a galactose-inserted erythrocyte membrane, which could actively target TAMs and showed enhanced cell uptake in vitro and targeting in vivo [239] (Table 2). In addition, the biomimetic NPs containing baicalin, Hgp and CpG, or baicalin- and galactose-inserted erythrocyte membranes significantly repolarized the TAMs from the M2-like type towards the M1-like type, both in vitro and in vivo, which led to a significant enhancement of the antitumor T cell response. Another approach focused on repolarizing TAMs to antitumor M1 macrophages by administration of redox/pH dual-responsive nanovectors encapsulating with PEG-PLL (sPEG) copolymers/galactose-functionalized n-butylamine-poly(l-lysine)-b-poly(l-cysteine) polypeptides (GLC) for delivery of miR155. miR155 had previously been shown to significantly attenuate the cytokine production in TAMs via targeting C/EBPβ [240] and repolarized protumoral [240] and re-polarized pro-tumoral M2 TAMs into antitumor M1 macrophages in vitro [241]. The results showed that administration of sPEG/GLC/155 nanocomplexes increased miR155 expression in TAMs about 100–400-fold both in vitro and in vivo, and efficiently repolarized immunosuppressive TAMs towards antitumor M1 macrophages by elevating M1 macrophage markers (IL-12, iNOS and MHC II) and suppressing M2 macrophage markers (Msr2 and Arg1) in TAMs [242]. Huang et al. developed PEG-histidine-modified alginate (PHA) NPs encapsulating galactosylated cationic dextran with nucleic acids. Using this system, the authors combined CpG oligonucleotides, anti-IL-10 and anti-IL-10 receptor oligonucleotides [243]. After intravenous injection in an allograft hepatoma murine model, the acidic environment of the TME triggered an alteration in the charge of PHA from negative to positive, which led to the dissociation of PHA from the complex and exposed the galactosylated cationic dextran–ODN complex, which led to the predominant uptake of NPs in TAMs via MGL. As a consequence, the authors observed suppression of protumor functions and stimulation of the antitumor activities of TAMs [243].
Moreover, self-assembled glyco-NPs, composed of hydrophobic block polystyrene (PS) and a hydrophilic glycoblock have been shown to successfully induce the polarization of mouse primary peritoneal macrophages from M2 to the inflammatory type M1 in vitro and in vivo [244]. The glyco-NPs were coated with three sugars (d-galactopyranoside, α-mannopyranoside and l-fucopyranoside), which served as ligands for C-type lectin receptors on macrophages. To increase the selectivity of M2 TAM targeting, Cieslewicz et al. screened a peptide phage library and identified M2pep as a promising alternative to the traditional ligands [245]. When conjugated to a proapoptotic peptide, it could induce toxic effects in M2 macrophages. Pang et al. modified the surface of PLGA NPs with M2pep and encapsulated the CSF-1/CSF-1R pathway inhibitor, PLX3397 [246]. The results showed that M2pep conjugation increased the uptake of PLGA NPs by M2-polarized macrophages, improved the antitumor efficacy of PLX3397 and helped to attenuate tumor growth in the mouse B16F10 melanoma model.

4.6.4. Scavenger Receptor B Type 1

In another approach, Qian et al. developed lipid NPs specifically targeting M2-like TAMs by including the α-peptide (a scavenger receptor B type 1 (SR-B1)-targeting peptide) linked to M2pep [126]. The lipid NPs were loaded with CSF-1R-siRNA to specifically block the survival signal of M2-like TAMs and to deplete them from melanoma tumors. The results showed a dramatic elimination of M2-like TAMs, decreased tumor size and prolonged survival. In addition, this targeting strategy overall reduced the immunosuppressive state of the TME by reducing expression of IL-10 and TGF-β, while increasing the immunostimulatory cytokines IL-12 and IFN-γ. Furthermore, siRNA-carrying lipid NPs increased CD8+ T cell infiltration (2.9-fold) and down-regulated expression of the exhaustion markers PD-1 and Tim-3.

4.6.5. Sialic Acid Binding Receptors

Sialic acid binding receptors (siglecs) are overexpressed on the surface of TAMs and thus can be utilized for TAM-specific targeting by decorating NPs with sialic acid. For example, epirubicin-loaded liposomes were decorated with sialic acid–cholesterol conjugates to improve the delivery of epirubicin to TAMs [247].

4.6.6. Legumain

In an alternative approach, hydrazinocurcumin-encapsulated liposomes, a synthetic analog of curcumin with improved water solubility, cell permeability and bioavailability, were used to target legumain, a protease specifically expressed on TAMs. Hydrazinocurcumin was used to inhibit STAT3 activity and to re-educate TAMs. Furthermore, hydrazinocurcumin down-regulated a series of immunosuppression-related proteins and downstream target molecules of STAT3, thereby limiting the proliferation and migration of 4T1 breast cancer cells in vitro, as well as tumor growth, angiogenesis and metastasis in vivo [248].
Instead of focusing on specific receptors, cell-membrane-coated NPs have emerged as a promising antitumor therapeutic strategy, due to their enhanced blood circulation, immune compatibility and tissue-targeting capacities [249]. Recently, this strategy has been expanded to TAMs. TAM membrane (TAMM) derived from 4T1 primary breast tumors was isolated and coated onto the surface of upconversion NPs (UCNPs) made of rare-earth metals (namely, NaYF4:Yb, Er@NaYF4). The TAMM was rich in CSF1-R, which could be exploited to capture the CSF1 secreted by tumor cells in the TME upon injection of UCNPs into tumor-bearing mice. This blocked the interaction between TAMs and cancer cells and the activation of downstream signaling pathways responsible for the polarization of TAMs to the immunosuppressive phenotype. In particular, after coating the TAMM-UCNPs with photosensitizer and applying photodynamic therapy, macrophages in the TME could be polarized from an immunosuppressive M2-like phenotype to a more inflammatory M1-like state, while the PDT effect induced immunogenic cell death. Overall, this dual approach enhanced the antitumor immune response by the activation of antigen-presenting cells and production of tumor-specific effector T cells in metastatic tumors. Natural NPs, such as exosomes, have been shown to transfer cargo between neighboring cells of the tumor microenvironment, and have been implicated in promoting tumor progression. Several non-coding RNA molecules, such as microRNAs, were shown to be present in these vesicles [250]. Trivedi et al. manipulated the miRNA content of secreted exosomes by the genetic transfection of tumor cells using dual-targeted hyaluronic-acid-based nanoparticles encapsulating plasmid DNA encoding for wild-type p53 (wt p53) and microRNA-125b [251]. These altered miRNA levels in the exosomes mediated macrophage repolarization towards a more proinflammatory/antitumor M1 phenotype.
Table 2. Overview of NP-based TAM-targeting strategies in cancer therapy.
Table 2. Overview of NP-based TAM-targeting strategies in cancer therapy.
CategoryNP MaterialsPayloadTargeted ReceptorTAM-Targeting StrategyRef.
Polymeric NPsHyaluronic-acid-coated Pei-PLGA NPsMethotrexate (MTX)CD44Reprogramming TAMs[117]
Chitosan NPsGal-1 siRNAPassive targetingModulation of TME[219]
PEG-sheddable, mannose-modified PLGA NPs Mannose receptorTargeting TAMs[231]
Cationic polymeric NPssiCCR2Passive targetingInhibition of monocyte recruitment[209]
Carboxyl- and amino-functionalized polystyrene NPs Passive targetingReprogramming TAMs[226]
PEGylated silk fibroin NPsSilk and fibroinPassive targetingReprogramming of TAMs[207,251]
Hyaluronic acid–PEG blend NPsp53 plasmidPassive targetingReprogramming of TAMs[251]
MMP-responsive PLGA NPsBisphosphonate-glucomannanMannose receptorInterference with TAMs survival[237]
Mannose-modified polymeric micellessiRNAMannose receptorDepletion of TAMs[229]
PLGA NPsRadachlorinPassive targetingDepletion of TAMs[252]
Mannose-modified polymeric NPssiRNAs to modulate NF-κB signalingMannose receptorReprogramming of TAMs[253]
PEG and mannose doubly modified trimethyl chitosan NPsVEGF siRNA/PIGF siRNAMannose receptor and passive targetingReprogramming TAMs[127]
Nanovectors made of galactose-functionalized n-butylamine-poly(l-lysine)-b-poly(l-cysteine) polypeptides coated with DCA-grafted sheddable PEG-PLL copolymersmiR155C-type lectin (MGL) receptorReprogramming TAMs[242]
Mannose/acid-sensitive
PEG-modified PLGA NPs
Doxorubicin (DOX)Mannose receptorDepletion of TAMs[233]
M2pep-coated PLGA NPsPLX3397Passive targetingDepletion of TAMs[246]
Biomimetic NPsCpG, baicalin and antigenic peptidePassive targetingReprogramming TAMs[239]
Microenvironment-responsive polymeric (P1) NPsIL-12Passive targetingReprogramming TAMs[218]
Mannan-coated PLGA NPsDidanosineMannose receptorDepletion of TAMs[206]
Mannose-modified PLGA NPsICG, NH4HCO3 and titanium dioxideMannose receptorReprogramming of TAMs[230]
miR155-loaded sPEG/GLC (sPEG/GLC/155) NPsmiR155Passive targetingReprogramming TAMs[242]
Arginine NPsCas9 and gRNAPassive targetingIncreased phagocytosis of tumor cells by macrophages[254,255]
PEG-histidine-modified alginate NPsOligodeoxynucleotide and galactosylated cationic dextranMacrophage galactose-type lectin (Mgl)Reprogramming TAMs[243]
Cyclodextrin NPsR848Passive targetingReprogramming of TAMs[220]
Lipid-based NPsLiposomesC6-ceramide (LipC6)Passive targetingDepletion and reprogramming of TAMs[31]
MMP2-sensitive apoptotic body-mimicking NPs and phosphatidylserine-modified NPsPhosphatidylserineMMP2-sensitiveDepletion of TAMs[256]
Cationic liposomesCurcumin and anti-STAT3 siRNAPassive targetingReprogramming of TAMs[210]
Sialic acid–cholesterol conjugate-modified liposomesEpirubicin and sialic acidSiglec receptorsDepletion of TAMs[247]
LiposomesCCR2 siRNAPassive targetingDepletion of TAMs[208]
LiposomesHydrazinocurcuminLegumainReprogramming of TAMs[248]
Clodronate liposomesClodrolipPassive targetingDepletion of TAMs[212,213]
LiposomesZoledronatePassive targetingDepletion of TAMs[214]
PEGylated liposomesAlendronatePassive targetingReprogramming of TAMs[215]
Long-circulating liposomesSimvastatinPassive targetingInterference with TAM survival[257]
PEGylated liposomesAnti-CD40 and CpG oligonucleotidesCD40Reprogramming of TAMs[216]
Supramolecular lipid NPsDual-kinase (MEK and CSF1R) inhibitorPassive targetingReprogramming of TAMs[225]
Sorafenib-loaded lipid NPs (SLNPs)SorafenibPassive targetingRelieve of the immunosuppressive TME[258]
α-peptide- and
M2pep-linked lipid NPs
Anti-CSF-1R
siRNA
Scavenger receptor B type 1 (SR-B1) and M2 macrophage binding peptideDepletion of TAMs[126]
Clodronate-loaded liposomesBindarit and clodronatePassive targetingInhibition of macrophage recruitment[213]
Inorganic NPsMesoporous silica NPsIL-4Passive targetingReprogramming of TAMs[223]
Fe2O3 NPTLR3 agonist, poly (I:C)Passive targetingReprogramming of TAMs[221]
Fe3O4/PLGA NPsAnti-CD206CD206Reprogramming of TAM[149]
Ca-CO3/Au-NPs Passive targetingReprogramming of TAMs[227]
Au@SiO2 NPs (GNPs)Anti-CD163CD163Depletion of TAMs[259]
Stöber silica NPs Passive targetingSequestration in M1 TAMs[200]
Peptide/hyaluronic acid/protamine/CaCO3/DNA-NPspDNA and IL-12CD44Reprogramming of TAMs[260]
Hyaluronic acid-manganese dioxide (MnO2) NPsDoxorubicinCD44Reprogramming of TAMs[261]
αvβ3-integrin-targeting (αvβ3-antagonist, a quinalone nonpeptide)-perfluorocarbon NPsMYC inhibitorVitronectin receptor αvβ3Depletion and reprogramming of TAMs[262]
MnO2 NPsSorafeniPassive targetingReprogramming of TAMs[263]
Iron oxide NPsTrastuzumab, hIgGPassive targeting and active targetingDepletion of TAMs[264]
Zinc oxide (ZnO) NPsDoxorubicin (Dox)Passive targetingReprogramming of TAMs[265]
Mesoporous silica-coated
white-light emitting
NaYbF4:Tm@NaYF4:Yb/Er upconversion NPs
Roussin’s black salt and doxorubicinPassive targetingInterference with TAMs survival[266]
NaYF4: Yb, Er@NaYF4 rare-earth upconversion-NPs conjugated with rose bengal (NPR)Conjugated photosensitizerPassive targetingReprogramming of TAMs[249]
Peptide/hyaluronic acid/protamine/CaCO3/DNA-NPspDNA and IL-12CD44Reprogramming of TAMs[260,265]
Ca-CO3/Au-NPs Passive targetingReprogramming of TAMs[227]
Calcium-biphosphonate-PEG NPsChelator-free radiolabelingPassive targetingDepletion of TAMs[211]
Porous silicon multistage nanovectors (MSV)Albumin-bound paclitaxelPassive targetingReprogramming of TAMs[267]
MUC-1-modified AuNPsMucin-1 peptidePassive targetingReprogramming of TAMs[268]
Carbon NPsMulti-walled carbon nantotubesFluorescent dyePassive targetingReprogramming of TAMs[222,269]
Single-walled carbon nanotubes Passive targetingDepletion of TAMs[270]
Graphene oxide NPsFITC and PEGPassive targetingReprogramming of TAMs[201]
PEG and polyethylenimine (PEI) dual-polymer-functionalized graphene oxide NPsCpGPassive targetingReprogramming of TAMs[271]
PEG and polyethylenimine (PEI) dual-polymer-functionalized graphene oxide NPsCpGPassive targetingReprogramming of TAMs[271]
Polyhydroxylated fullerenols Passive targetingReprogramming of TAMs[272]
Hybrid NPsLipid polymer (7C1) NPCX3CL1 siRNAPassive targetingInhibition of macrophage recruitment[217]
Mannosylated cationic nano hydrogel particles (ManNP)siRNAMannose receptorDepletion of TAMs[232]
Mannose-decorated Pluronic®-F127 polymer and tannic acid NPsF127-TAMannose receptorTargeting TAMs[234]
Block copolymers with a hydrophobic block polystyrene (PS) and a hydrophilic glycoblock (PR) self-assembled into glyco-NPs with an inert organic core PS and a glycopolymer shell Passive targetingReprogramming of TAMs[244]
Legumain-sheddable PEG5k and tuftsin dual-modified NPsPEG5k and tuftsinFc receptorDepletion of TAMs[273]
OtherHydroxyl dendrimerCSF-1R inhibitor BLZ945Passive targetingReprogramming of TAMs[224]

5. NPs as a Delivery Platform for Gene-Editing Toolsdelivery Platform for Gene Editing Tools to Macrophagesmacrophages

The development of gene-editing tools, such as CRISPR, has opened novel avenues for the treatment of cancer. Theoretically, this revolutionary technology could bring novel therapeutic modalities to many diseases, including cancer, by precisely manipulating cellular DNA sequences. However, to date, the low efficiency of in vivo delivery and gene editing, as well as the lack of specificity towards target cells and/or tissues, must be enhanced before its therapeutic potential can be fully realized. NP-based CRISPR/Cas9 delivery systems represent a novel technology in which NPs instead of viruses are employed to directly deliver the CRISPR/Cas9 complex, or plasmids expressing Cas9 and guide RNAs, to the nuclei of targeted cells. In recent years, many NP formulations have been described as suitable non-viral delivery systems for CRISPR/Cas9, including lipid, polymeric and inorganic NPs [274]. Several molecules, such as CD47 (“do not eat me” signal), which are implicated in tumor immune escape mechanisms and can be targeted by small-molecule drugs [4], have now become targets of gene-editing approaches utilizing NPs as a delivery system for CRISPR/Cas9. Ray et al. described an integrated nanotechnology immunotherapy approach that uses arginine NPs (Arg NPs) to deliver the CRISPR/Cas9 complex into cells to generate SIRP-α knockout macrophages [254]. Knockout of SIRP-α prevented CD47:SIRP-α interactions between cancer cells and macrophages, and increased the phagocytosis of tumor cells by macrophages 4-fold. Lee et al. fabricated nanocomposites consisting of engineered Cas9 proteins that included a 20-glutamic acid tag (Cas9E20)—a strategy that has been shown to lead to the formation of hierarchical nanocomposites with NPs featuring arginine head groups (Arg NPs) through the carboxylate–guanidinium interaction [255]. The authors complexed Arg NPs, Cas9E20 and sgRNAs targeting the PTEN gene to formulate a single nanocomposite encapsulating the ribonucleoprotein complex. Systemic injection of this complex provided efficient (>8% and >4%) in vivo gene editing of the PTEN gene, specifically in macrophages of the liver and spleen, respectively. In summary, NPs as platform for gene-editing tools that specifically target TAMs open new avenues for selective immunotherapy for the treatment of cancer.

6. Comparative Analysis of TAM-Targeting Strategies Targeting Strategies

As discussed in Section 4.6, active targeting strategies have been widely employed to improve NPs drug delivery to TAMs. However, drug delivery to TAMs can also be achieved without linking targeting motifs to the surface of NPs [252]. In the following paragraph we will discuss advantages and disadvantages of active and passive NP-NPs targeting strategies that need to be considered when designing NPs-based therapeutics targeting TAMs.

6.1. Active Versus Passive Targetingversus Passive Targeting of TAMs by NPs: Advantages and Disadvantages

In contrast to the notion that targeting increases tumor localization, several studies in mice have shown that linking targeting motifs to the surface of NPs does not affect NP biodistribution or increase NP accumulation in the TME [243,275,276,277]. Rather, the dominant factors determining the accumulation of NPs in the TME are the leakiness of the vasculature and the presence of lymphatic drainage. In contrast to most normal tissues, which are characterized by tight endothelial linings that prevent the penetration of macromolecules and NPs, murine tumors possess enlarged endothelial gaps that facilitate the preferential accumulation of NPs at the tumor site. We and others have recently shown that systemically applied PEGylated PLGA NPs were able to localize to the TME in mouse models of colorectal cancer, and could be detected in different cellular subsets, in particular, myeloid cells, such as CD45+ CD11b+ F4/80+ macrophages [278]. While the importance and the existence of the EPR effect in humans is heavily debated, accumulating evidence suggests that it also occurs in human patients, but that it greatly varies between patients and type of tumor [279]; thus, tumor-targeting approaches beyond the EPR effect are equally important. Recently, it became clear that the way that NPs enter solid tumors is more complex than previously thought, and that immune cells in the TME play important roles in NP accumulation, retention and intratumoral distribution. For example, the intratumor retention of antibody-labeled NPs was shown to be determined mainly by tumor-associated dendritic cells, neutrophils, monocytes and macrophages, and not by antibody–antigen interaction [264].
Based on their phenotype, mobility and localization in the TME, different types of TAMs have been identified both in murine and human tumors. For instance, a subset of M2-like perivascular (PV) macrophages resides proximal to the abluminal surface of blood vessels and is highly efficient in sampling particular substances after intravenous injection [280]. Intravital imaging studies have shown that PV macrophages interact closely with the endothelial cells and actively remodel the permeability of neighboring blood vessels by the expression of VEGFA [281]. Due to their proximity to blood vessels and their phagocytotic nature, PV TAMs are ideally positioned to take up NPs, even in the absence of targeting motifs on the surface of NPs. Compared to non-targeted NPs, a number of active targeting approaches have resulted in decreased blood circulation times due to opsonization. For example, Salvati et al. found that transferrin-conjugated NPs placed in a biological environment were immediately coated with a protein corona, which shielded transferrin on the surface of NPs, resulting in a marked reduction in binding specificity [174]. Thus, the global delivery of drugs to the TME and TAMs can be efficiently achieved by passive targeting. This strategy might in particular be suited for the delivery of membrane- permeable drugs, such as cyclodextrins, which upon NPs accumulation in the TME diffuse from the core of NPs and are taken up by surrounding cells [282].
As outlined above, passive targeting circumvents some drawbacks of targeted NPs approaches, however, the delivery of NPs to target cells by the recognition of surface receptors also shows clear advantages. First, specific payload delivery to target cells reduces drug toxicity and keeps adverse immune reactions at minimum. Second, active targeting strategies have demonstrated increased NPs internalization inside target cells and intracellular drug delivery, compared to non-targeted NPs. Thus, the active targeting approach is in particular suited for the cell-specific delivery of membrane- impermeable drugs, and/or co-delivery of compounds, such as guide RNA and Cas9 [283], to specific cell types in vivo. For example, DSPE-PEG-EGFR-siRNA NPs were coated with the ligand anisamide to target sigma receptors and administered intravenouslyintravenous in mice bearing a sigma receptor expressing NCI-H460 xenograft tumors [275]. While the targeting ligand did not influence the distribution of the NPs, it increased the intratumoral localization of the siRNA and silencing of EGFR expression. The targeted NPs formulation efficiently delivered siRNA into the cytoplasm of NCI-H460 tumor cells, whereas free siRNA and non-targeted NPs showed little uptake. Third, targeting specific cell types instead of global administration might overcome limitations of some EPR-based therapies. For instance, passive drug delivery strategies based on the EPR effect are only applicable to solid tumors above a certain size [284], whereas active targeting of cell populations provides the possibility of drug delivery to a variety of tumor and non-tumor tissues. Fourth, targeting NPs via surface receptors to myeloid cells has been shown to increase adjuvanticity and immunogenicity in immunotherapy [14,18,285,286].
Active targeting provides the possibility to target TAM subsets. In this context, it has been shown that spatially distinct TAM subsets with different phenotypes and behavior co-exist in the TME. The spatial heterogeneity is most likely as an adaptation to the different tumor compartments, such as the tumor nest, stroma and PV niche, and TAM subsets show differential expression of cell surface (such as angiopoietin-1 receptor, MR, CD163, TLR4, PD-L1. CD100, CD80 and CD86) and intracellular/secreted proteins (cathepsins, urokinase-type plasminogen activator, nitric oxide synthase 2, pro-MMP7, VEGFA, pro-MMP10 and IL-10). Evidence is emerging that TAM subsets in certain tumor areas limit tumor responses to treatment. For example, PV TAMs have been implicated in contributing to relapse after chemotherapy, and hypoxic TAMs in the tumor resistance to anticancer therapy [287]. Thus, active targeting by NPs could provide the tool to deplete or re-educate those TAM subsets that are tumor-promoting, while keeping those TAMs that are tumoricidal and/or promote antitumor immunity.

6.2. Comparison of NP-Based TAM-Targeting Therapies and Current Challenges

Current NP-based TAM-targeting strategies have shown great potential in preclinical and clinical cancer research, and can greatly improve therapeutic effects while reducing systemic side effects [288]. NPs that treat TAMs in cancer therapy have been designed to inhibit TAM recruitment or accumulation, or to re-educate TAMs to restore functional phagocytosis and M1 phenotype. In the following paragraph we compare the most common TAM-targeting approaches utilizing NPs and discuss associated challenges. TAM depletion is often achieved by administration of drug-loaded NPs (Table 2); however, administration of non-targeted drug-loaded NPs can evoke immune toxicity through undesired NP interaction with immune cells; thus, in-depth toxicity studies must be carried out to carefully evaluate potential off-target effects. For example, acute toxicity has been observed when administering liposomal zoledronic acid, most likely due to increased cytokine production [289]. In addition, NPs with immunomodulatory functions should be designed to target specific cells to limit off-target effects and to provide a sustained antitumor effect in patients. The long-term consequences of TAM depletion are still unclear; therefore, it might not be desired to eliminate both proinflammatory and anti-inflammatory types of TAMs. Rather, subset-specific targeting ligands are needed to deplete or re-educate those that are tumor-promoting, while leaving or increasing those that promote antitumor immunity. Another drawback of this method is that it is likely some TAMs will remain after treatment which could drive tumor relapse [43]. Mannose and galactose are often used as a targeting ligand for NP-mediated drug delivery to TAMs (Table 2); however, mannose/galactose receptors are present on both M1- and M2-type TAMs. In addition, TAMs often display intermediate states (in-between M1 and M2 phenotypes) and other cell types may display these receptors as well [290,291]. Therefore, targeting specificity remains a key challenge. Thus, ligands that are more specific to M2 TAMs, such as M2pep [245,246], might represent better targeting moieties for NP-mediated drug delivery to TAMs. A lack of specificity might lead to the undesired targeting of tissue-resident macrophages, such as red pulp macrophages in the spleen or Kupffer cells in the liver. Another strategy focuses on the depletion of TAMs or inhibition of TAM recruitment, which is often achieved by the delivery of molecules (e.g., siRNA) that inhibit trafficking receptors, such as CCR2. However, this strategy does not take into account tissue-resident macrophages that do not seem to be recruited in the same way as monocyte-derived macrophages. Thus, strategies that can differentially block the recruitment and proliferation of tissue-resident macrophages are needed. Finally, most NP formulations constructed to treat TAMs focus on the modulation of the polarization status of TAMs towards an antitumorigenic phenotype. However, more research is needed to evaluate the long-term effect of this approach. Excessive macrophage activation might lead to substantial cell-mediated toxicity, such as macrophage activation syndrome and hemophagocytic syndrome.
Therapeutic approaches to deplete or reprogram TAMs, including inhibitors targeting the CSF-1-CSF-1R and the CCL-CCR2 axes, have largely failed to show efficacy in cancer clinical trials as monotherapies [292,293]. One explanation could be that the inhibition of one component might be compensated by another. For this reason, multimodal treatment approaches are expected to improve the outcome of immunotherapies in different types of cancer. NPs carrying different drugs for combination therapy have been shown to improve the therapeutic efficacy and to overcome multidrug resistance [252,294]. For example, a humanized anti-SEMA4D antibody (pepinemab) is currently being tested in multiple clinical trials for various cancers (NCT03320330, NCT03690986 and NCT03769155), in combination with PD1/PDL1 checkpoint inhibitors [295]. M2 -TAMs produce SEMA4D, which among other effects has been reported to inhibit immune cell movement and to cause vascular destabilization. Combinatorial treatment strategies targeting these two pathways in a synergistic manner are expected to have a greater impact on the inhibition of tumor growth, than PD1/PDL1 checkpoint inhibitors or anti- SEMA4D monotherapies. The combination of TAM-targeted drugs with chemotherapy could also improve the therapeutic effect of chemotherapeutic drugs, which when administered alone can increase the infiltration of TAMs into tumor tissues [258,296]. Furthermore, the combination of M2- TAM-targeting by Au@SiO2 (GNPs) NPs conjugated with a CD163 antibody with radiation has resulted in more effective tumor growth delay than radiation alone [259]. Synergistic effects were also observed when gadofullerene NPs were combined with anti-PD-L1 immune checkpoint inhibition [259]. Gadofullerene NPs were shown to reprogram TAMs to a M1-like phenotype and increase the infiltration of cytotoxic T lymphocytes. Combinational treatment with anti-PD-L1 achieved effective inhibition of tumor growth in the 4T1 murine breast cancer model in vivo.

7. Conclusions and Future Outlook

As described in this review, the TME plays an important role in tumorigenesis. As one of the main cell populations of tumor-infiltrating immune cells, macrophages play a central role in homeostasis in inflammatory diseases and immunotherapy.
More recently, TAMs have emerged as a target of NPs-mediated cancer immunotherapy. The internalization mechanism of targeted NPs in macrophages and their capacity to repolarize the TAMs phenotype of TAMs in cancer treatment have received increasing attention. More and more evidence shows that eliminating or manipulating TAMs is beneficial for cancer treatment. The inhibition of macrophage CSF1R activity or the enhancement of STAT3 activation in the TME can promote the repolarization of M2 macrophages to a phenotype similar to M1 to overcome the immunosuppressive state that promotes tumor growth and metastasis. However, pan-macrophage therapy targeting all macrophages will cause systemic toxicity. Therefore, immunotherapy strategies specifically targeting immunosuppressive TAMs in the TME by exploiting TAM-specific cell surface receptors, combined with immunotherapy, provide new opportunities for tumor diagnosis and treatment.
Because certain characteristics of in vitro and in vivo macrophages are not completely consistent, multi-level verification in animal models is required, as outlined in this review. In vivo, NPs can be passively targeted through the EPR effect or actively targeted to tumors through conjugated ligands for specific drug delivery. The physicochemical properties of NPs (size, shape, surface charge and, targeting moieties) affect NPs uptake in different TAM subsets and consequently determine the NP efficacy. Thus, NPs preparation and the use of targeting moieties are important points to consider when designing NPs to target TAMs. Additionally, the choice of payload, such as drug and adjuvants, will require thorough consideration when designing novel NP-based treatments with low systemic toxicity. Thus, to design more efficient delivery systems that specifically target TAMs and to improve the concentration of drugs at the tumor site in the course of immunotherapy.
To overcome some limitations of current TAM-targeting NP formulations, such as off-target effects, stimuli-responsive NPs that selectively release therapeutic agents in target tissues or cells have been designed. Such NP systems with spatio-temporal release kinetics can release therapeutics “on-demand” in a controlled fashion, and are often combined with molecular imaging to monitor the biodistribution or activation of these systems. Examples of such a system are NPs that are built from materials that “respond” to intrinsic stimuli, such as pH, enzymes and reducing agents, as well as extrinsic stimuli such as heat, light, electric fields, ultrasound and magnetic fields [297]. For example, Liang and coworkers recently developed a novel dually functionalized NP platform by the surface co-modification of NPs with tuftsin, a natural activator of phago-cytosis, and legumain-protease-sheddable PEG to achieve selective targeted delivery to TAMs [273]. In another example, MMP2-sensitive phosphatidylserine-modified NPs loaded with the model drug dasatinib were developed [256]. In this design, the phosphatidylserine was externalized to the NPs’ surface only when the NPs reached the MMP2-overexpressing TME, allowing for the TAM-specific phagocytosis. The NPs showed excellent specificity towards TAMs in various biological models. Importantly, the specificity of NPs towards TAMs remarkably enhanced TAM depletion. Overall, such novel NPs platforms with an “on-demand” release kinetics design could greatly improve the selectivity and efficacy of targeted NP therapeutics towards TAMs.

Author Contributions

Y.H., C.E., R.F.d.A.J. and L.J.C. structured the review. Y.H. and C.E. wrote the review. C.E., R.F.d.A.J. and L.J.C. revised the text. All authors have read and agreed to the published version of the manuscript.

Funding

L.J.C. was supported by project grants from the European Commission: H2020-MSCA-RISE (644373—PRISAR), H2020-MSCA-RISE (777682—CANCER), H2020-WIDESPREAD-05-2017-Twinning (807281—ACORN), H2020-WIDESPREAD-2018-03 (852985—SIMICA), H2020-SCA-RISE-2016 (734684—CHARMED), MSCA-ITN-2015-ETN (675743-ISPIC), 861190 (PAVE), 857894 (CAST), 859908 (NOVA-MRI) and 872860 (PRISAR2). R.F.A.J. was supported by project grants from the European Commission under grant agreement 872391 (CONCRETE) and by CAPES [grant number 88881.119850)/CNPq (grant number 301877/2019-0). C.E. was supported by the research program VENI with the project number 916.181.54, which is (partly) financed by the Dutch Research Council, the H2020-WIDESPREAD-2018-03 (852985—SIMICA) project grant from the European Commission and the Dutch PPS allowance made available by Health~Holland, Top Sector Life Sciences & Health for the project NANOCAST 2. Y.H. received funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska Curie grant agreement, no. 777682 (CANCER) and 734684 (CHARMED).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

No conflict of interest exists in the submission of this manuscript, and the manuscript is approved by all authors for publication.

References

  1. Urruticoechea, A.; Alemany, R.; Balart, J.; Villanueva, A.; Vinals, F.; Capella, G. Recent Advances in Cancer Therapy: An Overview. Curr. Pharm. Des. 2010, 16, 3–10. [Google Scholar] [CrossRef] [PubMed]
  2. Khalil, D.N.; Smith, E.L.; Brentjens, R.J.; Wolchok, J.D. The future of cancer treatment: Immunomodulation, CARs and combination immunotherapy. Nat. Rev. Clin. Oncol. 2016, 13, 273–290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Miller, K.D.; Nogueira, L.; Mariotto, A.B.; Rowland, J.H.; Yabroff, K.R.; Alfano, C.M.; Jemal, A.; Kramer, J.L.; Siegel, R.L. Cancer treatment and survivorship statistics, 2019. CA Cancer J. Clin. 2019, 69, 363–385. [Google Scholar] [CrossRef] [Green Version]
  4. Zhou, J.; Tang, Z.; Gao, S.; Li, C.; Feng, Y.; Zhou, X. Tumor-Associated Macrophages: Recent Insights and Therapies. Front. Oncol. 2020, 10, 188. [Google Scholar] [CrossRef] [PubMed]
  5. DeNardo, D.G.; Ruffell, B. Macrophages as regulators of tumour immunity and immunotherapy. Nat. Rev. Immunol. 2019, 19, 369–382. [Google Scholar] [CrossRef]
  6. Yang, Y.; Guo, J.; Huang, L. Tackling TAMs for Cancer Immunotherapy: It’s Nano Time. Trends Pharmacol. Sci. 2020, 41, 701–714. [Google Scholar] [CrossRef]
  7. Reichel, D.; Tripathi, M.; Perez, J.M. Biological Effects of Nanoparticles on Macrophage Polarization in the Tumor Microenvironment. Nanotheranostics 2019, 3, 66–88. [Google Scholar] [CrossRef] [Green Version]
  8. Liu, L.; Chen, Q.; Ruan, C.; Chen, X.; Zhang, Y.; He, X.; Zhang, Y.; Lu, Y.; Guo, Q.; Sun, T.; et al. Platinum-Based Nanovectors Engineered with Immuno-Modulating Adjuvant for Inhibiting Tumor growth and Promoting Immunity. Theranostics 2018, 8, 2974–2987. [Google Scholar] [CrossRef]
  9. Mu, W.; Chu, Q.; Liu, Y.; Zhang, N. A Review on Nano-Based Drug Delivery System for Cancer Chemoimmunotherapy. Nano-Micro Lett. 2020, 12, 1–24. [Google Scholar] [CrossRef]
  10. Zhou, Q.; Zhang, L.; Wu, H. Nanomaterials for cancer therapies. Nanotechnol. Rev. 2017, 6, 473–496. [Google Scholar] [CrossRef]
  11. Chen, Y.; Xianyu, Y.; Jiang, X. Surface Modification of Gold Nanoparticles with Small Molecules for Biochemical Analysis. Acc. Chem. Res. 2017, 50, 310–319. [Google Scholar] [CrossRef]
  12. de S.L. Oliveira, A.L.C.; de Araujo Junior, R.F.; Gomes de Carvalho, T.; Chan, A.B.; Schomann, T.; Tamburini, F.; de Geus-Oei, L.F.; Cruz, L.J. Effect of Oxaliplatin-Loaded Poly (d,l-Lactide-co-Glycolic Acid) (PLGA) Nanoparticles Combined with Retinoic Acid and Cholesterol on Apoptosis, Drug Resistance, and Metastasis Factors of Colorectal Cancer. Pharmaceutics 2020, 12, 193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Cruz, L.J.; Stammes, M.A.; Que, I.; van Beek, E.R.; Knol-Blankevoort, V.T.; Snoeks, T.J.A.; Chan, A.; Kaijzel, E.L.; Lowik, C. Effect of PLGA NP size on efficiency to target traumatic brain injury. J. Control Release 2016, 223, 31–41. [Google Scholar] [CrossRef]
  14. Cruz, L.J.; Tacken, P.J.; Pots, J.M.; Torensma, R.; Buschow, S.I.; Figdor, C.G. Comparison of antibodies and carbohydrates to target vaccines to human dendritic cells via DC-SIGN. Biomaterials 2012, 33, 4229–4239. [Google Scholar] [CrossRef] [PubMed]
  15. Rudramurthy, G.R.; Swamy, M.K. Potential applications of engineered nanoparticles in medicine and biology: An update. J. Biol. Inorg. Chem. 2018, 23, 1185–1204. [Google Scholar] [CrossRef]
  16. Khan, I.; Saeed, K.; Khan, I. Nanoparticles: Properties, applications and toxicities. Arab. J. Chem. 2019, 12, 908–931. [Google Scholar] [CrossRef]
  17. Rueda, F.; Eich, C.; Cordobilla, B.; Domingo, P.; Acosta, G.; Albericio, F.; Cruz, L.J.; Domingo, J.C. Effect of TLR ligands co-encapsulated with multiepitopic antigen in nanoliposomes targeted to human DCs via Fc receptor for cancer vaccines. Immunobiology 2017, 222, 989–997. [Google Scholar] [CrossRef]
  18. Cruz, L.J.; Tacken, P.J.; van der Schoot, J.M.S.; Rueda, F.; Torensma, R.; Figdor, C.G. ICAM3-Fc Outperforms Receptor-Specific Antibodies Targeted Nanoparticles to Dendritic Cells for Cross-Presentation. Molecules 2019, 24, 1825. [Google Scholar] [CrossRef] [Green Version]
  19. Butts, C.; Murray, N.; Maksymiuk, A.; Goss, G.; Marshall, E.; Soulieres, D.; Cormier, Y.; Ellis, P.; Price, A.; Sawhney, R.; et al. Randomized phase IIB trial of BLP25 liposome vaccine in stage IIIB and IV non-small-cell lung cancer. J. Clin. Oncol. 2005, 23, 6674–6681. [Google Scholar] [CrossRef] [Green Version]
  20. Anselmo, A.C.; Mitragotri, S. Nanoparticles in the clinic: An update. Bioeng. Transl. Med. 2019, 4, e10143. [Google Scholar] [CrossRef] [Green Version]
  21. Wang, Y.; Wen, Q.; Choi, S. FDA’s Regulatory Science Program for Generic PLA/PLGA-Based Drug Products. Am. Pharm. Rev. 2016, 19, 5–9. [Google Scholar]
  22. Palanikumar, L.; Al-Hosani, S.; Kalmouni, M.; Nguyen, V.P.; Ali, L.; Pasricha, R.; Barrera, F.N.; Magzoub, M. pH-responsive high stability polymeric nanoparticles for targeted delivery of anticancer therapeutics. Commun. Biol. 2020, 3, 95. [Google Scholar] [CrossRef] [PubMed]
  23. Da Silva, C.G.; Camps, M.G.M.; Li, T.; Zerrillo, L.; Lowik, C.W.; Ossendorp, F.; Cruz, L.J. Effective chemoimmunotherapy by co-delivery of doxorubicin and immune adjuvants in biodegradable nanoparticles. Theranostics 2019, 9, 6485–6500. [Google Scholar] [CrossRef]
  24. Mogosanu, G.D.; Grumezescu, A.M.; Bejenaru, C.; Bejenaru, L.E. Polymeric protective agents for nanoparticles in drug delivery and targeting. Int. J. Pharm. 2016, 510, 419–429. [Google Scholar] [CrossRef]
  25. Reardon, P.J.; Parhizkar, M.; Harker, A.H.; Browning, R.J.; Vassileva, V.; Stride, E.; Pedley, R.B.; Edirisinghe, M.; Knowles, J.C. Electrohydrodynamic fabrication of core-shell PLGA nanoparticles with controlled release of cisplatin for enhanced cancer treatment. Int. J. Nanomed. 2017, 12, 3913–3926. [Google Scholar] [CrossRef] [Green Version]
  26. Cruz, L.J.; Tacken, P.J.; Fokkink, R.; Figdor, C.G. The influence of PEG chain length and targeting moiety on antibody-mediated delivery of nanoparticle vaccines to human dendritic cells. Biomaterials 2011, 32, 6791–6803. [Google Scholar] [CrossRef] [PubMed]
  27. Rosalia, R.A.; Cruz, L.J.; van Duikeren, S.; Tromp, A.T.; Silva, A.L.; Jiskoot, W.; de Gruijl, T.; Lowik, C.; Oostendorp, J.; van der Burg, S.H.; et al. CD40-targeted dendritic cell delivery of PLGA-nanoparticle vaccines induce potent anti-tumor responses. Biomaterials 2015, 40, 88–97. [Google Scholar] [CrossRef]
  28. Kinoshita, R.; Ishima, Y.; Chuang, V.T.G.; Nakamura, H.; Fang, J.; Watanabe, H.; Shimizu, T.; Okuhira, K.; Ishida, T.; Maeda, H.; et al. Improved anticancer effects of albumin-bound paclitaxel nanoparticle via augmentation of EPR effect and albumin-protein interactions using S-nitrosated human serum albumin dimer. Biomaterials 2017, 140, 162–169. [Google Scholar] [CrossRef]
  29. Bobo, D.; Robinson, K.J.; Islam, J.; Thurecht, K.J.; Corrie, S.R. Nanoparticle-Based Medicines: A Review of FDA-Approved Materials and Clinical Trials to Date. Pharm. Res. 2016, 33, 2373–2387. [Google Scholar] [CrossRef]
  30. Gao, S.; Yang, D.; Fang, Y.; Lin, X.; Jin, X.; Wang, Q.; Wang, X.; Ke, L.; Shi, K. Engineering Nanoparticles for Targeted Remodeling of the Tumor Microenvironment to Improve Cancer Immunotherapy. Theranostics 2019, 9, 126–151. [Google Scholar] [CrossRef] [PubMed]
  31. Li, G.; Liu, D.; Kimchi, E.T.; Kaifi, J.T.; Qi, X.; Manjunath, Y.; Liu, X.; Deering, T.; Avella, D.M.; Fox, T.; et al. Nanoliposome C6-Ceramide Increases the Anti-tumor Immune Response and Slows Growth of Liver Tumors in Mice. Gastroenterology 2018, 154, 1024–1036.e9. [Google Scholar] [CrossRef] [Green Version]
  32. Shi, J.; Kantoff, P.W.; Wooster, R.; Farokhzad, O.C. Cancer nanomedicine: Progress, challenges and opportunities. Nat. Rev. Cancer 2017, 17, 20–37. [Google Scholar] [CrossRef]
  33. Italiani, P.; Boraschi, D. New Insights into Tissue Macrophages: From Their Origin to the Development of Memory. Immune Netw. 2015, 15, 167–176. [Google Scholar] [CrossRef] [Green Version]
  34. Franklin, R.A.; Liao, W.; Sarkar, A.; Kim, M.V.; Bivona, M.R.; Liu, K.; Pamer, E.G.; Li, M.O. The cellular and molecular origin of tumor-associated macrophages. Science 2014, 344, 921–925. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Cortez-Retamozo, V.; Etzrodt, M.; Newton, A.; Rauch, P.J.; Chudnovskiy, A.; Berger, C.; Ryan, R.J.; Iwamoto, Y.; Marinelli, B.; Gorbatov, R.; et al. Origins of tumor-associated macrophages and neutrophils. Proc. Natl. Acad. Sci. USA 2012, 109, 2491–2496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. van Furth, R.; Cohn, Z.A.; Hirsch, J.G.; Humphrey, J.H.; Spector, W.G.; Langevoort, H.L. The mononuclear phagocyte system: A new classification of macrophages, monocytes, and their precursor cells. Bull. World Health Organ. 1972, 46, 845–852. [Google Scholar]
  37. van de Laar, L.; Saelens, W.; De Prijck, S.; Martens, L.; Scott, C.L.; Van Isterdael, G.; Hoffmann, E.; Beyaert, R.; Saeys, Y.; Lambrecht, B.N.; et al. Yolk Sac Macrophages, Fetal Liver, and Adult Monocytes Can Colonize an Empty Niche and Develop into Functional Tissue-Resident Macrophages. Immunity 2016, 44, 755–768. [Google Scholar] [CrossRef] [Green Version]
  38. Hoeffel, G.; Ginhoux, F. Ontogeny of Tissue-Resident Macrophages. Front. Immunol. 2015, 6, 486. [Google Scholar] [CrossRef] [Green Version]
  39. Salmaninejad, A.; Valilou, S.F.; Soltani, A.; Ahmadi, S.; Abarghan, Y.J.; Rosengren, R.J.; Sahebkar, A. Tumor-associated macrophages: Role in cancer development and therapeutic implications. Cell Oncol. 2019, 42, 591–608. [Google Scholar] [CrossRef] [PubMed]
  40. De Palma, M.; Biziato, D.; Petrova, T.V. Microenvironmental regulation of tumour angiogenesis. Nat. Rev. Cancer 2017, 17, 457–474. [Google Scholar] [CrossRef]
  41. Wei, C.; Yang, C.; Wang, S.; Shi, D.; Zhang, C.; Lin, X.; Liu, Q.; Dou, R.; Xiong, B. Crosstalk between cancer cells and tumor associated macrophages is required for mesenchymal circulating tumor cell-mediated colorectal cancer metastasis. Mol. Cancer 2019, 18, 64. [Google Scholar] [CrossRef] [PubMed]
  42. Lin, Y.; Xu, J.; Lan, H. Tumor-associated macrophages in tumor metastasis: Biological roles and clinical therapeutic applications. J. Hematol. Oncol. 2019, 12, 76. [Google Scholar] [CrossRef] [PubMed]
  43. Hughes, R.; Qian, B.Z.; Rowan, C.; Muthana, M.; Keklikoglou, I.; Olson, O.C.; Tazzyman, S.; Danson, S.; Addison, C.; Clemons, M.; et al. Perivascular M2 Macrophages Stimulate Tumor Relapse after Chemotherapy. Cancer Res. 2015, 75, 3479–3491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Kloc, M.; Uosef, A.; Kubiak, J.Z.; Ghobrial, R.M. Macrophage Proinflammatory Responses to Microorganisms and Transplanted Organs. Int. J. Mol. Sci. 2020, 21, 9669. [Google Scholar] [CrossRef] [PubMed]
  45. Fernando, M.R.; Reyes, J.L.; Iannuzzi, J.; Leung, G.; McKay, D.M. The pro-inflammatory cytokine, interleukin-6, enhances the polarization of alternatively activated macrophages. PLoS ONE 2014, 9, e94188. [Google Scholar] [CrossRef] [PubMed]
  46. Perry, C.J.; Munoz-Rojas, A.R.; Meeth, K.M.; Kellman, L.N.; Amezquita, R.A.; Thakral, D.; Du, V.Y.; Wang, J.X.; Damsky, W.; Kuhlmann, A.L.; et al. Myeloid-targeted immunotherapies act in synergy to induce inflammation and antitumor immunity. J. Exp. Med. 2018, 215, 877–893. [Google Scholar] [CrossRef] [Green Version]
  47. Chen, Y.; Song, Y.; Du, W.; Gong, L.; Chang, H.; Zou, Z. Tumor-associated macrophages: An accomplice in solid tumor progression. J. Biomed. Sci. 2019, 26, 78. [Google Scholar] [CrossRef]
  48. Belguendouz, H.; Lahmar-Belguendouz, K.; Messaoudene, D.; Djeraba, Z.; Otmani, F.; Hakem, D.; Lahlou-Boukoffa, O.S.; Youinou, P.; Touil-Boukoffa, C. Cytokines Modulate the “Immune-Metabolism” Interactions during Behcet Disease: Effect on Arginine Metabolism. Int. J. Inflam. 2015, 2015, 241738. [Google Scholar] [CrossRef]
  49. Li, X.; Su, X.; Liu, R.; Pan, Y.; Fang, J.; Cao, L.; Feng, C.; Shang, Q.; Chen, Y.; Shao, C.; et al. HDAC inhibition potentiates anti-tumor activity of macrophages and enhances anti-PD-L1-mediated tumor suppression. Oncogene 2021, 40, 1836–1850. [Google Scholar] [CrossRef]
  50. Chen, Z.; Malhotra, P.S.; Thomas, G.R.; Ondrey, F.G.; Duffey, D.C.; Smith, C.W.; Enamorado, I.; Yeh, N.T.; Kroog, G.S.; Rudy, S.; et al. Expression of Proinflammatory and Proangiogenic Cytokines in Patients with Head and Neck Cancer. Mol. Oncol. 1999, 5, 1369–1379. [Google Scholar]
  51. Qian, B.Z.; Pollard, J.W. Macrophage diversity enhances tumor progression and metastasis. Cell 2010, 141, 39–51. [Google Scholar] [CrossRef] [Green Version]
  52. Verdeil, G.; Lawrence, T.; Schmitt-Verhulst, A.M.; Auphan-Anezin, N. Targeting STAT3 and STAT5 in Tumor-Associated Immune Cells to Improve Immunotherapy. Cancers 2019, 11, 1832. [Google Scholar] [CrossRef] [Green Version]
  53. Biswas, S.K.; Gangi, L.; Paul, S.; Schioppa, T.; Saccani, A.; Sironi, M.; Bottazzi, B.; Doni, A.; Vincenzo, B.; Pasqualini, F.; et al. A distinct and unique transcriptional program expressed by tumor-associated macrophages (defective NF-kappaB and enhanced IRF-3/STAT1 activation). Blood 2006, 107, 2112–2122. [Google Scholar] [CrossRef] [Green Version]
  54. Zhu, Y.; Herndon, J.M.; Sojka, D.K.; Kim, K.W.; Knolhoff, B.L.; Zuo, C.; Cullinan, D.R.; Luo, J.; Bearden, A.R.; Lavine, K.J.; et al. Tissue-Resident Macrophages in Pancreatic Ductal Adenocarcinoma Originate from Embryonic Hematopoiesis and Promote Tumor Progression. Immunity 2017, 47, 323–338.e6. [Google Scholar] [CrossRef]
  55. Pollard, J.W. The Yolk Sac Feeds Pancreatic Tumors. Immunity 2017, 47, 217–218. [Google Scholar] [CrossRef]
  56. Giurisato, E.; Lonardi, S.; Telfer, B.; Lussoso, S.; Risa-Ebri, B.; Zhang, J.; Russo, I.; Wang, J.; Santucci, A.; Finegan, K.G.; et al. Extracellular-Regulated Protein Kinase 5-Mediated Control of p21 Expression Promotes Macrophage Proliferation Associated with Tumor Growth and Metastasis. Cancer Res. 2020, 80, 3319–3330. [Google Scholar] [CrossRef] [PubMed]
  57. Giurisato, E.; Xu, Q.; Lonardi, S.; Telfer, B.; Russo, I.; Pearson, A.; Finegan, K.G.; Wang, W.; Wang, J.; Gray, N.S.; et al. Myeloid ERK5 deficiency suppresses tumor growth by blocking protumor macrophage polarization via STAT3 inhibition. Proc. Natl. Acad. Sci. USA 2018, 115, E2801–E2810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Zhang, M.Z.; Wang, X.; Wang, Y.; Niu, A.; Wang, S.; Zou, C.; Harris, R.C. IL-4/IL-13-mediated polarization of renal macrophages/dendritic cells to an M2a phenotype is essential for recovery from acute kidney injury. Kidney Int. 2017, 91, 375–386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Bao, L.; Li, X. MicroRNA-32 targeting PTEN enhances M2 macrophage polarization in the glioma microenvironment and further promotes the progression of glioma. Mol. Cell Biochem. 2019, 460, 67–79. [Google Scholar] [CrossRef] [PubMed]
  60. Arranz, A.; Doxaki, C.; Vergadi, E.; Martinez de la Torre, Y.; Vaporidi, K.; Lagoudaki, E.D.; Ieronymaki, E.; Androulidaki, A.; Venihaki, M.; Margioris, A.N.; et al. Akt1 and Akt2 protein kinases differentially contribute to macrophage polarization. Proc. Natl. Acad. Sci. USA 2012, 109, 9517–9522. [Google Scholar] [CrossRef] [Green Version]
  61. Wu, H.; Zheng, J.; Xu, S.; Fang, Y.; Wu, Y.; Zeng, J.; Shao, A.; Shi, L.; Lu, J.; Mei, S.; et al. Mer regulates microglial/macrophage M1/M2 polarization and alleviates neuroinflammation following traumatic brain injury. J. Neuroinflamm. 2021, 18, 2. [Google Scholar] [CrossRef]
  62. Jarnicki, A.G.; Lysaght, J.; Todryk, S.; Mills, K.H. Suppression of antitumor immunity by IL-10 and TGF-beta-producing T cells infiltrating the growing tumor: Influence of tumor environment on the induction of CD4+ and CD8+ regulatory T cells. J. Immunol. 2006, 177, 896–904. [Google Scholar] [CrossRef] [Green Version]
  63. Dennis, K.L.; Blatner, N.R.; Gounari, F.; Khazaie, K. Current status of interleukin-10 and regulatory T-cells in cancer. Curr. Opin. Oncol. 2013, 25, 637–645. [Google Scholar] [CrossRef]
  64. Llopiz, D.; Ruiz, M.; Infante, S.; Sarobe, P.; Hervas-Stubbs, S.; Alignani, D.; Lasarte, J.J.; Villanueva, L.; Silva, L.; Guruceaga, E. IL-10 expression defines an immunosuppressive dendritic cell population induced by antitumor therapeutic vaccination. Oncotarget 2017, 8, 2659. [Google Scholar] [CrossRef]
  65. Noe, J.T.; Mitchell, R.A. MIF-Dependent Control of Tumor Immunity. Front. Immunol. 2020, 11, 609948. [Google Scholar] [CrossRef]
  66. Collison, L.W.; Workman, C.J.; Kuo, T.T.; Boyd, K.; Wang, Y.; Vignali, K.M.; Cross, R.; Sehy, D.; Blumberg, R.S.; Vignali, D.A. The inhibitory cytokine IL-35 contributes to regulatory T-cell function. Nature 2007, 450, 566–569. [Google Scholar] [CrossRef] [PubMed]
  67. Yoo, S.H.; Keam, B.; Ock, C.Y.; Kim, S.; Han, B.; Kim, J.W.; Lee, K.W.; Jeon, Y.K.; Jung, K.C.; Chung, E.J.; et al. Prognostic value of the association between MHC class I downregulation and PD-L1 upregulation in head and neck squamous cell carcinoma patients. Sci. Rep. 2019, 9, 7680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Wang, X.; Yang, X.; Zhang, C.; Wang, Y.; Cheng, T.; Duan, L.; Tong, Z.; Tan, S.; Zhang, H.; Saw, P.E.; et al. Tumor cell-intrinsic PD-1 receptor is a tumor suppressor and mediates resistance to PD-1 blockade therapy. Proc. Natl. Acad. Sci. USA 2020, 117, 6640–6650. [Google Scholar] [CrossRef]
  69. Smyth, M.J.; Teng, M.W. The Nobel Prize in Physiology and Medicine. J. Aust. N. Z. Soc. Immunol. 2018. [Google Scholar] [CrossRef]
  70. Chen, X.; Gao, A.; Zhang, F.; Yang, Z.; Wang, S.; Fang, Y.; Li, J.; Wang, J.; Shi, W.; Wang, L.; et al. ILT4 inhibition prevents TAM- and dysfunctional T cell-mediated immunosuppression and enhances the efficacy of anti-PD-L1 therapy in NSCLC with EGFR activation. Theranostics 2021, 11, 3392–3416. [Google Scholar] [CrossRef]
  71. Anfray, C.; Ummarino, A.; Andon, F.T.; Allavena, P. Current Strategies to Target Tumor-Associated-Macrophages to Improve Anti-Tumor Immune Responses. Cells 2019, 9, 46. [Google Scholar] [CrossRef] [Green Version]
  72. Baghdadi, M.; Endo, H.; Takano, A.; Ishikawa, K.; Kameda, Y.; Wada, H.; Miyagi, Y.; Yokose, T.; Ito, H.; Nakayama, H.; et al. High co-expression of IL-34 and M-CSF correlates with tumor progression and poor survival in lung cancers. Sci. Rep. 2018, 8, 418. [Google Scholar] [CrossRef]
  73. Magkouta, S.F.; Vaitsi, P.C.; Pappas, A.G.; Iliopoulou, M.; Kosti, C.N.; Psarra, K.; Kalomenidis, I.T. CSF1/CSF1R Axis Blockade Limits Mesothelioma and Enhances Efficiency of Anti-PDL1 Immunotherapy. Cancers 2021, 13, 2546. [Google Scholar] [CrossRef]
  74. Pyonteck, S.M.; Akkari, L.; Schuhmacher, A.J.; Bowman, R.L.; Sevenich, L.; Quail, D.F.; Olson, O.C.; Quick, M.L.; Huse, J.T.; Teijeiro, V.; et al. CSF-1R inhibition alters macrophage polarization and blocks glioma progression. Nat. Med. 2013, 19, 1264–1272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Ali, S.; Borin, T.F.; Piranlioglu, R.; Ara, R.; Lebedyeva, I.; Angara, K.; Achyut, B.R.; Arbab, A.S.; Rashid, M.H. Changes in the tumor microenvironment and outcome for TME-targeting therapy in glioblastoma: A pilot study. PLoS ONE 2021, 16, e0246646. [Google Scholar]
  76. Denny, W.A.; Flanagan, J.U. Small-molecule CSF1R kinase inhibitors; review of patents 2015-present. Expert Opin. Ther. Pat. 2021, 31, 107–117. [Google Scholar] [CrossRef] [PubMed]
  77. Dowlati, A.; Harvey, R.D.; Carvajal, R.D.; Hamid, O.; Klempner, S.J.; Kauh, J.S.W.; Peterson, D.A.; Yu, D.; Chapman, S.C.; Szpurka, A.M.; et al. LY3022855, an anti-colony stimulating factor-1 receptor (CSF-1R) monoclonal antibody, in patients with advanced solid tumors refractory to standard therapy: Phase 1 dose-escalation trial. Investig. New Drugs 2021, 39, 1–15. [Google Scholar] [CrossRef] [PubMed]
  78. Kumar, V.; Donthireddy, L.; Marvel, D.; Condamine, T.; Wang, F.; Lavilla-Alonso, S.; Hashimoto, A.; Vonteddu, P.; Behera, R.; Goins, M.A.; et al. Cancer-Associated Fibroblasts Neutralize the Anti-tumor Effect of CSF1 Receptor Blockade by Inducing PMN-MDSC Infiltration of Tumors. Cancer Cell 2017, 32, 654–668.e5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Fantuzzi, L.; Tagliamonte, M.; Gauzzi, M.C.; Lopalco, L. Dual CCR5/CCR2 targeting: Opportunities for the cure of complex disorders. Cell Mol. Life Sci. 2019, 76, 4869–4886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Nywening, T.M.; Wang-Gillam, A.; Sanford, D.E.; Belt, B.A.; Panni, R.Z.; Cusworth, B.M.; Toriola, A.T.; Nieman, R.K.; Worley, L.A.; Yano, M.; et al. Targeting tumour-associated macrophages with CCR2 inhibition in combination with FOLFIRINOX in patients with borderline resectable and locally advanced pancreatic cancer: A single-centre, open-label, dose-finding, non-randomised, phase 1b trial. Lancet Oncol. 2016, 17, 651–662. [Google Scholar] [CrossRef] [Green Version]
  81. Qian, B.Z.; Li, J.; Zhang, H.; Kitamura, T.; Zhang, J.; Campion, L.R.; Kaiser, E.A.; Snyder, L.A.; Pollard, J.W. CCL2 recruits inflammatory monocytes to facilitate breast-tumour metastasis. Nature 2011, 475, 222–225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Fridlender, Z.G.; Kapoor, V.; Buchlis, G.; Cheng, G.; Sun, J.; Wang, L.C.; Singhal, S.; Snyder, L.A.; Albelda, S.M. Monocyte chemoattractant protein-1 blockade inhibits lung cancer tumor growth by altering macrophage phenotype and activating CD8+ cells. Am. J. Respir. Cell Mol. Biol. 2011, 44, 230–237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. He, M.; Yu, W.; Chang, C.; Miyamoto, H.; Liu, X.; Jiang, K.; Yeh, S. Estrogen receptor alpha promotes lung cancer cell invasion via increase of and cross-talk with infiltrated macrophages through the CCL2/CCR2/MMP9 and CXCL12/CXCR4 signaling pathways. Mol. Oncol. 2020, 14, 1779–1799. [Google Scholar] [CrossRef] [PubMed]
  84. Brummer, G.; Fang, W.; Smart, C.; Zinda, B.; Alissa, N.; Berkland, C.; Miller, D.; Cheng, N. CCR2 signaling in breast carcinoma cells promotes tumor growth and invasion by promoting CCL2 and suppressing CD154 effects on the angiogenic and immune microenvironments. Oncogene 2020, 39, 2275–2289. [Google Scholar] [CrossRef] [PubMed]
  85. Xi, S.; Zheng, X.; Li, X.; Jiang, Y.; Wu, Y.; Gong, J.; Jie, Y.; Li, Z.; Cao, J.; Sha, L.; et al. Activated Hepatic Stellate Cells Induce Infiltration and Formation of CD163(+) Macrophages via CCL2/CCR2 Pathway. Front. Med. 2021, 8, 627927. [Google Scholar] [CrossRef] [PubMed]
  86. Noel, M.; O’Reilly, E.M.; Wolpin, B.M.; Ryan, D.P.; Bullock, A.J.; Britten, C.D.; Linehan, D.C.; Belt, B.A.; Gamelin, E.C.; Ganguly, B.; et al. Phase 1b study of a small molecule antagonist of human chemokine (C-C motif) receptor 2 (PF-04136309) in combination with nab-paclitaxel/gemcitabine in first-line treatment of metastatic pancreatic ductal adenocarcinoma. Investig. New Drugs 2020, 38, 800–811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Nie, Y.; Huang, H.; Guo, M.; Chen, J.; Wu, W.; Li, W.; Xu, X.; Lin, X.; Fu, W.; Yao, Y.; et al. Breast Phyllodes Tumors Recruit and Repolarize Tumor-Associated Macrophages via Secreting CCL5 to Promote Malignant Progression, Which Can Be Inhibited by CCR5 Inhibition Therapy. Clin. Cancer Res. 2019, 25, 3873–3886. [Google Scholar] [CrossRef] [Green Version]
  88. Wang, S.W.; Wu, H.H.; Liu, S.C.; Wang, P.C.; Ou, W.C.; Chou, W.Y.; Shen, Y.S.; Tang, C.H. CCL5 and CCR5 interaction promotes cell motility in human osteosarcoma. PLoS ONE 2012, 7, e35101. [Google Scholar]
  89. Xiang, P.; Jin, S.; Yang, Y.; Sheng, J.; He, Q.; Song, Y.; Yu, W.; Hu, S.; Jin, J. Infiltrating CD4+ T cells attenuate chemotherapy sensitivity in prostate cancer via CCL5 signaling. Prostate 2019, 79, 1018–1031. [Google Scholar] [CrossRef]
  90. Jiao, X.; Velasco-Velazquez, M.A.; Wang, M.; Li, Z.; Rui, H.; Peck, A.R.; Korkola, J.E.; Chen, X.; Xu, S.; DuHadaway, J.B.; et al. CCR5 Governs DNA Damage Repair and Breast Cancer Stem Cell Expansion. Cancer Res. 2018, 78, 1657–1671. [Google Scholar] [CrossRef] [Green Version]
  91. Long, H.; Xie, R.; Xiang, T.; Zhao, Z.; Lin, S.; Liang, Z.; Chen, Z.; Zhu, B. Autocrine CCL5 signaling promotes invasion and migration of CD133+ ovarian cancer stem-like cells via NF-kappaB-mediated MMP-9 upregulation. Stem Cells 2012, 30, 2309–2319. [Google Scholar] [CrossRef] [PubMed]
  92. Casagrande, N.; Borghese, C.; Visser, L.; Mongiat, M.; Colombatti, A.; Aldinucci, D. CCR5 antagonism by maraviroc inhibits Hodgkin lymphoma microenvironment interactions and xenograft growth. Haematologica 2019, 104, 564–575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Frankenberger, C.; Rabe, D.; Bainer, R.; Sankarasharma, D.; Chada, K.; Krausz, T.; Gilad, Y.; Becker, L.; Rosner, M.R. Metastasis Suppressors Regulate the Tumor Microenvironment by Blocking Recruitment of Prometastatic Tumor-Associated Macrophages. Cancer Res. 2015, 75, 4063–4073. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Knowles, H.J.; Harris, A.L. Macrophages and the hypoxic tumour microenvironment. Front. Biosci. 2007, 12, 4298–4314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Schioppa, T.; Uranchimeg, B.; Saccani, A.; Biswas, S.K.; Doni, A.; Rapisarda, A.; Bernasconi, S.; Saccani, S.; Nebuloni, M.; Vago, L.; et al. Regulation of the chemokine receptor CXCR4 by hypoxia. J. Exp. Med. 2003, 198, 1391–1402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Mota, J.M.; Leite, C.A.; Souza, L.E.; Melo, P.H.; Nascimento, D.C.; de-Deus-Wagatsuma, V.M.; Temporal, J.; Figueiredo, F.; Noushmehr, H.; Alves-Filho, J.C.; et al. Post-Sepsis State Induces Tumor-Associated Macrophage Accumulation through CXCR4/CXCL12 and Favors Tumor Progression in Mice. Cancer Immunol. Res. 2016, 4, 312–322. [Google Scholar] [CrossRef] [Green Version]
  97. Yusen, W.; Xia, W.; Shengjun, Y.; Shaohui, Z.; Hongzhen, Z. The expression and significance of tumor associated macrophages and CXCR4 in non-small cell lung cancer. J. BUON 2018, 23, 398–402. [Google Scholar] [PubMed]
  98. Ma, J.; Su, H.; Yu, B.; Guo, T.; Gong, Z.; Qi, J.; Zhao, X.; Du, J. CXCL12 gene silencing down-regulates metastatic potential via blockage of MAPK/PI3K/AP-1 signaling pathway in colon cancer. Clin. Transl. Oncol. 2018, 20, 1035–1045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Wu, A.; Maxwell, R.; Xia, Y.; Cardarelli, P.; Oyasu, M.; Belcaid, Z.; Kim, E.; Hung, A.; Luksik, A.S.; Garzon-Muvdi, T.; et al. Combination anti-CXCR4 and anti-PD-1 immunotherapy provides survival benefit in glioblastoma through immune cell modulation of tumor microenvironment. J. Neurooncol. 2019, 143, 241–249. [Google Scholar] [CrossRef]
  100. Zarif, J.C.; Baena-Del Valle, J.A.; Hicks, J.L.; Heaphy, C.M.; Vidal, I.; Luo, J.; Lotan, T.L.; Hooper, J.E.; Isaacs, W.B.; Pienta, K.J.; et al. Mannose Receptor-positive Macrophage Infiltration Correlates with Prostate Cancer Onset and Metastatic Castration-resistant Disease. Eur. Urol. Oncol. 2019, 2, 429–436. [Google Scholar] [CrossRef]
  101. Bockorny, B.; Semenisty, V.; Macarulla, T.; Borazanci, E.; Wolpin, B.M.; Stemmer, S.M.; Golan, T.; Geva, R.; Borad, M.J.; Pedersen, K.S.; et al. BL-8040, a CXCR4 antagonist, in combination with pembrolizumab and chemotherapy for pancreatic cancer: The COMBAT trial. Nat. Med. 2020, 26, 878–885. [Google Scholar] [CrossRef]
  102. Zhang, W.; Huang, Q.; Xiao, W.; Zhao, Y.; Pi, J.; Xu, H.; Zhao, H.; Xu, J.; Evans, C.E.; Jin, H. Advances in Anti-Tumor Treatments Targeting the CD47/SIRPalpha Axis. Front. Immunol. 2020, 11, 18. [Google Scholar] [CrossRef] [Green Version]
  103. Cioffi, M.; Trabulo, S.; Hidalgo, M.; Costello, E.; Greenhalf, W.; Erkan, M.; Kleeff, J.; Sainz, B., Jr.; Heeschen, C. Inhibition of CD47 Effectively Targets Pancreatic Cancer Stem Cells via Dual Mechanisms. Clin. Cancer Res. 2015, 21, 2325–2337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Tseng, D.; Volkmer, J.P.; Willingham, S.B.; Contreras-Trujillo, H.; Fathman, J.W.; Fernhoff, N.B.; Seita, J.; Inlay, M.A.; Weiskopf, K.; Miyanishi, M.; et al. Anti-CD47 antibody-mediated phagocytosis of cancer by macrophages primes an effective antitumor T-cell response. Proc. Natl. Acad. Sci. USA 2013, 110, 11103–11108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Ring, N.G.; Herndler-Brandstetter, D.; Weiskopf, K.; Shan, L.; Volkmer, J.P.; George, B.M.; Lietzenmayer, M.; McKenna, K.M.; Naik, T.J.; McCarty, A.; et al. Anti-SIRPalpha antibody immunotherapy enhances neutrophil and macrophage antitumor activity. Proc. Natl. Acad. Sci. USA 2017, 114, E10578–E10585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Alvey, C.M.; Spinler, K.R.; Irianto, J.; Pfeifer, C.R.; Hayes, B.; Xia, Y.; Cho, S.; Dingal, P.; Hsu, J.; Smith, L.; et al. SIRPA-Inhibited, Marrow-Derived Macrophages Engorge, Accumulate, and Differentiate in Antibody-Targeted Regression of Solid Tumors. Curr. Biol. 2017, 27, 2065–2077.e6. [Google Scholar] [CrossRef] [Green Version]
  107. Petrova, P.S.; Viller, N.N.; Wong, M.; Pang, X.; Lin, G.H.; Dodge, K.; Chai, V.; Chen, H.; Lee, V.; House, V.; et al. TTI-621 (SIRPalphaFc): A CD47-Blocking Innate Immune Checkpoint Inhibitor with Broad Antitumor Activity and Minimal Erythrocyte Binding. Clin. Cancer Res. 2017, 23, 1068–1079. [Google Scholar] [CrossRef] [Green Version]
  108. Sikic, B.I.; Lakhani, N.; Patnaik, A.; Shah, S.A.; Chandana, S.R.; Rasco, D.; Colevas, A.D.; O’rourke, T.; Narayanan, S.; Papadopoulos, K.; et al. First-in-Human, First-in-Class Phase I Trial of the Anti-CD47 Antibody Hu5F9-G4 in Patients With Advanced Cancers. J. Clin. Oncol. 2019, 37, 946. [Google Scholar] [CrossRef]
  109. Zhang, J.Q.; Zeng, S.; Vitiello, G.A.; Seifert, A.M.; Medina, B.D.; Beckman, M.J.; Loo, J.K.; Santamaria-Barria, J.; Maltbaek, J.H.; Param, N.J.; et al. Macrophages and CD8(+) T Cells Mediate the Antitumor Efficacy of Combined CD40 Ligation and Imatinib Therapy in Gastrointestinal Stromal Tumors. Cancer Immunol. Res. 2018, 6, 434–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Vitale, L.A.; Thomas, L.J.; He, L.Z.; O’Neill, T.; Widger, J.; Crocker, A.; Sundarapandiyan, K.; Storey, J.R.; Forsberg, E.M.; Weidlick, J.; et al. Development of CDX-1140, an agonist CD40 antibody for cancer immunotherapy. Cancer Immunol. Immunother. 2019, 68, 233–245. [Google Scholar] [CrossRef]
  111. Djureinovic, D.; Wang, M.; Kluger, H.M. Agonistic CD40 Antibodies in Cancer Treatment. Cancers 2021, 13, 1302. [Google Scholar] [CrossRef]
  112. Kaneda, M.M.; Messer, K.S.; Ralainirina, N.; Li, H.; Leem, C.J.; Gorjestani, S.; Woo, G.; Nguyen, A.V.; Figueiredo, C.C.; Foubert, P.; et al. PI3Kgamma is a molecular switch that controls immune suppression. Nature 2016, 539, 437–442. [Google Scholar] [CrossRef] [Green Version]
  113. Mohan, S.; Vander Broek, R.; Shah, S.; Eytan, D.F.; Pierce, M.L.; Carlson, S.G.; Coupar, J.F.; Zhang, J.; Cheng, H.; Chen, Z.; et al. MEK Inhibitor PD-0325901 Overcomes Resistance to PI3K/mTOR Inhibitor PF-5212384 and Potentiates Antitumor Effects in Human Head and Neck Squamous Cell Carcinoma. Clin. Cancer Res. 2015, 21, 3946–3956. [Google Scholar] [CrossRef] [Green Version]
  114. Li, H.; Somiya, M.; Kuroda, S. Enhancing antibody-dependent cellular phagocytosis by Re-education of tumor-associated macrophages with resiquimod-encapsulated liposomes. Biomaterials 2021, 268, 120601. [Google Scholar] [CrossRef]
  115. Yang, L.; Zhang, Y. Tumor-associated macrophages: From basic research to clinical application. J. Hematol. Oncol. 2017, 10, 58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Sui, A.; Chen, X.; Demetriades, A.M.; Shen, J.; Cai, Y.; Yao, Y.; Yao, Y.; Zhu, Y.; Shen, X.; Xie, B. Inhibiting NF-kappaB Signaling Activation Reduces Retinal Neovascularization by Promoting a Polarization Shift in Macrophages. Investig. Ophthalmol. Vis. Sci. 2020, 61, 4. [Google Scholar] [CrossRef]
  117. Cavalcante, R.S.; Ishikawa, U.; Silva, E.S.; Silva Junior, A.A.; Araujo, A.A.; Cruz, L.J.; Chan, A.B.; Araujo Junior, R.F. M2 TAM-associated STAT3/NF-kappaB signalling suppression as major target of immunomodulatory therapy with PLGA-based nanocarriers and anti-PD-L1 in breast cancer. Br. J. Pharmacol. 2021, 178, 2284–2304. [Google Scholar] [CrossRef] [PubMed]
  118. de Araujo Junior, R.F.; Eich, C.; Jorquera, C.; Schomann, T.; Baldazzi, F.; Chan, A.B.; Cruz, L.J. Ceramide and palmitic acid inhibit macrophage-mediated epithelial-mesenchymal transition in colorectal cancer. Mol. Cell Biochem. 2020, 468, 153–168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Mao, Y.; Liu, X. Bioresponsive Nanomedicine: The Next Step of Deadliest Cancers’ Theranostics. Front. Chem. 2020, 8, 257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Lobera, M.; Madauss, K.P.; Pohlhaus, D.T.; Wright, Q.G.; Trocha, M.; Schmidt, D.R.; Baloglu, E.; Trump, R.P.; Head, M.S.; Hofmann, G.A.; et al. Selective class IIa histone deacetylase inhibition via a nonchelating zinc-binding group. Nat. Chem. Biol. 2013, 9, 319–325. [Google Scholar] [CrossRef]
  121. Guerriero, J.L.; Sotayo, A.; Ponichtera, H.E.; Castrillon, J.A.; Pourzia, A.L.; Schad, S.; Johnson, S.F.; Carrasco, R.D.; Lazo, S.; Bronson, R.T.; et al. Class IIa HDAC inhibition reduces breast tumours and metastases through anti-tumour macrophages. Nature 2017, 543, 428–432. [Google Scholar] [CrossRef]
  122. Frank, A.C.; Raue, R.; Fuhrmann, D.C.; Sirait-Fischer, E.; Reuse, C.; Weigert, A.; Lutjohann, D.; Hiller, K.; Syed, S.N.; Brune, B. Lactate dehydrogenase B regulates macrophage metabolism in the tumor microenvironment. Theranostics 2021, 11, 7570–7588. [Google Scholar] [CrossRef]
  123. Zheng, P.; Chen, L.; Yuan, X.; Luo, Q.; Liu, Y.; Xie, G.; Ma, Y.; Shen, L. Exosomal transfer of tumor-associated macrophage-derived miR-21 confers cisplatin resistance in gastric cancer cells. J. Exp. Clin. Cancer Res. 2017, 36, 53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Wayne, E.C.; Long, C.; Haney, M.J.; Batrakova, E.V.; Leisner, T.M.; Parise, L.V.; Kabanov, A.V. Targeted Delivery of siRNA Lipoplexes to Cancer Cells Using Macrophage Transient Horizontal Gene Transfer. Adv. Sci. 2019, 6, 1900582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Zhao, J.; Li, H.; Zhao, S.; Wang, E.; Zhu, J.; Feng, D.; Zhu, Y.; Dou, W.; Fan, Q.; Hu, J.; et al. Epigenetic silencing of miR-144/451a cluster contributes to HCC progression via paracrine HGF/MIF-mediated TAM remodeling. Mol. Cancer 2021, 20, 46. [Google Scholar] [CrossRef] [PubMed]
  126. Qian, Y.; Qiao, S.; Dai, Y.; Xu, G.; Dai, B.; Lu, L.; Yu, X.; Luo, Q.; Zhang, Z. Molecular-Targeted Immunotherapeutic Strategy for Melanoma via Dual-Targeting Nanoparticles Delivering Small Interfering RNA to Tumor-Associated Macrophages. ACS Nano 2017, 11, 9536–9549. [Google Scholar] [CrossRef] [PubMed]
  127. Song, Y.; Tang, C.; Yin, C. Combination antitumor immunotherapy with VEGF and PIGF siRNA via systemic delivery of multi-functionalized nanoparticles to tumor-associated macrophages and breast cancer cells. Biomaterials 2018, 185, 117–132. [Google Scholar] [CrossRef] [PubMed]
  128. Rodon, J.; Brana, I.; Siu, L.L.; De Jonge, M.J.; Homji, N.; Mills, D.; Di Tomaso, E.; Sarr, C.; Trandafir, L.; Massacesi, C.; et al. Phase I dose-escalation and -expansion study of buparlisib (BKM120), an oral pan-Class I PI3K inhibitor, in patients with advanced solid tumors. Investig. New Drugs 2014, 32, 670–681. [Google Scholar] [CrossRef] [PubMed]
  129. Bendell, J.C.; Rodon, J.; Burris, H.A.; de Jonge, M.; Verweij, J.; Birle, D.; Demanse, D.; De Buck, S.S.; Ru, Q.C.; Peters, M.; et al. Phase I, dose-escalation study of BKM120, an oral pan-Class I PI3K inhibitor, in patients with advanced solid tumors. J. Clin. Oncol. 2012, 30, 282–290. [Google Scholar] [CrossRef] [PubMed]
  130. Lu, Y.S.; Lee, K.S.; Chao, T.Y.; Tseng, L.M.; Chitapanarux, I.; Chen, S.C.; Liu, C.T.; Sohn, J.; Kim, J.H.; Chang, Y.C.; et al. A Phase Ib Study of Alpelisib or Buparlisib Combined with Tamoxifen Plus Goserelin in Premenopausal Women with HR-Positive HER2-Negative Advanced Breast Cancer. Clin. Cancer Res. 2021, 27, 408–417. [Google Scholar] [CrossRef]
  131. Treilleux, I.; Arnedos, M.; Cropet, C.; Wang, Q.; Ferrero, J.M.; Abadie-Lacourtoisie, S.; Levy, C.; Legouffe, E.; Lortholary, A.; Pujade-Lauraine, E.; et al. Translational studies within the TAMRAD randomized GINECO trial: Evidence for mTORC1 activation marker as a predictive factor for everolimus efficacy in advanced breast cancer. Ann. Oncol. 2015, 26, 120–125. [Google Scholar] [CrossRef]
  132. Wagner, U.; du Bois, A.; Pfisterer, J.; Huober, J.; Loibl, S.; Luck, H.J.; Sehouli, J.; Gropp, M.; Stahle, A.; Schmalfeldt, B.; et al. Gefitinib in combination with tamoxifen in patients with ovarian cancer refractory or resistant to platinum-taxane based therapy--a phase II trial of the AGO Ovarian Cancer Study Group (AGO-OVAR 2.6). Gynecol. Oncol. 2007, 105, 132–137. [Google Scholar] [CrossRef] [PubMed]
  133. Mayer, E.L.; DeMichele, A.; Rugo, H.S.; Miller, K.; Waks, A.G.; Come, S.E.; Mulvey, T.; Jeselsohn, R.; Overmoyer, B.; Guo, H.; et al. A phase II feasibility study of palbociclib in combination with adjuvant endocrine therapy for hormone receptor-positive invasive breast carcinoma. Ann. Oncol. 2019, 30, 1514–1520. [Google Scholar] [CrossRef] [PubMed]
  134. Tap, W.D.; Gelderblom, H.; Palmerini, E.; Desai, J.; Bauer, S.; Blay, J.-Y.; Alcindor, T.; Ganjoo, K.; Martín-Broto, J.; Ryan, C.W.; et al. Pexidartinib versus placebo for advanced tenosynovial giant cell tumour (ENLIVEN): A randomised phase 3 trial. Lancet 2019, 394, 478–487. [Google Scholar] [CrossRef]
  135. Tap, W. ENLIVEN study: Pexidartinib for tenosynovial giant cell tumor (TGCT). Futur. Oncol. 2020, 16, 1875–1878. [Google Scholar] [CrossRef]
  136. Gelhorn, H.L.; Tong, S.; McQuarrie, K.; Vernon, C.; Hanlon, J.; Maclaine, G.; Lenderking, W.; Ye, X.; Speck, R.M.; Lackman, R.D.; et al. Patient-reported Symptoms of Tenosynovial Giant Cell Tumors. Clin. Ther. 2016, 38, 778–793. [Google Scholar] [CrossRef] [Green Version]
  137. Lewis, J.H.; Gelderblom, H.; van de Sande, M.; Stacchiotti, S.; Healey, J.H.; Tap, W.D.; Wagner, A.J.; Pousa, A.L.; Druta, M.; Lin, C.C.; et al. Pexidartinib Long-Term Hepatic Safety Profile in Patients with Tenosynovial Giant Cell Tumors. Oncologist 2021, 26, e863–e873. [Google Scholar] [CrossRef]
  138. Liu, J.; Wang, L.; Zhao, F.; Tseng, S.; Narayanan, C.; Shura, L.; Willingham, S.; Howard, M.; Prohaska, S.; Volkmer, J.; et al. Pre-Clinical Development of a Humanized Anti-CD47 Antibody with Anti-Cancer Therapeutic Potential. PLoS ONE 2015, 10, e0137345. [Google Scholar] [CrossRef] [Green Version]
  139. Johnson, P.; Challis, R.; Chowdhury, F.; Gao, Y.; Harvey, M.; Geldart, T.; Kerr, P.; Chan, C.; Smith, A.; Steven, N.; et al. Clinical and biological effects of an agonist anti-CD40 antibody: A Cancer Research UK phase I study. Clin. Cancer Res. 2015, 21, 1321–1328. [Google Scholar] [CrossRef] [Green Version]
  140. Beatty, G.L.; Chiorean, E.G.; Fishman, M.P.; Saboury, B.; Teitelbaum, U.R.; Sun, W.; Huhn, R.D.; Song, W.; Li, D.; Sharp, L.L.; et al. CD40 agonists alter tumor stroma and show efficacy against pancreatic carcinoma in mice and humans. Science 2011, 331, 1612–1616. [Google Scholar] [CrossRef] [Green Version]
  141. Vonderheide, R.H.; Burg, J.M.; Mick, R.; Trosko, J.A.; Li, D.; Shaik, M.N.; Tolcher, A.W.; Hamid, O. Phase I study of the CD40 agonist antibody CP-870,893 combined with carboplatin and paclitaxel in patients with advanced solid tumors. Oncoimmunology 2013, 2, e23033. [Google Scholar] [CrossRef] [Green Version]
  142. Wolchok, J.D.; Hoos, A.; O’Day, S.; Weber, J.S.; Hamid, O.; Lebbe, C.; Maio, M.; Binder, M.; Bohnsack, O.; Nichol, G.; et al. Guidelines for the evaluation of immune therapy activity in solid tumors: Immune-related response criteria. Clin. Cancer Res. 2009, 15, 7412–7420. [Google Scholar] [CrossRef] [Green Version]
  143. Vassilios, E.K.; Dardoufas, C.E.; Kouvaris, J.R.; Gennatas, C.S.; Polyzos, A.K.; Gogas, H.J.; Sandilos, P.H.; Uzunoglu, N.K.; Malas, E.G.; Vlahos, L.J. Liposomal Doxorubicin in Conjunction with Reirradiation and Local Hyperthermia Treatment in Recurrent Breast Cancer: A Phase I/II Trial. Clin. Cancer Res. 2002, 8, 374–382. [Google Scholar]
  144. Ngambenjawong, C.; Gustafson, H.H.; Pun, S.H. Progress in tumor-associated macrophage (TAM)-targeted therapeutics. Adv. Drug Deliv. Rev. 2017, 114, 206–221. [Google Scholar] [CrossRef] [Green Version]
  145. Vu, M.N.; Rajasekhar, P.; Poole, D.P.; Khor, S.Y.; Truong, N.P.; Nowell, C.J.; Quinn, J.F.; Whittaker, M.; Veldhuis, N.A.; Davis, T.P. Rapid Assessment of Nanoparticle Extravasation in a Microfluidic Tumor Model. ACS Appl. Nano Mater. 2019, 2, 1844–1856. [Google Scholar] [CrossRef] [Green Version]
  146. Mahlbacher, G.; Curtis, L.T.; Lowengrub, J.; Frieboes, H.B. Mathematical modeling of tumor-associated macrophage interactions with the cancer microenvironment. J. Immunother. Cancer 2018, 6, 10. [Google Scholar] [CrossRef] [Green Version]
  147. Ovais, M.; Guo, M.; Chen, C. Tailoring Nanomaterials for Targeting Tumor-Associated Macrophages. Adv. Mater. 2019, 31, e1808303. [Google Scholar] [CrossRef] [PubMed]
  148. Binnemars-Postma, K.; Storm, G.; Prakash, J. Nanomedicine Strategies to Target Tumor-Associated Macrophages. Int. J. Mol. Sci. 2017, 18, 979. [Google Scholar] [CrossRef] [PubMed]
  149. Zhou, Y.; Que, K.T.; Tang, H.M.; Zhang, P.; Fu, Q.M.; Liu, Z.J. Anti-CD206 antibody-conjugated Fe3O4-based PLGA nanoparticles selectively promote tumor-associated macrophages to polarize to the pro-inflammatory subtype. Oncol. Lett. 2020, 20, 298. [Google Scholar] [CrossRef] [PubMed]
  150. Yuan, Y.; Long, L.; Lin, Y.; Peng, C.; Tang, Y.; Zhou, X.; Li, S.; Zhang, C.; Li, X.; Zhou, X. The Double-edged Sword Effect of Macrophage Targeting Delivery System in Different Macrophage Subsets Related Diseases. J. Nanobiotechnol. 2020, 18, 1–14. [Google Scholar] [CrossRef]
  151. Thomann, J.S.; Heurtault, B.; Weidner, S.; Braye, M.; Beyrath, J.; Fournel, S.; Schuber, F.; Frisch, B. Antitumor activity of liposomal ErbB2/HER2 epitope peptide-based vaccine constructs incorporating TLR agonists and mannose receptor targeting. Biomaterials 2011, 32, 4574–4583. [Google Scholar] [CrossRef]
  152. Ni, L.; Gayet, I.; Zurawski, S.; Duluc, D.; Flamar, A.L.; Li, X.H.; O’Bar, A.; Clayton, S.; Palucka, A.K.; Zurawski, G.; et al. Concomitant activation and antigen uptake via human dectin-1 results in potent antigen-specific CD8+ T cell responses. J. Immunol. 2010, 185, 3504–3513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Leidi, M.; Gotti, E.; Bologna, L.; Miranda, E.; Rimoldi, M.; Sica, A.; Roncalli, M.; Palumbo, G.A.; Introna, M.; Golay, J. M2 macrophages phagocytose rituximab-opsonized leukemic targets more efficiently than m1 cells in vitro. J. Immunol. 2009, 182, 4415–4422. [Google Scholar] [CrossRef] [Green Version]
  154. Kuhn, D.A.; Vanhecke, D.; Michen, B.; Blank, F.; Gehr, P.; Petri-Fink, A.; Rothen-Rutishauser, B. Different endocytotic uptake mechanisms for nanoparticles in epithelial cells and macrophages. Beilstein J. Nanotechnol. 2014, 5, 1625–1636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. dos Santos, T.; Varela, J.; Lynch, I.; Salvati, A.; Dawson, K.A. Effects of transport inhibitors on the cellular uptake of carboxylated polystyrene nanoparticles in different cell lines. PLoS ONE 2011, 6, e24438. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Yeh, Y.M.; Hsu, S.J.; Lin, P.C.; Hsu, K.F.; Wu, P.Y.; Su, W.C.; Chang, J.Y.; Shen, M.R. The c.1085A>G Genetic Variant of CSF1R Gene Regulates Tumor Immunity by Altering the Proliferation, Polarization, and Function of Macrophages. Clin. Cancer Res. 2017, 23, 6021–6030. [Google Scholar] [CrossRef] [Green Version]
  157. Tashiro-Yamaji, J.; Kubota, T.; Yoshida, R. Macrophage MHC receptor 2: A novel receptor on allograft (H-2D(d)K(d))-induced macrophage (H-2D(b)K(b)) recognizing an MHC class I molecule, H-2K(d), in mice. Gene 2006, 384, 1–8. [Google Scholar] [CrossRef] [PubMed]
  158. Toshchakov, V.; Jones, B.W.; Perera, P.Y.; Thomas, K.; Cody, M.J.; Zhang, S.; Williams, B.R.; Major, J.; Hamilton, T.A.; Fenton, M.J.; et al. TLR4, but not TLR2, mediates IFN-beta-induced STAT1alpha/beta-dependent gene expression in macrophages. Nat. Immunol. 2002, 3, 392–398. [Google Scholar] [CrossRef]
  159. Roszer, T. Understanding the Mysterious M2 Macrophage through Activation Markers and Effector Mechanisms. Mediat. Inflamm. 2015, 2015, 816460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Xiong, S.; Zhao, X.; Heng, B.C.; Ng, K.W.; Loo, J.S. Cellular uptake of Poly-(d,l-lactide-co-glycolide) (PLGA) nanoparticles synthesized through solvent emulsion evaporation and nanoprecipitation method. Biotechnol. J. 2011, 6, 501–508. [Google Scholar] [CrossRef]
  161. Augustine, R.; Hasan, A.; Primavera, R.; Wilson, R.J.; Thakor, A.S.; Kevadiya, B.D. Cellular uptake and retention of nanoparticles: Insights on particle properties and interaction with cellular components. Mater. Today Commun. 2020, 25, 101692. [Google Scholar] [CrossRef]
  162. Cruz, L.J.; Tacken, P.J.; Rueda, F.; Domingo, J.C.; Albericio, F.; Figdor, C.G. Targeting nanoparticles to dendritic cells for immunotherapy. Methods Enzymol. 2012, 509, 143–163. [Google Scholar] [PubMed]
  163. Sousa de Almeida, M.; Susnik, E.; Drasler, B.; Taladriz-Blanco, P.; Petri-Fink, A.; Rothen-Rutishauser, B. Understanding nanoparticle endocytosis to improve targeting strategies in nanomedicine. Chem. Soc. Rev. 2021, 50, 5397–5434. [Google Scholar] [CrossRef]
  164. Cruz, L.J.; Tacken, P.J.; Fokkink, R.; Joosten, B.; Stuart, M.C.; Albericio, F.; Torensma, R.; Figdor, C.G. Targeted PLGA nano- but not microparticles specifically deliver antigen to human dendritic cells via DC-SIGN in vitro. J. Control Release 2010, 144, 118–126. [Google Scholar] [CrossRef] [PubMed]
  165. Kanchan, V.; Panda, A.K. Interactions of antigen-loaded polylactide particles with macrophages and their correlation with the immune response. Biomaterials 2007, 28, 5344–5357. [Google Scholar] [CrossRef] [PubMed]
  166. Ehrlich, M.; Boll, W.; Van Oijen, A.; Hariharan, R.; Chandran, K.; Nibert, M.L.; Kirchhausen, T. Endocytosis by random initiation and stabilization of clathrin-coated pits. Cell 2004, 118, 591–605. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Manolova, V.; Flace, A.; Bauer, M.; Schwarz, K.; Saudan, P.; Bachmann, M.F. Nanoparticles target distinct dendritic cell populations according to their size. Eur. J. Immunol. 2008, 38, 1404–1413. [Google Scholar] [CrossRef] [PubMed]
  168. Alber, F.; Dokudovskaya, S.; Veenhoff, L.M.; Zhang, W.; Kipper, J.; Devos, D.; Suprapto, A.; Karni-Schmidt, O.; Williams, R.; Chait, B.T.; et al. The molecular architecture of the nuclear pore complex. Nature 2007, 450, 695–701. [Google Scholar] [CrossRef]
  169. Zhu, M.; Nie, G.; Meng, H.; Xia, T.; Nel, A.; Zhao, Y. Physicochemical Properties Determine Nanomaterial Cellular Uptake, Transport, and Fate. Acc. Chem. Res. 2013, 46, 622–631. [Google Scholar] [CrossRef] [Green Version]
  170. Jones, S.W.; Christison, R.; Bundell, K.; Voyce, C.J.; Brockbank, S.M.; Newham, P.; Lindsay, M.A. Characterisation of cell-penetrating peptide-mediated peptide delivery. Br. J. Pharmacol. 2005, 145, 1093–1102. [Google Scholar] [CrossRef] [PubMed]
  171. Lutsiak, M.E.; Kwon, G.S.; Samuel, J. Biodegradable nanoparticle delivery of a Th2-biased peptide for induction of Th1 immune responses. J. Pharm. Pharmacol. 2006, 58, 739–747. [Google Scholar] [CrossRef]
  172. Lundqvist, M.; Stigler, J.; Elia, G.; Lynch, I.; Cedervall, T.; Dawson, K.A. Nanoparticle size and surface properties determine the protein corona with possible implications for biological impacts. Proc. Natl. Acad. Sci. USA 2008, 105, 14265–14270. [Google Scholar] [CrossRef] [Green Version]
  173. Konduru, N.V.; Molina, R.M.; Swami, A.; Damiani, F.; Pyrgiotakis, G.; Lin, P.; Andreozzi, P.; Donaghey, T.C.; Demokritou, P.; Krol, S.; et al. Protein corona: Implications for nanoparticle interactions with pulmonary cells. Part. Fibre Toxicol. 2017, 14, 42. [Google Scholar] [CrossRef] [Green Version]
  174. Salvati, A.; Pitek, A.S.; Monopoli, M.P.; Prapainop, K.; Bombelli, F.B.; Hristov, D.R.; Kelly, P.M.; Aberg, C.; Mahon, E.; Dawson, K.A. Transferrin-functionalized nanoparticles lose their targeting capabilities when a biomolecule corona adsorbs on the surface. Nat. Nanotechnol. 2013, 8, 137–143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Hoppstadter, J.; Seif, M.; Dembek, A.; Cavelius, C.; Huwer, H.; Kraegeloh, A.; Kiemer, A.K. M2 polarization enhances silica nanoparticle uptake by macrophages. Front. Pharmacol. 2015, 6, 55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Fedeli, C.; Segat, D.; Tavano, R.; Bubacco, L.; De Franceschi, G.; de Laureto, P.P.; Lubian, E.; Selvestrel, F.; Mancin, F.; Papini, E. The functional dissection of the plasma corona of SiO(2)-NPs spots histidine rich glycoprotein as a major player able to hamper nanoparticle capture by macrophages. Nanoscale 2015, 7, 17710–17728. [Google Scholar] [CrossRef] [PubMed]
  177. Binnemars-Postma, K.A.; Ten Hoopen, H.W.; Storm, G.; Prakash, J. Differential uptake of nanoparticles by human M1 and M2 polarized macrophages: Protein corona as a critical determinant. Nanomedicine 2016, 11, 2889–2902. [Google Scholar] [CrossRef] [PubMed]
  178. Zuckerman, J.E.; Choi, C.H.; Han, H.; Davis, M.E. Polycation-siRNA nanoparticles can disassemble at the kidney glomerular basement membrane. Proc. Natl. Acad. Sci. USA 2012, 109, 3137–3142. [Google Scholar] [CrossRef] [Green Version]
  179. Longmire, M.; Choyke, P.L.; Kobayashi, H. Clearance properties of nano-sized particles and molecules as imaging agents: Considerations and caveats. Nanomedicine 2008, 3, 703–717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Cabral, H.; Matsumoto, Y.; Mizuno, K.; Chen, Q.; Murakami, M.; Kimura, M.; Terada, Y.; Kano, M.R.; Miyazono, K.; Uesaka, M.; et al. Accumulation of sub-100 nm polymeric micelles in poorly permeable tumours depends on size. Nat. Nanotechnol. 2011, 6, 815–823. [Google Scholar] [CrossRef]
  181. Matsumura, Y.; Maeda, H. A New Concept for Macromolecular Therapeutics in Cancer Chemotherapy: Mechanism of Tumoritropic Accumulation of Proteins and the Antitumor Agent Smancs. Cancer Res. 1986, 46, 6387–6392. [Google Scholar]
  182. Ren, H.; He, Y.; Liang, J.; Cheng, Z.; Zhang, M.; Zhu, Y.; Hong, C.; Qin, J.; Xu, X.; Wang, J. Role of Liposome Size, Surface Charge, and PEGylation on Rheumatoid Arthritis Targeting Therapy. ACS Appl. Mater. Interfaces 2019, 11, 20304–20315. [Google Scholar] [CrossRef]
  183. Liu, Y.; Tan, J.; Thomas, A.; Ou-Yang, D.; Muzykantov, V.R. The shape of things to come: Importance of design in nanotechnology for drug delivery. Ther. Deliv. 2012, 3, 181–194. [Google Scholar] [CrossRef] [Green Version]
  184. Chen, L.; Li, X.; Zhang, Y.; Chen, T.; Xiao, S.; Liang, H. Morphological and Mechanical Determinants of Cellular Uptake of Deformable Nanoparticles. Nanoscale 2018, 10, 11969–11979. [Google Scholar] [CrossRef] [Green Version]
  185. Chen, L.; Xiao, S.; Zhu, H.; Wang, L.; Liang, H. Shape-dependent internalization kinetics of nanoparticles by membranes. Soft Matter 2016, 12, 2632–2641. [Google Scholar] [CrossRef] [PubMed]
  186. Li, Z.; Sun, L.; Zhang, Y.; Dove, A.P.; O’Reilly, R.K.; Chen, G. Shape Effect of Glyco-Nanoparticles on Macrophage Cellular Uptake and Immune Response. ACS Macro Lett. 2016, 5, 1059–1064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Niikura, K.; Matsunaga, T.; Suzuki, T.; Kobayashi, S.; Yamaguchi, H.; Orba, Y.; Kawaguchi, A.; Hasegawa, H.; Kajino, K.; Ninomiya, T.; et al. Gold Nanoparticles as a Vaccine Platform: Influence of Size and Shape on Immunological Responses in Vitro and in Vivo. ACS Nano 2013, 7, 3926–3938. [Google Scholar] [CrossRef] [PubMed]
  188. Bartneck, M.; Keul, H.A.; Singh, S.; Czaja, K.; Bornemann, J.; Bockstaller, M.; Moeller, M.; Zwadlo-Klarwasser, G.; Groll, J. Rapid Uptake of Gold Nanorods by Primary Human Blood Phagocytes and Immunomodulatory Effects of Surface Chemistry. ACS Nano 2010, 4, 3073–3086. [Google Scholar] [CrossRef]
  189. Hinde, E.; Thammasiraphop, K.; Duong, H.T.; Yeow, J.; Karagoz, B.; Boyer, C.; Gooding, J.J.; Gaus, K. Pair correlation microscopy reveals the role of nanoparticle shape in intracellular transport and site of drug release. Nat. Nanotechnol. 2017, 12, 81–89. [Google Scholar] [CrossRef]
  190. Du, X.J.; Wang, J.L.; Iqbal, S.; Li, H.J.; Cao, Z.T.; Wang, Y.C.; Du, J.Z.; Wang, J. The effect of surface charge on oral absorption of polymeric nanoparticles. Biomater. Sci. 2018, 6, 642–650. [Google Scholar] [CrossRef]
  191. Hsu, F.F.; Turk, J. Studies on phosphatidylserine by tandem quadrupole and multiple stage quadrupole ion-trap mass spectrometry with electrospray ionization: Structural characterization and the fragmentation processes. J. Am. Soc. Mass. Spectrom. 2005, 16, 1510–1522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Xiao, K.; Li, Y.; Luo, J.; Lee, J.S.; Xiao, W.; Gonik, A.M.; Agarwal, R.G.; Lam, K.S. The effect of surface charge on in vivo biodistribution of PEG-oligocholic acid based micellar nanoparticles. Biomaterials 2011, 32, 3435–3446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Brannon-Peppas, L.; Blanchette, J.O. Nanoparticle and targeted systems for cancer therapy. Adv. Drug Deliv. Rev. 2012, 64, 206–212. [Google Scholar] [CrossRef]
  194. Getts, D.R.; Martin, A.J.; McCarthy, D.P.; Terry, R.L.; Hunter, Z.N.; Yap, W.T.; Getts, M.T.; Pleiss, M.; Luo, X.; King, N.J.; et al. Microparticles bearing encephalitogenic peptides induce T-cell tolerance and ameliorate experimental autoimmune encephalomyelitis. Nat. Biotechnol. 2012, 30, 1217–1224. [Google Scholar] [CrossRef]
  195. Song, Q.; Wang, X.; Hu, Q.; Huang, M.; Yao, L.; Qi, H.; Qiu, Y.; Jiang, X.; Chen, J.; Chen, H.; et al. Cellular internalization pathway and transcellular transport of pegylated polyester nanoparticles in Caco-2 cells. Int. J. Pharm. 2013, 445, 58–68. [Google Scholar] [CrossRef]
  196. Peracchia, M.T.; Fattal, E.; Desmaele, D.; Besnard, M.; Noel, J.P.; Gomis, J.M.; Appel, M.; d’Angelo, J.; Couvreur, P. Stealth PEGylated polycyanoacrylate nanoparticles for intravenous administration and splenic targeting. J. Control. Release 1999, 60, 121–128. [Google Scholar] [CrossRef]
  197. Yu, Q.; Zhao, L.; Guo, C.; Yan, B.; Su, G. Regulating Protein Corona Formation and Dynamic Protein Exchange by Controlling Nanoparticle Hydrophobicity. Front. Bioeng. Biotechnol. 2020, 8, 210. [Google Scholar] [CrossRef] [Green Version]
  198. Jones, S.W.; Roberts, R.A.; Robbins, G.R.; Perry, J.L.; Kai, M.P.; Chen, K.; Bo, T.; Napier, M.E.; Ting, J.P.; Desimone, J.M.; et al. Nanoparticle clearance is governed by Th1/Th2 immunity and strain background. J. Clin. Investig. 2013, 123, 3061–3073. [Google Scholar] [CrossRef] [Green Version]
  199. MacParland, S.A.; Tsoi, K.M.; Ouyang, B.; Ma, X.Z.; Manuel, J.; Fawaz, A.; Ostrowski, M.A.; Alman, B.A.; Zilman, A.; Chan, W.C.; et al. Phenotype Determines Nanoparticle Uptake by Human Macrophages from Liver and Blood. ACS Nano 2017, 11, 2428–2443. [Google Scholar] [CrossRef]
  200. Herd, H.L.; Bartlett, K.T.; Gustafson, J.A.; McGill, L.D.; Ghandehari, H. Macrophage silica nanoparticle response is phenotypically dependent. Biomaterials 2015, 53, 574–582. [Google Scholar] [CrossRef] [Green Version]
  201. Feito, M.J.; Diez-Orejas, R.; Cicuendez, M.; Casarrubios, L.; Rojo, J.M.; Portoles, M.T. Characterization of M1 and M2 polarization phenotypes in peritoneal macrophages after treatment with graphene oxide nanosheets. Colloids Surf B Biointerfaces 2019, 176, 96–105. [Google Scholar] [CrossRef] [Green Version]
  202. Pajarinen, J.; Kouri, V.P.; Jamsen, E.; Li, T.F.; Mandelin, J.; Konttinen, Y.T. The response of macrophages to titanium particles is determined by macrophage polarization. Acta Biomater. 2013, 9, 9229–9240. [Google Scholar] [CrossRef] [PubMed]
  203. Barua, S.; Mitragotri, S. Challenges associated with Penetration of Nanoparticles across Cell and Tissue Barriers: A Review of Current Status and Future Prospects. Nano Today 2014, 9, 223–243. [Google Scholar] [CrossRef] [PubMed]
  204. Katifelis, H.; Mukha, I.; Bouziotis, P.; Vityuk, N.; Tsoukalas, C.; Lazaris, A.C.; Lyberopoulou, A.; Theodoropoulos, G.E.; Efstathopoulos, E.P.; Gazouli, M. Ag/Au Bimetallic Nanoparticles Inhibit Tumor Growth and Prevent Metastasis in a Mouse Model. Int. J. Nanomed. 2020, 15, 6019–6032. [Google Scholar] [CrossRef] [PubMed]
  205. Golombek, S.K.; May, J.N.; Theek, B.; Appold, L.; Drude, N.; Kiessling, F.; Lammers, T. Tumor targeting via EPR: Strategies to enhance patient responses. Adv. Drug Deliv. Rev. 2018, 130, 17–38. [Google Scholar] [CrossRef] [PubMed]
  206. Kaur, A.; Jain, S.; Tiwary, A. Mannan-coated gelatin nanoparticles for sustained and targeted delivery of didanosine: In vitro and in vivo evaluation. Acta Pharm. 2008, 58, 61–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Totten, J.D.; Wongpinyochit, T.; Carrola, J.; Duarte, I.F.; Seib, F.P. PEGylation-Dependent Metabolic Rewiring of Macrophages with Silk Fibroin Nanoparticles. ACS Appl. Mater. Interfaces 2019, 11, 14515–14525. [Google Scholar] [CrossRef] [Green Version]
  208. Leuschner, F.; Dutta, P.; Gorbatov, R.; Novobrantseva, T.I.; Donahoe, J.S.; Courties, G.; Lee, K.M.; Kim, J.I.; Markmann, J.F.; Marinelli, B.; et al. Therapeutic siRNA silencing in inflammatory monocytes in mice. Nat. Biotechnol. 2011, 29, 1005–1010. [Google Scholar] [CrossRef]
  209. Shen, S.; Zhang, Y.; Chen, K.G.; Luo, Y.L.; Wang, J. Cationic Polymeric Nanoparticle Delivering CCR2 siRNA to Inflammatory Monocytes for Tumor Microenvironment Modification and Cancer Therapy. Mol. Pharm. 2018, 15, 3642–3653. [Google Scholar] [CrossRef]
  210. Jose, A.; Labala, S.; Ninave, K.M.; Gade, S.K.; Venuganti, V.V.K. Effective Skin Cancer Treatment by Topical Co-delivery of Curcumin and STAT3 siRNA Using Cationic Liposomes. AAPS PharmSciTech 2018, 19, 166–175. [Google Scholar] [CrossRef]
  211. Tian, L.; Yi, X.; Dong, Z.; Xu, J.; Liang, C.; Chao, Y.; Wang, Y.; Yang, K.; Liu, Z. Calcium Bisphosphonate Nanoparticles with Chelator-Free Radiolabeling to Deplete Tumor-Associated Macrophages for Enhanced Cancer Radioisotope Therapy. ACS Nano 2018, 12, 11541–11551. [Google Scholar] [CrossRef]
  212. Zeisberger, S.M.; Odermatt, B.; Marty, C.; Zehnder-Fjällman, A.H.M.; Ballmer-Hofer, K.; Schwendener, R.A. Clodronate-liposome-mediated depletion of tumour-associated macrophages: A new and highly effective antiangiogenic therapy approach. Br. J. Cancer 2006, 95, 272–281. [Google Scholar] [CrossRef] [PubMed]
  213. Gazzaniga, S.; Bravo, A.I.; Guglielmotti, A.; van Rooijen, N.; Maschi, F.; Vecchi, A.; Mantovani, A.; Mordoh, J.; Wainstok, R. Targeting Tumor-Associated Macrophages and Inhibition of MCP-1 Reduce Angiogenesis and Tumor Growth in a Human Melanoma Xenograft. J. Investig. Dermatol. 2007, 127, 2031–2041. [Google Scholar] [CrossRef] [Green Version]
  214. Sousa, S.; Auriola, S.; Mönkkönen, J.; Määttä, J. Liposome encapsulated zoledronate favours M1-like behaviour in murine macrophages cultured with soluble factors from breast cancer cells. BMC Cancer 2015, 15, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Rajan, R.; Sabnani, M.K.; Mavinkurve, V.; Shmeeda, H.; Mansouri, H.; Bonkoungou, S.; Le, A.D.; Wood, L.M.; Gabizon, A.A.; La-Beck, N.M. Liposome-induced immunosuppression and tumor growth is mediated by macrophages and mitigated by liposome-encapsulated alendronate. J. Control. Release Off. J. Control. Release Soc. 2018, 271, 139–148. [Google Scholar] [CrossRef]
  216. Kwong, B.; Liu, H.; Irvine, D.J. Induction of potent anti-tumor responses while eliminating systemic side effects via liposome-anchored combinatorial immunotherapy. Biomaterials 2011, 32, 5134–5147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Jung, K.; Heishi, T.; Khan, O.F.; Kowalski, P.S.; Incio, J.; Rahbari, N.N.; Chung, E.; Clark, J.W.; Willett, C.G.; Luster, A.D.; et al. Ly6Clo monocytes drive immunosuppression and confer resistance to anti-VEGFR2 cancer therapy. J. Clin. Investig. 2017, 127, 3039–3051. [Google Scholar] [CrossRef] [PubMed]
  218. Wang, Y.; Lin, Y.X.; Qiao, S.L.; An, H.W.; Ma, Y.; Qiao, Z.Y.; Rajapaksha, R.P.; Wang, H. Polymeric nanoparticles promote macrophage reversal from M2 to M1 phenotypes in the tumor microenvironment. Biomaterials 2017, 112, 153–163. [Google Scholar] [CrossRef]
  219. van Woensel, M.; Mathivet, T.; Wauthoz, N.; Rosière, R.; Garg, A.D.; Agostinis, P.; Mathieu, V.; Kiss, R.; Lefranc, F.; Boon, L.; et al. Sensitization of glioblastoma tumor micro-environment to chemo- and immunotherapy by Galectin-1 intranasal knock-down strategy. Sci. Rep. 2017, 7, 1217. [Google Scholar] [CrossRef]
  220. Rodell, C.B.; Arlauckas, S.P.; Cuccarese, M.F.; Garris, C.S.; Li, R.; Ahmed, M.S.; Kohler, R.H.; Pittet, M.J.; Weissleder, R. TLR7/8-agonist-loaded nanoparticles promote the polarization of tumour-associated macrophages to enhance cancer immunotherapy. Nat. Biomed. Eng. 2018, 2, 578–588. [Google Scholar] [CrossRef]
  221. Zhao, J.; Zhang, Z.; Xue, Y.; Wang, G.; Cheng, Y.; Pan, Y.; Zhao, S.; Hou, Y. Anti-tumor macrophages activated by ferumoxytol combined or surface-functionalized with the TLR3 agonist poly (I: C) promote melanoma regression. Theranostics 2018, 8, 6307–6321. [Google Scholar] [CrossRef] [PubMed]
  222. VanHandel, M.; Alizadeh, D.; Zhang, L.; Kateb, B.; Bronikowski, M.; Manohara, H.; Badie, B. Selective uptake of multi-walled carbon nanotubes by tumor macrophages in a murine glioma model. J. Neuroimmunol. 2009, 208, 3–9. [Google Scholar] [CrossRef] [PubMed]
  223. Kwon, D.; Cha, B.G.; Cho, Y.; Min, J.; Park, E.-B.; Kang, S.-J.; Kim, J. Extra-Large Pore Mesoporous Silica Nanoparticles for Directing in Vivo M2 Macrophage Polarization by Delivering IL-4. Nano Lett. 2017, 17, 2747–2756. [Google Scholar] [CrossRef] [PubMed]
  224. Liaw, K.; Reddy, R.; Sharma, A.; Li, J.; Chang, M.; Sharma, R.; Salazar, S.; Kannan, S.; Kannan, R.M. Targeted systemic dendrimer delivery of CSF-1R inhibitor to tumor-associated macrophages improves outcomes in orthotopic glioblastoma. Bioeng. Transl. Med. 2021, 6, e10205. [Google Scholar] [CrossRef]
  225. Ramesh, A.; Brouillard, A.; Kumar, S.; Nandi, D.; Kulkarni, A. Dual inhibition of CSF1R and MAPK pathways using supramolecular nanoparticles enhances macrophage immunotherapy. Biomaterials 2020, 227, 119559. [Google Scholar] [CrossRef]
  226. Fuchs, A.K.; Syrovets, T.; Haas, K.A.; Loos, C.; Musyanovych, A.; Mailander, V.; Landfester, K.; Simmet, T. Carboxyl- and amino-functionalized polystyrene nanoparticles differentially affect the polarization profile of M1 and M2 macrophage subsets. Biomaterials 2016, 85, 78–87. [Google Scholar] [CrossRef]
  227. Yang, S.; Zhang, Y.; Lu, S.; Yang, L.; Yu, S.; Yang, H. CaCO3-Encapsulated Au Nanoparticles Modulate Macrophages toward M1-like Phenotype. ACS Appl. Bio Mater. 2021, 4, 3214–3223. [Google Scholar] [CrossRef]
  228. Etzerodt, A.; Tsalkitzi, K.; Maniecki, M.; Damsky, W.; Delfini, M.; Baudoin, E.; Moulin, M.; Bosenberg, M.; Graversen, J.H.; Auphan-Anezin, N.; et al. Specific targeting of CD163(+) TAMs mobilizes inflammatory monocytes and promotes T cell-mediated tumor regression. J. Exp. Med. 2019, 216, 2394–2411. [Google Scholar] [CrossRef] [PubMed]
  229. Yu, S.S.; Lau, C.M.; Barham, W.J.; Onishko, H.M.; Nelson, C.E.; Li, H.; Smith, C.A.; Yull, F.E.; Duvall, C.L.; Giorgio, T.D. Macrophage-Specific RNA Interference Targeting via “Click”, Mannosylated Polymeric Micelles. Mol. Pharm. 2013, 10, 975–987. [Google Scholar] [CrossRef] [Green Version]
  230. Shi, C.; Liu, T.; Guo, Z.; Zhuang, R.; Zhang, X.; Chen, X. Reprogramming Tumor-Associated Macrophages by Nanoparticle-Based Reactive Oxygen Species Photogeneration. Nano Lett. 2018, 18, 7330–7342. [Google Scholar] [CrossRef]
  231. Zhu, S.; Niu, M.; O’Mary, H.; Cui, Z. Targeting of tumor-associated macrophages made possible by PEG-sheddable, mannose-modified nanoparticles. Mol. Pharm. 2013, 10, 3525–3530. [Google Scholar] [CrossRef] [Green Version]
  232. Leber, N.; Kaps, L.; Yang, A.; Aslam, M.; Giardino, M.; Klefenz, A.; Choteschovsky, N.; Rosigkeit, S.; Mostafa, A.; Nuhn, L.; et al. Alpha-Mannosyl-Functionalized Cationic Nanohydrogel Particles for Targeted Gene Knockdown in Immunosuppressive Macrophages. Macromol. Biosci. 2019, 19, e1900162. [Google Scholar] [CrossRef]
  233. Niu, M.; Naguib, Y.W.; Aldayel, A.M.; Shi, Y.C.; Hursting, S.D.; Hersh, M.A.; Cui, Z. Biodistribution and in vivo activities of tumor-associated macrophage-targeting nanoparticles incorporated with doxorubicin. Mol. Pharm. 2014, 11, 4425–4436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Hatami, E.; Mu, Y.; Shields, D.N.; Chauhan, S.C.; Kumar, S.; Cory, T.J.; Yallapu, M. Mannose-decorated hybrid nanoparticles for enhanced macrophage targeting. Biochem. Biophys. Rep. 2019, 17, 197–207. [Google Scholar] [CrossRef] [PubMed]
  235. Chiang, C.F.; Chao, T.T.; Su, Y.F.; Hsu, C.C.; Chien, C.Y.; Chiu, K.C.; Shiah, S.G.; Lee, C.H.; Liu, S.Y.; Shieh, Y.S. Metformin-treated cancer cells modulate macrophage polarization through AMPK-NF-κB signaling. Oncotarget 2017, 8, 20706–20718. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Wei, Z.; Zhang, X.; Yong, T.; Bie, N.; Zhan, G.; Li, X.; Liang, Q.; Li, J.; Yu, J.; Huang, G.; et al. Boosting anti-PD-1 therapy with metformin-loaded macrophage-derived microparticles. Nat. Commun. 2021, 12, 440. [Google Scholar] [CrossRef]
  237. Wang, Y.; Guo, G.; Feng, Y.; Long, H.; Ma, D.-L.; Leung, C.-H.; Dong, L.; Wang, C. A tumour microenvironment-responsive polymeric complex for targeted depletion of tumour-associated macrophages (TAMs). J. Mater. Chem. B 2017, 5, 7307–7318. [Google Scholar] [CrossRef] [PubMed]
  238. Maalej, M.; Forgione, R.E.; Marchetti, R.; Bulteau, F.; Thépaut, M.; Lanzetta, R.; Laguri, C.; Simorre, J.P.; Fieschi, F.; Molinaro, A.; et al. Human Macrophage Galactose-Type Lectin (MGL) Recognizes the Outer Core of Escherichia coli Lipooligosaccharide. ChemBioChem 2019, 20, 1778–1782. [Google Scholar] [PubMed]
  239. Han, S.; Wang, W.; Wang, S.; Wang, S.; Ju, R.; Pan, Z.; Yang, T.; Zhang, G.; Wang, H.; Wang, L. Multifunctional biomimetic nanoparticles loading baicalin for polarizing tumor-associated macrophages. Nanoscale 2019, 11, 20206–20220. [Google Scholar] [CrossRef] [PubMed]
  240. He, M.; Xu, Z.; Ding, T.; Kuang, D.M.; Zheng, L. MicroRNA-155 regulates inflammatory cytokine production in tumor-associated macrophages via targeting C/EBPbeta. Cell Mol. Immunol. 2009, 6, 343–352. [Google Scholar] [CrossRef] [Green Version]
  241. Cai, X.; Yin, Y.; Li, N.; Zhu, D.; Zhang, J.; Zhang, C.Y.; Zen, K. Re-polarization of tumor-associated macrophages to pro-inflammatory M1 macrophages by microRNA-155. J. Mol. Cell Biol. 2012, 4, 341–343. [Google Scholar] [CrossRef] [PubMed]
  242. Liu, L.; Yi, H.; He, H.; Pan, H.; Cai, L.; Ma, Y. Tumor associated macrophage-targeted microRNA delivery with dual-responsive polypeptide nanovectors for anti-cancer therapy. Biomaterials 2017, 134, 166–179. [Google Scholar] [CrossRef] [PubMed]
  243. Huang, Z.; Zhang, Z.; Jiang, Y.; Zhang, D.; Chen, J.; Dong, L.; Zhang, J. Targeted delivery of oligonucleotides into tumor-associated macrophages for cancer immunotherapy. J. Control Release 2012, 158, 286–292. [Google Scholar] [CrossRef]
  244. Su, L.; Zhang, W.; Wu, X.; Zhang, Y.; Chen, X.; Liu, G.; Chen, G.; Jiang, M. Glycocalyx-Mimicking Nanoparticles for Stimulation and Polarization of Macrophages via Specific Interactions. Small 2015, 11, 4191–4200. [Google Scholar] [CrossRef]
  245. Cieslewicz, M.; Tang, J.; Yu, J.L.; Cao, H.; Zavaljevski, M.; Motoyama, K.; Lieber, A.; Raines, E.W.; Pun, S.H. Targeted delivery of proapoptotic peptides to tumor-associated macrophages improves survival. Proc. Natl. Acad. Sci. USA 2013, 110, 15919–15924. [Google Scholar] [CrossRef] [Green Version]
  246. Pang, L.; Pei, Y.; Uzunalli, G.; Hyun, H.; Lyle, L.T.; Yeo, Y. Surface Modification of Polymeric Nanoparticles with M2pep Peptide for Drug Delivery to Tumor-Associated Macrophages. Pharm. Res. 2019, 36, 65. [Google Scholar] [CrossRef]
  247. Zhou, S.; Zhang, T.; Peng, B.; Luo, X.; Liu, X.; Hu, L.; Liu, Y.; Di, D.; Song, Y.; Deng, Y. Targeted delivery of epirubicin to tumor-associated macrophages by sialic acid-cholesterol conjugate modified liposomes with improved antitumor activity. Int. J. Pharm. 2017, 523, 203–216. [Google Scholar] [CrossRef]
  248. Zhang, X.; Tian, W.; Cai, X.; Wang, X.; Dang, W.; Tang, H.; Cao, H.; Wang, L.; Chen, T. Hydrazinocurcumin Encapsuled nanoparticles “re-educate” tumor-associated macrophages and exhibit anti-tumor effects on breast cancer following STAT3 suppression. PLoS ONE 2013, 8, e65896. [Google Scholar] [CrossRef]
  249. Chen, C.; Song, M.; Du, Y.; Yu, Y.; Li, C.; Han, Y.; Yan, F.; Shi, Z.; Feng, S. Tumor-Associated-Macrophage-Membrane-Coated Nanoparticles for Improved Photodynamic Immunotherapy. Nano Lett. 2021, 21, 5522–5531. [Google Scholar] [CrossRef]
  250. Brase, J.C.; Wuttig, D.; Kuner, R.; Sültmann, H. Serum microRNAs as non-invasive biomarkers for cancer. Mol. Cancer 2010, 9, 306. [Google Scholar] [CrossRef] [Green Version]
  251. Trivedi, M.; Talekar, M.; Shah, P.; Ouyang, Q.; Amiji, M. Modification of tumor cell exosome content by transfection with wt-p53 and microRNA-125b expressing plasmid DNA and its effect on macrophage polarization. Oncogenesis 2016, 5, e250. [Google Scholar] [CrossRef] [Green Version]
  252. Huis In ‘t Veld, R.V.; Ritsma, L.; Kleinovink, J.W.; Que, I.; Ossendorp, F.; Cruz, L.J. Photodynamic cancer therapy enhances accumulation of nanoparticles in tumor-associated myeloid cells. J. Control Release 2020, 320, 19–31. [Google Scholar] [CrossRef]
  253. Ortega, R.A.; Barham, W.; Sharman, K.; Tikhomirov, O.; Giorgio, T.D.; Yull, F.E. Manipulating the NF-kappaB pathway in macrophages using mannosylated, siRNA-delivering nanoparticles can induce immunostimulatory and tumor cytotoxic functions. Int. J. Nanomed. 2016, 11, 2163–2177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Ray, M.; Lee, Y.-W.; Hardie, J.; Mout, R.; Yeşilbag Tonga, G.; Farkas, M.E.; Rotello, V.M. CRISPRed Macrophages for Cell-Based Cancer Immunotherapy. Bioconjug. Chem. 2018, 29, 445–450. [Google Scholar] [CrossRef]
  255. Lee, Y.-W.; Mout, R.; Luther, D.C.; Liu, Y.; Castellanos-García, L.; Burnside, A.S.; Ray, M.; Tonga, G.Y.; Hardie, J.; Nagaraj, H.; et al. In Vivo Editing of Macrophages through Systemic Delivery of CRISPR-Cas9-Ribonucleoprotein-Nanoparticle Nanoassemblies. Adv. Ther. 2019, 2, 1900041. [Google Scholar] [CrossRef]
  256. Liu, Y.; Wang, J.; Zhang, J.; Marbach, S.; Xu, W.; Zhu, L. Targeting Tumor-Associated Macrophages by MMP2-Sensitive Apoptotic Body-Mimicking Nanoparticles. ACS Appl. Mater. Interfaces 2020, 12, 52402–52414. [Google Scholar] [CrossRef]
  257. Alupei, M.C.; Licarete, E.; Patras, L.; Banciu, M. Liposomal simvastatin inhibits tumor growth via targeting tumor-associated macrophages-mediated oxidative stress. Cancer Lett. 2015, 356, 946–952. [Google Scholar] [CrossRef]
  258. Hou, T.; Wang, T.; Mu, W.; Yang, R.; Liang, S.; Zhang, Z.; Fu, S.; Gao, T.; Liu, Y.; Zhang, N. Nanoparticle-Loaded Polarized-Macrophages for Enhanced Tumor Targeting and Cell-Chemotherapy. Nanomicro Lett. 2020, 13, 6. [Google Scholar] [PubMed]
  259. Kim, M.S.; Lee, J.S.; Kim, J.E.; Kim, J.-W.; Bok, S.; Keum, K.C.; Koh, W.-G.; Koom, W.S. Enhancement of antitumor effect of radiotherapy via combination with Au@SiO2 nanoparticles targeted to tumor-associated macrophages. J. Ind. Eng. Chem. 2020, 84, 349–357. [Google Scholar] [CrossRef]
  260. He, X.Y.; Liu, B.Y.; Xu, C.; Zhuo, R.X.; Cheng, S.X. A multi-functional macrophage and tumor targeting gene delivery system for the regulation of macrophage polarity and reversal of cancer immunoresistance. Nanoscale 2018, 10, 15578–15587. [Google Scholar] [CrossRef]
  261. Song, M.; Liu, T.; Shi, C.; Zhang, X.; Chen, X. Bioconjugated Manganese Dioxide Nanoparticles Enhance Chemotherapy Response by Priming Tumor-Associated Macrophages toward M1-like Phenotype and Attenuating Tumor Hypoxia. ACS Nano 2016, 10, 633–647. [Google Scholar] [CrossRef] [Green Version]
  262. Esser, A.K.; Ross, M.H.; Fontana, F.; Su, X.; Gabay, A.; Fox, G.C.; Xu, Y.; Xiang, J.; Schmieder, A.H.; Yang, X.; et al. Nanotherapy delivery of c-myc inhibitor targets Protumor Macrophages and preserves Antitumor Macrophages in Breast Cancer. Theranostics 2020, 10, 7510–7526. [Google Scholar] [CrossRef]
  263. Chang, C.C.; Dinh, T.K.; Lee, Y.A.; Wang, F.N.; Sung, Y.C.; Yu, P.L.; Chiu, S.C.; Shih, Y.C.; Wu, C.Y.; Huang, Y.D.; et al. Nanoparticle Delivery of MnO2 and Antiangiogenic Therapy to Overcome Hypoxia-Driven Tumor Escape and Suppress Hepatocellular Carcinoma. ACS Appl. Mater. Interfaces 2020, 12, 44407–44419. [Google Scholar] [CrossRef] [PubMed]
  264. Korangath, P.; Barnett, J.D.; Sharma, A.; Henderson, E.T.; Stewart, J.; Yu, S.-H.; Kandala, S.K.; Yang, C.-T.; Caserto, J.S.; Hedayati, M.; et al. Nanoparticle interactions with immune cells dominate tumor retention and induce T cell–mediated tumor suppression in models of breast cancer. Sci. Adv. 2020, 6, eaay1601. [Google Scholar] [CrossRef] [Green Version]
  265. Wang, J.; Lee, J.S.; Kim, D.; Zhu, L. Exploration of Zinc Oxide Nanoparticles as a Multitarget and Multifunctional Anticancer Nanomedicine. ACS Appl. Mater. Interfaces 2017, 9, 39971–39984. [Google Scholar] [CrossRef] [PubMed]
  266. Zhang, X.; Tian, G.; Yin, W.; Wang, L.; Zheng, X.; Yan, L.; Li, J.; Su, H.; Chen, C.; Gu, Z.; et al. Controllable Generation of Nitric Oxide by Near-Infrared-Sensitized Upconversion Nanoparticles for Tumor Therapy. Adv. Funct. Mater. 2015, 25, 3049–3056. [Google Scholar] [CrossRef]
  267. Leonard, F.; Curtis, L.T.; Ware, M.J.; Nosrat, T.; Liu, X.; Yokoi, K.; Frieboes, H.B.; Godin, B. Macrophage Polarization Contributes to the Anti-Tumoral Efficacy of Mesoporous Nanovectors Loaded with Albumin-Bound Paclitaxel. Front. Immunol. 2017, 8, 693. [Google Scholar] [CrossRef] [Green Version]
  268. Mocan, T.; Matea, C.; Tabaran, F.; Iancu, C.; Orasan, R.; Mocan, L. In Vitro Administration of Gold Nanoparticles Functionalized with MUC-1 Protein Fragment Generates Anticancer Vaccine Response via Macrophage Activation and Polarization Mechanism. J. Cancer 2015, 6, 583–592. [Google Scholar] [CrossRef]
  269. Dong, J.; Ma, Q. Macrophage polarization and activation at the interface of multi-walled carbon nanotube-induced pulmonary inflammation and fibrosis. Nanotoxicology 2018, 12, 153–168. [Google Scholar] [CrossRef]
  270. Noonan, D.M.; Principi, E.; Girardello, R.; Bruno, A.; Manni, I.; Gini, E.; Pagani, A.; Grimaldi, A.; Ivaldi, F.; Congiu, T.; et al. Systemic distribution of single-walled carbon nanotubes in a novel model: Alteration of biochemical parameters, metabolic functions, liver accumulation, and inflammation in vivo. Int. J. Nanomed. 2016, 11, 4299–4316. [Google Scholar] [CrossRef] [Green Version]
  271. Tao, Y.; Ju, E.; Ren, J.; Qu, X. Immunostimulatory oligonucleotides-loaded cationic graphene oxide with photothermally enhanced immunogenicity for photothermal/immune cancer therapy. Biomaterials 2014, 35, 9963–9971. [Google Scholar] [CrossRef]
  272. Tang, J.; Chen, Z.; Sun, B.; Dong, J.; Liu, J.; Zhou, H.; Wang, L.; Bai, R.; Miao, Q.; Zhao, Y.; et al. Polyhydroxylated fullerenols regulate macrophage for cancer adoptive immunotherapy and greatly inhibit the tumor metastasis. Nanomedicine 2016, 12, 945–954. [Google Scholar] [CrossRef] [PubMed]
  273. Liang, D.-S.; Wen, Z.-J.; Wang, J.-H.; Zhu, F.-F.; Guo, F.; Zhou, J.-L.; Xu, J.-J.; Zhong, H.-J. Legumain protease-sheddable PEGylated, tuftsin-modified nanoparticles for selective targeting to tumor-associated macrophages. J. Drug Target. 2021, 1–25. [Google Scholar]
  274. Duan, L.; Ouyang, K.; Xu, X.; Xu, L.; Wen, C.; Zhou, X.; Qin, Z.; Xu, Z.; Sun, W.; Liang, Y. Nanoparticle Delivery of CRISPR/Cas9 for Genome Editing. Front. Genet. 2021, 12, 673286. [Google Scholar] [CrossRef]
  275. Li, S.D.; Chen, Y.C.; Hackett, M.J.; Huang, L. Tumor-targeted delivery of siRNA by self-assembled nanoparticles. Mol. Ther. 2008, 16, 163–169. [Google Scholar] [CrossRef]
  276. Cardoso, M.M.; Peca, N.I.; Roque, C.A. A Antibody-Conjugated Nanoparticles for Therapeutic Applications. Curr. Med. Chem. 2012, 19, 3103–3127. [Google Scholar] [CrossRef]
  277. Ngoune, R.; Peters, A.; von Elverfeldt, D.; Winkler, K.; Putz, G. Accumulating nanoparticles by EPR: A route of no return. J. Control Release 2016, 238, 58–70. [Google Scholar] [CrossRef] [Green Version]
  278. Huis In ‘t Veld, R.V.; Da Silva, C.G.; Jager, M.J.; Cruz, L.J.; Ossendorp, F. Combining Photodynamic Therapy with Immunostimulatory Nanoparticles Elicits Effective Anti-Tumor Immune Responses in Preclinical Murine Models. Pharmaceutics 2021, 13, 1470. [Google Scholar] [CrossRef]
  279. Harrington, K.J.; Mohammadtaghi, S.; Uster, P.S.; Glass, D.; Peters, A.M.; Vile, R.G.; Stewart, J.S.W. Effective Targeting of Solid Tumors in Patients With Locally Advanced Cancers by Radiolabeled Pegylated Liposomes. Clin. Cancer Res. 2021, 7, 243–254. [Google Scholar]
  280. Venneri, M.A.; De Palma, M.; Ponzoni, M.; Pucci, F.; Scielzo, C.; Zonari, E.; Mazzieri, R.; Doglioni, C.; Naldini, L. Identification of proangiogenic TIE2-expressing monocytes (TEMs) in human peripheral blood and cancer. Blood 2007, 109, 5276–5285. [Google Scholar] [CrossRef] [Green Version]
  281. Rohan, T.E.; Xue, X.; Lin, H.M.; D’Alfonso, T.M.; Ginter, P.S.; Oktay, M.H.; Robinson, B.D.; Ginsberg, M.; Gertler, F.B.; Glass, A.G.; et al. Tumor microenvironment of metastasis and risk of distant metastasis of breast cancer. J. Natl. Cancer Inst. 2014, 106, 106. [Google Scholar] [CrossRef] [Green Version]
  282. Másson, M.; Loftsson, T.; Másson, G.; Stefánsson, E. Cyclodextrins as permeation enhancers: Some theoretical evaluations and in vitro testing. J. Control. Release 1999, 59, 107–118. [Google Scholar] [CrossRef]
  283. Cruz, L.J.; van Dijk, T.; Vepris, O.; Li, T.; Schomann, T.; Baldazzi, F.; Kurita, R.; Nakamura, Y.; Grosveld, F.; Philipsen, S.; et al. PLGA-Nanoparticles for Intracellular Delivery of the CRISPR-Complex to Elevate Fetal Globin Expression in Erythroid Cells. Biomaterials 2021, 268, 120580. [Google Scholar] [CrossRef] [PubMed]
  284. Maeda, H.; Bharate, G.Y.; Daruwalla, J. Polymeric drugs for efficient tumor-targeted drug delivery based on EPR-effect. Eur. J. Pharm. Biopharm. 2009, 71, 409–419. [Google Scholar] [CrossRef]
  285. Cruz, L.J.; Rueda, F.; Tacken, P.; Albericio, F.; Torensma, R.; Figdor, C.G. Enhancing immunogenicity and cross-reactivity of HIV-1 antigens by in vivo targeting to dendritic cells. Nanomedicine 2012, 7, 1591–1610. [Google Scholar] [CrossRef] [PubMed]
  286. Tacken, P.J.; Zeelenberg, I.S.; Cruz, L.J.; van Hout-Kuijer, M.A.; van de Glind, G.; Fokkink, R.G.; Lambeck, A.J.; Figdor, C.G. Targeted delivery of TLR ligands to human and mouse dendritic cells strongly enhances adjuvanticity. Blood 2011, 118, 6836–6844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Lewis, C.E.; Harney, A.S.; Pollard, J.W. The Multifaceted Role of Perivascular Macrophages in Tumors. Cancer Cell 2016, 30, 18–25. [Google Scholar] [CrossRef] [Green Version]
  288. Zhao, Z.; Zheng, L.; Chen, W.; Weng, W.; Song, J.; Ji, J. Delivery strategies of cancer immunotherapy: Recent advances and future perspectives. J. Hematol. Oncol. 2019, 12, 126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  289. Shmeeda, H.; Amitay, Y.; Tzemach, D.; Gorin, J.; Gabizon, A. Liposome encapsulation of zoledronic acid results in major changes in tissue distribution and increase in toxicity. J. Control Release 2013, 167, 265–275. [Google Scholar] [CrossRef] [PubMed]
  290. Mantovani, A.; Marchesi, F.; Malesci, A.; Laghi, L.; Allavena, P. Tumour-associated macrophages as treatment targets in oncology. Nat. Rev. Clin. Oncol. 2017, 14, 399–416. [Google Scholar] [CrossRef]
  291. Condeelis, J.; Pollard, J.W. Macrophages: Obligate partners for tumor cell migration, invasion, and metastasis. Cell 2006, 124, 263–266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  292. von Tresckow, B.; Morschhauser, F.; Ribrag, V.; Topp, M.S.; Chien, C.; Seetharam, S.; Aquino, R.; Kotoulek, S.; de Boer, C.J.; Engert, A. An Open-Label, Multicenter, Phase I/II Study of JNJ-40346527, a CSF-1R Inhibitor, in Patients with Relapsed or Refractory Hodgkin Lymphoma. Clin. Cancer Res. 2015, 21, 1843–1850. [Google Scholar] [CrossRef] [Green Version]
  293. Butowski, N.; Colman, H.; De Groot, J.F.; Omuro, A.M.; Nayak, L.; Wen, P.Y.; Cloughesy, T.F.; Marimuthu, A.; Haidar, S.; Perry, A.; et al. Orally administered colony stimulating factor 1 receptor inhibitor PLX3397 in recurrent glioblastoma: An Ivy Foundation Early Phase Clinical Trials Consortium phase II study. Neuro. Oncol. 2016, 18, 557–564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Da Silva, C.G.; Camps, M.G.M.; Li, T.; Chan, A.B.; Ossendorp, F.; Cruz, L.J. Co-delivery of immunomodulators in biodegradable nanoparticles improves therapeutic efficacy of cancer vaccines. Biomaterials 2019, 220, 119417. [Google Scholar] [CrossRef]
  295. Basu, A.; Kodumudi, K. Multimodal approaches to improve immunotherapy in breast cancer. Immunotherapy 2020, 12, 161–165. [Google Scholar] [CrossRef]
  296. Malesci, A.; Bianchi, P.; Celesti, G.; Basso, G.; Marchesi, F.; Grizzi, F.; Di Caro, G.; Cavalleri, T.; Rimassa, L.; Palmqvist, R.; et al. Tumor-associated macrophages and response to 5-fluorouracil adjuvant therapy in stage III colorectal cancer. OncoImmunology 2017, 6, e1342918. [Google Scholar] [CrossRef]
  297. Fang, Z.; Shen, Y.; Gao, D. Stimulus-responsive nanocarriers for targeted drug delivery. New J. Chem. 2021, 45, 4534–4544. [Google Scholar] [CrossRef]
Figure 1. Applications of nanoparticles (NPs) in the field of biomedicine. NPs can be used for in vitro and in vivo targeted drug delivery, therapy and imaging. Multimodal imaging refers to the production of signals that can be detected by more than one imaging technique, for example, the combination of magnetic and radioactive substances to be detected by PET/MRI or PET/CT. Many NP formulations are multimodal and multifunctional. The latter refers to NPs that combine multiple objectives, such therapy and imaging. This figure was prepared using a template on the Servier medical art website (http://www.servier.fr/servier-medical-art, accessed on 10 June 2021).
Figure 1. Applications of nanoparticles (NPs) in the field of biomedicine. NPs can be used for in vitro and in vivo targeted drug delivery, therapy and imaging. Multimodal imaging refers to the production of signals that can be detected by more than one imaging technique, for example, the combination of magnetic and radioactive substances to be detected by PET/MRI or PET/CT. Many NP formulations are multimodal and multifunctional. The latter refers to NPs that combine multiple objectives, such therapy and imaging. This figure was prepared using a template on the Servier medical art website (http://www.servier.fr/servier-medical-art, accessed on 10 June 2021).
Pharmaceutics 13 01670 g001
Figure 2. The origin and polarization state of tumor-associated macrophages (TAMs) in the tumor microenvironment (TME). (A) Origin: TAMs in the tumor are largely derived from blood monocytes that have their origin in hematopoietic stem cells in the bone marrow, or, at minor contributions, from locally derived, tissue-resident macrophages with their origin in progenitor cells generated in the fetal liver and yolk sac during embryonic development. (B) Tumor-derived factors: tumor cells can produce a variety of factors (such as CCL2/3/4, CSF1, IL-4 and IL-10) that act on the polarization of TAMs. (C) Immune microenvironment: Treg-derived IL-10, MDSC-derived TGFβ, B-cell-derived Ig, Th2-derived IL-4 and NK-cell-derived IL-13 enhance the M2-like (immunosuppressive) phenotype of TAMs. (D) Homeostatic imbalance: Hypoxia promotes the malignant transformation of tumors under the action of hypoxia-inducible factors. Tumor-derived ECM and HMGB1 increase the number of TAMs. The abbreviations are CSF1, colony-stimulating factor 1; CSF1R, CSF1 receptor; ECM, tumor-derived extracellular matrix; HMGB1, high-mobility group box 1 protein; Tregs, regulatory T cells; Ig, immunoglobulin; TGF, transforming growth factor; IL, interleukin; TNF, tumor necrosis factor; VEGF, vascular endothelial growth factor; VEGFR, VEGF receptor; MDSCs, myeloid suppressor cells; TLR, Toll-like receptor; CCL, chemotaxis chemokine ligand; CXCL, chemokine; and CXCR, chemokine receptor. This figure was prepared using a template on the Servier medical art website (http://www.servier.fr/servier-medical-art, accessed on 10 June 2021).
Figure 2. The origin and polarization state of tumor-associated macrophages (TAMs) in the tumor microenvironment (TME). (A) Origin: TAMs in the tumor are largely derived from blood monocytes that have their origin in hematopoietic stem cells in the bone marrow, or, at minor contributions, from locally derived, tissue-resident macrophages with their origin in progenitor cells generated in the fetal liver and yolk sac during embryonic development. (B) Tumor-derived factors: tumor cells can produce a variety of factors (such as CCL2/3/4, CSF1, IL-4 and IL-10) that act on the polarization of TAMs. (C) Immune microenvironment: Treg-derived IL-10, MDSC-derived TGFβ, B-cell-derived Ig, Th2-derived IL-4 and NK-cell-derived IL-13 enhance the M2-like (immunosuppressive) phenotype of TAMs. (D) Homeostatic imbalance: Hypoxia promotes the malignant transformation of tumors under the action of hypoxia-inducible factors. Tumor-derived ECM and HMGB1 increase the number of TAMs. The abbreviations are CSF1, colony-stimulating factor 1; CSF1R, CSF1 receptor; ECM, tumor-derived extracellular matrix; HMGB1, high-mobility group box 1 protein; Tregs, regulatory T cells; Ig, immunoglobulin; TGF, transforming growth factor; IL, interleukin; TNF, tumor necrosis factor; VEGF, vascular endothelial growth factor; VEGFR, VEGF receptor; MDSCs, myeloid suppressor cells; TLR, Toll-like receptor; CCL, chemotaxis chemokine ligand; CXCL, chemokine; and CXCR, chemokine receptor. This figure was prepared using a template on the Servier medical art website (http://www.servier.fr/servier-medical-art, accessed on 10 June 2021).
Pharmaceutics 13 01670 g002
Figure 3. The role of tumor-associated macrophages (TAMs) in tumor progression. Macrophages are recruited into tumors under the action of chemokines and growth factors and gradually become TAMs. TAMs can be activated as (A) M1 macrophages with pro-inflammatory effects (antitumor) and (B) M2 macrophages with non-inflammatory properties (protumor). Among them, a variety of factors expressed and secreted by M2-like macrophages (such as VEGF, TNF-α, matrix metalloproteinase (MMP), CCL2 and CSF1) can induce tumor growth, migration, invasion, immunosuppression, treatment resistance and intravascular invasion. For example, M2 macrophages promote tumor migration and invasion through EGF, TNF, CSF-1 and MMPs; promote angiogenesis and tumor growth through the secretion of VEGF, MCSF, IL10 and TGF-β; promote immunosuppression through the secretion of CCL17/18/22, CSF -1, IL-10, TGF-β and PEG; promote tumor metastasis by secreting CSF-1, IL-1 and MMPs; and induce cancer cell proliferation through NF-κB or STAT3 signaling pathways. In contrast, the “classical activated” M1 macrophages are immune-stimulating, induce antitumor immunity and kill intracellular pathogens by the actions of IL-1, IL-6, IL-12, IL23, TNF-α, NOS, ROI and RNI. CTL, cytotoxic T lymphocytes; CAF, cancer-related fibroblasts; CSF1, colony-stimulating factor 1; TGF, transforming growth factor; IL, interleukin; TNF, tumor necrosis factor; MMP, matrix metalloproteinase; TLR, toll-type receptor; VEGF, vascular endothelial growth factor; LPS, lipopolysaccharide; NOS, inducible nitric oxide synthase; RNI, reactive nitrogen intermediates; and ROI, reactive oxygen intermediates. This figure was prepared using a template on the Servier medical art website (http://www.servier.fr/servier-medical-art, accessed on 10 July 2021).
Figure 3. The role of tumor-associated macrophages (TAMs) in tumor progression. Macrophages are recruited into tumors under the action of chemokines and growth factors and gradually become TAMs. TAMs can be activated as (A) M1 macrophages with pro-inflammatory effects (antitumor) and (B) M2 macrophages with non-inflammatory properties (protumor). Among them, a variety of factors expressed and secreted by M2-like macrophages (such as VEGF, TNF-α, matrix metalloproteinase (MMP), CCL2 and CSF1) can induce tumor growth, migration, invasion, immunosuppression, treatment resistance and intravascular invasion. For example, M2 macrophages promote tumor migration and invasion through EGF, TNF, CSF-1 and MMPs; promote angiogenesis and tumor growth through the secretion of VEGF, MCSF, IL10 and TGF-β; promote immunosuppression through the secretion of CCL17/18/22, CSF -1, IL-10, TGF-β and PEG; promote tumor metastasis by secreting CSF-1, IL-1 and MMPs; and induce cancer cell proliferation through NF-κB or STAT3 signaling pathways. In contrast, the “classical activated” M1 macrophages are immune-stimulating, induce antitumor immunity and kill intracellular pathogens by the actions of IL-1, IL-6, IL-12, IL23, TNF-α, NOS, ROI and RNI. CTL, cytotoxic T lymphocytes; CAF, cancer-related fibroblasts; CSF1, colony-stimulating factor 1; TGF, transforming growth factor; IL, interleukin; TNF, tumor necrosis factor; MMP, matrix metalloproteinase; TLR, toll-type receptor; VEGF, vascular endothelial growth factor; LPS, lipopolysaccharide; NOS, inducible nitric oxide synthase; RNI, reactive nitrogen intermediates; and ROI, reactive oxygen intermediates. This figure was prepared using a template on the Servier medical art website (http://www.servier.fr/servier-medical-art, accessed on 10 July 2021).
Pharmaceutics 13 01670 g003
Figure 4. The main strategies for targeting tumor-associated macrophages (TAMs) in cancer therapy. (A) TAM reprogramming is mainly achieved by modulating signaling pathways using agonists of TLRs and CD40, inhibitors of NF-κB, PI3Kγ, STAT and HDAC, and si/miRNA. (B) Depletion of TAMs. TAM depletion strategies focus on using anti-neoplastic drugs, such trabectedin, and monoclonal antibodies that block signaling axes, including CCL2/CCR2, CSF-1/CSF-1R, CXCL12/CXCR4, or the interaction of SIRP1α-CD47. This figure was prepared using a template on the Servier medical art website (http://www.servier.fr/servier-medical-art, accessed on 10 July 2021).
Figure 4. The main strategies for targeting tumor-associated macrophages (TAMs) in cancer therapy. (A) TAM reprogramming is mainly achieved by modulating signaling pathways using agonists of TLRs and CD40, inhibitors of NF-κB, PI3Kγ, STAT and HDAC, and si/miRNA. (B) Depletion of TAMs. TAM depletion strategies focus on using anti-neoplastic drugs, such trabectedin, and monoclonal antibodies that block signaling axes, including CCL2/CCR2, CSF-1/CSF-1R, CXCL12/CXCR4, or the interaction of SIRP1α-CD47. This figure was prepared using a template on the Servier medical art website (http://www.servier.fr/servier-medical-art, accessed on 10 July 2021).
Pharmaceutics 13 01670 g004
Figure 5. Cellular internalization pathways of nanoparticles (NPs). Depending on their size, NPs enter cells through mechanisms of caveolin- and clathrin-mediated endocytosis, macropinocytosis, phagocytosis or pinocytosis.
Figure 5. Cellular internalization pathways of nanoparticles (NPs). Depending on their size, NPs enter cells through mechanisms of caveolin- and clathrin-mediated endocytosis, macropinocytosis, phagocytosis or pinocytosis.
Pharmaceutics 13 01670 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

He, Y.; de Araújo Júnior, R.F.; Cruz, L.J.; Eich, C. Functionalized Nanoparticles Targeting Tumor-Associated Macrophages as Cancer Therapy. Pharmaceutics 2021, 13, 1670. https://doi.org/10.3390/pharmaceutics13101670

AMA Style

He Y, de Araújo Júnior RF, Cruz LJ, Eich C. Functionalized Nanoparticles Targeting Tumor-Associated Macrophages as Cancer Therapy. Pharmaceutics. 2021; 13(10):1670. https://doi.org/10.3390/pharmaceutics13101670

Chicago/Turabian Style

He, Yuanyuan, Raimundo Fernandes de Araújo Júnior, Luis J. Cruz, and Christina Eich. 2021. "Functionalized Nanoparticles Targeting Tumor-Associated Macrophages as Cancer Therapy" Pharmaceutics 13, no. 10: 1670. https://doi.org/10.3390/pharmaceutics13101670

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