Skip to main content

Generation, evolution, interfering factors, applications, and challenges of patient-derived xenograft models in immunodeficient mice

Abstract

Establishing appropriate preclinical models is essential for cancer research. Evidence suggests that cancer is a highly heterogeneous disease. This follows the growing use of cancer models in cancer research to avoid these differences between xenograft tumor models and patient tumors. In recent years, a patient-derived xenograft (PDX) tumor model has been actively generated and applied, which preserves both cell–cell interactions and the microenvironment of tumors by directly transplanting cancer tissue from tumors into immunodeficient mice. In addition to this, the advent of alternative hosts, such as zebrafish hosts, or in vitro models (organoids and microfluidics), has also facilitated the advancement of cancer research. However, they still have a long way to go before they become reliable models. The development of immunodeficient mice has enabled PDX to become more mature and radiate new vitality. As one of the most reliable and standard preclinical models, the PDX model in immunodeficient mice (PDX-IM) exerts important effects in drug screening, biomarker development, personalized medicine, co-clinical trials, and immunotherapy. Here, we focus on the development procedures and application of PDX-IM in detail, summarize the implications that the evolution of immunodeficient mice has brought to PDX-IM, and cover the key issues in developing PDX-IM in preclinical studies.

Introduction

Cancer, as a disease affecting human life and health worldwide, has been of high concern [1]. Among various treatments for cancer, chemotherapy has become an increasingly mature therapy, unfortunately the efficacy has been discounted. In the early preclinical stage for developing new therapies, it is necessary to adopt appropriate in vitro or in vivo preclinical models to estimate the efficacy and possible toxicity of anticancer drugs to cancer patients [2]. Current tumor models for drug evaluation generally are to implant xenografts derived from well-established human cancer cell lines into immunodeficient mice. However, cell line models not only lack heterogeneity and tissue structure, but also fail to accurately mimic the complex tumor environment involving tumor deterioration under hypoxic conditions, excessive hypoxia-induced transcription factor activation, defective immune evasion mechanisms and angiogenesis [3]. Therefore, cell line models might not be the most appropriate models for evaluating the efficacy of novel drugs.

As a possible solution, xenografts derived from engrafting fresh surgical specimens directly into immunodeficient mice have facilitated the development of in vivo models of human tumors [4]. Such patient-derived xenograft models (PDXs) established by direct metastasis of tumor tissue retain morphological, structural, and molecular characteristics similar to those of primary cancer [5]. The interaction from host microenvironment, genetic characteristics, gene expression patterns, and histological characteristics of the original patient were reproduced in immunodeficient mice that have received transplants [6, 7]. The above characteristics will directly support PDXs as a reliable strategy to anticipate clinical findings and rapidly screen potential therapies, which provides guidance for optimizing personalized treatment in advanced cancer, and suggest new treatment opportunities for patients without other treatment options [8].

In recent years, many attempts have been made to provide additional models, such as genetically engineered mouse models (GEMMs), alteration of the host in xenograft models (zebrafish), innovations in culture patterns (conditionally reprogrammed cell cultures and induced pluripotent stem cells, etc.), and several in vitro models (organoids, spheroids, and microfluidics) [9,10,11,12]. These platforms have their own characteristics in terms of tumor architecture, microenvironment, cellular composition and heterogeneity, stem differentiation status, growth patterns, and response to treatment. However, partially significant limitations have led to low scoring of these traditional models in the evaluation criteria for animal models (Table 1). They still have a long way to go to become reliable models leading to a better understanding of fundamental cancer biology and future applications to translational cancer research. PDX models established in immunodeficient mice (PDXs-IM), as one of the most reliable and standard models in preclinical studies approximately a century after the first tumor model, are more expected to effectively connect nonclinical and clinical data in translational research, ultimately becoming the standard "Avatar" model for human cancer research.

Table 1 Characteristics of different cancer models

With the development of highly immunocompromised mouse receptors, PDXs-IM have been successfully constructed by domestic and foreign research institutions including breast cancer, liver cancer, pancreatic cancer, esophageal cancer, gastric cancer, colorectal cancer, cervical carcinoma, bladder cancer, non-small cell lung cancer, pleural mesothelioma, squamous cell carcinoma of head and neck, glioblastoma and small cell lung cancer, etc. [13]. In this review, we will describe in detail the impact of the evolution of immunodeficient mice on PDX-IM, elucidate the points to consider when establishing such models, and investigate the application of PDX-IM models to cancer research.

Patient-derived xenograft models in immunodeficiency mice

Tumor biopsies or tissues from patients were implanted into immunodeficient mice to generate PDX-IM models that better reflect the tumor-stromal interactions present in the primary tumors (Fig. 1A), although the matrix is derived from the host [14]. In contrast to cell line-derived xenograft models, the PDX-IM model has great potential for efficacy assessment and co-clinical studies. Previous studies have identified that tumors formed in PDX-IM models are histologically and genetically similar to patients' original tumors, in addition to their high genomic fidelity [15]. For example, Dong et al. established 33 RCC mouse xenograft models by biopsy transplantation and tumor resection transplantation, accompanied by differences in transplantation rates between the two approaches. The PDX-IM tumors were highly consistent with the primary tumors in regard to histology, mutation status, copy number changes and targeted therapy response [16]. Furthermore, fifteen PDX-IM models were successfully established from 62 gastric cancer patients and passaged to maintain tumors in immune-compromised mice. The histological and genetic characteristics of PDX-IM models were relatively stable in passage through the comparison of genomes and other characteristics of later generations [17]. Despite some drawbacks, including low transplant rates and high costs, the PDX-IM model has been widely used in personalized medicine, drug screening, combination clinical therapy, and especially in drug efficacy prediction. In terms of metastatic models, cell line-derived xenograft models, as well as transgenic mouse models, often fail to reproduce key mechanisms. Patient-derived xenograft models have become an attractive alternative as they more effectively reflect the diversity and heterogeneity of tumors, while considerable progress has been made in metastasis research [18]. Therefore, the PDX-IM model is currently the most powerful tool for assessing tumor-related mechanisms. The success rate of grafting primary tumor specimens into PDX-IM models is affected by the type of immunodeficient mice, as detailed later in the review.

Fig. 1
figure 1

The process of creating PDX models in immunodeficient mice (A) and human immunodeficient mice (B)

Patient-derived xenograft models in Humanized immunodeficiency mice

With the rapid development of cancer immunotherapy in recent years, in addition to the models for screening chemotherapeutic drugs, researchers also need a PDX-IM model that can be used for immunotherapy evaluation, known as humanized PDX-IM model [19]. This model is currently the best pre-clinical model for evaluating immunotherapy, providing numerous insights into the behavior of different cancers in their own tumor microenvironment under the action of human immune cells. This necessitates the creation of mice with a human immune system on the basis of immunodeficient mice as a vehicle to better test the efficacy of immunotherapy (Table 2) [20]. In general, human derived mice that contributed to the establishment of the humanized PDX-IM model by combining with the PDX-IM model (Fig. 1B), can be obtained by injecting peripheral blood mononuclear cells (PBMC) or CD34+ human hematopoietic stem cells (HSCs) (obtained from bone marrow, cord blood, fetal liver or thymus tissue) [21]. However, PBMCs may cause severe graft-versus-host responses, meanwhile obtaining CD34+ cells from clinical cancer patients is challenging [22]. At present, several strains of immunodeficient mice have been utilized to produce humanized mice: the NOD.Cg-PrkdcscidIl2rgtm1WjlTg (CMV IL-3, CSF2, KITLG)1Eav/MloySzJ (also known as NSG-SGM3) mice, the NOD, B6. SCID Il2rγ − / − KitW41/W41(NBSGW) mice and the human SIRPA and IL15 knockin(SRG-15) mice [23,24,25]. Unfortunately, various deficiencies remain in the humanized mice, in particular different degrees of immune system reconstruction, which may lead to different efficacy of immunotherapy. Several types of human hematopoietic cells could not sufficiently differentiate with hematopoietic stem cells in any humanized mouse strain, such as erythrocytes, platelets, neutrophils, NKT cells, and ILC2 [20]. In addition, conventional humanized mouse models suffer from incomplete replacement of the hemato-lymphoid system and inefficient myelopoiesis in humans [26].

Table 2 Different types, construction methods, characteristics, and applications of humanized mice

This status has been partially improved by the advent of cytokine humanized mice (MISTRG mice), which combine genetic preconditioning and cytokine-mediated support by knocking in gene replacement, removing mouse cytokine-encoding genes and replacing them with their human counterparts [24, 27]. In these mice, there was a clear increase in the level of human hematopoietic engraftment in organs. For example, human phenotypically defined heat shock protein cells in the bone marrow, T cells in the thymus, and myeloid cells in non-hematopoietic organs have elevated levels of engraftment that approach those in the human system [28]. In this way, the innate and adaptive immune responses to diseases such as COVID-19 and myelodysplastic syndromes in humans have been faithfully recapitulated [29, 30]. Radtke et al. developed the first "monkeyized" mouse xenografts through the MISTRG mouse model, which allowed for pre-evaluation of novel HSC-mediated gene therapies, thereby enabling more facile and fewer costly evaluation of promising strategies [31]. In addition, the establishment of a mouse model of humanized immune system has been further improved: such as irradiation or chemical pretreatment, depletion of auto-immune cells in mice, injection of human cytokines, construction of viral vectors, and injection of gene expression plasmids.

Development of immunodeficient mice

Immune-deficient mice implanted with the human immune system provide powerful models for the study of human immunology in vivo, and the PDX-IM model using these mice is a critical tool for discussing the interaction of human immunity with various cancers. Therefore, the use of the most appropriate host mouse strain to generate PDX-IM models is an essential consideration, with different applications and research benefits in different tumors. At present, several types of immunodeficient mice can be used to establish xenograft models: nude mice, severe combined immunodeficient (SCID) mice, SCID/Beige mice, non-obese diabetic (NOD) mice, NOD/SCID mice, NSG mice, BALB/c mice, etc. (Table 3). Studies have reported that NOD/SCID mice are mainly used for lung cancer and melanoma, NSG mice for breast, SCCHN and ovarian cancer, Balb/c nude mice for colon, pancreatic and gastric cancer and renal cell cancer, and SCID mice for prostate cancer [32].

Table 3 Strains of immunocompromised mice used to develop the PDX-IM model

Nude mice have been used as recipients of human tumor xenograft, accompanied by high implantation rate of gastrointestinal tumors, easy observation of subcutaneous tumors and low price, thus they are still an important resource for PDX-IMs establishment with an efficiency of 75% [3, 33, 34]. However, because the complete (or activated) innate immunity and leakage of T cells in nude mice restrict the options for human cancer transplantation, SCID mice have been developed to improve the efficiency of tumor transplantation [35, 36]. Affected by the function of remnant natural killer (NK) cells that prevent homing and maintenance of human cells, transplantation efficiencies of human blood cells and tumor cells in this mouse models are not as high as expected [37]. Fortunately, SCID/Beige mice were established by crossing SCID mice with Beige mice to overcome the effects of NK cells, increase uptake of human tumor cells, and are more commonly used in F1 [38].

NOD/SCID mice with IL2rg mutations, such as NOD.Cg-PrkdcscidIl2rgtm1Wjl (NSG) or NODShi.Cg-PrkdcscidIl2rgtm1Sug (NOG) mice, have highly enhanced immunodeficiency and are able to engraft almost all types of human cancers [39]. Moreover, NOD.Cg‐PrkdcscidIl2rgtm1Sug/Jic, NOD.Cg‐PrkdcscidIl2rgtm1Wjl/SzJ and NOD.Cg‐PrkdcscidJak3tm1card are also the major tools for establishing PDX-IMs [14, 33], which is characterized by a higher degree of immunodeficiency due to the decline or complete absence of natural killer (NK) cell function [33]. For hematological malignancies, such as leukemia and multiple myeloma, it is necessary to implant them directly into the blood or bone marrow of NOG/NSG mice. The BRJ mice have been used as alternative recipients of cholangiocarcinoma PDX-IMs with a high engraftment ratio (75%) [39, 40]. Additional solid tumors, such as head and neck tumors, gastric cancers, and bladder cancers, were also transplanted into the BRJ mice with relatively high engraftment rates compared to currently available models. Since BRJ mice are easy to breed and maintain, and have the benefits of both BRJ and nude mice, they may be ideal models for passaging and drug evaluation [41]. This, in part, explains the importance of selecting mouse strains for cancer research as the number of immunodeficient strains increases.

Interfering factors to consider of establishing PDX-IM model

The PDX-IM model is promising and has led to some exciting breakthroughs in oncology research. However, not all patient tissues can be successfully established as PDX-IM models, and the obstacles that hinder the establishment of PDX-IM model normalization still deserve to be considered. In addition to mouse strain, several factors (Fig. 2), such as (a) tumor type, subtypes, and hormone supplement, (b) tumor microenvironment, (c) Matrigle, (d) xenograft material, (e) implantation site, and (f) gender gap need to be investigated. In order to better utilize PDX-IM models, effective establishment of them is particularly important.

Fig. 2
figure 2

The interfering factors to consider of establishing PDX-IM model, including mouse strain, tumor type, subtypes, hormone supplement, tumor microenvironment, Matrigle, xenograft material (tumor debris, single-cell suspensions, surgical excision, biopsy or metastatic samples, tumor stage and size, treatment status), implantation site, and gender gap

Tumor type, subtypes, and hormone supplement

The success rate of establishing PDX-IMs varies depending on the source of the tumor and the characteristics of the disease, such as tumor invasiveness, relapse/recurrence status, primary or metastatic tumors, and higher transplantation rates tend to manifest in more aggressive, recurrent, and highly metastatic tumors (Table 4). The reported success rates of PDX-IM models generally range from 23 to 75% [42]. Gastrointestinal cancer, such as colon cancer (64–89%) and pancreatic cancer (62%), seems to have higher transplant rates than other cancers. However, the response rate varies greatly among different subtypes of gastric cancer [43]. The success rate of breast cancer implantation is very low (13–27%), especially for estrogen receptor-positive breast cancer, even in the most successful laboratory of PDX-IMs engraftment [44]. Among human breast cancers, triple-negative breast cancer has the highest acceptance rate (51.3%), followed by HER2-positive (26.5%) and luminal B (5.0%). Furthermore, the stability rate of ER-negative and PR-negative (52% and 37%) was significantly higher than that of ER-positive and PR-positive tumors (2% and 3%) [45]. In general, orthotropic transplantation is required for this cancer model [14]. For hormone-dependent tumors, the success of PDX-IM model transplantation is also affected by the experimental protocol, and human hormone replacement therapy can be used to improve the transplantation rate [46]. Most ER-positive PDX-IM models of breast cancer remain estrogen-dependent in tumor growth, and their response to endocrine therapy is consistent with the clinical response of tumor origin. Supplementation of estradiol pellets increased the stability rate of xenografts from 2.6% to 21.4% [42]. In prostate cancer, the development of the prostate cancer xenograft model has been hampered by low success rates, in part by the poor vascularization of the transplantation site, which is successful only when applied to advanced cancers with high growth rates [47]. Supplementation of exogenous androgen shortened the latency of tumorigenesis and improved the rate of tumorigenesis [48]. Lin et al. supplemented testosterone in non-obese diabetic (NOD)/SCID male mice, resulting in a 44% transplantation success rate [47].

Table 4 Current status of transplantation based on multiple interfering factors in different tumors

Tumor microenvironment

The tumor microenvironment is the internal environment in which tumor cells are produced and live. In addition to cancer cells, the microenvironment includes surrounding lymphatics and capillaries, stromal cells (immune cells and cancer-associated fibroblasts), additional normal cells, extracellular matrix (ECM) and various signaling molecules. For the growth and maintenance of cancer cells, changes in microenvironmental conditions play an irreplaceable role [49], involving the promotion of unrestricted cell proliferation, tumorigenesis and direct metastasis of tumors [50]. Among these components of the tumor microenvironment, stromal cells are able to directly regulate the behavior of tumor cells. Tumor growth is accompanied by tumor-specific T-cell maturation and tumor-specific T-cell activation under normal conditions, and accumulation of NK cells is also observed at the tumor site. Tauriello et al. reported that several driver mutations in a mouse model of colorectal cancer were specifically modified in intestinal stem cells to develop metastatic tumors. The quadruple-mutant mice exhibited hallmarks of human colorectal cancer, including T-cell exclusion and TGFb-activated stromal cells. Inhibition of TGF β induced a cytotoxic T-cell response to tumor cells, thus preventing metastasis [51].

Similarly, cancer-associated fibroblasts (CAFs), as one of the most important components of the tumor microenvironment, secrete a variety of cytokines to facilitate tumor growth. Ohlund and his colleagues found that pancreatic stellate cells (PSCs) first differentiate into CAFs and then form stroma. Interestingly, these pancreatic stellate cells have two subtypes: the former could increase the expression of α-smooth muscle actin (α SMA) in tumor cells adjacent to mouse and human PDA tissues, while the latter is located far away from tumor cells and secreting IL-6 and other inflammatory mediators, but lacking α SMA expression. In accordance with this, Seino and colleagues established a co-culture of PDAC organoids and CAFs, indicated that the CAFs provide a WNT niche for PDAC [52], which provides direct evidence for the heterogeneity of CAFs in PDA tumor biology and highlights the importance of CAFs in the tumor microenvironment [53].

Abnormal tumor vascular function, including irregular and premature vascular networks, inadequate microcirculation, and high vascular permeability, may also contribute to the formation of an adverse pathophysiological tumor microenvironment. Hypoxia, in particular, is a common condition in most tumor masses, generally resulting in mutations, inhibition of apoptosis, and epithelial-mesenchymal transition [54]. Furthermore, exosomes produced by cancer cells have been proved to be an active communication mechanism between tumors and their microenvironment, making some breakthroughs in the treatment of drug resistance, metastasis and immunosuppression [55]. Given the high complexity of the composition of the tumor microenvironment, numerous microenvironmental factors should also be considered in the modeling.

Matrigel

During the establishment of PDX-IM model, a mouse basement membrane extract (matrix gel) was used to improve the xenograft rate by combining with patient-derived biopsy materials. The procedure is simple and requires only attention to hold the needle in place for a few seconds after injection to allow the mixture to condense and prevent leakage. Fridman et al. directly mixed tumor cells with BME/ Matrigel at low temperature and injected to increase uptake and growth of cancer cells, cancer stem cells and non-cancerous cells [56, 57]. Generally, tumors grow nicely initially and then "stagnate.", and injection of BME/Matrigel near the tumor center will reinitiate growth. Countless lines of tumor cells that do not grow individually in mice can grow, and tumors that are already well-grown grow faster. The higher the concentration of BME/Matrigel, the faster the growth of tumor cells, because the presence of growth factors in Matrigel is conducive to the xenograft of primary tumor cells, and additional growth factors are added to further promote growth [58]. This not only accelerates the growth of tumor cells, saving time and animal costs, but also considerably increases the number of animal models of human cancer. The uptake rate is classically greater than 80%, which is much higher for most cancers [59].

The addition of Matrigel made cell lines more accessible from xenografts, improved the proximity between the tumor spheroid environment and the tumor growth environment in vivo [60]. Many research groups have mixed BME/Matrigel with tumor cells and injected them in vivo to investigate the therapy of tumors, as well as determine whether genetically modified tumor cells could form tumors [61]. BME/Matrigel could also be supplemented with type I collagen in the orthotopic mammary fat pad model to promote the growth and reduce differentiation of breast cancer MCF7 cells [62]. In order to better analyze the effect of Matrigel, Michael with colleagues adopted a bilateral planting method in the PDX-IM model of colorectal cancer: one side was pre-soaked in Matrigel, which significantly ameliorated the tumor extraction rate compared to the remaining side without Matrigel [63]. More definitely, the combination of BME/Matrigel with patient biopsy materials has been commercialized by several companies (Champions Oncology, Oncostat, and Crown Bio) for "precise" or "personalized" drugs.

Implantation site

The PDX-IM model is divided into ectopic and orthotopic implantation. Ectopic implantation is the implantation of material into an area unrelated to the original tumor site, mostly subcutaneously. Subcutaneous transplantation of PDX-IM models has been widely developed due to its simple operation, high success rate and more accurate monitoring of tumor size [42]. The most common implant site is the dorsal side of the mouse, which is particularly suitable for situations that require large transplants over a short period of time. Alternatively, implantation in the same organ (orthotopic transplantation) as the primary tumor could be selected, such as pancreas, oral cavity, ovary, breast fat pad, brain [44]. Orthotopic transplantation might be an ideal approach, with the advantage that the tumor could develop in the same anatomical microenvironment and thus exhibit more similar behavior to the patient's tumor, especially in terms of metastasis. For several tumor types (such as ovarian, lung and testicular cancers), orthotopic transplantation has significantly increased the incidence of tumors [64]. However, this approach requires trained surgical techniques to generate appropriate PDX-IM models, which are complex and expensive, and imaging techniques are commonly used to monitor tumor growth. Modeling typically takes 2 to 4 months, and failure to transplant is indicated if no tumor growth is observed for 6 months.

Several approaches unrelated to the source of the tumor have implanted primary tumors into the renal capsule to improve the success rate of transplantation. The blood vessels in the subrenal capsule (SRC) site are more abundant compared to the subcutaneous transplantation site, and the fertilization rate of most intact transplanted tissues is high, including benign prostate tissue [47]. The implantation rate of non-small cell lung cancer (NSCLC) in the renal capsule is as high as 90%, while that after subcutaneous implantation is only 25%, and these results are not derived from a single comparative study [65]. Recently, Wu et al. have standardized a transplant protocol and established PDX-IM models of hormone-naive (D17225) and castration-resistant (B45354) PC by implanting fresh tumor samples that obtained from patients with advanced PC under the renal capsule of immune-compromised mice, thus demonstrating the significant effectiveness of the infrarenal zone in the modeling of localized prostate tumors [48]. Furthermore, renal capsule implantation shortens time for engraftment, which is one of the most significant variables in the research of seeking real-time PDX-IMs data for personalized cancer treatment [66].

Xenograft material

Tumor debris or single-cell suspensions

Two different graft materials were utilized for the generation of PDX-IM models, including tumor debris or single-cell suspensions digested by tumors. The application of tumor discrete fragments and single cell suspensions in PDX-IM models has their own characteristics. Tumor debris retains the interconnection between tumor cells and some structural characteristics of the original tumor, thus mimicking the microenvironment of the tumor. Alternatively, the single-cell suspensions could avoid biased sampling of the entire tumor and achieve indiscriminate selection of subclones during analysis or tumor passage. However, single-cell suspensions expose tumor cells to chemical or mechanical forces that may sensitize cells to anoikis, thereby affecting cell survival and transplantation success [67]. Dong et al. established 33 RCC mice xenograft models by tumor debris and single cell suspension with a total implantation rate of 45%, but the success rate of single cell suspension transplantation was 17% different from that of surgical resection of tumor debris transplantation [17]. In addition, the transplanted tissue, typically 1–2mm3 in size, needs to be kept fresh during the transplantation process, which means that the time from the operating room to the laboratory should be as short as possible. After the operation, the tissue was immediately preserved in a cold, fresh medium. In a recent study, the ex vivo times of successful cases of gastric cancer PDX-IM differed greatly from unsuccessful cases (median time for successful cases was 75 min vs. 135 min for unsuccessful cases, P = 0.003). Similarly, shorter overall procedure time was associated with engraftment success (123 min for successful engraftment vs. 167 min for unsuccessful engraftment, P = 0.01) [17]. In addition, the process of sample collection, preservation, and transportation is critical to ensure maximum freshness of samples. This process also takes into account the number of transplanted tissues and the appropriate percentage of tumor cells in the tissues, with the greater the number of fragments, the higher the success rate [67]. There are differences in genome and gene expression levels among different isolates of the same cell line. A cell line represents only one tumor type and actually only one patient in many cases, thus the success rate of transplantation will be relatively changed with different isolates [68]. In one study, Madhavi et al. compared the growth and metastasis of estrogen receptor negative (ER) breast cancer cell lines (MDA-MB-231, SUM1315, CN34BrM) and an ER+ cell line (T47D) in immune mice. The results demonstrated that the weight and size of tumors of each cell line were significantly different at different time points, and the mice also showed different signs of pain [69].

Surgical excision, biopsy or metastatic samples

In malignant tumors, radical surgical resection is superior to partial resection or biopsy in preserving tumor integrity. This is exemplified in a comparison of transurethral resection of the prostate (TURP) with radical prostatectomy for the treatment of prostate cancer. TURP is more prone to generate tissue debris, leading to the destruction of tissue structure, thereby reducing tumor heterogeneity and reducing tumor cell invasion. In a prostate cancer study, successful PDX-IM models were all developed from tissues derived from radical prostatectomy [70]. Furthermore, Lawrence et al. investigated the factors that determine the initial engraftment of patient tissue extracted from TURP specimens and confirmed that only 21% of the grafts contained cancer at the time of harvest. Grafts were most successful when the original patient specimen contained significant amounts of viable cancers, defined as a specimen with (I) at least 50% cancer cells, (II) no physical damage, and (III) detectable Ki67 expression [71]. In colorectal cancer, katsiampoura et al. found a higher success rate of modeling tumor tissue obtained by surgical resection (36/50 = 72%) than biopsy (14/40 = 35%). In short, specimens resected with surgical integrity would be preferable [72]. However, the existence of a clinical biopsy is indispensable in order to open up xenotransplantation to a broader population of cancer patients in some unresectable primary tumors [73].

PDX-IMs can be successfully created from clinical biopsy specimens that are metastatic or primary, and metastatic cancers exhibit higher engraftment rates. In one study, biopsy specimens from 29 patients were used for engraftment of PDX-IMs, and PDX-IM models created from metastatic biopsies had higher engraftment rates compared with unresectable primary tumor tissue (69 vs. 15.4%, P = 0.001) [74]. Masanori et al. established a PDX-IM model of human brain metastases from breast cancer in the mouse brain with an engraftment rate of 100% (10/10) [75]. In a study of colon cancer, the engraftment rate of the PDX-IM models of metastatic tumors was similarly high at 100% (8/8), as compared with 84% (27/32) in primary cancers [76]. A higher engraftment rate was also observed upon engraftment of primary tumors from distant metastases into the PDX-IM models [77]. These data suggest that the ability of tumors to grow continuously in mice is associated with tumor metastasis, due to the fact that metastasized tumors may be more active and invasive. Of course, there is also related to the degree of differentiation of the tumor. The growth rate of metastases is not limited, and PDX-IM models have indeed demonstrated genomic and transcriptomic signatures of metastatic and recurrent carcinomas in some cases [78].

Tumor stage and size

Different tumor stages play a crucial role in transplantation rates. It was found that non-small cell tumor samples from stage II (43/96, 45%) and stage III (25/49, 51%) patients showed a higher engraftment rate than stage I (32/145, 22%) [79]. Oh et al. demonstrated that the same results were seen in colorectal cancer xenograft mice, with transplantation rates corresponding to different tumor stages of 4 of 15 (26.7%) stage I tumors, 41 of 72 (56.9%) stage II tumors, and 50 of 84 (59.5%) stage III tumors, and 55 of 70 (78.6%) stage IV tumors [80]. Advanced tumors tended to correspond to larger sized tumor volumes, and hepatocellular carcinoma samples taken from patients with large-sized tumors (> 5 cm) showed a higher engraftment rate of PDX-IMs (87/130, 67%) than those with small-sized tumors (≤ 5 cm) (16/124, 12.9%) [81]. Jung et al. successfully produced 20 PDX-IMs of pancreatic cancer and also found that tumor size was an important factor in the success of PDX-IM [82]. However, tumors are routinely inoculated into mice to grow the next generation of PDX-IM when the primary tumor volume transplanted into F1 generation mice approaches 1000 to 2000 mm3 [83, 84]. Excessively large tumors easily affect the survival state of mice, resulting in the lack of nutrition of tumor cells, and the transplantation ability of the tumor is weakened. If the tumor is too small in size, there will be insufficient stromal cells to form the next generation of tumors. The multiple tumor fragments being transplanted are mostly 1–2 mm in diameter [85,86,87]. Excessive tumor volume easily affects the accuracy of tumor transplantation, leading to a shift in the transplantation position, reducing the transplantation rate. On the contrary, tumor fragments that are too small may not adequately reflect the heterogeneity of the primary tumor, thereby affecting the predictive value of PDX-IM in drug screening.

Treatment status

Whether patients treated prior to tumor resection will hinder the successful establishment of PDX-IM models remains controversial. In gastric cancer, 63 PDX-IM models were successfully established from 185 fresh gastroscopic biopsies and maintained in vivo through passage. The results showed that the implantation rate of the biopsy tissues inoculated before chemotherapy (52.1%, 37/71) was higher than that of the biopsy tissues inoculated after chemotherapy (21.9%, 25/114) [88]. Kuwata et al. also found that the success rate of PDX-IM establishment was higher in gastric cancer patients who received chemotherapy than in those who did not (26.4% (9/34) vs. 13.1% (26/198)) [83]. In addition, samples from 133 patients with resected pancreatic duct adenocarcinoma were successfully transplanted into mice, of which 42 samples (32%) received chemotherapy, and the remaining 91 samples (68%) did not [89]. However, in a study of NSCLC PDX-IMs, the engraftment rate without preoperative chemotherapy was 32% (81/247) compared with 37.3% (22/59) in the chemotherapy group, indicating that preoperative chemotherapy did not significantly affect the engraftment rate [15]. The reason may be that part of lung cancer patients treated with chemotherapy have a high degree of tumor differentiation, and even after chemotherapy the malignant tumors are strongly invasive and metastatic. Taken together, chemotherapy may have an impact on the activity of tumor samples, which will reduce the engraftment rate of biopsied tissue. However, this conclusion is not universal across different tumors. The core is to consider the invasiveness, metastasis, as well as the actual situation of the tumor samples.

Gender gap

With the exception of hormone-dependent prostate and breast cancer, the engraftment rates of PDX-IM models in most other tumor types are independent of the sex of the sample or that of the mouse [90, 91]. However, mouse models of gastric cancer tissue or intestinal gastric cancer tissue from male patients are more likely to be successfully established [92]. The androgen receptor (AR) was demonstrated to directly regulate miR-125b expression and the AR-miR-125b signaling pathway inhibits apoptosis and promotes proliferation, thus may improve transplantation efficiency.

Overall, as these points are handled properly, PDX-IM will be constructed more effectively, which in turn provide more potentially predictive value. There are some critical factors in the establishment of PDX-IM models of partial cancers (Table 5), which still require further study.

Table 5 Key points for establishing PDX-IM models in different tumors

PDX-IM models in cancer research

Screening of drugs and diagnosis of biomarkers

PDX-IMs are substantial for clinical decision-making before human clinical trials, the development of anti-cancer agents, and diagnosis of biomarkers (Fig. 3). One of the major problems in oncology drug development is the low success rate of new drugs, with only 5% of preclinical anticancer drugs eventually approved for clinical treatment. Many anticancer drugs failed due to lack of efficacy in phase II and III clinical trials and wasted a lot of resources, mainly because of the low predictive value of conventional preclinical models for screening new formulations for clinical development [39]. As a preclinical model with high predictive value, the PDX-IM model plays an irreplaceable role in preclinical screening of new anticancer drugs. In the absence of appropriate biomarkers to detect patient selection and response monitoring for new drugs, PDX-IM models could alter this status quo, both for targeted drugs and for classical cytotoxic drugs. Studies have demonstrated that the drug response rate of PDX-IM model in breast cancer, renal cell carcinoma, non-small cell lung cancer, head and neck squamous cell carcinoma, colorectal cancer and other cancers is very similar to the clinically observed effective rate [67].

Fig. 3
figure 3

PDX-IM applications in cancer therapy, such as screening and evaluation of drugs, bioinformatics databases, diagnosis of biomarkers and sensitive indicators, co-clinical trials (mechanisms of drug resistance), metastasis of malignant tumors, precision medicine, and immunotherapy

In addition to identifying potential clinical indications, PDX-IM models have contributed greatly to the diagnosis of potential biomarkers. In fact, PDX-IMs provide a virtually unlimited source of tumor tissue for multi-dimensional molecular profiles and allow detecting responses to multiple drugs in the same model. According to the consistency between PDX-IM model and human trials, the corresponding drug biomarkers were screened out. In one study, PDX-IM models of colorectal cancers treated with an epidermal growth factor receptor inhibitor, cetuximab, showed comparable response rates to those of the patients in whom the tumor originated [127]. Because of the consistency of drug response in PDX-IM model with clinical practice, the analysis aims to identify tumor-specific and matrix-specific biomarkers, and ultimately to achieve the discovery of drug efficacy research and specific biomarkers [128].

Co-clinical trials and precision medicine

Once a drug enters clinical trials, the opportunity to analyze and integrate useful information for the development of the formulation on a real-time basis is limited [129]. In the absence of comprehensive and more in-depth clinical observations, patients may experience extreme responses or rapid drug resistance, and thus the concept of co-clinical trials has been proposed to address these issues. Co-clinical trials refer to concurrent preclinical and clinical trials (parallel studies between mouse models and patients) that comprehensively analyze and integrate relevant clinical, biological and pharmacological information to identify predictive biomarkers for specific therapeutic responses, even in rare types of cancer. These studies initially applied the results of genetically engineered mouse models to clinical trials and were validated in the treatment of acute promyelocytic leukemia. This trial, considered a model of personalized care or precision medicine [130], has shown positive results in parallel clinical trials, including clear cell adenocarcinoma [131], melanoma [132], oral squamous cell carcinoma [133] and small cell lung cancer [134]. For example, the PDX-IM model of 85 patients with metastatic colorectal cancer was established to predict resistance to targeted anti-EGFR therapies using a combination of anti-HER-2 and anti-EGFR therapies. The results suggest that patients with metastatic colorectal cancer resistant to cetuximab and those with concomitant HER-2 amplification and ineffective clinical treatment may benefit from the combination of HER-2 inhibitors and EGFR inhibitors [135].

Most human solid tumors did not metastasize after subcutaneous implantation in nude mice, whereas the PDOX (Patient-derived orthotopic xenograft) model could recapitulate the local aggressive growth as well as metastasis behavior of primary tumors, and is commonly used to establish a metastatic tumor model. Metastasis models typically involve orthotopic transplantation of primary tumor fragments waiting for transplanted tumor growth to produce spontaneous metastasis [136], or direct orthotopic transplantation of metastases [137]. Drug sensitivity may differ between primary and metastatic tumors grown in the PDOX model, which were previously undetectable in the subcutaneous graft tumor model. For example, neither the subcutaneous PDX-IM model nor the PDOX model of HER-2-positive cervical cancer nude mice were sensitive to the benzamide histone delactase inhibitor eninostat, but in the PDOX model, the drug significantly reduced the load of metastatic tumors compared with the control group [138]. PDOX model could better reflect the biological process of tumor metastasis by retaining the microenvironment of tumor, including the role of stroma in the process of tumor treatment, the mechanism of tumor metastasis, the drug response of metastatic tumors and other related clinical studies [139], which is of great significance to the precise treatment of tumor metastasis.

Oncology research has evolved in parallel with the improved understanding of the cancer genotype and phenotype. Multiple potential targets have been identified in some patients, making it difficult to select the most appropriate target, ushering in a new era of precision medicine. Different from traditional chemotherapy, precision medicine combines the characteristics of individual patients, that is, investigating the genomic profiles of tumors through molecular targeted drugs or immunotherapy to maximize therapeutic efficacy and minimize side effects [140]. Practically, the concept of precision medicine is to divide patients into different Gene subpopulations based on sophisticated genomic profiling, enabling certain therapies to target specific subgroups [42]. In view of this, PDX-IM models, which play an increasingly important role in personalized medicine, not only represent subpopulations with similar genetic profiles, but also recapitulate the intratumoral heterogeneity of tumors in primary patients (Table 6). Meanwhile, its genomics, metabolomics and microbiome analysis are the closest to entering clinical practice [141].

Table 6 Application of PDX-IM model in preclinical or clinical studies of several common tumors

Immunotherapy and bioinformatics databases

In recent years, immunotherapy has achieved widespread success against a variety of malignancies. The humanized PDX-IM model facilitates the study of tumor biology and immune system function by reconstructing the human immune system and tumor growth. Zhao et al. developed a PDX-IM model matching the human immune system as an immuno-oncology model using NOD-SCID Il2rg (NSG) mice and investigated immunotherapy approaches utilizing type I humanized leukocyte antigen in mice. Among them, the treatment and side effects of phenylpropanolizumab and ipilimumab have been investigated in this model [35]. In cell therapy, PDX-IM models can be used to evaluate various aspects of CAR-T cell therapy and biology. In particular, for the interaction between CAR-T and other immune cells (such as Tregs and bone marrow derived suppressor cells (MDSC)) in the tumor microenvironment, PDX-IM models will show more accurate and acceptable results [166]. The humanized PDX-IM model is a future tool for personalized medicine that will support clinical decision-making. In an avatar of human melanoma patients (hIL2-NOG mice), anti-PD-1 (programmed cell death protein 1) antibody responses and tumor-infiltrating T cells support clinical decision making for immunotherapy [167]. However, the humanized immune PDX-IM model still needs more validation.

Many institutions and organizations were committed to creating a large number of PDX-IM or PDX-IM bioinformatics databases [139]. These bioinformatics databases with patient clinical data, pathology, gene profiles and drug response data are essential for predicting and validating drug response information from tumors with similar genetic backgrounds. The successful establishment of a global PDX-IM bioinformatics database has contributed to the rapid acquisition of similar PDX-IM models by comparing data related to specific patients, thereby transforming traditional clinical treatment concepts and facilitating the transition from individualized to programmed therapies. By comparing clinical samples with those in the database, the optimal therapy regimen can be determined from a shared database when patient genomic characteristics are similar or consistent [44, 168]. Currently, PDX-IM bioinformatics databases are available in the United States and Europe (Table 7), and most PDX-IMS are derived from common cancers [139, 169]. Similarly, PDX-IM bioinformatics databases in Asia and PDX-IM for rare cancers are indispensable, which facilitates the sharing between global PDX-IM bioinformatics databases, as well as the popularization of PDX-IM across all tumors.

Table 7 PDX-IM bioinformatics databases

Challenges and prospects in PDX-IM models

The ability to directly transfer human tumors into mice and perform multiple in-vivo passages provides unique opportunities for cancer research and drug discovery, making PDX-IM a valuable cancer model. However, like other model systems, understanding the limitations is necessary for optimal application [170]. Firstly, the longer time of model establishment limits the application in patients with a shorter expected survival [67]. Second, the establishment and maintenance are costly, and the amount of tissue available for implantation is limited. In order to improve transplantation rates, the next phase of PDX-IMs development aims to identify the most appropriate conditions and methodologies to maximize tumor formation [3], sometimes requiring transplantation of smaller samples for personalized medication, such as fine needle aspiration. Third, the incidence of developing EBV-related B-cell lymphoma was as high as 68% when PDX-IM models were generated using severe combined immunodeficient mice NOD/SCID, NSG, or NOD, especially in the F1 generation (33.3%) [17]. However, lymphomagenesis can be reduced using nude mice, which do not form lymphoma even when NOD (F2) mice are used in subsequent transplants. Fourth, the pharmacodynamic evaluation system of the PDX-IM model needs to be improved. Tumor growth retardation during therapy typically results in a larger tumor volume at the end point than before treatment, but smaller than in the control group, indicating that the treatment is biologically active. In fact, this response does not imply clinical efficacy, but is only clinically defined as "disease progression" or even "disease stabilization" [139], requiring quantitative indicators to categorize responses in order to more accurately assess treatment effects in trials. Some researchers have proposed improved evaluation criteria [168], which combines reaction speed, intensity and persistence to further unify the interpretation of treatment response. Fifth, human stromal components are rapidly lost during implantation and replaced by the microenvironment of mice [171], thus genetic heterogeneity cannot be fully manifested in dissected tumors of passage, which results in tumors with genetic heterogeneity that cannot always be reproduced in successive passages. PDX-IM models have been reported to undergo mouse-pecific tumor evolution with rapid accumulation of copy number alterations during PDX-IM passaging, which differed from those acquired during tumor evolution in patients by the strong selection pressures in the mice [170]. Consequently, PDX-IMs need to find a solution in the context of simulating a fully accurate human tumor microenvironment. In this respect, advanced real-time imaging systems are able to quantitatively assess the growth and metastatic progression of primary tumors. Bioluminescence imaging of PDX-IMs from organs is a highly sensitive approach for detecting micrometastasis lesions, but relies on the use of imaging modalities [172, 173]. Finally, further research is necessary to develop strategies for evaluating the efficacy of immunosuppressive checkpoint inhibitors, as the PDX-IM model was established only in immunodeficient mouse strains. Long-term preservation of PDX-IM models is difficult and requires the establishment of efficient long-term cryopreservation conditions to prevent microbial infection. Previous studies have shown that special cryoprotectants exhibit superior performance over traditional media [174].

Conclusion

Different tumor molecular signatures correspond to different therapeutic responses, which are not well represented in most preclinical models. Since the advent of the first tumor models, PDX-IM models have demonstrated significant tumor heterogeneity and are among the most reliable and standard models in preclinical studies. In addition, as a promising and innovative preclinical tool, PDX-IMs are available for the study of tumor initiation, progression, and metastasis (generally orthotopic transplantation). However, despite the increasing relevance of PDX-IM models in cancer research and treatment, patient-derived models also suffer from limitations due to the lack of human immune cells and stromal cells, which contribute to tumor progression by interacting with tumor cells. We set up a humanized PDX-IM mouse model to recapitulate immune cell interactions in the human tumor microenvironment. It is important to note that, while emphasizing the individualization of tumor PDX-IM models, targeted therapies based on the genomic characteristics of PDX-IM models are required to improve the efficiency of the model's application and even to obtain more tumor subtypes and more effective targeting options. Meanwhile, the confounding factors that affect the efficiency of model establishment need to be considered as well as addressing these key issues: (1) improving the engraftment success rate of the models, (2) accelerating the generation rate, (3) ameliorating long-term preservation conditions, (4) reducing microbial infection, (5) perfecting the pharmacodynamic evaluation system, (6) promoting the application of matrix and immune system related research, etc., is conducive to making PDX-IMs the mainstream model for studying tumor biology, investigating genetic heterogeneity and therapeutic targets. Although there are certain limitations of such models, they hold promise for developing more applications in cancer research.

Availability of data and materials

All data generated or analyzed during this study are included in this manuscript and its Additional file.

References

  1. Sun W, Shi Q, Zhang H, Yang K, Ke Y, Wang Y, Qiao L. Advances in the techniques and methodologies of cancer gene therapy. Discov Med. 2019;27(146):45–55.

    CAS  PubMed  Google Scholar 

  2. Zou J, Liu Y, Wang J, Liu Z, Lu Z, Chen Z, Li Z, Dong B, Huang W, Li Y, Gao J, Shen L. Establishment and genomic characterizations of patient-derived esophageal squamous cell carcinoma xenograft models using biopsies for treatment optimization. J Transl Med. 2018;16(1):15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Xu C, Li X, Liu P, Li M, Luo F. Patient-derived xenograft mouse models: A high fidelity tool for individualized medicine. Oncol Lett. 2019;17(1):3–10.

    CAS  PubMed  Google Scholar 

  4. Goto T. Patient-derived tumor xenograft models: toward the establishment of precision cancer medicine. J Pers Med. 2020;10(3):64.

    Article  PubMed  PubMed Central  Google Scholar 

  5. Krepler C, Sproesser K, Brafford P, Beqiri M, Garman B, Xiao M, et al. A comprehensive patient-derived xenograft collection representing the heterogeneity of melanoma. Cell Rep. 2017;21:1953–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Palechor-Ceron N, Krawczyk E, Dakic A, Simic V, Yuan H, Blancato J, et al. Conditional reprogramming for patient-derived cancer models and next-generation living biobanks. Cells. 2019;8:1327.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Chijiwa T, Kawai K, Noguchi A, Sato H, Hayashi A, Cho H, et al. Establishment of patient-derived cancer xenografts in immunodeficient NOG mice. Int J Oncol. 2015;47:61–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Eoh KJ, Chung YS, Lee SH, Park SA, Kim HJ, Yang W, et al. Comparison of clinical features and outcomes in epithelial ovarian cancer according to tumorigenicity in patient-derived xenograft models. Cancer Res Treat. 2018;50:956–63.

    Article  CAS  PubMed  Google Scholar 

  9. Zhu Y, Huang W, Wu Y, Jia L, Li Y, Chen R, et al. Establishment of a patient-derived xenotransplantation animal model for small cell lung cancer and drug resistance model. Zhongguo Fei Ai Za Zhi. 2019;22:6–14.

    PubMed  Google Scholar 

  10. Mahmoudian RA, Farshchian M, Abbaszadegan MR. Genetically engineered mouse models of esophageal cancer. Exp Cell Res. 2021;406(2): 112757.

    Article  CAS  PubMed  Google Scholar 

  11. Brown HK, Schiavone K, Tazzyman S, Heymann D, Chico TJ. Zebrafish xenograft models of cancer and metastasis for drug discovery. Expert Opin Drug Discov. 2017;12(4):379–89.

    Article  CAS  PubMed  Google Scholar 

  12. Wu GD, Xiao YG, Fang FY, Yao D, Liu J, Cao YH, Mao Y, Yu B, Yao TR, Wu YM, Qian YH, Yu DH. Monitoring of lung malignant epithelial cells by gene methylation analysis in the conditionally reprogrammed cell cultures. Neoplasma. 2020;67(3):692–9.

    Article  CAS  PubMed  Google Scholar 

  13. Drost J, Clevers H. Organoids in cancer research. Nat Rev Cancer. 2018;18(7):407–18.

    Article  CAS  PubMed  Google Scholar 

  14. Lallo A, Schenk MW, Frese KK, Blackhall F, Dive C. Circulating tumor cells and CDX models as a tool for preclinical drug development. Transl Lung Cancer Res. 2017;6:397–408.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Cuppens T, Depreeuw J, Annibali D, Thomas D, Hermans E, Gommé E, et al. Establishment and characterization of uterine sarcoma and carcinosarcoma patient-derived xenograft models. Gynecol Oncol. 2017;146:538–45.

    Article  PubMed  Google Scholar 

  16. Dong Y, Manley BJ, Becerra MF, Redzematovic A, Casuscelli J, Tennenbaum DM, et al. Tumor xenografts of human clear cell renal cell carcinoma but not corresponding cell lines recapitulate clinical response to sunitinib: feasibility of using biopsy samples. Other. 2017;3:590–8.

    Google Scholar 

  17. Choi YY, Lee JE, Kim H, Sim MH, Kim KK, Lee G, et al. Establishment and characterisation of patient-derived xenografts as paraclinical models for gastric cancer. Sci Rep. 2016;6:22172.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Liu X, Taftaf R, Kawaguchi M, Chang YF, Chen W, Entenberg D, et al. Homophilic CD44 interactions mediate tumor cell aggregation and polyclonal metastasis in patient-derived breast cancer models. Cancer Discov. 2019;9:96–113.

    Article  PubMed  Google Scholar 

  19. Tavakoli F, Sartakhti JS, Manshaei MH, Basanta D. Cancer immunoediting: a game theoretical approach. In Silico Biol. 2021;14:1–12.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Ito R, Takahashi T, Ito M. Humanized mouse models: application to human diseases. J Cell Physiol. 2018;233:3723–8.

    Article  CAS  PubMed  Google Scholar 

  21. Morton JJ, Bird G, Keysar SB, Astling DP, Lyons TR, Anderson RT, et al. XactMice: humanizing mouse bone marrow enables microenvironment reconstitution in a patient-derived xenograft model of head and neck cancer. Oncogene. 2016;35:290–300.

    Article  CAS  PubMed  Google Scholar 

  22. Abraham S, Pahwa R, Ye C, Choi JG, Pahwa S, Jaggaiahgari S, et al. Long-term engraftment of human natural T regulatory cells in NOD/SCID IL2rγc(null) mice by expression of human IL-2. PLoS ONE. 2012;7: e51832.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Magnotti E, Marasco WA. The latest animal models of ovarian cancer for novel drug discovery. Expert Opin Drug Discov. 2018;13:249–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Martinov T, McKenna KM, Tan WH, Collins EJ, Kehret AR, Linton JD, et al. Building the next generation of humanized hemato-lymphoid system mice. Front Immunol. 2021;12: 643852.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Herndler-Brandstetter D, Shan L, Yao Y, Stecher C, Plajer V, Lietzenmayer M, et al. Humanized mouse model supports development, function, and tissue residency of human natural killer cells. Proc Natl Acad Sci U S A. 2017;114:E9626–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Rongvaux A, Willinger T, Martinek J, Strowig T, Gearty SV, Teichmann LL, et al. Corrigendum: development and function of human innate immune cells in a humanized mouse model. Nat Biotechnol. 2017;35(12):1211.

    Article  CAS  PubMed  Google Scholar 

  27. Sippel TR, Radtke S, Olsen TM, Kiem HP, Rongvaux A. Human hematopoietic stem cell maintenance and myeloid cell development in next-generation humanized mouse models. Blood Adv. 2019;3(3):268–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Saito Y, Ellegast JM, Rafiei A, Song Y, Kull D, Heikenwalder M, et al. Peripheral blood CD34+ cells efficiently engraft human cytokine knock-in mice. Blood. 2016;128(14):1829–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Sefik E, Israelow B, Zhao J, Qu R, Song E, Mirza H, et al. A humanized mouse model of chronic covid-19 to evaluate disease mechanisms and treatment options. Res Sq. 2021;rs.3.rs-279341.

  30. Song Y, Rongvaux A, Taylor A, Jiang T, Tebaldi T, Balasubramanian K, et al. A highly efficient and faithful MDS patient-derived xenotransplantation model for pre-clinical studies. Nat Commun. 2019;10(1):366.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Radtke S, Chan YY, Sippel TR, Kiem HP, Rongvaux A. MISTRG mice support engraftment and assessment of nonhuman primate hematopoietic stem and progenitor cells. Exp Hematol. 2019;70:31–41.

    Article  CAS  PubMed  Google Scholar 

  32. Chen C, Lin W, Huang Y, Chen X, Wang H, Teng L. The essential factors of establishing patient-derived tumor model. J Cancer. 2021;12(1):28–37.

    Article  PubMed  PubMed Central  Google Scholar 

  33. Okada S, Vaeteewoottacharn K, Kariya R. Establishment of a patient-derived tumor xenograft model and application for precision cancer medicine. Chem Pharm Bull (Tokyo). 2018;66:225–30.

    Article  CAS  PubMed  Google Scholar 

  34. Collins AT, Lang SH. A systematic review of the validity of patient derived xenograft (PDX) models: the implications for translational research and personalised medicine. Other. 2018;6: e5981.

    Google Scholar 

  35. Zhao Y, Shuen TWH, Toh TB, Chan XY, Liu M, Tan SY, et al. Development of a new patient-derived xenograft humanised mouse model to study human-specific tumour microenvironment and immunotherapy. Gut. 2018;67:1845–54.

    Article  CAS  PubMed  Google Scholar 

  36. Okada S, Vaeteewoottacharn K, Kariya R. Application of highly immunocompromised mice for the establishment of patient-derived xenograft (PDX) models. Cells. 2019;8:889.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Giovanella BC, Fogh J. The nude mouse in cancer research. Adv Cancer Res. 1985;44:69–120.

    Article  CAS  PubMed  Google Scholar 

  38. Roder J, Duwe A. The beige mutation in the mouse selectively impairs natural killer cell function. Nature. 1979;278:451–3.

    Article  CAS  PubMed  Google Scholar 

  39. Lai Y, Wei X, Lin S, Qin L, Cheng L, Li P. Current status and perspectives of patient-derived xenograft models in cancer research. J Hematol Oncol. 2017;10:106.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Vaeteewoottacharn K, Pairojkul C, Kariya R, Muisuk K, Imtawil K, Chamgramol Y, et al. Establishment of highly transplantable cholangiocarcinoma cell lines from a patient-derived xenograft mouse model. Cells. 2019;8:496.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Sittithumcharee G, Suppramote O, Vaeteewoottacharn K, Sirisuksakun C, Jamnongsong S, Laphanuwat P, et al. Dependency of cholangiocarcinoma on cyclind-dependent kinase activity. Hepatology. 2019;70:1614–30.

    Article  CAS  PubMed  Google Scholar 

  42. Cho SY, Kang W, Han JY, Min S, Kang J, Lee A, et al. An integrative approach to precision cancer medicine using patient-derived xenografts. Mol Cells. 2016;39:77–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Zeng M, Pi C, Li K, Sheng L, Zuo Y, Yuan J, Zou Y, Zhang X, Zhao W, Lee RJ, Wei Y, Zhao L. Patient-derived xenograft: a more standard “avatar” model in preclinical studies of gastric cancer. Front Oncol. 2022;12: 898563.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Hidalgo M, Amant F, Biankin AV, Budinská E, Byrne AT, Caldas C, et al. Patient-derived xenograft models: an emerging platform for translational cancer research. Cancer Discov. 2014;4:998–1013.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Goetz MP, Kalari KR, Suman VJ, Moyer AM, Yu J, Visscher DW. Tumor sequencing and patient-derived xenografts in the neoadjuvant treatment of breast cancer. J Natl Cancer Inst. 2017. https://doi.org/10.1093/jnci/djw306.

    Article  PubMed  PubMed Central  Google Scholar 

  46. Centenera MM, Hickey TE, Jindal S, Ryan NK, Ravindranathan P, Mohammed H, et al. A patient-derived explant (pde) model of hormone- dependent cancer. Mol Oncol. 2018;12:1608–22.

    Article  PubMed  PubMed Central  Google Scholar 

  47. Lin D, Wyatt AW, Xue H, Wang Y, Dong X, Haegert A, et al. High fidelity patient-derived xenografts for accelerating prostate cancer discovery and drug development. Cancer Res. 2014;74:1272–83.

    Article  CAS  PubMed  Google Scholar 

  48. Wu P, Xu R, Chen X, Zhao Y, Tan D, Zhao Y, et al. Establishment and characterization of patient-derived xenografts for hormone-nave and castrate-resistant prostate cancers to improve treatment modality evaluation. Aging (Albany NY). 2020;12:3848–61.

    Article  CAS  PubMed  Google Scholar 

  49. Choi J, Gyamfi J, Jang H, Koo JS. The role of tumor-associated macrophage in breast cancer biology. Histol Histopathol. 2018;33:133–45.

    CAS  PubMed  Google Scholar 

  50. Ren B, Cui M, Yang G, Wang H, Feng M, You L, et al. Tumor microenvironment participates in metastasis of pancreatic cancer. Mol Cancer. 2018;17:108.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Tauriello DVF, Palomo-Ponce S, Stork D, Berenguer-Llergo A, Badia-Ramentol J, Iglesias M, et al. TGFβ drives immune evasion in genetically reconstituted colon cancer metastasis. Nature. 2018;554:538–43.

    Article  CAS  PubMed  Google Scholar 

  52. Seino T, Kawasaki S, Shimokawa M, Tamagawa H, Toshimitsu K, Fujii M, et al. Human pancreatic tumor organoids reveal loss of stem cell niche factor dependence during disease progression. Cell Stem Cell. 2018;22:454–67.

    Article  CAS  PubMed  Google Scholar 

  53. Öhlund D, Handly-Santana A, Biffi G, Elyada E, Almeida AS, Ponz-Sarvise M, et al. Distinct populations of inflammatory fibroblasts and myofibroblasts in pancreatic cancer. J Exp Med. 2017;214:579–96.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Jing X, Yang F, Shao C, Wei K, Xie M, Shen H, et al. Role of hypoxia in cancer therapy by regulating the tumor microenvironment. Mol Cancer. 2019;18:157.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Meldolesi J. Exosomes and ectosomes in intercellular communication. Curr Biol. 2018;28:435–44.

    Article  Google Scholar 

  56. Fridman R, Benton G, Aranoutova I, Kleinman HK, Bonfil RD. Increased initiation and growth of tumor cell lines, cancer stem cells and biopsy material in mice using basement membrane matrix protein (Cultrex or Matrigel) co-injection. Nat Protoc. 2012;7:1138–44.

    Article  CAS  PubMed  Google Scholar 

  57. Zhan T, Ambrosi G, Wandmacher AM, Rauscher B, Betge J, Rindtorff N, et al. MEK inhibitors activate Wnt signalling and induce stem cell plasticity in colorectal cancer. Nat Commun. 2019;10:2197.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Kastana P, Zahra FT, Ntenekou D, Katraki-Pavlou S, Beis D, Lionakis MS, et al. Matrigel plug assay for in vivo evaluation of angiogenesis. Methods Mol Biol. 2019;1952:219–32.

    Article  CAS  PubMed  Google Scholar 

  59. Badea MA, Balas M, Hermenean A, Ciceu A, Herman H, Ionita D, et al. Influence of matrigel on single- and multiple-spheroid cultures in breast cancer research. SLAS Discov. 2019;24:563–78.

    Article  CAS  PubMed  Google Scholar 

  60. Lewis MT, Landua JD, Adams-III HC, Medina D. A mystery wrapped in an enigma: matrigel enhancement of mammary cell growth and morphogenesis. J Mammary Gland Biol Neoplasia. 2012;17:99–101.

    Article  PubMed  Google Scholar 

  61. Grigorieva EV, Prudnikova TY, Domanitskaya NV, Mostovich LA, Pavlova TV, Kashuba VI, et al. D-Glucuronyl C5-epimerase suppresses small-cell lung cancer cell proliferation in vitro and tumour growth in vivo. Br J Cancer. 2011;105:74–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Samineni S, Glackin C, Shively JE. Role of CEACAM1, ECM, and mesenchymal stem cells in an orthotopic model of human breast cancer. Int J Breast Cancer. 2011;2011: 381080.

    PubMed  Google Scholar 

  63. Gock M, Kühn F, Mullins CS, Krohn M, Prall F, Klar E, et al. Tumor take rate optimization for colorectal carcinoma patient-derived xenograft models. Biomed Res Int. 2016;2016:1–7.

    Article  Google Scholar 

  64. Okano M, Oshi M, Butash A, Okano I, Saito K, Kawaguchi T, et al. Orthotopic implantation achieves better engraftment and faster growth than subcutaneous implantation in breast cancer patient-derived xenografts. J Mammary Gland Biol Neoplasia. 2020;25:27–36.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Morgan KM, Riedlinger GM, Rosenfeld J, Ganesan S, Pine SR. Patient-derived xenograft models of non-small cell lung cancer and their potential utility in personalized medicine. Front Oncol. 2017;7:2.

    Article  PubMed  PubMed Central  Google Scholar 

  66. Davies AH, Wang Y, Zoubeidi A. Patient-derived xenografts: a platform for accelerating translational research in prostate cancer. Mol Cell Endocrinol. 2018;462:17–24.

    Article  CAS  PubMed  Google Scholar 

  67. Jung J, Seol HS, Chang S. The generation and application of patient-derived xenograft model for cancer research. Cancer Res Treat. 2018;50:1–10.

    Article  PubMed  Google Scholar 

  68. Zhang X, Claerhout S, Prat A, Dobrolecki LE, Petrovic I, Lai Q, et al. A renewable tissue resource of phenotypically stable, biologically and ethnically diverse, patient-derived human breast cancer xenograft models. Cancer Res. 2013;73:4885–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Puchalapalli M, Zeng X, Mu L, Anderson A, Hix Glickman L, Zhang M, et al. NSG mice provide a better spontaneous model of breast cancer metastasis than athymic (nude) mice. PLoS ONE. 2016;11: e0163521.

    Article  PubMed  PubMed Central  Google Scholar 

  70. Wetterauer C, Vlajnic T, Schuler J, Gsponer JR, Thalmann GN, Cecchini M, et al. Early development of human lymphomas in a prostate cancer xenograft program using triple knock-out immunocompromised mice. Prostate. 2015;75:585–92.

    Article  PubMed  Google Scholar 

  71. Lawrence MG, Pook DW, Wang H, Porter LH, Frydenberg M, Kourambas J, et al. Establishment of primary patient-derived xenografts of palliative TURP specimens to study castrate-resistant prostate cancer. Prostate. 2015;75:1475–83.

    Article  CAS  PubMed  Google Scholar 

  72. Katsiampoura A, Raghav K, Jiang Z-Q, Menter DG, Varkaris A, Morelli MP, et al. Modeling of patient-derived xenografts in colorectal cancer. Mol Cancer Ther. 2017;16:1435–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Roife D, Kang Y, Wang L, Fang B, Swisher SG, Gershenwald JE, et al. Generation of patient-derived xenografts from fine needle aspirates or core needle biopsy. Surgery. 2017;161:1246–54.

    Article  PubMed  Google Scholar 

  74. Hernandez MC, Bergquist JR, Leiting JL, Ivanics T, Yang L, Smoot RL, et al. Patient-derived xenografts can be reliably generated from patient clinical biopsy specimens. J Gastrointest Surg. 2019;23(4):818–24.

    Article  PubMed  Google Scholar 

  75. Oshi M, Okano M, Maiti A, Rashid OM, Saito K, Kono K. Novel breast cancer brain metastasis patient-derived orthotopic xenograft model for preclinical studies. Cancers. 2020;12:444.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Puig I, Chicote I, Tenbaum SP, Arques O, Herance JR, Gispert JD, et al. A personalized preclinical model to evaluate the metastatic potential of patient-derived colon cancer initiating cells. Clin Cancer Res. 2013;19:6787–801.

    Article  CAS  PubMed  Google Scholar 

  77. Sivanand S, Pena-Llopis S, Zhao H, Kucejova B, Spence P, Pavia-Jimenez A. et al. A Validated tumorgraft model reveals activity of dovitinib against renal cell carcinoma. Sci Transl Med2012;4:137ra75–ra75.

  78. Rosfjord E, Lucas J, Li G, Gerber HP. Advances in patient-derived tumor xenografts: from target identification to predicting clinical response rates in oncology. Biochem Pharmacol. 2014;91:135–43.

    Article  CAS  PubMed  Google Scholar 

  79. Chen Y, Zhang R, Wang L, Correa AM, Pataer A, Xu Y, et al. Tumor characteristics associated with engraftment of patient-derived non-small cell lung cancer xenografts in immunocompromised mice. Cancer. 2019;125:3738–48.

    Article  CAS  PubMed  Google Scholar 

  80. Oh BY, Lee WY, Jung S, Hong HK, Nam DH, Park YA, et al. Correlatibetween tumor engraftment in patieon nt-derived xenograft models and clinical outcomes in colorectal cancer patients. Oncotarget. 2015;6:16059–68.

    Article  PubMed  PubMed Central  Google Scholar 

  81. Hu B, Li H, Guo W, Sun YF, Zhang X, Tang WG, et al. Establishment of a hepatocellular carcinoma patient-derived xenograft platform and its application in biomarker identification. Int J Cancer. 2019;146:1606–17.

    Article  PubMed  Google Scholar 

  82. Jung J, Lee HC, Seol HS, Choi YS, Kim E, Lee EJ, et al. Generation and molecular characterization of pancreatic cancer patient-derived xenografts reveals their heterologous nature. Oncotarget. 2016;7:62533–46.

    Article  PubMed  PubMed Central  Google Scholar 

  83. Kuwata T, Yanagihara K, Iino Y, Komatsu T, Ochiai A, Sekine S, et al. Establishment of novel gastric cancer patient-derived xenografts and cell lines: pathological comparison between primary tumor, patient-derived, and cell-line derived xenografts. Cells. 2019;8(6):585.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Lu JH, Wang Y, Meng Q, Zeng ZL. Establishment of gastric cancer patient-derived xenograft models and primary cell lines. J Vis Exp. 2019. https://doi.org/10.3791/59871.

    Article  PubMed  Google Scholar 

  85. Illert B, Otto C, Thiede A, Timmermann W. Detection of disseminated tumor cells in nude mice with human gastric cancer. Clin Exp Metastasis. 2003;20(6):549–54.

    Article  PubMed  Google Scholar 

  86. Illert B, Otto C, Braendlein S, Thiede A, Timmermann W. Optimization of a metastasizing human gastric cancer model in nude mice. Microsurgery. 2003;23(5):508–12.

    Article  PubMed  Google Scholar 

  87. Jones-Bolin S, Ruggeri B. Orthotopic models of human gastric carcinoma in nude mice: applications for study of tumor growth and progression. Curr Protoc Pharmacol. 2007. https://doi.org/10.1002/0471141755.ph1404s37.

    Article  PubMed  Google Scholar 

  88. Zhu Y, Tian T, Li Z, Tang Z, Wang L, Wu J, et al. Establishment and characterization of patient-derived tumor xenograft using gastroscopic biopsies in gastric cancer. Sci Rep. 2015;5:8542.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Pergolini I, Morales-Oyarvide V, Mino-Kenudson M, Honselmann KC, Rosenbaum MW, Nahar S, et al. Tumor engraftment in patient-derived xenografts of pancreatic ductal adenocarcinoma is associated with adverse clinicopathological features and poor survival. PLoS ONE. 2017;12:e0182855.

    Article  PubMed  PubMed Central  Google Scholar 

  90. Xu H, Zheng H, Zhang Q, et al. A multicentre clinical study of sarcoma personalised treatment using patient-derived tumour xenografts. Clin Oncol (R Coll Radiol). 2022. https://doi.org/10.1016/j.clon.2022.06.002.

    Article  PubMed  Google Scholar 

  91. Lin Qi, Kogiso Mari Du, Yuchen, et al. Impact of SCID mouse gender on tumorigenicity, xenograft growth and drug-response in a large panel of orthotopic PDX models of pediatric brain tumors. Cancer Lett. 2020;493:197–206.

    Article  Google Scholar 

  92. Liu B, Zhou M, Li X, Zhang X, Wang Q, Liu L, et al. Interrogation of gender disparity uncovers androgen receptor as the transcriptional activator for oncogenic Mir-125b in gastric cancer. Cell Death Dis. 2021;12(5):441.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Cavalloni G, Peraldo-Neia C, Sassi F, Chiorino G, Sarotto I, Aglietta M, et al. Establishment of a patient-derived intrahepatic cholangiocarcinoma xenograft model with KRAS mutation. BMC Cancer. 2016;16:90.

    Article  PubMed  PubMed Central  Google Scholar 

  94. Cottu P, Marangoni E, Assayag F, de Cremoux P, VincentSalomon A, Guyader C, et al. Modeling of response to endocrine therapy in a panel of human luminal breast cancer xenografts. Breast Cancer Res Treat. 2012;133:595–606.

    Article  CAS  PubMed  Google Scholar 

  95. Sprouffske K, Kerr G, Li C, Prahallad A, Rebmann R, Waehle V, et al. Genetic heterogeneity and clonal evolution during metastasis in breast cancer patient-derived tumor xenograft models. Comput Struct Biotechnol J. 2020;18:323–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Mattie M, Christensen A, Chang MS, Yeh W, Said S, Shostak Y, et al. Molecular characterization of patient-derived human pancreatic tumor xenograft models for preclinical and translational development of cancer therapeutics. Neoplasia. 2013;15:1138–50.

    Article  PubMed  PubMed Central  Google Scholar 

  97. Guo S, Gao S, Liu R, Shen J, Shi X, Bai S, et al. Oncological and genetic factors impacting PDX model construction with NSG mice in pancreatic cancer. FASEB J. 2019;33:873–84.

    Article  CAS  PubMed  Google Scholar 

  98. Chen Q, Wei T, Wang J, Zhang Q, Li J, Zhang J, et al. Patient-derived xenograft model engraftment predicts poor prognosis after surgery in patients with pancreatic cancer. Pancreatology. 2020;20:485–92.

    Article  CAS  PubMed  Google Scholar 

  99. Wang H, Lu J, Tang J, Chen S, He K, Jiang X, et al. Establishment of patient-derived gastric cancer xenografts: A useful tool for preclinical evaluation of targeted therapies involving alterations in her-2, met and fgfr2 signaling pathways. BMC Cancer. 2017;17:191.

    Article  PubMed  PubMed Central  Google Scholar 

  100. Zhang T, Zhang L, Fan S, Zhang M, Fu H, Liu Y, et al. Patient-derived gastric carcinoma xenograft mouse models faithfully represent human tumor molecular diversity. PLoS ONE. 2015;10: e0134493.

    Article  PubMed  PubMed Central  Google Scholar 

  101. Han C, Shen J, Wang H, Yu L, Qian X, Liu B, et al. Personalized primary tumor xenograft model established for the pre-clinical trial to guide postoperative chemotherapy. Med Hypotheses. 2012;79:705–8.

    Article  CAS  PubMed  Google Scholar 

  102. Schütte M, Risch T, Abdavi-Azar N, Boehnke K, Schumacher D, Keil M, et al. Molecular dissection of colorectal cancer in pre-clinical models identifies biomarkers predicting sensitivity to EGFR inhibitors. Nat Commun. 2017;8:14262.

    Article  PubMed  PubMed Central  Google Scholar 

  103. Prasetyanti PR, van Hooff SR, van Herwaarden T, de Vries N, Kalloe K, Rodermond H, et al. Capturing colorectal cancer inter-tumor heterogeneity in patient-derived xenograft (PDX) models. Int J Cancer. 2019;144:366–71.

    Article  CAS  PubMed  Google Scholar 

  104. Guo J, Xu G, Mao C, Wei R. Low Expression of smurf1 enhances the chemosensitivity of human colorectal cancer to gemcitabine and cisplatin in patient-derived xenograft models. Transl Oncol. 2020;13: 100804.

    Article  PubMed  PubMed Central  Google Scholar 

  105. Zhang Y, Lee SH, Wang C, Gao Y, Li J, Xu W. Establishing metastatic patient-derived xenograft model for colorectal cancer. Jpn J Clin Oncol. 2020;50:1108–16.

    Article  PubMed  Google Scholar 

  106. Wang D, Pham NA, Tong J, Sakashita S, Allo G, Kim L, et al. Molecular heterogeneity of non-small cell lung carcinoma patient-derived xenografts closely reflect their primary tumors. Int J Cancer. 2017;140:662–73.

    Article  CAS  PubMed  Google Scholar 

  107. Moro M, Bertolini G, Tortoreto M, Pastorino U, Sozzi G, Roz L. Patient-derived xenografts of non small cell lung cancer: Resurgence of an old model for investigation of modern concepts of tailored therapy and cancer stem cells. J Biomed Biotechnol. 2012;2012: 568567.

    Article  PubMed  PubMed Central  Google Scholar 

  108. Nakajima T, Geddie W, Anayama T, Ko HM, da Cunha SG, Boerner S, et al. Patient-derived tumor xenograft models established from samples obtained by endobronchial ultrasound-guided transbronchial needle aspiration. Lung Cancer. 2015;89:110–4.

    Article  PubMed  Google Scholar 

  109. Almer AC, Plana D, Gao H, Korn JM, Yang G, Green J, et al. A proof of concept for biomarker-guided targeted therapy against ovarian cancer based on patient-derived tumor xenografts. Cancer Res. 2020;80:4278–87.

    Article  Google Scholar 

  110. Heo EJ, Cho YJ, Cho WC, Hong JE, Jeon HK, Oh DY, et al. Patient-derived xenograft models of epithelial ovarian cancer for preclinical studies. Cancer Res Treat. 2017;49(4):915–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Weroha SJ, Becker MA, Enderica-Gonzalez S, Harrington SC, Oberg AL, Maurer MJ, et al. Tumor grafts as in vivo surrogates for women with ovarian cancer. Clin Cancer Res. 2014;20:1288–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Topp MD, Hartley L, Cook M, Heong V, Boehm E, McShane L, et al. Molecular correlates of platinum response in human high-grade serous ovarian cancer patient-derived xenografts. Mol Oncol. 2014;8:656–68.

    Article  PubMed  PubMed Central  Google Scholar 

  113. Boone JD, Dobbin ZC, Straughn JM, Buchsbaum DJ. Ovarian and cervical cancer patient derived xenografts: The past, present, and future. Gynecol Oncol. 2015;138:486–91.

    Article  PubMed  Google Scholar 

  114. Wu J, Zheng Y, Tian Q, Yao M, Yi X. Establishment of patient-derived xenograft model in ovarian cancer and its influence factors analysis. J Obstet Gynaecol Res. 2019;45:2062–73.

    Article  CAS  PubMed  Google Scholar 

  115. Kang HN, Kim JH, Park AY, Choi JW, Lim SM, Kim J, et al. Establishment and characterization of patient-derived xenografts as paraclinical models for head and neck cancer. BMC Cancer. 2020;20:316.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Facompre ND, Sahu V, Montone KT, Harmeyer KM, Nakagawa H, Rustgi AK, et al. Barriers to generating PDX models of HPV-related head and neck cancer. Laryngoscope. 2017;127:2777–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Brabetz S, Leary SES, Gröbner SN, Nakamoto MW, Şeker-Cin H, Girard EJ, et al. A biobank of patient-derived pediatric brain tumor models. Nat Med. 2018;24(11):1752–61.

    Article  CAS  PubMed  Google Scholar 

  118. Priolo C, Agostini M, Vena N, Ligon AH, Fiorentino M, Shin E, et al. Establishment and genomic characterization of mouse xenografts of human primary prostate tumors. Am J Pathol. 2010;176:1901–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Nguyen HM, Vessella RL, Morrissey C, Brown LG, Coleman IM, Higano CS, et al. LuCaP prostate cancer patient-derived xenografts reflect the molecular heterogeneity of advanced disease an–d serve as models for evaluating cancer therapeutics. Prostate. 2017;77:654–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Einarsdottir BO, Bagge RO, Bhadury J, Jespersen H, Mattsson J, Nilsson LM, et al. Melanoma patient-derived xenografts accurately model the disease and develop fast enough to guide treatment decisions. Oncotarget. 2014;5(20):9609–18.

    Article  PubMed  PubMed Central  Google Scholar 

  121. Lang H, Béraud C, Bethry A, Danilin S, Lindner V, Coquard C, et al. Establishment of a large panel of patient-derived preclinical models of human renal cell carcinoma. Oncotarget. 2016;7:59336–59.

    Article  PubMed  PubMed Central  Google Scholar 

  122. Dong Y, Manley BJ, Becerra MF, Redzematovic A, Casuscelli J, Tennenbaum DM, et al. Tumor xenografts of human clear cell renal cell carcinoma but not corresponding cell lines recapitulate clinical response to sunitinib: Feasibility of using biopsy samples. Eur Urol Focus. 2017;3:590–8.

    Article  PubMed  Google Scholar 

  123. Mohseni MJ, Amanpour S, Muhammadnejad S, Sabetkish S, Muhammadnejad A, Heidari R, et al. Establishment of a patient-derived Wilms’ tumor xenograft model: a promising tool for individualized cancer therapy. J Pediatr Urol. 2014;10:123–9.

    Article  PubMed  Google Scholar 

  124. Zhao X, Liu Z, Yu L, Zhang Y, Baxter P, Voicu H, et al. Global gene expression profiling confirms the molecular fidelity of primary tumor-based orthotopic xenograft mouse models of medulloblastoma. Neuro Oncol. 2012;14:574–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Larmour LI, Cousins FL, Teague JA, Deane JA, Jobling TW, Gargett CE. A patient derived xenograft model of cervical cancer and cervical dysplasia. PLoS ONE. 2018;13: e0206539.

    Article  PubMed  PubMed Central  Google Scholar 

  126. Wu L, Allo G, John T, Li M, Tagawa T, Opitz I, et al. Patient-derived xenograft establishment from human malignant pleural mesothelioma. Clin Cancer Res. 2017;23:1060–7.

    Article  PubMed  Google Scholar 

  127. Lee HW, Son E, Lee K, Lee Y, Kim Y, Lee JC, et al. Promising therapeutic efficacy of GC1118, an anti-EGFR antibody, against KRAS mutation-driven colorectal cancer patient-derived xenografts. Int J Mol Sci. 2019;20:5894.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Bradford JR, Wappett M, Beran G, Logie A, Delpuech O, Brown H, et al. Whole transcriptome profiling of patient-derived xenograft models as a tool to identify both tumor and stromal specific biomarkers. Oncotarget. 2016;7:20773–87.

    Article  PubMed  PubMed Central  Google Scholar 

  129. Junttila MR, de Sauvage FJ. Influence of tumour micro-environment heterogeneity on therapeutic response. Nature. 2013;501:346–54.

    Article  CAS  PubMed  Google Scholar 

  130. Clohessy JG, Pandolfi PP. The mouse hospital and its integration in ultra-precision approaches to cancercare. Front Oncol. 2018;8:340.

    Article  PubMed  PubMed Central  Google Scholar 

  131. Vargas R, Gopal P, Kuzmishin GB, DeBernardo R, Koyfman SA, Jha BK, et al. Case study: patient-derived clear cell adenocarcinoma xenograft model longitudinally predicts treatment response. NPJ Precis Oncol. 2018;2:14.

    Article  PubMed  PubMed Central  Google Scholar 

  132. Frankel AE, Eskiocak U, Gill JG, Yuan S, Ramesh V, Froehlich TW, et al. Digoxin plus trametinib therapy achieves disease control in braf wild-type metastatic melanoma patients. Neoplasia. 2017;19:255–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Campbell KM, Lin T, Zolkind P, Barnell EK, Skidmore ZL, Winkler AE, et al. Oral cavity squamous cell carcinoma xenografts retain complex genotypes and intertumor molecular heterogeneity. Cell Rep. 2018;24:2167–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Owonikoko TK, Zhang G, Kim HS, Stinson RM, Bechara R, Zhang C, et al. Patient-derived xenografts faithfully replicated clinical outcome in a phase II co-clinical trial of arsenic trioxide in relapsed small cell lung cancer. J Transl Med. 2016;14:111.

    Article  PubMed  PubMed Central  Google Scholar 

  135. Bertotti A, Migliardi G, Galimi F, Sassi F, Torti D, Isella C, et al. A molecularly annotated platform of patient-derived xenografts (“Xenopatients”) identifies HER2 as an effective therapeutic target in cetuximab-resistant colorectal cancer. Cancer Discov. 2011;1:508–23.

    Article  CAS  PubMed  Google Scholar 

  136. Zhang Z, Hu K, Miyake K, Kiyuna T, Oshiro H, Wangsiricharoen S, et al. A novel patient-derived orthotopic xenograft (PDOX) mouse model of highly-aggressive liver metastasis for identification of candidate effective drug-combinations. Sci Rep. 2020;10:20105.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Roque-Lima B, Roque CCTA, Begnami MD, Peresi P, Lima ENP, Mello CAL, et al. Development of patient-derived orthotopic xenografts from metastatic colorectal cancer in nude mice. J Drug Target. 2019;27:943–9.

    Article  CAS  PubMed  Google Scholar 

  138. Hiroshima Y, Maawy A, Zhang Y, Zhang N, Murakami T, Chishima T, et al. Patient-derived mouse models of cancer need to be orthotopic in order to evaluate targeted anti-metastatic therapy. Oncotarget. 2016;7:71696–702.

    Article  PubMed  PubMed Central  Google Scholar 

  139. Byrne AT, Alférez DG, Amant F, Annibali D, Arribas J, Biankin AV, et al. Interrogating open issues in cancer precision medicine with patient-derived xenografts. Nat Rev Cancer. 2017;17:254–68.

    Article  CAS  PubMed  Google Scholar 

  140. Collins DC, Sundar R, Lim JS, Yap TA. Towards precision medicine in the clinic: from biomarker discovery to novel therapeutics. Trends Pharmacol Sci. 2017;38:25–40.

    Article  CAS  PubMed  Google Scholar 

  141. Burney IA, Lakhtakia R. Precision medicin: where have we reached and where are we headed? Sultan Qaboos Univ Med J. 2017;17:e255–8.

    Article  PubMed  PubMed Central  Google Scholar 

  142. Schueler J, Tschuch C, Klingner K, Bug D, Peille AL, de Koning L, et al. Induction of acquired resistance towards EGFR inhibitor gefitinib in a patient-derived xenograft model of non-small cell lung cancer and subsequent molecular characterization. Cells. 2019;8(7):740.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Zhang T, Qu R, Chan S, Lai M, Tong L, Feng F, et al. Discovery of a novel third-generation EGFR inhibitor and identification of a potential combination strategy to overcome resistance. Mol Cancer. 2020;19(1):90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Odintsov I, Mattar MS, Lui AJW, Offin M, Kurzatkowski C, Delasos L, et al. Novel preclinical patient-derived lung cancer models reveal inhibition of HER3 and MTOR signaling as therapeutic strategies for NRG1 fusion-positive cancers. J Thorac Oncol. 2021;16(7):1149–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Haikala HM, Lopez T, Köhler J, Eser PO, Xu M, Zeng Q, et al. EGFR inhibition enhances the cellular uptake and antitumor-activity of the HER3 antibody-drug conjugate HER3-DXd. Cancer Res. 2022;82(1):130–41.

    Article  CAS  PubMed  Google Scholar 

  146. Zhang XC, Zhang J, Li M, Huang XS, Yang XN, Zhong WZ, et al. Establishment of patient-derived non-small cell lung cancer xenograft models with genetic aberrations within EGFR, KRAS and FGFR1: useful tools for preclinical studies of targeted therapies. J Transl Med. 2013;11:168.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Chew NJ, Lim Kam Sian TCC, Nguyen EV, Shin SY, Yang J, Hui MN, et al. Evaluation of FGFR targeting in breast cancer through interrogation of patient-derived models. Breast Cancer Res. 2021;23(1):82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Coussy F, El Botty R, Lavigne M, Gu C, Fuhrmann L, Briaux A, et al. Combination of PI3K and MEK inhibitors yields durable remission in PDX models of PIK3CA-mutated metaplastic breast cancers. J Hematol Oncol. 2020;13(1):13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Coussy F, Lavigne M, de Koning L, Botty RE, Nemati F, Naguez A, et al. Response to mTOR and PI3K inhibitors in enzalutamide-resistant luminal androgen receptor triple-negative breast cancer patient-derived xenografts. Theranostics. 2020;10(4):1531–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Hsu PY, Wu VS, Kanaya N, Petrossian K, Hsu HK, Nguyen D, et al. Dual mTOR kinase inhibitor MLN0128 sensitizes HR+/HER2+ breast cancer patient-derived xenografts to trastuzumab or fulvestrant. Clin Cancer Res. 2018;24(2):395–406.

    Article  CAS  PubMed  Google Scholar 

  151. Hashimoto Y, Koyama K, Kamai Y, Hirotani K, Ogitani Y, Zembutsu A, et al. A novel HER3-targeting antibody-drug conjugate, u3–1402, exhibits potent therapeutic efficacy through the delivery of cytotoxic payload by efficient internalization. Clin Cancer Res. 2019;25(23):7151–61.

    Article  CAS  PubMed  Google Scholar 

  152. Reddy TP, Choi DS, Anselme AC, Qian W, Chen W, Lantto J, et al. Simultaneous targeting of HER family pro-survival signaling with Pan-HER antibody mixture is highly effective in TNBC: a preclinical trial with PDXs. Breast Cancer Res. 2020;22(1):48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Marangoni E, Vincent-Salomon A, Auger N, Degeorges A, Assayag F, de Cremoux P, et al. A new model of patient tumor-derived breast cancer xenografts for preclinical assays. Clin Cancer Res. 2007;13(13):3989–98.

    Article  CAS  PubMed  Google Scholar 

  154. Kinsey CG, Camolotto SA, Boespflug AM, Guillen KP, Foth M, Truong A, et al. Protective autophagy elicited by RAF→MEK→ERK inhibition suggests a treatment strategy for RAS-driven cancers. Nat Med. 2019;25(4):620–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Garrido-Laguna I, Uson M, Rajeshkumar NV, Tan AC, de Oliveira E, Karikari C, et al. Tumor engraftment in nude mice and enrichment in stroma- related gene pathways predict poor survival and resistance to gemcitabine in patients with pancreatic cancer. Clin Cancer Res. 2011;17(17):5793–800.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Harris FR, Zhang P, Yang L, Hou X, Leventakos K, Weroha SJ, et al. Targeting HER2 in patient-derived xenograft ovarian cancer models sensitizes tumors to chemotherapy. Mol Oncol. 2019;13(2):132–52.

    Article  CAS  PubMed  Google Scholar 

  157. Ricci F, Bizzaro F, Cesca M, Guffanti F, Ganzinelli M, Decio A, et al. Patient-derived ovarian tumor xenografts recapitulate human clinicopathology and genetic alterations. Cancer Res. 2014;74(23):6980–90.

    Article  CAS  PubMed  Google Scholar 

  158. Chen L, Zhang J, Wang X, Li Y, Zhou L, Lu X, et al. Discovery of novel KRAS-PDEδ inhibitors with potent activity in patient-derived human pancreatic tumor xenograft models. Acta Pharm Sin B. 2022;12(1):274–90.

    Article  CAS  PubMed  Google Scholar 

  159. Knudsen ES, Kumarasamy V, Chung S, Ruiz A, Vail P, Tzetzo S, et al. Targeting dual signalling pathways in concert with immune checkpoints for the treatment of pancreatic cancer. Gut. 2021;70(1):127–38.

    Article  CAS  PubMed  Google Scholar 

  160. Gebreyohannes YK, Wozniak A, Zhai ME, Wellens J, Cornillie J, Vanleeuw U, et al. Robust activity of avapritinib, potent and highly selective inhibitor of mutated KIT, in patient-derived xenograft models of gastrointestinal stromal tumors. Clin Cancer Res. 2019;25(2):609–18.

    Article  CAS  PubMed  Google Scholar 

  161. Karalis JD, Yoon LY, Hammer STG, Hong C, Zhu M, Nassour I, et al. Lenvatinib inhibits the growth of gastric cancer patient-derived xenografts generated from a heterogeneous population. J Transl Med. 2022;20(1):116.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Huynh H, Ong R, Zopf D. Antitumor activity of the multikinase inhibitor regorafenib in patient-derived xenograft models of gastric cancer. J Exp Clin Cancer Res. 2015;34:132.

    Article  PubMed  PubMed Central  Google Scholar 

  163. Yao YM, Donoho GP, Iversen PW, Zhang Y, Van Horn RD, Forest A, et al. Mouse PDX trial suggests synergy of concurrent inhibition of RAF and EGFR in colorectal cancer with BRAF or KRAS mutations. Clin Cancer Res. 2017;23(18):5547–60.

    Article  CAS  PubMed  Google Scholar 

  164. Nunes M, Vrignaud P, Vacher S, Richon S, Lièvre A, Cacheux W, et al. Evaluating patient-derived colorectal cancer xenografts as preclinical models by comparison with patient clinical data. Cancer Res. 2015;75(8):1560–6.

    Article  CAS  PubMed  Google Scholar 

  165. Kim MK, Osada T, Barry WT, Yang XY, Freedman JA, Tsamis KA, et al. Characterization of an oxaliplatin sensitivity predictor in a preclinical murine model of colorectal cancer. Mol Cancer Ther. 2012;11(7):1500–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Rodriguez-Garcia A, et al. CAR-T cells hit the tumor microenvironment: strategies to overcome tumor escape. Front Immunol. 2020. https://doi.org/10.3389/fimmu.2020.01109.

    Article  PubMed  PubMed Central  Google Scholar 

  167. Ny L, et al. Supporting clinical decision making in advanced melanoma by preclinical testing in personalized immune-humanized xenograft mouse models. Ann Oncol. 2020;31(2):266–73.

    Article  CAS  PubMed  Google Scholar 

  168. Gao H, Korn JM, Ferretti S, Monahan JE, Wang Y, Singh M, et al. High-throughput screening using patient-derived tumor xenografts to predict clinical trial drug response. Nat Med. 2015;21:1318–25.

    Article  CAS  PubMed  Google Scholar 

  169. Evrard YA, Srivastava A, Randjelovic J, Doroshow JH, Dean DA 2nd, Morris JS, et al. Systematic establishment of robustness and standards in patient-derived xenograft experiments and analysis. Cancer Res. 2020;80(11):2286–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Ben-David U, Ha G, Tseng YY, Greenwald NF, Oh C, Shih J, et al. Patient-derived xenografts undergo mouse-specific tumor evolution. Nat Genet. 2017;49(11):1567–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Liu Y, Chanana P, Davila JI, Hou X, Zanfagnin V, McGehee CD, et al. Gene expression differences between matched pairs of ovarian cancer patient tumors and patient-derived xenografts. Sci Rep. 2019;9:6314.

    Article  PubMed  PubMed Central  Google Scholar 

  172. Pompili L, Porru M, Caruso C, Biroccio A, Leonetti C. Patient-derived xenografts: a relevant preclinical model for drug development. J Exp Clin Cancer Res. 2016;35:189.

    Article  PubMed  PubMed Central  Google Scholar 

  173. Asadzadeh F, Ferrucci V, Antonellis DE, P, Zollo M. In vivo bioluminescence imaging using orthotopic xenografts towards patient’s derived-xenograft Medulloblastoma models. Q J Nucl Med Mol Imaging. 2017;61:95–101.

    Article  PubMed  Google Scholar 

  174. Ivanics T, Bergquist JR, Liu G, Kim MP, Kang Y, Katz MH, et al. Patient-derived xenograft cryopreservation and reanimation outcomes are dependent on cryoprotectant type. Lab Invest. 2018;98:947–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This study was funded by the Key Research and Development Project of the Department of Science and Technology of Sichuan Province (No. 2021YFS0274).

Author information

Authors and Affiliations

Authors

Contributions

XD and PF contributed to the conception of the study; MZ performed the data analyses and wrote the manuscript; ZR, JT, and CH helped perform the analysis with constructive discussions. ML mainly contributed to the revision of this paper. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Ping Fan or Xinhua Dai.

Ethics declarations

Ethics approval and consent to participate

All animal studies (including the mice euthanasia procedure) were done in compliance with the regulations and guidelines of Southwest Medical University institutional animal care and conducted according to the AAALAC and the IACUC guidelines.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zeng, M., Ruan, Z., Tang, J. et al. Generation, evolution, interfering factors, applications, and challenges of patient-derived xenograft models in immunodeficient mice. Cancer Cell Int 23, 120 (2023). https://doi.org/10.1186/s12935-023-02953-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12935-023-02953-3

Keywords