Young woman in head scarf after cancer treatment

Cancer Treatment: The Critical Factors

Cancer Treatment: The Critical Factors

Last Section Update: 09/2020

1 Overview

Summary and Quick Facts for Cancer Treatment: Critical Factors

  • The mainstream medical establishment treats the majority of cancer cases with a one size fits all strategy that may deprive many patients of a greater chance of successful treatment.
  • In this protocol, Life Extension® discusses 9 critical steps that may increase the likelihood of a successful outcome in the treatment of many cancers. Of critical importance to treatment-naïve patients is implementing as many of these 9 critical steps as can safely be done concurrently with conventional therapy.
  • Implementing strategies to address the 9 critical factors of cancer treatment identified in this Life Extension protocol provides an evidenced-based approach to aid physicians in…
    • Determining which medical therapies are most likely to be effective for each individual's unique cancer
    • Targeting multiple biochemical pathways known to be aberrant in many cancers
    • Providing a more thorough prognostic analysis that can help physicians make informed decisions about how aggressive treatment should be
    • Educating patients about some potential side effects associated with conventional cancer treatment options and what they can do to minimize risk

In order to derive the greatest potential benefit from any cancer treatment regimen, both patients and physicians must respect and adapt to the complex and multidimensional nature of each individual's unique cancer. Sadly, the mainstream medical establishment treats the majority of cancer cases with a "one size fits all" strategy that may deprive many patients of a greater chance of successful treatment.

Implementing strategies to address each of the nine critical factors of cancer treatment identified in this Life Extension protocol provides an evidenced based approach that will:

  • Aid physicians in determining which medical therapies are most likely to be effective for each individual's unique cancer;
  • Pharmacogenomically and nutrigenomically target multiple biochemical pathways known to be aberrant in many cancers;
  • Provide a more thorough prognostic analysis that can help physicians make informed decisions about how aggressive treatment should be, and to properly inform their patients about the state of their health, and;
  • Educate patients about some potential side effects associated with conventional cancer treatment options, and what they can do to minimize risk.

In this protocol, Life Extension will discuss the following nine critical steps that may increase the likelihood of a successful outcome in the treatment of many cancers:

  1. Evaluating the molecular biology of the tumor cell population
  2. Analyzing the patient's living tumor cells to determine sensitivity or resistance to chemotherapy
  3. Circulating tumor cell (CTC) testing
  4. Inhibiting the cyclooxygenase enzymes (COX-1 and COX-2)
  5. Suppressing Ras oncogene expression
  6. Correcting coagulation abnormalities
  7. Inhibiting angiogenesis
  8. Inhibiting the 5-lipoxygenase (5-LOX) enzyme
  9. Inhibiting cancer metastasis

Of critical importance to treatment-naïve patients is implementing as many of these nine critical steps as can safely be done concurrently with conventional therapy. In newly diagnosed patients who have not yet been treated, the objective is to eradicate the primary tumor and metastatic cells with a multi-pronged "first strike therapy" so that residual tumor cells are not given an opportunity to evolve survival mechanisms that make them resistant to further treatments.

2 Step One: Evaluating the Molecular Biology of the Tumor Cell Population

Throughout this protocol, you will see terminology relating to the molecular aspects of the cancer cell. When we use the term molecular, we are referring to specific characteristics of cancer cells such as:

  • Tumor-promoting genes (oncogenes)
  • Tumor suppressor genes
  • Receptors or docking sites on the cell membrane where communication with proteins occur
  • Cellular differentiation—the degree of aggressiveness of the cancer cell (poorly differentiated cancer cells are more aggressive, while highly differentiated cells are less aggressive).

These individual variations—the unique biology of the cancer cell—help explain why a particular therapy may be highly effective for some cancer patients but fail others.

People typically think of their disease based on the organ it affects (eg, lung cancer or colon cancer). The problem with that rationale is not all cancers are the same, even if they affect the same organ. With the advent of advanced molecular diagnostic profiling, the specific strengths and vulnerabilities of each patient's cancer can be identified in order to design an individually tailored treatment program.

It is critically important to obtain a description of the type of cells that populate your tumor. Not only does this assist the oncologist in recommending the most effective conventional therapy, but it also helps determine what adjuvant nutritional and/or off-label drug therapies to consider. The human eye can serve to provide the most basic information about a cancer cell through the microscopic examination of the cell's general characteristics. Taking this one step further is evaluation by an immunohistochemistry test. This test detects markers of diagnostic value, on and within the cell surface, through the application of colored dye or stains. In order to perform this and other tests, it is necessary for a sample of your tumor to be sent to a specialized laboratory.

Caris Life Sciences provides a comprehensive range of customized analyses to help cancer specialists correctly identify difficult-to-diagnose tumors, establish prognosis in many cancers (including breast, prostate, and colon), and determine optimal treatment. By providing this information, Caris Life Sciences starts treatment on the right course and helps avoid unnecessary and potentially ineffective therapies. Caris Life Sciences performs more specialized analyses for cancer than any other laboratory in the world.

When a person has cancer, the physician confronts a chain of pressing questions: What type of cancer is it? Where did it originate? Which treatments are most likely to be effective? Caris Life Sciences can help clinicians with answers to many of these questions.

As far as diagnosis is concerned, many cancers defy classification by visual examination. In fact, the diagnosis of "metastatic cancer of unknown primary site" accounts for 2–6% of cancer diagnoses.1 In a majority of these difficult cases, Caris Life Sciences' medical expertise and advanced technologies lead to an accurate diagnosis.

Visual examination of tumors provides very little information about their growth rate or the type of treatment to which they will respond. Caris Life Sciences' prognostic expertise can accurately evaluate the aggressiveness of the cancer and predict the effects of therapy.

Cancers have traditionally been treated as follows: if one therapy proves ineffective, then try another until a successful therapy is found or all options are exhausted. Caris Life Sciences helps eliminate the need for this trial-and-error method by providing individualized information to help determine the optimal therapy before initiating treatment. This can save the patient time, money, and most importantly, it may provide a better opportunity for "first strike" eradication.

Caris Life Sciences provides highly sensitive patient monitoring for the follow-up care of many cancers. For example, Caris Life Sciences can determine whether certain types of lymphomas have recurred before they can be detected by any other method. The earlier tumor recurrence is detected, the greater the likelihood of therapeutic success.

Typically, within 48 hours after receiving a specimen, Caris Life Sciences returns the stained slides along with a thorough and detailed case report to a physician. Your oncologist can also consult with a member of the Caris Life Sciences staff by phone.

Contact information for Caris Life Sciences is as follows:

Telephone: 888-979-8669

Website: http://www.carislifesciences.com/oncology-molecular-intelligence

How to Implement Step One

Make certain your surgeon sends a specimen of your tumor to Caris Life Sciences for immunohistochemistry testing. Be sure to follow the instructions that Caris Life Sciences provides for proper shipping of the surgical specimen. You may have to pay out of pocket for this test because not all insurance plans reimburse for it. Please note that this test may not be of benefit to all cancer patients. While it provides a basis for improved treatment, not all cancers are effectively treated with current therapies.

3 Step Two: Determine Sensitivity or Resistance to Chemotherapy

When a person is prescribed a treatment for their cancer, they might assume that the treatment was chosen based on the uniqueness of their cancer. For instance, when a woman with early-stage breast cancer is told that her chemotherapy treatment regimen will consist of the drugs doxorubicin (Adriamycin), Cytoxan, and Taxol (ACT), she might think this treatment was individually tailored for her cancer. In actuality, ACT is a standard chemotherapy protocol given to breast cancer patients. This "one-size-fits-all" approach to breast cancer treatment would work well if superior results were obtained from this routine practice. Sadly, this has not been the case.

The "one-size-fits-all" approach to prescribing chemotherapy has failed to improve survival for the vast majority of women with breast cancer. In a shocking study of women with breast cancer over the age of 50 who had cancer present in their lymph nodes, standard chemotherapy regimens were shown to increase 10-year survival by only 3%.2,3 Other studies have determined that standard chemotherapy does not improve survival in women with estrogen-receptor positive breast cancer.4,5

A critical flaw of the "one-size-fits-all" approach rests in treating all breast cancers as if they are one in the same. Although traditional oncology does make distinctions in a few obvious qualities, such as size of the cancer, lymph node status, and estrogen receptor status, we now know there can be substantial individual differences in cancer cell genetics among those with "similar" breast cancers. These differences can dramatically influence the response to treatment. A powerful illustration of the lack of appreciation for individual differences in cancer treatment was clearly revealed in a landmark study published in the New England Journal of Medicine in 2007. Researchers compared women with lymph node positive breast cancer who received ACT chemotherapy to those who did not receive chemotherapy.6 Their HER2 status was also determined—which refers to a genetic characteristic of the cancer. The researchers discovered that the group of women who were HER2 negative and estrogen receptor positive did not benefit at all from taking Taxol. The ramifications of this study are immense, as a large percentage of women with breast cancer fall into this category. In recognition of the failure of Taxol to benefit this large group of women with breast cancer, oncologist Anne Moore, M.D., Professor of Clinical Medicine at the Weill Medical College of Cornell University in New York stated, "The days of 'one size fits all' therapy for patients with breast cancer are coming to an end."7

A further indictment of the "one-size-fits-all" approach was prominently displayed in a study published in the Journal of the National Cancer Institute in 2008. In this investigation,8 scientists measured the effectiveness of an anthracycline-based chemotherapy regimen in 5,354 women with early-stage breast cancer. Anthracyclines are a class of chemotherapy drugs of which Adriamycin is a key member. The scientists determined that women with early-stage breast cancer who were HER2 negative derived absolutely no benefit from taking Adriamycin or other anthracycline drugs. Given that approximately 80% of breast cancers are HER2 negative,7 these results suggest that only one of five women with breast cancer may benefit from these drugs that have considerable toxicity associated with their use. For example, in a large-scale study, 5% of patients treated with Adriamycin developed congestive heart failure.9

Breast cancer is not the only type of cancer in which resistance to chemotherapy may impair treatment; in fact, all cancers may display interindividual variability in chemosensitivity. For example, assessing expression of a chemoresistance protein (IGFBP2) in leukemia patients helped identify those who were likely to respond to standard chemotherapy, and those who were not.10

If chemotherapy is being considered, it is crucial to ascertain which chemotherapy drugs will have the highest probability of being effective against your particular cancer. A company called Nagourney Cancer Institute performs chemo-sensitivity testing on your living cancer cells to determine how your cancer cells respond when exposed to various drugs in the laboratory.

Nagourney Cancer Institute was founded in 1993 by Dr. Robert Nagourney, a prominent hematologist and oncologist. Nagourney Cancer Institute pioneers cancer therapies specifically tailored for each individual patient, and is a leader in individualized cancer strategies. With no financial ties to outside healthcare organizations, recommendations are made without financial bias.

Nagourney Cancer Institute develops and provides cancer therapy recommendations that have been designed scientifically for each patient. Following the collection of living cancer cells obtained at the time of biopsy or surgery, Nagourney Cancer Institute performs an ex-vivo analysis of programmed cell death (EVA-PCD) functional profile on your tumor sample to measure drug activity (sensitivity and resistance). Ex-vivo analyses mean your living tumor cells are maintained outside of your body for the purpose of determining which drug or drug combination most effectively induces cell death in the laboratory. Each patient is highly individualized with regard to his or her sensitivity to chemotherapy drugs. Your responsiveness to chemotherapy is as unique as your fingerprints. Therefore, this test will help to determine which drug(s) will be most effective for you. Nagourney Cancer Institute will then make a treatment recommendation based on these findings.

Nagourney Cancer Institute provides custom-tailored, assay-directed therapy based on your tumor response in the laboratory. This eliminates much of the guesswork prior to your undergoing the potentially toxic side effects of chemotherapy regimens that could prove to be of little value against your cancer.

Contact information for Nagourney Cancer Institute:

Nagourney Cancer Institute
750 East 29th Street
Long Beach, CA 90806
Telephone: (562) 989-6455
Email: http://www.nagourneycancerinstitute.com

How to Implement Step Two

Contact Nagourney Cancer Institute so that your surgeon can follow the precise instructions required to send a living specimen of your tumor for chemo- sensitivity testing. It is important that your surgeon carefully coordinate with Nagourney Cancer Institute in order to ensure your cells arrive in a viable condition. You may have to pay for this test yourself because your insurance may not reimburse for it. Please note that this test may not be of benefit to all cancer patients. While it provides a basis for improved treatment, not all cancers are effectively treated with current therapies.

4 Step Three: Circulating Tumor Cell Testing

Circulating tumor cell (CTC) testing involves the detection of cancer cells in the bloodstream. These circulating tumor cells are the "seeds" that break away from the primary site of cancer and spread to other parts of the body. Understanding circulating tumor cells is critically important since it is the spread of cancer to other parts of the body—and not the primary cancer—that is very often responsible for the death of a person with cancer.

Historically, medical science has been focused on the primary tumor, basing treatment decisions on the specific genetic characteristics of the primary cancer cells—which assumes that the metastatic cancer cells are genetically identical to the primary tumor. This assumption might be ill-advised, as research has demonstrated that metastatic cancer cells can be genetically dissimilar from the primary tumor as they become more highly differentiated.

In an illuminating study conducted with metastatic breast cancer patients, researchers compared the genetic composition of the cancer cells that had formed distant metastasis to the genetic composition of the corresponding cancer cells in the primary breast tumor. The findings were alarming—in 31% of the comparisons, the genetic composition of the metastatic cancer cells differed almost completely from that of the primary breast tumors.11 Amazingly, further analysis revealed that none of the pairs of primary breast tumors with its corresponding metastatic cancer were identical. Based on these findings, the authors remarked that "because metastatic cells often have a completely different genetic composition, their phenotype [biological behavior], including aggressiveness and therapy responsiveness, may also vary substantially from that seen in the primary tumors," leading to their conclusion that "the resulting heterogeneity [genetic variability] of metastatic breast cancer may underlie its poor responsiveness to therapy..." To further support the evidence that metastatic cancer cells can vary genetically from the primary tumor, two additional studies with breast cancer patients have demonstrated that CTC can be HER2 positive while the primary breast tumor can be HER2 negative.12,13

This research suggests that directing treatment towards the cancer cells of the primary tumor can, in some cases, be "barking up the wrong tree." Treatments designed to attack the primary tumor could fail to destroy the circulating tumor cells. For this reason, focusing on the metastatic cancer cells could potentially lead to better results. CTC testing provides us with the means with which we can now focus our attention on these potential metastatic cancer cells.

CTC testing has been shown to improve prognostic accuracy. German scientists studied 35 women with non-metastatic breast cancer who had their levels of CTC measured before they had received any treatment for their cancer.13 Seventeen tested positive for CTC, while 18 tested negative for CTC. The group that tested negative for CTC had a median overall survival of 125 months. The group with five or more CTC present in their blood sample had a median overall survival of only 61 months—less than half as long. In a related study, researchers at the University of Texas M. D. Anderson Cancer Center measured CTC in 151 women with metastatic breast cancer.14 These patients were also evaluated for other prognostic cancer markers, such as hormone receptor status, CA 27.29, and HER2 status. Those who had five or more CTC per blood sample had a median overall survival of 13.5 months. The median overall survival for those with less than five CTC per blood sample was over 29 months. The researchers also discovered that the presence of five or more CTC in a blood sample had the highest predictive value compared to all other tumor markers. The researchers went on to state that "circulating tumor cells have superior and independent prognostic value…"

Furthermore, recent research indicates that CTC evaluation can be used to predict prognosis for men with prostate cancer. Researchers at Thomas Jefferson University compared the levels of CTC in 37 men with metastatic prostate cancer.15 Their findings were remarkable—for the men with five or more CTC per blood sample, the median overall survival was only 8.4 months. For those men with fewer than five CTC the median overall survival was 48 months. Yet another study measured CTC in 55 men with a rising PSA after surgery for prostate cancer.16 A rising PSA after surgery is strongly predictive of prostate cancer recurrence.17 Radiation therapy was administered to 15 patients. 60% of those who were CTC positive had progression of their disease during radiation therapy, while there were no disease progressions in the CTC negative group. Additional studies have confirmed these results.18,19

One of the most exciting potential uses of CTC technology is to allow doctors to evaluate treatment effectiveness during the early phase of therapy. With regard to chemotherapy, doctors have often had to wait at least a few months before they can assess the effectiveness of treatment. This inability to evaluate a treatment’s efficacy during the early phase of therapy can have disastrous consequences for the person with cancer. Those three months of waiting to know if the treatment is working can make the difference between altering therapy to reflect the lack of response, or continuing with an ineffective treatment and allowing the cancer to progress. This waiting may become a thing of the past, as recent studies have demonstrated that CTC testing can reliably predict the response to treatment during the early phase of therapy. In an important study,20 163 women with metastatic breast cancer were tested for CTC at four different times during the course of treatment. The first measurement of CTC was taken approximately four weeks after treatment had begun. At the first measurement the researchers discovered that those patients with less than five CTC per sample had a median overall survival of greater than 18.5 months. Those with five or more CTC per sample had a median overall survival of only seven months. Additionally, 66% of those with five or more CTC per sample had died after one year, compared to only 19% of those who had less than five CTC per sample. Thus, as early as four weeks into therapy CTC testing determined which patients were not responding and whose cancer would continue to progress with ineffective treatment. The authors of this study concluded that "detection of elevated CTC at any time during therapy is an accurate indication of subsequent rapid disease progression and mortality for metastatic breast cancer patients."

In a study published in the Journal of Clinical Oncology in 2008, 430 patients with metastatic colon cancer had CTC testing before and 3‒5 weeks after the initiation of treatment with chemotherapy.21 For patients who initially started with three or more CTC detected in their blood sample, if they converted to less than three CTC then the median survival was 11.0 months. However, if they continued to have three or more CTC on follow-up testing then the median survival was only 3.7 months. For patients whose cancers were deemed to be non-progressing by imaging studies, median survival was 18.8 months if they had less than three CTC on follow-up testing, whereas those with three or more CTC on follow-up testing had a median survival of only 7.1 months. The authors concluded that "the number of CTCs before and during treatment is an independent predictor of…overall survival in patients with metastatic colorectal cancer. CTCs provide prognostic information in addition to that of imaging studies."

CTC testing can also predict the likelihood of recurrence after initial cancer treatment. In 2006, scientists in Spain measured the presence of CTC in 84 high-risk breast cancer patients after they received initial chemotherapy.22 The researchers found dramatic differences in the relapse rates between those who tested positive for CTC, as compared to those that did not have any CTC detected in their blood samples. The group testing positive for CTC had a 269% increased risk of relapse, and a 300% increased risk of death, compared to the group testing negative for CTC. Further analysis showed a striking 53-month difference in the time to relapse between the groups. In a related study, German scientists in 2008 studied 25 women with breast cancer that had not metastasized.23 They measured CTC levels before and after the patients received chemotherapy. The results showed that relapse occurred in only 9% of patients whose CTC levels indicated a decline, no change, or minor increase when compared to baseline CTC levels. There was a substantially higher relapse rate of 40% in the group with a CTC increase at the end of therapy.

Given that CTC can be the seeds that eventually form metastatic disease, then CTC analysis provides medical science with an excellent opportunity to examine the genetic features of these cancer cells before metastasis occurs, when treatment is far more likely to be successful. In addition to detecting the presence and quantity of CTC in the bloodstream, recent advances in technology now allows the examination of CTC for a large number of tumor cell markers and genetic expressions. The information obtained from this analysis can provide vital insight as to which chemotherapy drugs are best suited to exploit the genetic weaknesses of the CTC, as well as which chemotherapy agents are likely to be powerless against the genetic strengths of the CTC.

Let’s illustrate the benefits of CTC analysis of tumor markers and genetic expressions with a few examples. Chemotherapy drugs can exert their therapeutic effects by inhibiting essential enzymes within the cancer cell. The over expression of these enzymes—called drug targets—can enhance the tumor destructive effects of these drugs. Adriamycin is a prime example of this mechanism of action. The main drug target for Adriamycin is topoisomerase 2. Studies have demonstrated that those patients with cancers expressing higher levels of topoisomerase 2 can benefit from treatment with Adriamycin.24 Conversely, those patients with cancers that produce smaller amounts of topoisomerase 2 are less likely to respond to Adriamycin. Cancer cells also have the ability to produce enzymes that convert chemotherapy drugs into less potent forms, which weakens the anti-tumor activity of these drugs. 5-FU is commonly used in the treatment of breast and colon cancer. DPD is an enzyme that degrades 5-FU to an inactive metabolite. Cancer cells expressing higher levels of DPD can be resistant to 5-FU. One study of colorectal cancer patients treated with 5-FU revealed that those with high DPD levels had significantly shorter overall survival compared to patients with low DPD expression.25

As an added benefit, genetic analysis of CTC can inform us as to which natural supplements might be best indicated. For instance, NF-kB promotes the growth of cancer. Curcumin is an inhibitor of NF-kB.26 So, a person whose cancer is expressing high levels of NF-kB might consider including curcumin as part of their supplement program.

Some cancers are able to produce GSTpi, which confers resistance to multiple chemotherapy drugs. Ellagic acid—from pomegranate—inhibits GSTpi.27 Supplementation with ellagic acid may be wise if CTC analysis demonstrates over production of GSTpi.

How to Implement Step Three

Note: Quantitative CTC testing is only available for breast, colon, and prostate cancer.

To test for the presence and quantity of CTC in your blood, speak with your physician regarding the CELLSEARCH® CTC test.

Qualitative CTC testing can perform a genetic analysis of CTC to identify the most effective treatments. Qualitative CTC testing can be performed with Guardant360®.

Note: Qualitative CTC testing is most feasible with any kind of carcinoma (eg, lung, colon, breast, ovary, cervix, prostate, stomach, gastric, esophagus, liver...), mesothelioma or melanoma.

It is also possible to test for synovial sarcoma. However, other types of sarcoma (eg, liposarcoma, leiomyosarcoma, chondrosarcoma...) gave lower success rates for isolation of tumor cells (<50%).

Testing of tumors of the central nervous system (eg, glioblastoma) is limited, because in the majority of cases it is not successful to isolate CTCs, since these tumors rarely shed tumor cells into the bloodstream.

Hematological cancers (eg, Hodgkin- or non-Hodgkin lymphomas, B-cell lymphomas, myelogenous and lymphocytic leukemias) can be assayed as well. Please note that we cannot test for T-cell cancers at all. Testing for multiple myeloma (MM) is not recommended since we observed low detection rates in MM.

5 Step Four: Inhibiting the Cyclooxygenase Enzymes (COX-1 and COX-2)

Inflammation plays a pivotal role in the formation and progression of cancer. There are many inflammatory pathways in the body. The cyclooxygenase (COX-2) enzyme is a particular inflammatory pathway that has been the focus of research in the realm of oncology. Initially, scientists believed COX-2 was merely an inducible response to inflammation. It is now speculated that COX-2 performs biological functions in the body, particularly in the brain and kidneys as well as the immune system. COX-2 becomes troublesome when upregulated (sometimes 10- to 80-fold) by pro-inflammatory stimuli (interleukin-1, growth factors, tumor necrosis factor, and endotoxins). When overexpressed, COX-2 participates in various pathways that could promote cancer (ie, angiogenesis), cell proliferation, and the production of inflammatory prostaglandins.28-30

A growing body of research has documented the relationship between COX-2 and cancer:

  • An article in the journal Cancer Research showed that COX-2 levels in pancreatic cancer cells are 60 times greater than in adjacent normal tissue.31
  • Solid tumors contain oxygen-deficient or hypoxic areas (a reduced oxygen supply to a tissue below physiological levels). Hypoxia promotes up-regulation of COX-2 and angiogenesis, and establishes resistance to ionizing radiation.32
  • Within the nonsteroidal anti-inflammatory drug (NSAIDs) class is a subclass referred to as COX-2 inhibitors (cyclooxygenase inhibitors). COX-2 inhibitors were popularly prescribed to relieve pain but now have found a place in oncology. It began when scientists recognized that people who regularly take NSAIDs lowered their risk of colon cancer by as much as 50%.33
  • JAMA reported that a 9.4-year epidemiological study showed that COX-2 upregulation was related to more advanced tumor stage, tumor size, and lymph node metastasis as well as diminished survival rates among colorectal cancer patients.34 With more regular use of aspirin (a COX-2 inhibitor), the risk of dying from the disease decreased.35,36 The journal Gastroenterology reported additional encouragement, showing that three different colon cell lines underwent apoptosis (cell death) when deprived of COX-2; when lovastatin was added to the COX-2 inhibitor the kill rate increased another five-fold.37 The benefits observed with COX-2 inhibitors extend beyond colon protection to the cardiovascular system, where they help sustain endothelial cell function.38
  • A groundbreaking study published in 2009 revealed that breast cancer patients treated with COX-2 inhibitors had a greatly reduced risk of bone metastases. In this investigation, the incidence of bone metastases were recorded in breast cancer patients not treated with a COX-2 inhibitor, as well as in individuals who received a COX-2 inhibitor for at least six months following the diagnosis of breast cancer. The findings were astounding—those who were treated with a COX-2 inhibitor were 90% less likely to develop bone metastases than those who were not treated with a COX-2 inhibitor.39
  • 134 patients with advanced lung cancer were treated with chemotherapy alone or combined with Celebrex (a COX-2 inhibitor). For those patients with cancers expressing increased amounts of COX-2, treatment with Celebrex dramatically prolonged survival.40
  • Celebrex slowed cancer progression in men with recurrent prostate cancer.41,42
  • Celebrex prevented weight loss and improved quality of life in individuals with head and neck cancers.43
  • Regular intake of over-the-counter (OTC) NSAIDs produced highly significant composite risk reductions of 43% for colon cancer, 25% for breast cancer, 28% for lung cancer, and 27% for prostate cancer. Furthermore, in a series of case control studies, daily use of a selective COX-2 inhibitor, either celecoxib or rofecoxib, significantly reduced the risk for each of these malignancies. The evidence is compelling that anti-inflammatory agents with selective or non-selective activity against cycloooxygenase-2 (COX-2) have strong potential for the chemoprevention of cancers of the colon, breast, prostate and lung. Results confirming that COX-2 blockade is effective for cancer prevention have been tempered by observations that some selective COX-2 inhibitors pose a risk to the cardiovascular system.44

Life Extension recognizes the value of COX-2 inhibitors in cancer treatment. Some progressive oncologists include COX-2 inhibitors in their anticancer protocols, but the numbers are still relatively few. The risks associated with traditional NSAIDs include gastrointestinal perforation, ulceration and bleeding, and less frequently, renal and liver damage, but the benefits of for certain cancer patients may outweigh these risks.

A number of natural COX-2 inhibitors are discussed in the protocol entitled “Cancer Adjuvant Therapy.”

Like COX-2, the COX-1 enzyme also catalyzes (mediates) the conversion of certain fatty acids into inflammatory end products in some cell types. Cancer cells are sometimes genetically altered in such a way that causes them to express higher levels of COX-1; this has been observed in several types of cancer including ovarian, colon, and head and neck cancers.45-47 Moreover, selectively inhibiting COX-1 with experimental compounds has demonstrated a marked reduction in viability of colon cancer cells and ovarian cancer cells.47,48

Experimental studies on breast cancer cells revealed that simultaneous inhibition of both COX-1 and COX-2 might exert a synergistic effect to combat cancer cell growth.49 The authors of this study concluded that "The significant and additive effects exhibited by the combination of COX-1 and -2 inhibitors and their effects on cell cycle suggest that these agents could become an effective treatment modality for carcinoma of the breast."

Aspirin is an NSAID, but its actions are unique in that it selectively inhibits COX-1 activity while modulating the expression of COX-2.50 The net result of this dualistic action is diminished production of harmful metabolites via COX-1 and a reduction in total COX-2 activity. Since both COX-1 and COX-2 are drivers of inflammatory cancer cell growth, aspirin is an important yet underappreciated anticancer drug.

At the forefront of the growing field of research into aspirin’s role as a cancer fighter is Professor Peter Rothwell of Oxford University. Having specialized primarily in cardiovascular medical research, he and his colleagues had at their disposal a trove of information compiled from eight massive studies examining the effect of aspirin therapy on cardiovascular health.

Among the most compelling of their findings:

  • Aspirin reduced the overall risk of death from cancer by approximately 20%.
  • Most of that benefit was due to a 30‒40% reduction in deaths occurring after five years of daily aspirin intake.
  • The reduction in deaths due to solid cancers was maintained for 20 years in studies in which data was available for that period of time.
  • These effects were consistent across all populations studied—despite their diversity in health histories.
  • A dose of just 75 mg daily was all that was required for the protective effect—higher doses did not increase the benefit.
  • The reduction in cancer deaths increased with age: peak effects were observed in people aged 55‒64 and remained high in those 65 years or older.
  • The effect of aspirin on reducing risk of fatal cancers was powerful enough to contribute to a significant reduction in mortality rates from all causes.

The data correlating aspirin therapy with colon cancer prevention proved particularly compelling. Rothwell’s team saw a 24% reduction in the risk of developing colon cancer over a 20-year period in patients who took aspirin daily and a 35% reduction in the risk of dying from colon cancer. The most potent preventive benefit was observed in cancers of the upper colon (the ascending and transverse colon).51

Separate observational studies have suggested a preventive effect for cancers of the esophagus, stomach, lung, breast, and ovaries.52-54 A 2010 study revealed that men taking regular aspirin supplements attained a 10% reduction in prostate cancer risk compared to men who took no aspirin.55 Another study showed a risk reduction of 24% in long-term users (greater than five years), and 29% in daily aspirin users.56

How to Implement Step Four

  • Take a low-dose aspirin each day, and;
  • Ask your physician to prescribe one of the following COX-2 inhibiting drugs:
    • Lodine XL, 1000 mg once daily, or
    • Celebrex, 100‒200 mg every 12 hours

Note: The use of Lodine and Celebrex has been associated with an increased risk of heart attack and stroke. The anti-cancer benefits of these drugs have to be weighed against these increased cardiovascular risks. Using aspirin in combination with a COX-2 inhibitor may increase the risk for bleeding, but also reduce cardiovascular risks; speak with your physician before combining aspirin with a COX-2 inhibitor.

6 Step Five: Suppressing Ras Oncogene Expression

The family of proteins known as Ras plays a central role in the regulation of cell growth. It fulfills this fundamental role by integrating the regulatory signals that govern the cell cycle and proliferation.

Defects in the Ras-Raf pathway can result in cancerous growth. Mutant Ras genes were among the first oncogenes identified for their ability to transform cells into a cancerous phenotype (ie, a cell observably altered because of distorted gene expression). Mutations in one of three genes (H, N, or K-Ras) encoding Ras proteins are associated with upregulated cell proliferation and are found in an estimated 30‒40% of all human cancers. The highest incidences of Ras mutations are found in cancers of the pancreas (80%), colon (50%), thyroid (50%), lung (40%), liver (30%), melanoma (30%), and myeloid leukemia (30%).57-63

The differences between oncogenes and normal genes can be slight. The mutant protein that an oncogene ultimately creates may differ from the healthy version by only a single amino acid, but this subtle variation can radically alter the protein's functionality.

The Ras-Raf pathway is used by human cells to transmit signals from the cell surface to the nucleus. Such signals direct cells to divide, differentiate, or even undergo programmed cell death (apoptosis).

A Ras gene usually behaves as a relay switch within the signal pathway that instructs the cell to divide. In response to stimuli transmitted to the cell from outside, cell-signaling pathways are activated. In the absence of stimulus, the Ras protein remains in the "off" position. A mutated Rasprotein gene behaves like a switch stuck on the "on" position, continuously misinforming the cell, instructing it to divide when the cycle should be turned off.64,65 Researchers have known for some time that injecting anti-Ras antibodies, specific for amino acid 12, cause a reversal of excessive proliferation and a transient alteration of the mutated cell to one of a normal appearance.66 Recently, scientists have taken advantage of the high frequency at which K-Ras is mutated in several types of cancer by developing vaccines that trigger the immune system to attack cells harboring this mutant protein. For example, a 2011 study found that patients with resected pancreatic cancer were much more likely to be alive 10 years post vaccination (20%) than those who did not receive the vaccine (0%).67

To establish new methods for diagnosing pancreatic cancer, K-Ras mutations were examined in the pancreatic juice of pancreatic cancer patients. Pancreatic juice was positive for K-Ras in 87.8% (36/41) of patients.68 When combined with p53 mutations in the stool and CA 19-9 (a blood marker for pancreatic cancer), it may be possible to identify the disease much earlier than by conventional diagnostic methods.69

Greater understanding regarding the activity of mutant Ras genes opens exciting avenues of treatment. Researchers found that precursor Ras genes must undergo several biochemical modifications to become mature, active versions. After such maturation, the Ras proteins attach to the inner surface of the cells outer membrane where they can interact with other cellular proteins and stimulate cell growth.

The events resulting in mature Ras genes take place in three steps, the most critical being the first—referred to as the farnesylation step. A specific enzyme, farnesyl-protein transferase (FPTase), speeds up the reaction. One strategy for blocking Ras protein activity has been to inhibit FPTase. Inhibitors of this enzyme block the maturation of Ras protein and reverse the cancerous transformation induced by mutant Ras genes.65

A number of natural substances impact the activity of Ras oncogenes. For example, limonene is a substance found in the essential oils of citrus products. Limonene has been shown to act as a farnesyl transferase inhibitor. Administering high doses of limonene to cancer-bearing animals blocks the farnesylation of Ras, thus inhibiting cell replication.70,71 Curcumin inhibited the farnesylation of RAS, and caused cell death in breast cancer cells expressing RAS mutations.72,73

Japanese researchers examined the effects of vitamin E on the presence of K-Ras mutations in mice with lung cancer. Prior to treatment with vitamin E, K-Ras mutations were present in 64% of the mice. After treatment with vitamin E, only 18% of the mice expressed K-Ras mutations.74 Vitamin E decreased levels of H-Ras proteins in cultured melanoma cells.75 A study conducted at Mercy Hospital of Pittsburgh also showed that diallyl disulfide, a naturally occurring organosulfide from garlic, inhibits p21 H-Ras oncogenes, displaying a significant restraining effect on tumor growth.76

Researchers at Rutgers University investigated the ability of different green and black tea polyphenols to inhibit H-Ras oncogenes. The Rutgers team found that all the major polyphenols contained in green and black tea except epicatechin showed strong inhibition of cell growth.77 Investigators at Texas A&M University also found that fish oil decreased colonic Ras membrane localization and reduced tumor formation in rats. In view of the central role of oncogenic Ras in the development of colon cancer, the finding that omega-3 fatty acids modulate Ras activation could explain why dietary fish oil protects against colon cancer.78

Statins are a class of popular cholesterol-lowering drugs. Mevacor (lovastatin), Zocor (simvastatin), and Pravachol (pravastatin) are statin drugs shown to inhibit the activity of Ras oncogenes.79 Statin drugs block the HMG-COA) reductase enzyme, which depletes cells of farnesyl pyrophosphate. This results in a reduction of activated farnesylated Ras.80

Illustrative of the potential of statin therapy, patients with primary liver cancer were treated with either the chemotherapeutic drug 5-FU or a combination of 5-FU and 40 mg/day of pravastatin.81 Median survival increased from nine months, among patients treated with only 5-FU, to 18 months when using 5-FU combined with the statin drug pravastatin (Pravachol). In 2008, German researchers studied the effects of pravastatin in patients with advanced liver cancer.82 One hundred thirty-one patients received chemoembolization alone, while 52 patients received chemoembolization plus pravastatin (20‒40mg). During the observation period of up to five years, 23.7% of the patients treated with chemoembolization alone had survived, compared to a 36.5% survival for the chemoembolization and pravastatin group. Median survival was 12 months for the chemoembolization only group, while the pravastatin group had a median survival of 20.9 months.

Statin drugs are known to deplete coenzyme Q10 (CoQ10) levels, therefore those taking a statin drug should supplement with CoQ10. For a detailed explanation, please consult the coenzyme Q10 section in the “Cancer Adjuvant Therapy” protocol.

Individuals with cancer should consider an immunohistochemistry test of their cancer tissue for mutated ras genes at Caris Life Sciences Laboratories (see Step One of this protocol), a recommendation the Life Extension first made in 1997. Life Extension strongly believes all cancer patients should undergo immunohistochemical testing to determine Ras status.

How to Implement Step Five

Ask your physician to prescribe one of the following statin drugs to inhibit the activity of Ras oncogenes:

  • Mevacor (lovastatin)
  • Zocor (simvastatin)
  • Pravachol (pravastatin)

Note: Statin drugs may generate adverse side effects. Physician oversight and careful surveillance with monthly blood tests (at least initially) to evaluate liver function, muscle enzymes, and lipid levels are suggested.

In addition to statin drug therapy, consider supplementing with the following nutrients to further suppress the expression of Ras oncogenes:

  • Curcumin: 400 – 800 mg daily
  • Fish oil: 2,100 mg of EPA and 1,500 mg of DHA daily with meals
  • Green tea; standardized extract: 725 – 1,450 mg EGCG daily
  • Aged garlic extract: 2,400 mg daily with meals
  • Vitamin E: 400 – 1,000 IU of natural alpha tocopherol along with at least 200 mg of gamma tocopherol daily with meals

7 Step Six: Correcting Coagulation Abnormalities

Both experimental and clinical data have determined that coagulation disorders are common in patients with cancer. Many cancer patients reportedly have a hypercoagulable state, with recurrent thrombosis (blood clot) due to the impact of cancer cells and chemotherapy on the coagulation cascade.83 Pulmonary embolism (blood clot in the lung) is a particular problem for patients with pancreatic and gastric cancer, colon cancer, and ovarian cancer.84 Thus, momentum is building for anticoagulant therapy through reports, the vast majority of which are derived from secondary analyses of clinical trials on the treatment of thromboembolism.

Research on low-molecular-weight heparin (LMWH)—an anticoagulant—shows promise in regard to increasing cancer survival rates. Data comparing unfractionated heparin to LMWH indicate that LMWH is equally beneficial if not more beneficial to cancer patients in terms of survival. The improved life expectancy gathered from anticoagulant therapy is not solely a result of the reduced complications from thromboembolism, but also from enzyme interactions, cellular growth modifications, and anti-angiogenic factors.85,86 It appears heparin inhibits the formation of cancer's vascular network by binding to angiogenic promoters (ie, basic fibroblast growth factor and VEGF).87

Another important aspect of anticoagulant therapy involves breaking down fibrin, a coagulation protein found in blood. Cancers employ various strategies to utilize fibrin for their own benefit. For example, fibrin covers cancer cells with a protective coat, hindering recognition by the immune system. In addition, fibrin relays a signal to the cancer to initiate angiogenesis—the growth of new blood vessels. As fibrin encourages a healthy vascular network and tumor growth increases, it sets the stage for metastasis.

German scientists evaluated whether cancer fatalities in women with previously untreated breast cancer were reduced using LMWH therapy. The study showed that breast cancer patients receiving LMWH had a lower rate of mortality during the first 650 days following surgery, compared to women receiving unfractionated heparin. The survival advantage was apparent after even a short course of therapy.88 In another study of 300 breast cancer patients, none of the trial participants developed metastasis while receiving anticoagulant therapy although 37 (12.3%) died from the disease.89

Similar advantages were evidenced among small cell lung cancer patients undergoing heparin therapy in conjunction with conventional treatments. When subjects were treated with heparin they enjoyed a better prognosis, with greater numbers of complete responses, longer median survival, and higher survival rates at 1, 2, and 3 years compared to patients who did not receive heparin.90

A comprehensive analysis of the data pertaining to all studies published on the impact of heparin treatment on survival in cancer patients determined that treatment with heparin (both unfractionated heparin and LMWH) decreased the risk of death by 23%, compared to those who did not receive heparin.91

How to Implement Step Six

Ascertain if you are in a hypercoagulable state by having your blood tested for prothrombin time (PT), partial thromboplastin time (PTT), and D-dimers. A hypercoagulable state is suggested if the shortening of the PT and PTT are seen in conjunction with elevation of D-dimers (see table after next paragraph on laboratory tests for hypercoagulability).

If there is any evidence of a hypercoagulable (prethrombotic) state, ask your physician to prescribe the appropriate individualized dose of low-molecular-weight heparin (LMWH). Repeat the prothrombin blood test every two weeks.

Lab Tests for Hypercoagulability

Tests Routinely Available

Results if Hypercoagulable

Tests Requiring Dedicated Coagulation Laboratory

Results if Hypercoagulable

Prothrombin time (PT)

Less than normal

Alpha-1 antitrypsin (A1AT)

Elevated

Partial thromboplastin time (PTT)

Less than normal

Euglobulin clot lysis time (ECLT)

Prolonged

Platelet count (part of CBC)

Elevated

Factor VIII levels

Elevated

8 Step Seven: Inhibiting Angiogenesis

Angiogenesis—the growth of new blood vessels—is critical during fetal development but occurs minimally in healthy adults. Exceptions occur during wound healing, inflammation, following a myocardial infarction, in female reproductive organs, and in pathologic conditions such as cancer.110,111

Angiogenesis is a strictly controlled process in the healthy adult human body, a process regulated by endogenous angiogenic promoters and inhibitors. Dr. Judah Folkman, the father of the angiogenesis theory of cancer stated, "Blood vessel growth is controlled by a balancing of opposing factors. A tilt in favor of stimulators over inhibitors might be what trips the lever and begins the process of tumor angiogenesis."112

Solid tumors cannot grow beyond the size of a pinhead without inducing the formation of new blood vessels to supply the nutritional needs of the tumor.113 Since rapid vascularization and tumor growth appear to occur concurrently, interrupting the formation of new blood vessels is paramount to overcoming the malignancy.114

Tumor angiogenesis results from a cascade of molecular and cellular events, usually initiated by the release of angiogenic growth factors. At a critical phase in the growth of a cancer, signal molecules are secreted from the cancer to nearby endothelial cells to activate new blood vessel growth. These angiogenic growth factors diffuse in the direction of preexisting blood vessels, encouraging the formation of new blood vessel growth.115,116 VEGF and basic fibroblast growth factors are expressed by many tumors and appear to be particularly important for angiogenesis.117

A number of natural substances, such as curcumin, green tea, N-acetyl cysteine (NAC), resveratrol, grape seed-skin extract, and vitamin D have anti-angiogenic properties. For further discussion, see the “Cancer Adjuvant Therapy” protocol.

The FDA has approved an anti-angiogenesis drug called Avastin (bevacizumab), but it has demonstrated severe side effects and often only mediocre efficacy. Several other drugs inhibit angiogensis as secondary mechanisms and are sometimes utilized in cancer therapy. These included sorafenib, sunitinb, pazopanib, and everolimus. These options should be discussed with a healthcare professional because these drugs may cause considerable side effects, and are only FDA approved for specific types of cancer.

How to Implement Step Seven

  • There are clinical trials using other anti-angiogenesis agents. Log on to www.cancer.gov/clinicaltrials to find out if you are eligible to participate.
  • Several nutrients have demonstrated potential anti-angiogenesis effects such as green tea extract and curcumin.
    • Green tea; standardized extract: 725 – 1,450 mg daily
    • Curcumin: 400 – 800 mg daily
    • Vitamin D: 5,000 – 8,000 IU daily (depending on blood levels)
    • Grape extract (seed and skin): 150 – 300 mg daily
    • N-acetyl cysteine: 600 – 1200 mg daily

9 Step Eight: Inhibiting the 5-Lipoxygenase (5-LOX) Enzyme

As discussed in step 4 of this protocol: Inhibiting the Cyclooxygenase-2 (COX-2) Enzyme, the scientific literature has demonstrated that inflammation plays a pivotal role in the formation and progression of cancer.

The 5-lipoxygenase (5-LOX) enzyme is another inflammatory enzyme that can contribute to the formation and progression of cancer. Arachidonic acid—a polyunsaturated fat found in high concentrations in meat and dairy products—promotes elevation of the 5-LOX enzyme. A growing number of studies have documented that 5-LOX directly stimulates prostate cancer cell proliferation via several well-defined mechanisms.118-126 In addition, arachidonic acid is metabolized by 5-LOX to 5-HETE, a potent survival factor that prostate cancer cells utilize to escape destruction.121,127-130

In response to arachidonic acid overload, the body increases its production of enzymes like 5-lipooxygenase (5-LOX) to degrade arachidonic acid. Not only does 5-LOX directly stimulate cancer cell propagation,121,131-140 but the breakdown products that 5-LOX produces from arachidonic acid (such as leukotriene B4, 5-HETE, and hydroxylated fatty acids) cause tissue destruction, chronic inflammation, and increased resistance of tumor cells to apoptosis (programmed cell destruction).120,127,141-145

Based on studies showing that consumption of foods rich in arachidonic acid is greatest in regions with high incidences of prostate cancer,119,120,125,131 scientists sought to determine how much of the 5-LOX enzyme is present in malignant versus benign prostate tissues. Using prostate biopsy samples, the researchers found that 5-LOX levels were an astounding six-fold greater in malignant prostate tissues compared to benign tissues. This study also found that levels of 5-HETE were 2.2-fold greater in malignant versus benign prostate tissues.123 The scientists concluded this study by stating that selective inhibitors of 5-LOX may be useful in the prevention or treatment of patients with prostate cancer.

As the evidence mounts that consuming saturated fats increase prostate cancer risk, scientists are evaluating the effects of 5-LOX on various growth factors involved in the progression, angiogenesis, and metastasis of cancer cells. One study found that 5-LOX activity is required to stimulate prostate cancer cell growth by epidermal growth factor (EGF) and other cancer cell proliferating factors produced in the body. When 5-LOX levels were reduced, the cancer cell stimulatory effect of EGF and other growth factors was diminished.120

In a mouse study, an increase in 5-LOX resulted in a corresponding increase in vascular endothelial growth factor (VEGF), a key growth factor that tumor cells use to stimulate new blood vessel formation (angiogenesis) into the tumor. 5-LOX inhibitors were shown to reduce tumor angiogenesis along with a host of other growth factors.146 Chronic inflammation is tightly linked to the induction of aberrant angiogenesis used by cancer cells to facilitate the growth of new blood vessels (angiogenesis) into tumors.147

In both androgen-dependent and androgen-independent human prostate cancer cell lines, the inhibition of 5-lipoxygenase (5-LOX) has consistently been shown to induce rapid and massive apoptosis (cancer cell destruction).119,131,148-151

As humans age, chronic inflammatory processes can cause the over-expression of 5-LOX in the body. Excess 5-LOX may contribute to the development and progression of prostate cancer in aging males.152

Based on the cumulative knowledge that 5-LOX can promote the invasion and metastasis of prostate cancer cells, it would appear advantageous to take aggressive steps to suppress this lethal enzyme. A critical approach to decreasing 5-LOX activity in the body is to decrease the consumption of saturated and omega-6 fats that contain high concentrations of arachidonic acid and high glycemic carbohydrates that contribute to arachidonic acid formation. Another worthwhile approach is to supplement with fish oil, which reduces 5-LOX activity in the body.153,154 Studies show that lycopene and saw palmetto extract also help to suppress 5-LOX.148,155-168 The suppression of 5-LOX by these nutrients may partially account for their favorable effects on the prostate gland.

Specific extracts from the boswellia plant selectively inhibit 5-lipoxygenase (5-LOX).169,170 In several well-controlled human studies, boswellia has been shown to be effective in alleviating various chronic inflammatory disorders.123,171-179 Scientists have discovered that the specific constituent in boswellia responsible for suppressing 5-LOX is AKBA (3-O-acetyl-11-keto-B-boswellic acid). Boswellia-derived AKBA binds directly to 5-LOX and inhibits its activity. Other boswellic acids only partially and incompletely inhibit 5-LOX.170,180

Researchers have discovered how to obtain an economically viable boswellia extract standardized to contain a greater than 20% concentration of AKBA. A novel boswellia extract has been developed that is 52% more bioavailable compared to standard boswellia extracts,181,182 thus providing a greater opportunity to suppress deadly 5-LOX and other cancer-promoting byproducts of arachidonic acid. This more bioavailable AKBA extraction discovery was patented and given the trademark name AprèsFlex.

How to Implement Step Eight

Decrease the consumption of saturated and omega-6 fats that contain high concentrations of arachidonic acid, such as meats, dairy products, and egg yolks, along with high-glycemic carbohydrates.

Consider supplementing with the following nutrients to suppress 5-LOX enzyme activity:

  • AprèsFlex: 100 to 400 mg daily
  • Fish oil: 2,100 mg of EPA and 1,500 mg of DHA daily with meals
  • Lycopene: 30 mg daily with meals
  • Curcumin: 400 – 800 mg daily

10 Step Nine: Inhibiting Cancer Metastasis

The surgical removal of the primary tumor has been the cornerstone of treatment for the great majority of cancers. The rationale for this approach is straightforward: if you can get rid of the cancer by simply removing it from the body, then a cure can likely be achieved. Unfortunately, this approach does not take into account that after surgery the cancer will frequently metastasize (spread to different organs). Quite often, the metastatic recurrence is far more serious than the original tumor. In fact, for many cancers, it is the metastatic recurrence—and not the primary tumor—that ultimately proves to be fatal.183

One mechanism by which surgery increases the risk of metastasis is by enhancing cancer cell adhesion.184 Cancer cells that have broken away from the primary tumor utilize adhesion to boost their ability to form metastases in distant organs. These cancer cells must be able to clump together and form colonies that can expand and grow. It is unlikely that a single cancer cell will form a metastatic tumor, just as one person is unlikely to form a thriving community. Cancer cells use adhesion molecules—such as galectin-3—to facilitate their ability to clump together. Present on the surface of cancer cells, these molecules act like velcro by allowing free-standing cancer cells to adhere to each other.185

Cancer cells circulating in the bloodstream also make use of galectin-3 surface adhesion molecules to latch onto the lining of blood vessels.186 The adherence of CTCs to the blood vessel walls is an essential step for the process of metastasis. A cancer cell that cannot adhere to the blood vessel wall will just continue to wander through the blood stream incapable of forming metastases. Unable to latch onto the wall of the blood vessel, these circulating tumor cells become like "ships without a port" and are unable to dock. Eventually, white blood cells circulating in the bloodstream will target and destroy the CTC. If the CTCs successfully bind to the blood vessel wall and burrow their way through the basement membrane, they will then utilize galectin-3 adhesion molecules to adhere to the organ to form a new metastatic cancer.185

Regrettably, research has shown that cancer surgery increases tumor cell adhesion.187 Therefore, it is critically important for the person undergoing cancer surgery to take measures that can help to neutralize the surgery-induced increase in cancer cell adhesion.

Fortunately, a natural compound called modified citrus pectin (MCP) can do just that. Citrus pectin—a type of dietary fiber—is not absorbed from the intestine. However, MCP has been altered so that it can be absorbed into the blood and exert its anti-cancer effects. The mechanism by which MCP inhibits cancer cell adhesion is by binding to galectin-3 adhesion molecules on the surface of cancer cells, thereby preventing cancer cells from sticking together and forming a cluster. MCP can also inhibit circulating tumor cells from latching onto the lining of blood vessels. This was demonstrated by an experiment in which MCP blocked the adhesion of galectin-3 to the lining of blood vessels by an astounding 95%. MCP also substantially decreased the adhesion of breast cancer cells to the blood vessel walls.188

After these exciting findings in animal research, MCP was then put to the test in men with prostate cancer. In this trial, 10 men with recurrent prostate cancer received MCP (14.4 grams per day). After one year, a considerable improvement in cancer progression was noted, as determined by a reduction of the rate at which the prostate-specific antigen (PSA) level increased.189 This was followed by a study in which 49 men with prostate cancer of various types were given MCP for a four-week cycle. After two cycles of treatment with MCP, 22% of the men experienced a stabilization of their disease or improved quality of life; 12% had stable disease for more than 24 weeks. The authors of the study concluded that "MCP (modified citrus pectin) seems to have positive impacts especially regarding clinical benefit and life quality for patients with far advanced solid tumor."190

In addition to MCP, a well-known OTC medication can also play a pivotal role in reducing cancer cell adhesion. Cimetidine—commonly known as Tagamet—is a drug historically used to alleviate heartburn. A growing body of scientific evidence has revealed that cimetidine also possesses potent anti-cancer activity.

Cimetidine inhibits cancer cell adhesion by blocking the expression of an adhesive molecule—called E-selectin—on the surface of cells lining blood vessels. Cancers cells latch onto E-selectin in order to adhere to the lining of blood vessels.191 By preventing the expression of E-selectin, cimetidine significantly limits the ability of cancer cell adherence to the blood vessel walls. This effect is analogous to removing the velcro from the blood vessels walls that would normally enable circulating tumor cells to bind.

Cimetidine’s potent anti-cancer effects were clearly displayed in a report published in the British Journal of Cancer in 2002. In this study, 64 colon cancer patients received chemotherapy with or without cimetidine (800 mg per day) for one year. The 10-year survival for the cimetidine group was almost 90%. This is in stark contrast to the control group, which had a 10-year survival of only 49.8%. Remarkably, for those patients with a more aggressive form of colon cancer, the 10-year survival was 85% in those treated with cimetidine compared to a dismal 23% in the control group.192 The authors of the study concluded, "Taken together, these results suggested a mechanism underlying the beneficial effect of cimetidine on colorectal cancer patients, presumably by blocking the expression of E-selectin on vascular endothelial [lining of blood vessels] cells and inhibiting the adhesion of cancer cells." These findings are supported by another study with colorectal cancer patients wherein cimetidine given for just seven days at the time of surgery increased three-year survival from 59% to 93%.193

Another major contributor to cancer metastasis is immune dysfunction; primarily that which occurs immediately following a surgical procedure such as removal of a primary tumor.194 Specifically, surgery suppresses the number of specialized immune cells called natural killer (NK) cells, which are a type of white blood cell tasked with seeking out and destroying cancer cells.

To illustrate the importance of NK cell activity in fighting cancer, a study published in the journal Breast Cancer Research and Treatment examined NK cell activity in women shortly after surgery for breast cancer. The researchers reported that low levels of NK cell activity were associated with an increased risk of death from breast cancer.191 In fact, reduced NK cell activity was a better predictor of survival than the actual stage of the cancer. In another alarming study, individuals with reduced NK cell activity before surgery for colon cancer had a 350% increased risk of metastasis during the following 31 months.195

One prominent natural compound that can increase NK cell activity is PSK, (protein-bound polysaccharide K) a specially prepared extract from the mushroom Coriolus versicolor. PSK has been shown to enhance NK cell activity in multiple studies.196,197 PSK’s ability to enhance NK cell activity helps to explain why it has been shown to dramatically improve survival in cancer patients. For example, 225 patients with lung cancer received radiation therapy with or without PSK (3 grams per day). For those with more advanced Stage 3 cancers, more than three times as many individuals taking PSK were alive after five years (26%), compared to those not taking PSK (8%). PSK more than doubled five-year survival in those individuals with less advanced stage 1 or 2 disease (39% vs.17%).198

In a 2008 study, a group of colon cancer patients were randomized to receive chemotherapy alone or chemotherapy plus PSK, which was taken for two years. The group receiving PSK had an exceptional 10-year survival of 82%. Sadly, the group receiving chemotherapy alone had a 10-year survival of only 51%.199 In a similar trial reported in the British Journal of Cancer, colon cancer patients received chemotherapy alone or combined with PSK (3 grams per day) for two years. In the group with a more dangerous stage 3 colon cancer, the five-year survival was 75% in the PSK group. This compared to a five-year survival of only 46% in the group receiving chemotherapy alone.200 Additional research has shown that PSK improves survival in cancers of the breast, stomach, esophagus, and uterus as well.201-203

Methylmalonic Acid and Metastasis

Cancer susceptibility and mortality increase substantially with age.204,205 This is due, in part, to accumulated mutations and exposure to mutagens over the course of a lifetime. However, metabolic changes due to age also appear to play a significant role in promoting cancer cell aggressiveness and metastasis. A 2020 study published in the journal Nature provided evidence that methylmalonic acid (MMA), a metabolic by-product of protein and fat digestion, and a marker of vitamin B12 deficiency, increases with age and can endow cancer cells with more aggressive properties.206,207 The researchers cultured cancer cells with serum from young and old healthy donors and found that the cells cultured in serum from older donors tended to undergo a transition that resembled the epithelial-mesenchymal transition (EMT), a process through which cancer forms and progresses, and becomes aggressive and metastatic.208 In addition, the older serum promoted resistance to two common chemotherapeutic drugs.206

The researchers then analyzed the metabolic composition of the older and younger serums to determine what caused the change; they found the old serum had significantly higher levels of three compounds, including MMA, and that of them, only MMA induced an epithelial-mesenchymal-like transition in cancer cells. The elevation in MMA could be due to deregulation of the enzymes in the propionate metabolic pathway, which breaks down certain amino acids and fats, and/or vitamin B12 deficiency, which is a necessary cofactor in the pathway. Further study indicated that MMA appeared to upregulate the transcription factor gene SOX4—a “master regulator” of EMT that is overexpressed in many aggressive cancers.206,209,210 Interestingly, MMA relied on lipids in the serum in order to penetrate cancer cells.206 Given the results of this study, it may be wise for older individuals and/or cancer patients to take steps to reduce MMA levels, such as maintaining adequate B12 levels, avoiding excessive protein intake (especially from animal sources), and possibly lowering blood lipids like cholesterol. However, more research is needed before firm conclusions can be drawn regarding the utility of these approaches to lower MMA in the context of cancer and cancer prevention.

How to Implement Step Nine

The following three novel compounds have shown efficacy in inhibiting several mechanisms that contribute to cancer metastasis. It is especially important to consider these compounds during the perioperative period (period before and after surgery), because a known consequence of surgery is an enhanced proclivity for metastasis.

  • Modified citrus pectin: 15 grams daily, in three divided doses
  • Cimetidine: 800 mg daily, in two divided doses
  • Coriolus versicolor; standardized extract: 1,200 – 3,600 mg daily

Note: Of critical importance to treatment-naïve patients is implementing as many of the 9 critical steps as can safely be done concurrently with conventional therapy. In newly diagnosed patients who have not yet been treated, the objective is to eradicate the primary tumor and metastatic cells with a multi-pronged "first strike therapy" so that residual tumor cells are not given an opportunity to evolve survival mechanisms that make them resistant to further treatments. Omitting any of the 9 steps may provide an opening for residual cancer cells to mutate in a way that makes them very difficult to treat a second time.

Life Extension oncology Wellness Specialists are available to provide clarification on any of the steps in this protocol; they can be reached at 800-226-2370.

2020

  • Sep: Added section on methylmalonic acid and metastasis to Step Nine: Inhibiting Cancer Metastasis

2011

  • Nov: Comprehensive update & review

Disclaimer and Safety Information

This information (and any accompanying material) is not intended to replace the attention or advice of a physician or other qualified health care professional. Anyone who wishes to embark on any dietary, drug, exercise, or other lifestyle change intended to prevent or treat a specific disease or condition should first consult with and seek clearance from a physician or other qualified health care professional. Pregnant women in particular should seek the advice of a physician before using any protocol listed on this website. The protocols described on this website are for adults only, unless otherwise specified. Product labels may contain important safety information and the most recent product information provided by the product manufacturers should be carefully reviewed prior to use to verify the dose, administration, and contraindications. National, state, and local laws may vary regarding the use and application of many of the therapies discussed. The reader assumes the risk of any injuries. The authors and publishers, their affiliates and assigns are not liable for any injury and/or damage to persons arising from this protocol and expressly disclaim responsibility for any adverse effects resulting from the use of the information contained herein.

The protocols raise many issues that are subject to change as new data emerge. None of our suggested protocol regimens can guarantee health benefits. Life Extension has not performed independent verification of the data contained in the referenced materials, and expressly disclaims responsibility for any error in the literature.

  1. Tan WW et al. Metastatic Cancer With Unknown Primary Site. Medscape web page. Background. Available at: http://emedicine.medscape.com/article/280505-overview. Accessed 10/24/2011.
  2. Lancet. [no authors listed] Polychemotherapy for early breast cancer: an overview of the randomised trials. Early Breast Cancer Trialists' Collaborative Group. Lancet. 1998 Sep 19;352(9132):930-42.
  3. Clarke MJ. WITHDRAWN: Multi-agent chemotherapy for early breast cancer. Cochrane Database Syst Rev. 2008 Oct 8;(4):CD000487.
  4. Gelber RD, Cole BF, Goldhirsch A, Rose C, Fisher B, Osborne CK, Boccardo F, Gray R, Gordon NH, Bengtsson NO, Sevelda P. Adjuvant chemotherapy plus tamoxifen compared with tamoxifen alone for postmenopausal breast cancer: meta-analysis of quality-adjusted survival. Lancet. 1996 Apr 20;347(9008):1066-71.
  5. International Breast Cancer Study Group (IBCSG). Endocrine responsiveness and tailoring adjuvant therapy for postmenopausal lymph node-negative breast cancer: a randomized trial. J Natl Cancer Inst. 2002; 94(14):1054-65.
  6. Hayes DF, Thor AD, Dressler LG, Weaver D, Edgerton S, Cowan D, Broadwater G, Goldstein LJ, Martino S, Ingle JN, Henderson IC, Norton L, Winer EP, Hudis CA, Ellis MJ, Berry DA; Cancer and Leukemia Group B (CALGB) Investigators. HER2 and response to paclitaxel in node-positive breast cancer. N Engl J Med. 2007 Oct 11;357(15):1496-506.
  7. Moore, A. Breast-cancer therapy—looking back to the future. N Engl J Med. 2007 Oct 11;357(15):1547-9.
  8. Gennari A, Sormani MP, Pronzato P, Puntoni M, Colozza M, Pfeffer U, Bruzzi P. HER2 status and efficacy of adjuvant anthracyclines in early breast cancer: a pooled analysis of randomized trials. J Natl Cancer Inst. 2008 Jan 2;100(1):14-20.
  9. Swain SM, Whaley FS, Ewer MS. Congestive heart failure in patients treated with doxorubicin: a retrospective analysis of three trials. Cancer. 2003 Jun 1;97(11):2869-79.
  10. Kuhnl A et al. High expression of IGFBP2 is associated with chemoresistance in adult acute myeloid leukemia. Leuk Res. 2011 Sep 5.
  11. Kuukasjärvi T, Karhu R, Tanner M, Kähkönen M, Schäffer A, Nupponen N, Pennanen S, Kallioniemi A, Kallioniemi OP, Isola J. Genetic heterogeneity and clonal evolution underlying development of asynchronous metastasis in human breast cancer. Cancer Res. 1997 Apr 15;57(8):1597-604.
  12. Meng S, Tripathy D, Shete S, Ashfaq R, Haley B, Perkins S, Beitsch P, Khan A, Euhus D, Osborne C, Frenkel E, Hoover S, Leitch M, Clifford E, Vitetta E, Morrison L, Herlyn D, Terstappen LW, Fleming T, Fehm T, Tucker T, Lane N, Wang J, Uhr J. HER-2 gene amplification can be acquired as breast cancer progresses. Proc Natl Acad Sci U S A. 2004 Jun 22;101(25):9393-8.
  13. Wülfing P, Borchard J, Buerger H, et al. HER2-positive circulating tumor cells indicate poor clinical outcome in stage I and III breast cancer patients. Clin Cancer Res. 2006;12(6):1715-20.
  14. Cristofanilli M et al. Circulating tumor cells in metastatic breast cancer: biologic staging beyond tumor burden. Clin Breast Cancer. 2007 Feb;7(6):471-9.
  15. Moreno JG, Miller MC, Gross S, Allard WJ, Gomella LG, Terstappen LW. Circulating tumor cells predict survival in patients with metastatic prostate cancer. Urology. 2005 Apr;65(4):713-8.
  16. Tombal B, Van Cangh PJ, Loric S, Gala JL. Prognostic value of circulating prostate cells in patients with a rising PSA after radical prostatectomy. Prostate. 2003 Aug 1;56(3):163-70.
  17. Pound CR, Partin AW, Eisenberger MA, Chan DW, Pearson JD, Walsh PC. Natural history of progression after PSA elevation following radical prostatectomy. JAMA. 1999;281:1591-1597
  18. Halabi S, Small EJ, Hayes DF, Vogelzang NJ, Kantoff PW. Prognostic significance of reverse transcriptase polymerase chain reaction for prostate-specific antigen in metastatic prostate cancer: a nested study within CALGB 9583. J Clin Oncol. 2003 Feb 1;21(3):490-5.
  19. Danila DC, Heller G, Gignac GA, Gonzalez-Espinoza R, Anand A, Tanaka E, Lilja H, Schwartz L, Larson S, Fleisher M, Scher HI. Circulating tumor cell number and prognosis in progressive castration-resistant prostate cancer. Clin Cancer Res. 2007 Dec 1;13(23):7053-8.
  20. Hayes DF, Cristofanilli M, Budd GT, Ellis MJ, Stopeck A, Miller MC, Matera J, Allard WJ, Doyle GV, Terstappen LW. Circulating tumor cells at each follow-up time point during therapy of metastatic breast cancer patients predict progression-free and overall survival. Clin Cancer Res. 2006 Jul 15;12(14 Pt 1):4218-24.
  21. Cohen SJ, Punt CJ, Iannotti N, Saidman BH, Sabbath KD, Gabrail NY, Picus J, Morse M, Mitchell E, Miller MC, Doyle GV, Tissing H, Terstappen LW, Meropol NJ. Relationship of circulating tumor cells to tumor response, progression-free survival, and overall survival in patients with metastatic colorectal cancer. J Clin Oncol. 2008 Jul 1;26(19):3213-21.
  22. Quintela-Fandino M, López JM, Hitt R, Gamarra S, Jimeno A, Ayala R, Hornedo J, Guzman C, Gilsanz F, Cortés-Funes H. Breast cancer-specific mRNA transcripts presence in peripheral blood after adjuvant chemotherapy predicts poor survival among high-risk breast cancer patients treated with high-dose chemotherapy with peripheral blood stem cell support. J Clin Oncol. 2006 Aug 1;24(22):3611-8.
  23. Pachmann K, Dengler R, Lobodasch K, Fröhlich F, Kroll T, Rengsberger M, Schubert R, Pachmann U. An increase in cell number at completion of therapy may develop as an indicator of early relapse: Quantification of circulating epithelial tumor cells (CETC) for monitoring of adjuvant therapy in breast cancer. J Cancer Res Clin Oncol. 2008 Jan;134(1):59-65.
  24. Di Leo A, Gancberg D, Larsimont D, Tanner M, Jarvinen T, Rouas G, Dolci S, Leroy JY, Paesmans M, Isola J, Piccart MJ. HER-2 amplification and topoisomerase IIalpha gene aberrations as predictive markers in node-positive breast cancer patients randomly treated either with an anthracycline-based therapy or with cyclophosphamide, methotrexate, and 5-fluorouracil. Clin Cancer Res. 2002 May;8(5):1107-16
  25. Ciaparrone M, Quirino M, Schinzari G, et al. Predictive role of thymidylate synthase, dihydropyrimidine dehydrogenase and thymidine phosphorylase expression in colorectal cancer patients receiving adjuvant 5-fluorouracil. Oncology. 2006;70(5):366-77.
  26. Siwak DR, Shishodia S, Aggarwal BB, Kurzrock R. Curcumin-induced antiproliferative and proapoptotic effects in melanoma cells are associated with suppression of IkappaB kinase and nuclear factor kappaB activity and are independent of the B-Raf/mitogen-activated/extracellular signal-regulated protein kinase pathway and the Akt pathway. Cancer. 2005 Aug 15;104(4):879-90.
  27. Hayeshi R, Mutingwende I, Mavengere W, Masiyanise V, Mukanganyama S. The inhibition of human glutathione S-transferases activity by plant polyphenolic compounds ellagic acid and curcumin. Food Chem Toxicol. 2007 Feb;45(2):286-95.
  28. Sears B. A Dietary Road Map.1995.
  29. Newmark T. Beyond Aspirin. 2000.
  30. Chakraborti AK, Garg SK, Kumar R, Motiwala HF, Jadhavar PS. Progress in COX-2 Inhibitors: A Journey So Far. Curr Med Chem. 2010 Feb 18.
  31. Tucker ON, Dannenberg AJ, Yang EK, et al. Cyclooxygenase-2 expression is up-regulated in human pancreatic cancer. Cancer Res. 1999 Mar 1;59(5):987-90.
  32. Gately S. The contributions of cyclooxygenase-2 to tumor angiogenesis. Cancer Metastasis Rev. 2000;19(1-2):19-27.
  33. Reddy BS, Rao CV. Colon cancer: a role for cyclo-oxygenase-2-specific nonsteroidal anti-inflammatory drugs. Drugs Aging. 2000 May;16(5):329-34.
  34. Sheehan KM, Sheahan K, O'Donoghue DP, et al. The relationship between cyclooxygenase-2 expression and colorectal cancer. JAMA. 1999 Oct 6;282(13):1254-7.
  35. Brody JE. Aspirin linked to aid in colon cancer fight. New York Times. 1991;141(48, 812):A12.
  36. Knorr J. Aspirin: the multi-purpose compound not just for headaches anymore. Life Extension Magazine 2000 Feb. 2000; Life Extension Magazine 2000 Feb6(2):50.
  37. Agarwal B, Rao CV, Bhendwal S, et al. Lovastatin augments sulindac-induced apoptosis in colon cancer cells and potentiates chemopreventive effects of sulindac. Gastroenterology. 1999 Oct;117(4):838-47.
  38. Tsujii M, Kawano S, Tsuji S, et al. Cyclooxygenase regulates angiogenesis induced by colon cancer cells. Cell. 1998 May 29;93(5):705-16.
  39. Valsecchi ME, Pomerantz SC, Jaslow R, Tester W. Reduced risk of bone metastasis for patients with breast cancer who use COX-2 inhibitors. Clin Breast Cancer. 2009 Nov;9(4):225-30.
  40. Edelman MJ, Watson D, Wang X, et al. Eicosanoid modulation in advanced lung cancer: cyclooxygenase-2 expression is a positive predictive factor for celecoxib + chemotherapy—Cancer and Leukemia Group B Trial 30203. J Clin Oncol. 2008 Feb;26(6):848-55.
  41. Pruthi RS, Derksen JE, Moore D, e al. Phase II trial of celecoxib in prostate-specific antigen recurrent prostate cancer after definitive radiation therapy or radical prostatectomy. Clin Cancer Res. 2006 Apr;12(7 Pt 1):2172-7.
  42. Manola J, Atkins M, Ibrahim J, et al. Prognostic factors in metastatic melanoma: a pooled analysis of Eastern Cooperative Oncology Group trials. J Clin Oncol. 2000 Nov;18(22):3782-93.
  43. Lai V, George J, Richey L, et al. Results of a pilot study of the effects of celecoxib on cancer cachexia in patients with cancer of the head, neck, and gastrointestinal tract. Head Neck. 2008 Jan;30(1):67-74.
  44. Harris RE. Cyclooxygenase-2 (cox-2) blockade in the chemoprevention of cancers of the colon, breast, prostate, and lung. Inflammopharmacology. 2009 Apr;17(2):55-67.
  45. Erovic BM et al. Strong evidence for up-regulation of cyclooxygenase-1 in head and neck cancer. Eur J Clin Invest. 2008 Jan;38(1):61-6.
  46. Khunnarong J et al. Expression of cyclooxygenase-1 in epithelial ovarian cancer: a clinicopathological study. Asian Pac J Cancer Prev. 2008 Oct-Dec;9(4):757-62.
  47. Wu WK et al. Inhibition of cyclooxygenase-1 lowers proliferation and induces macroautophagy in colon cancer cells. Biochem Biophys Res Commun. 2009 Apr 24;382(1):79-84. Epub 2009 Mar 1.
  48. Li W et al. Effects of a cyclooxygenase-1-selective inhibitor in a mouse model of ovarian cancer, administered alone or in combination with ibuprofen, a nonselective cyclooxygenase inhibitor. Med Oncol. 2009;26(2):170-7. Epub 2008 Nov 6.
  49. McFadden DW et al. Additive effects of Cox-1 and Cox-2 inhibition on breast cancer in vitro. Int J Oncol. 2006 Oct;29(4):1019-23.
  50. Wu KK et al. Aspirin and other cyclooxygenase inhibitors: new therapeutic insights. Semin Vasc Med. 2003 May;3(2):107-12.
  51. Rothwell PM, Wilson M, Elwin CE, et al. Long-term effect of aspirin on colorectal cancer incidence and mortality: 20-year follow-up of five randomised trials. Lancet. 2010 Nov 20;376(9754):1741-50.
  52. Sjodahl R. Nonsteroidal anti-inflammatory drugs and the gastrointestinal tract. Extent, mode, and dose dependence of anticancer effects. Am J Med. 2001 Jan 8;110(1A):66S-69S.
  53. Corley DA, Kerlikowske K, Verma R, Buffler P. Protective association of aspirin/NSAIDs and esophageal cancer: a systematic review and meta-analysis. Gastroenterology. 2003 Jan;124(1):47-56.
  54. Bosetti C, Gallus S, La Vecchia C. Aspirin and cancer risk: a summary review to 2007. Recent Results Cancer Res. 2009;181:231-51.
  55. Dhillon PK, Kenfield SA, Stampfer MJ, Giovannucci EL. Long-term aspirin use and the risk of total, high-grade, regionally advanced and lethal prostate cancer in a prospective cohort of health professionals, 1988-2006. Int J Cancer. 2010 Dec 2.
  56. Salinas CA, Kwon EM, FitzGerald LM, et al. Use of aspirin and other nonsteroidal antiinflammatory medications in relation to prostate cancer risk. Am J Epidemiol. 2010 Sep 1;172(5):578-90.
  57. Duursma AM, Agami R. Ras interference as cancer therapy. Semin Cancer Biol. 2003 Aug;13(4):267-73.
  58. Minamoto T, Mai M, Ronai Z. K-ras mutation: early detection in molecular diagnosis and risk assessment of colorectal, pancreas, and lung cancers--a review. Cancer Detect Prev. 2000;24(1):1-12.
  59. Vachtenheim J. Occurrence of ras mutations in human lung cancer. Minireview. Neoplasma. 1997;44(3):145-9.
  60. Bartram CR. Mutations in ras genes in myelocytic leukemias and myelodysplastic syndromes. Blood Cells. 1988;14(2-3):533-8.
  61. Bos JL. ras oncogenes in human cancer: a review. Cancer Res. 1989 Sep;49(17):4682-9.
  62. Hsieh JS, Lin SR, Chang MY, Chen FM, Lu CY, Huang TJ, Huang YS, Huang CJ, Wang JY. APC, K-ras, and p53 gene mutations in colorectal cancer patients: correlation to clinicopathologic features and postoperative surveillance. Am Surg. 2005 Apr;71(4):336-43.
  63. Däbritz J, Preston R, Hänfler J, Oettle H. Follow-up study of K-ras mutations in the plasma of patients with pancreatic cancer: correlation with clinical features and carbohydrate antigen 19-9. Pancreas. 2009 Jul;38(5):534-41.
  64. Gibbs JB, Kohl NE, Koblan KS, et al. Farnesyltransferase inhibitors and anti-Ras therapy. Breast Cancer Res Treat. 1996;38(1):75-83.
  65. Oliff A. Therapies of the Future: New Molecular Targets for Cancer Therapy. 1996.
  66. Feramisco JR, Clark R, Wong G, et al. Transient reversion of ras oncogene-induced cell transformation by antibodies specific for amino acid 12 of ras protein. Nature. 1985 Apr 18;314(6012):639-42.
  67. Weden S et al. Long-term follow-up of patients with resected pancreatic cancer following vaccination against mutant K-ras. Int J Cancer. 2011 Mar 1;128(5):1120-8. doi: 10.1002/ijc.25449.
  68. Lu X, Xu T, Qian J. [An application value of detecting K-ras and p53 gene mutation in the stool and pure pancreatic juice for diagnosis of early pancreatic cancer]. Zhonghua Yi Xue Za Zhi. 2001 Sep 10;81(17):1050-3.
  69. Wu X, Lu XH, Xu T, Qian JM, Zhao P, Guo XZ, Yang XO, Jiang WJ. Evaluation of the diagnostic value of serum tumor markers, and fecal k-ras and p53 gene mutations for pancreatic cancer. Chin J Dig Dis. 2006;7(3):170-4.
  70. Bland J. Farnesyl transferase inhibitors.2001;October 17-21, 2001.
  71. Asamoto M, Ota T, Toriyama-Baba H, et al. Mammary carcinomas induced in human c-Ha-ras proto-oncogene transgenic rats are estrogen-independent, but responsive to d-limonene treatment. Jpn J Cancer Res. 2002 Jan;93(1):32-5.
  72. Kim Ms, Kang HJ, Moon A. Inhibition of invasion and induction of apoptosis by curcumin in H-ras-transformed MCF10A human breast epithelial cells. Arch Pharm Res. 2001 Aug;24(4):349-54.
  73. Chen X, Hasuma T, Yano Y, et al. Inhibition of farnesyl protein transferase by monoterpene, curcumin derivatives and gallotannin. Anticancer Res. 1997 Jul-Aug;17(4A):2555-64.
  74. Yano T, Uchida M, Yuasa M, et al. The inhibitory effect of vitamin E on K-ras mutation at an early stage of lung carcinogenesis in mice. Eur J Pharmacol. 1997 Mar;323(1):99-102.
  75. Prasad KN, Cohrs RJ, Sharma OK. Decreased expressions of c-myc and H-ras oncogenes in vitamin E succinate induced morphologically differentiated murine B-16 melanoma cells in culture. Biochem Cell Biol. 1990 Nov;68(11):1250-5.
  76. Singh A, Purohit A, Hejaz HA, et al. Inhibition of deoxyglucose uptake in MCF-7 breast cancer cells by 2-methoxyestrone and 2-methoxyestrone-3-O-sulfamate. Mol Cell Endocrinol. 2000 Feb 25;160(1-2):61-6.
  77. Chung JY, Huang C, Meng X, et al. Inhibition of activator protein 1 activity and cell growth by purified green tea and black tea polyphenols in H-ras-transformed cells: structure-activity relationship and mechanisms involved. Cancer Res. 1999 Sep 15;59(18):4610-7.
  78. Collett ED, Davidson LA, Fan YY, et al. n-6 and n-3 polyunsaturated fatty acids differentially modulate oncogenic Ras activation in colonocytes. Am J Physiol Cell Physiol. 2001 May;280(5):C1066-C1075.
  79. Wang IK, Lin-Shiau SY, Lin JK. Suppression of invasion and MMP-9 expression in NIH 3T3 and v-H-Ras 3T3 fibroblasts by lovastatin through inhibition of ras isoprenylation. Oncology. 2000 Sep;59(3):245-54.
  80. Hohl RJ, Lewis K. Differential effects of monoterpenes and lovastatin on RAS processing. J Biol Chem. 1995 Jul 21;270(29):17508-12.
  81. Kawata S, Yamasaki E, Nagase T, Inui Y, Ito N, Matsuda Y, Inada M, Tamura S, Noda S, Imai Y, Matsuzawa Y. Effect of pravastatin on survival in patients with advanced hepatocellular carcinoma. A randomized controlled trial. Br J Cancer. 2001 Apr 6;84(7):886-91.
  82. Graf H, Jüngst C, Straub G, Dogan S, Hoffmann RT, Jakobs T, Reiser M, Waggershauser T, Helmberger T, Walter A, Walli A, Seidel D, Goke B, Jüngst D. Chemoembolization combined with pravastatin improves survival in patients with hepatocellular carcinoma. Digestion. 2008;78(1):34-8.
  83. Samuels AJBHRSSB. Increased alpha 1 antitrypsin, decreased systemic fibrinolysis, hypercoaguable state and increased fibrin split products in patients with lung cancer. Proc Am Soc Clin Oncol. 1975;(16):238.
  84. Cafagna D, Ponte E. [Pulmonary embolism of paraneoplastic origin]. Minerva Med. 1997 Dec;88(12):523-30.
  85. Ahmad S et al. Therapeutic roles of heparin anticoagulants in cancer and related disorders. Med Chem. 2011 Sep 1;7(5):504-17.
  86. Cosgrove RH, Zacharski LR, Racine E, et al. Improved cancer mortality with low-molecular-weight heparin treatment: a review of the evidence. Semin Thromb Hemost. 2002 Feb;28(1):79-87.
  87. Mousa SA. Anticoagulants in thrombosis and cancer: the missing link. Expert Rev Anticancer Ther. 2002 Apr;2(2):227-33.
  88. von Tempelhoff GF, Harenberg J, Niemann F, et al. Effect of low molecular weight heparin (Certoparin) versus unfractioned heparin on cancer survival following breast and pelvic cancer surgery: A prospective randomized double-blind trial. Int J Oncol. 2000;16(4):815-24.
  89. Wellness Directory of Minnesota. 2002
  90. Lebeau B, Chastang C, Brechot JM, et al. Subcutaneous heparin treatment increases survival in small cell lung cancer. "Petites Cellules" Group. Cancer. 1994 Jul 1;74(1):38-45.
  91. Van Doormaal FF, Büller HR, Middeldrop S. Development of anticoagulant therapy. Crit Rev Oncol Hematol. 2008 May;66(2):145-54.
  92. Spivak JL. Cancer-related anemia: its causes and characteristics. Semin Oncol. 1994 Apr;21(2 Suppl 3):3-8.
  93. Caro JJ, Salas M, Ward A, et al. Anemia as an independent prognostic factor for survival in patients with cancer: a systemic, quantitative review. Cancer. 2001 Jun 15;91(12):2214-21.
  94. Cazzola M. Mechanisms of anaemia in patients with malignancy: implications for the clinical use of recombinant human erythropoietin. Med Oncol. 2000 Nov;17 Suppl 1:S11-S16.
  95. Reis ST et al. Tgf-β1 expression as a biomarker of poor prognosis in prostate cancer. Clinics (Sao Paulo). 2011;66(7):1143-7.
  96. Zacharski LR, Henderson WG, Rickles FR, et al. Effect of warfarin anticoagulation on survival in carcinoma of the lung, colon, head and neck, and prostate. Final report of VA Cooperative Study #75. Cancer. 1984 May 15;53(10):2046-52.
  97. Zacharski LR, Memoli VA, Rousseau SM, et al. Occurrence of blood coagulation factors in situ in small cell carcinoma of the lung. Cancer. 1987 Dec 1;60(11):2675-81.
  98. Chahinian AP, Propert KJ, Ware JH, et al. A randomized trial of anticoagulation with warfarin and of alternating chemotherapy in extensive small-cell lung cancer by the Cancer and Leukemia Group B. J Clin Oncol. 1989 Aug;7(8):993-1002.
  99. Saad F, Abrahamsson PA, Miller K. Preserving bone health in patients with hormone-sensitive prostate cancer: the role of bisphosphonates. BJU Int. 2009 Dec;104(11):1573-9.
  100. Mystakidou K, Katsouda E, Parpa E, Kelekis A, Galanos A, Vlahos L. Randomized, open label, prospective study on the effect of zoledronic acid on the prevention of bone metastases in patients with recurrent solid tumors that did not present with bone metastases at baseline. Med Oncol. 2005;22(2):195-201.
  101. Gnant BT, Amenedo C, Freeman K, et al. Outcomes of placing dental implants in patients taking oral bisphosphonates: a review of 115 cases. J Oral Masxillofac Surg. 2008;66:223-230.
  102. Coleman RE, Thorpe HC, Cameron D, et al. Adjuvant Treatment with Zoledronic Acid in Stage II/III Breast Cancer. The AZURE Trial (BIG 01/04). http://www.abstracts2view.com/sabcs10/view.php?nu=SABCS10L_226&terms. 12/10/2010.
  103. Morgan G, Davies F, Gregory W, et al. Evaluating the effects of zoledronic acid (ZOL) on overall survival (OS) in patients (Pts) with multiple myeloma (MM): results of the Medical Research Council (MRC) Myeloma IX study. J Clin Oncol. 2010;28(suppl):578s. (abstr 8021).
  104. Hwang SH, Lee SA, Ahn SG, Lee HM, Jeong J, Lee HD. Effects of zoledronic acid on bone mineral density during aromatase inhibitor treatment of Korean postmenopausal breast cancer patients. Breast Cancer Res Treat. 2011 Aug 23.
  105. Stopeck AT, Lipton A, Body JJ, et al. Denosumab compared with zoledronic acid for the treatment of bone metastases in patients with advanced breast cancer: a randomized, double-blind study. J Clin Oncol. 2010;28(35):5132-9.
  106. Weitzman R, Sauter N, Eriksen EF, et al. Critical review: updated recommendations for the prevention, diagnosis, and treatment of osteonecrosis of the jaw in cancer patients—May 2006. Crit Rev Oncol Hematol. 2007 May;62(2):148-52.
  107. Heckbert SR, Li G, Cummings SR, Smith NL, Psaty BM. Use of alendronate and risk of incident atrial fibrillation in women. Arch Intern Med. 2008 Apr 28;168(8):826-31.
  108. Miller PD. Curr Osteoporos Rep. 2009 Mar;7(1):18-22. Denosumab: anti-RANKL antibody.
  109. Fizazi K, Carducci M, Smith M, Damião R, Brown J, Karsh L, Milecki P, Shore N, Rader M, Wang H, Jiang Q, Tadros S, Dansey R, Goessl C. Lancet. 2011 Mar 5;377(9768):813-22. Denosumab versus zoledronic acid for treatment of bone metastases in men with castration-resistant prostate cancer: a randomised, double-blind study.
  110. Shammas NW, Moss AJ, Sullebarger JT, et al. Acquired coronary angiogenesis after myocardial infarction. Cardiology. 1993;83(3):212-6.
  111. Suh DY. Understanding angiogenesis and its clinical applications. Ann Clin Lab Sci. 2000 Jul;30(3):227-38.
  112. Cooke R. Dr. Folkman's War: Angiogenesis and the Struggle to Defeat Cancer. Nature Medicine. 2001 May;7(5):525-6.
  113. Folkman J. Tumor angiogenesis: therapeutic implications. N Engl J Med. 1971 Nov 18;285(21):1182-6
  114. Cao Y, Langer R. A review of Judah Folkman's remarkable achievements in biomedicine. Proc Natl Acad Sci U S A. 2008 Sep 9;105(36):13203-5.
  115. Folkman J. The role of angiogenesis in tumor growth. Semin Cancer Biol. 1992b Apr;3(2):65-71.
  116. Folkman J, Shing Y. Control of angiogenesis by heparin and other sulfated polysaccharides. Adv Exp Med Biol. 1992a;313:355-64.
  117. NIH/NCI. Angiogenesis Inhibitors in Cancer Research. NIH/NCI Angiogenesis Inhibitors in Cancer Research 1998 Jul 7 Bethesda, MD: National Cancer Institute (Http://Www Slip Net/~Mcdavis/Database/Angio163 Htm).1998;Jul 7
  118. Ghosh J. Rapid induction of apoptosis in prostate cancer cells by selenium: reversal by metabolites of arachidonate 5-lipoxygenase. Biochem Biophys Res Commun. 2004 Mar 12;315(3):24-35.
  119. Moretti RM, Montagani Marelli M, Sala A, et al. Activation of the orphan nuclear receptor RORalpha counteracts the proliferative effect of fatty acids on prostate cancer cells: crucial role of 5-lipoxygenase. Int J Cancer. 2004 Oct;121(1):87-93.
  120. Hassan S, Carraway RE. Involvement of arachidonic acid metabolism and EGF receptor in neurotensin-induced prostate cancer PC3 cell growth. Regul Pept. 2006 Jan;133(1-3):105-14.
  121. Matsuyama M, Yoshimura R, Mitsuhashi M, et al. Expression of lipoxygenase in human prostate cancer and growth reduction by its inhibitors. Int J Oncol. 2004a Apr;24(4):821-7.
  122. Kelavkar UP, Glasgow W, Olson SJ, Foster BA, Shappell SB. Overexpression of 12/15-lipoxygenase, an ortholog of human 15-lipoxygenase-1, in the prostate tumors of TRAMP mice. Neoplasia. 2004 Nov;6(6):821-30.
  123. Gupta S, Srivastava M, Ahmad N, et al. Lipoxygenase-5 is overexpressed in prostate adenocarcinoma. Cancer. 2001 Feb 15;91(4):737-43
  124. Kelavkar UP, Nixon JB, Cohen C, et al. Overexpression of 15-lipoxygenase-1 in PC-3 human prostate cancer cells increases tumorigenesis. Carcinogenesis. 2001 Nov;22(11):1765-73.
  125. Ghosh J, Myers CE. Arachidonic acid stimulates prostate cancer cell growth: critical role of 5-lipoxygenase. Biochem Biophys Res Commun. 1997 Jun;235(2):418-23.
  126. Gao X, Grignon DJ, Chbihi T, et al. Elevated 12-lipoxygenase mRNA expression correlates with advanced stage and poor differentiation of human prostate cancer. Urology. 1995;46(2):227-37.
  127. Sundaram S, Ghosh J. Expression of 5-oxoETE receptor in prostate cancer cells: critical role in survival. Biochem Biophys Res Commun. 2006 Jan 6;339(1):93-8.
  128. Myers CE, Ghosh J. Lipoxygenase inhibition in prostate cancer. Eur Urol. 1999;35(5-6):395-8.
  129. Nakao-Hayashi J, Ito H, Kanayasu T, Morita I, Murota S. Stimulatory effects of insulin and insulin-like growth factor I on migration and tube formation by vascular endothelial cells. Atherosclerosis. 1992 Feb;92(2-3):141-9.
  130. Cohen P, Peehl DM, Lamson G, Rosenfeld RG. Insulin-like growth factors (IGFs), IGF receptors, and IGF-binding proteins in primary cultures of prostate epithelial cells. J Clin Endocrinol Metab. 1991 Aug;73(2):401-7.
  131. Ghosh J. Inhibition of arachidonate 5-lipoxygenase triggers prostate cancer cell death through rapid activation of c-Jun N-terminal kinase. Biochem Biophys Res Commun. 2003 Jul 25;307(2):342-9.
  132. Jiang WG, Douglas-Jones AG, Mansel RE. Aberrant expression of 5-lipoxygenase-activating protein (5-LOXAP) has prognostic and survival significance in patients with breast cancer. Prostaglandins Leukot Essent Fatty Acids. 2006 Feb;74(2):125-34.
  133. Yoshimura R, Matsuyama M, Mitsuhashi M, et al. Relationship between lipoxygenase and human testicular cancer. Int J Mol Med. 2004 Mar;13(3):389-93.
  134. Zhang L, Zhang WP, Hu H, et al. Expression patterns of 5-lipoxygenase in human brain with traumatic injury and astrocytoma. Neuropathology. 2006 Apr;26(2):99-106.
  135. Soumaoro LT, Iida S, Uetake H, et al. Expression of 5-Lipoxygenase in human colorectal cancer. World J Gastroenterol. 2006 Oct 21;12(39):6355-60.
  136. Hayashi T, Nishiyama K, Shirahama T. Inhibition of 5-lipoxygenase pathway suppresses the growth of bladder cancer cells. Int J Urol. 2006 Aug;13(8):1086-91.
  137. Hoque A, Lippman SM, Wu TT, et al. Increased 5-lipoxygenase expression and induction of apoptosis by its inhibitors in esophageal cancer: a potential target for prevention. Carcinogenesis. 2005 Apr;26(4):785-91.
  138. Hennig R, Ding XZ, Tong WG, et al. 5-Lipoxygenase and leukotriene B(4) receptor are expressed in human pancreatic cancers but not in pancreatic ducts in normal tissue. Am J Pathol. 2002 Aug;161(2):421-8.
  139. Ding XZ, Iversen P, Cluck MW, Knezetic JA, Adrian TE. Lipoxygenase inhibitors abolish proliferation of human pancreatic cancer cells. Biochem Biophys Res Commun. 1999 Jul 22;261(1):218-23.
  140. Matsuyama M, Yoshimura R, Mitsuhashi M, et al. 5-Lipoxygenase inhibitors attenuate growth of human renal cell carcinoma and induce apoptosis through arachidonic acid pathway. Oncol Rep. 2005 Jul;14(1):73-9.
  141. Zhi H, Zhang J, Hu G, et al. The deregulation of arachidonic acid metabolism-related genes in human esophageal squamous cell carcinoma. Int J Cancer. 2003 Sep 1;106(3):327-33.
  142. Penglis PS, Cleland LG, Demasi M, Caughey GE, James MJ. Differential regulation of prostaglandin E2 and thromboxane A2 production in human monocytes: implications for the use of cyclooxygenase inhibitors. J Immunol. 2000 Aug 1;165(3):1605-11.
  143. Rubinsztajn R, Wronska J, Chazan R. [Urinary leukotriene E4 concentration in patients with bronchial asthma and intolerance of non-steroids anti-inflammatory drugs before and after oral aspirin challenge]. Pol Arch Med Wewn. 2003 Aug;110(2):849-854
  144. Subbarao K, Jala VR, Mathis S, et al. Role of leukotriene B4 receptors in the development of atherosclerosis: potential mechanisms. Arterioscler Thromb Vasc Biol. 2004 Feb;24(2):369-75.
  145. Hu Y, Sun H, O’Flaherty JT, et al. 15-Lipoxygenase-1-mediated metabolism of docosahexaenoic acid is required for syndecan-1 signaling and apoptosis in prostate cancer cells. Carcinogenesis. 2013;34(1):176-82.
  146. Ye YN, Liu ES, Shin VY, Wu WK, Cho CH. Contributory role of 5-lipoxygenase and its association with angiogenesis in the promotion of inflammation-associated colonic tumorigenesis by cigarette smoking. Toxicology. 2004 Oct 15;203(1-3):179-88.
  147. Rajashekhar G, Willuweit A, Patterson CE, et al. Continuous endothelial cell activation increases angiogenesis: evidence for the direct role of endothelium linking angiogenesis and inflammation. J Vasc Res. 2006;43(2):193-204.
  148. Hazai E, Bikadi Z, Zsila F, Lockwood SF. Molecular modeling of the non-covalent binding of the dietary tomato carotenoids lycopene and lycophyll, and selected oxidative metabolites with 5-lipoxygenase. Bioorg Med Chem. 2006 Oct 15;14(20):6859-67.
  149. Ghosh J, Myers CE. Inhibition of arachidonate 5-lipoxygenase triggers massive apoptosis in human prostate cancer cells. Proc Natl Acad Sci USA. 1998 Oct 27;95(22):13182-7.
  150. Yang P, Collin P, Madden T, et al. Inhibition of proliferation of PC3 cells by the branched-chain fatty acid, 12-methyltetradecanoic acid, is associated with inhibition of 5-lipoxygenase. Prostate. 2003 Jun 1;55(4):281-91.
  151. Anderson KM, Seed T, Vos M, et al. 5-Lipoxygenase inhibitors reduce PC-3 cell proliferation and initiate nonnecrotic cell death. Prostate. 1998;37(3):161-73.
  152. Steinhilber D et al. 5-lipoxygenase: underappreciated role of a pro-inflammatory enzyme in tumorigenesis. Front Pharmacol. 2010;1:143. Epub 2010 Dec 24.
  153. Taccone-Gallucci M, Manca-di-Villahermosa S, Battistini L, et al. N-3 PUFAs reduce oxidative stress in ESRD patients on maintenance HD by inhibiting 5-lipoxygenase activity. Kidney Int. 2006 Apr;69(8):1450-4.
  154. Calder PC. N-3 polyunsaturated fatty acids and inflammation: from molecular biology to the clinic. Lipids. 2003 Apr;38(4):343-52.
  155. Jian L, Du CJ, Lee AH, Binns CW. Do dietary lycopene and other carotenoids protect against prostate cancer? Int J Cancer. 2005 Mar 1;113(6):1010-4.
  156. Campbell JK, Canene-Adams K, Lindshield BL, et al. Tomato phytochemicals and prostate cancer risk. J Nutr. 2004 Dec;134(12 Suppl):3486S-92S.
  157. Wertz K, Siler U, Goralczyk R. Lycopene: modes of action to promote prostate health. Arch Biochem Biophys. 2004 Oct 1;430(1):127-34.
  158. Binns CW, LJ LJ, Lee AH. The relationship between dietary carotenoids and prostate cancer risk in Southeast Chinese men. Asia Pac J Clin Nutr. 2004;13(Suppl):S117.
  159. Kim L, Rao AV, Rao LG. Effect of lycopene on prostate LNCaP cancer cells in culture. J Med Food. 2002;5(4):181-7.
  160. Giovannucci E. A review of epidemiologic studies of tomatoes, lycopene, and prostate cancer. Exp Biol Med (Maywood.). 2002 Nov;227(10):852-9.
  161. Vogt TM, Mayne ST, Graubard BI, et al. Serum lycopene, other serum carotenoids, and risk of prostate cancer in US blacks and whites. Am J Epidemiol. 2002 Jun 1;155(11):1023-32.
  162. Bosetti C, Tzonou A, Lagiou P, et al. Fraction of prostate cancer incidence attributed to diet in Athens, Greece. Eur J Cancer Prev. 2000 Apr;9(2):119-23.
  163. Blumenfeld AJ, Fleshner N, Casselman B, Trachtenberg J. Nutritional aspects of prostate cancer: a review. Can J Urol. 2000 Feb;7(1):927-35.
  164. Agarwal S, Rao AV. Tomato lycopene and its role in human health and chronic diseases. CMAJ. 2000 Sep 19;163(6):739-44.
  165. Gann PH, Ma J, Giovannucci E, et al. Lower prostate cancer risk in men with elevated plasma lycopene levels: results of a prospective analysis. Cancer Res. 1999 Mar 15;59(6):1225-30.
  166. Giovannucci E, Ascherio A, Rimm EB, et al. Intake of carotenoids and retinol in relation to risk of prostate cancer. J Natl Cancer Inst. 1995 Dec;87(23):1767-76.
  167. Hill B, Kyprianou N. Effect of permixon on human prostate cell growth: lack of apoptotic action. Prostate. 2004 Sep 15;61(1):73-80.
  168. Paubert-Braquet M, Mencia Huerta JM, Cousse H, et al. Effect of the lipidic lipidosterolic extract of Serenoa repens (Permixon) on the ionophore A23187-simulated production of leukotriene B4 (LTB4) from human polymorphonuclear neutrophils. Prostaglandins Leukot Essent Fatty Acids. 1997;57(3):299-304.
  169. Safayhi H, Rall B, Sailer ER, Ammon HP. Inhibition by boswellic acids of human leukocyte elastase. J Pharmacol Exp Ther. 1997 Apr;281(1):460-3.
  170. Safayhi H, Sailer ER, Ammon HP. Mechanism of 5-lipoxygenase inhibition by acetyl-11-keto-beta-boswellic acid. Mol Pharmacol. 1995 Jun;47(6):1212-6.
  171. Kimmatkar N, Thawani V, Hingorani L, Khiyani R. Efficacy and tolerability of Boswellia serrata extract in treatment of osteoarthritis of knee—a randomized double blind placebo controlled trial. Phytomedicine. 2003 Jan;10(1):3-7.
  172. Ammon HP. Boswellic acids (components of frankincense) as the active principle in treatment of chronic inflammatory disease. Wien Med Wochenschr. 2002;152(15-16):373-8.
  173. Wallace JM. Nutritional and botanical modulation of the inflammatory cascade--eicosanoids, cyclooxygenases, and lipoxygenases—as an adjunct in cancer therapy. Integr Cancer Ther. 2002 Mar;1(1):7-37.
  174. Gerhardt H, Seifert F, Buvari P, Vogelsang H, Repges R. Therapy of active Crohn disease with Boswellia serrata extract H 15. Z Gastroenterol. 2001 Jan;39(1):11-7.
  175. Gupta I, Gupta V, Parihar A, et al. Effects of Boswellia serrata gum resin in patients with bronchial asthma: results of a double-blind, placebo-controlled, 6-week clinical study. Eur J Med Res. 1998 Nov 17;3(11):511-4.
  176. Kulkarni RR, Patki PS, Jog VP, Gandage SG, Patwardhan B. Treatment of osteoarthritis with a herbomineral formulation: a double-blind, placebo-controlled, cross-over study. J Ethnopharmacol. 1991 May;33(1-2):91-5.
  177. Park YS, Lee JH, Harwalkar JA, et al. Acetyl-11-keto-beta-boswellic acid (AKBA) is cytotoxic for meningioma cells and inhibits phosphorylation of the extracellular-signal regulated kinase 1 and 2. Adv Exp Med Biol. 2002;507:387-93.
  178. Liu JJ, Nilsson A, Oredsson S, Badmaev V, Duan RD. Keto- and acetyl-keto-boswellic acids inhibit proliferation and induce apoptosis in Hep G2 cells via a caspase-8 dependent pathway. Int J Mol Med. 2002 Oct;10(4):501-5.
  179. Syrovets T, Büchele B, Gedig E, et al. Acetyl-boswellic acids are novel catalytic inhibitors of human topoisomerases I and IIalpha. Mol Pharmacol. 2000 Jul;58(1):71-81.
  180. Sailer ER, Subramanian LR, Rall B, et al. Acetyl-11-keto-beta-boswellic acid (AKBA): structure requirements for binding and 5-lipoxygenase inhibitory activity. Br J Pharmacol. 1996 Feb;117(4):615-8.
  181. Sengupta K, Kolla JN, Krishnaraju AV, et al. Cellular and molecular mechanisms of anti-inflammatory effect of Aflapin: a novel Boswellia serrata extract. Mol Cell Biochem. 2011 Aug;354(1-2):189-97.
  182. Krishnaraju AV, Sundararaju D, Vamsikrishna U, et al. Safety and toxicological evaluation of Aflapin: a novel Boswellia-derived anti-inflammatory product. Toxicol Mech Methods. 2010;20(9):556-63.
  183. Bird NC et al. Biology of colorectal liver metastases: A review. J Surg Oncol. 2006 Jul 1;94(1):68-80.
  184. Dowdall JF et al. Soluble interleukin 6 receptor (sIL-6R) mediates colonic tumor cell adherence to the vascular endothelium: a mechanism for metastatic initiation? J Surg Res. 2002 Sep;107(1):1-6.
  185. Raz A et al. Endogenous galactoside-binding lectins: a new class of functional tumor cell surface molecules related to metastasis. Cancer Metastasis Rev. 1987;6(3):433-52.
  186. Yu LG et al. Galectin-3 interaction with Thomsen-Friedenreich disaccharide on cancer-associated MUC1 causes increased cancer cell endothelial adhesion. J Biol Chem. 2007 Jan 5;282(1):773-81. Epub 2006 Nov 7.
  187. ten Kate M et al. Influence of proinflammatory cytokines on the adhesion of human colon carcinoma cells to lung microvascular endothelium. Int J Cancer. 2004 Dec 20;112(6):943-50.
  188. Nangia-Makker P et al. Inhibition of human cancer cell growth and metastasis in nude mice by oral intake of modified citrus pectin. J Natl Cancer Inst. 2002 Dec 18;94(24):1854-62.
  189. Guess BW et al. Modified citrus pectin (MCP) increases the prostate-specific antigen doubling time in men with prostate cancer: a phase II pilot study. Prostate Cancer Prostatic Dis. 2003;6(4):301-4.
  190. Jackson CL et al. Pectin induces apoptosis in human prostate cancer cells: correlation of apoptotic function with pectin structure. Glycobiology. 2007;17(8):805-819.
  191. Eichbaum C et al. Breast Cancer Cell-derived Cytokines, Macrophages and Cell Adhesion: Implications for Metastasis. Anticancer Res. 2011 Oct;31(10):3219-27.
  192. Matsumoto S et al. Cimetidine increases survival of colorectal cancer patients with high levels of sialyl Lewis-X and sialyl Lewis-A epitope expression on tumour cells. Br J Cancer. 2002 Jan 21;86(2):161-7.
  193. Adams WJ et al. Short-course cimetidine and survival with colorectal cancer. Lancet. 1994 Dec 24-31;344(8939-8940):1768-9.
  194. Shakhar G et al. Potential prophylactic measures against postoperative immunosuppression: could they reduce recurrence rates in oncological patients? Ann Surg Oncol. 2003 Oct;10(8):972-92.
  195. Koda K et al. Preoperative natural killer cell activity: correlation with distant metastases in curatively research colorectal carcinomas. Int Surg. 1997 Apr-Jun;82(2):190-3.
  196. Fisher M et al. Anticancer effects and mechanisms of polysaccharide-K (PSK): implications of cancer immunotherapy. Anticancer Res. 2002 May-Jun;22(3):1737-54.
  197. Garcia-Lora A et al. Protein-bound polysaccharide K and interleukin-2 regulate different nuclear transcription factors in the NKL human natural killer cell line. Cancer Immunol Immunother. 2001 Jun;50(4):191-8.
  198. Hayakawa K et al. Effect of Krestin as adjuvant treatment following radical radiotherapy in non-small cell lung cancer patients. Cancer Detect Prev. 1997;21(1):71-7.
  199. Sakai T et al. Immunochemotherapy with PSK and fluoropyrimidines improves long-term prognosis for curatively resected colorectal cancer. Cancer Biother Radiopharm. 2008 Aug;23(4):461-7.
  200. Ohwada S et al. Adjuvant immunochemotherapy with oral Tegafur/Uracil plus PSK in patients with stage II or III colorectal cancer: a randomised controlled study. Br J Cancer. 2004 Mar 8;90(5):1003-10.
  201. Okazaki A et al. [The effects of PS-K on long-term survival of uterine cervical cancer patients treated with radiation]. Gan No Rinsho. 1986 Feb;32(2):181-5.
  202. Nakazato H et al. Efficacy of immunochemotherapy as adjuvant treatment after curative resection of gastric cancer. Study Group of Immunochemotherapy with PSK for Gastric Cancer. Lancet. 1994 May 7;343(8906):1122-6.
  203. Toi M et al. Randomized adjuvant trial to evaluate the addition of tamoxifen and PSK to chemotherapy in patients with primary breast cancer. 5-Year results from the Nishi-Nippon Group of the Adjuvant Chemoendocrine Therapy for Breast Cancer Organization. Cancer. 1992 Nov 15;70(10):2475-83.
  204. White MC, Holman DM, Boehm JE, Peipins LA, Grossman M, Henley SJ. Age and cancer risk: a potentially modifiable relationship. American journal of preventive medicine. 2014;46(3 Suppl 1):S7-S15.
  205. Harding C, Pompei F, Wilson R. Peak and decline in cancer incidence, mortality, and prevalence at old ages. Cancer. 2012;118(5):1371-1386.
  206. Gomes AP, Ilter D, Low V, et al. Age-induced accumulation of methylmalonic acid promotes tumour progression. Nature. 2020.
  207. Vashi P, Edwin P, Popiel B, Lammersfeld C, Gupta D. Methylmalonic Acid and Homocysteine as Indicators of Vitamin B-12 Deficiency in Cancer. PloS one. 2016;11(1):e0147843-e0147843.
  208. Ye X, Weinberg RA. Epithelial-Mesenchymal Plasticity: A Central Regulator of Cancer Progression. Trends in cell biology. 2015;25(11):675-686.
  209. Wang L, Zhang J, Yang X, et al. SOX4 is associated with poor prognosis in prostate cancer and promotes epithelial–mesenchymal transition in vitro. Prostate cancer and prostatic diseases. 2013;16(4):301-307.
  210. Zhang J, Liang Q, Lei Y, et al. SOX4 induces epithelial-mesenchymal transition and contributes to breast cancer progression. Cancer Res. 2012;72(17):4597-4608.