Next Article in Journal
Quality of Sleep and Work Productivity among White-Collar Workers during the COVID-19 Pandemic
Next Article in Special Issue
Anesthetics and Long Term Cancer Outcomes: May Epigenetics Be the Key for Pancreatic Cancer?
Previous Article in Journal
Novel Animal Model of Spontaneous Cerebral Petechial Hemorrhage Using Focused Ultrasound in Rats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Impact of Local Anesthetics on Cancer Behavior and Outcome during the Perioperative Period: A Review

1
Balgrist Campus, University of Zurich, Lengghalde 5, 8008 Zurich, Switzerland
2
Institute of Anesthesiology, Triemli City Hospital, 8063 Zurich, Switzerland
*
Author to whom correspondence should be addressed.
Medicina 2022, 58(7), 882; https://doi.org/10.3390/medicina58070882
Submission received: 21 April 2022 / Revised: 16 June 2022 / Accepted: 21 June 2022 / Published: 30 June 2022
(This article belongs to the Special Issue Impact of Anesthetics on Cancer Behavior and Outcome)

Abstract

:
There is a growing interest regarding the impact of the perioperative period and the application of anesthetic drugs on the recurrence of cancer metastases. Among them, the use of amide-type local anesthetics seems promising since in vitro studies and animal models have shown their potential to inhibit the Intercellular Adhesion Molecule 1 (ICAM-1) expression and Src activity, which are clearly implicated in the process of inflammation and cancer metastases. This review emphasizes the potential of amide-type local anesthetics in this context.

1. Introduction

Lidocaine is an old drug, which has been acknowledged to be exclusively indicated for the management of pain and cardiac arrhythmias through its ability to block the sodium channel; however, new research has demonstrated that this drug was able to block or interfere with other receptors and systems leading to new clinical development. One of them could be the potential to reduce cancer metastases.
Cancer is still one of the major causes of morbidity and mortality worldwide [1]. Surgical removal is still used in more than 60% of cases [2]; however, the perioperative period is increasingly recognized as a time point with the release of circulating tumor cells and a great influence on the immune system [3].
Recently more and more evidence has shown that inflammation and metastases are closely connected [4]. The presence of leucocytes within and around tumors was already observed by Virchow in the 19th century leading to the evidence that an inflammatory microenvironment was an essential component of all tumors [5].
Metastases are the most crucial aspect of tumorigenesis because approximately 2/3 of cancer mortality is caused by metastases [6,7]. This process requires a close link between cancer cells and immune and inflammatory cells. Among the different steps necessary for the occurrence of metastases, one involves direct inflammation, which, through the release of mediators, increases vascular permeability. In this context, it was shown that local anesthetics (LAs) possess anti-inflammatory properties [8,9] by acting, among others, on the G protein-coupled receptors [10].
Surgery is a known model to favor the development of cancer: There is a great inflammatory response, and the immune defenses are greatly impaired. Lin et al. [11] have extensively described surgical-induced systemic inflammation. Primarily, this process should eradicate microorganisms and enhance healing; however, this response can be exaggerated, which will impair the immune system, leading to multiple organ failure and patient death [12]. This occurrence may have the potential to favor metastases.

2. Circulating Tumor Cells (CTCs)

The role of CTCs has been hypothesized to play a major role in the formation of metastases [13]. Recently CTCs have been studied as a biomarker for cancer detection and prognosis. It was shown that the number of CTCs was associated with patient survival in different types of cancer, including breast, colon, and many others [14]. Asworth [15] was the first to describe the occurrence of CTCs, but the clinical interest in this finding was postponed until recently due to technical issues for isolating the CTCs since they are ultra-rare events. Moreover, they are derived from different types of tissues, knowing that each one distinguishes from the other considering the size, shape, and immune phenotyping profile [16].
The CTCs are considered to be shed from different locations within tumors; however, this occurrence is accelerated during surgery. Sergeant et al. [17] have found that after a successful pancreatectomy for an adenocarcinoma, the outcome was poor due to new distant metastases related to a large increase in CTCs [17,18]; however, the diameter of CTCs is large, greater than the bores of capillaries, in which we would expect that these cells would be blocked [19]. This concept suggests that other mechanisms should be activated to allow the CTCs to transfer directly into the tissues from the vascular endothelia. To allow this occurrence, some biomolecular changes should occur within the CTCs, making them more aggressive. This can occur secondary to the severe inflammatory reaction induced by the surgery.

3. Anti-Inflammatory Effects of Local Anesthetics Potentially Affecting Metastases Immune Modulation/Natural Killer (NK) Cell Activity

Each surgery or trauma will create new body homeostasis, called the surgical-induced stress response [20,21]. This phenomenon is characterized by the production of catecholamine and the activation of the corticotropic axis leading to a redistribution of circulating leucocytes, resulting in lymphopenia [22]; therefore, during this period, a generalized pro-inflammatory state occurs. It is associated with the release of different cytokines, including, among others, IL-6 and tumor necrosis factor, [23] which can potentially favor tumor progression. This inflammatory surrounding also negatively influences the immune system [24].
In humans, NK cells have been the most studied. NK cells were first described by Kiessling et al., 1975 [25], and it became rapidly evident that these cells play an important role in establishing anti-tumor immunity [26]. In humans, Iannone et al. found that NK cell cytotoxicity was significantly inhibited in pancreatic cancer patients after pancreatectomy. It was noted that only after 30 days the NK cell activity was restored [27]. A reduction in NK cell cytotoxicity was also observed after surgery for pulmonary, hepatocellular, and breast cancer [28]. In animals in whom tumor cells were administered IV, tumor cell retention was significantly increased after laparotomy compared to those having sham surgery [29].
The interaction of different anesthetics on the NK cell activity has been tested. Among the volatiles, halothane has been shown in rats to reduce the cytotoxicity of NK cells [30]. In the same investigation, retention of tumor cells was increased with halothane, but not after administration of propofol [30]. Among analgesics, morphine has been shown in animals and humans to have a negative effect on NK cell activity (Table 1).
A dual effect of lidocaine has been observed. At a very high concentration in vitro, lidocaine has been shown to have a negative effect [31], whereas, at clinical concentration, lidocaine demonstrated an increase in the cytotoxic potential of NK cells [32]. The presumed beneficial action was due to the release of granzyme b and perforin [32]. Moreover, the serum from patients receiving lidocaine during cancer surgery was greatly competent in killing cancer cells [33,34]. Another beneficial effect of lidocaine is the action of the Th1/Th2 ratio. CD4 and T-cells, after antigen stimulation, will differentiate into Th1 or Th2 cells. Interferon-γ (IFN-γ) is secreted by Th1 cells and is responsible for cellular immunity. On the other hand, Th2 cells are associated with humoral immunity through the release of IL-4. The proportion of these mediators determines the Th1/Th2 ratio. Knowing that cellular immunity is a key factor in controlling the immune response toward the tumor, any reduction in this ratio will favor tumor progression. In patients undergoing hysterectomy or hepatectomy, the addition of lidocaine inclined the ratio Th1/Th2 towards a Th1 profile, including the production of IFN-γ [35,36]. Numerous studies have demonstrated the negative effect of certain cytokines, especially IL-6, which has been associated with tumor progression [37,38]. These mediators increase dramatically after all surgeries and/or trauma. Investigations have shown that the perioperative application of lidocaine significantly decreased the release of pro-inflammatory cytokines [39]. This suggests an intensive cross-talking between inflammatory and tumor cells, making a common pathway between these two systems likely and emphasizing the interest in giving lidocaine during this period to stimulate the immune system and protect against tumor cell dissemination during the perioperative and first postoperative days.

4. Endothelial Barrier, Leucocyte Activation, Leukocyte/Tumor Cell Adhesion, and Transmigration

The prerequisites for metastatic dissemination of solid tumors are migration, invasion, and adhesion of cancer cells. The movement mechanism of the cells is migration, while direct extension and penetration of cancer cells into tissues is invasion. In between, the process of adhesion also plays a critical aspect in the formation of metastases. It seems that this process is involved in the great majority of metastases. Amide-type LAs have been shown to interact positively with each of these steps. In a model of endotoxin-induced lung injury in rats, the application of ropivacaine IV was shown to significantly decrease ICAM-1 expression, as well as neutrophil adhesion and the concentration of albumin in bronchoalveolar lavage [40]. Piegeler et al. [41] exposed mice to either nebulized normal saline or lipopolysaccharide. The addition of ropivacaine showed a significant reduction in excess lung water, permeability index, and myeloperoxidase activity compared to the control. In the treatment group significant reduction in Src activation/expression, as well as ICAM-1 expression and caveolin-1 phosphorylation, was observed. It was concluded that ropivacaine was efficient in this model to treat the cause of pulmonary edema. Another study demonstrated that lidocaine and ropivacaine blocked inflammatory TNFα signaling in pulmonary endothelial cells by attenuating p85 recruitment to TNF-receptor 1, resulting in decreasing Akt, endothelial nitric oxide synthase, and Src phosphorylation [42]. Src activation has been shown to be responsible for a massive loss of endothelial barrier function, [43] which may also contribute to an increase in the extravasation of CTCs [44]. ICAM-1 is expressed by many different types of cancer cells [45,46,47,48] and may play an important role in the adhesion of CTCs on the endothelium, resulting in enhancing the extravasation of CTCs [47,48,49,50,51]. Other investigations have also shown that polymorphonuclear neutrophils (PNMs) activation and priming were significantly reduced by LAs [52,53,54,55]. If most of the studies pointed out beneficial effects, some did not show any benefit, which could be explained by methodological issues [56].
There is some evidence in vitro and animal models that LAs in cancer patients undergoing surgery may have the potential to reduce CTCs extravasation during the perioperative period by preserving the endothelial barrier function, reducing the PMNs adhesion, and therefore having the potential to reduce the transmigration of CTCs.

5. Direct Effects of LAs on Cancer Cells

LAs may have the potential to reduce/inhibit the occurrence of metastases by targeting different receptors/mechanisms (Table 2). Piegeler et al. [57] investigated in vitro the effects of lidocaine and ropivacaine on human lung cancer cells in the presence or not of an inflammatory stimulus (TNFα). The authors found that at clinical concentrations, both drugs in both conditions inhibited the Src expression and the ICAM-1 expression. The migration of cancer cells through a biological membrane was significantly reduced when the cells were in contact with the drugs for 4 h, but not if the exposure was limited to 15 min, followed by a washout. In these experiments, it was shown that these effects were not influenced by the addition of veratridine or tetrodotoxine, suggesting that these observations were independent of the sodium channel function. Moreover, chloroprocaine, an ester-type LA, did not show any effect. This work suggested that the positive effects of lidocaine and ropivacaine are time- and concentration-dependent, are not linked to any interaction with the sodium channel and are specific for the amide-type of LAs. Local anesthetics have also been shown to interfere with the regulation of gene expression by changing the methylation of DNA of breast cancer cells in vitro. In this investigation, lidocaine was able to demethylate DNA of these cells, a mechanism that will help to eliminate the cancer cells. Another investigation demonstrated that colon cancer cells’ invasiveness was decreased in vitro by the sodium channel 1.5 blocking effect [58]. Koh et al. [59] investigated the effect of lidocaine added to cisplatin in patients having highly aggressive triple-negative breast cancer. They found that this combination significantly increased the apoptosis ratio and the inhibitory effects of cisplatin. A higher expression of activated caspase-3 was observed. Lidocaine is also considered an ion channel regulator, which can regulate channel or membrane potential, such as mitochondrial membrane potential resulting in the mitochondria-related apoptosis [60]. Du and Coll [61] investigated in vitro the effects of lidocaine on the proliferation of colorectal cancer cells. The authors found that the proliferation of these cells was suppressed by the blockade of aerobic glycolysis.

6. Human Retrospective Studies

The interest in lidocaine application in cancer patients has been the subject of many clinical studies in recent years.
A retrospective analysis of medical records of a patient who had a mastectomy and axillary dissection for breast cancer was performed by Exadaktylos et al. [67]. Two groups were compared; one received general anesthesia (GA) alone, in the other, a paravertebral block was added. The results showed that the paravertebral group had at 37 months a better metastasis/recurrence-free survival of 94% compared to 77% in the control group.
In a study including 2239 patients undergoing pancreatectomy for adenocarcinoma, Zhang et al. [68] found that the addition of a bolus of lidocaine followed by a continuous infusion until the end of the surgery was associated with prolonged overall survival but without prolonged disease-free survival. The major weakness of this investigation was the short time of lidocaine administration, knowing that in vitro studies have shown that the positive effect of lidocaine in this context is time- and concentration-dependent.
Another retrospective study was performed by Christopherson et al. [69]. Patients undergoing surgery for colon cancer received a GA alone or GA with an epidural. The results showed the epidural group had a better survival only for the first two years.
Patients undergoing prostate cancer surgery receiving GA alone or GA with peridural anesthesia/analgesia were retrospectively compared. The results demonstrated that the epidural group had a better cancer recurrence-free of 57% compared to the control group.
The same type of anesthesia (with or without epidural) was used by de Oliveira et al. [70] to investigate retrospectively the recurrence-free interval in patients undergoing surgery for ovarian cancer. The results demonstrated that the addition of epidural was beneficial in terms of prolonging the recurrence-free interval.
Similar results were found in a retrospective analysis of patients receiving spinal anesthesia for excision of primary malignant melanoma of the lower extremity compared to the control group [71].
However, more recent retrospective studies did not show any benefits of adding a neuraxial block in terms of recurrence-free interval or mortality [70,71,72].
A recent meta-analysis [73] confirmed the discrepancy between the positive and negative results observed in the investigations dealing with neuraxial anesthesia and cancer outcome. Figure 1 shows the trial outcome from no effect to significant benefit versus year of publication and study size. It shows that the trends are first contrary to the first publications, more recent studies are rather negative and second, the larger study sample is associated with more negative results.
To summarize, it is well-known that the accuracy of medical records, the absence of randomization, blinding and standardization, and confounding factors are all serious bias, which can directly or indirectly influence the results and therefore do not allow for drawing valid conclusions.

7. Prospective Human Studies

Karmakar et al. [72] undertook a 5-year prospective follow-up of patients randomized to GA with or without thoracic paravertebral block undergoing a modified radical mastectomy. They found that the addition of a paravertebral block had little to no appreciable effect on local recurrence, metastases, or mortality. A similar study was conducted by Sessler et al. [74] In this trial, they found no difference in terms of breast cancer recurrence or mortality between those who received a paravertebral block or not; however, the major weakness of these two studies was the performance of only a single shot of LA in the paravertebral space, knowing that, after such a procedure, LA blood levels would be too low to modify the biology of the CTCs.

8. Conclusions

The perioperative period should be considered a crucial time in surgical oncology because chemotherapy will be delayed for a few weeks, opening a window for the occurrence of metastases. The application during this period of amide-type LAs may add a new option in the armentarium for the prevention of metastases.
In vitro and animal investigations have demonstrated the positive effect of lidocaine in the context of metastases on the biomolecular levels, including receptors and mechanisms.
Lidocaine, via different mechanisms, could inhibit/reduce cancer growth in vitro and in animal models. This provides valuable potential for its further application in cancer therapy and opens new insight for new drug discovery.
However, well-documented human studies in terms of dosage concentration and duration of time of application are needed to confirm the real benefit of this technique.

Author Contributions

A.B.: writing; A.B., J.A.: original draft preparation, A.B., J.A.: review. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Torre, L.A.; Bray, F.; Siegel, R.L.; Ferlay, J.; Lortet-Tieulent, J.; Jemal, A. Global cancer statistics, 2012. CA Cancer J. Clin. 2015, 65, 87–108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Christiansen, P.; Vejborg, I.; Kroman, N.; Holten, I.; Garne, J.P.; Vedsted, P.; Moller, S.; Lynge, E. Position paper: Breast cancer screening, diagnosis, and treatment in Denmark. Acta Oncol. 2014, 53, 433–444. [Google Scholar] [CrossRef] [PubMed]
  3. Horowitz, M.; Neeman, E.; Sharon, E.; Ben-Eliyahu, S. Exploiting the critical perioperative period to improve long-term cancer outcomes. Nat. Rev. Clin. Oncol. 2015, 12, 213–226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Grivennikov, S.I.; Greten, F.R.; Karin, M. Immunity, inflammation, and cancer. Cell 2010, 140, 883–899. [Google Scholar] [CrossRef] [Green Version]
  5. Mantovani, A.; Allavena, P.; Sica, A.; Balkwill, F. Cancer-related inflammation. Nature 2008, 454, 436–444. [Google Scholar] [CrossRef]
  6. Dillekas, H.; Rogers, M.S.; Straume, O. Are 90% of deaths from cancer caused by metastases? Cancer Med. 2019, 8, 5574–5576. [Google Scholar] [CrossRef] [Green Version]
  7. Guan, X. Cancer metastases: Challenges and opportunities. Acta Pharm. Sin. B 2015, 5, 402–418. [Google Scholar] [CrossRef] [Green Version]
  8. Hollmann, M.W.; Durieux, M.E. Local anesthetics: Effects on inflammation, wound healing and coagulation. Prog. Anesthesiol. 2000, 14, 291–304. [Google Scholar]
  9. Hollmann, M.W.; Durieux, M.E. Effects on the central nervous system and bronchial reactivity. Prog. Anesthesiol. 2000, 14, 323–336. [Google Scholar]
  10. Hollmann, M.W.; Difazio, C.A.; Durieux, M.E. Ca-signaling G-protein-coupled receptors: A new site of local anesthetic action? Reg. Anesth. Pain Med. 2001, 26, 565–571. [Google Scholar] [CrossRef]
  11. Lin, E.; Calvano, S.E.; Lowry, S.F. Cytokine response in abdominal surgery. In Cytokines and the Abdominal Surgeon; Landes Bioscience: Austin, TX, USA, 1998; pp. 17–34. [Google Scholar]
  12. Brochner, A.C.; Toft, P. Pathophysiology of the systemic inflammatory response after major accidental trauma. Scand. J. Trauma Resusc. Emerg. Med. 2009, 17, 43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. De Wit, S.; vaevon Dalum, G.; Terstappen, L.W. Detection of circulating tumor cells. Scientifica 2014, 2014, 819362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Gorges, T.M.; Pantel, K. Circulating tumor cells as therapy-related biomarkers in cancer patients. Cancer Immunol. Immunother. 2013, 62, 931–939. [Google Scholar] [CrossRef] [PubMed]
  15. Asworth, T.R. A case of cancer in which cells similar to those in tumors were seen in the blood after death. Aust. Med. J. 1869, 14, 146–149. [Google Scholar]
  16. Glodblatt, S.A.; Nadel, E.M. Cancer Cells in the Circulating Blood. Cancer Prog. 1963, 92, 119–140. [Google Scholar]
  17. Sergeant, G.; Roskams, T.; van Pelt, J.; Houtmeyers, F.; Aerts, R.; Topal, B. Perioperative cancer cell dissemination detected with a real-time RT-PCR assay for EpCAM is not associated with worse prognosis in pancreatic ductal adenocarcinoma. BMC Cancer 2011, 11, 47. [Google Scholar] [CrossRef] [Green Version]
  18. Lurje, G.; Schiesser, M.; Claudius, A.; Schneider, P.M. Circulating tumor cells in gastrointestinal malignancies: Current techniques and clinical implications. J. Oncol. 2010, 2010, 392652. [Google Scholar] [CrossRef] [Green Version]
  19. Plaks, V.; Koopman, C.D.; Werb, Z. Cancer. Circulating tumor cells. Science 2013, 341, 1186–1188. [Google Scholar] [CrossRef]
  20. Alsina, E.; Matute, E.; Ruiz-Huerta, A.D.; Gilsanz, F. The effects of sevoflurane or remifentanil on the stress response to surgical stimulus. Curr. Pharm. Des. 2014, 20, 5449–5468. [Google Scholar] [CrossRef]
  21. Angele, M.K.; Faist, E. Clinical review: Immunodepression in the surgical patient and increased susceptibility to infection. Crit. Care 2002, 6, 298–305. [Google Scholar] [CrossRef]
  22. Salmon, P.; Kaufman, L. Preoperative anxiety and endocrine response to surgery. Lancet 1990, 335, 1340. [Google Scholar] [CrossRef]
  23. Chachkhiani, I.; Gurlich, R.; Maruna, P.; Frasko, R.; Lindner, J. The postoperative stress response and its reflection in cytokine network and leptin plasma levels. Physiol. Res. 2005, 54, 279–285. [Google Scholar] [CrossRef] [PubMed]
  24. Angka, L.; Khan, S.T.; Kilgour, M.K.; Xu, R.; Kennedy, M.A. Auer RC: Dysfunctional Natural Killer Cells in the Aftermath of Cancer Surgery. Int. J. Mol. Sci. 2017, 18, 1787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Kiessling, R.; Klein, E.; Wigzell, H. “Natural” killer cells in the mouse. I. Cytotoxic cells with specificity for mouse Moloney leukemia cells. Specificity and distribution according to genotype. Eur. J. Immunol. 1975, 5, 112–117. [Google Scholar] [CrossRef]
  26. Baginska, J.; Viry, E.; Paggetti, J.; Medves, S.; Berchem, G.; Moussay, E.; Janji, B. The critical role of the tumor microenvironment in shaping natural killer cell-mediated anti-tumor immunity. Front. Immunol. 2013, 4, 490. [Google Scholar] [CrossRef] [Green Version]
  27. Iannone, F.; Porzia, A.; Peruzzi, G.; Birarelli, P.; Milana, B.; Sacco, L.; Dinatale, G.; Peparini, N.; Prezioso, G.; Battella, S.; et al. Effect of surgery on pancreatic tumor-dependent lymphocyte asset: Modulation of natural killer cell frequency and cytotoxic function. Pancreas 2015, 44, 386–933. [Google Scholar] [CrossRef] [Green Version]
  28. Ramirez, M.F.; Ai, D.; Bauer, M.; Vauthey, J.N.; Gottumukkala, V.; Kee, S.; Shon, D.; Truty, M.; Kuerer, H.M.; Kurz, A.; et al. Innate immune function after breast, lung, and colorectal cancer surgery. J. Surg. Res. 2015, 194, 185–193. [Google Scholar] [CrossRef]
  29. Ben-Eliyahu, S.; Page, G.G.; Yirmiya, R.; Shakhar, G. Evidence that stress and surgical interventions promote tumor development by suppressing natural killer cell activity. Int. J. Cancer 1999, 80, 880–888. [Google Scholar] [CrossRef]
  30. Melamed, R.; Bar-Yosef, S.; Shakhar, G.; Shakhar, K.; Ben-Eliyahu, S. Suppression of natural killer cell activity and promotion of tumor metastasis by ketamine, thiopental, and halothane, but not by propofol: Mediating mechanisms and prophylactic measures. Anesth. Analg. 2003, 97, 1331–1339. [Google Scholar] [CrossRef]
  31. Krog, J.; Hokland, M.; Ahlburg, P.; Parner, E.; Tonnesen, E. Lipid solubility- and concentration-dependent attenuation of in vitro natural killer cell cytotoxicity by local anesthetics. Acta Anaesthesiol. Scand. 2002, 46, 875–881. [Google Scholar] [CrossRef]
  32. Ramirez, M.F.; Tran, P.; Cata, J.P. The effect of clinically therapeutic plasma concentrations of lidocaine on natural killer cell cytotoxicity. Reg. Anesth. Pain Med. 2015, 40, 43–48. [Google Scholar] [CrossRef] [PubMed]
  33. Jaura, A.I.; Flood, G.; Gallagher, H.C.; Buggy, D.J. Differential effects of serum from patients administered distinct anaesthetic techniques on apoptosis in breast cancer cells in vitro: A pilot study. Br. J. Anaesth. 2014, 113 (Suppl. S1), i63–i67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Dong, H.; Zhang, Y.; Xi, H. The effects of epidural anaesthesia and analgesia on natural killer cell cytotoxicity and cytokine response in patients with epithelial ovarian cancer undergoing radical resection. J. Int Med. Res. 2012, 40, 1822–1829. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, H.L.; Yan, H.D.; Liu, Y.Y.; Sun, B.Z.; Huang, R.; Wang, X.S.; Lei, W.F. Intraoperative intravenous lidocaine exerts a protective effect on cell-mediated immunity in patients undergoing radical hysterectomy. Mol. Med. Rep. 2015, 12, 7039–7044. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Zhou, D.; Gu, F.M.; Gao, Q.; Li, Q.L.; Zhou, J.; Miao, C.H. Effects of anesthetic methods on preserving anti-tumor T-helper polarization following hepatectomy. World J. Gastroenterol. 2012, 18, 3089–3098. [Google Scholar] [CrossRef] [PubMed]
  37. Bernabe, D.G.; Tamae, A.C.; Biasoli, E.R.; Oliveira, S.H. Stress hormones increase cell proliferation and regulates interleukin-6 secretion in human oral squamous cell carcinoma cells. Brain Behav. Immun. 2011, 25, 574–583. [Google Scholar] [CrossRef] [Green Version]
  38. Yang, E.V.; Kim, S.J.; Donovan, E.L.; Chen, M.; Gross, A.C.; Webster Marketon, J.I.; Barsky, S.H.; Glaser, R. Norepinephrine upregulates VEGF, IL-8, and IL-6 expression in human melanoma tumor cell lines: Implications for stress-related enhancement of tumor progression. Brain Behav. Immun. 2009, 23, 267–275. [Google Scholar] [CrossRef] [Green Version]
  39. Ortiz, M.P.; Godoy, M.C.; Schlosser, R.S.; Ortiz, R.P.; Godoy, J.P.; Santiago, E.S.; Rigo, F.K.; Beck, V.; Duarte, T.; Duarte, M.M.; et al. Effect of endovenous lidocaine on analgesia and serum cytokines: Double-blinded and randomized trial. J. Clin. Anesth. 2016, 35, 70–77. [Google Scholar] [CrossRef]
  40. Blumenthal, S.; Borgeat, A.; Pasch, T.; Reyes, L.; Booy, C.; Lambert, M.; Schimmer, R.C.; Beck-Schimmer, B. Ropivacaine decreases inflammation in experimental endotoxin-induced lung injury. Anesthesiology 2006, 104, 961–969. [Google Scholar] [CrossRef]
  41. Piegeler, T.; Dull, R.O.; Hu, G.; Castellon, M.; Chignalia, A.Z.; Koshy, R.G.; Votta-Velis, E.G.; Borgeat, A.; Schwartz, D.E.; Beck-Schimmer, B.; et al. Ropivacaine attenuates endotoxin plus hyperinflation-mediated acute lung injury via inhibition of early-onset Src-dependent signaling. BMC Anesthesiol. 2014, 14, 57. [Google Scholar] [CrossRef] [Green Version]
  42. Piegeler, T.; Votta-Velis, E.G.; Bakhshi, F.R.; Mao, M.; Carnegie, G.; Bonini, M.G.; Schwartz, D.E.; Borgeat, A.; Beck-Schimmer, B.; Minshall, R.D. Endothelial barrier protection by local anesthetics: Ropivacaine and lidocaine block tumor necrosis factor-alpha-induced endothelial cell Src activation. Anesthesiology 2014, 120, 1414–1428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Hu, G.; Minshall, R.D. Regulation of transendothelial permeability by Src kinase. Microvasc. Res. 2009, 77, 21–25. [Google Scholar] [CrossRef] [PubMed]
  44. Garcia-Roman, J.; Zentella-Dehesa, A. Vascular permeability changes involved in tumor metastasis. Cancer Lett. 2013, 335, 259–269. [Google Scholar] [CrossRef]
  45. Fakler, C.R.; Wu, B.; McMicken, H.W.; Geske, R.S.; Welty, S.E. Molecular mechanisms of lipopolysaccharide induced ICAM-1 expression in A549 cells. Inflamm. Res. 2000, 49, 63–72. [Google Scholar] [CrossRef] [PubMed]
  46. Guney, N.; Soydinc, H.O.; Derin, D.; Tas, F.; Camlica, H.; Duranyildiz, D.; Yasasever, V.; Topuz, E. Serum levels of intercellular adhesion molecule ICAM-1 and E-selectin in advanced stage non-small cell lung cancer. Med. Oncol. 2008, 25, 194–200. [Google Scholar] [CrossRef]
  47. Lin, Y.C.; Shun, C.T.; Wu, M.S.; Chen, C.C. A novel anticancer effect of thalidomide: Inhibition of intercellular adhesion molecule-1-mediated cell invasion and metastasis through suppression of nuclear factor-kappaB. Clin. Cancer Res. 2006, 12, 7165–7173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Melis, M.; Spatafora, M.; Melodia, A.; Pace, E.; Gjomarkaj, M.; Merendino, A.M.; Bonsignore, G. ICAM-1 expression by lung cancer cell lines: Effects of upregulation by cytokines on the interaction with LAK cells. Eur. Respir. J. 1996, 9, 1831–1888. [Google Scholar] [CrossRef]
  49. Beck-Schimmer, B.; Schimmer, R.C.; Warner, R.L.; Schmal, H.; Nordblom, G.; Flory, C.M.; Lesch, M.E.; Friedl, H.P.; Schrier, D.J.; Ward, P.A. Expression of lung vascular and airway ICAM-1 after exposure to bacterial lipopolysaccharide. Am. J. Respir. Cell Mol. Biol. 1997, 17, 344–352. [Google Scholar] [CrossRef] [Green Version]
  50. Liu, G.; Place, A.T.; Chen, Z.; Brovkovych, V.M.; Vogel, S.M.; Muller, W.A.; Skidgel, R.A.; Malik, A.B.; Minshall, R.D. ICAM-1-activated Src and eNOS signaling increase endothelial cell surface PECAM-1 adhesivity and neutrophil transmigration. Blood 2012, 120, 1942–1952. [Google Scholar] [CrossRef] [Green Version]
  51. Wu, Q.D.; Wang, J.H.; Condron, C.; Bouchier-Hayes, D.; Redmond, H.P. Human neutrophils facilitate tumor cell transendothelial migration. Am. J. Physiol. Cell Physiol. 2001, 280, C814–C822. [Google Scholar] [CrossRef] [Green Version]
  52. Fischer, L.G.; Bremer, M.; Coleman, E.J.; Conrad, B.; Krumm, B.; Gross, A.; Hollmann, M.W.; Mandell, G.; Durieux, M.E. Local anesthetics attenuate lysophosphatidic acid-induced priming in human neutrophils. Anesth. Analg. 2001, 92, 1041–1047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Picardi, S.; Cartellieri, S.; Groves, D.; Hahnenkamp, K.; Gerner, P.; Durieux, M.E.; Stevens, M.F.; Lirk, P.; Hollmann, M.W. Local anesthetic-induced inhibition of human neutrophil priming: The influence of structure, lipophilicity, and charge. Reg. Anesth. Pain Med. 2013, 38, 9–15. [Google Scholar] [CrossRef] [PubMed]
  54. Hollmann, M.W.; Gross, A.; Jelacin, N.; Durieux, M.E. Local anesthetic effects on priming and activation of human neutrophils. Anesthesiology 2001, 95, 113–122. [Google Scholar] [CrossRef] [PubMed]
  55. Hollmann, M.W.; Kurz, K.; Herroeder, S.; Struemper, D.; Hahnenkamp, K.; Berkelmans, N.S.; den Bakker, C.G.; Durieux, M.E. The effects of S(−)-, R(+)-, and racemic bupivacaine on lysophosphatidate-induced priming of human neutrophils. Anesth. Analg. 2003, 97, 1053–1058. [Google Scholar] [CrossRef] [PubMed]
  56. Li, T.; Chen, L.; Zhao, H.; Wu, L.; Masters, J.; Han, C.; Hirota, K.; Ma, D. Both Bupivacaine and Levobupivacaine inhibit colon cancer cell growth but not melanoma cells in vitro. J. Anesth. 2019, 33, 17–25. [Google Scholar] [CrossRef] [Green Version]
  57. Piegeler, T.; Votta-Velis, E.G.; Liu, G.; Place, A.T.; Schwartz, D.E.; Beck-Schimmer, B.; Minshall, R.D.; Borgeat, A. Antimetastatic potential of amide-linked local anesthetics: Inhibition of lung adenocarcinoma cell migration and inflammatory Src signaling independent of sodium channel blockade. Anesthesiology 2012, 117, 548–559. [Google Scholar] [CrossRef] [Green Version]
  58. Baptista-Hon, D.T.; Robertson, F.M.; Robertson, G.B.; Owen, S.J.; Rogers, G.W.; Lydon, E.L.; Lee, N.H.; Hales, T.G. Potent inhibition by ropivacaine of metastatic colon cancer SW620 cell invasion and NaV1.5 channel function. Br. J. Anaesth. 2014, 113 (Suppl. S1), i39–i48. [Google Scholar] [CrossRef] [Green Version]
  59. Koh, S.Y.; Moon, J.Y.; Unno, T.; Cho, S.K. Baicalein Suppresses Stem Cell-Like Characteristics in Radio- and Chemoresistant MDA-MB-231 Human Breast Cancer Cells through Up-Regulation of IFIT2. Nutrients 2019, 11, 624–643. [Google Scholar] [CrossRef] [Green Version]
  60. Ye, L.; Zhang, Y.; Chen, Y.J.; Liu, Q. Anti-tumor effects of lidocaine on human gastric cancer cells in vitro. Bratisl. Lek Listy 2019, 120, 212–217. [Google Scholar] [CrossRef]
  61. Du, J.; Zhang, L.; Ma, H.; Wang, Y.; Wang, P. Lidocaine Suppresses Cell Proliferation and Aerobic Glycolysis by Regulating circHOMER1/miR-138-5p/HEY1 Axis in Colorectal Cancer. Cancer Manag. Res. 2020, 12, 5009–5022. [Google Scholar] [CrossRef]
  62. House, C.D.; Vaske, C.J.; Schwartz, A.M.; Obias, V.; Frank, B.; Luu, T.; Sarvazyan, N.; Irby, R.; Strausberg, R.L.; Hales, T.G.; et al. Voltage-gated Na+ channel SCN5A is a key regulator of a gene transcriptional network that controls colon cancer invasion. Cancer Res. 2010, 70, 6957–6967. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Sakaguchi, M.; Kuroda, Y.; Hirose, M. The antiproliferative effect of lidocaine on human tongue cancer cells with inhibition of the activity of epidermal growth factor receptor. Anesth. Analg. 2006, 102, 1103–1107. [Google Scholar] [CrossRef] [PubMed]
  64. Xing, W.; Chen, D.T.; Pan, J.H.; Chen, Y.H.; Yan, Y.; Li, Q.; Xue, R.F.; Yuan, Y.F.; Zeng, W.A. Lidocaine Induces Apoptosis and Suppresses Tumor Growth in Human Hepatocellular Carcinoma Cells In Vitro and in a Xenograft Model In Vivo. Anesthesiology 2017, 126, 868–881. [Google Scholar] [CrossRef] [PubMed]
  65. Lirk, P.; Berger, R.; Hollmann, M.W.; Fiegl, H. Lidocaine time- and dose-dependently demethylates deoxyribonucleic acid in breast cancer cell lines in vitro. Br. J. Anaesth. 2012, 109, 200–207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. D’Agostino, G.; Saporito, A.; Cecchinato, V.; Silvestri, Y.; Borgeat, A.; Anselmi, L.; Uguccioni, M. Lidocaine inhibits cytoskeletal remodelling and human breast cancer cell migration. Br. J. Anaesth. 2018, 121, 962–968. [Google Scholar] [CrossRef] [Green Version]
  67. Exadaktylos, A.K.; Buggy, D.J.; Moriarty, D.C.; Mascha, E.; Sessler, D.I. Can anesthetic technique for primary breast cancer surgery affect recurrence or metastasis? Anesthesiology 2006, 105, 660–664. [Google Scholar] [CrossRef] [Green Version]
  68. Zhang, H.; Yang, L.; Zhu, X.; Zhu, M.; Sun, Z.; Cata, J.P.; Chen, W.; Miao, C. Association between intraoperative intravenous lidocaine infusion and survival in patients undergoing pancreatectomy for pancreatic cancer: A retrospective study. Br. J. Anaesth. 2020, 125, 141–148. [Google Scholar] [CrossRef]
  69. Christopherson, R.; James, K.E.; Tableman, M.; Marshall, P.; Johnson, F.E. Long-term survival after colon cancer surgery: A variation associated with choice of anesthesia. Anesth. Analg. 2008, 107, 325–332. [Google Scholar] [CrossRef] [Green Version]
  70. De Oliveira, G.S., Jr.; Ahmad, S.; Schink, J.C.; Singh, D.K.; Fitzgerald, P.C.; McCarthy, R.J. Intraoperative neuraxial anesthesia but not postoperative neuraxial analgesia is associated with increased relapse-free survival in ovarian cancer patients after primary cytoreductive surgery. Reg. Anesth. Pain Med. 2011, 36, 271–277. [Google Scholar] [CrossRef]
  71. Gottschalk, A.; Brodner, G.; Van Aken, H.K.; Ellger, B.; Althaus, S.; Schulze, H.J. Can regional anaesthesia for lymph-node dissection improve the prognosis in malignant melanoma? Br. J. Anaesth. 2012, 109, 253–259. [Google Scholar] [CrossRef] [Green Version]
  72. Karmakar, M.K.; Samy, W.; Lee, A.; Li, J.W.; Chan, W.C.; Chen, P.P.; Tsui, B.C.H. Survival Analysis of Patients with Breast Cancer Undergoing a Modified Radical Mastectomy With or Without a Thoracic Paravertebral Block: A 5-Year Follow-up of a Randomized Controlled Trial. Anticancer Res. 2017, 37, 5813–5820. [Google Scholar] [PubMed]
  73. Weng, M.; Chen, W.; Hou, W.; Li, L.; Ding, M.; Miao, C. The effect of neuraxial anesthesia on cancer recurrence and survival after cancer surgery: An updated meta-analysis. Oncotarget 2016, 7, 15262–15273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Sessler, D.I.; Pei, L.; Huang, Y.; Fleischmann, E.; Marhofer, P.; Kurz, A.; Mayers, D.B.; Meyer-Treschan, T.A.; Grady, M.; Tan, E.Y.; et al. Breast Cancer Recurrence C: Recurrence of breast cancer after regional or general anaesthesia: A randomised controlled trial. Lancet 2019, 394, 1807–1815. [Google Scholar] [CrossRef]
Figure 1. The outcome from 2 = positive result to 0 = null result is shown on the y axis. The year of publication is on the x axis.
Figure 1. The outcome from 2 = positive result to 0 = null result is shown on the y axis. The year of publication is on the x axis.
Medicina 58 00882 g001
Table 1. Natural killer cell activity: Influence of commonly used drugs in anesthesia.
Table 1. Natural killer cell activity: Influence of commonly used drugs in anesthesia.
Human Studies
Without SurgeryWith Surgery
Opioids
  • Morphine
↑, =, ↓=, ↓
  • Fentanyl
↑, ==, ↓
  • Remifentanil
=?
Ketamine?↑,↓
Propofol?
NSAID
Local anesthetics
  • IV infusion
  • Regional application
(=)(=)
↑ increase; = neutral effect; ↓ decrease; ? no available data.
Table 2. Biomolecular Mechanisms and Antimetastatic Properties of Amide-Type Local Anesthetics.
Table 2. Biomolecular Mechanisms and Antimetastatic Properties of Amide-Type Local Anesthetics.
Inhibition of Src-Kinase/ICAM-1 PhosphorylationPiegeler et al., 2014, 2012 [42,57]
Downregulation of VGSC (Voltage-Gated Sodium Channel)House et al., 2010 [62]
Antiproliferative effectsSakaguchi et al., 2006 [63]
Increase the apoptotic effectXing et al., 2017 [64]
Increase the demethylationLirk et al., 2012 [65]
Inhibition of cytoskeletal remodelingD’Agostino et al., 2018 [66]
Potentiation in vitro and in vivo of NK cell activityRamirez et al., 2015 [32]
Jaura et al., 2014 [33]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Borgeat, A.; Aguirre, J. Impact of Local Anesthetics on Cancer Behavior and Outcome during the Perioperative Period: A Review. Medicina 2022, 58, 882. https://doi.org/10.3390/medicina58070882

AMA Style

Borgeat A, Aguirre J. Impact of Local Anesthetics on Cancer Behavior and Outcome during the Perioperative Period: A Review. Medicina. 2022; 58(7):882. https://doi.org/10.3390/medicina58070882

Chicago/Turabian Style

Borgeat, Alain, and José Aguirre. 2022. "Impact of Local Anesthetics on Cancer Behavior and Outcome during the Perioperative Period: A Review" Medicina 58, no. 7: 882. https://doi.org/10.3390/medicina58070882

Article Metrics

Back to TopTop